Acessibilidade / Reportar erro

A quantitative relationship between Tgs and chain segment structures of polystyrenes

Abstract

The glass transition temperature (Tg) is a fundamental characteristic of an amorphous polymer. A quantitative structure-property relationship (QSPR) based on error back-propagation artificial neural network (ANN) was constructed to predict Tgs of 107 polystyrenes. Stepwise multiple linear regression (MLR) analysis was adopted to select an optimal subset of molecular descriptors. The chain segments (or motion units) of polymer backbones with 20 carbons in length (10 repeating units) were used to calculate these molecular descriptors reflecting polymer structures. The relative optimal conditions of ANN were obtained by adjusting various network paramters by trial-and-error. Compared to the model already published in the literature, the optimal ANN model with [4-7-1] network structure in this paper is accurate and acceptable, although our model has more samples in the test set. The results demonstrate the feasibility and powerful ability of the chain segment structures as representative of polymers for developing Tg models of polystyrenes.

Keywords:
chain segments; glass transition temperature; polystyrenes; structure-property relationship

1. Introduction

The glass transition temperature (Tg) is known as the glass temperature or the transition temperature between glass and rubber states of amorphous materials. Tg is a fundamental characteristic and is taken as the most crucial property of amorphous polymeric materials[11 Katritzky, A. R., Kuanar, M., Slavov, S., Hall, C. D., Karelson, M., Kahn, I., & Dobchev, D. A. (2010). Quantitative correlation of physical and chemical properties with chemical structure: utility for prediction. Chemical Reviews, 110(10), 5714-5789. PMid:20731377. http://dx.doi.org/10.1021/cr900238d.
http://dx.doi.org/10.1021/cr900238d...
]. The nature of the theory in the glass and glass transition is unsolved, however, is taken as the deepest and most interesting problem in solid stated theory. Though Tg can be determined experimentally, the discrepancies in reported Tg values in the literature may be quite large, because (1) the transition happens over a comparatively wide temperature range, and (2) many factors affect Tg values, which include the structural, constitutional and conformational features of polymers, molecular weight, and experimental conditions such as the measuring method, duration of the experiment, and pressure during the measurement[22 Bicerano, J. (1996). Prediction of polymer properties. 2nd ed. New York: Marcel Dekker.]. In addition, experimental determination of Tgs cannot apply to those polymers that are not yet synthesized. Hence, it is necessary to develop theoretical methods for the prediction of Tgs.

Quantitative structure-property relationship (QSPR) models can be used to predict Tg values of polymers. This approach is based on the assumption that the variation of physicochemical properties of the compounds is dependent on changes of molecular structure, which can be characterized with descriptors. A major goal of QSPR approach is to develop a mathematical relationship between the property of interest and structural features[33 Karelson, M., Lobanov, V. S., & Katritzky, A. R. (1996). Quantum-chemical descriptors in QSAR/QSPR studies. Chemical Reviews, 96(3), 1027-1044. PMid:11848779. http://dx.doi.org/10.1021/cr950202r.
http://dx.doi.org/10.1021/cr950202r...
].

Some researchers have predicted Tgs of polymers with QSPR models. Van Krevelen[44 Van Krevelen, D. W. (1976). Properties of polymers their estimation and correlation with chemical structure. 2nd ed. New York: Elsevier.] predicted Tgs by using the group additive property theory. This method is only applicable to polymers whose contribution values are known. Bicerano[22 Bicerano, J. (1996). Prediction of polymer properties. 2nd ed. New York: Marcel Dekker.] developed a more universally QSPR model with R2 (the square of the correlation coefficient R) being 0.95 and standard error (s) being 24.65 K for a data set of 320 polymers. The Tg model was based on the solubility parameter and the weighted sum of 13 topological bond connectivity parameters of the monomer structures. But the model is not validated with the test set. Joyce et al.[55 Joyce, S. J., Osguthorpe, D. J., Padgett, J. A., & Price, G. J. (1995). Neural network prediction of glass-transition temperatures from monomer structure. Journal of the Chemical Society, Faraday Transactions, 91(16), 2491-2496. http://dx.doi.org/10.1039/ft9959102491.
http://dx.doi.org/10.1039/ft9959102491...
] built models for Tg prediction based on the monomer structures of 360 polymers. The model predicted the Tg values for a test set of polymers with a root mean square (rms) error of 35 K. Katritzky et al.[66 Katritzky, A. R., Rachwal, P., Law, K. W., Karelson, M., & Lobanov, V. S. (1996). Prediction of polymer glass transition temperatures using a general quantitative structure-property relationship treatment. Journal of Chemical Information and Computer Sciences, 36(4), 879-884. http://dx.doi.org/10.1021/ci950156w.
http://dx.doi.org/10.1021/ci950156w...
] introduced a four-parameter model with R2 0.928 for 21 medium molecular weight polymers and copolymers based on their repeat units. On a larger data set, Katritzky et al.[77 Katritzky, A. R., Sild, S., Lobanov, V., & Karelson, M. (1998). Quantitative Structure-Property Relationship (QSPR) correlation of glass transition temperatures of high molecular weight polymers. Journal of Chemical Information and Computer Sciences, 38(2), 300-304. http://dx.doi.org/10.1021/ci9700687.
http://dx.doi.org/10.1021/ci9700687...
] developed a QSPR for the molar glass transition temperature (Tg/M) of 88 uncross-linked linear homopolymers. The model has five molecular descriptors and the s for Tg is 32.9 K. On the same data, Cao and Lin[88 Cao, C., & Lin, Y. (2003). Correlation between the glass transition temperatures and repeating unit structure for high molecular weight polymers. Journal of Chemical Information and Computer Sciences, 43(2), 643-650. PMid:12653533. http://dx.doi.org/10.1021/ci0202990.
http://dx.doi.org/10.1021/ci0202990...
] developed a QSPR model (R2 = 0.9056) by using five molecular descriptors that focus on the influence of chain stiffness and intermolecular forces. Yu et al.[99 Yu, X. L., Wang, X. Y., Li, X. B., Gao, J. W., & Wang, H. L. (2006). Prediction of glass transition temperatures for polystyrenes by a four-descriptors QSPR model. Macromolecular Theory and Simulations, 15(1), 94-99. http://dx.doi.org/10.1002/mats.200500057.
http://dx.doi.org/10.1002/mats.200500057...
] developed stepwise multiple linear regression (MLR) for 107 polystyrenes and generated a QSPR model (R = 0.959 and s = 15.20 K) from the training set of 96 polystyrenes. The MLR model produced a rms error of 20.5 K for the test set comprising 11 polystyrenes. Recently, some quantum chemical descriptors calculated from repeating units or monomers were used to develop QSPR models for Tgs of polymers[1010 Liu, A. H., Wang, X. Y., Wang, L., Wang, H. L., & Wang, H. L. (2007). Prediction of dielectric constants and glass transition temperatures of polymers by quantitative structure property relationships. European Polymer Journal, 43(3), 989-995. http://dx.doi.org/10.1016/j.eurpolymj.2006.12.029.
http://dx.doi.org/10.1016/j.eurpolymj.20...

11 Liu, W. Q., Yi, P. G., & Tang, Z. L. (2006). QSPR models for various properties of polymethacrylates based on quantum chemical descriptors. QSAR & Combinatorial Science, 25(10), 936-943. http://dx.doi.org/10.1002/qsar.200510177.
http://dx.doi.org/10.1002/qsar.200510177...
-1212 Yu, X. L., Yi, B., Wang, X. Y., & Xie, Z. M. (2007). Correlation between the glass transition temperatures and multipole moments for polymers. Chemical Physics, 332(1), 115-118. http://dx.doi.org/10.1016/j.chemphys.2006.11.029.
http://dx.doi.org/10.1016/j.chemphys.200...
].

Due to the large and variable size of polymer molecules, the QSPR models stated above, together with QSPR models of other polymer properties, are modeled by extrapolation from monomer structures or repeating units[11 Katritzky, A. R., Kuanar, M., Slavov, S., Hall, C. D., Karelson, M., Kahn, I., & Dobchev, D. A. (2010). Quantitative correlation of physical and chemical properties with chemical structure: utility for prediction. Chemical Reviews, 110(10), 5714-5789. PMid:20731377. http://dx.doi.org/10.1021/cr900238d.
http://dx.doi.org/10.1021/cr900238d...
]. These methods fail to account for the influences from neighboring repeating units. Especially for the Tg, the glass transition is resulted from Brownian motion of chain segments subjected to freezing or thawing. In this work, the chain segments (localized units or motion units) with 20 carbons (10 repeated units) in length were used to calculate descriptors for their corresponding polystyrenes and to develop QSPR models for their Tgs.

2. Materials and Methods

2.1 Data set

Table 1 shows the experimental Tg data for 107 polystyrenes, which are taken from Brandrup et al.[1313 Brandrup, J., Immergut, E. H., & Grulke, E. A. (1999). Polymer handbook. 4th ed. New York: Wiley-Interscience.]. The entire set contains a Tg value range of 208-490 K. The pendant groups presented in the benzene ring include halides, carbonyls, ethers, hydrocarbon chains, hydroxyl, hydroxyimino, aromatic rings, and other functional groups. These polystyrenes were randomly divided into a training set (70 polystyrenes) and a test set (37 polystyrenes). The training set was used to build a QSPR model, and the test set was adopted to evaluate the model.

Table 1
Molecular descriptors and Tg data of 107 polystyrenes.

2.2 Descriptor computation

A polymeric material consists of a mixture of giant molecules. Therefore, it is impossible to calculate descriptors directly from molecular structures of the polymeric material. Two approaches have been adopted to resolve this problem. One is using the repeating unit to calculate descriptors for the corresponding polymer. The other is using the monomer as representative of the corresponding polymer[11 Katritzky, A. R., Kuanar, M., Slavov, S., Hall, C. D., Karelson, M., Kahn, I., & Dobchev, D. A. (2010). Quantitative correlation of physical and chemical properties with chemical structure: utility for prediction. Chemical Reviews, 110(10), 5714-5789. PMid:20731377. http://dx.doi.org/10.1021/cr900238d.
http://dx.doi.org/10.1021/cr900238d...
].

Tg is a temperature point used to express transition region, where polymer chain segments can move from frozen to movement (or vice versa). Below the glass transition region, 1-4 chain atoms are involved in motion. Further, these motions are largely restricted to vibrations and short-range rotational motions. During the glass transition region, 10-50 chain atoms attain sufficient thermal energy to move in a coordinated manner. In Tg region, these chain atoms (motion units) are first mobilized before the whole molecule starts moving. On further heating, the increased energy allotted to the chains permits them reptate out through entanglements rapidly and flow as individual molecules[1414 Sperling, L. H. (1992). Introduction to physical polymer science. New York: John Wiley & Sons.,1515 Gedde, U. W. (1995). Polymer physics. London: Chapman & Hall.].

The structures of polymer chain segment have an effect on its glass transition and are correlated to Tgs. According to above theory of glass transition, descriptors calculated from the chain segments are more accurate in describing structures affecting polymer Tgs than that from repeating units and monomers. From a theoretical point of view, the chain segment used to calculate descriptors is longer, the descriptors are more accurate in characterizing polymers. The motion units related to glass transition of polymers usually contain 10-50 carbons in length. In addition, a too long segment taken into account may cause difficulty in calculating descriptors, and a too short segment cannot sufficiently represent the structure of motion units. Thus, chain segments with 20 carbons in chain length were used to calculate molecular descriptors for the corresponding polymers.

Polymeric chain segments containing 20 main chain carbons of polystyrenes were first sketched using ChemBioDraw Ultra 11.0 in ChemBioOffice 2008 program. For example, the structure model consisting of 10 repeating units end-capped by two hydrogens (see Figure 1) was adopted as the representative structure of poly(styrene) (No. 65 in Table 1) to calculate the descriptors.

Figure 1
The calculated models of poly(styrene).

Subsequently, the sketched 2D molecular structures were converted to 3D structures and optimized using a molecular mechanics (MM2 force field) in ChemBio3D Ultra 11.0 with the convergence criterion of minimum rms of gradient value being 0.01 kcal/molÅ. The optimized molecules were saved in Sybyl mol2 (.mol2) format as the input files for Dragon software[1616 Talete SRL. (2012). Dragon: software for molecular descriptor calculation. Version 6.0. Milano: Talete SRL. Retrieved in 20 February 2016, from http://www.talete.mi.it/
http://www.talete.mi.it/...
]. Lastly, 4885 descriptors were calculated for each energy-minimized motion unit with Dragon software. Descriptors with constant or near constant values and with pair correlation greater than or equal to 0.90 were removed in order to reduce redundant and non-useful information. After excluding redundant and non-useful variables, 551 descriptors were remained to undergo descriptor selection. A relative optimal subset of descriptors was obtained by applying MLR analysis in IBM SPSS Statistics 19.

2.3 Artificial neural network

The optimal descriptors subset was fed to artificial neural network (ANN) as input vectors. ANNs are computational models, which simulate the human brain behavior. The common networks consist of an input layer, some number of hidden layers (intermediate layers) and an output layer. Each layer includes a number of processing nodes, called neurons or units. Each node in the network is influenced by those nodes to which it is connected in a highly complex and parallel way. The degree of influence is dictated by the values of the links or connections. Through a training algorithm, the overall behavior of ANNs can be modified by adjusting the weights (or the values of the links or connections). After learning from the input dataset, ANNs acquire knowledge and can be applied on test set data not present in the training set. The output layer produces the prediction values of properties interested. One of the most popular algorithms applied in the training phase is the error back-propagation (BP) algorithm. The number of neurons in the hidden layer shouled be optimized by trial and validation until no obvious improvement was seen for that model[1717 Yu, X. L., Yi, B., Liu, F., & Wang, X. Y. (2008). Prediction of the dielectric dissipation factor tan δ of polymers with an ANN model based on the DFT calculation. Reactive & Functional Polymers, 68(11), 1557-1562. http://dx.doi.org/10.1016/j.reactfunctpolym.2008.08.009.
http://dx.doi.org/10.1016/j.reactfunctpo...
].

3. Results and Discussions

By analyzing the correlation between the 551 descriptors and Tgs of 70 polystyrenes in the training set with stepwise MLR analysis in IBM SPSS Statistics 19, Equation 1 and the corresponding statistical results were obtained.

T g = 1346.200 4546.871 C h i A _ B ( e ) + 100.47 S p M a x _ E A ( b o ) 12.653 H 7 s 122.231 D L S _ 01 (1)
n = 70, R = 0.955, R 2 = 0.912, s = 16.717, F = 167.652

where n is the number of samples from the training set; s is the standard error of estimate; R is the correlation coefficient; F is the Fischer ratio.

The four molecular descriptors, ChiA_B(e), SpMax_EA(bo), H7s and DLS_01 appearing in above MLR model and the corresponding descriptor values are shown in Table 1. Their descriptor characteristics are listed in Table 2; and their definitions[1818 Todeschini, R., & Consonni, V. (2009). Molecular descriptors for chemoinformatics. Weinheim: Wiley-VCH. 2 v.] are shown in Table 3. Calculated results with Equation 1 are depicted in Figure 2A. The rms errors of Tgs of the training and test sets are 16.1 and 22.4 K, respectively.

Table 2
Characteristics of descriptors appearing in MLR model.
Table 3
The symbol, class and definition for descriptors appearing in MLR models.
Figure 2
Plots of calculated vs. experimental Tg values of polystyrenes: (A) for MLR model; (B) for ANN model.

The four descriptors are then fed to ANN as input vectors. The optimal condition of the neural network was obtained by adjusting various parameters by trial-and-error. The architecture of the final optimum BP neural network is [4-7-1], with the number of hidden layer being 1, the nodes in hidden layer being 7, the permission error being 0.00001, the momentum being 0.6, and the sigmoid parameter being 0.9. The results from ANN method are listed in Table 1 and depicted in Figure 2B, which indicate that the predicted Tg values are close to the experimental ones. The rms error of training set is 13.6 K (R = 0.939). The test set rms error is 17.1 K (R = 0.902) which is less than the errors from the test set in previous model (20.5 K)[99 Yu, X. L., Wang, X. Y., Li, X. B., Gao, J. W., & Wang, H. L. (2006). Prediction of glass transition temperatures for polystyrenes by a four-descriptors QSPR model. Macromolecular Theory and Simulations, 15(1), 94-99. http://dx.doi.org/10.1002/mats.200500057.
http://dx.doi.org/10.1002/mats.200500057...
]. The mean relative error for the 107 polystyrenes in Table 1 is 3.4%, less than that from the model of Yu et al.[99 Yu, X. L., Wang, X. Y., Li, X. B., Gao, J. W., & Wang, H. L. (2006). Prediction of glass transition temperatures for polystyrenes by a four-descriptors QSPR model. Macromolecular Theory and Simulations, 15(1), 94-99. http://dx.doi.org/10.1002/mats.200500057.
http://dx.doi.org/10.1002/mats.200500057...
] (3.7%). Furthermore, it should be noted that the test set in this paper possesses 37 polystyrenes, more than the number of samples (11 polystyrenes)[99 Yu, X. L., Wang, X. Y., Li, X. B., Gao, J. W., & Wang, H. L. (2006). Prediction of glass transition temperatures for polystyrenes by a four-descriptors QSPR model. Macromolecular Theory and Simulations, 15(1), 94-99. http://dx.doi.org/10.1002/mats.200500057.
http://dx.doi.org/10.1002/mats.200500057...
]. And it is much easier to obtain better results on small test set of polymers. In comparison to previous model on Tgs of polystyrenes[99 Yu, X. L., Wang, X. Y., Li, X. B., Gao, J. W., & Wang, H. L. (2006). Prediction of glass transition temperatures for polystyrenes by a four-descriptors QSPR model. Macromolecular Theory and Simulations, 15(1), 94-99. http://dx.doi.org/10.1002/mats.200500057.
http://dx.doi.org/10.1002/mats.200500057...
], the statistic qualities of our model is accurate and acceptable. Therefore, it is feasible to calculate molecular descriptors from the chain segments of polymer backbones comprising 10 repeating units for developing Tg model of polystyrenes.

Table 2 shows that each descriptor in Equation 1 has a Sig.-value near to 0, and less than the default level of 0.05, which suggest that these descriptors are significant for Tgs. Moreover, all variance inflation factor (VIF) values are less than 2, far less than the default value of 10. Thus these descriptors are “pure” without “mixing” or contamination from other descriptors, and each descriptor reflects some particular molecular structures affecting Tgs.

According to the t-test, the most significant descriptor in the MLR model is ChiA_B(e) (2D matrix-based descriptors)[1616 Talete SRL. (2012). Dragon: software for molecular descriptor calculation. Version 6.0. Milano: Talete SRL. Retrieved in 20 February 2016, from http://www.talete.mi.it/
http://www.talete.mi.it/...
]. ChiA_B(e) denotes the average randic-like index from burden matrix weighted by Sanderson electronegativity and is defined as follow:

C h i A _ B ( e ) = C h i _ M ( e ) n B O (2)

Where nBO is the number of graph edges. Chi_M(e) is the Randic-like index calculated by applying Sanderson electronegativity as the vertex weighting scheme and a H-depleted molecular graph as a square matrix:

C h i _ M ( e ) = i = 1 n S K 1 j = i + 1 n S K α i j [ V S i ( M ; e ) V S j ( M ; e ) ] 1 / 2 (3)

Here nSK means the number of graph vertices; VSi(M) is the ith matrix row sum; αij are the elements of the adjacency matrix, which are equal to one for pairs of adjacent vertices, and zero otherwise. ChiA_B(e) reflects information about interatomic distances, bond distances, ring types, planar and non-planar systems and atom types[1616 Talete SRL. (2012). Dragon: software for molecular descriptor calculation. Version 6.0. Milano: Talete SRL. Retrieved in 20 February 2016, from http://www.talete.mi.it/
http://www.talete.mi.it/...
]. A small ChiA_B(e) indicates a small interatomic distances, which results in a low degree of freedom for rotation and leads to high Tg.

The second significant descriptor is the GETAWAY (GEometry, Topology, and Atom-Weights AssemblY) descriptor, H7s (H autocorrelation of lag 7 / weighted by I-state). The descriptor H7s encodes information on structural fragments, such as the effective position of substituents and fragments in the molecular space, and accounts information on molecular size and shape as well as for specific atomic properties[1616 Talete SRL. (2012). Dragon: software for molecular descriptor calculation. Version 6.0. Milano: Talete SRL. Retrieved in 20 February 2016, from http://www.talete.mi.it/
http://www.talete.mi.it/...
]. A large H7s suggests that a polymer has a large side group, which decreases the volume ratio of phenyl ring to other substituent groups. While the aromatic or cyclic structure in bulky side groups increases rotational barrier for backbone chain and leads to high Tg. Therefore, a polymer with large H7s may have a low Tg.

The next significant descriptor is the Drug-like indice DLS_01. The descriptor DLS_01, being modified drug-like score from Lipinski (4 rules), is calculated as 1 minus Lipinski Alert Index (LAI), while LAI is defined as the ratio between the number of satisfied conditions over the total number of conditions, i.e., (1) there are more than 5 H-bond donors; (2) there are more than 10 H-bond acceptors (N and O atoms); (3) molecular weight (MW) is over 500; and (4) Moriguchi's logP (MLogP) is over 4.15[1616 Talete SRL. (2012). Dragon: software for molecular descriptor calculation. Version 6.0. Milano: Talete SRL. Retrieved in 20 February 2016, from http://www.talete.mi.it/
http://www.talete.mi.it/...
]. DLS_01 is related to the number of intermolecular hydrogen bonds, which increase intermolecular force and determine the magnitude of molecular aggregates. Polymer molecules with small DLS_01 hold together more strongly due to intermolecular hydrogen bonds and are unable to mover that easily, and possess high Tgs.

According to the t-test, the last significant descriptor in the MLR model is SpMax_EA(bo). Edge adjacency index, SpMax_EA(bo), is derived from the H-depleted molecular graph and encodes the connectivity between graph edges. It is leading eigenvalue from edge adjacency matrix weighted by bond order. SpMax_EA(bo) reflects molecular shape and implies the substituent position in the phenyl ring for styrenes[1616 Talete SRL. (2012). Dragon: software for molecular descriptor calculation. Version 6.0. Milano: Talete SRL. Retrieved in 20 February 2016, from http://www.talete.mi.it/
http://www.talete.mi.it/...
]. Compared to styrenes with substituents lying in p-or m-positions of the phenyl ring, a styrene with a substituent lying in o-positions usually has a larger SpMax_EA(bo), which can be seen from Table 1. The substituents in o-positions will enhance rotational barrier for backbone chain, increase rigidity of polymer chains and result in higher Tgs[99 Yu, X. L., Wang, X. Y., Li, X. B., Gao, J. W., & Wang, H. L. (2006). Prediction of glass transition temperatures for polystyrenes by a four-descriptors QSPR model. Macromolecular Theory and Simulations, 15(1), 94-99. http://dx.doi.org/10.1002/mats.200500057.
http://dx.doi.org/10.1002/mats.200500057...
].

Despite a variety of factors affecting the Tg values of polymeric materials, intermolecular forces and molecular flexibility (or rigidity) are two important factors related to Tgs. The descriptor DLS_01 reflects the intermolecular forces, while descriptors ChiA_B(e), SpMax_EA(bo) and H7s indicate the stiffness of polymer. Therefore, the four descriptors can predict Tgs sufficiently.

Figure 3 (Williams plot) was obtained to visualize the applicability domain of the ANN model in this paper. According to Williams plot based on standardized residuals vs. leverages, predictions for only those samples that fall into this domain may be considered reliable[1919 Tropsha, A., Gramatica, P., & Gombar, V. K. (2003). The importance of being earnest: validation is the absolute essential for successful application and interpretation of QSPR Models. QSAR & Combinatorial Science, 22(1), 69-77. http://dx.doi.org/10.1002/qsar.200390007.
http://dx.doi.org/10.1002/qsar.200390007...
,2020 Wang, Y. N., Chen, J. W., Li, X. H., Wang, B., Cai, X. Y., & Huang, L. P. (2009). Predicting rate constants of hydroxyl radical reactions with organic pollutants: algorithm, validation, applicability domain, and mechanistic interpretation. Atmospheric Environment, 43(5), 1131-1135. http://dx.doi.org/10.1016/j.atmosenv.2008.11.012.
http://dx.doi.org/10.1016/j.atmosenv.200...
]. Figure 3 shows that only the two samples No. 44, poly(4-propoxysulfonylstyrene) and No. 67, poly(2,3,4,5,6,-pentafluorostyrene) in the training set have larger leverage h values (0.448 and 0.456, respectively), greater than the warning leverage h* (= 0.214). But their standardized residual values (0.430 and 0.075, respectively) are less than 3. Thus the two samples, poly(4-propoxysulfonylstyrene) and poly(2,3,4,5,6,-pentafluorostyrene), can stabilize the ANN model of polystyrenes and make it more accurate.

Figure 3
Williams plot for polystyrenes with a warning leverage of 0.214.

4. Conclusions

Four molecular descriptors calculated from the chain segments of main chains comprising 10 repeating units were adopted for developing QSPR model of Tgs for polystyrenes. MLR analysis was used to select the optimal subset of descriptors after molecular descriptor generation for each chain segment. The developed ANN model was proved to be accurate and acceptable, with the absolute mean errors for the whole data set is 3.4%, which is less than that of the model published in the literature, although our model possesses more samples for the test set. Therefore, it is feasible calculating molecular descriptors from the chain segments comprising 10 repeating units in length to develop ANN model of Tgs for polystyrenes.

5. Acknowledgements

The project was supported by the National Natural Science Foundation of China (No. 21472040) and Scientific Research Fund of the Hunan Provincial Education Department (No. 16A047).

6. References

  • 1
    Katritzky, A. R., Kuanar, M., Slavov, S., Hall, C. D., Karelson, M., Kahn, I., & Dobchev, D. A. (2010). Quantitative correlation of physical and chemical properties with chemical structure: utility for prediction. Chemical Reviews, 110(10), 5714-5789. PMid:20731377. http://dx.doi.org/10.1021/cr900238d
    » http://dx.doi.org/10.1021/cr900238d
  • 2
    Bicerano, J. (1996). Prediction of polymer properties. 2nd ed. New York: Marcel Dekker.
  • 3
    Karelson, M., Lobanov, V. S., & Katritzky, A. R. (1996). Quantum-chemical descriptors in QSAR/QSPR studies. Chemical Reviews, 96(3), 1027-1044. PMid:11848779. http://dx.doi.org/10.1021/cr950202r
    » http://dx.doi.org/10.1021/cr950202r
  • 4
    Van Krevelen, D. W. (1976). Properties of polymers their estimation and correlation with chemical structure. 2nd ed. New York: Elsevier.
  • 5
    Joyce, S. J., Osguthorpe, D. J., Padgett, J. A., & Price, G. J. (1995). Neural network prediction of glass-transition temperatures from monomer structure. Journal of the Chemical Society, Faraday Transactions, 91(16), 2491-2496. http://dx.doi.org/10.1039/ft9959102491
    » http://dx.doi.org/10.1039/ft9959102491
  • 6
    Katritzky, A. R., Rachwal, P., Law, K. W., Karelson, M., & Lobanov, V. S. (1996). Prediction of polymer glass transition temperatures using a general quantitative structure-property relationship treatment. Journal of Chemical Information and Computer Sciences, 36(4), 879-884. http://dx.doi.org/10.1021/ci950156w
    » http://dx.doi.org/10.1021/ci950156w
  • 7
    Katritzky, A. R., Sild, S., Lobanov, V., & Karelson, M. (1998). Quantitative Structure-Property Relationship (QSPR) correlation of glass transition temperatures of high molecular weight polymers. Journal of Chemical Information and Computer Sciences, 38(2), 300-304. http://dx.doi.org/10.1021/ci9700687
    » http://dx.doi.org/10.1021/ci9700687
  • 8
    Cao, C., & Lin, Y. (2003). Correlation between the glass transition temperatures and repeating unit structure for high molecular weight polymers. Journal of Chemical Information and Computer Sciences, 43(2), 643-650. PMid:12653533. http://dx.doi.org/10.1021/ci0202990
    » http://dx.doi.org/10.1021/ci0202990
  • 9
    Yu, X. L., Wang, X. Y., Li, X. B., Gao, J. W., & Wang, H. L. (2006). Prediction of glass transition temperatures for polystyrenes by a four-descriptors QSPR model. Macromolecular Theory and Simulations, 15(1), 94-99. http://dx.doi.org/10.1002/mats.200500057
    » http://dx.doi.org/10.1002/mats.200500057
  • 10
    Liu, A. H., Wang, X. Y., Wang, L., Wang, H. L., & Wang, H. L. (2007). Prediction of dielectric constants and glass transition temperatures of polymers by quantitative structure property relationships. European Polymer Journal, 43(3), 989-995. http://dx.doi.org/10.1016/j.eurpolymj.2006.12.029
    » http://dx.doi.org/10.1016/j.eurpolymj.2006.12.029
  • 11
    Liu, W. Q., Yi, P. G., & Tang, Z. L. (2006). QSPR models for various properties of polymethacrylates based on quantum chemical descriptors. QSAR & Combinatorial Science, 25(10), 936-943. http://dx.doi.org/10.1002/qsar.200510177
    » http://dx.doi.org/10.1002/qsar.200510177
  • 12
    Yu, X. L., Yi, B., Wang, X. Y., & Xie, Z. M. (2007). Correlation between the glass transition temperatures and multipole moments for polymers. Chemical Physics, 332(1), 115-118. http://dx.doi.org/10.1016/j.chemphys.2006.11.029
    » http://dx.doi.org/10.1016/j.chemphys.2006.11.029
  • 13
    Brandrup, J., Immergut, E. H., & Grulke, E. A. (1999). Polymer handbook. 4th ed. New York: Wiley-Interscience.
  • 14
    Sperling, L. H. (1992). Introduction to physical polymer science. New York: John Wiley & Sons.
  • 15
    Gedde, U. W. (1995). Polymer physics. London: Chapman & Hall.
  • 16
    Talete SRL. (2012). Dragon: software for molecular descriptor calculation. Version 6.0. Milano: Talete SRL. Retrieved in 20 February 2016, from http://www.talete.mi.it/
    » http://www.talete.mi.it/
  • 17
    Yu, X. L., Yi, B., Liu, F., & Wang, X. Y. (2008). Prediction of the dielectric dissipation factor tan δ of polymers with an ANN model based on the DFT calculation. Reactive & Functional Polymers, 68(11), 1557-1562. http://dx.doi.org/10.1016/j.reactfunctpolym.2008.08.009
    » http://dx.doi.org/10.1016/j.reactfunctpolym.2008.08.009
  • 18
    Todeschini, R., & Consonni, V. (2009). Molecular descriptors for chemoinformatics. Weinheim: Wiley-VCH. 2 v.
  • 19
    Tropsha, A., Gramatica, P., & Gombar, V. K. (2003). The importance of being earnest: validation is the absolute essential for successful application and interpretation of QSPR Models. QSAR & Combinatorial Science, 22(1), 69-77. http://dx.doi.org/10.1002/qsar.200390007
    » http://dx.doi.org/10.1002/qsar.200390007
  • 20
    Wang, Y. N., Chen, J. W., Li, X. H., Wang, B., Cai, X. Y., & Huang, L. P. (2009). Predicting rate constants of hydroxyl radical reactions with organic pollutants: algorithm, validation, applicability domain, and mechanistic interpretation. Atmospheric Environment, 43(5), 1131-1135. http://dx.doi.org/10.1016/j.atmosenv.2008.11.012
    » http://dx.doi.org/10.1016/j.atmosenv.2008.11.012

Publication Dates

  • Publication in this collection
    Jan-Mar 2017

History

  • Received
    20 Feb 2016
  • Reviewed
    02 May 2016
  • Accepted
    24 May 2016
Associação Brasileira de Polímeros Rua São Paulo, 994, Caixa postal 490, São Carlos-SP, Tel./Fax: +55 16 3374-3949 - São Carlos - SP - Brazil
E-mail: revista@abpol.org.br