EMAN RESEARCH PUBLISHING | <p>Prediction of Protein–Metal Ion-Binding Sites Using Sequence Homology and Machine-Learning Methods</p>
System and chemical biology
RESEARCH ARTICLE   (Open Access)

Prediction of Protein–Metal Ion-Binding Sites Using Sequence Homology and Machine-Learning Methods

Zihan Tian 1, Cao Wei 1, Yutaka Moriwaki 1, Tohru Terada 1, Shugo Nakamura 1, Kazuya Sumikoshi 1, Fang Chun 1, and Kentaro Shimizu 1*

+ Author Affiliations

Advanced Bioinformatics & Chemistry 1 (1) 025-036 https://doi.org/10.25163/abc.11208022130119

Submitted: 21 July 2019 Revised: 22 August 2019  Published: 06 September 2019 


Abstract

Metal ions are essential for metalloproteins to perform their catalytic or structural functions. To understand their role in protein function, it is important to identify metal ion-binding sites. Because experimental identification is labor-intensive and time-consuming, computational methods are expected to be used in the prediction of protein–metal ion-binding sites. A range of computational methods have been proposed to predict metal ion-binding sites from protein sequences. In this study, we implemented two methods of predicting metal ion-binding sites for Ca2+, Co2+, Cu2+, Cu+, Fe3+, Fe2+, Hg2+, Mg2+, Mn2+, Ni2+, and Zn2+ from amino acid sequences. One is a homology-based method, and the other is a machine-learning method. The homology-based method predicts the binding sites from homologous sequences obtained by a protein–protein basic local alignment search tool (BLASTP) search. The machine-learning method uses a support vector machine with three protein sequence features. Our results showed that the homology-based method achieved an accuracy of 0.9905 and a specificity of 0.9978, while the machine-learning method showed balanced performance with regard to accuracy, sensitivity, and specificity. Especially, the sensitivity of the machine-learning method was 0.8239, and many metal ion-binding sites were predicted only by the machine-learning method.

Keywords: protein, metal ion, binding site prediction, machine learning, homology search
 

Introduction

GO

Metalloproteins are proteins that can bind one or more metal ions, which are essential for the proteins to perform their catalytic or structural functions (Degtyarenko et al., 2000). Approximately one-third of the structures in the Protein Data Bank (PDB) (Berman  et al., 2000) contain at least one metal ion. Metal ions are important for the function of proteins and typically have catalytic, transfer, regulatory, structure, recognition, transcription, and transducer roles. The presence of metal ions in metalloproteins is not only helpful for maintaining spatial stability, but also important for executing the physiological functions of the proteins (Holm et al., 1996, Matthews et al., 2008). With the rapid expansion of protein databases, it has become important to identify metal ion-binding sites in metalloproteins in order to understand their role in protein function. Metal ion-binding proteins are experimentally identified and characterized using nuclear magnetic resonance spectroscopy (Jensen et al., 2005; Zhu et al., 2004), gel electrophoresis (Greenough et al., 2015), metal-affinity column chromatography (Herald et al., 2003), electrophoretic mobility shift assay (EMSA) (Hellman et al., 2007), absorbance spectroscopy (Korshin et al., 2009), and mass spectrometry (Binet et al., 2003). However, most of these methods require complex steps and specialized equipment, making them labor-intensive and time-consuming. Hence, prediction of metal ion-binding sites using bioinformatics methods will not only be useful for annotation of experimentally uncharacterized proteins, but also will have a significant role in the prediction of protein structure, function, and genome annotation.

