Next Article in Journal
Acylated mono-, bis- and tris- Cinchona-Based Amines Containing Ferrocene or Organic Residues: Synthesis, Structure and in Vitro Antitumor Activity on Selected Human Cancer Cell Lines
Previous Article in Journal
Evaluation of Rheological Properties and Swelling Behaviour of Sonicated Scleroglucan Samples
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Methodologies and Evaluations of Computer-Aided Drug Design Methodologies Applied to a Series of 2-Aminothiophene Derivatives with Antifungal Activities

by
Luciana Scotti
1,*,
Marcus Tullius Scotti
2,
Edeltrudes De Oliveira Lima
3,
Marcelo Sobral da Silva
1,
Maria Do Carmo Alves de Lima
4,
Ivan Da Rocha Pitta
4,
Ricardo Olímpio de Moura
5,
Jaismary Gonzaga Batista de Oliveira
5,
Rayssa Marques Duarte da Cruz
5 and
Francisco Jaime Bezerra Mendonça, Júnior
5,*
1
Centro de Biotecnologia, Universidade Federal da Paraíba, João Pessoa 50670-910, PB, Brazil
2
Departamento de Engenharia e Meio Ambiente, Universidade Federal da Paraíba, Campus IV, Rio Tinto 58297-000, PB, Brazil
3
Laboratório Micologia, Departamento de Ciências Farmacêuticas, Centro de Ciências da Saúde, Universidade Federal da Paraíba, João Pessoa 50670-910, PB, Brazil
4
Laboratório de Planejamento e Síntese de Fármacos, Departamento de Antibióticos, Universidade Federal de Pernambuco, Recife 50670-910, PE, Brazil
5
Laboratório de Síntese e Vetorização de Moléculas, Departamento de Ciências Biológicas, Universidade Estadual da Paraíba, Rua Horácio Trajano de Oliveira s/n, Cristo Redentor, João Pessoa 58070-450, PB, Brazil
*
Authors to whom correspondence should be addressed.
Molecules 2012, 17(3), 2298-2315; https://doi.org/10.3390/molecules17032298
Submission received: 30 November 2011 / Revised: 3 February 2012 / Accepted: 21 February 2012 / Published: 24 February 2012

Abstract

:
Fifty 2-[(arylidene)amino]-4,5-cycloalkyl[b]thiophene-3-carbonitrile derivatives were screened for their in vitro antifungal activities against Candida krusei and Cryptococcus neoformans. Based on experimentally determined minimum inhibitory concentration (MIC) values, we conducted computer-aided drug design studies [molecular modelling, chemometric tools (CPCA, PCA, PLS) and QSAR-3D] that enable the prediction of three-dimensional structural characteristics that influence the antifungal activities of these derivatives. These predictions provide direction with regard to the syntheses of new derivatives with improved biological activities, which can be used as therapeutic alternatives for the treatment of fungal infections.

Graphical Abstract

1. Introduction

Systemic fungal infections have gradually increased over the last three decades, leading to considerable rates of morbidity and mortality, principally due to the low effectiveness of available medications and the development of resistant strains. Immunocompromised patients, such as those with previous diseases like tuberculosis, cancer, diabetes and AIDS and patients undergoing chemotherapy or who have had transplants, are more susceptible to opportunistic fungal infections, which often prove fatal [1].
Conventional treatment regimes for diseases caused by fungi are based on only a few chemotherapies (i.e., azoles and polyenes), and these are not completely effective and present serious problems related to their narrow spectrum, high toxicity, low potency (they are generally fungistatic) and inadequate pharmacokinetic properties. Taken together, these characteristics result in many cases of microbial resistance, making treatment even more difficult [2].
Thus, there is a clear need to develop new antifungal agents as therapeutic alternatives for the control of fungal infections. In this context, compounds derived from heterocyclic thiophenes have emerged as promising antifungal drugs [3,4,5,6,7,8,9,10], some of which have become commercially available (for example, sertaconazol, a benzothiophene derivative).
The microorganisms Candida krusei and Criptococcus neoformans were selected for the present study because they presented better profiles of antifungal sensitivity compared to other 2-amino-thiophene derivatives previously synthesised by our group [10]. Furthermore, these microorganisms have additional medical relevance in that Candida krusei shows natural resistance to the standard medication fluconazole, one of the most widely used antifungals in clinical medicine. The same microorganism shows reduced sensitivity to both flucytosine and amphotericin B [11,12,13] and causes greater mortality than the species C. albicans [14] and C. neoformans due to its unusual ability to evolve into meningitis. Candida krusei and Criptococcus neoformans are also important markers for HIV infection [15,16].
This study investigates a series of fifty 2-[(arylidene)-amino]-4,5-cycloalkyl[b]thiophene-3-carbonitrile derivatives (generically called 2-aminothiophenes, Figure 1). After experimentally assessing their antifungal activities in vitro by testing their sensitivities to the fungal species Candida krusei and Cryptococcus neoformans, the derivatives were subjected to molecular modelling, chemometric tools and QSAR-3D in order to extract structural information relating to their antifungal properties [17,18,19,20,21].

2. Results and Discussion

2.1. Synthesis and Antifungal Activity

The fifty 2-aminothiophenes evaluated in this study (Figure 1) were previously synthesised and spectroscopically characterised by this group [10,22,23] using the classic Gewald reaction [24] followed by substitution (via condensation) with different aromatic aldehydes. Table 1 presents the chemical structures of the 50 derivatives studied (compounds 150) as well as their minimum inhibitory concentrations (MIC) expressed in μg/mL (referring to their antifungal activities with respect to the fungi Candida krusei and Cryptococcus neoformans).
Figure 1. General structure of the studied 2-aminothiophenes.
Figure 1. General structure of the studied 2-aminothiophenes.
Molecules 17 02298 g001
Table 1. Anti-Candida and anti-Criptococcus activities and chemical formulas of the studied 2-aminothiophene derivatives.
Table 1. Anti-Candida and anti-Criptococcus activities and chemical formulas of the studied 2-aminothiophene derivatives.
Comp.nArMIC (μg/mL)
C. neoformans ICB59Candida krusei LM13
11Ph2,5001,250
214-NO2-Ph1,2502,500
312-Thiophenyl156>10,000
412-Indole1,250>10,000
514-F-Ph2,500>10,000
614-Br-Ph2,500>10,000
714-Morpholine312>10,000
814-OEt-Ph2,500>10,000
914-Cl-Ph5,0002,500
1014-OMe-Ph1,2502,500
1112,3-Cl-Ph156>10,000
1213,4-Cl-Ph>10,0005,000
1312,4-Cl-Ph7878
1413,4,5-OMe-Ph15678
1512-OMe,5-Br-Ph15678
1618-Quinoline1565,000
1724-NO2-Ph1,250128
1822-Indole78>10,000
1924-OBz-Ph>10,000625
2024-Br-Ph>10,000>10,000
2124-Cl-Ph>10,000>10,000
2224-OMe-Ph1561,250
2324-F-Ph5,000625
2422,3-Cl-Ph156156
2523,4-Cl-Ph1,250312
2622,4-Cl-Ph2,500>10,000
2723,4,5-OMe-Ph1562,500
2822-OMe,5-Br-Ph10,00010,000
2924-Pyrrolidine-Ph2,5002,500
3023-Thiophenyl2,500>10,000
3123,4-OBz-Ph2,5001,250
3228-Quinoline78625
3324-Morpholine-Ph2,5002,500
3434-Me-Ph1562,500
3534-iPr-Ph1,2502,500
363Ph>10,0002,500
3734-NO2-Ph10,000625
3832-indole3122,500
3934-OBz-Ph10,000>10,000
4034-Br-Ph2,5005,000
4134-Cl-Ph1,2502,500
4234-OMe-Ph3122,500
4334-OEt-Ph5,0002,500
4433,4-Cl-Ph1,2502,500
4532,4-Cl-Ph>10,0002,500
4633,4,5-OMe-Ph1,2501,250
4732-OMe,5-Br-Ph10,0005,000
4833-Thiophenyl2,500>10,000
4938-Quinoline78625
5034-Et-Ph2,5002,500

2.2. Electronic Surfaces

The density functional theory of electrostatic potential and ionisation potential and the HOMO and LUMO orbital surface were determined for each compound. Figure 2 shows some striking examples regarding ionisation potential. To obtain and compare surfaces’ ionisation potentials, we adopted a single band (red 14800–215000 blue). To evaluate the electronic surfaces, the molecules were represented by the tube model. Specific atoms are represented by colours as follows: carbon (grey), oxygen (red) and hydrogen (white). Using the Spartan program, these electronic-surface studies revealed that derivatives with the same aryl substituent linked to the 2-amino portion show significant differences in their reactivities (compounds 10, 22 and 42–4-methoxyphenyl substituent; 2, 17 and 37–4-nitrophenyl substituent), depending on the type of cycloalkyl ring attached to each thiophene nucleus. The cyclopenta[b]thiophene derivatives exhibited lower reactivities compared to the cyclohexa- and cyclohepta[b]thiophene derivatives. This difference in reactivities was demonstrated principally by comparing the ionisation surface potentials, with cyclopenta[b]thiophenes exhibiting negative ionisation potentials (red) and cyclohexa- and cyclohepta-[b]thiophenes exhibiting positive ionisation potentials (blue) (Figure 2).
Using the VolSurf+ program, the compounds were subjected to the GRID force field, and their interactions with the following probes were evaluated: H2O (light blue), O (red), N (dark blue) and DRY (green) (Figure 3).
Figure 2. Ionisation potentials obtained for compounds 10, 22, 42, 2, 17 and 37.
Figure 2. Ionisation potentials obtained for compounds 10, 22, 42, 2, 17 and 37.
Molecules 17 02298 g002
Figure 3. Maps of interaction with the probes H2O, N1 and O obtained for inactive compounds 20 and 45 and for active compounds 15 and 17.
Figure 3. Maps of interaction with the probes H2O, N1 and O obtained for inactive compounds 20 and 45 and for active compounds 15 and 17.
Molecules 17 02298 g003
The interaction maps generated with the GRID force field show that, independent of the reactivities and electronic behaviours exhibited by cyclopenta-, cyclohexa- and cyclohepta-[b]thiophenes, the inactive compounds (for example: compounds 20 and 45) had little or no interaction with the H2O probe. In this study, the cyclopenta[b]thiophene derivatives exhibited greater interactions with the H2O probe and better antifungal profiles, indicating that increases in water solubility correlate with increases in antifungal activity for these derivatives.
At first glance, the cyclopenta[b]thiophenes appear more active, leading us to believe that they are also more water soluble. This belief was confirmed by subsequent analyses. Thus, we conclude that more active compounds exhibit greater interactions with H2O.

2.3. Chemometric Studies Using the Program VolSurf+

After conformational analyses, data obtained for the fifty 2-aminothiophene derivatives shown in Table 1 were subjected to chemometric analyses [17,18,19] (PCA and CPCA) using the VolSurf+ program for Windows [25,26].

2.3.1. CPCA (Consensus PCA)

A preliminary exploratory analysis, CPCA, which considered 128 independent variables or descriptors, was developed. Pre-processing (autoscaling) of data was performed, and 13 blocks of descriptors were calculated. Table 2 and Table 3 present the values obtained for the variance in anti-Candida and anti-Cryptococcus activities.
Table 2. Results of CPCA analyses of anti-Candida activities.
Table 2. Results of CPCA analyses of anti-Candida activities.
PC% explained variance from original data
141.26
222.69
33.06
41.05
50.00
Table 3. Results of CPCA analyses of anti-Cryptococcus activities.
Table 3. Results of CPCA analyses of anti-Cryptococcus activities.
PC% explained variance from original data
145.03
220.11
34.00
40.98
50.00
The blocks of descriptors that presented higher weights are shown in the Figure 4 graphs, which were constructed taking into account the orthogonal properties of the CPCA: LogS (ADME descriptors) algorithm and H2O, resulting in a total of 34 variables.
Regarding CPCA formalism, 128 independent variables were considered, and no biological data were used as input to the model. The orthogonal properties of the CPCA algorithm were explored. The use of CPCA in decentralised process monitoring and diagnosis is derived from regular PCA scores and residuals. Two significant principal components (PCs) were found using a cross-validation technique, explaining more than 60% of the total variance in anti-Candida and anti-Criptococcus activities (Table 3 and Table 4).
In these CPCA studies, we observed super block-weights and, comparing several blocks of descriptor variables measured on the same objects, became aware of the important influence that each block has on the associated calculations. Thirteen blocks of descriptors were calculated, and their weights were plotted considering two factors: PC1 and PC2.
Figure 4. The relationship between the weights of the blocks of descriptors and the PC1 and PC2 components of the CPCA calculated using activities against C. krusei and C. neoformans.
Figure 4. The relationship between the weights of the blocks of descriptors and the PC1 and PC2 components of the CPCA calculated using activities against C. krusei and C. neoformans.
Molecules 17 02298 g004
Table 4. Results of PCA analyses of anti-Candida activities.
Table 4. Results of PCA analyses of anti-Candida activities.
PC% explained variance from original data
143.77
220.58
32.33
41.71
50.09
As shown in Figure 4, the LogS and H2O blocks presented significant weights in relation to PC2 and PC1, respectively. As previously mentioned, the CPCA algorithm is essentially equivalent to the regular PCA, but new definitions of blocks and variables of larger contributions were investigated with PCA.

2.3.2. PCA

With regard to the interaction of 3D structures and a GRID force field, PCA results were obtained using the H2O and LogS probes. Thirty-four descriptors were calculated. The data were autoscaled (pre-processed). Once again, for both activities, approximately 60% of the total variance was explained by the PC1 and PC2 components (Table 4 and Table 5).
Table 5. Results of PCA analyses of anti-Cryptococcus activities.
Table 5. Results of PCA analyses of anti-Cryptococcus activities.
PC% explained variance from original data
139.80
222.75
33.56
41.00
50.08
Good separations were observed between active and inactive compounds for both anti-Candida activities and anti-Cryptococcus activities (Figure 5). The PCA studies considered the 3D interaction energies calculated using the LogS and H2O probes in a GRID force field. The PCA method also served to refine the data. Thirty-four total descriptors were calculated. Using the leave-one-out (LOO) cross-validation technique (Table 4 and Table 5), the PC1 and PC2 components captured approximately 60% of the total variance from the original data. Active and inactive compounds were well distinguished (see Figure 5). Defined clusters of active and inactive compounds were observed when the LogS and H2O VolSurf descriptors were used.
The VolSurf descriptors give the possibility of representing the molecular interaction regions of interest graphically, a big advantage in pharmacological studies and are independent of the molecular orientation within the grid. The main advantage of the GRID method is the great variety of available probes, represented by several functional groups among them there are probes that can both accept and donate a hydrogen bond (e.g., water), probes that cannot turn around (e.g., carbonyl probe), and a hydrophobic probe. These results indicate the strong predictability of the model.
Figure 5. Positions of objects in relation to PC1 and PC2 components. Blue represents more active compounds, lilac represents compounds that are moderately active, and pink represents less active compounds.
Figure 5. Positions of objects in relation to PC1 and PC2 components. Blue represents more active compounds, lilac represents compounds that are moderately active, and pink represents less active compounds.
Molecules 17 02298 g005
The presence of descriptors indicating the following pharmacokinetic properties influenced antifungal activity: CACO2, SKIN, L1, 2GS, SOLY, LOGS10, LOGS11. In addition, FLEX_RB highlights conformational characteristics. These abbreviations are defined as follows:
  • SOLY-intrinsic solubility,
  • LOGS10 and 11-intrinsic solubilities calculated at different pHs (pH = 10 and pH = 11),
  • CACO2 and SKIN-permeability/solubility in skin cells and CACO2,
  • L1 and 2 GS-solubility for Legendre-polynomials that are used to differentiate solubility between similar compounds,
  • FLEX_RB–A parameter related to flexibility and the number of rotational links. The Flex descriptor represents the maximum flexibility of a molecule. It is the result of the average of the differences between the maximum and minimum distance of every atom with the others searched on 50 random conformers). Flex_RB descriptor is the ratio between Flex and the number of rotable bonds.
Most of these descriptors relate to the influence of solubility on antifungal activity, consistent with the interaction maps for the GRID force field. Studies regarding the toxicity of these compounds are underway; in the future, we expect to use these active agents in formulations that have higher solubilities.

2.4. QSAR-3D and PLS

Significant results were obtained using the antifungal activity against C. krusei. The best model and its statistical indices are presented in equations (1) and (2). Figure 6 reports the arrangement of selected objects to form the training set (experimental versus calculated activity values).
Figure 6. Graph of experimental versus calculated activity values, predicted by Equation (1).
Figure 6. Graph of experimental versus calculated activity values, predicted by Equation (1).
Molecules 17 02298 g006
−log(MFC) = + 15.99 (± 5.34) H7m − 47.48 (±19.81) R4u+ + 141.78 (± 36.53) R7u+ + 44.17 (± 17.80) R3v+ − 1.73 (± 1.37)
(n = 20; r2 = 0.87; s = 0.32; F = 26.04; Q2cv = 0.78; SPRESS = 0.36)
The descriptors of the DRAGON [27,28] in the best model were:
  • H7m–related to the probability of interaction between atoms and topological distance 7.
  • R4u+, R7u+ and R3v+–related to the maximum steric contributions to the molecule shape related to the atoms with topological distance of 4, 7 and 3, respectively. These descriptor have greater values for higher influence on the shape and the lower geometric distance between them.
The evaluation set was made by five compounds, rationally selected as described by Golbraikh and colleagues [29]. Table 6 presents the calculated pIC50 values for the set of assessment, considering the descriptors of the selected QSAR model for the training set, the experimental pIC50 values, the residues values (difference between the values of calculated and observed activity) and the standard deviation (SD) of the mean of the residues. The compound that showed residues value greater than the average deviation was not well predicted by the selected model.
Table 6. pIC50 values, residues and standard deviation (SD) of the mean of the residues of the tested series.
Table 6. pIC50 values, residues and standard deviation (SD) of the mean of the residues of the tested series.
pIC50 calculatedpIC50 observedResiduesSD
2.542.380.160.26
3.873.700.17
3.723.670.05
2.281.530.75
2.272.090.18
The partial least squares regression (PLS), was applied to the same set with twenty compounds. The PLS analysis, with five LVs ad 580 descriptors selected, yields a model with an r2 of 0.91 and a q2cv of 0.77 after variable selection via fractional factorial design (FFD) (Table 7 and Figure 7). The internal validation (cross-validation test) of PLS model was performed by the leave-one-out (LOO).
The Pentacle software is a computational tool for computing alignment-free molecular descriptors, also called GRid-INdependent descriptors or GRIND [30]. The software is based on Molecular Interaction Fields, describe the ability of the molecules to interact with other molecules and do not require to superimpose the compounds. After treatment of the descriptors with CLACC algorithm, we observed that the auto-correlogram of the probe shape TIP-TIP produced variables of greater values in active compounds, emphasizing physical and chemical characteristics, mainly steric, similar to the QSAR model.
Table 7. Statistical parameters of PLS model.
Table 7. Statistical parameters of PLS model.
LVSDECSDEPR2Q2
50.180.480.910.77
Figure 7. Best fit model obtained in PLS.
Figure 7. Best fit model obtained in PLS.
Molecules 17 02298 g007

3. Experimental

3.1. Antifungal Activity

The in vitro antifungal activities of compounds 150 were investigated with regard to two species of pathogenic fungi, Candida krusei LM08 and Criptococcus neoformans ICB59. These strains were supplied by the URM Culture Collection of the Department of Mycology, Department of Pharmaceutical Sciences of the Federal University of Paraíba, Brazil. MIC values were determined by the microdilution broth method using microdilution plates according to the guidelines of the National Committee for Clinical and Laboratory Standards (NCCLS) [31]. Tests were performed according to Nascente et al. [32].
All strains were stocked in mineral oil at 18 °C. Viability tests and subsequent taxonomic confirmations were carried out according to Barnett et al. [33]. To obtain an inoculum of 2.5 × 103 cells/mL, each strain was cultured in a tube containing 20 mL of Sabouraud Dextrose Agar plus yeast extract at 35 °C for two days. Subsequently, suspensions were prepared in sterile physiological solution (0.85%) and mixed in a rotator shaker. The inoculum was adjusted to 90% transmittance at 530 nm (measured via a spectrophotometer). Stock solutions of tested compounds were freshly prepared and dissolved in dimethylsulfoxide (DMSO), aliquoted and stored at −20 °C in stock solution with a final concentration of 1,024 μg/mL. Decimal dilutions of thiophene stock solutions were prepared in RPMI 1640 (Sigma Chemical Co., St., Louis, MO, USA) and buffered to pH 7.0 with 0.165 M morpholinopropanesulphonic acid (MOPS, Sigma). Microdilution plates containing serial dilutions of each compound (from 1,024 µg/mL to 1 µg/mL and from 10,000 µg/mL to 1,250 µg/mL (for the inactive compounds in the first serial dilution)) were inoculated with each organism; each plate included positive controls (fungi without a compound) and a negative control (medium only). The microdilution plates were incubated at 35 °C and were read visually after 24 and 72 h of incubation. For each strain, the MIC was defined as the lowest concentration of antifungal agent that completely inhibited the growth of the organism, as detected by the unaided eye and compared to the control group. All tests were performed in duplicate, and the results were expressed as the geometric mean of the MIC values obtained in the two trials [34,35].

3.2. Molecular Modelling

Using the program Hyperchem v. 8.0 [36], the chemical structures of the compounds of interest were drawn, and their geometry was optimised using MM+ force field [37]. Afterwards, we performed a new geometry optimisation based on the semi-empirical method AM1 (Austin Model 1) [38]. The optimised structures were subjected to conformational analyses using the random search method [20,21] with 1,000 interactions, 100 cycles of optimisation, and 10 conformers of lowest minimum energy. The selected dihedrals were evaluated by rotation in accordance with the standard (default) conditions of the program, in which the number of simultaneous variations was 1 to 8, acyclic chains were submitted to rotations from 60 to 180° and torsion rings were in the range of 30 to 120°.
The conformers of lowest minimum energy were selected, saved as .mol and exported to the Spartan 8 program [39]. The lowest energy conformers were then saved in .sdf format and imported into VolSurf+ for Windows [25,26]. They were subjected to the GRID force field, and interaction maps with the following “probes” were generated: H2O (light blue), O (red), N (dark blue) and DRY (green).

3.3. Chemometrics

The structures modelled as described above were used as the initial structures to calculate the molecular descriptors via the VolSurf+ program [25,26]. PCA and CPCA methodologies were applied to the set of interest using the VolSurf+ software [40,41,42,43,44,45].

3.4. QSAR

The conformers of lowest minimum energy obtained in the molecular modelling were exported to the DRAGON 5.4 program [27]. The selection of the descriptors was performed by multiple linear regression with the genetic algorithm GA-VSS of the MobyDigs 1.1 program [46]. Twenty compounds were used as training set and five compounds were used as a test (all of them rationally selected).

3.5. PLS

The conformers of lowest minimum energy obtained in the molecular modelling were imported by the Pentacle 1.5 program [30]. The partial least squares regression (PLS) was applied to a set of twenty compounds. The treatment of the descriptors was done with CLACC algorithm, using the Pentacle program.

4. Conclusions

Studies regarding the density functional theory emphasised the influence of water solubilities on the antifungal activities (with respect to strains of Candida krusei and Criptococcus neoformans) of 2-aminothiophene derivatives. CPCA and PCA analyses corroborated these results. The chemometric tools applied in this study generated good exploratory results. The VolSurf+ descriptors showed that hydrophilic properties are strongly correlated to the biological data. Our studies using Dragon and Pentacle programs emphasized the importance of the form and steric parameters for the antifungal activity.

Conflict of Interest

The authors declare no conflicts of interest.

Acknowledgements

We are indebted to the Brazilian National Research Council (CNPq) for financial support. L. Scotti, a researcher, is grateful to this organisation for a fellowship. The authors are grateful to Prof. Gabriele Cruciani from the University of Perugia, Italy, for the VolSurf+ license. We also thank UEPB (Research and Post-Graduation Incentive Program/PRPGP) for financial support.
  • Sample Availability: Samples of the compounds 150 are available from the authors.

References and Notes

  1. Brown, J.M. Fungal infections in bone marrow transplant patients. Curr. Opin. Infect. Dis. 2004, 17, 347–352. [Google Scholar] [CrossRef]
  2. Bergold, A.M.; Georgiadis, S. New antifungic drugs: A review. Visão Acadêmica 2004, 5, 159–172. [Google Scholar]
  3. Ram, V.J.; Goel, A.; Shukla, P.K.; Kapil, A. Synthesis of thiophenes and thieno[3,2-c]pyran-4-ones as antileishmanial and antifungal agents. Bioorg. Med. Chem. Lett. 1997, 7, 3101–3106. [Google Scholar] [CrossRef]
  4. Chobot, V.; Buchta, V.; Jahodárová, H.; Pour, M.; Opletal, L.; Jahodár, L.; Harant, P. Antifungal activity of a thiophenepolyine from Leuzeacarthamoides. Phytotherapy 2003, 74, 288–290. [Google Scholar]
  5. Wardakhan, W.W.; Louca, N.A.; Kamel, M.M. The reaction of 2 Aminocyclohexeno[b]thiophene derivatives with ethoxycarbonylisothiocyanate: Synthesis of fused thiophene derivatives with antibacterial and antifungal activities. Acta Chim. Slov. 2007, 54, 229–241. [Google Scholar]
  6. Shaaban, M.R.; Saleh, T.S.; Farag, A.M. Synthesis and antimicrobial evaluation of new thiophene and 1,3,4-thiadiazole derivatives. 2009, 78, 151–159. [Google Scholar]
  7. Islor, A.M.; Kalluraya, B.; Pai, K.S. Synthesis, characterization and biological activities of some new benzo[b]thiophene derivatives. Eur. J. Med. Chem. 2010, 45, 825–830. [Google Scholar] [CrossRef]
  8. Fokialakis, N.; Cantrell, C.L.; Duke, S.O.; Skaltsounis, A.L.; Wedge, D.E. Antifungal activity of thiophenes from Echinops ritro. J. Agric. Food Chem. 2006, 54, 1651–1655. [Google Scholar] [CrossRef]
  9. Pinto, E.; Queiroz, M.J.R.P.; Vale-Silva, L.A.; Oliveira, J.F.; Begoin, A.; Begoin, J.M.; Kirsch, G. Antifungal activity of synthetic di(hetero)arylamines based on the benzo[b]thiophene moiety. Bioorg. Med. Chem. 2008, 16, 8172–8177. [Google Scholar] [CrossRef]
  10. Mendonça Junior, F.J.B.; Lima-Neto, R.G.; de Oliveira, T.B.; Lima, M.C.A.; Pitta, I.R.; Galdino, S.L.; da Cruz, R.M.D.; de Araújo, R.S.A.; Neves, R.P. Synthesis and evaluation of the antifungal activity of 2-(substituted-amino)-4,5-dialkyl-thiophene-3-carbonitrile derivatives. Lat. Am. J. Pharm. 2011, 30, 1492–1499. [Google Scholar]
  11. Abi-Said, D.; Anaissie, E.; Uzum, O.; Raad, I.; Pinzcowski, H.; Vartivarian, S. The epidemiology of hematogenouscandidiasis caused by different Candida species. Clin. Infect. Dis. 1997, 24, 1122–1128. [Google Scholar]
  12. Berrouane, Y.F.; Hollis, R.J.; Pfaller, M.A. Strain variation among and antifungal susceptibilities of isolates of Candida krusei. J. Clin. Microbiol. 1996, 34, 1856–1858. [Google Scholar]
  13. Pfaller, M.A.; Diekema, D.J.; Gibbs, D.L.; Newell, V.A.; Nagy, E.; Dobiasova, S.; Rinaldi, M.; Barton, R.; Veselov, A. Global Antifungal Surveillance Group. Candida krusei, a multidrug-resistant opportunistic fungal pathogen: Geographic and temporal trends from the ARTEMIS DISK antifungal surveillance program, 2001 to 2005. J. Clin. Microbiol. 2008, 46, 515–521. [Google Scholar]
  14. Merz, W.G.; Karp, J.E.; Schron, D.; Saral, R. Increased incidence of fungemia caused by candida krusei. J. Clin. Microbiol. 1986, 24, 581–584. [Google Scholar]
  15. Cheung, M.C.; Rachlis, A.R.; Shumak, S.L. A cryptic cause of cryptococcal meningitis. Can. Med. Assoc. J. 2003, 168, 451–452. [Google Scholar]
  16. Wormley, F.L., Jr.; Perfect, J.R.; Steele, C.; Cox, G.M. Protection against Cryptococcosis by Using a Murine Gamma Interferon-Producing Cryptococcus neoformans Strain. Infect. Immun. 2007, 75, 1453–1462. [Google Scholar] [CrossRef]
  17. Beebe, K.R.; Pell, R.J.; Seasholtz, M.B. Chemometrics: A Practical Guide; Wiley: New York, NY, USA, 1998. [Google Scholar]
  18. Sharaf, M.A.; Illman, D.L.; Kowalski, B.R. Chemometrics; John Wiley & Sons: New York, NY, USA, 1986. [Google Scholar]
  19. Westerhuis, J.A.; Kourti, T.; Macgregor, J.F. Analysis of multiblock and hierarchical PCA and PLS models. J. Chemometr. 1998, 12, 301–321. [Google Scholar] [CrossRef]
  20. Cohen, N.C. Guidebook on Molecular Modeling in Drug Design; Academic Press: San Diego, CA, USA, 1996. [Google Scholar]
  21. Leach, A.R. Molecular Modeling: Principles and Applications; Prentice Hall: London, UK, 2001. [Google Scholar]
  22. de Oliveira, T.B. Síntese, elucidação estrutural e avaliação biológica de novos derivados tiazolidínicos e cicloalquil-tiofênicos. M.Sc. Thesis, Graduate Program in Pharmaceutical Sciences. Federal University of Pernambuco, Recife, Brazil, 2010. [Google Scholar]
  23. Souza, B.C.C. Síntese e Avaliação das atividades Antifúngicas e Antitumorais de novos derivados 2-[(arilideno)amino]-cicloalquil[b]tiofenos-3-carbonitrila. M.Sc. Thesis. Graduate Program in Pharmaceutical Sciences. Federal University of Pernambuco, Recife, Brazil, 2011. [Google Scholar]
  24. Abu-Hashem, A.A.; El-Shehry, M.F.; Badria, F.A. Design and synthesis of novel thiophenecarbohydrazide, thienopyrazole and thienopyrimidine derivatives as antioxidant and antitumor agents. Acta Pharm. 2010, 60, 311–323. [Google Scholar]
  25. Cruciani, G.; Crivori, P.; Carrupt, P.-A.; Testa, B. Molecular fields in quantitative structure-permeation relationships: The VolSurf approach. J. Mol. Struct. 2000, 503, 17–30. [Google Scholar]
  26. Cruciani, G.; Pastor, M.; Guba, W. VolSurf: A new tool for the pharmacokinetic optimization of lead compounds. Eur. J. Pharm. Sci. 2000, 11, S29–S39. [Google Scholar]
  27. Dragon, Version 5.4; Talete Inc.:: Milano, Italy,, 2006. Available online: http://www.talete.mi.it/products/dragon_description.htm (accessed on 2 January 2012).
  28. Todeschini, R.; Consonni, V. Handbook of Molecular Descriptors; Wiley-VCH: Weinheim, Germany, 2000. [Google Scholar]
  29. Golbraikh, A.; Shen, M.; Xiao, Z.; Xiao, Y.-D.; Lee, K.-H.; Tropsha, A. Rational selection of training and test sets for the development of validated QSAR models. J. Comput. Aid. Mol. Des. 2003, 17, 241–253. [Google Scholar] [CrossRef]
  30. Pastor, M.; Mclay, I.; Pickett, S.; Clementi, S. Grid-Independent descriptors (GRIND): A novel class of alignment-independent three-dimensional molecular descriptors. Med. Chem. 2000, 43, 3233–3243. [Google Scholar] [CrossRef]
  31. National Committee for Clinical and Laboratory Standards; Method M27-A2; NCCLS: Wayne, MI, USA, 2002; pp. 1–29.
  32. Nascente, P.S.; Meinerz, A.R.M.; Faria, R.O.; Schuch, L.F.D.; Meireles, M.C.A.; de Mello, J.R.B. CLSI broth microdilution method for testing susceptibility of Malassezia pachydermatis to thiabendazole. Braz. J. Microbiol. 2009, 40, 222–226. [Google Scholar] [CrossRef]
  33. Barnett, J.A.; Paire, R.W.; Yarrow, D. Yeasts: Characteristics and Identification; Cambridge University Press: Cambridge, UK, 2000. [Google Scholar]
  34. Eloff, J.N. A sensitive and quick microplate method to determine the minimal inhibitory concentration of plant extracts for bacteria. Planta Med. 1978, 64, 711–713. [Google Scholar] [CrossRef]
  35. Souza, E.L.; Stamford, T.L.M.; Lima, E.O.; Trajano, V.N. Effectiveness of Origanum vulgare L. essential oil to inhibit the growth of food spoiling yeasts. Food Control 2007, 18, 409–413. [Google Scholar] [CrossRef]
  36. HyperChem, Version 8.0, Hybercube Inc.: Gainesville, FL, USA, 2009.
  37. Allinger, N.L. A hydrocarbon force-field utilizing V1 and V2 torsional terms. J. Am. Chem. Soc. 1977, 99, 8127–8134. [Google Scholar] [CrossRef]
  38. Dewar, M.J.S.E.; Zoebisch, G.; Healy, E.F.; Stewart J.J.P. AM1: A new general purpose quantum mechanical molecular model. J. Am. Chem. Soc. 1985, 107, 3902–3909. [Google Scholar] [CrossRef]
  39. Spartan Model, Version 8.0; Wavefunction, Inc.: Irvine, CA, USA, 2008. Available online: http://www.wavefun.com/products/windows/SpartanModel/win_model.html (accessed on 2 October 2011).
  40. Zamora, I.; Oprea, T.; Cruciani, G.; Pastor, M.; Ungell, A.-L. Surface descriptors for protein-ligand affinity prediction. J. Med. Chem. 2003, 46, 25–33. [Google Scholar] [CrossRef]
  41. Crivori, P.; Cruciani, G.; Carrupt, P.-A.; Testa, B. Predicting blood-brain barrier permeation from three-dimensional molecular structure. J. Med. Chem. 2000, 43, 2204–2216. [Google Scholar] [CrossRef]
  42. Kovatcheva, A.; Golbraikh, A.; Oloff, S.; Xiao, Y.-D.; Zheng, W.; Wolschann, P.; Buchbauer, G.; Tropsha, A. Combinatorial QSAR of ambergris fragrance compounds. J. Chem. Inf. Comp. Sci. 2004, 44, 582–595. [Google Scholar] [CrossRef]
  43. Oprea, T.I.; Zamora, I.; Ungell, A.-L. Pharmacokinetically based mapping device for chemical space navigation. J. Comb. Chem. 2002, 4, 258–266. [Google Scholar] [CrossRef]
  44. Cruciani, G.; Pastor, M.; Mannhold, R. Suitability of molecular descriptors for database mining. A comparative analysis. J. Med. Chem. 2002, 45, 2685–2694. [Google Scholar] [CrossRef]
  45. Cianchetta, G.; Mannhold, R.; Cruciani, G.; Baroni, M.; Cecchetti, V.J. Chemometric studies on the bactericidal activity of quinolones via an extended volsurf approach. J. Med. Chem. 2004, 47, 3193–3201. [Google Scholar] [CrossRef]
  46. MobyDigs, Version 1.1; Talete Inc.:: Milano, Italy,, 2009. Available online: http://www.talete.mi.it/products/moby_description.htm (accessed on 2 January 2012).

Share and Cite

MDPI and ACS Style

Scotti, L.; Tullius Scotti, M.; De Oliveira Lima, E.; Sobral da Silva, M.; Do Carmo Alves de Lima, M.; Da Rocha Pitta, I.; Olímpio de Moura, R.; Gonzaga Batista de Oliveira, J.; Duarte da Cruz, R.M.; Bezerra Mendonça, F.J., Júnior. Experimental Methodologies and Evaluations of Computer-Aided Drug Design Methodologies Applied to a Series of 2-Aminothiophene Derivatives with Antifungal Activities. Molecules 2012, 17, 2298-2315. https://doi.org/10.3390/molecules17032298

AMA Style

Scotti L, Tullius Scotti M, De Oliveira Lima E, Sobral da Silva M, Do Carmo Alves de Lima M, Da Rocha Pitta I, Olímpio de Moura R, Gonzaga Batista de Oliveira J, Duarte da Cruz RM, Bezerra Mendonça FJ Júnior. Experimental Methodologies and Evaluations of Computer-Aided Drug Design Methodologies Applied to a Series of 2-Aminothiophene Derivatives with Antifungal Activities. Molecules. 2012; 17(3):2298-2315. https://doi.org/10.3390/molecules17032298

Chicago/Turabian Style

Scotti, Luciana, Marcus Tullius Scotti, Edeltrudes De Oliveira Lima, Marcelo Sobral da Silva, Maria Do Carmo Alves de Lima, Ivan Da Rocha Pitta, Ricardo Olímpio de Moura, Jaismary Gonzaga Batista de Oliveira, Rayssa Marques Duarte da Cruz, and Francisco Jaime Bezerra Mendonça, Júnior. 2012. "Experimental Methodologies and Evaluations of Computer-Aided Drug Design Methodologies Applied to a Series of 2-Aminothiophene Derivatives with Antifungal Activities" Molecules 17, no. 3: 2298-2315. https://doi.org/10.3390/molecules17032298

Article Metrics

Back to TopTop