Next Article in Journal
Isolation and Identification of Myo-Inositol Crystals from Dragon Fruit (Hylocereus polyrhizus)
Previous Article in Journal
Systematic Evaluation of Structure-Activity Relationships of the Riminophenazine Class and Discovery of a C2 Pyridylamino Series for the Treatment of Multidrug-Resistant Tuberculosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Support Vector Machine Classification Model for Benzo[c]phenathridine Analogues with Topoisomerase-I Inhibitory Activity

Department of Medicinal Chemistry, School of Pharmacy, University of Medicine and Pharmacy at Ho Chi Minh City, 41 Dinh Tien Hoang St., District 1, Ho Chi Minh City, Vietnam
*
Author to whom correspondence should be addressed.
Molecules 2012, 17(4), 4560-4582; https://doi.org/10.3390/molecules17044560
Submission received: 13 March 2012 / Revised: 8 April 2012 / Accepted: 10 April 2012 / Published: 17 April 2012
(This article belongs to the Section Medicinal Chemistry)

Abstract

:
Benzo[c]phenanthridine (BCP) derivatives were identified as topoisomerase I (TOP-I) targeting agents with pronounced antitumor activity. In this study, a support vector machine model was performed on a series of 73 analogues to classify BCP derivatives according to TOP-I inhibitory activity. The best SVM model with total accuracy of 93% for training set was achieved using a set of 7 descriptors identified from a large set via a random forest algorithm. Overall accuracy of up to 87% and a Matthews coefficient correlation (MCC) of 0.71 were obtained after this SVM classifier was validated internally by a test set of 15 compounds. For two external test sets, 89% and 80% BCP compounds, respectively, were correctly predicted. The results indicated that our SVM model could be used as the filter for designing new BCP compounds with higher TOP-I inhibitory activity.

Graphical Abstract

1. Introduction

The topoisomerases (TOP) are enzymes involved in processes such as replication, repair, transcription, recombination and segregation of DNA. Those of type I are the target of several anticancer agents based on their ability to stabilize the DNA-enzyme cleavage complex that causes DNA damage and cytotoxicity [1,2]. Among the agents expressing a targeted anti-topoisomerase activity, alkaloids of the benzo[c]phenanthridines (BCPs) family are well known [1,2,3,4]. Many BCP analogues were synthesized and evaluated for their activity on topoisomerase I as well as their cytotoxicity. Of those, ethoxidine, NK-109 and topovale (ARC 111) are potential candidates for cancer chemotherapy [1,2,3,4,5] (Figure 1).
Figure 1. Chemical structures of benzo[c]phenanthridine derivatives.
Figure 1. Chemical structures of benzo[c]phenanthridine derivatives.
Molecules 17 04560 g001
The in vitro TOP-I inhibitory activity is valued by REC value, which is the relative effective concentration of TOP-I related to topotecan [6,7,8,9,10,11,12,13,14,15,16,17]. Therefore, compounds having TOP-I inhibitory activity could be divided into two classes of compounds based on the topotecan threshold. In this study, a support vector machine (SVM) approach was used to build up a classification model based on the anti-topoisomerase-I activity for BCP analogues. The SVM model could be applied to seek out new BCP analogues which are inhibitors of TOP-I for cancer treatment.

2. Results and Discussion

2.1. Feature Selection

Three feature selection approaches, namely mRMR (Max-Relevance, Min-Redundancy), GA (genetic algorithm) and RF (random forest) were applied to the dataset to figure out a set of chemical descriptors related to bioactivity properties from 2,032 molecular descriptors calculated by Dragon. Based on the calculation, three sets of molecular descriptors were selected, including a set of 10, 16 and seven descriptors for the mRMR, GA and RF methods, respectively. Table 1 shows the molecular descriptors selected via the RF method. After that, these sets of descriptors were used to develop the SVM models. For any SVM, it is necessary to select the optimal parameters of the Kernel function (C, γ) and these values for each set of descriptors were figured out and are shown in Table 2. The results of classification based on three different descriptor sets (by mRMR, GA and RF) by the SVM algorithm (the e1071 package in R) are shown in Table 3. The total accuracies of these SVM models are greater than or equal to 0.91 on training sets and are greater than or equal to 0.74 for cross-validation. However, the basic principle of SVM is a supervised learning approach. Hence, the classification results regarding test set and external set showed the relatively accurate performance of models. The SVM model which was developed from descriptors selected by RF method gave a better classification power than the two other methods.
Table 1. Molecular descriptors from RF feature selection method.
Table 1. Molecular descriptors from RF feature selection method.
ClassSymbolDefinition
Topological descriptors

D/Dr05distance/detour ring index of order 5
D/Dr06distance/detour ring index of order 6
Walk and path counts

MPC06molecular path count of order 06
MPC08molecular path count of order 08
MPC10molecular path count of order 10
2D frequency fingerprintsF05[C–C]frequency of C–C at topological distance 05
F08[NO]frequency of N–O at topological distance 08
Table 2. Optimal parameters (C, γ) for SVM approach.
Table 2. Optimal parameters (C, γ) for SVM approach.
Feature selection methodNumber of features selectedRange of γBest γ Range of CBest CCross-validation error
mRMR105[−10:10]3.1255[−10:10]10.21
GA162[−10:10]0.1252[−10:10]160.15
RF72[−10:10]0.252[−10:10]40.17
Table 3. Classification results of 3 SVM models corresponding with 3 descriptor sets selected successively via mRMR, GA and RF methods.
Table 3. Classification results of 3 SVM models corresponding with 3 descriptor sets selected successively via mRMR, GA and RF methods.
Feature selection methodTraining setTest setExternal set
mRMRGARFmRMRGARFmRMRGARF
Number of support vectors514035
Total accuracy0.980.910.930.800.930.870.670.780.89
Sensitivity0.930.730.870.331.001.000.400.600.80
Specificity1.000.980.950.920.920.831.001.001.00
Positive precision 1.000.920.870.500.750.601.001.001.00
Negative precision 0.980.910.950.851.001.000.570.670.80
Matthews correlation cofficient (MCC)0.960.770.820.290.830.710.480.630.80
Total accuracy of cross-validation0.760.780.74

2.2. SVM Classification Model

A total of 82 compounds were used for SVM and classified into actives or inactives based on the the relative effective concentration (REC) compared to topotecan and the respective threshold points. For the separation of active/inactive, REC of topotecan (REC = 1) was selected as a threshold point. TOP-I active compound having equal or stronger activity than that of topotecan is presented as “1” and vice versa, “0” is presented TOP-I inactive compound having weaker activity than that of topotecan. The set of molecular descriptors selected via RF method was used to create the final SVM (the e1071 package in R) classification model based on anti-topoisomerase-1 activity. Moreover, two other classification approaches namely the SVM kernlab and “randomForest” (RF) in R are also applied and the classification results are presented in Table 4. In general, both of SVM packages (the e1071 package, the package kernlab) gave better results than those of observed for random forest methods with the different values in the total accuracy of 20%.
Table 4. Classification results from three classified approaches.
Table 4. Classification results from three classified approaches.
Training setTest setExternal setApplication set
Evaluation criteriaSVM e1071SVM KernlabRFSVM e1071SVM kernlabRFSVM e1071SVM kernlabRFSVM e1071
TP131363334335
TN414136108104433
FP2290001220
FN2272420012
Total accuracy0.930.930.720.870.730.870.890.780.670.80
Sensitivity0.870.870.460.600.430.601.001.000.750.71
Specificity0.950.950.801.001.001.000.800.670.601.00
Positive precision 0.870.870.401.001.001.000.800.600.601.00
Negative precision 0.950.950.840.830.670.831.001.000.750.60
MCC0.820.820.250.710.540.710.800.630.350.65
Cross-validation error0.220.200.28 a
Y-scrambling total accuracy0.59
a “out of bags” of RF method is based on principle of cross-validation.
Regarding the training set, two SVM packages (e1071, kernlab) expressed similar classification power. However, regarding the test set and the external set, the SVM package e1071 showed the better results. The Matthews correlation coefficient (MCC) of this package was stable for training, test and external test sets with values 0.82, 0.71 and 0.80, respectively. For the kernlab package, the MCC value for the training set was as high as 0.82 whereas the MCC of the test set was only 0.54. Normally, a model with MCC larger than 0.4 indicates that it has the predictive power [18]. The RF classification method in combination with a set of descriptors chosen by RF gave MCC = 0.35 on the external set and this result indicates that the RF classification model has no ability to predict the biological activity of compounds in this study. The SVM-e1071 in R classification model with a set of descriptors identified from a large set via a RF algorithm showed the best results.
To further validate the use of SVM-e1071 for TOP-I classification, a 50-fold Y-scrambling procedure for the training sets was performed [19,20,21]. The total accuracy of 0.59 was obtained from the Y-scrambling analyses. The results showed that our models show a significantly better performance than those obtained when class assignments are randomly achieved with total accuracy values of 0.93 versus 0.59, respectively. However, it has to be noted that the principle of SVM is supervised learning and the learning strategy tries to keep the error to a minimum values which explains the relative “good” performance of this method for the y scrambled data sets.

2.3. Validation and Application

The final SVM model was validated not only by cross-validation procedure but also by an external dataset (not belonging to the dataset used to create the SVM classification model). The power for classification of BCPs by anti-TOP-I activity was 0.87–0.93 for total accuracy and 0.22 for cross-validation error. Moreover, the final SVM model was applied to classify a set of 10 BCP analogues recently synthesized by Lavoie et al., which so-called application set or external test set 2 and detailed chemical structures and TOP-I activities are shown in Table 5 [22,23,24,25,26]. According to literature, among these 10 BCP analogues, seven compounds have stronger anti-TOP-I activity than topotecan and three compounds are weaker than topotecan. The results indicated that the classification model was achieved a correct prediction of 80% (8/10) and the detailed results are presented in Table 4. The positive accuracy gained the value of 100% i.e., the classification model is more accurate for predicting compounds wtith stronger activity than topotecan.
Table 5. Chemical structure often benzo[c]phenanthridine derivatives in application set and their topoisomerase I inhibitory activity REC and classification results from final SVM model. Classification term: “1” presented TOP-I active compound having equal or stronger activity than that of topotecan; “0” presented TOP-I inactive compound having weaker activity than that of topotecan.
Table 5. Chemical structure often benzo[c]phenanthridine derivatives in application set and their topoisomerase I inhibitory activity REC and classification results from final SVM model. Classification term: “1” presented TOP-I active compound having equal or stronger activity than that of topotecan; “0” presented TOP-I inactive compound having weaker activity than that of topotecan.
NoChemical structureNameREC TOP-I mediated DNA cleavage (Experimental)TOP-I Active/InactiveTOP-I classified result from final SVM model
A1 Molecules 17 04560 i001BMC_08_7824_7a0.0311
A2BMC_08_7824_7b0.0811
A3BMC_08_7824_90.211
A4BMC_08_7824_110.111
A5BMC_08_8598_9>1000
A6BMC_08_8598_100.210
A7BMC_08_8598_120.210
A8BMC_08_8598_130.211
A9BMC_08_8598_14>1000
A10BMC_08_8598_15>1000
This result completely meets our goal, which is aiming to look for new inhibitors of TOP-I for cancer chemotherapy. However, one limitation of all machine learning approaches is their inability to indicate the important role of functional groups related to biological activity. Hence, the combination of this SVM classification model with molecular docking studies [27,28] and also related 2D- and 3D-QSAR model on cytotoxicity of BCPs [5] could provide insight into the molecular basis of TOP-I inhibitors.

2.4. Discussions

In this study, the TOP-I inhibitory activity of topotecan, the synthetic derivative of camptothecin and the most potent anticancer drugs in clinical use, is used as threshold points for SVM classification models. Topotecan, ethoxidine, fagaronine and BCP related compounds indicated the selectivity on TOP-I than TOP-II. These novels acted as DNA intercalators and having two mechanisms including (i) TOP-I poison as fagaronine; and (ii) TOP-I suppressor as ethoxidine [27,28]. Our preliminary results from in silico modeling indicated that BCP compounds may inhibit the TOP-I activity via suppression mechanism.
SVM is a machine learning method, which has been used for many kinds of pattern recognition problems. SVMs have been successfully adapted to treat both regression and classification [29]. In this study, mRMR, GA and RF algorithms were used for the selection of descriptors from large descriptors set derived from Dragon software and the results showed that the descriptors selected from the RF method gave a better classification power than others. The SVM also indicated the better results on classification of BCP analogues with anti-TOP-I activity than that of RF. This final SVM model with its high accuracy (80–90%), fast calculation without the need of 3D conformation and accurate prediction on the substances having positive activity could be applied to look for and design new analogues of BCPs with higher topoisomerase I inhibitory activity. SVMs are a black box technique and deliver information about their predictions other than the relationship between molecular descriptors and bioactivity. The SVM model’s explanatory in term molecular descriptors and bioactivity could be performed if the weights of each descriptor are explicitly solved [29]. It should be noted that our SVM model could be applied in screening prior to synthesis procedures for new BCP-like compouds to identify potential TOP-I inhibitors. However, the development of a useful drug has to be dealt on systems level of drug discovery and development and take into consideration many factors including solubility, ease of drug formulation, selectivity and ADME-Tox.

3. Experimental

3.1. Dataset

BCP analogues in the study (82 compounds) with topoisomerase I inhibitory activity evaluated by testing of DNA cleavage were collected from the work of LaVoie et al. [6,7,8,9,10,11,12,13,14,15,16,17]. The biological activity data is represented by the relative effective concentration (REC) compared to topotecan whose value is arbitrarily set at 1.0 (as reference compound). This parameter allows the comparison of molecules based on the cleavage of plasmid DNA in the presence of human topoisomerase I. Therefore, with a compound having the REC value higher than 1, that means its topoisomerase I inhibitory activity lower than topotecan and vice versa. Based upon the REC values, the data set was grouped into two classes: 58 analogues were assigned to inactive class (REC value > 1, negatives or inactives) and 24 analogues were classified as active class with REC value ≤ 1 (high TOP-I blockade, positives or actives). Chemical structure of 82 benzo[c]phenanthridine derivatives and their topoisomerase I inhibitory activity REC values were presented in Table 6.
Table 6. Chemical structure of82 benzo[c]phenanthridine derivatives and their topoisomerase I inhibitory activity REC and classification results from final SVM model. Classification term: “1” presented TOP-I active compound having equal or stronger activity than that of topotecan; “0” presented TOP-I inactive compound having weaker activity than that of topotecan.
Table 6. Chemical structure of82 benzo[c]phenanthridine derivatives and their topoisomerase I inhibitory activity REC and classification results from final SVM model. Classification term: “1” presented TOP-I active compound having equal or stronger activity than that of topotecan; “0” presented TOP-I inactive compound having weaker activity than that of topotecan.
NoChemical structureNameREC TOP-I mediated DNA cleavage (Experimental)TOP-I Active/InactiveTOP-I classified result from final SVM model
1 Molecules 17 04560 i002Nitidine1000
2BMC_03_3795_10a800
3BMC_03_3795_10b20000
4BMC_03_3795_10c20000
5BMC_03_3795_10d>100000
6BMC_03_3795_10e50000
7BMC_03_3795_10f1000
8BMC_03_3795_11a>100000
9BMC_03_3795_11b10000
10BMC_03_3795_12d>100000
11BMC_03_2061_03a0.510
12BMC_03_2061_03b>100000
13BMC_03_2061_03c0.311
14BMC_03_2061_03d100000
15BMC_03_2061_03e110
16BMC_03_2061_03f100000
17BMCL_02_3333_03c1000
18BMC_03_2061_03h5000
19BMC_03_2061_03i110
20BMC_03_2061_03j>100000
21BMC_03_2061_03k0.8–1.011
22BMC_03_2061_04a0.810
23BMC_03_2061_04b10001
24 Molecules 17 04560 i003BMC_03_2061_09k1000
25JMC_03_2254_020.310
26JMC_03_2254_03600
27JMC_03_2254_05a100001
28JMC_03_2254_05b0.111
29JMC_03_2254_06a1500
30JMC_03_2254_06b0.510
31JMC_03_2254_05c0.210
32JMC_03_2254_06c8.000
33JMC_03_2254_16a1001
34JMC_03_2254_17a50000
35JMC_03_2254_020.511
36BMCL_02_3333_02>100000
37BMCL_02_3333_0320000
38BMCL_02_3333_04a0.311
39BMCL_02_3333_04b100000
40BMCL_02_3333_04c3000
41BMCL_02_3333_04d1.010
42LDDD_04_198_010.0310
43 Molecules 17 04560 i004LDDD_04_198_022.000
44LDDD_04_198_032.000
45BMC_04_3731_03a900
46BMC_04_3731_03b600
47BMC_04_3731_03c200
48BMC_04_3731_03d>30000
49BMC_04_3731_04a10000
50BMC_04_3731_04b1200
51BMC_04_3731_04c600
52BMC_04_0795_04>30000
53BMC_04_3731_051000
54BMC_04_0795_01b0.310
55BMC_04_0795_01f0.211
56BMC_04_0795_01g2.000
57BMC_04_0795_01h2000
58BMC_04_0795_031000
59BMC_04_0795_02>100000
60BMC_04_0795_040.811
61 Molecules 17 04560 i005BMC_04_0795_05500
62BMC_05_6782_07c1.010
63BMC_05_6782_07d800
64BMC_05_6782_080.611
65BMC_05_6782_09c0.410
66BMC_05_6782_09d1000
67BMC_05_6782_09e0.711
68BMC_05_6782_09f1200
69BMC_05_6782_09g6000
70BMC_05_6782_09h0.611
71BMC_05_6782_09i1.500
72BMC_05_6782_09j0.311
73BMC_05_6782_10301
74BMC_06_3131_10c1.200
75BMC_06_3131_10d0.0711
76BMC_06_3131_10g900
77BMC_06_3131_10i0.4511
78 Molecules 17 04560 i006BMC_06_3131_10j>10000
79BMC_06_3131_10f300
80BMC_06_3131_10k1300
81BMC_06_3131_10l0.3511
82BMC_06_3131_10m0.1511

3.2. Training and Test Sets

The training and test sets were generated by random division. Firstly, a set of nine compounds was selected and not used to develop the models. These compounds were separated from the others and considered as external set. The remaining analogues were split randomly for five times into 80% for training sets and 20% for test sets using the R software [30,31]. The numbers of compounds in each subset are presented in Table 7.
Table 7. BCPs datasets division.
Table 7. BCPs datasets division.
DatasetTotal compoundsActives aInactives b
Whole set822458
Training set581543
Test set15510
External set945
Application set (External set 2)1073
a Actives: Compounds whose activity is equal or stronger than topotecan; b Inactives: compounds whose activity is weaker than topotecan.

3.3. Molecular Descriptors and Feature Selection

The Dragon software was applied to calculate 2,032 molecular descriptors [32,33]. In this study, several feature selection algorithms were performed to reduce dimensionality of descriptor space. Firstly, chemical descriptors only containing the value of zero were eliminated before using the mathematic methods. Only 533 descriptors were selected for next steps. In addition, highly correlated descriptors (r > 0.90) were removed to avoid redundancy and to manage the data more efficiently in terms of computation resources and intuitive perception of the chemical space. A total of 103 descriptors were selected and scaled to unit variance to be used as input value to SVMs. Finally, three feature selection methods namely mRMR (Max-Relevance, Min-Redundancy), GA (genetic algorithm) and RF (random forest) were applied in order to select the optimum set for the molecular descriptors [30,34].

3.4. Support Vector Machine

Support vector machine (SVM), the most successfully applied new classification algorithms, were introduced to the machine learning community by Vapnik. Smola and Schölkopf provided an extensive tutorial on SVMs [30,31,32,33,34]. The underlying idea of an SVM classifier is to map linearly inseparable input data into a higher dimensional space where the data can be linearly separated, using a maximal separating hyperplane. Support Vector Machines (SVM) is the classification system based on the supervised learning approach. In this study, SVM algorithm in the e1071 package in R with Kernel function was used [30,31]. The strategies for the SVM classification model are shown in Figure 2.
Selection of optimal parameters of Kernel function: RBF Kernel function has two parameters (C and γ) and the selection of these parameters is one of two critical issues to develop the good SVM model (the other is the feature selection) [35,36,37]. To find the optimal parameters, an algorithm Grid (function tune in R) is made following the process described in Figure 3. The training set and test set are used to find a pair of optimal parameters (C and γ) of the Kernel function. Pairs of parameters were tested in intervals reduced step by step an algorithm Grid. The pair is chosen when the error of cross validation is minimal [30,34].
Figure 2. Process of SVM classification model.
Figure 2. Process of SVM classification model.
Molecules 17 04560 g002
Figure 3. Selection of optimal parameters of Kernel function using Grid algorithm.
Figure 3. Selection of optimal parameters of Kernel function using Grid algorithm.
Molecules 17 04560 g003

3.5. Evaluation Criteria for Classification Model

Performance of the SVM models was measured by using standard parameters for classification models [18,19,38,39,40]. They are described as follows:
  • (i) The overall classification accuracy of a prediction model, accuracy = (tp + tn)/(tp + fp + tn + fn),
  • (ii) Sensitivity (recall, accuracy on actives) = tp/(tp + fn),
  • (iii) Specificity (accuracy on inactives) = tn/(tn + fp),
  • (iv) Precision on actives = tp/(tp + fp),
  • (v) Precision on inactives = tn/(tn + fn) and,
  • (vi) Matthews correlation coefficient (MCC):
Molecules 17 04560 i007
The quality of a classification model is generally measured by using these parameters, which are estimated for the whole set and by applying a cross-validation protocol based on a leave-one-out (LOO) procedure. In all equations, tp = number of true positives, tn = number of true negatives, fp = number of false positives, and fn = number of false negatives.

4. Conclusions

In this study, the SVM was used to build up a model for prediction and classification of 73 BCP analogues based on their anti-topoisomerase-1 activity. The best model was derived from the SVM-e1071 package in R with the optimal settings of the Kernel function (C = 4, γ = 0.25) and the set of descriptors selected by RF method. This final SVM model is able to correctly predict the anti-topoisomerase activity for 93% of compounds in the training set and 87% of those in a test set. It was also validated on external sets not involving in the dataset used with the developed model and then on the application set. Total accuracies of 89% (the prediction is correct for eight out of nine compounds) and 80% (the prediction is correct for eight out of 10 compounds) were obtained for the external set and application sets, respectively. Furthermore, this model has also proved its ability to classify correctly BCP analogues that have the positive activity, with an accuracy from 80 to 100% overall. With its high accuracy (80–90%), fast and accurate prediction on the substances having positive activity, our SVM model could be applied to look for and design new analogues of BCPs with higher topoisomerase I inhibitory activity.

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/17/4/4560/s1.

Acknowledgments

This work was supported by the Vietnam’s National Foundation for Science and Technology Development—NAFOSTED (Grant # 104.01.21.09 to Khac-Minh Thai). The authors wish to thank Le Thi Tuong Vy and Thi-Cam-Van Vo for critical reading.

Conflict of Interest

The authors declare no conflict of interest.

References and Notes

  1. Fleury, F.; Sukhanova, A.; Ianoul, A.; Devy, J.; Kudelina, I.; Duval, O.; Alix, A.J.P.; Jardillier, J.C.; Nabiev, I. Molecular determinants of site-specific inhibition of human DNA topoisomerase I by fagaronine and ethoxidine. Relation to DNA binding. J. Biol. Chem. 2000, 275, 3501–3509. [Google Scholar]
  2. Pommier, Y. DNA topoisomerase I inhibitors: Chemistry, biology, and interfacial inhibition. Chem. Rev. 2009, 109, 2894–2902. [Google Scholar] [CrossRef]
  3. Nakanishi, T.; Suzuki, M.; Mashiba, A.; Ishikawa, K.; Yokotsuka, T. Synthesis of NK109, an anticancer benzo[c]phenanthridine alkaloid. J. Org. Chem. 1998, 63, 4235–4239. [Google Scholar]
  4. Li, T.K.; Houghton, P.J.; Desai, S.D.; Daroui, P.; Liu, A.A.; Hars, E.S.; Ruchelman, A.L.; LaVoie, E.J.; Liu, L.F. Characterization of ARC-111 as a novel topoisomerase I-targeting anticancer drug. Cancer Res. 2003, 63, 8400–8407. [Google Scholar]
  5. Liao, S.Y.; Qian, L.; Lu, H.L.; Shen, Y.; Zheng, K.C. A combined 2D- and 3D-QSAR study on analogues of ARC-111 with antitumor activity. QSAR Comb. Sci. 2008, 27, 740–749. [Google Scholar] [CrossRef]
  6. Li, D.; Zhao, B.; Sim, S.-P.; Li, T.-K.; Liu, A.; Liu, L.F.; LaVoie, E.J. 2,3-Dimethoxybenzo(i)-phenanthidines: Topoisomerase I-targeting anticancer agents. Bioorg. Med. Chem. 2003, 11, 521–528. [Google Scholar] [CrossRef]
  7. Yu, Y.; Singh, S.K.; Liu, A.; Li, T.K.; Liu, L.F.; LaVoie, E.J. Substituted dibenzo(c,h)cinnolines: Topoisomerase I-targeting anticancer agents. Bioorg. Med. Chem. 2003, 11, 1475–1491. [Google Scholar] [CrossRef]
  8. Makhey, D.; Li, D.; Zhao, B.; Sim, S.-P.; Li, T.-K.; Liu, A.; Liu, L.F.; LaVoie, E.J. Substituted benzo(i)phenanthridines as mammalian topoisomerase I-targeting anticancer agents. Bioorg. Med. Chem. 2003, 11, 1809–1820. [Google Scholar] [CrossRef]
  9. Ruchelman, A.L.; Singh, S.K.; Ray, A.; Wu, X.H.; Yang, J.-M.; Li, T.-K.; Liu, A.; Liu, L.F.; LaVoie, E.J. 5H-Dibenzo(c,h)1,6-napthiridin-6-ones: Novel topo-isomerase I-targeting anticancer agents with potent cytotoxic activity. Bioorg. Med. Chem. 2003, 11, 2061–2073. [Google Scholar] [CrossRef]
  10. Li, D.; Zhao, B.; Sim, S.-P.; Li, T.-K.; Liu, A. 8,9-Methylenedioxybenzo-(i)phenanthridines: Topoisomerase I-targeting activity and cytotoxicity. Bioorg. Med. Chem. 2003, 11, 3795–3805. [Google Scholar] [CrossRef]
  11. Ruchelman, A.L.; Singh, S.K.; Ray, A.; Wu, X.; Yang, J.M.; Zhou, N.; Liu, A.; Liu, L.F.; LaVoie, E.J. 11H-Isoquino(4,3-c)cinnolin-12-ones: Novel anticancer agents with potent topoisomerase I-targeting activity and cytotoxicity. Bioorg. Med. Chem. 2004, 12, 795–806. [Google Scholar]
  12. Ruchelman, A.L.; Kerrigan, J.E.; Li, T.-K.; Zhou, N.; Liu, A.; Liu, L.F.; LaVoie, E.J. Nitro and amino substitution within the A-ring of 5H-8,9-dimethoxy-5-(2-N,N-dimethylamino-ethyl)dibenzo(c,h)(1,6)naph-thyridin-6-ones: Influence on topoisomerase I-targeting activity and cytotoxicity. Bioorg. Med. Chem. 2004, 12, 3731–3742. [Google Scholar]
  13. Zhu, S.; Ruchelman, A.L.; Zhou, N.; Liu, A.; Liu, L.F.; LaVoie, E.J. Esters and amides of 2,3-dime-thoxy-8,9-methylene-dioxybenzo(i)phenanthridine-12-carboxylic acid: Potent topoisomerase I-targeting agents. Bioorg. Med. Chem. 2005, 13, 6782–6794. [Google Scholar] [CrossRef]
  14. Zhu, S.; Ruchelman, A.L.; Zhou, N.; Liu, A.; Liu, L.F.; LaVoie, E.J. 6-Substituted 6H-dibenzo(c,h)(2,6)naph-thyridin-5-ones: Reversed lactam analogues of ARC-111 with potent topoisomerase I-targeting activity and cytotoxicity. Bioorg. Med. Chem. 2006, 14, 3131–3143. [Google Scholar] [CrossRef]
  15. Ruchelman, A.L.; Singh, S.K.; Wu, X.; Ray, A.; Yang, J.M.; Li, T.-K.; Liu, A.; Liu, L.F.; LaVoie, E.J. Diaza- and triazachrysenes: Potent topoisomerase I-targeting agents with exceptional antitumor activity against the human tumor xenograf, MDA-MB-435. Bioorg. Med. Chem. Lett. 2002, 12, 3333–3336. [Google Scholar]
  16. Ruchelman, A.L.; Zhu, S.; Zhou, N.; Liu, A.; Liu, L.F.; LaVoie, E.J. Dimethoxybenzo(i)-phenanthridine-12-carboxylic acid derivatives and 6H-dibenzo(c,h) (2,6)naphthyridin-5-ones with potent topoisomerase I-targeting activity and cytotoxicity. Bioorg. Med. Chem. Lett. 2004, 14, 5585–5589. [Google Scholar] [CrossRef]
  17. Feng, W.; Satyanarayana, M.; Cheng, L.; Liu, A.; Tsai, Y.-C.; Liu, L.F.; LaVoie, E.J. Synthesis of N-substituted 5-[2-(N-alkylamino)ethyl]dibenzo[c,h][1,6]-naphthyridines as novel topoisomerase I-targeting anti-tumor agents. Bioorg. Med. Chem. 2008, 16, 9295–9301. [Google Scholar]
  18. Yap, C.W.; Chen, Y.Z. Prediction of cytochrome P450 3A4, 2D6, and 2C9 Inhibitors and substrates by using support vector machines. J. Chem. Inf. Model. 2005, 45, 982–992. [Google Scholar] [CrossRef]
  19. Thai, K.-M.; Ecker, G.F. Classification models for hERG inhibitors by counter-propagation neural networks. Chem. Biol. Drug Des. 2008, 72, 279–289. [Google Scholar] [CrossRef]
  20. Kubinyi, H. Validation and predictivity of QSAR models. In QSAR & Molecular Modelling in Rational Design of Bioactive Molecules, Proceedings of the 15th European Symposium on QSAR & Molecular Modelling, Istanbul, Turkey, 2004; Sener, E.A., Yalcin, I., Eds.; CADDD Society: Ankara, Turkey, 2006; pp. 30–33. [Google Scholar]
  21. Gramatica, P. Principles of QSAR models validation: Internal and external. QSAR Comb. Sci. 2007, 26, 694–701. [Google Scholar] [CrossRef]
  22. Feng, W.; Satyanarayana, M.; Tsai, Y.-C.; Liu, A.A.; Liu, L.F.; LaVoie, E.J. 11-Substituted 2,3-dimethoxy-8,9-methylenedioxybenzo[i]phenanthridine derivatives as novel topoisomerase I-targeting agents. Bioorg. Med. Chem. 2008, 16, 8598–8606. [Google Scholar] [CrossRef]
  23. Satyanarayana, M.; Tsai, Y.-C.; Liu, A.A.; Liu, L.F.; LaVoie, E.J. Syntheses and biological evaluation of topoisomerase I-targeting agents related to 11-[2-(N,N-dimethylamino)ethyl]-2,3-dimethoxy-8,9-methylene-dioxy-11H-isoquino[4,3-c]cinnolin-12-one (ARC-31). Bioorg. Med. Chem. 2008, 16, 7824–7831. [Google Scholar] [CrossRef]
  24. Satyanarayana, M.; Tsai, Y.-C.; Liu, A.A.; Liu, L.F.; LaVoie, E.J. Synthesis of N-substituted 5-[2-(N-alkyl-amino)ethyl]-dibenzo[c,h][1,6]-naphthyridines as novel topoisomerase I-targeting antitumor agents. Bioorg. Med. Chem. 2008, 16, 9295–9301. [Google Scholar] [CrossRef]
  25. Scaglioni, L.; Mazzini, S.; Mondelli, R.; Dallavalle, S.; Gattinoni, S.; Tinelli, S.; Beretta, G.L.; Zunino, F.; Ragg, E. Interaction between double helix DNA fragments and a new topopyrone acting as human topoisomerase I poison. Bioorg. Med. Chem. 2009, 17, 484–491. [Google Scholar]
  26. Satyanarayana, M.; Tsai, Y.-C.; Liu, A.A.; Liu, L.F.; LaVoie, E.J. 12-Substituted 2,3-dimethoxy-8,9-methylenedioxybenzo-[i]phenanthridines as novel topoisomerase I-targeting antitumor agents. Bioorg. Med. Chem. 2009, 17, 2877–2885. [Google Scholar] [CrossRef]
  27. Clark, R.L.; Deane, F.M.; Anthony, N.G.; Johnston, B.F.; McCarthy, F.O.; Mackay, S.P. Exploring DNA topoisomerase I inhibition by the benzo[c]phenanthridines fagaronine and ethoxidine using steered molecular dynamics. Bioorg. Med. Chem. 2007, 15, 4741–4752. [Google Scholar]
  28. Khadka, D.B.; Cho, W.J. 3-Arylisoquinolines as novel topoisomerase I inhibitors. Bioorg. Med. Chem. 2011, 19, 724–734. [Google Scholar] [CrossRef]
  29. Muegge, I.; Oloff, S. Advances in virtual screening. Drug Dis. Today Tech. 2006, 3, 405–411. [Google Scholar] [CrossRef]
  30. The R Project for Statistical Computing. Available online: http://www.r-project.org (accessed on 26 December 2011).
  31. Karatzoglou, A.; Meyer, D.; Hornik, K. Support vector machines in R. J. Stat. Softw. 2006, 15, 1–28. [Google Scholar]
  32. Tetko, I.V.; Gasteiger, J.; Todeschini, R.; Mauri, A.; Livingstone, D.; Ertl, P.; Palyulin, V.A.; Radchenko, E.V.; Zefirov, N.S.; Makarenko, A.S.; et al. Virtual computational chemistry laboratory—Design and description. J. Comput. Aid. Mol. Des. 2005, 19, 453–463. [Google Scholar] [CrossRef]
  33. VCCLAB, Virtual Computational Chemistry Laboratory. Available online: http://www.vcclab.org (accessed on 30 January 2010).
  34. Hsu, C.-W.; Chang, C.-C.; Lin, C.-J. LIBSVM: A Practical Guide to Support Vector Classification. Available online: http://www.csie.ntu.edu.tw/~cjlin/libsvm (accessed on 5 November 2010).
  35. Fröhlich, H.; Chapelle, O.; Schölkopf, B. Feature selection for support vector machines by means of genetic algorithms. In. In Proceeding ICTAI '03, Proceedings of the 15th IEEE International Conference on Tools with Artificial Intelligence, Sacramento, CA, USA, 3–5 November 2003; IEEE Computer Society: Washington, DC, USA, 2003; pp. 142–148. [Google Scholar]
  36. Fröhlich, H.; Chapelle, O.; Schölkopf, B. Feature selection for support vector machines using genetic algorithms. Int. J. Artif. Intell. T. 2004, 13, 791–800. [Google Scholar] [CrossRef]
  37. Demel, M.A.; Janecek, A.G.K.; Thai, K.-M.; Ecker, G.F.; Gansterer, W.N. Predictive QSAR models for polyspecific drug targets: The importance of feature selection. Curr. Comput. Aided Drug Des. 2008, 4, 91–110. [Google Scholar] [CrossRef]
  38. Thai, K.-M.; Ecker, G.F. A binary QSAR model for classification of hERG potassium channel blockers. Bioorg. Med. Chem. 2008, 16, 4107–4119. [Google Scholar] [CrossRef]
  39. Thai, K.-M.; Ecker, G.F. Similarity-based SIBAR descriptors for classification of chemically diverse hERG blockers. Mol. Divers. 2009, 13, 321–336. [Google Scholar] [CrossRef]
  40. Ji, L.; Wang, X.; Qin, L.; Luo, S.; Wang, L. Predicting the androgenicity of structurally diverse compounds from molecular structure using different classifiers. QSAR Comb. Sci. 2009, 28, 542–550. [Google Scholar] [CrossRef]
  • Samples Availability: Not available.

Share and Cite

MDPI and ACS Style

Thai, K.-M.; Nguyen, T.-Q.; Ngo, T.-D.; Tran, T.-D.; Huynh, T.-N.-P. A Support Vector Machine Classification Model for Benzo[c]phenathridine Analogues with Topoisomerase-I Inhibitory Activity. Molecules 2012, 17, 4560-4582. https://doi.org/10.3390/molecules17044560

AMA Style

Thai K-M, Nguyen T-Q, Ngo T-D, Tran T-D, Huynh T-N-P. A Support Vector Machine Classification Model for Benzo[c]phenathridine Analogues with Topoisomerase-I Inhibitory Activity. Molecules. 2012; 17(4):4560-4582. https://doi.org/10.3390/molecules17044560

Chicago/Turabian Style

Thai, Khac-Minh, Thuy-Quyen Nguyen, Trieu-Du Ngo, Thanh-Dao Tran, and Thi-Ngoc-Phuong Huynh. 2012. "A Support Vector Machine Classification Model for Benzo[c]phenathridine Analogues with Topoisomerase-I Inhibitory Activity" Molecules 17, no. 4: 4560-4582. https://doi.org/10.3390/molecules17044560

Article Metrics

Back to TopTop