Computational methods have been used to predict protein­­–metal ion-binding sites. Lin et al. 2005 used a neural networks method to predict metal-binding sites in proteins. The target metal ions of Lin’s work are Ca2+, Mg2+, K+, and Na+. Kumar, 2017 used the random forest algorithm to predict protein–metal ion–binding sites for multiple metal ions: Ca2+, Co2+, Cu2+, Fe3+, Mg2+, Mn2+, Ni2+, and Zn2+. We compared our method with Kumar’s. Many other studies have also predicted specific metal ion–binding sites. For example, Zn2+–binding site predictors (Passerini et al., 2007; Srivastava et al., 2018) used the support vector machine and sequence profile information. Metal ion–binding sites have also be predicted using 3D protein structures (Chen et al., 2013; Lu et al., 2012; Yan et al., 2019; Schymkowitz et al., 2005; Goyal et al., 2008). For example, Chen et al. 2013, searched for a triad of amino acids having ligand atoms within specific distances to predict protein–Zn2+–binding sites while Lu et al. 2012 used the fragments transformation method based on the binding site templates. The approach using 3D protein structures can improve prediction performance; however, 3D protein structures are not always available. Predicting metal ion–binding sites only from sequence information is widely applied. Studies have also identified metal ion–binding histidine and cysteine residues in a protein (Passerini et al., 2006; Haberal et al., 2019). However, since these residues are important in metal ion binding, this approach is different from our study.

In this study, we implemented and compared two methods to predict metal ion-binding sites for Ca2+, Co2+, Cu2+, Cu+, Fe3+, Fe2+, Hg2+, Mg2+, Mn2+, Ni2+, and Zn2+ from amino acid sequences: a homology-based method and a machine-learning method. Homology search is a popular method in which multiple alignment of the homologous sequences enables the prediction of metal ion-binding sites. We implemented the homology-based method, which performs this prediction procedure automatically. In the machine-learning method, a support vector machine was employed, and three protein sequence features were selected as its input. To the best of our knowledge, these two methods have not been directly compared using the same dataset. We therefore compared and discussed the results of prediction for each metal ion.

Materials and methods

GO

We developed two kinds of method to predict metal-binding sites from amino acid sequences only: a homology-based method and a machine-learning method (Figure 1). In the homology-based method, BLASTP search (Altschul et al., 1997) was used to identify sequence homology. In the machine-learning method, a support vector machine (SVM) (Boser et al., 1992 was employed with the sequence features amino acid type, position specific scoring matrix (PSSM), and 13 groups of amino acids based on physicochemical properties described below.

Basic flows of the homology-based method and the machine-learning method are shown in orange and blue, respectively.

Dataset

All protein structures in PDB files and mmCIF files and sequences in FASTA files were extracted from the website of the PDB database (Berman et al., 2000) using the following criteria: (1) the protein complex contained at least one metal ion; (2) the x-ray protein structure had a resolution less than or equal to 2.5 Å; and (3) the chain length was greater than 50 residues. A total of 11,903 protein structures and 33,271 chains were collected. For the homology-based method, these chains were used as the positive dataset for BLASTP search. For the machine-learning method, these chains were clustered by CD-HIT (Fu et al., 2012) to remove redundancy for sequence identity thresholds of 30%.

Homology-based method

In the homology-based method, we used BLASTP to obtain the homologous sequences of the query sequence in the positive dataset for this method. The homologous sequences were selected based on the E-value; sequences with E-values lower than a threshold value are defined as homologous sequences.

For instance, when 1AH7:A (A chain of PDB structure 1AH7) was input as a query sequence, assuming the binding sites of this sequence were unknown, three sequences of known structures, 1KHO:A, 2WXU:A, and 1OLP:A, were obtained as the homologous sequences. If an aligned residue in one of the homologous sequences was a metal-binding residue, this residue in the query sequence was predicted as a metal-binding site; if not, this residue was predicted as a nonbinding site. Figure 2 shows an example of this prediction. As shown, 2WXU:A lacks two Zn2+, and the corresponding residues are not defined as binding sites. However, residues of the target aligned to these positions are predicted as binding residues because the other homologs contain Zn2+-binding residues.

Figure 3 shows the superposition of the target and homolog structures. In this example, Zn2+-binding sites (pockets) of the homologs have similar structures and are well aligned.

 

Machine-learning methods and feature extraction

We used SVM as a machine-learning algorithm. SVM shows good performance and generalization abilities for classification and regression analysis. The radial basis function (RBF) kernel was adopted as a kernel function since it seemed appropriate to capture the nonlinearity of multiple binding site properties. We used scikit learn (Pedregosa et al., 2011) to implement the SVM. SVM requires a fixed length of the feature vector for training and testing. We extracted three features to train SVM: amino acid type, position specific scoring matrix (PSSM), and side chain type.

Amino acid type

Each amino acid residue was encoded as a vector of 20 elements; an element that corresponds to an amino acid type is one, and the others are zero. The all zero vector represents a spacer for N- and C-terminals of the sequences. Twenty types of amino acids are decoded into a binary pattern.

PSSM

PSSM is the amino acid substitution score for each position in a protein multiple sequence alignment. The PSSMs for each sequence in the dataset were obtained by the PSI-BLAST (Cooper et al., 2004) program with three iterations of search against the nonredundant data in the National Center for Biotechnology Information (NCBI). For each sequence, PSSM was represented by an N × 20  matrix, where N  is the length of the amino acid sequence. PSSM scores are generally shown as positive or negative integers.

Side chain type

Amino acids are organic compounds containing amine (-NH2) and carboxyl (-COOH) functional groups, along with a side chain (R group) specific to each amino acid.

Amino acids can be classified according to the physicochemical properties of their side chains. In this study, we implemented the grouping scheme proposed in (Cooper et al., 2004); according to side chain polarity, amino acids are divided into four groups; according to side chain class, amino acids are divided into nine groups. The definitions of these features are described in Supplementary material A. Using this feature, each residue was represented by a vector of 13 elements.

Measurement of performance

In the machine-learning method, we employed five-fold cross-validation. The final performance at each parameter was obtained by averaging the performance of all five test sets. The performance of the SVM model at a particular training parameter was assessed using threshold-dependent parameters, namely accuracy, sensitivity, specificity, precision, and Matthew’s correlation coefficient (MCC).

These parameters were calculated using the true positive (TP), true negative (TN), false positive (FP), and false negative (FN), where TP is correctly predicted metal ion-binding amino acids, TN is correctly predicted metal ion-nonbinding amino acids, FP is wrongly predicted metal ion-binding amino acids, and FN is wrongly predicted metal ion-nonbinding amino acids.

Accuracy is the ratio of the correct predictions (TP + TN) to all predictions, and it is calculated as follows:

Accuracy = TP + TNTP + FP + TN + FN .

Sensitivity is the ratio of correctly predicted binding residues (TP) to actual binding residues (TP + FN), and it is calculated as follows:

Sensitivity = TPTP + FN .

Specificity is the ratio of correctly predicted nonbinding residues to actual nonbinding residues (TN + FP), and it is calculated as follows:

Specificity = TNTN + FP .

Precision is the ratio of correctly predicted binding residues (TP) to correctly predicted binding and nonbinding residues (TP + FP), and it is calculated as follows:

Precision = TPTP + FP .

MCC is a balanced measurement that is used to assess the effectiveness of the performance, and it was calculated as follows:

MCC = TP × TN - (FP × FN)(TP + FP)(TP + FN)(TN + FP)(TN + FN) .

Results and discussion

GO

Amino acid frequencies of metal ion-binding sites

We calculated the frequencies of 20 types of amino acid residues in the metal ion-binding sites. The frequencies of each amino acid in the binding sites of each metal ion are represented in Supplementary material B.

Metal ions are commonly coordinated by nitrogen, oxygen, or sulfur centers belonging to side chains on the amino acid residues of the protein, where metal ions provide empty orbits and amino acids provide electrons. The imidazole substituents of histidine residues, the thiolate substituents of cysteine residues, and the carboxylate groups of aspartic acid and glutamic acid can provide electrons as donor groups. This is consistent with the results of the present study, which showed that metal ions preferentially bind certain residues, namely cysteine, histidine, aspartic acid, and glutamic acid. As for Ca2+ and Mg2+, cysteine and histidine do not form cross-links with amino acid residues for constructing a specific structure; therefore, the frequencies of cysteine and histidine are low compared to other ions. Negatively charged residues (aspartic acid and glutamic acid) have high frequencies because of electrostatic interaction with Ca2+ and Mg2+. These ions also bind to oxygen atoms of the backbone. The imidazole substituents of histidine residues, the thiolate substituents of cysteine residues, and the carboxylate groups of aspartic acid and glutamic acid can provide electrons as donor groups, and metal ions can provide empty orbits. Nonpolar residues, such as leucine, isoleucine, and valine, as well as less polar amino acids, such as proline and threonine, show no preference for metal coordination. However, they show certain frequencies because we define a metal ion–binding site not as interacting residues but as neighboring residues within 3.5 Å of a metal ion center.

Performance of the homology-based method

In the homology-based method, we used different E-value thresholds: 0.0001. 0.001, 0.01, and 0.1. The results are shown in Supplementary material C. The number of true positive cases increases and the number of true negative cases decreases as the E-value threshold increases. The accuracy depends on the number of true positives and true negatives, and in many metal ions, except for Ca2+, Mg2+, and Ni2+, the increase in true positives is larger than the increase in true negatives, and thus the accuracy is best for the E-value threshold 0.1. We set this rather high E-value threshold because the sequences of metalloproteins for each metal ion are not so similar as can be detected by BLASTP although some kinds of motifs may exist for binding sites.

Table 2 shows the results of homology-based prediction for each of 11 metal ions when the E-value threshold is 0.1. The last column, “All,” shows the performance for all 11 metal ions (not all metal ions in natural proteins). Metal ions bound to some specific protein families have high sensitivity. In terms of accuracy, the homology-based method performed excellently, with an overall accuracy of 0.9905. It should be noted that with this method, the amount of negative data, which was easier to predict, was much larger than the amount of positive data. For instance, the sensitivity was 0.3544 and the specificity was 0.9961 when an E-value of 0.1 was used to predict Ca2+ binding sites, which means that only 35.44% of actual binding sites (positive data) were correctly predicted. Since substantial numbers of nonbinding sites were correctly predicted, the accuracy became higher.

 

Performance of the machine-learning method

An SVM classifier with RBF kernel has at least two parameters that need to be tuned for good performance: the cost parameter C , which determines the misclassification penalty; and the gamma parameter ? , which is used in the RBF kernel function. We used grid search for obtaining optimal values of ?  and C  to train the SVM models. As a result, we set ?  and C  to 0.07 and 100, respectively, to train the SVM models.

In order to determine an optimal window size, we used the PSSM features of metal ion-binding proteins with window sizes of 9 to 19 to train SVM models. The results are shown in Supplementary material D. As can be seen in the results, for Ca2+ and Mg2+, the accuracy increased from window size 9 to window size 15; for Co2+, Cu+, Fe3+, Hg2+, Mn2+, and Zn2+, the accuracy increased up to window size 13; for Cu2+ and Fe2+, the accuracy increased up to window size 11; for Ni2+, the best performance was shown at window size 9. More than half of the models performed best at window size 13, and the accuracy of the others did not change much in the range between their own optimal window size and window size 13. Therefore, we selected 13 residues as the optimal window size of all 11 types of metal ion-binding proteins.

Table 3 shows the performance of the machine-learning (SVM) method with the PSSM feature. The accuracy was 0.8017 and the MCC was 0.61 overall, and the performance was best for Cu+, with an accuracy of 0.8846 and an MCC of 0.77. It is interesting that the accuracy for Cu+ was the worst in the homology-based method. Table 4 shows the performance of the machine-learning method with the PSSM, the amino acid type, and the side chain type features. The accuracy increased to 0.8336 and the MCC increased to 0.67 overall. The performance was best for Zn2+, with an accuracy of 0.8901 and an MCC of 0.78.

 

Comparison of the two methods

We compared the results of the homology-based method and the machine-learning method. Figure 4 shows the number of chains predicted by the two methods for each metal ion, and Table 5 shows the number of unpredictable chains in the homology-based method. (In the homology-based method, we used the E-value threshold of 0.01.) For 11 types of metal ion-binding protein, the numbers of chains predicted by the SVM method were larger than those predicted by the homology-based method, and only 78% of chains hit their homologous sequences. For instance, the Cu+ binding sites of three chains (PDB IDs: 4BZ4-A, 4MAI-A, 5FJE-B) were successfully predicted by the SVM method and were not predicted by the homology-based method. The SVM method can better predict Ca2+-, Mg2+-, and Zn2+-binding chains compared to the homology-based method. Since Ca2+- and Mg2+-binding proteins exist in various families, the sensitivity of the homology-based method is low. Zn2+-binding sites can often be represented as motifs, and their sequence features tend to be local. The SVM method can recognize these features, while the homology-based method cannot align these sequence patterns.

 

The blue bar represents the number of chains predicted by the homology-based method and the red bar represents the number of chains predicted by the SVM method. The SVM shows higher sensitivity for all metal ions.

Table 5 summarizes the performances of the homology-based method and the machine-learning method. The high accuracy and low sensitivity of the homology-based method were caused by the inequality between negative and positive data. By contrast, the SVM method, which predicted with a balanced performance of accuracy, sensitivity, and specificity, was more effective.

We applied Student’s t-test to compare the two methods used in this study. The p-values of accuracy, sensitivity, specificity, precision, and MCC were 1.551 ´ 10–5, 3.840 ´ 10–8, 8.556 ´ 10–6, 1.684 ´ 10–1, and 1.765 ´ 10–2, respectively. For P < 0.05, the performance measures, except for precision, were significantly different.

Comparison with other work

We also compared our results with those of Kumar’s method (Kumar et al., 2017), which used simplified amino acid alphabets and a random forest model (Table 6). Our method using an SVM model targeted the prediction of three types of metal ion-binding protein (Cu+-, Fe2+-, and Hg2+-binding proteins) which are not available in Kumar’s method. As for Hg2+ ion, however, our dataset contains Hg2+ ions for the soaking in the X-Ray crystal structure analysis, and the results are not clear. When the same types of metal ion-binding proteins were compared, a considerable increase in the accuracy of the Zn2+-, Mn2+-, and Fe3+-binding proteins was observed. The disparity between the accuracy of the Ca2+-, Co2+-, Cu2+-, and Mg2+-binding proteins and the accuracy of Kumar’s method was not apparent. For Ni2+-binding proteins, the performance of our method was unsatisfactory.

Conclusions

GO

Metal ions preferentially bind certain amino acid residues, and many sequence motifs are known in the metalloproteins. In this study, the homology-based method achieved higher accuracy and specificity compared to the machine-learning (SVM) method, while the machine-learning method showed balanced performance with regard to accuracy, sensitivity, and specificity. Especially, the sensitivity of the machine-learning method was high, and we found that it could predict some metal ion-binding sites that were not predicted by the homology-based method and achieved a balanced performance of accuracy, sensitivity, and specificity. We can conceive the following integration of the machine-learning and homology-based methods. Since the machine-learning method achieves high sensitivity, we first obtain candidates of metal ion–binding sites using the machine-learning method. Then, if the homology-based method predicts that the candidates are nonbinding sites, we discard them. This result can be reliable for very high specificity of the homology-based method. In addition, if the homology-based method predicts metal ion–binding sites when homologous sequences are hit by a BLASTP search with low E-values, the result can also be used as candidates of metal ion–binding sites. A detailed design of the integration would be part of future research.

Acknowledgments

GO

The authors would like to thank Yan Chen for valuable comments. This research is supported by Basis for Supporting Innovative Drug Discovery and Life Science Research from the Japan Agency for Medical Research and Development, and the Uehara Memorial Foundation.

Competing Financial Interest

GO

The Auhtor(s) declare no competing financial interest. 

References


Altschul, S.F., Madden, T.L., Schäffer, A.A., Zhang, J., Zhang, Z., Miller, W., Lipman, DJ. (1997). Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 25, 3389-3402.
https://doi.org/10.1093/nar/25.17.3389
PMid:9254694 PMCid:PMC146917
 
Berman, H.M., Westbrook, J., Feng, Z., Gilliland, G., Bhat, T.N., Weissig, H., Shindyalov, I.N., Bourne, P.E. (2000). The protein data bank. Nucleic Acids Res. 28, 235-242.
https://doi.org/10.1093/nar/28.1.235
PMid:10592235 PMCid:PMC102472
 
Binet, M.R.B., Ma, R., McLeod, C.W., Poole, R.K. (2003). Detection and characterization of zinc-and cadmium-binding proteins in Escherichia coli by gel electrophoresis and laser ablation-inductively coupled plasma-mass spectrometry. Anal. Biochem. 318, 30-38.
https://doi.org/10.1016/S0003-2697(03)00190-8
 
Boser, B.E., Guyon, I.M., Vapnik, V.N. (1992). A training algorithm for optimal margin classifiers. Proc. the fifth Annual Workshop on Computational Learning Theory. ACM. 25, 144-152.
https://doi.org/10.1145/130385.130401
 
Chen, Z., Wang, Y., Zhai, Y.F., Song, J., Zhang, Z. (2013). ZincExplorer: an accurate hybrid method to improve the prediction of zinc-binding sites from protein sequences. Mol. Biosyst. 9, 2213-2222.
https://doi.org/10.1039/c3mb70100j
PMid:23861030
 
Cooper, G.M., Hausman, R.E. (2007). The cell: Molecular approach. ASM Press, Washington, D.C.
 
Degtyarenko, K. (2000). Bioinorganic motifs: towards functional classification of metalloproteins. Bioinformatics. 16, 851-864.
https://doi.org/10.1093/bioinformatics/16.10.851
PMid:11120676
 
Fu, L., Niu, B., Zhu, Z., Wu, S., Li, W. (2012). CD-HIT: Accelerated for clustering the next-generation sequencing data. Bioinformatics. 28, 3150-3152.
https://doi.org/10.1093/bioinformatics/bts565
PMid:23060610 PMCid:PMC3516142
 
Goyal, K., Mande, S.C. (2008). Exploiting 3D structural templates for detection of metal-binding sites in protein structures. Proteins. 70, 1206-1218.
https://doi.org/10.1002/prot.21601
PMid:17847089
 
Greenough, L., Schermerhorn, K.M., Mazzola, L., Bybee, J., Rivizzigno, D., Cantin, E., Slatko, B.E., Gardner, A.F. (2015). Adapting capillary gel electrophoresis as a sensitive, high-throughput method to accelerate characterization of nucleic acid metabolic enzymes. Nucleic Acids Res. 44, e15-e15.
https://doi.org/10.1093/nar/gkv899
PMid:26365239 PMCid:PMC4737176
 
Haberal, I., Ogul, H. (2019). Prediction of Protein Metal Binding Sites Using Deep Neural Networks. Mol. Inform. 38, e1800169.
https://doi.org/10.1002/minf.201800169
PMid:30977960
 
Hellman, L.M., Fried, M.G. (2007). Electrophoretic mobility shift assay (EMSA) for detecting protein-nucleic acid interactions. Nat. Protoc. 2, 1849.
https://doi.org/10.1038/nprot.2007.249
PMid:17703195 PMCid:PMC2757439
 
Herald, V.L., Heazlewood, J.L., Day, D.A., Millar, A.H. (2003). Proteomic identification of divalent metal cation-binding proteins in plant mitochondria. FEBS Lett. 537, 96-100.
https://doi.org/10.1016/S0014-5793(03)00101-7
 
Holm, R.H., Kennepohl, P., Solomon, E.I. (1996). Structural and functional aspects of metal sites in biology. Chem. Rev. 96, 2239-2314.
https://doi.org/10.1021/cr9500390
PMid:11848828
 
Jensen, M.R., Petersen, G., Lauritzen, C., Pedersen, J., Led, J.J. (2005). Metal binding sites in proteins: identification and characterization by paramagnetic NMR relaxation. Biochemistry. 44, 11014-11023.
https://doi.org/10.1021/bi0508136
PMid:16101285
 
Korshin, G., Chow, C.W.K., Fabris, R., Drikas, M. (2009). Absorbance spectroscopy-based examination of effects of coagulation on the reactivity of fractions of natural organic matter with varying apparent molecular weights. Water Res. 43, 1541-1548.
https://doi.org/10.1016/j.watres.2008.12.041
PMid:19131089
 
Kumar, S. (2017). Prediction of metal ion binding sites in proteins from amino acid sequences by using simplified amino acid alphabets and random forest model. Genomics Inform. 15, 162-169.
https://doi.org/10.5808/GI.2017.15.4.162
PMid:29307143 PMCid:PMC5769865
 
Lin, C.T., Lin, K.L., Yang, C.H., Chung, I.F., Huang, C.D., Yang, Y.S. (2005). Protein metal binding residue prediction based on neural networks. Int. J. Neural Syst. 15, 71-84.
https://doi.org/10.1142/S0129065705000116
PMid:15912584
 
Lu, C., Lin, Y., Lin, J., Yu, C. (2012). Prediction of Metal Ion-Binding Sites in Proteins Using the Fragment Transformation Method. PLoS ONE. 7, e39252.
https://doi.org/10.1371/journal.pone.0039252
PMid:22723976 PMCid:PMC3377655
 
Matthews, J.M., Loughlin, F.E., Mackay, J.P. (2008). Designed metal-binding sites in biomolecular and bioinorganic interactions. Curr. Opin. Struct. Biol. 18, 484-490.
https://doi.org/10.1016/j.sbi.2008.04.009
PMid:18554898
 
Passerini, A., Andreini, C., Menchetti, S., Rosato, A., Frasconi, P. (2007). Predicting zinc binding at the proteome level. BMC Bioinformatics. 8, 39.
https://doi.org/10.1186/1471-2105-8-39
PMid:17280606 PMCid:PMC1800866
 
Passerini, A., Punta, M., Ceroni, A., Rost, B., Frasconi, P. (2006). Identifying cysteines and histidines in transition-metal-binding sites using support vector machines and neural networks. Proteins. 65, 305-316.
https://doi.org/10.1002/prot.21135
PMid:16927295
 
Pedregosa, F., Varoquaux, G., Gramfort, A., Michel, V., Thirion, B., Grisel, O., Blondel, M., Müller, A., Nothman, J., Louppe, G., Prettenhofer, P., Weiss, R., Dubourg, V., Vanderplas, J., Passos, A., Cournapeau, D., Brucher, M., Perrot, M., Duchesnay, É. (2011). Scikit-learn: Machine Learning in Python, J. Mach. Learn. Res. 12, 2825-2830.
 
Schymkowitz, J.W.H., Rousseau, F., Martins, I.C., Ferkinghoff-Borg, J., Stricher, F., Serrano, L. (2005). Prediction of water and metal binding sites and their affinities by using the Fold-X force field. Nucleic Acids Res. 102, 10147-10152.
https://doi.org/10.1073/pnas.0501980102
PMid:16006526 PMCid:PMC1177371
 
Srivastava, A., Kumar, M. (2018). Prediction of zinc binding sites in proteins using sequence derived information. J. Biomol. Struct. Dyn. 36, 4413-4423.
https://doi.org/10.1080/07391102.2017.1417910
PMid:29241411
 
Yan, R., Wang, X., Tian, Y., Xu, J., Xu, X., Lin, J. (2019). Prediction of zinc-binding sites using multiple sequence profiles and machine learning methods. Molecular Omics. 15, 205-215.
https://doi.org/10.1039/C9MO00043G
PMid:31046040
 
Zhu,D., Herbert, B.E., Schlautman, M.A., Carraway, E.R. (2004). Characterization of cation-π interactions in aqueous solution using deuterium nuclear magnetic resonance spectroscopy. J. Environ. Qual. 33, 276-284.
https://doi.org/10.2134/jeq2004.2760
PMid:14964382

Committee on Publication Ethics

PDF
Abstract
Export Citation

View Dimensions


View Plumx


View Altmetric




Save
0
Citation
983
View

Share