Processing math: 100%
Research article

QSAR study and theoretical investigation on the lethality of halogenated aliphatic hydrocarbons toward Aspergillus (A.) Nidulans

  • Received: 20 August 2024 Revised: 06 January 2025 Accepted: 08 May 2025 Published: 19 May 2025
  • The prediction of Aspergillus (A.) nidulans toxicities (log1/D37) for a set of 55 halogenated aliphatic hydrocarbons (HAHs) was thoroughly investigated using density functional theory (DFT) computations. Different multiple linear regression (MLR)methods were employed to assess the reliability of the proposed quantitative structure-activity relationships (QSAR) model. The obtained ELUMO, Egap, molecular polarizability (α), and molar refractivity (MR) values offered informative indications in determining the toxicity of the HAHs. A promising three-descriptor linear model was constructed using 41 molecules as a training set; then, the model was validated on the remaining 14 molecules. Statistical comparisons between these models and others quoted from the literature were presented. Furthermore, the potential causes of the outlier molecules in the proposed QSAR models were explored. The most preferable interactions were obviously noticed within the 1-bromo-2-methylpropane…α-glucan complex, followed by 2-bromo-2-methylpropane…α-glucan and 2-chloro-2-methylpropane…α-glucan complexes. Compared to other analogs, the higher number of bond paths and bond critical points within the 1-bromo-2-methylpropane…α-glucan complex highlighted its high preferability.

    Citation: Jabir H. Al-Fahemi, Faten A. Aljiffrey, Elshafie A. M. Gad, Mahmoud A. A. Ibrahim. QSAR study and theoretical investigation on the lethality of halogenated aliphatic hydrocarbons toward Aspergillus (A.) Nidulans[J]. AIMS Environmental Science, 2025, 12(3): 419-434. doi: 10.3934/environsci.2025019

    Related Papers:

    [1] Navid Ahmadi, Mozhgan Ahmadi Nadoushan, Mohammad Hadi Abolhasani, Abbas Hosseini . Investigating the efficiency of biological treatment process of oil pollutants using mix of Scenedesmus obliquus and Chlamydomonas reinhardtii algae: A case study. AIMS Environmental Science, 2021, 8(3): 221-237. doi: 10.3934/environsci.2021015
    [2] Arriya Mungsunti, Kevin A. Parton . The sustainability of the muang fai irrigation system of northern Thailand. AIMS Environmental Science, 2019, 6(2): 77-93. doi: 10.3934/environsci.2019.2.77
    [3] Alma Sobrino-Figueroa, Sergio H. Álvarez Hernandez, Carlos Álvarez Silva C . Evaluation of the freshwater copepod Acanthocyclops americanus (Marsh, 1983) (Cyclopidae) response to Cd, Cr, Cu, Hg, Mn, Ni and Pb. AIMS Environmental Science, 2020, 7(6): 449-463. doi: 10.3934/environsci.2020029
    [4] Ewelina Nerek, Barbara Sokołowska . Pseudomonas spp. in biological plant protection and growth promotion. AIMS Environmental Science, 2022, 9(4): 493-504. doi: 10.3934/environsci.2022029
    [5] Karolina Nowocień, Barbara Sokołowska . Bacillus spp. as a new direction in biocontrol and deodorization of organic fertilizers. AIMS Environmental Science, 2022, 9(2): 95-105. doi: 10.3934/environsci.2022007
    [6] Delianis Pringgenies, Wilis Ari Setyati, Nirwani Soenardjo, Rini Pramesti . Investigation of extra-cellular protease in indigenous bacteria of sea cucumbers as a candidate for bio-detergent material in bio-industry. AIMS Environmental Science, 2020, 7(4): 335-349. doi: 10.3934/environsci.2020022
    [7] Wen-Hung Lin, Kuo-Hua Lee, Liang-Tu Chen . The effects of Ganoderma lucidum compound on goat weight and anti-inflammatory: a case study of circular agriculture. AIMS Environmental Science, 2021, 8(6): 553-566. doi: 10.3934/environsci.2021035
    [8] Riri Novita Sunarti, Sri Budiarti, Marieska Verawaty, Bayo Alhusaeri Siregar, Poedji Loekitowati Hariani . Diversity of Antibiotic-Resistant Escherichia coli from Rivers in Palembang City, South Sumatra, Indonesia. AIMS Environmental Science, 2022, 9(5): 721-734. doi: 10.3934/environsci.2022041
    [9] Cristina Calderón-Tapia, Daniel Chuquín-Vasco, Alex Guambo-Galarza, Soledad Núñez-Moreno, Cristina Silva-Cisneros . Bioelectricity production from anaerobically treated leachate in microbial fuel cell using Delftia acidovorans spp.. AIMS Environmental Science, 2023, 10(6): 847-867. doi: 10.3934/environsci.2023046
    [10] HsiaoDao Chang, XiuYou Wan, HsiaoLan Huang, YiSu Chen, ChaoYing Chen . Anaerobic enrichment of Bacillus alkylbenzene remedial consortia from waste biomass melanoid sources. AIMS Environmental Science, 2021, 8(4): 341-357. doi: 10.3934/environsci.2021022
  • The prediction of Aspergillus (A.) nidulans toxicities (log1/D37) for a set of 55 halogenated aliphatic hydrocarbons (HAHs) was thoroughly investigated using density functional theory (DFT) computations. Different multiple linear regression (MLR)methods were employed to assess the reliability of the proposed quantitative structure-activity relationships (QSAR) model. The obtained ELUMO, Egap, molecular polarizability (α), and molar refractivity (MR) values offered informative indications in determining the toxicity of the HAHs. A promising three-descriptor linear model was constructed using 41 molecules as a training set; then, the model was validated on the remaining 14 molecules. Statistical comparisons between these models and others quoted from the literature were presented. Furthermore, the potential causes of the outlier molecules in the proposed QSAR models were explored. The most preferable interactions were obviously noticed within the 1-bromo-2-methylpropane…α-glucan complex, followed by 2-bromo-2-methylpropane…α-glucan and 2-chloro-2-methylpropane…α-glucan complexes. Compared to other analogs, the higher number of bond paths and bond critical points within the 1-bromo-2-methylpropane…α-glucan complex highlighted its high preferability.



    Halogenated aliphatic hydrocarbons (HAHs) are of growing concern because of their effectual toxic and carcinogenic features [1,2]. Nevertheless, HAHs are commonly utilized as industrial and home solvents, chemical synthesis intermediates, and have a range of other applications. Within the literature, HAHs were addressed with a potent versatility to react and covalently bind with DNA, thus causing genetic damage in versatile experimental organisms.

    In this regard, developing new models capable of promptly screening for potential hazards of any conceivable substance in the aquatic habitat has become an essential demand. Among the developed models, there has been a continued interest in estimating the toxicity data via quantitative structure-activity/-toxicity relationships (QSAR/QSTR) [3,4,5,6]. Furthermore, the QSAR model attempts to link molecular structures to biological endpoints such as toxicity [7]. These findings are considerably distinct from the well-known in vivo data, which would be attributed to a number of toxic action mechanisms. In contrast to in vivo observations, in vitro toxicity data are easy to collect and often have a direct mechanistic significance. Consequently, the QSAR concept is one of the ligand-based drug design strategies that has effectively led to the development of new drug candidates for a wide range of diseases [8].

    Based on the literature, the QSAR can be mechanistically meaningful by using suitable molecular descriptors to mediate toxicity [9]. Except for the toxic effects caused by receptor binding, toxicity might be described as an outcome of the chemical compound's capacity to reach and covalently interact at the active site [5]. Considering a common biophysical mechanism, variations in the chemical structure of a set of known compounds are linked to toxicity alterations. Thence, the QSAR is utilized to extrapolate into other compounds.

    As a point of departure, Crebelli et al. [3] presented a QSAR model based on the molar refractivity (MR) and energy gap (Egap) for the prediction of 55 haloalkenes towards Aspergillus (A.) nidulans. This regression model was characterized by R2 and F values of 0.64 and 45.5, respectively, and outlined the effectual role of the electrophilicity in predicting the aneugenic potential of the aliphatic hydrocarbons. Furthermore, the toxicity of 52 halogenated hydrocarbons toward A. nidulans was well-characterized by Trohalaki et al. using descriptors, including polarizability (α), lowest unoccupied molecular orbital energy (ELUMO), and molecular volume (V) [4]. The most preferential model was detected with an R2 value of 0.79 and an F value of 0.59, thereby utilizing the B3LYP/6-31G** and HF/6-31G** levels of theory. Afterward, the lethal effect of 55 aliphatic molecules using descriptors as the logarithm of the octanol-water partition coefficient (log P) and the ELUMO was addressed by Cronin et al. [10]. Subsequently, this two-descriptors-based regression model was characterized by R2 and standard error values of 0.615 and 0.413, respectively.

    While QSAR models are widely used, there is a lack of predictive models which specifically target the toxicity of HAHs, and this gap necessitates further research. Therefore, the current study aims to construct QSAR models to predict A. nidulans toxicity of 55 halogenated aliphatic hydrocarbons. A possible source of outlier molecules in the developed models is proposed along with the predictions.

    The utilized data on A. nidulans toxicity (log1/D37) of 55 halogenated aliphatic hydrocarbons were previously measured. In particular, the data set splitting was performed using random division as the standard method. Based on the literature [11], the rational division methods produced better statistical outcomes for the test sets compared to models that used random division; however, the predictive powers of random and rational models are similar. Therefore, the selected halogenated hydrocarbons were randomly partitioned into two sets, namely, training and validation sets, which involve 41 and 14 molecules, respectively. The latter set was herein devoted to thoroughly confirming the accuracy of the developed QSTR model.

    Herein, all executed quantum chemical calculations were performed using the Amsterdam density functional program package (ADF 2010.02) [12,13]. Geometry optimization was first carried out for each molecule using the quadruple-ξ Slater basis set with four polarization functions (STO-QZ4P) in conjunction with the generalized gradient approximation (GGA) within the PW91 exchange and correlation functional [14,15]. The inner electrons within the atomic shells were treated as a frozen-core approximation to accelerate the computations. In the realm of the DFT calculations, various quantum chemical parameters were calculated, including the energies of molecular orbitals (EHOMO and ELUMO) along with the energy gap (Egap = ELUMOEHOMO). Afterward, HyperChem (version 8.0) was employed to compute various physical properties, including the surface area grid (S) octanol-water partition coefficient (log P), molecular volume (V), molar refractivity (MR), and molecular weight (M). Additionally, the molecular ovality (O) [16] was computed as a fundamental indicator to represent how the molecular shape approaches a sphere, which can be defined as follows:

    O=S/4π(3V4π)2/3 (1)

    For the statistical analysis, a principal component analysis (PCA) and multiple linear regression (MLR) were adopted [17,18,19]. To study the correlations among the variables, the PCA was executed using XLSTAT. The PCA method was used to study the individual correlation between the log1/D37 and the molecular descriptors and excluded any variables that had a trivial effect on the log1/D37 value (low R value). The MLR was used to generate linear models where no descriptor was excluded, and was performed using SPSS [20]. Due to the vast number of molecular descriptors, a stepwise multiple linear regression approach was employed to find the pertinent parameters based on the forward-selection and the backward-elimination procedures. Statistical outliers were established as those compounds with definite standardized residuals greater than 2. The feature of the developed MLR models was judged using the correlation coefficient (R2), the standard error of the estimate (s), the adjusted R2 (Radj2), and Fisher's criterion (F).

    The versatility of 1-bromo-2-methylpropane (1Br), 2-bromo-2-methylpropane (2Br), and 2-chloro-2-methylpropane (2Cl) to attractively interact with α-glucan (G) was thoroughly assessed and comparatively investigated. The geometry optimization calculations were first executed for the 1Br/2Br/2Cl/G monomers and the 1Br…/2Br…/2ClG complexes at the M06-2X/6-31+G* level of theory [21]. An electrostatic potential (ESP) analysis was carried out for the optimized systems to extract the molecular electrostatic potential (MEP) maps. The electron density envelope with a value of 0.002 au was employed to generate the MEP maps, as previously reported in the literature [22].

    Relying on the optimized complexes, the interaction (Eint) energies were computed. The counterpoise correction method was invoked to eliminate the basis set superposition error (BSSE) [23], which is illustrated as follows:

    Eint=E1Br/2Br/2ClG(E1Br/2Br/2Clincomplex+EGincomplex)+EBSSE (2)

    where E1Br/2Br/2ClG, E1Br/2Br/2Clincomplex, and EGincomplex refer to the energies of the complex, 1Br/2Br/2Cl, and the G structures pertinent to their coordinates in the optimized complexes.

    Moreover, the quantum theory of atoms in molecules (QTAIM) calculations were carried out to deeply recognize the studied interactions by extracting the bond paths (BPs) and the bond critical points (BCPs). All executed DFT computations were performed using the Gaussian 09 package [24]. The QTAIM analysis was conducted by utilizing the Multiwfn 3.7 [25] and visualized by Visual Molecular Dynamics (VMD) [26] packages.

    In order to select the molecular descriptors and to build the QSAR models, the PCA and MLR methods were applied. The correlation coefficients (R) between the observed Aspergillus (A.) nidulans toxicities (log1/D37) of 41 halogenated aliphatic hydrocarbons and the molecular descriptors are gathered in Table 1. The PCA, including the correlation between the A. nidulans toxicities and the molecular descriptors, is illustrated in the biplot depicted in Figure 1. The data clusters were depicted by exhibiting the scores for the first components (F1) versus the second ones (F2).

    Table 1.  Correlation coefficients matrix of log 1/D37 and the computed parameters.
    Variables log1/D37 EHOMO ELUMO Egap S V log P MR MW α O
    log 1/D37 1
    EHOMO 0.019 1
    ELUMO −0.529 0.338 1
    Egap −0.561 −0.206 0.851 1
    S 0.536 0.090 0.256 0.216 1
    V 0.563 0.079 0.223 0.187 0.997 1
    log P 0.247 −0.202 0.183 0.303 0.672 0.667 1
    MR 0.027 0.107 0.372 0.327 0.650 0.654 0.621 1
    MW 0.756 −0.138 −0.753 −0.705 0.250 0.297 −0.041 −0.008 1
    α 0.694 0.013 −0.001 −0.008 0.934 0.951 0.666 0.650 0.468 1
    O 0.297 0.122 0.422 0.371 0.890 0.852 0.612 0.555 − 0.049 0.715 1

     | Show Table
    DownLoad: CSV
    Figure 1.  Correlation circle of log1/D37 and the computed descriptors.

    According to the data displayed in Figure 1, the molecular weight (MW) and the A. nidulans toxicities (log1/D37) were somewhat near to each other and were highly related (R = 0.756). This significant correlation coefficient demonstrated that the MW had a prominent influence on the A. nidulans toxicities. Furthermore, the surface area grid (S) and molecular volume (V) were superimposed on each other, thereby outlining the prominent positive correlation between these molecular descriptors with R = 0.997. Additionally, the biplot revealed that the S, V, log P, α, and O had positive correlations with log1/D37 (0.70 〉 R 〉 0.25).

    However, ELUMO and Egap had negative correlations with log1/D37 and were considered as important descriptors. Notwithstanding the importance of EHOMO and MR in various QSAR studies [27,28], the EHOMO and MR values were herein neglected due to their trivial effect on the log1/D37 value, which was attributed to their low correlation coefficients (R 〈 0.03). After excluding EHOMO and MR, the preferred regression model was constructed using a backward method based on three descriptors (See Eq 3). Particularly, considerable correlations were found between the MW and both ELUMO and Egap, with R values of −0.753 and −0.705, respectively.

    log1/D37=3.6309.175ELUMO10.052Egap+0.016S (3)
    n=41,R2=0.806,R2adj=0.790,s=0.332,F=51.261

    where n represents the number of halogenated aliphatic hydrocarbon molecules.

    Toward investigating the statistical significance of the selected descriptors in the considered model, the t-test values were calculated for each descriptor. Furthermore, the t-test is defined as the regression coefficient for the relevant descriptor divided by its error. The highest t-test value identifies the most important descriptors, where a molecular descriptor is considered significant when its t-test value exceeds |2|.

    In model 1 (See Eq 3), the t-test values were −2.697, −2.871, and 9.589 for ELUMO, Egap, and S, respectively. The obtained t-test values showed that all three descriptors played an important role in the regression models, with the surface area grid being the most relevant. It is worth mentioning that the positive values of the regression coefficient outlined that the molecular descriptors in the QSTR model provided a positive behavior towards the toxicity values. Meanwhile, negative values suggested that the log1/D37 values diminished by augmenting the value of the computed descriptor. From Eq 3, the log1/D37 value was disclosed to decline when the ELUMO and Egap values increased. Notwithstanding, the enhanced S value leads, in turn, to an increase in the value of log1/D37. These results were consistent with the results obtained using the PCA. The statistical results of Eq 3 for the validation set (14 molecules) were poor and were characterized by (R2 = 0.443, s = 0.2449, F = 9.5267).

    As mentioned earlier, the EHOMO and MR were found to be essential descriptors for the prediction of various biological activities. Consequently, the two aforesaid descriptors were considered, which gave rise to using all the calculated descriptors in the construction of the adopted QSAR models, bearing in mind that these two insignificant descriptors (EHOMO and MR) could account for the residual. Consequently, the preferred model was obtained based on three descriptors, which are described as follows:

    log1/D37=1.2429.769Egap0.047MR+0.329α, (4)
    n=41,R2=0.912,R2adj=0.904,s=0.224,F=127.106.

    The t-test values for Egap, MR, and α were −7.080, −7.157, and 15.137, respectively, thus indicating the significant contributions of the three considered descriptors to the log1/D37 value, with a highly appreciated role for the molecular polarizability descriptor. According to Eq 4, the log1/D37 decreases along with increasing the Egap and MR. On the other hand, increasing the values of the molecular polarizability causes a boost in the value of the log1/D37. These results were in good agreement with those obtained using the PCA. As an essential issue, Egap plays decisive roles in the chemical reactivity, biological activity, hydrophobicity, and electrophilicity of chemicals pertinent to the living cell activity and the associated mechanistic interactions. In that spirit, the Egap is considered one of the most important parameters that can inversely affect the chemical reactivity.

    The physical justification for using molar refractivity and polarizability in the developed QSAR models might be understood by considering that most of the biochemical process occurs in aqueous media. Therefore, the MR and α of aqueous toxic solutions are informative descriptors, which are useful in QSTR studies. As judged by the four statistical criteria (R2, Radj2, s, and F values), Eq 4 is statistically superior to Eq 3. Table 2 compares the observed log1/D37 values for 41 molecules to the predictions generated by Eq 4. Furthermore, Figure 2 visually depicts the relationship between these expectations and the experimental results of log1/D37.

    Table 2.  Observed and predicted values of log1/D37 for the training set. Predictions were made using Eq 4.
    Molecule name log1/D37 (Observed) log1/D37 (Predicted) MAE* Molecule name log1/D37 (Observed) log1/D37 (Predicted) MAE*
    Dichloromethane –1.97 –1.96 0.17 1,3-Dichloropropene –0.72 –0.83 0.16
    Chloroform –1.39 –1.36 0.18 1,1,3-Trichloropropene –0.38 –0.43 0.16
    Tetrachloromethane –0.49 –0.82 0.18 3-Chloro-2-chloromethylpropene –0.43 –0.49 0.17
    1,2-Dichloroethane –1.71 –1.64 0.18 1-Chloro-2-methylpropene –1.10 –0.85 0.17
    1,1,2-Trichloroethane –1.03 –1.12 0.18 Chlorodibromofluoromethane –0.49 0.07 0.17
    1,1,1,2-Tetrachloroethane –0.45 –0.56 0.18 Bromoform –0.72 –0.58 0.14
    Pentachloroethane –0.23 –0.11 0.18 Bromochloromethane –1.79 –1.60 0.14
    Hexachloroethane 0.10 0.27 0.18 Bromotrichloromethane 0.10 –0.05 0.14
    1,1,2-Trichloroethylene –1.05 –0.68 0.19 Bromodichloromethane –1.03 –1.02 0.14
    Tetrachloroethylene –0.08 –0.23 0.18 Chlorodibromomethane –0.46 –0.78 0.15
    1,1-Dichloroethylene –1.40 –1.23 0.18 1-Bromo-2-chloroethane –1.38 –1.26 0.13
    1,2-Dichloropropane –1.19 –1.21 0.18 1-Bromobutane –0.75 –0.95 0.14
    1,3-Dichloropropane –1.02 –1.13 0.19 2-Bromo-1-chloropropane –0.99 –0.84 0.13
    1,2,3-Trichloropropane –0.97 –0.73 0.19 1-Bromo-2-methylpropane –1.26 –0.93 0.12
    2-Chlorobutane –1.28 –1.38 0.19 1-Bromo-4-chlorobutane –0.52 –0.55 0.09
    1,3-Dichlorobutane –0.94 –0.76 0.19 1-Bromo-3-methylbutane –0.49 –0.54 0.10
    1-Chloro-2-methylpropane –1.16 –1.32 0.19 Dibromodichloromethane 0.70 0.61 0.12
    1-Chloropentane –0.70 –0.95 0.19 Dibromomethane –1.33 –1.39 0.13
    1-Chlorohexane –0.18 –0.56 0.19 Tetrabromomethane 2.00 1.79 0.16
    1-Chlorooctane –0.18 0.22 0.18 1,1,2,2-Tetrabromoethane 0.40 0.29 0.11
    2,3-Dichloropropene –0.43 –0.82 0.17
    * MAE stands for mean absolute error.

     | Show Table
    DownLoad: CSV
    Figure 2.  Observed values of log1/D37 for 41 molecules vis predictions made by Eq 4.

    For a comparison purpose between the developed model (See Eq 4) and those constructed by Cronin et al. [10], the QSAR model was built for the whole data set, and the model was characterized by R2 = 0.897, s = 0.219, and F = 148.766. The currently developed model was superior to those developed by Cronin et al., who reported R2 = 0.615, s = 0.413, and F = 44.2 (Table S1).

    The regression coefficients for the 41 molecules were adopted to compute the log1/D37 for 14 halogenated aliphatic molecules to assess the predictive performance of the three-variable QSTR model (See Eq 4). Table 3 contains the gathered data, which is graphically displayed in Figure 3.

    Table 3.  Observed and predicted values of log1/D37 for the validation set. Predictions were made using Eq 4.
    Molecule name log1/D37
    (Observed)
    log1/D37
    (Predicted)
    MAE*
    1,1-Dichloroethane –1.68 –1.61 0.16
    1,1,1-Trichloroethane –1.00 –0.96 0.17
    1,1,2,2-Tetrachloroethane –0.45 –0.70 0.18
    1,2-Dichloroethylene –1.48 –1.11 0.17
    2,2-Dichloropropane –1.28 –1.19 0.15
    1-Chlorobutane –1.16 –1.31 0.16
    2,3-Dichlorobutane –0.94 –0.73 0.16
    2-Chloro-2-methylpropane –1.26 –1.32 0.15
    1,1-Dichloropropene –0.82 –0.89 0.17
    3-Chloro-2-methylpropene –0.88 –0.95 0.19
    2-Bromobutane –1.06 –0.93 0.21
    1-Bromo-3-chloropropane –0.88 –0.82 0.24
    2-Bromo-2-methylpropane –1.41 –0.86 0.33
    1,2-Dibromoethylene –0.96 –0.84 0.12
    * MAE stands for mean absolute error.

     | Show Table
    DownLoad: CSV
    Figure 3.  Observed values of log1/D37 for validation set versus predictions made by Eq 4. The outlier is represented as a triangle.

    According to Table 3, the results were labeled with an R2 value of 0.575 and a standard error of 0.214. Therefore, 2-bromo-2-methylpropane (2Br) had an outlier behavior in the verification set (See Figure 3) with a standard residual of –2.167. The mathematical performance of the model was marginally improved by excluding this particular compound from the predictive equation (R2 = 0.737, s = 0.168). Still, of course, there were no posterior justifications for reselections of the data set.

    In accordance with the strong statistical significance, the presented QSTR models had a substantial reliability and an internal extrapolative capacity. As illustrated in Figure 3, the observed log1/D37 values for the verification set were consistent with the projected values.

    Furthermore, additional calculations were performed on 2-bromo-2-methyl propane (2Br) in order to clarify the potential causes for its outlier behavior in the QSTR models. Due to the structural similarities of 1-bromo-2-methyl propane (1Br) and 2-chloro-2-methyl propane (2Cl) to 2Br, the same procedure was performed on 2Cl and 1Br. Several studies have claimed that the outlier molecules can emerge for a variety of reasons [29,30,31,32,33]. Outliers, for instance, may be caused by inaccurate molecular descriptor values, or they may suggest an error in the experimental toxicities. They could either be the result of an uncommon mechanism or a distinct binding type. First, a deep inspection of the values of the Egap energy was performed to examine the effect of the stability of the HAHs on the value of log 1/D37. A stable molecule is indicated by an alkyl halide with a high value of Egap, while a compound with a low Egap value has a high effectiveness. Remarkable correlations (R = –0.561) were observed between the stability of the alkyl halide and its toxicity. More stable molecules turned out to have a high log 1/D37 value. These results were previously outlined using a PCA (See Figure 1). The Egap for 2Br, 1Br, and 2Cl were 0.187, 0.195, and 0.225 au, respectively. These values indicated that 2Br had the lowest energy gap and, therefore, demonstrated the highest efficacy (i.e., lowest stability). On the other hand, 2Cl turned out to be the most stable one. The log 1/D37 values of 2Br, 1Br, and 2Cl were −1.41, −1.26, and −1.26, respectively. The identical results of log1/D37 for the latter two molecules may suggest errors in the experimental toxicities' values.

    The structural compositions of A. nidulans cell walls have been studied and reported in various studies [34,35,36,37,38]. The cell wall is necessary for fungi to survive in their natural habitat [34]. Fungi cell walls account for approximately one-quarter of the total fungal biomass [35], and approximately one-third of the fungal genome is engaged in cell wall formation and maintenance [36]. A carbohydrate investigation of the A. nidulans wall concluded that it contains roughly 40% α-glucan and β-glucan [34,37,38]. In this study, α-glucan was utilized to explore the binding interactions between the A. nidulans cell wall and the selected halide alkyls.

    An electrostatic potential (ESP) analysis was performed to unveil the nucleophilic and electrophilic regions over the molecular systems. By employing the ESP analysis, the molecular electrostatic potential (MEP) maps of the optimized monomers were extracted utilizing a 0.002 au electron density envelope. Figure 4 displays the MEP maps for the optimized geometries of 1Br, 2Br, 2Cl, and G.

    Figure 4.  MEP maps of 1-bromo-2-methylpropane (1Br), 2-bromo-2-methylpropane (2Br), 2-chloro-2-methylpropane (2Cl), and α-glucan (G). The color scale is extended from the red (–0.01 au) to the blue (0.01 au) scope.

    As depicted in Figure 4, the molecular surfaces with relative electrophilic and nucleophilic regions over the molecular surface of the studied systems were conspicuously observed by the presence of blue-coded surfaces (i.e., positive ESP) and red-coded surfaces (i.e., negative ESP), respectively. Notably, negative ESP regions were observed over the surfaces of the Br, Cl, and O atoms of the investigated molecules. On the other hand, the positive ESP areas were found around the surfaces of the H atoms. The obtained findings paraded the supreme penchant of the 1Br, 2Br, and 2Cl molecules to attractively interact with the G molecule through hydrogen bonding interactions.

    The preferability of the 1Br, 2Br, and 2Cl molecules to interact with the G molecule was herein thoroughly addressed through the 1Br…/2Br…/2ClG complexes. First, geometry optimization calculations were performed. Using optimized complexes, the interaction energy (Eint) values were computed. Figure S1 shows the optimized geometries of the studied complexes within all the plausible configurations and their Eint values. Relying on the negative Eint values, the most preferable configurations of the 1Br…/2Br…/2ClG complexes were selected and gathered in Figure 5.

    Figure 5.  Optimized complexes within the most preferable configurations of the 1Br…, 2Br…, and 2ClG complexes. Interaction energy (Eint) values are in kcal/mol.

    As shown in Figure S1, the 1Br, 2Br, and 2Cl molecules showed a superior potentiality to interact with the G molecule through the depicted configurations with respectable negative interaction energy values. Notably, the stability of the scouted complexes was generally explained as an upshot to the presence of numerous hydrogen bonds, which were found to vary in numbers and distances. It is worth noting that the preferable interactions within the studied complexes were ascribed to the 1BrG complex, followed by 2BrG and 2ClG complexes (Figure 5). Numerically, Eint of the 1Br…, 2Br…, and 2ClG complexes were –10.49, –10.21, and –9.99 kcal/mol, respectively.

    To gain further insight into the origin of the studied interactions, a quantum theory of atoms in molecules (QTAIM) analysis was executed for the optimized 1Br…/2Br…/2ClG complexes within the most preferable configurations. In that spirit, the BPs and BCPs were generated and are illustrated in Figure 6.

    Figure 6.  QTAIM plots of the most preferable configurations of the optimized 1Br…, 2Br…, and 2ClG complexes.

    As shown in Figure 6, the occurrence of interactions within the investigated complexes was assured by the presence of BPs and BCPs (i.e., hydrogen bonds). The higher number of the BPs and BCPs within the 1BrG complex over the other analogs could interpret its higher preferability in comparison to the others. This observation was in line with the interaction energy affirmations.

    Aspergillus (A.) nidulans toxicities (log1/D37) were precisely predicted for a set of 55 halogenated aliphatic hydrocarbons (HAHs) using DFT calculations. Besides, the individual correlations between a series of evaluated descriptors and the A. nidulans toxicities (log1/D37) were unveiled. The importance of quantum-chemical parameters along with various physical properties, including the surface area grid (S), molecular volume (V), molar refractivity (MR), molecular weight (MW), and molecular ovality (O) were investigated to predict the A. nidulans toxicities of halogenated aliphatic hydrocarbons. A potential three-variable QSTR model was trained using 41 molecules and subsequently validated for a set of 14 compounds. Based on the Egap, MR, and α, the linear model had correlation coefficient (R2) values of 0.912 and 0.575 for the training and the validation sets, respectively. The generated QSAR model could be used for the log1/D37 predictions of HAHs provided in this paper; however, more research would be required to test more aliphatic compounds for an external validation. The most preferred interactions were clearly identified within the 1-bromo-2-methylpropane…α-glucan complex. The greater number of bond pathways and bond critical sites within the 1-bromo-2-methylpropane…α-glucan complex over other analogs affirmed its significant favorability.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    All authors have no conflict of interest to report.



    [1] Premkumar R, Hussain S, Koyambo-Konzapa SJ, et al. (2021) SERS and DFT studies of 2-(trichloroacetyl)pyrrole chemisorbed on the surface of silver and gold coated thin films: In perspective of biosensor applications. J Mol Recognit 34: e2921. https://doi.org/10.1002/jmr.2921 doi: 10.1002/jmr.2921
    [2] MacFarland HN (1986) Toxicology of solvents. Am Ind Hyg Assoc J 47: 704–707. https://doi.org/10.1080/15298668691390511 doi: 10.1080/15298668691390511
    [3] Crebelli R, Andreoli C, Carere A, et al. (1995) Toxicology of halogenated aliphatic hydrocarbons: structural and molecular determinants for the disturbance of chromosome segregation and the induction of lipid peroxidation. Chem-Biol Interact 98: 113–129. https://doi.org/10.1016/0009-2797(95)03639-3 doi: 10.1016/0009-2797(95)03639-3
    [4] Trohalaki S, Gifford E, Pachter R (2000) Improved QSARs for predictive toxicology of halogenated hydrocarbons. Comput Chem 24: 421–427. https://doi.org/10.1016/s0097-8485(99)00093-5 doi: 10.1016/s0097-8485(99)00093-5
    [5] Gadaleta D, Vukovic K, Toma C, et al. (2019) SAR and QSAR modeling of a large collection of LD(50) rat acute oral toxicity data. J Cheminformatics 11: 58. https://doi.org/10.1186/s13321-019-0383-2 doi: 10.1186/s13321-019-0383-2
    [6] Li F, Wang P, Fan T, et al. (2024) Prioritization of the ecotoxicological hazard of PAHs towards aquatic species spanning three trophic levels using 2D-QSTR, read-across and machine learning-driven modelling approaches. J Hazard Mater 465: 133410. https://doi.org/10.1016/j.jhazmat.2023.133410 doi: 10.1016/j.jhazmat.2023.133410
    [7] Wang QQ, Fan TJ, Jia RQ, et al. (2024) First report on the QSAR modelling and multistep virtual screening of the inhibitors of nonstructural protein Nsp14 of SARS-CoV-2: Reducing unnecessary chemical synthesis and experimental tests. Arab J Chem 17: 105614. https://doi.org/10.1016/j.arabjc.2024.105614 doi: 10.1016/j.arabjc.2024.105614
    [8] Oubahmane M, Hdoufane I, Delaite C, et al. (2023) Design of Potent Inhibitors Targeting the Main Protease of SARS-CoV-2 Using QSAR Modeling, Molecular Docking, and Molecular Dynamics Simulations. Pharmaceuticals.
    [9] Dearden JC, Cronin MT, Dobbs AJ (1995) Quantitative structure-activity relationships as a tool to assess the comparative toxicity of organic chemicals. Chemosphere 31: 2521–2528. https://doi.org/10.1016/0045-6535(95)00121-n doi: 10.1016/0045-6535(95)00121-n
    [10] Cronin MT, Dearden JC, Duffy JC, et al. (2002) The importance of hydrophobicity and electrophilicity descriptors in mechanistically-based QSARs for toxicological endpoints. SAR QSAR Environ Res 13: 167–176. https://doi.org/10.1080/10629360290002316 doi: 10.1080/10629360290002316
    [11] Martin TM, Harten P, Young DM, et al. (2012) Does rational selection of training and test sets improve the outcome of QSAR modeling? J Chem Inf Model 52: 2570–2578. https://doi.org/10.1021/ci300338w doi: 10.1021/ci300338w
    [12] te Velde G, Bickelhaupt FM, Baerends EJ, et al. (2001) Chemistry with ADF. J Comput Chem 22: 931–967. https://doi.org/10.1002/jcc.1056 doi: 10.1002/jcc.1056
    [13] Fonseca Guerra C, Snijders JG, te Velde G, et al. (1998) Towards an order- N DFT method. Theor Chem Acc 99: 391–403. https://doi.org/10.1007/s002140050353 doi: 10.1007/s002140050353
    [14] Perdew JP, Wang Y (1992) Accurate and simple analytic representation of the electron-gas correlation energy. Phys Rev B 45: 13244–13249. https://doi.org/10.1103/physrevb.45.13244 doi: 10.1103/physrevb.45.13244
    [15] Perdew JP, Chevary JA, Vosko SH, et al. (1992) Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation. Phys Rev B 46: 6671–6687. https://doi.org/10.1103/physrevb.46.6671 doi: 10.1103/physrevb.46.6671
    [16] Bodor N, Gabanyi Z, Wong CK (1989) A New Method for the Estimation of Partition-Coefficient. J Am Chem Soc 111: 3783–3786. https://doi.org/10.1021/ja00193a003 doi: 10.1021/ja00193a003
    [17] Elmi Z, Faez K, Goodarzi M, et al. (2009) Feature selection method based on fuzzy entropy for regression in QSAR studies. Mol Phys 107: 1787–1798. https://doi.org/10.1080/00268970903078559 doi: 10.1080/00268970903078559
    [18] Goudarzi N, Goodarzi M, Chen T (2012) QSAR prediction of HIV inhibition activity of styrylquinoline derivatives by genetic algorithm coupled with multiple linear regressions. Med Chem Res 21: 437–443. https://doi.org/10.1007/s00044-010-9542-8 doi: 10.1007/s00044-010-9542-8
    [19] Cai Z, Zafferani M, Akande OM, et al. (2022) Quantitative Structure-Activity Relationship (QSAR) Study Predicts Small-Molecule Binding to RNA Structure. J Med Chem 65: 7262–7277. https://doi.org/10.1021/acs.jmedchem.2c00254 doi: 10.1021/acs.jmedchem.2c00254
    [20] Nie NH, Bent DH, Hull CH (1970) SPSS: Statistical Package for the Social Sciences: McGraw-Hill.
    [21] Zhao Y, Truhlar DG (2008) Exploring the limit of accuracy of the global hybrid meta density functional for main-group thermochemistry, kinetics, and noncovalent interactions. J Chem Theory Comput 4: 1849–1868. https://doi.org/10.1021/ct800246v doi: 10.1021/ct800246v
    [22] Ibrahim MAA (2012) Molecular mechanical perspective on halogen bonding. J Mol Model 18: 4625–4638. https://doi.org/10.1007/s00894-012-1454-8 doi: 10.1007/s00894-012-1454-8
    [23] Boys SF, Bernardi F (2006) The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors. Mol Phys 19: 553–566. https://doi.org/10.1080/00268977000101561
    [24] Frisch MJ, Trucks GW, Schlegel HB, et al. (2009) Gaussian 09. Revision E01 ed. Wallingford CT, USA.: Gaussian09, Gaussian Inc.
    [25] Lu T, Chen F (2012) Multiwfn: a multifunctional wavefunction analyzer. J Comput Chem 33: 580–592. https://doi.org/10.1002/jcc.22885 doi: 10.1002/jcc.22885
    [26] Humphrey W, Dalke A, Schulten K (1996) VMD: Visual molecular dynamics. J Mol Graph 14: 33–38. https://doi.org/10.1016/0263-7855(96)00018-5 doi: 10.1016/0263-7855(96)00018-5
    [27] Al-Fahemi JH (2012) The use of quantum-chemical descriptors for predicting the photoinduced toxicity of PAHs. J Mol Model 18: 4121–4129. https://doi.org/10.1007/s00894-012-1417-0 doi: 10.1007/s00894-012-1417-0
    [28] Al-Fahemi JH (2013) Structural descriptors for the correlation of human blood: air partition coefficient of volatile organic molecules by QSPRs. Struct Chem 24: 2155–2161. https://doi.org/10.1007/s11224-013-0224-2 doi: 10.1007/s11224-013-0224-2
    [29] Kim KH (2021) Outliers in SAR and QSAR: 3. Importance of considering the role of water molecules in protein-ligand interactions and quantitative structure-activity relationship studies. J Comput-Aid Mol Des 35: 371-396. https://doi.org/10.1007/s10822-021-00377-7
    [30] Zhao L, Wang W, Sedykh A, et al. (2017) Experimental Errors in QSAR Modeling Sets: What We Can Do and What We Cannot Do. ACS Omega 2: 2805–2812. https://doi.org/10.1021/acsomega.7b00274 doi: 10.1021/acsomega.7b00274
    [31] Kim KH (2007) Outliers in SAR and QSAR: 2. Is a flexible binding site a possible source of outliers? J Comput-Aid Mol Des 21: 421–435. https://doi.org/10.1007/s10822-007-9126-y
    [32] Kim KH (2007) Outliers in SAR and QSAR: is unusual binding mode a possible source of outliers? J Comput-Aid Mol Des 21: 63–86. https://doi.org/10.1007/s10822-007-9106-2 doi: 10.1007/s10822-007-9106-2
    [33] Kim KH (2022) Outliers in SAR and QSAR: 4. effects of allosteric protein-ligand interactions on the classical quantitative structure-activity relationships. Mol Divers 26: 3057–3092. https://doi.org/10.1007/s11030-021-10365-6
    [34] Alam MK, van Straaten KE, Sanders DA, et al. (2014) Aspergillus nidulans cell wall composition and function change in response to hosting several Aspergillus fumigatus UDP-galactopyranose mutase activity mutants. PLoS One 9: e85735. https://doi.org/10.1371/journal.pone.0085735 doi: 10.1371/journal.pone.0085735
    [35] Gastebois A, Clavaud C, Aimanianda V, et al. (2009) Aspergillus fumigatus: cell wall polysaccharides, their biosynthesis and organization. Future Microbiol 4: 583–595. https://doi.org/10.2217/fmb.09.29 doi: 10.2217/fmb.09.29
    [36] de Groot PW, Brandt BW, Horiuchi H, et al. (2009) Comprehensive genomic analysis of cell wall genes in Aspergillus nidulans. Fungal Genet Biol 46 Suppl 1: S72–81. https://doi.org/10.1016/j.fgb.2008.07.022 doi: 10.1016/j.fgb.2008.07.022
    [37] Guest GM, Momany M (2000) Analysis of cell wall sugars in the pathogen Aspergillus fumigatus and the saprophyte Aspergillus nidulans. Mycologia 92: 1047–1050. https://doi.org/10.2307/3761470 doi: 10.2307/3761470
    [38] Free SJ (2013) Chapter Two - Fungal Cell Wall Organization and Biosynthesis. In: Friedmann T, Dunlap JC, Goodwin SF, editors. Advances in Genetics: Academic Press. pp. 33–82.
  • Environ-12-03-019-s001.pdf
  • Reader Comments
  • © 2025 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(162) PDF downloads(57) Cited by(0)

Figures and Tables

Figures(6)  /  Tables(3)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog