Next Article in Journal
Fatty Acids as a Tool to Understand Microbial Diversity and Their Role in Food Webs of Mediterranean Temporary Ponds
Next Article in Special Issue
Leishmanicidal Evaluation of Tetrahydroprotoberberine and Spirocyclic Erythrina-Alkaloids
Previous Article in Journal
Pharmacokinetics, Tissue Distribution, Excretion and Plasma Protein Binding Studies of Wogonin in Rats
Previous Article in Special Issue
Two Trypanocidal Dipeptides from the Roots of Zapoteca portoricensis (Fabaceae)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In-Silico Analyses of Sesquiterpene-Related Compounds on Selected Leishmania Enzyme-Based Targets

by
Freddy A. Bernal
and
Ericsson Coy-Barrera
*
Laboratorio de Química Bioorgánica, Departamento de Química, Facultad de Ciencias Básicas y Aplicadas, Universidad Militar Nueva Granada, Cundinamarca 250240, AA 49300, Colombia
*
Author to whom correspondence should be addressed.
Molecules 2014, 19(5), 5550-5569; https://doi.org/10.3390/molecules19055550
Submission received: 17 March 2014 / Revised: 14 April 2014 / Accepted: 22 April 2014 / Published: 29 April 2014

Abstract

:
A great number of sesquiterpenes are reported in the available literature as good antileishmanial leads. However, their mode of action at the molecular level has not been elucidated. The lack of molecular studies could be considered an impediment for studies seeking to improve sesquiterpene-based drug design. The present in silico study allows us to make important observations about the molecular details of the binding modes of a set of antileishmanial sesquiterpenes against four drug-enzyme targets [pteridine reductase-1 (PTR1), N-myristoyl transferase (NMT), cysteine synthase (CS), trypanothione synthetase (TryS)]. Through molecular docking it was found that two sesquiterpene coumarins are promising leads for the PTR1 and TryS inhibition purposes, and some xanthanolides also exhibited better affinity towards PTR1 and CS binding. In addition, the affinity values were clustered by Principal Component Analysis and drug-like properties were analyzed for the strongest-docking sesquiterpenes. The results are an excellent starting point for future studies of structural optimization of this kind of compounds.

Graphical Abstract

1. Introduction

Parasites of the genus Leishmania are trypanosomatid protozoa that cause the neglected disease known as leishmaniasis, which infects some 15 million people around the world in three clinical forms: cutaneous, mucocutaneous and visceral [1,2]. Its emergence as an opportunistic pathogen has generated a public health interest and an ongoing need to control it. However, its treatment still remains based on the use of pentavalent antimony salts as first-line drugs, or the use of amphotericin B and pentamidine as second-line drugs, which are often toxic, some have an unknown mode of action, and are usually marginally effective, plus the added problem that there are outbreaks of resistance [3,4]. So far, there are not enough advances in the replacement of these drugs, with some cases of uncertainly effective therapy, although there are some clinical trials based on known drugs such as allopurinol (a protein synthesis inhibitor), AmBisome® (a formulation of amphotericin B in liposomes) and ketoconazole (an inhibitor of sterol synthesis) [5].
Several metabolic pathways are currently under study in order to identify critical enzymes for inhibition purposes with the aim of exploiting them for the treatment of parasitic diseases. Some of the most interesting enzymes are those involved in metabolism of glucose, sterols, fatty acids and nucleotides and those key enzymes for protein biosynthesis and for the maintenance of trypanothione and polyamine levels [6].
A highly valuable reaction to protect and maintain vitally important intracellular amounts of tetrahydropterin—which has been proven as an essential part of the growth—is the reduction of conjugated and non-conjugated pterins, such as reduced biopterin to dihydrobiopterin, which is catalyzed by pteridine reductase (PTR1) [7,8]. Because PTR1 is less sensitive to the effect of methotrexate, but also catalyzes the reduction of folates, it explains the therapeutic failure of these drugs against trypanosomatid parasites [7].
It is well-known that Leishmania requires cysteine for protein biosynthesis and as a precursor of trypanothione, with an essential role in redox metabolism and antioxidant defense [9]. A de novo route to produce cysteine, selectively present in some microorganisms (but absent in mammals), is catalyzed by cysteine synthase (CS), in a two-stage process from serine [10]. In addition, the bifunctional trypanothione synthetase-amidase catalyzes biosynthesis and hydrolysis of the trypanothione, a glutathione-spermidine conjugate, which is critical for intracellular thiol-redox unique to trypanosomatids. The synthetase C-terminal domain displays the ATP-grasp fold, which binds nucleotide in a well-defined fashion [11].
Some studies have identified N-myristoyltransferase (NMT) as an appropriate drug target against protozoan parasitic diseases. NMT is ubiquitous in eukaryotic cells catalyzing the addition of the myristate to the N-terminal glycine residue of some proteins [12]. Protein N-myristoylation is important for targeting proteins to membrane sites, mediating protein–protein interactions and stabilizing protein structures [13].
Several sesquiterpene-related compounds isolated from natural sources have shown in vitro effectivity against promastigotes and amastigotes of Leishmania and in vivo activity [14,15]. However, their mode of action has not been elucidated. Recently, a work reported an in silico study through molecular docking of a large number of terpenoids within the active site of 29 enzymes from L. major, L. donovani, L. mexicana and L. infantum [16]. In this study, ca. 100 different sesquiterpenoids (mainly sesquiterpene lactones) were also docked, with the results indicating target selectivity, e.g., germacranolide sesquiterpenoids exhibited selectivity to L. major methionyl t-RNA synthetase (LmMRS) and dihydroorotate dehydrogenase (LmDHODH). Among the targeted enzymes, LmPTR1 and LdNMT were also docked with test sesquiterpenes with reasonable affinity. However, a number of bioactive sesquiterpenoids were not included in that study and LmCS and LmTryS enzymes are not targeted. Therefore, as part of our interest in finding new prototypes against leishmaniasis, in this work a study through molecular docking and residual interactions of the molecular basis of action at the in silico level of 123 sesquiterpene-related compounds (possessing antiparasitic activity) within the active site of LmPTR1, LdNMT, LmCS, and LmTryS, is described. These analyses showed very important details at the molecular level that should prove useful for the development of therapeutics from sesquiterpene-based prototypes.

2. Results and Discussion

2.1. Molecular Docking of Sesquiterpenes

Structural information about the test sesquiterpenoids were acquired from different published works reporting antileishmanial activity of natural-occurring sesquiterpenes. The IUPAC and common names can be found in the Supplementary Material. The compounds were further subdivided by common moieties such as sesquiterpene-coumarins, agarofurans, drimanes, pseudoguaianolides, guaianolides, xanthanolides, germacranolides, eudesmanolides and miscellaneous, whose structures are shown in Figure 1, Figure 2, Figure 3, Figure 4, Figure 5, Figure 6, Figure 7, Figure 8 and Figure 9. Each ligand structure was separately submitted to a conformational search using MMFF and the geometry of the most stable conformer was then optimized by semi-empirical AM1 parametrization. The most stable conformer of each optimized structure was imported into Autodock/Vina to be docked into the active site of LmPTR1, LdNMT, LmCS and LmTryS enzymes. The structural data of the enzymes were imported from the Protein Data Bank (PDB) (Table 1).
Table 1. Data of the test enzymes.
Table 1. Data of the test enzymes.
EnzymePDB ID aResolution [Å]SourceAbbreviationReference b
Pteridine reductase2QHX2.61L. majorLmPTR1[17]
Cysteine synthase4AIR1.80L. majorLdNMT[10]
N-Myristoyl transferase2WUU1.42L. donovaniLmCS[13]
Trypanothione synthetase2VOB2.30L. majorLmTryS[11]
a Assigned code in the Protein Data Bank (PDB); b Reference including the active site.
Figure 1. Structures of the tested sesquiterpene coumarins [18].
Figure 1. Structures of the tested sesquiterpene coumarins [18].
Molecules 19 05550 g001
Figure 2. Structures of the tested agarofuran sesquiterpenes [19,20,21,22,23].
Figure 2. Structures of the tested agarofuran sesquiterpenes [19,20,21,22,23].
Molecules 19 05550 g002
Figure 3. Structures of the tested drimane sesquiterpenes [24].
Figure 3. Structures of the tested drimane sesquiterpenes [24].
Molecules 19 05550 g003
Figure 4. Structures of the tested pseudoguaianolide sesquiterpenes [25,26,27].
Figure 4. Structures of the tested pseudoguaianolide sesquiterpenes [25,26,27].
Molecules 19 05550 g004
Figure 5. Structures of the tested guaianolide sesquiterpenes [26,28,29].
Figure 5. Structures of the tested guaianolide sesquiterpenes [26,28,29].
Molecules 19 05550 g005
Figure 6. Structures of the tested xanthanolide sesquiterpenes [27].
Figure 6. Structures of the tested xanthanolide sesquiterpenes [27].
Molecules 19 05550 g006
Figure 7. Structures of the tested germacranolide sesquiterpenes [27,30,31,32].
Figure 7. Structures of the tested germacranolide sesquiterpenes [27,30,31,32].
Molecules 19 05550 g007
Figure 8. Structures of the tested eudesmanolide sesquiterpenes [27].
Figure 8. Structures of the tested eudesmanolide sesquiterpenes [27].
Molecules 19 05550 g008
Figure 9. Structures of the tested miscellaneous sesquiterpenes [24,33,34,35,36,37,38,39,40,41,42].
Figure 9. Structures of the tested miscellaneous sesquiterpenes [24,33,34,35,36,37,38,39,40,41,42].
Molecules 19 05550 g009
The resulting docking energies (vina scores as affinity values in kcal/mol) of sesquiterpenoids are summarized in Table 2. Szowitsiacoumarin B (2) was found to be the strongest-docking sesquiterpene coumarin. The sesquiterpene coumarins exhibited overall stronger docking energies towards the Leishmania protein targets. This fact is observed by the mean value of the docking energies for each sesquiterpene-type skeletal (Table 3). The RSD percentages were found to be in a 5.1%–11.4% range indicating dispersion between the affinities for each type of sesquiterpenes which might be correlated with enzyme selectivity. In contrast, pseudoguaianolides, guaianolides, and germacranolides revealed the highest-docking energies and lowest-RSD percentages. Agarofurans 4850, 55, 9091 also exhibited stronger docking energies, although they showed selectivity towards PTR1 and NMT.
In the case of the docking with PTR1, dimeric xanthanolide compounds 7576 exhibited the best affinity values (‒10.6 kcal/mol), with even lower values than that of the control used, DB07765 (‒10.2 kcal/mol). Their lowest-energy docked poses are shown in Figure 10. Both complexes share identical H-bond interactions with Ser111, Ser227 and Arg17, being the crucial contacts for the ligand-enzyme docking. In addition, as shown in the 2D representation of protein ligand interactions (Figure 10c), a receptor contact through Phe113 contributes to the non-polar stabilization in the hydrophobic guaianolide moiety. These two compounds were reported by Schmidt et al. [27] with reasonable antileishmanial activity against L. donovani amastigotes with IC50 values in the 22–27 µM range. Our results indicate that compounds 7576 could serve as leads for further structural optimization studies towards PTR1 inhibition.
In addition, compounds 7576 share as structural feature an α-methylene γ-lactone moiety. This moiety in sesquiterpene lactones (STLs) is considered to be capable for reacting in nucleophilic Michael additions (acting as Michael acceptor) with sulfhydryl-(-SH) or amino-(-NH2)-containing residues, such as cysteine, lysine and histidine [43,44]. Several studies have indicated that the formation of this type of Michael adducts is correlated with the biological activity of STLs [45]. Analyzing the 2D representation of protein ligand interactions (Figure 10c–d) for lowest-energy docking pose of 7576, and other test α-methylene-containing STLs (data not shown), Lys198 was found to be the only residue within the LmPTR1 binding pocket that satisfies the structural requirement for a possible Michael addition, but due to the H-bond interaction (2.1 Å) between the lactone carbonyl group and Ser222, the Lys198 residue is placed at the other face of the molecule, which eventually might hinder the formation of a Michael adduct.
Table 2. Docking energies a (kcal/mol) of sesquiterpenoids with Leishmania enzyme targets.
Table 2. Docking energies a (kcal/mol) of sesquiterpenoids with Leishmania enzyme targets.
NoLmPTR1LmCSLdNMTLmTrySNoLmPTR1LmCSLdNMTLmTrySNoLmPTR1LmCSLdNMTLmTryS
sesquiterpene coumarinsdrimanesgermacranolides cont.
1−9.9−11.2−6.6−8.945−6.9−7.6−4.5−7.086−8.0−9.0−6.2−8.3
2−10.0−11.8−6.8−8.546−8.9−9.9−5.9−7.987−8.1−9.5−5.5−7.7
3−7.8−10.0−5.6−7.447−8.6−9.9−6.1−8.388−8.1−8.7−5.2−7.5
4−7.4−9.5−5.1−7.948−8.9−10.8−5.5−7.889−8.4−9.0−5.7−8.7
5−7.5−10.0−4.9−8.2pseudoguaianolides90−8.0−9.6−5.7−7.9
6−9.5−10.2−6.4−8.449−8.6−8.5−5.1−7.991−8.2−9.0−6.3−7.4
7−8.9−10.2−6.3−8.450−7.9−8.5−5.2−7.092−7.6−8.7−5.8−7.4
8−9.2−10.2−6.0−8.151−8.1−9.6−5.6−8.693−7.7−8.9−5.8−7.3
9−8.6−10.7−6.2−8.252−8.2−10.5−5.7−7.994−8.1−8.6−5.7−7.7
agarofurans53−8.0−8.5−5.5−7.395−7.7−8.7−5.9−7.5
10−8.8−8.5−5.5−7.054−8.3−9.2−6.1−7.696−8.6−8.7−5.7−7.1
11−7.5−7.4−4.7−6.655−8.3−8.9−5.4−7.697−8.8−9.2−5.5−7.3
12−7.8−7.6−5.1−6.556−8.2−9.7−5.7−8.098−8.3−9.0−5.1−7.2
13−7.4−8.9−5.1−3.257−7.8−9.1−5.7−7.3eudesmanolides
14−7.2−9.0−5.2−3.158−8.1−8.5−5.9−7.599−7.8−8.4−5.1−7.1
15−8.2−8.1−5.0−5.659−8.3−9.4−5.9−8.0100−7.8−8.4−5.1−7.1
16−8.3−9.8−4.9−5.360−8.0−8.7−6.4−8.1101−8.1−8.1−5.5−7.8
17−7.9−8.7−5.6−5.661−8.2−9.1−6.2−8.0102−7.9−8.1−5.9−6.9
18−8.3−7.7−5.3−6.062−7.9−8.5−5.3−7.4103−7.9−8.4−5.2−6.7
19−9.2−8.3−5.8−6.863−8.3−10.5−6.2−8.3104−8.4−9.0−5.2−7.1
20−9.2−10.4−6.7−6.164−8.9−9.2−5.7−8.6105−8.0−9.4−5.7−7.5
21−9.2−10.7−6.7−5.965−8.0−8.8−5.8−7.9106−7.8−8.8−5.6−7.1
22−9.0−10.2−6.9−6.566−8.8−8.7−6.0−7.7miscellaneous
23−8.7−10.4−5.8−5.667−8.3−9.0−5.4−7.4107−8.1−9.9−5.2−7.6
24−9.0−11.9−5.6−7.4guaianolides108−7.8−9.6−5.8−8.0
25−8.7−10.8−6.7−5.368−8.1−9.8−5.5−7.4109−7.3−9.5−5.0−7.9
26−8.9−9.4−5.9−8.569−7.8−8.9−5.8−7.8110−6.5−8.5−3.8−6.9
27−8.6−10.1−6.9−6.070−8.0−9.7−5.8−7.0111−8.8−9.4−5.3−7.8
28−9.4−10.8−6.5−8.5xanthanolides112−7.4−7.9−5.2−7.0
29−8.6−11.4−5.5−7.771−7.2−9.4−5.4−6.8113−6.9−8.8−5.2−6.7
30−8.9−12.1−5.8−7.072−7.6−8.5−5.8−7.2114−7.1−9.1−5.0−7.1
31−8.8−11.4−5.9−8.573−7.8−8.9−5.2−7.4115−7.0−9.6−5.0−6.7
32−9.3−10.2−5.8−8.774−7.5−7.9−5.1−7.1116−7.7−8.1−4.7−6.9
33−8.6−9.8−6.0−7.675−10.6−11.4−6.5−0.5117−6.5−8.5−4.7−6.7
34−8.9−11.2−6.1−8.776−10.6−9.4−6.4−3.3118−8.7−9.2−5.7−7.6
35−9.3−10.6−6.7−7.077−8.0−9.5−4.9−7.3119−7.8−8.4−5.6−7.6
36−9.2−9.5−6.1−7.278−7.5−8.6−5.5−7.4120−8.0−9.0−5.4−7.2
37−8.8−9.5−6.0−8.4germacranolides121−8.1−7.5−4.9−7.1
38−8.9−10.0−6.7−6.679−7.7−8.6−5.8−7.4122−8.3−9.2−5.8−7.7
39−9.5−9.4−6.9−6.180−8.1−9.3−5.6−8.5123−7.9−8.6−5.2−7.7
40−9.3−10.4−6.8−5.981−8.5−9.1−6.4−6.7known inhibitors b
41−7.7−7.0−4.7−6.682−8.3−9.7−5.5−7.8I1−10.2
42−6.2−6.3−5.5−7.083−8.0−10.6−5.7−8.3I2−9.6
43−7.3−7.1−6.1−5.384−8.4−8.9−5.3−8.2I3−8.0
44−7.8−6.8−5.1−5.685−8.3−8.8−5.8−8.1I4−8.2
a Vina scores (affinity values); b Reported inhibitors against each enzyme: I1 = DB07765; I2 = ZINC01690699; I3 = DDD64558; I4 = DDD66604.
Table 3. Mean values of docking energies (kcal/mol) and standard deviation for each skeletal type of sesquiterpenes with Leishmania enzyme targets.
Table 3. Mean values of docking energies (kcal/mol) and standard deviation for each skeletal type of sesquiterpenes with Leishmania enzyme targets.
Sesquiterpene−type skeletalLmPTR1LmCSLdNMTLmTryS
MDE a%RSD bMDE a%RSD bMDE a%RSD bMDE a%RSD b
sesquiterpene coumarins−8.811.4−10.46.7−6.011.0−8.25.1
agarofurans−8.59.0−9.516.1−5.911.5−6.620.6
drimanes−8.311.5−9.614.3−5.512.9−7.87.0
pseudoguaianolides−8.23.5−9.16.7−5.76.2−7.85.9
guaianolides−8.01.9−9.55.2−5.73.0−7.45.4
xanthanolides−8.416.9−9.211.4−5.610.5−5.943.8
germacranolides−8.13.9−9.15.3−5.75.8−7.76.7
eudesmanolides−8.02.6−8.65.3−5.45.6−7.24.8
miscellaneous−7.69.0−8.97.5−5.19.5−7.36.2
a Mean values of docking energies (kcal/mol) for each skeletal type of sesquiterpenes; b Relative Standard Deviation (RSD) percentage of docking energy values for each skeletal type of sesquiterpenes.
On the other hand, the bicyclic sesquiterpene coumarins, szowitsiacoumarins A (1) and B (2), also exhibited good affinity with PTR1 (‒9.9 and ‒10.0 kcal/mol, respectively) very close to that of the control. The lowest-energy docked pose of compound 1 is shown in Figure 11. An identical H-bond interaction with Ser111 was revealed by the PTR1···1 and PTR1···2 complexes, which seems to be a key interaction for the complexes stabilization. Another polar interaction with Lys198 could also stabilize the complex as well as a receptor contact through Phe113 contributes to stabilization via the hydrophobic sesquiterpene centroid, which is located between three leucine (Leu226, Leu229, and Leu188) and one serine (Ser227) residues. These two compounds were reported by Iranshahi et al. [18]. However, their antiparasitic activity was not evaluated while compounds 4-5 were reported to be active against L. major promastigotes with IC50 values in the 13–17 µM range, but the PTR1 affinity values were not significant.
Regarding docking with LmCS, compounds 2, 30, and 75 exhibited the lowest docking energies (−11.8, −12.1, and ‒11.4 kcal/mol, respectively), possessing significant lower affinity values than that of control used in this study, ZINC01690699 (−9.6 kcal/mol). The agarofuran 30 was the most affine ligand for LmCs. The most-stable pose of 30 is shown in Figure 12. Two H-bond-type interactions are present in the CS···30 complex when the ligand interacts with Gly186 and Ser274 through cinnamoyl and benzoyl oxygens at C-9 and C-8 carbons (2.4 and 2.1 Å, respectively). A structural requirement was observed on comparing the affinity values of isomers 29 and 30. These compounds differ solely by the geometrical isomerism of the cinnamoyl double bond (Z-29 and E-30). However, the docking energy of 30 was lower than that of 29 (−12.4 vs. ‒11.4 kcal/mol). The cynnamoyl aromatic ring is apparently necessary for a hydrophobic contact with Gly184 and Thr190 in 30. The presence of the two cinnamoyl and benzoyl ester groups in 30 seems to be an essential requirement for the docking, since the best pose has these substitutions oriented towards the binding pocket, formed by Val50, Pro301, Phe273 and Ser274.
Figure 10. Lowest-energy docked pose with LmPTR1 of: (a) sesquiterpene 75; (b) sesquiterpene 76; 2D representation of protein ligand interactions of: (c) compound 75; (d) compound 76.
Figure 10. Lowest-energy docked pose with LmPTR1 of: (a) sesquiterpene 75; (b) sesquiterpene 76; 2D representation of protein ligand interactions of: (c) compound 75; (d) compound 76.
Molecules 19 05550 g010
Interactions are represented as: van der Waals (green circle); polar (magenta circle); H-bond with amino acid side chains (blue dashed line); H-bond with amino acid backbone (green dashed line); pi interaction (orange line).
Agarofurans are considered as good therapies for Multidrug Resistant Drug (MDR)-agents [20,21]. The fact of 16 out of 35 agarofurans exhibited affinity values lower than −10 kcal/mol (with compound 30 as the strongest-docking sesquiterpene toward CS enzyme) indicates a rational starting point for the inhibition of the de novo route to produce cysteine in trypanosomatids.
Figure 11. (a) Lowest-energy docked pose of sesquiterpene coumarin, szowitsiacoumarin A (1), with LmPTR1; (b) 2D representation of protein ligand interactions of compound 1 with LmPTR1.
Figure 11. (a) Lowest-energy docked pose of sesquiterpene coumarin, szowitsiacoumarin A (1), with LmPTR1; (b) 2D representation of protein ligand interactions of compound 1 with LmPTR1.
Molecules 19 05550 g011
Interactions are represented as: van der Waals (green circle); polar (magenta circle); H-bond with amino acid side chains (blue dashed line); H-bond with amino acid backbone (green dashed line); pi interaction (orange line).
In addition, the germacranolide neurolenin A, 83, also showed a better affinity value (−10.6 kcal/mol) than that of control for docking with LmCS. Compounds 81 and 83 (which differ by the position of the isovaleoryl group attached at C8 and C9 carbons) exhibited affinities which were significantly different. On comparing the ligand interactions of these germacranolides, it is evidenced that the location of this ester group, although it is not determinant in the docking, is important since in the germacrene moiety in 83 allows suitable H-bond interactions with Gly186, Ser274 and Lys51 (ca. 2 Å) (Figure 13), while the lowest-energy pose of 81 is oriented in such a way that H-bond interactions with Gly186, Ser274 and Asn278 are presented. Seemingly, the hydrophobic proximity with Phe273, Asn278 and Glu211 contributes to the better stability of the LmCS···83 complex. Our results indicate that compounds 30 (in fact, several agarofurans) and 83 are promissory leads for further studies towards CS inhibition.
Figure 12. (a) Lowest-energy docked pose of agarofuran 30 with LmCS; (b) 2D representation of protein ligand interactions of compound 30 with LmCS.
Figure 12. (a) Lowest-energy docked pose of agarofuran 30 with LmCS; (b) 2D representation of protein ligand interactions of compound 30 with LmCS.
Molecules 19 05550 g012
Interactions are represented as: van der Waals (green circle); polar (magenta circle); H-bond with amino acid side chains (blue dashed line); H-bond with amino acid backbone (green dashed line); pi interaction (orange line).
Germacranolide sesquiterpenes 81 and 83 also have an α-methylene γ-lactone moiety, but their respective lowest-energy poses are docked with different orientation. The lactone carbonyl of compound 83 exhibits a H-bond interaction (2.3 Å) with Lys51, which might promote a nucleophilic Michael addition. In contrast, the ligand interactions profile of compound 81 within the LmCS active site does not exhibit a free –SH or –NH2-containing residue near to the ligand.
Figure 13. 2D representation of protein ligand interactions with LmCS of: (a) compound 83; (b) compound 81.
Figure 13. 2D representation of protein ligand interactions with LmCS of: (a) compound 83; (b) compound 81.
Molecules 19 05550 g013
Interactions are represented as: van der Waals (green circle); polar (magenta circle); H-bond with amino acid side chains (blue dashed line); H-bond with amino acid backbone (green dashed line); pi interaction (orange line).
Figure 14. (a) Lowest-energy docked pose of agarofuran 22 with LdNMT; (b) 2D representation of protein ligand interactions of compound 22 with LdNMT.
Figure 14. (a) Lowest-energy docked pose of agarofuran 22 with LdNMT; (b) 2D representation of protein ligand interactions of compound 22 with LdNMT.
Molecules 19 05550 g014
Interactions are represented as: van der Waals (green circle); polar (magenta circle); H-bond with amino acid side chains (blue dashed line); H-bond with amino acid backbone (green dashed line); pi interaction (orange line).
Agarofurans 22, 27, and 39 exhibited the best docking energies (−6.9 kcal/mol) with LdNMT, but the vina score of the control used, DDD64558, was substantially better (−8.0 kcal/mol). A pi-interaction was found between nicotinoyl moiety and Arg179 for all nicotinoyl-containing agarofurans evaluated in this study (Figure 14). The presence of this substituent was found to be critical to the docking of the agarofuran-related sesquiterpenes. Similarly, sesquiterpene coumarin 2, exhibited almost identical affinity (−6.8 kcal/mol) with LdNMT to that of 22. However, after detailed analysis of the vina scores and binding modes of sesquiterpene···LdNMT complexes, this kind of compounds could be considered as non-suitable ligands for NMT inhibition.
Sesquiterpene coumarins also exhibited good affinity values in the docking with LmTryS, with compound 1 being the strongest-docking sesquiterpene coumarin (‒8.9 kcal/mol). Similarly, pseudoguaianolides 51 and 64 and germacranolides 80 and 89 exhibited good vina scores in the −8.5–−8.7 kcal/mol range. The complexes of all these compounds displayed lower docking energies than the control used, DDD66604, (−8.2 kca/mol). Figure 15 shows the 2D representation of protein ligand interactions of compounds 1, 64 and 80. The residue Gln5 was found to be a substantial polar contact in order to stabilize the complex for compounds 64 and 80, while Lys590 resulted as the indispensable side chain donor contact for compound 1. In addition, the lowest-energy pose of 80 has Lys590 very close to the α-methylene γ-lactone moiety.
Figure 15. 2D representation of protein ligand interactions with LmTryS of: (a)compound 1; (b) compound 64; (c) compound 80.
Figure 15. 2D representation of protein ligand interactions with LmTryS of: (a)compound 1; (b) compound 64; (c) compound 80.
Molecules 19 05550 g015
Interactions are represented as: van der Waals (green circle); polar (magenta circle); H-bond with amino acid side chains (blue dashed line); H-bond with amino acid backbone (green dashed line); pi interaction (orange line).
A Principal Component Analysis (PCA) was performed using the vina scores data of the test compounds with LmPTR1, LdNMT, LmCS, and LmTryS enzymes. The two first components of the PCA explain the 81% of the total variance. The resulting score plot (PC1 vs. PC2, Figure 16) revealed the clusters grouping the most affined ligands (red ellipses), on the one hand, and the ligands possessing the highest docking energies (green ellipses), on the other. The blue ellipse grouped the compounds with selectivity against one or two enzymes, but lowest docking values for the other enzymes. The resulting loading plot (PC1 vs. PC2) indicated that the docking values with LmTryS have a substantially different behavior than other enzymes, whereas vina scores values with LmPTR1, LdNMT, and LmCS are more highly correlated. Thus, compounds 75 and 76 are separated due to the higher docking values with LmTryS but good affinity with the other enzymes, while sesquiterpene coumarins 12 are clustered since they exhibited overall stronger docking energies towards the Leishmania protein targets as mentioned above. These PCA results indicated the good in silico information-based discrimination as a reasonable tool for ligand design, structural optimization and combinatorial purposes. Additionally, the majority of sesquiterpene lactones is located at the center of the score plot, indicating that, although they may have a certain affinity level towards the test enzymes, in general their structural features not fit for competition of the active sites.
Figure 16. Principal Component Analysis for the docking energies of test compounds with LmPTR1, LdNMT, LmCS, and LmTryS. (a) Score plot; (b) Loading plot.
Figure 16. Principal Component Analysis for the docking energies of test compounds with LmPTR1, LdNMT, LmCS, and LmTryS. (a) Score plot; (b) Loading plot.
Molecules 19 05550 g016
In order to observe the relationship between published IC50 values against Leishmania parasites and the docking energies for ligand design purposes, a correlation analysis was performed between Log(1/IC50) (IC50 in µM) versus vina scores (kcal/mol). However, due to the assay conditions-dependent variability of the data source and the limitation of the accuracy of docking calculations to predict in vitro activity, no significant correlation was then obtained (R2 < 20%). Such limitations in docking studies are well-known by the construction of the scoring functions through the postulation that solvation, entropy and electrostatics are largely applicable to several enzyme systems, adding compound absorption, plasma protein binding, tissue distribution, excretion and other biological events, which limit the prediction of biological activity. Nevertheless, there are several cases that distinguish between active and inactive compounds in docking studies and these procedures remain a powerful tool to facilitate the process of drug discovery [46,47].
Finally, in order to expand our study for identification of sesquiterpene-base leads, drug-like properties, such as partition coefficient (LogP), molar refractivity (MR), polar surface area (PSA), number of H-donors (H-D), H-acceptors (H-A) and rotatable bonds (RB), were calculated for the strongest-docking sesquiterpenes (Table 4). On applying the rule of five (RO5) and its variants [48], epimers 2227 are violating more than one of RO5, such as MR, PSA, number of atoms and H-acceptors whereas agarofuran isomers 2930, agarofuran 24 and xanthanolides 7576 exhibit an greater MR value to that of RO5 criteria. In contrast, sesquiterpene coumarins 12 have good drug-like properties, but the LogP values are near to the upper limit of lipophilicity. Within the strongest-docking sesquiterpenes, compounds 64, 81 and 83 displayed the best lead-like properties, suggesting these compounds deserve additional studies for enzyme inhibition.
Table 4. Calculated druglikeness properties for strongest-docking sesquiterpenes.
Table 4. Calculated druglikeness properties for strongest-docking sesquiterpenes.
CompoundLogP aMR bMW cMF dPSA eH-D fH-A gRB h
15.50110.2382.5C24H30O452.6043
24.78112.2382.5C24H30O455.8143
222.03149.7595.6C32H37NO10147.611110
243.48152.7578.6C33H38O9128.6299
272.03149.7595.6C32H37NO10147.611110
294.61151.5562.7C33H38O8108.4189
304.61151.5562.7C33H38O8108.4189
313.22120.5442.5C26H34O682.1166
395.06168.7641.7C37H39NO9127.301011
511.0870.6262.3C15H18O463.6140
631.7095.6364.4C20H28O693.1263
641.7095.6364.4C20H28O693.1263
751.74139.4508.6C30H36O7107.0173
760.82137.9524.6C30H36O8119.5183
811.2299.0380.4C20H28O7110.1274
831.8697.6364.4C20H28O689.9164
a partition coefficient; b molar refractivity (cm3·mol−1); c molecular weight (g/mol); d molecular formula; e polar surface area (Å2); f number of H-donor elements; g number of H-acceptor elements; h number of rotatable bonds.

3. Experimental Section

3.1. Protein Preparation

The X-ray crystallographic structure of LmPTR1, LdNMT, LmCS, and LmTryS proteins were obtained from the Protein Data Bank at a resolution of <2.5Å. Water molecules, ligands and other hetero atoms were removed from the protein molecule along with the chains excepting chain A. Addition of hydrogen atoms to the protein was performed. Energy minimization was performed with the CHARMm force field by using conjugate gradient method with an RMS gradient of 0.01 kcal/Å mol on Discovery Studio 2.5 software [49].

3.2. Preparation of the Ligands

Structural information of 123 test sesquiterpenoids was acquired from a literature survey on natural sesquiterpenes with antileishmanial activity report (see supplementary material for the IUPAC and common names). A Monte-Carlo randomized conformational search, without any geometrical restrictions, was made using the Merck Molecular Force Field (MMFF94) included in the SPARTAN software [50] with a limit of 500 conformers. Energetically lowest stable conformers within a 6 kcal/mol energy range were geometrically optimized using the semi-empirical AM1 parametrization implemented in the software package. The minimized protein and ligands were saved in PDB format for further docking analysis.

3.3. Docking Analyses

Docking experiments were performed using the Autodock/Vina (1.1.2) package under Linux in a Python 2.7 environment, using the AMBER force field [51]. The protein and ligand molecules were prepared as described above. The docking experiment on test enzymes was carried out between the energy-minimized ligand into the active site through a cube at the geometrical center of the native ligand present in the evaluated PDB structure, with the dimensions 24 × 24 × 24Å, covering the ligand binding site with a grid point spacing of 0.375Å. The docking poses are ranked according to their docking scores (as free energy of binding) and both the ranked list of docked ligands and their corresponding binding poses were exported as a CSV file for further analysis. The docked structures were viewed using PyMOL (1.3r2) [51].

3.4. Ligand Interactions

Resulting docking information-containing files (in .pdbqt format) were imported into the Accelyrs Discovery Studio 2.5 software package. Once the docked enzyme-ligand complex was imported, an analysis of the binding site was performed in order to create a ligand interaction 2D diagram.

3.5. Drug-Like Properties Calculations

Drug-like properties of each sesquiterpene were separately calculated using the Molecular Topology and ChemProPro plugins available into the ChemBio3D® Ultra 11.0 software package (Cambridge Soft Corporation, Cambridge, MA, USA) starting from the 3D structure.

3.6. Statistics

Regression and PCA analyses were carried out using the algorithms included into the software R (3.0.1 version) [52].

4. Conclusions

A plethora of antileishmanial naturally-occurring compounds active against promastigotes and amastigotes have been identified and evaluated, but their mode of action has not been elucidated. A great number of sesquiterpenes remains in the available literature considered as good antileishmanial leads, e.g., sesquiterpene lactones. However, the lack of molecular studies is a barrier in the route for improving the sesquiterpene-based drug design. In the present in silico study, we have identified molecular details of the binding modes of a set of antileishmanial sesquiterpenes against four drug-enzyme targets (LmPTR1, LdNMT, LmCS, and LmTryS). Through molecular docking it was found that two sesquiterpene coumarins are good ligand candidates for the PTR1 and TryS inhibition purposes as well as some xanthanolides exhibited better affinity towards PTR1 and CS binding. The ligand interactions allowed establishing the critical contacts between the each ligand and the active site of the enzymes. The present study is an excellent starting point for future studies of structural optimization of these compounds, whose observed details at the molecular level allowing opening a working route towards developing therapeutic sesquiterpenes-based prototypes against Leishmania.

Acknowledgments

This study was financially supported by the Vicerrectoría de Investigaciones at Universidad Militar Nueva Granada—Validity 2013, through the Project CIAS-1172. This work is also an activity conducted within the Research Network Natural Products against Neglected Diseases (ResNet NPND) [53].

Author Contributions

Authors contributed to the present work as described in the following: Literature review for antiparasitic sesquiterpenes: ECB, FB. Conceptualization of the study and experimental design: ECB. Molecular modelling and docking calculations: ECB, FB. Data Analysis: ECB, FB. Manuscript preparation and proofreading: ECB, FB.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Loría-Cervera, E.N.; Andrade-Narváez, F.J. Animal models for the study of leishmaniasis immunology. Rev. Inst. Med. Trop. Sao Paulo 2014, 56, 1–11. [Google Scholar] [CrossRef]
  2. Reithinger, R.; Dujardin, J.C.; Louzir, H.; Pirmez, C.; Alexander, B.; Brooker, S. Cutaneous leishmaniasis. Lancet Infect. Dis. 2007, 7, 581–596. [Google Scholar] [CrossRef]
  3. Arboleda, M.; Jaramillo, L.; Ortiz, D.; Díaz, A. Leishmaniasis cutánea y herpes zoster multidermatómico. Rev. Chil. Infectol. 2013, 30, 680–682. [Google Scholar] [CrossRef]
  4. Sundar, S.; Rai, M. Advances in the treatment of leishmaniasis. Curr. Opin. Infect. Dis. 2002, 15, 593–598. [Google Scholar] [CrossRef]
  5. Jebran, A.F.; Schleicher, U.; Steiner, R.; Wentker, P.; Mahfuz, F.; Stahl, H.C.; Amin, F.M.; Bogdan, C.; Stahl, K.W. Rapid healing of cutaneous Leishmaniasis by high-frequency electrocauterization and hydrogel wound care with or without DAC N-055: A randomized controlled phase II a trial in kabul. PLoS Negl. Trop. Dis. 2014, 8, e2694. [Google Scholar] [CrossRef]
  6. Barrett, M.P.; Mottram, J.C.; Goombs, G.H. Recent advances in identifying and validating drug targets in trypanosomes and leishmanias. Trends Microbiol. 1999, 7, 82–88. [Google Scholar] [CrossRef]
  7. Ong, H.B.; Sienkiewicz, N.; Wyllie, S.; Fairlamb, A.H. Dissecting the metabolic roles of pteridine reductase 1 in Trypanosoma. brucei and Leishmania. major. J. Biol. Chem. 2001, 286, 10429–10438. [Google Scholar]
  8. Kheirandish, F.; Bandehpour, M.; Haghighi, A.; Mahboudi, F.; Mohebali, M.; Kazemi, B. Inhibition of Leishmania major PTR1 Gene Expression by Antisense in Escherichia coli. Iran. J. Public Health 2012, 41, 65–71. [Google Scholar]
  9. Krauth-Siegel, R.L.; Comini, M.A. Redox control in trypanosomatids, parasitic protozoa with trypanothione-based thiol metabolism. Biochim. Biophys. Acta 2008, 1780, 1236–1248. [Google Scholar]
  10. Fyfe, P.K.; Westrop, G.D.; Ramos, T.; Muller, S.; Coombs, G.H.; Hunter, W.N. Structure of Leishmania major cysteine synthase. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 2012, 68, 738–743. [Google Scholar] [CrossRef]
  11. Fyfe, P.K.; Oza, S.L.; Fairlamb, A.H.; Hunter, W.N. Leishmania trypanothione synthetase-amidase structure reveals a basis for regulation of conflicting synthetic and hydrolytic activities. J. Biol. Chem. 2008, 283, 17672–17680. [Google Scholar]
  12. Resh, M.D. Fatty acylation of proteins: New insights into membrane targeting of myristoylated and palmitoylated proteins. Biochim. Biophys. Acta 1999, 1451, 1–16. [Google Scholar] [CrossRef]
  13. Brannigan, J.A.; Smith, B.A.; Yu, Z.; Brzozowski, A.M.; Hodgkinson, M.R.; Maroof, A.; Price, H.P.; Meier, F.; Leatherbarrow, R.J.; Tate, E.W.; et al. N-Myristoyltransferase from Leishmania donovani: Structural and functional characterisation of a potential drug target for visceral leishmaniasis. J. Mol. Biol. 2010, 396, 985–999. [Google Scholar] [CrossRef]
  14. Schmidt, T.J.; Khalid, S.A.; Romanha, A.J.; Alves, T.M.A.; Biavatti, M.W.; Brun, R.; da Costa, F.B.; de Castro, S.L.; Ferreira, V.F.; de Lacerda, M.V.G.; et al. The potential of secondary metabolites from plants as drugs or leads against protozoan neglected ideases-Part I. Curr. Med. Chem. 2012, 19, 2128–2175. [Google Scholar]
  15. Schmidt, T.J.; Khalid, S.A.; Romanha, A.J.; Alves, T.M.A.; Biavatti, M.W.; Brun, R.; da Costa, F.B.; de Castro, S.L.; Ferreira, V.F.; de Lacerda, M.V.G.; et al. The potential of secondary metabolites from plants as drugs or leads against protozoan neglected ideases-Part II. Curr. Med. Chem. 2012, 19, 2176–2228. [Google Scholar]
  16. Ogungbe, I.V.; Setzer, W.N. In-silico Leishmania target selectivity of antiparasitic terpenoids. Molecules 2013, 18, 7761–7847. [Google Scholar] [CrossRef]
  17. Cavazzuti, A.; Paglietti, G.; Hunter, W.N.; Gamarro, F.; Piras, S.; Loriga, M.; Allecca, S.; Corona, P.; McLuskey, K.; Tulloch, L.; et al. Discovery of potent pteridine reductase inhibitors to guide antiparasite drug development. Proc. Natl. Acad. Sci. USA 2008, 105, 1448–1453. [Google Scholar] [CrossRef]
  18. Iranshahi, M.; Arfa, P.; Ramezani, M.; Jaafari, M.R.; Sadeghian, H.; Bassarello, C.; Piacente, S.; Pizza, C. Sesquiterpene coumarins from Ferula szowitsiana and in vitro antileishmanial activity of 7-prenyloxycoumarins against promastigotes. Phytochemistry 2007, 68, 554–561. [Google Scholar]
  19. Pérez-Victoria, J.M.; Tincusi, B.M.; Jiménez, I.A.; Bazzocchi, I.L.; Gupta, M.P.; Castanys, S.; Gamarro, F.; Ravelo, A.G. New natural sesquiterpenes as modulators of daunomycin resistance in a Multidrug-Resistant Leishmania tropica line. J. Med. Chem. 1999, 42, 4388–4393. [Google Scholar] [CrossRef]
  20. Torres-Romero, D.; Muñoz-Martínez, F.; Jiménez, I.A.; Castanys, S.; Gamarro, F.; Bazzocchi, I.L. Novel dihydro-β-agarofuran sesquiterpenes as potent modulators of human P-glycoprotein dependent multidrug resistance. Org. Biomol. Chem. 2009, 7, 5166–5172. [Google Scholar] [CrossRef]
  21. Torres-Romero, D.; Jiménez, I.A.; Rojas, R.; Gilman, R.H.; López, M.; Bazzocchi, I.L. Dihydro-β-agarofuran sesquiterpenes isolated from Celastrus vulcanicola as potential anti-Mycobacterium tuberculosis multidrug-resistant agents. Bioorg. Med. Chem. 2011, 19, 2182–2189. [Google Scholar] [CrossRef]
  22. Delgado-Méndez, P.; Herrera, N.; Chávez, H.; Estévez-Braun, A.; Ravelo, A.G.; Cortes, F.; Castanys, S.; Gamarro, F. New terpenoids from Maytenus. apurimacensis as MDR reversal agents in the parasite Leishmania. Bioorg. Med. Chem. 2008, 16, 1425–1430. [Google Scholar] [CrossRef]
  23. Santos, V.A.F.F.M.; Regasini, L.O.; Nogueira, C.R.; Passerini, G.D.; Martinez, I.; Bolzani, V.S.; Graminha, M.A.S.; Cicarelli, R.M.B.; Furlan, M. Antiprotozoal sesquiterpene pyridine alkaloids from Maytenus ilicifolia. J. Nat. Prod. 2012, 75, 991–995. [Google Scholar] [CrossRef]
  24. Claudino, V.D.; da Silva, K.C.; Cechinel Filho, V.; Yunes, R.A.; Delle Monache, F.; Giménez, A.; Salamanca, E.; Gutierrez-Yapu, D.; Malheiros, A. Drimanes from Drimys brasiliensis with leishmanicidal and antimalarial activity. Mem. Inst. Oswaldo Cruz 2013, 108, 140–144. [Google Scholar]
  25. Sülsen, V.P.; Frank, F.M.; Cazorla, S.I.; Anesini, C.A.; Malchiodi, E.L.; Freixa, B.; Vila, R.; Muschietti, L.V.; Martino, V.S. Trypanocidal and leishmanicidal activities of sesquiterpene lactones from Ambrosia tenuifolia Sprengel (Asteraceae). Antimicrob. Agents Chemother. 2008, 52, 2415–2419. [Google Scholar] [CrossRef]
  26. Barrera, P.; Sülsen, V.P.; Lozano, E.; Rivera, M.; Beer, M.F.; Tonn, C.; Martino, V.S.; Sosa, M. A natural sesquiterpene lactones induce oxidative stress in Leishmania mexicana. Evid. Based. Complement. Alternat. Med. 2013, 2013, 163404. [Google Scholar]
  27. Schmidt, T.J.; Nour, A.M.M.; Khalid, S.A.; Kaiser, M.; Brun, R. Quantitative structure—antiprotozoal activity relationships of sesquiterpene lactones. Molecules 2009, 14, 2062–2076. [Google Scholar] [CrossRef]
  28. Fournet, A.; Muñoz, V.; Roblot, F.; Hocquemiller, R.; Cavé, A.; Gantier, J.-C. Antiprotozoal activity of dehydrozaluzanin C, a sesquiterpene lactone isolated from Munnozia maronii (Asteraceae). Phytoterapy. Res. 1993, 7, 111–115. [Google Scholar] [CrossRef]
  29. Da Silva, B.P.; Cortez, D.A.; Violin, T.Y.; Dias Filho, B.P.; Nakamura, C.V.; Ueda-Nakamura, T.; Ferreira, I.C.P. Antileishmanial activity of a guaianolide from Tanacetum parthenium (L.) Schultz Bip. Parasitol. Int. 2010, 59, 643–646. [Google Scholar] [CrossRef]
  30. Tiuman, T.S.; Ueda-Nakamura, T.; Cortez, D.A.G.; Filho, B.P.D.; Morgado-Díaz, J.A.; de Souza, W.; Nakamura, C.V. Antileishmanial activity of parthenolide, a sesquiterpene lactone isolated from Tanacetum parthenium. Antimicrob. Agents Chemother. 2005, 49, 176–182. [Google Scholar] [CrossRef]
  31. Berger, I.; Passreiter, C.M.; Cáceres, A.; Kubelka, W. Antiprotozoal activity of Neurolaena lobata. Phytother. Res. 2001, 15, 327–30. [Google Scholar] [CrossRef]
  32. Passreiter, C.M.; Wendisch, D.; Gondol, D. Sesquiterpene lactones from Neurolaena lobata. Phytochemistry 1995, 39, 133–137. [Google Scholar] [CrossRef]
  33. Sharma, U.; Singh, D.; Kumar, P.; Dobhal, M.P.; Singh, S. Antiparasitic activity of plumericin & isoplumericin isolated from Plumeria bicolor against Leishmania donovani. Indian J. Med. Res. 2011, 709–716. [Google Scholar]
  34. Savoia, D.; Avanzini, C.; Allice, T.; Callone, E.; Guella, G.; Dini, F. Antimicrobial activity of euplotin C, the sesquiterpene taxonomic marker from the marine ciliate Euplotes crassus. Antimicrob. Agents Chemother. 2004, 48, 3828–3833. [Google Scholar] [CrossRef]
  35. Morales-Yuste, M.; Morillas-Márquez, F.; Martín-Sánchez, J.; Valero-López, A.; Navarro-Moll, M.C. Activity of (−)-α-bisabolol against Leishmania infantum promastigotes. Phytomedicine 2010, 17, 279–281. [Google Scholar] [CrossRef]
  36. Rojas-Silva, P.; Graziose, R.; Vesely, B.; Poulev, A.; Mbeunkui, F.; Grace, M.H.; Kyle, D.E.; Lila, M.A.; Raskin, I. Leishmanicidal activity of a daucane sesquiterpene isolated from Eryngium foetidum. Pharm. Biol. 2014, 52, 398–401. [Google Scholar] [CrossRef]
  37. Dos Santos, A.O.; Veiga-Santos, P.; Ueda-Nakamura, T.; Filho, B.P.D.; Sudatti, D.B.; Bianco, E.M.; Pereira, R.C.; Nakamura, C.V. Effect of elatol, isolated from red seaweed Laurencia dendroidea, on Leishmania amazonensi. Mar. Drugs 2010, 8, 2733–2743. [Google Scholar] [CrossRef]
  38. Gul, W.; Hammond, N.L.; Yousaf, M.; Peng, J.; Holley, A.; Hamann, M.T. Chemical transformation and biological studies of marine sesquiterpene (S)-(+)-curcuphenol and its analogs. Biochim. Biophys. Acta 2007, 1770, 1513–1519. [Google Scholar] [CrossRef]
  39. Odonne, G.; Herbette, G.; Eparvier, V.; Bourdy, G.; Rojas, R.; Sauvain, M.; Stien, D. Antileishmanial sesquiterpene lactones from Pseudelephantopus spicatus, a traditional remedy from the Chayahuita Amerindians (Peru), Part III. J. Ethnopharmacol. 2011, 137, 875–879. [Google Scholar] [CrossRef]
  40. Ganfon, H.; Bero, J.; Tchinda, A.T.; Gbaguidi, F.; Gbenou, J.; Moudachirou, M.; Frédérich, M.; Quetin-Leclercq, J. Antiparasitic activities of two sesquiterpenic lactones isolated from Acanthospermum hispidum D.C. J. Ethnopharmacol. 2012, 141, 411–417. [Google Scholar] [CrossRef]
  41. Karioti, A.; Skaltsa, H.; Kaiser, M.; Tasdemir, D. Trypanocidal, leishmanicidal and cytotoxic effects of anthecotulide-type linear sesquiterpene lactones from Anthemis auriculata. Phytomedicine 2009, 16, 783–787. [Google Scholar] [CrossRef]
  42. Chollet, C.; Crousse, B.; Bories, C.; Bonnet-Delpon, D.; Loiseau, P.M. In vitro antileishmanial activity of fluoro-artemisinin derivatives against Leishmania donovani. Biomed. Pharmacother. 2008, 62, 462–465. [Google Scholar] [CrossRef]
  43. Picman, A.K.; Rodríguez, E.; Towers, G.H.N. Formation of adducts of parthenin and related sesquiterpene lactones with cysteine and glutathione. Chem. Biol. Interactions 1979, 28, 83–89. [Google Scholar] [CrossRef]
  44. Salapovic, H.; Geier, J.; Reznicek, G. Quantification of sesquiterpene lactones in asteraceae plant extracts: Evaluation of their allergenic potential. Sci. Pharm. 2013, 81, 807–818. [Google Scholar] [CrossRef]
  45. Fuchino, H.; Koide, T.; Takahashi, M.; Sekita, S.; Satake, M. New sesquiterpene lactones from Elephantopus mollis and their leishmanicidal activities. Planta Med. 2001, 67, 647–53. [Google Scholar] [CrossRef]
  46. Kitchen, D.B.; Decornez, H.; Furr, J.R.; Bajorath, J. Docking and scoring in virtual screening for drug discovery: Methods and applications. Nat. Rev.Drug Discov. 2004, 3, 935–949. [Google Scholar]
  47. Vilar, S.; Costanzi, S. Predicting biological activities through QSAR analysis and Docking-based scoring. Methods Mol. Biol. 2012, 914, 271–284. [Google Scholar]
  48. Lipinski, C.A.; Lombardo, F.; Dominy, B.W.; Feeney, P.J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug. Delivery Rev. 1997, 23, 4–25. [Google Scholar]
  49. Discovery Studio Modeling Environment; Release 2.5; Accelrys Software Inc: San Diego, CA, USA, 2013.
  50. Spartan ‘08 V1.1.1; Wavefunction, Inc.: Irvine, CA, USA, 2008.
  51. Seeliger, D.; de Groot, B.L. Ligand docking and binding site analysis with PyMOL and autodock/Vina. J. Comput. Aided Mol. Des. 2010, 24, 417–422. [Google Scholar] [CrossRef]
  52. R Core Team. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing: Vienna, Austria, 2012. ISBN 3–900051–07–0. Available online: http://www.R-project.org/ (accessed on 16 March 2014).
  53. Research Network Natural Products against Neglected Diseases (ResNet NPND); Westfälische Wilhelms-Universität Münster, 2011. Available online: http://www.uni-muenster.de/Chemie.pb/forschen/ResNetNPND/ (accessed on 16 March 2014).
  • Sample Availability: Not available.

Share and Cite

MDPI and ACS Style

Bernal, F.A.; Coy-Barrera, E. In-Silico Analyses of Sesquiterpene-Related Compounds on Selected Leishmania Enzyme-Based Targets. Molecules 2014, 19, 5550-5569. https://doi.org/10.3390/molecules19055550

AMA Style

Bernal FA, Coy-Barrera E. In-Silico Analyses of Sesquiterpene-Related Compounds on Selected Leishmania Enzyme-Based Targets. Molecules. 2014; 19(5):5550-5569. https://doi.org/10.3390/molecules19055550

Chicago/Turabian Style

Bernal, Freddy A., and Ericsson Coy-Barrera. 2014. "In-Silico Analyses of Sesquiterpene-Related Compounds on Selected Leishmania Enzyme-Based Targets" Molecules 19, no. 5: 5550-5569. https://doi.org/10.3390/molecules19055550

Article Metrics

Back to TopTop