Next Article in Journal
Crustal Apparent Density Variations in the Middle Segment of the North Tianshan Mountains and Their Tectonic Significance
Previous Article in Journal
A Review of Industry 4.0 Assessment Instruments for Digital Transformation
Previous Article in Special Issue
Aromatized Wine-Type Beverages with Added Byproducts of Essential Rose Oil Industry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Assessment of Bioavailability Parameters of Mono- and Bistriazole Derivatives of Propynoylbetulin

1
Department of Organic Chemistry, Faculty of Pharmaceutical Sciences in Sosnowiec, Medical University of Silesia, Katowice, 4 Jagiellońska Str., 41-200 Sosnowiec, Poland
2
Students Scientific Group of Department of Organic Chemistry, Faculty of Pharmaceutical Sciences in Sosnowiec, Medical University of Silesia, Katowice, 4 Jagiellońska Str., 41-200 Sosnowiec, Poland
3
Students Research Club at the Department of Community Pharmacy, Faculty of Pharmaceutical Sciences in Sosnowiec, Medical University of Silesia, Katowice, 3 Kasztanowa Str., 41-200 Sosnowiec, Poland
4
Department of Community Pharmacy, Faculty of Pharmaceutical Sciences in Sosnowiec, Medical University of Silesia, Katowice, 3 Kasztanowa Str., 41-200 Sosnowiec, Poland
*
Author to whom correspondence should be addressed.
Appl. Sci. 2024, 14(5), 1695; https://doi.org/10.3390/app14051695
Submission received: 1 February 2024 / Revised: 16 February 2024 / Accepted: 19 February 2024 / Published: 20 February 2024

Abstract

:
Bioavailability describes the properties that determine the passage of a compound through biological barriers. In many cases, bioavailability depends on the lipophilicity of the compound. In this study, the lipophilicity as well as other bioavailability properties of the mono- and bistriazole derivatives of betulin are presented. The lipophilicity was determined using RP-TLC and theoretical methods. The experimental lipophilicity of mono- and bistriazole derivatives is in the range from 4.39 to 7.85 and from 3.75 to 8.83, respectively. The lipophilicity of mono- and bistriazoles is similar, and the logPTLC depends on the type of substituent at the triazole ring. The introduction of a substituent with oxygen and nitrogen atoms decreases lipophilicity. Comparing the experimental and theoretical lipophilicity shows that the milogP and XLOGP3 programs best reproduce the experimental values. The in silico-determined pharmacokinetic parameters show that monotriazole derivatives could be used as oral drugs while bistriazoles show low availability after oral administration. Triazoles could be used as transdermal drugs. The analysis of in silico bioavailability parameters shows that the type of substituent at the triazole ring influences the pharmacokinetic properties, while the number of triazole rings slightly affects the bioavailability properties of the compound.

1. Introduction

Bioavailability describes the properties of a drug which determine its absorption, distribution, metabolism, and elimination from an organism. One of the most important properties which determine bioavailability is lipophilicity, which describes the ability of a compound to cross the biological barrier. Often, with increased lipophilicity, the biological activity of a compound increases due to better permeability through biological membranes and a higher concentration of the compound inside the cell. However, too high lipophilicity may reduce the absorption of the compound due to its absorption in biological membranes [1,2,3,4]. A measure of lipophilicity is the partition coefficient between non-polar and polar phases defined as P or logP. The precursor of the determination of the P coefficient was Hansch, who designated the concentration of a compound in a composite mixture with two phases. As a non-polar and polar phase he used n-octanol and water, respectively. The method named “shake-flask” is laborious, time-consuming, and low-precision. For this reason, liquid chromatographic methods are most popular, namely, thin-layer chromatography in a normal phase system (NP-TLC) and its variant, thin-layer chromatography in a reversed phase system (RP-TLC). Thin-layer chromatography methods have many advantages compared with the “shake-flask” method. The most important is possibility of determined the lipophilicity of many compounds in the same time, low time-consuming, high-precisions and high reproducibility [5,6,7,8,9,10].
Lipophilicity is connected to other parameters, like hydrophobicity (φ0), permeability by Caco-2 membrane, skin, blood–brain barrier (BB), and penetration to the central nervous system (CNS). The design of a new compound is often supported by computational methods which determine the lipophilicity and permeability parameters using in silico methods [11,12]. However, most programs determining the logP parameter calculate lipophilicity using atomic methods which do not include every structural parameter, such as the spatial arrangement of atoms or the position of a heterocyclic atom in a ring. In the literature, many examples of compounds described for which the theoretical lipophilicity is very different from the experimental one. The obtained theoretical value could eliminate the compounds from further research. For this reason, determining the correlation between experimental and theoretical parameters is still an important problem [13,14,15,16,17].
In the past few decades, the isolation of compounds from plant material has attracted considerable attention due to pharmacologically active compounds using in natural medicine. In many cases, natural compounds have moderate biological activity and bioavailability [18,19,20,21]. One of the first compounds isolated in pure form from plant material was betulin (Figure 1). Betulin belongs to a pentacyclic triterpene alcohol with a lupane skeleton. The chemical structure of betulin contains four six-member rings and one five-member ring. Moreover, they have an isopropylidene group at the C19 position and two hydroxyl groups at the C3 and C28 positions. The compound is characterized by a broad spectrum of biological activity, like anticancer, antibacterial, antiviral, antifungal, hepatoprotective, and anti-inflammatory properties. The used of betulin in treatment is limited due to its low water solubility [22,23,24,25,26,27,28].
Chemical modifications of its structure allow new semisynthetic compounds to be obtained which have better biological properties. The replacement of one or two hydroxyl groups by an ester, amide, or carbamate group influences the activity and solubility in water of the obtained derivatives [29,30].
An interesting trend observed in chemical synthesis is the introduction of heterocyclic moieties. The insertion of moiety with nitrogen, oxygen, or/and sulfur atoms affects biological activity because heteroatoms could create a hydrogen bond with the biological target and increase water solubility [31,32]. An especially interesting moiety is the triazole scaffold which consists of two carbon and three nitrogen atoms in a five-membered unsaturated ring. The triazole ring is created between organic azide and alkyne derivatives in the reaction of 1,3-dipolar cycloaddition [33,34,35]. The triazole ring is found in many clinically used drugs, like Fluconazole, Bittertanol, Cyproconazole, Trazodone, Triazolam, Ribavirin, Isavuconazole, and Cefatrizine [35,36,37,38,39,40]. In the literature, the triazole moiety is often used as a linker between two compounds showing high biological activity, which allows hybrids with new biological properties to be obtained. In many cases, the modification of natural compounds by introducing a triazole substituent affects biological activities, like anticancer, antibacterial, antifungal, antiviral, antimalarial, and antioxidant [41,42,43,44,45].
In our scientist group, we deal with the modification of the betulin skeleton by introducing various substituents at the C3, C28, and C30 positions [46,47]. One of the important research directions was the introduction of the triazole moiety to the betulin scaffold. The obtained compounds exhibited anticancer, antibacterial, and antiviral activity [13,48,49,50]. Continuing research on the biological potential of triazole derivatives of betulin, we decided to designate the bioavailability parameters of mono- and bistriazole derivatives of propynoylbetulin. Lipophilicity was determined using experimental and theoretical methods. Moreover, the correlation between these two values was studied. The research was supplemented by in silico-determined bioavailability parameters.

2. Materials and Methods

2.1. General Method

Reversed phase thin-layer chromatography (RP-TLC) was carried out on silica gel RP-18 F254S plates (Merck, Darmstadt, Germany) using a different mixture of acetone and (tris-hydroxymethyl)aminomethane (Tris) as an eluent. The concentration of acetone in the mobile phase changed from 60% to 90%. The chromatographic spots were visualized using UV light (λ = 254 nm) and by spraying with a solution of 10% sulfuric acid for the reference compounds and compounds 118, respectively.

2.2. Data Set

Propynoilbetulin 12 and their triazole derivatives 318 were synthesised using the methods found in the literature [48,51]. In the first stage, betulin was converted to derivatives 12 in the presence of propiolic acid, N,N′-dicyclohexylcarbodiimide (DCC), and 4-dimethylaminopyridine (DMAP) at room temperature. After purification using column chromatography, 28-propynoilbetulin 1 and 3,28-dipropynoilbetulin 2 were obtained with 52% and 15% yields, respectively. Next, compounds 12 were treated with organic azide in the presence of copper iodide and toluene at reflux. The crude products were purified using column chromatography. Triazole derivatives 318 obtained with 48-78% yield. The molecular structure of derivatives 118 is presented in Figure 2.

2.3. Assessment of Experimental and Calculated Lipophilicity

The experimental lipophilicity was determined using the RP-TLC method. As a stationary phase, un-polar silicone oil placed on the silica gel layer was used, and as a mobile phase, the mixture of (tris-hydroxymethyl)aminomethane (0.2 M, pH = 7.4) with acetone was used. The concentration of acetone in the mobile phase was in the range from 60% to 90% in 5% increments.
A total of 2 µL of the ethanolic solutions of compounds 118 (1 mg/mL) and reference compounds was marked on the chromatographic plates. After visualization, the retardation factor (Rf) value for each compound was designated. The RM parameter was calculated using Equation (1) below.
R M = log 1 R f 1 ,
The relative lipophilicity parameter (RM0) was obtained by the extrapolation of linear regression between RM and acetone concentration (C) to the zero value of the organic solvent (Equation (2)).
R M = R M 0 + b C ,
The RM0 and the slope of the regression plot (b) were used to calculated the hydrophobic index (φ0) according to Equation (3).
φ 0 = R M 0 b ,
The experimental lipophilicity study was supplemented by theoretical lipophilicity, which was determined using vcclab (2005) [52] and molinspiration (v2022.09) [53] software.

2.4. Assessment of Bioavailability Parameters

The Lipinski and Veber parameters, such as molecular weight (M), topological polar surface (TPSA), the number of acceptors (HAs) and donors (HDs) of the hydrogen bond, and the number of rotatable bonds (RTs) were determined using SwissADMET (2022) platform [54]. Moreover, the absorption and neurotoxicity parameters, like Caco-2 permability (logPapp), human intestinal absorption (HIA), skin permeability (logKp), blood–brain barrier permeability (logBB), and central nervous system (logPS) permeability were designated using pkCSM (2015) software [55].

3. Results

In the first step of the research, the experimental lipophilicity was determined by using the RP-TLC method. As a mobile phase, the mixture of acetone and Tris buffer was used. The concentration of acetone in the mobile phase was in the range from 60% to 90%. The obtained Rf values for each compound 118 were converted to RM and RM0 parameters using Equations (1) and (2), and are presented in Table 1.
The experimental values of RM0 and slope b were used to determine the hydrophobicity index according to Equation (3), and are presented in Table 1.
The RM0 parameter was converted to the lipophilicity parameter (logPTLC) using the calibration curve. As reference compounds, 4-bromoacetophenone (A), benzophenone (B), anthracene (C), dibenzyl (D), 9-phenylanthracene (E), and dichlorodiphenyltrichloroethane (DDT) (F) were used whose lipophilicity (logPlit) in the literature is in the rage from 2.43 to 6.38 [13,14].
The experimental lipophilicity of the reference compounds was determined similarly to the tested compounds 118 (Table 2). The calibration curve (Equation (4)) was obtained by linear regression between the experimental RM0 parameter and lipophilicity (logPlit) in the literature.
LogPTLC = 1.0878 RM0 + 0.3972   (r = 0.988; SD = 0.263),
According to Equation (4), the experimental lipophilicity (logPTLC) of the reference compounds was calculated (Table 2). As seen in Table S1, the experimental results correlate well with the ones in the literature.
In the next step, the experimental lipophilicity was compared with the calculated lipophilicity which was determined using the ALOGPs, miLogP, XLOGP2, AClogP, and XLOGP3 programs. The results are shown in Table 3.
Lipophilicity is connected with other bioavailability parameters such as molecular mass (M), topological polar surface (TPSA), the number of acceptors (HAs) and donors (HDs) of the hydrogen bond, and the number of rotatable bonds (RTs). The research was supplemented with in silico-determined parameters responsible for the drug’s permeation of the biological barrier, like Caco-2 cell (logPapp), human intestinal absorption (HIA), skin permeability (logKp), blood–brain barrier permeability (logBB), and central nervous system (logPS) penetration [56,57,58]. The results obtained using pkCSM software are presented in Table 4.

4. Discussion

The use of new synthesised compounds in therapy depends on their biological activity and bioavailability. One of the most important properties of compounds is their lipophilicity which determines drug transport across the biological barrier. The determination of lipophilicity offers the opportunity to assess the availability and toxicity of a potential drug [1].
The propynoyl derivatives of betulin 1 and 2 are characterized by high lipophilicity which is equal to 6.68 and 9.20, respectively. Compounds 318 create two groups, the first of which includes the monotriazole derivatives of 1 and the second consists of the bistriazole derivatives of 2. In the group of monotriazole compounds 310, the lowest logPTLC value is shown by derivative 7 with the deoxy-β-D-glucopyranosyl substituent at the triazole ring. Derivatives of the benzyl group (35) exhibit comparable lipophilicity which is in the range from 7.09 to 7.24. The introduction of the phenyltiomethyl group at the triazole ring, compound 6, increases lipophilicity. Comparing the logPTLC value of compounds 1 and 810 shows that the introduction of a substituent with oxygen and nitrogen atoms decreases lipophilicity.
The transformation of 3,28-dipropynoylbetulin 2 to bistriazole derivatives 1118 decreases lipophilicity. Similar to the first group, in the group of bistriazole derivatives, the lowest value of lipophilicity is shown by compound 15 with the deoxy-β-D-glucopyranosyl substituent. In these groups of compounds (1118), similar correlations were observed as for 310. Comparing the lipophilicity of mono- and bistriazole compounds containing the same substituents at the triazole ring, bistriazole derivatives had higher log PTLC values. The exceptions were the compounds with a β-D-glucopyranosyl substituent (7 and 15) and a hydroxypropyl substituent (9 and 17), of which derivatives with one triazole ring (7 and 9) had higher lipophilicity.
The hydrophobicity index describes the water solubility of a compound; if the φ0 value is lower, then water solubility is better [59]. Comparing the φ0 value of triazole derivatives 318 and compounds 12 shows that the introduction of a triazole ring increases water solubility. The hydrophobicity index is in the range from 73.11 to 86.86 and from 59.69 to 86.29 for mono- 310 and bistriazoles 1118, respectively. The results show that mono- and bistriazole derivatives have comparable solubility in water.
Before determining the experimental lipophilicity, it is necessary to synthesize and purify the tested compound. The experimental method is not useful during the computer design of new active compounds. For this reason, the experimental lipophilicity is replaced by theoretical methods which use a different algorithm [60,61,62].
The calculated lipophilicity for triazole compounds 318 was determined using theoretical methods used in software that is available online [52,53]. The obtained values were in the range from 1.17 to 13.42 (Table 3). For all programs, the theoretical lipophilicity depends on the type of substituent at the triazole ring. A relationship was observed, namely, that monotriazoles 310 have lower lipophilicity than bistriazoles 1118. The exceptions are compounds 7 and 15 containing the deoxy-β-D-glucopyranosyl substituent. Bistriazole derivative 15 has lower lipophilicity due to the larger number of hydroxyl groups which reduces the logP parameter. Comparing the experimental and theoretical lipophilicity shows that logP values obtained using the AClogP and ALOGPs programs are lower than the experimental one (Figure 3). For other programs, the calculated lipophilicity is higher than the experimental logPTLC.
Due to the fact that different programs use different calculating algorithms, an important aspect of the research is to determine the correlation between the experimental and theoretical lipophilicity (Table 5). The correlation equation shows that the ALOGPs and XLOGP3 programs best reproduce the experimental lipophilicity, and the correlation cofactor is equal to 0.928 and 0.937, respectively (Table 5).
The lipophilicity values obtained by the correlation equation are presented in Table S1. The obtained values show that the experimental lipophilicity could be predicted using the milogP and XLOGP3 programs. The relationship between the lipophilicity and structure of the tested compounds 318 was analysed using the similarity analysis (Figure 4).
The cluster analysis shows that the compounds are localized in two main clusters. The derivatives localized in the first cluster have one or more hydroxyl groups as the substituent connected with the triazole ring. The second cluster is divided into three subclusters. In the first and third subcluster, mono- (36) and bistriazoles (1114) are localized which contain a substituent with a phenyl ring. The second subcluster contains derivatives 10 and 18 which contain an ethylacetyle substituent at the triazole ring. The measure of similarity of two compounds is the Euclidean distance (ED). If the ED distance is similar, it means that the compounds exhibit similar features [63,64,65,66]. As seen in Table S2, the derivatives in different subclusters have the same ED distance, which means that the compounds show high similarity. The analysis of Euclidean distance shows that lipophilicity strongly depends on the type of substituent connected with the triazole ring. Moreover, the replacement of a hydroxyl group at the C3 position of betulin by a triazole substituent slightly affects the logP parameter.
Lipophilicity is connected with other parameters which determine bioavailability. The first rules defining oral availability were described by Lipinski and Veber [67,68,69]. According to Lipinski’s rule, lipophilicity should be less than 5. Only the experimental lipophilicity of derivatives 7 and 15 meets this condition. However, the logPTLC of compounds 8, 9, and 17 is slightly greater than 5 (Table 1). Almost all derivatives meet the criteria that the number of donors (HDs) and acceptors (HAs) of the hydrogen bond are less than 5 and 10, respectively. Only compounds 1516 do not meet both rules. The molecular weight (M) of derivatives 318 is higher than 500, which means that none of tested compounds meets the mass criterion. Veber’s rules connect good oral bioavailability with topological polar surface (TPSA) and the number of rotatable bonds (RTs). Most of the tested monotriazoles 310 meet the Veber criteria, except for compounds 78 for which the TPSA value is more than 140 Å. None of the bistriazole derivatives 1118 meets the Veber criteria. In summary, monotriazole derivatives 310 show moderate oral availability, and bistriazole derivatives 1118 are characterized by low oral availability.
Multilinear regression (MLR) was used to determine the correlation equation between the experimental lipophilicity and the calculated parameters. The best correlation was obtained for the relationship between experimental lipophilicity, molecular mass (M), and topological polar surface (TPSA) (Equation (5)), for which the correlation cofactor is equal to 0.860. The lipophilicity (logPcalc) values calculated using Equation (5) are presented in Table S1.
logPcalc = 1.31M − 1.50 TPSA + 1.120
(r = 0.860; SD = 1.465; VIP = 2.13; F = 18.45),
Comparing the lipophilicity obtained by different methods shows that Equation (5) reproduces the experimental lipophilicity better than the theoretical methods.
The physicochemical parameters were used to determine the in silico pharmacokinetic properties determining the absorption of the compound [70]. The intestinal mucosa absorption was described using Caco-2 permeability (logPapp) and human intestinal absorption (HIA) [56,57]. The logPapp values calculated for mono- and bistriazoles are in the range from 0.342 to 0.808 and from 0.001 to 0.605, respectively (Table 4). The tested compounds show moderate Caco-2 permeability, because the logPaap values are lower than 0.9. Most monotriazoles (310) were characterized by better Caco-2 permeability than bistriazoles (1118). The higher HIA index shows that most derivatives could be absorbed from the gastrointestinal system into the bloodstream, except from triazole 15 which could be poorly absorbed. The tested derivatives 318 could be used as transdermal drugs because the calculated logKp for all compounds is lower than −2.5 [55]. The neurotoxicity of compounds depends on the passage through the blood–brain barrier and the penetration of the central nervous system (CNS), which was determined by logBB and logPS parameters, respectively [58]. The in silico values of logBB and logPS were determined using pkCSM software [55]. The logBB value for the tested compounds is in the range from −2.220 to −0.522, which means that triazole was poorly distributed to the brain (Table 4). Comparing the logBB parameter of mono- (310) and bistriazoles (1118) shows that the introduction of triazole substituent at C3 position of betulin scaffold reduces passage through blood–brain barrier. Most monotriazoles could penetrate the central nervous system because the logPS value is higher than −3, except for triazoles with a sugar moiety (7 and 8), which were unable to penetrate the central nervous system. In the series of bistriazoles (1118), only compounds with a phenyl substituent in the triazole moiety (1114) could penetrate the CNS. Based on the obtained results, it can be concluded that the introduction of a phenyl substituent to the triazole ring increases the neurotoxicity of the obtained compounds.
The ADME in silico-determined parameters were compared with the experimental lipophilicity (logPTLC) using similarity analysis (Figure 5).
As seen on Figure 5, compounds are localized in two main clusters. The first cluster includes bistriazoles with a sugar moiety (1516). The second cluster is divided into three subclusters which contain bistriazoles 1114, compounds 78 and 1718, and monotriazoles 36 and 910, respectively. The analysis of the Euclidean distance between subclusters shows that the type of substituent at the triazole ring influences the ADME (adsorption, distribution, metabolite, and extraction) properties, while the number of triazole rings slightly affects the bioavailability properties of the compound.

5. Conclusions

This research shows that triazole derivatives of betulin are characterized by high values of lipophilicity (logPTLC) which is in the range from 3.75 to 8.83. Mono- and bistriazoles have similar logPTLC values. The lowest values were obtained for compounds with a hydroxyl group as the triazole substituent (7, 9, 15, and 17). This research shows that lipophilicity depends on the type of substituent at the triazole moiety. The introduction of a triazole ring influences the water solubility of the compound. Comparing the hydrophobicity index of propynoyl betulin and its triazole derivatives shows that triazole compounds have better solubility. Moreover, bistriazoles have slightly better water solubility than monotriazoles. The experimental lipophilicity was compared with the lipophilicity determined using computer programs. The milogP and XLOGP3 programs best reproduce the experimental lipophilicity.
The bioavailability of compounds was determined using in silico-calculated Lipinski and Veber parameters. The analysis of parameters shows that monotriazoles 310 could be used as orally administered drugs while bistriazoles 1118 have low absorption after oral administration. The molecular weight and topological polar surface could be used to determine lipophilicity, and the obtained logPcalc better reproduces the experimental value than the theoretical logP.
The last part of the study was the determination of in silico pharmacokinetic parameters responsible for the permeability of the gastrointestinal system, skin, and blood–brain barrier. The parameters determining the absorption from the gastrointestinal system (logPapp and HIA index) show that monotriazole derivatives (310) were characterized by better Caco-2 membrane penetration than bistriazoles (1118). Triazoles could be used as transdermal drugs because the logKp is lower than −2.5. The tested compounds were poorly distributed in the central nervous system. The similarity analysis showed that the ADME parameters depend on the type of substituent in the triazole moiety. Mono- and bistriazoles have similar properties.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/app14051695/s1, Table S1: The theoretical value of lipophilicity of compounds 318. Table S2: The similarity parameter (ED) for the theoretical and experimental lipophilicity for compounds 318.

Author Contributions

Conceptualization, methodology, and writing—original draft preparation, M.K.-T.; methodology, A.S. and K.S.; validation and formal analysis, E.B. and E.C.; writing—review and editing, P.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Medical University of Silesia, grant number BNW-1-013/N/3/F.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article/supplementary material, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Rutkowska, E.; Pajak, K.; Jóźwiak, K. Lipophilicity-methods of determination and its role in medicinal chemistry. Acta Pol. Pharm. 2013, 70, 3–18. [Google Scholar] [PubMed]
  2. Arnott, J.A.; Planey, L.S. The influence of lipophilicity in drug discovery and design. Expert Opin. Drug Discov. 2012, 7, 863–875. [Google Scholar] [CrossRef] [PubMed]
  3. Ginex, T.; Vazquez, J.; Gilbert, E.; Herrero, E.; Luque, J.F. Lipophilicity in drug design: An overview of lipophilicity descriptors in 3D-QSAR studies. Future Med. Chem. 2019, 11, 1177–1193. [Google Scholar] [CrossRef]
  4. Lee, M.H.; Kim, H.D.; Jang, J.Y. Delivery systems designed to enhance stability and suitability of lipophilic bioactive compounds in food processing: A review. Food Chem. 2024, 437, 137910. [Google Scholar] [CrossRef] [PubMed]
  5. Erckes, V.; Steuer, C. A story of peptides, lipophilicity and chromatography—Back and forth in time. RSC Med. Chem. 2022, 16, 676–687. [Google Scholar] [CrossRef]
  6. Pyka-Pająk, A.; Parys, W.; Dołowy, M. Comparison of the utility of RP-TLC technique and different computational methods to assess the lipophilicity of selected antiparasitic, antihypertensive, and anti-inflammatory drugs. Molecules 2019, 24, 3187. [Google Scholar] [CrossRef]
  7. Sangster, J. Octanol-Water Partition Coefficients: Fundamentals and Physical Chemistry; Wiley&Sons: Chichester, UK, 1997; pp. 57–64. [Google Scholar]
  8. Cimpan, G. Lipophilicity determination of organic substances by reversed-phase TLC. In Encyclopedia of Chromatography, 2nd ed.; Taylor & Francis Group: Boca Raton, FL, USA, 2005; pp. 999–1001. [Google Scholar]
  9. Hubicka, U.; Żuromska-Witek, B.; Komsta, Ł.; Krzek, J. Lipophilicity study of fifteen fluoroquinolones by reversed-phase thin-layer chromatography. Anal. Methods 2015, 7, 3841–3848. [Google Scholar] [CrossRef]
  10. Bhatt, N.M.; Chavada, V.D.; Sanyal, M.; Shrivastav, P.S. Influence of organic modifier and separation modes for lipophilicity assessment of drugs using thin layer chromatography indices. J. Chromatogr. A 2018, 1571, 223–230. [Google Scholar] [CrossRef]
  11. Emonts, J.; Buyel, J.F. An overview of descriptors to capture protein properties—Tools and perspectives in the context of QSAR modeling. Comput. Struct. Biotechnol. J. 2023, 21, 3234–3247. [Google Scholar] [CrossRef]
  12. Lin, X.; Li, X.; Lin, X. A review on applications of computational methods in drug screening and design. Molecules 2020, 25, 1375. [Google Scholar] [CrossRef]
  13. Kadela-Tomanek, M.; Jastrzębska, M.; Marciniec, K.; Chrobak, E.; Bębenek, E.; Boryczka, S. Lipophilicity, pharmacokinetic properties, and molecular docking study on SARS-CoV-2 target for betulin triazole derivatives with attached 1,4-quinone. Pharmaceutics 2021, 13, 781. [Google Scholar] [CrossRef]
  14. Kadela-Tomanek, M.; Jastrzębska, M.; Marciniec, K.; Chrobak, E.; Bębenek, E.; Boryczka, S. Chromatographic and computational screening of lipophilicity and pharmacokinetics of newly synthesized betulin-1,4-quinone hybrids. Processes 2021, 9, 376. [Google Scholar] [CrossRef]
  15. Morak-Młodawska, B.; Jeleń, M.; Martula, E.; Korlacki, R. Study of lipophilicity and ADME properties of 1,9-diazaphenothiazines with anticancer action. Int. J. Mol. Sci. 2023, 24, 6970. [Google Scholar] [CrossRef] [PubMed]
  16. Šegan, S.; Jevtić, I.; Tosti, T.; Penjišević, J.; Šukalović, V.; Kostić-Rajačić, S.; Milojković-Opsenica, D. Determination of lipophilicity and ionization of fentanyl and its 3-substituted analogs by reversed-phase thin-layer chromatography. J. Chromatogr. B Analyt. Technol. Biomed. Life Sci. 2022, 1211, 123481. [Google Scholar] [CrossRef] [PubMed]
  17. Ciura, K.; Kawczak, P.; Greber, K.E.; Kapica, H.; Nowakowska, J.; Bączek, T. Application of reversed-phase thin layer chromatography and QSRR modelling for prediction of protein binding of selected β-blockers. J. Pharm. Biomed. Anal. 2019, 176, 112767. [Google Scholar] [CrossRef]
  18. Deep, A.; Kumar, D.; Bansal, N.; Narasimhan, B.; Marwaha, R.K.; Sharma, P.C. Understanding mechanistic aspects and therapeutic potential of natural substances as anticancer agents. Phytomed. Plus 2023, 3, 100418. [Google Scholar] [CrossRef]
  19. Mazumder, K.; Aktar, A.; Roy, P.; Biswas, B.; Hossain, M.E.; Sarkar, K.K.; Bachar, S.C.; Ahmed, F.; Monjur-Al-Hossain, A.S.M.; Fukase, K. A review on mechanistic insight of plant derived anticancer bioactive phytocompounds and their structure activity relationship. Molecules 2022, 27, 3036. [Google Scholar] [CrossRef]
  20. Lo Faro, F.A.; Di Trana, A.; La Maida, N.A.; Tagliabracci, R.; Giorgetti, F.; Busardò, P. Biomedical analysis of new psychoactive substances (NPS) of natural origin. J. Pharm. Biomed. Anal. 2020, 179, 112945. [Google Scholar] [CrossRef]
  21. Mancuso, C.; Siciliano, R.; Barone, E.; Preziosi, P. Natural substances and Alzheimer’s disease: From preclinical studies to evidence based medicine. Biochim. Biophys. Acta 2012, 1822, 616–624. [Google Scholar] [CrossRef]
  22. Zhao, J.; Li, R.; Pawlak, A.; Henklewska, M.; Sysak, A.; Wen, L.; Yi, J.E.; Obmińska-Mrukowicz, B. Antitumor activity of betulinic acid and betulin in canine cancer cell lines. In Vivo 2018, 32, 1081–1088. [Google Scholar] [CrossRef]
  23. Schwiebs, A.; Radeke, H.H. Immunopharmacological activity of betulin in inflammation-associated carcinogenesis. Anticancer Agents Med. Chem. 2018, 18, 645–651. [Google Scholar] [CrossRef]
  24. Buko, V.; Zavodnik, I.; Palecz, B.; Stepniak, A.; Kirko, S.; Shlyahtun, A.; Misiuk, W.; Belonovskaya, E.; Lukivskaya, O.; Naruta, E.; et al. Betulin/2-hydroxypropyl-β-cyclodextrin inclusion complex: Physicochemical characterization and hepatoprotective activity. J. Mol. Liq. 2020, 309, 113118. [Google Scholar] [CrossRef]
  25. Chaniad, P.; Sudsai, T.; Septama, A.W.; Chukaew, A.; Tewtrakul, S. Evaluation of anti-HIV-1 integrase and anti-inflammatory activities of compounds from betula alnoides buch-ham. Adv. Pharmacol. Sci. 2019, 2019, 2573965. [Google Scholar] [CrossRef] [PubMed]
  26. Javed, S.; Mahmood, Z.; Khan, K.M.; Sarker, S.D.; Javaid, A.; Khan, I.H.; Shoaib, A. Lupeol acetate as a potent antifungal compound against opportunistic human and phytopathogenic mold Macrophomina phaseolina. Sci. Rep. 2021, 11, 8417. [Google Scholar] [CrossRef]
  27. Zhang, L.; Ma, Z.; Wang, R.; Zhu, M. Synthesis and characterization of methacrylate-functionalized betulin derivatives as antibacterial comonomer for dental restorative resins. ACS Biomater. Sci. Eng. 2021, 7, 3132–3140. [Google Scholar] [CrossRef]
  28. Li, J.; Jiang, B.; Chen, C.; Fan, B.; Huang, H.; Chen, G. Biotransformation of betulin by Mucor subtilissimus to discover anti-inflammatory derivatives. Phytochemistry 2019, 166, 112076. [Google Scholar] [CrossRef]
  29. Amiri, S.; Dastghaib, S.; Ahmadi, M.; Mehrbod, P.; Khadem, F.; Behrouj, H.; Aghanoori, M.R.; Machaj, F.; Ghamsari, M.; Rosik, J.; et al. Betulin and its derivatives as novel compounds with different pharmacological effects. Biotechnol. Adv. 2020, 38, 107409. [Google Scholar] [CrossRef]
  30. Guo, W.B.; Zhang, H.; Yan, W.Q.; Liu, Y.M.; Zhou, F.; Cai, D.S.; Zhang, W.X.; Huang, X.M.; Jia, X.H.; Chen, H.S.; et al. Design, synthesis, and biological evaluation of ligustrazine—Betulin amino-acid/dipeptide derivatives as anti-tumor agents. Eur. J. Med. Chem. 2020, 185, 111839. [Google Scholar] [CrossRef]
  31. Chen, G.Q.; Guo, H.Y.; Quan, Z.S.; Shen, Q.K.; Li, X.; Luan, T. Natural products-pyrazine hybrids: A review of developments in medicinal chemistry. Molecules 2023, 28, 7440. [Google Scholar] [CrossRef]
  32. Csuk, R.; Deigner, H.P. The potential of click reactions for the synthesis of bioactive triterpenes. Bioorg. Med. Chem. Lett. 2019, 29, 949–958. [Google Scholar] [CrossRef]
  33. Dudchak, R.; Podolak, M.; Holota, S.; Szewczyk-Roszczenko, O.; Roszczenko, P.; Bielawska, A.; Lesyk, R.; Bielawski, K. Click chemistry in the synthesis of antibody-drug conjugates. Bioorg. Chem. 2024, 143, 106982. [Google Scholar] [CrossRef]
  34. Huisgen, R. 1,3-Dipolar cycloaddition. Chemistry 1984, 1, 1. [Google Scholar]
  35. Salma, U.; Ahmad, S.; Alam, M.Z.; Khan, S.A. A review: Synthetic approaches and biological applications of triazole derivatives. J. Mol. Struct. 2024, 1301, 137240. [Google Scholar] [CrossRef]
  36. Hashemi, S.M.; Badali, H.; Irannejad, H.; Shokrzadeh, M.; Emami, S. Synthesis and biological evaluation of fluconazole analogs with triazole-modified scaffold as potent antifungal agents. Bioorg. Med. Chem. 2015, 23, 1481–1491. [Google Scholar] [CrossRef]
  37. Zhou, C.H.; Wang, Y. Recent researches in triazole compounds as medicinal drugs. Curr. Med. Chem. 2012, 19, 239–280. [Google Scholar] [CrossRef]
  38. Borden, K.L.; Culjkovic-Kraljacic, B. Ribavirin as an anti-cancer therapy: Acute myeloid leukemia and beyond? Leuk. Lymphoma 2010, 51, 1805–1815. [Google Scholar] [CrossRef]
  39. Denis, J.; Ledoux, M.P.; Nivoix, Y.; Herbrecht, R. Isavuconazole: A new broad-spectrum azole. Part 1: In vitro activity. J. Mycol. Med. 2018, 28, 8–14. [Google Scholar] [CrossRef] [PubMed]
  40. Ramos, P.; Pepliński, P.; Pilawa, B. Impact of UVC radiation on interaction of selected antifungal drugs (azole derivatives) with model DPPH free radical. Ann. Acad. Med. Siles 2020, 74, 77–90. [Google Scholar] [CrossRef]
  41. Slaihim, M.M.; Al-Suede, F.S.R.; Khairuddean, M.; Ahamed, M.B.K.; Majid, A.M.S.A. Synthesis, characterisation of new derivatives with mono ring system of 1,2,4-triazole scaffold and their anticancer activities. J. Mol. Struct. 2019, 1196, 78–87. [Google Scholar] [CrossRef]
  42. Roy, K.K. Targeting the active sites of malarial proteases for antimalarial drug discovery: Approaches, progress and challenges. Int. J. Antimicrob. Agents 2017, 50, 287–302. [Google Scholar] [CrossRef] [PubMed]
  43. Kaproń, B.; Czarnomysy, R.; Wysokiński, M.; Andrys, R.; Musilek, K.; Angeli, A.; Supuran, C.T.; Plech, T. 1,2,4-Triazole-based anticonvulsant agents with additional ROS scavenging activity are effective in a model of pharmacoresistant epilepsy. J. Enzyme Inhib. Med. Chem. 2020, 35, 993–1002. [Google Scholar] [CrossRef]
  44. Bozorov, K.; Zhao, J.; Aisa, H.A. 1,2,3-Triazole-containing hybrids as leads in medicinal chemistry: A recent overview. Bioorg. Med. Chem. 2019, 27, 3511–3531. [Google Scholar] [CrossRef] [PubMed]
  45. Guo, H.Y.; Chen, Z.A.; Shen, Q.K.; Quan, Z.S. Application of triazoles in the structural modification of natural products. J. Enzyme Inhib. Med. Chem. 2021, 36, 1115–1144. [Google Scholar] [CrossRef]
  46. Kadela-Tomanek, M.; Bębenek, E.; Chrobak, E.; Marciniec, K.; Latocha, M.; Kuśmierz, D.; Jastrzębska, M.; Boryczka, S. Betulin-1,4-quinone hybrids: Synthesis, anticancer activity and molecular docking study with NQO1 enzyme. Eur. J. Med. Chem. 2019, 177, 302–315. [Google Scholar] [CrossRef] [PubMed]
  47. Chrobak, E.; Jastrzębska, M.; Bębenek, E.; Kadela-Tomanek, M.; Marciniec, K.; Latocha, M.; Wrzalik, R.; Kusz, J.; Boryczka, S. Molecular Structure, in vitro anticancer study and molecular docking of new phosphate derivatives of betulin. Molecules 2021, 26, 737. [Google Scholar] [CrossRef]
  48. Bębenek, E.; Jastrzębska, M.; Kadela-Tomanek, M.; Chrobak, E.; Orzechowska, B.; Zwolińska, K.; Latocha, M.; Mertas, A.; Czuba, Z.; Boryczka, S. Novel triazole hybrids of betulin: Synthesis and biological activity profile. Molecules 2017, 22, 1876. [Google Scholar] [CrossRef]
  49. Kadela-Tomanek, M.; Jastrzębska, M.; Marciniec, K.; Chrobak, E.; Bębenek, E.; Latocha, L.; Kuśmierz, D.; Boryczka, S. Design, synthesis and biological activity of 1,4-quinone moiety attached to betulin derivatives as potent DT-diaphorase substrate. Bioorg. Chem. 2021, 106, 104478. [Google Scholar] [CrossRef]
  50. Bębenek, E.; Kadela-Tomanek, M.; Chrobak, E.; Jastrzębska, M.; Książek, M. Synthesis and structural characterization of a new 1,2,3-triazole derivative of pentacyclic triterpene. Crystals 2022, 12, 422. [Google Scholar] [CrossRef]
  51. Boryczka, S.; Bębenek, E.; Wietrzyk, J.; Kempińska, K.; Jastrzębska, M.; Kusz, J.; Nowak, M. Synthesis, structure and cytotoxic activity of new acetylenic derivatives of betulin. Molecules 2013, 18, 4526–4543. [Google Scholar] [CrossRef]
  52. Virtual Computational Chemistry Laboratory. Available online: http://www.vcclab.org (accessed on 10 January 2024).
  53. Molinspiration. Available online: https://www.molinspiration.com/services/logp.html (accessed on 10 January 2024).
  54. SwissADME. Available online: http://www.swissadme.ch/index.php (accessed on 10 January 2024).
  55. Biosig. Available online: https://biosig.lab.uq.edu.au/pkcsm/prediction (accessed on 10 January 2024).
  56. Kulkarni, A.; Han, Y.; Hopfinger, A. Predicting Caco-2 cell permeation coefficients of organic molecules using membrane-interaction QSAR analysis. J. Chem. Inf. Comput. Sci. 2002, 42, 331–342. [Google Scholar] [CrossRef]
  57. Feher, M.; Schmidt, J. Property distributions: Differences between drugs, natural products, and molecules from combinatorial chemistry. J. Chem. Inf. Comput Sci. 2003, 43, 218–227. [Google Scholar] [CrossRef] [PubMed]
  58. Hutter, M. Prediction of blood-brain barrier permeation using quantum chemically derived information. J. Comput. Aided Mol. Des. 2003, 17, 415–433. [Google Scholar] [CrossRef]
  59. Rosés, M.; Bosch, E.; Ràfols, C.; Fuguet, E. Chromatographic hydrophobicity index (CHI). Adv. Chromatogr. 2012, 50, 377–414. [Google Scholar] [PubMed]
  60. Mannhold, R.; Poda, G.I.; Ostermann, C.; Tetko, I.V. Calculation of molecular lipophilicity: State-of-the-art and comparison of log P methods on more than 96,000 compounds. J. Pharm. Sci. 2009, 98, 861–893. [Google Scholar] [CrossRef]
  61. Lin, S.Y.; Yang, Y.H.; Liu, H.H. Development of liposome/water partition coefficients predictive models for neutral and ionogenic organic chemicals. Ecotoxicol. Environ. Saf. 2019, 179, 40–49. [Google Scholar] [CrossRef]
  62. Endo, S.; Escher, B.I.; Goss, K.U. Capacities of membrane lipids to accumulate neutral organic chemicals. Environ. Sci. Technol. 2011, 45, 5912–5921. [Google Scholar] [CrossRef]
  63. Anderberg, M.R. Cluster Analysis for Applications; Elsevier: Amsterdam, The Netherlands, 1973. [Google Scholar]
  64. Santini, S.; Jain, R. Similarity measures. IEEE Trans. Pattern Anal. Mach. Intell. 1999, 21, 871–883. [Google Scholar] [CrossRef]
  65. Liberti, L.; Lavor, C.; Maculan, N.; Mucherino, A. Euclidean distance geometry and applications. SIAM Rev. 2014, 56, 1–62. [Google Scholar] [CrossRef]
  66. Kadela-Tomanek, M.; Jastrzębska, M.; Chrobak, E.; Bębenek, E. Lipophilicity and ADMET analysis of quinoline-1,4-quinone hybrids. Pharmaceutics 2022, 15, 34. [Google Scholar] [CrossRef]
  67. Walters, W.P. Going further than Lipinski’s rule in drug design. Expert Opin. Drug Discov. 2012, 7, 99–107. [Google Scholar] [CrossRef]
  68. Jangam, C.S.; Bhowmick, S.; Chorge, R.D.; Bharatrao, L.D.; Patil, P.C.; Chikhale, R.V.; AlFaris, N.A.; ALTamimi, J.Z.; Wabaidur, S.M.; Islam, M.A. Pharmacoinformatics-based identification of anti-bacterial catalase-peroxidase enzyme inhibitors. Comput. Biol. Chem. 2019, 83, 107136. [Google Scholar] [CrossRef] [PubMed]
  69. Pluskota, R.; Jaroch, K.; Kośliński, P.; Ziomkowska, B.; Lewińska, A.; Kruszewski, S.; Bojko, B.; Koba, M. Selected drug-likeness properties of 2-arylidene-indan-1,3-dione derivatives-chemical compounds with potential anti-cancer activity. Molecules 2021, 26, 256. [Google Scholar] [CrossRef] [PubMed]
  70. Johnson, T.; Dress, K.; Edwards, M. Using the Golden Triangle to optimize clearance and oral absorption. Bioorg. Med. Chem. Lett. 2009, 19, 5560–5564. [Google Scholar] [CrossRef]
Figure 1. Chemical structure of betulin.
Figure 1. Chemical structure of betulin.
Applsci 14 01695 g001
Figure 2. Molecular structure of compounds 118.
Figure 2. Molecular structure of compounds 118.
Applsci 14 01695 g002aApplsci 14 01695 g002b
Figure 3. The profile of changes in theoretical lipophilicity for derivatives 318.
Figure 3. The profile of changes in theoretical lipophilicity for derivatives 318.
Applsci 14 01695 g003
Figure 4. The cluster analysis of the experimental and theoretical lipophilicity for compounds 318.
Figure 4. The cluster analysis of the experimental and theoretical lipophilicity for compounds 318.
Applsci 14 01695 g004
Figure 5. The similarity analyses of ADME parameters and experimental lipophilicity for compounds 318.
Figure 5. The similarity analyses of ADME parameters and experimental lipophilicity for compounds 318.
Applsci 14 01695 g005
Table 1. The experimental values of lipophilicity and hydrophobicity parameters.
Table 1. The experimental values of lipophilicity and hydrophobicity parameters.
CompoundRM0brlogPTLCφ0CompoundRM0brlogPTLCφ0
15.78−0.060.9956.6890.1728.09−0.090.9929.2090.09
36.15−0.070.9977.0986.86117.05−0.080.9928.0786.29
46.31−0.070.9947.2686.68127.15−0.080.9938.1786.04
56.29−0.070.9937.2485.12136.57−0.080.9957.5483.38
66.85−0.080.9887.8587.15147.75−0.090.9948.8387.18
73.67−0.050.9884.3973.11153.08−0.050.9503.7559.69
84.56−0.060.9925.3678.76165.68−0.080.9986.5870.65
94.44−0.050.9945.2381.92174.35−0.060.9905.1375.65
105.57−0.070.9946.4685.04186.59−0.080.9877.5782.89
Table 2. The LogPlit and LogPTLC values for reference compounds AF.
Table 2. The LogPlit and LogPTLC values for reference compounds AF.
CompoundLogPlitRM0brLogPTLC
A2.431.87−0.020.9852.43
B3.182.42−0.020.9813.03
C4.454.08−0.030.9924.84
D4.794.11−0.050.9704.87
E6.014.87−0.040.9725.69
F6.385.50−0.060.9856.38
Table 3. The calculated lipophilicity of compounds 318.
Table 3. The calculated lipophilicity of compounds 318.
CompoundALOGPsAClogPmiLogPXLOGP2XLOGP3
36.606.978.709.8710.21
46.607.028.7710.3010.31
56.426.788.569.599.93
66.679.118.9010.2910.85
74.113.515.086.116.39
85.014.216.586.717.72
95.385.267.057.818.22
105.775.297.658.299.00
117.398.099.3511.9312.14
127.458.219.4312.2512.34
137.407.729.2111.3711.58
147.568.389.5612.7713.42
152.871.173.004.404.49
164.692.596.005.617.17
175.194.696.937.818.16
186.104.748.128.769.73
Table 4. The ADME parameters of compounds 318.
Table 4. The ADME parameters of compounds 318.
CompoundMTPSAHAHDRTlogPappHIAlogKplogBBlogPS
3627.9077.245170.737100−2.730−0.522−1.288
4648.8977.246170.808100−2.732−0.741−2.240
5652.91101.036170.485100−2.733−0.689−1.355
6659.95102.545180.66099−2.731−0.699−1.251
7699.92167.3951070.34260−2.735−1.853−3.671
8761.00152.879380.25099−2.736−1.314−3.001
9595.8697.476280.627100−2.780−0.714−2.606
10623.87103.547190.658100−2.762−0.928−2.677
11813.08114.0280120.528100−2.735−1.406−2.289
12849.06114.02100120.484100−2.735−1.845−2.536
13863.10161.60100120.001100−2.735−1.742−2.402
14877.21164.6280140.529100−2.735−1.764−2.193
15957.13294.34208100.09711−2.735−3.534−5.467
161079.31252.65181240.60588−2.735−2.454−4.070
17748.99154.48102140.071100−2.736−1.815−3.280
18805.01166.62120160.08991−2.735−2.220−3.420
Table 5. Correlation equation for the experimental (logPTLC) and theoretical (logPcalc) values of lipophilicity for derivatives 318.
Table 5. Correlation equation for the experimental (logPTLC) and theoretical (logPcalc) values of lipophilicity for derivatives 318.
ProgramCorrelation EquationrSD
ALOGPslogPTLC = 0.928 LogPcalc + 0.5800.9280.670
ACLOGPlogPTLC = 0.841 LogPcalc + 3.9070.8400.827
milogPlogPTLC = 0.916 LogPcalc + 0.9790.9160.682
XLOGP2logPTLC = 0.889 LogPcalc + 1.9260.8890.675
XLOGP3logPTLC = 0.937 LogPcalc + 1.1570.9370.565
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kadela-Tomanek, M.; Sokal, A.; Stocerz, K.; Bębenek, E.; Chrobak, E.; Olczyk, P. Assessment of Bioavailability Parameters of Mono- and Bistriazole Derivatives of Propynoylbetulin. Appl. Sci. 2024, 14, 1695. https://doi.org/10.3390/app14051695

AMA Style

Kadela-Tomanek M, Sokal A, Stocerz K, Bębenek E, Chrobak E, Olczyk P. Assessment of Bioavailability Parameters of Mono- and Bistriazole Derivatives of Propynoylbetulin. Applied Sciences. 2024; 14(5):1695. https://doi.org/10.3390/app14051695

Chicago/Turabian Style

Kadela-Tomanek, Monika, Arkadiusz Sokal, Klaudia Stocerz, Ewa Bębenek, Elwira Chrobak, and Paweł Olczyk. 2024. "Assessment of Bioavailability Parameters of Mono- and Bistriazole Derivatives of Propynoylbetulin" Applied Sciences 14, no. 5: 1695. https://doi.org/10.3390/app14051695

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop