Translate this page into:
Evaluation of non-covalent interactions of chlorambucil (monomer and dimer) and its interaction with biological targets: Vibrational frequency shift, electron density topological and automated docking analysis
⁎Corresponding author. karthiphy84@gmail.com (T. Karthick)
-
Received: ,
Accepted: ,
This article was originally published by Elsevier and was migrated to Scientific Scholar after the change of Publisher.
Peer review under responsibility of King Saud University.
Abstract
Chlorambucil is a well-known chemotherapy drug that is being used to treat chronic myelogenous leukemia. As it contains ten flexible rotational bonds, the possible spatial conformations have been identified theoretically. The spectral signatures of monomer and dimer structures of chlorambucil and the frequency shifts due to non-covalent interactions (NCIs) have been illustrated using FT-IR and FT-Raman spectra. The bond correlation between carbon and hydrogen nuclei of chlorambucil has been obtained using 2D-HSQC NMR spectrum. The assignments of harmonic normal modes have been done in order to find the vibrational contributions of each functional group. Besides the spectroscopic studies, the electron density based quantum topological atoms in molecule analysis have been performed to explore the possible interactions between the non-bonded atoms. The reduced density gradient and isosurface plots have been used in this study to understand the strength of NCIs. The charge delocalization patterns of monomer and dimer structures were explained so as to investigate the chemical stability profile. The active sites for the electrophilic and nucleophilic attack on the monomer conformers have been determined by applying Hirshfeld charges and atomic spin densities into Fukui and Parr functions, respectively. From the automated docking analysis, it is found that chlorambucil interacts with the aldo-keto reductase family 1 (AKR1B1, AKR1B10, AKR1B15) and FAD-linked sulfhydryl oxidase ALR proteins through strong hydrogen bonds and shows a potential inhibition. In order to take into account the interactions ranging from short to long range, the modern density functionals viz. M06-2X, wB97XD, B97D which includes dispersion-corrected repulsion terms have been employed and the theoretical results were found coincide with the experimental observations.
Keywords
Chlorambucil
Conformational analysis
Non-covalent interactions
FT-IR
FT-Raman
Protein-ligand interactions
1 Introduction
The study of non-covalent interactions (NCIs) plays a key role to unlock the possibility of rational drug design. In the recent years, the utilization of spectroscopic and theoretical topological tools for exploring the NCIs was increased considerably due to its ease and accuracy (Karthick et al., 2017a, 2017b; Srivastava et al., 2016). The fundamental spectral features such as IR and Raman allow us to know the occurrence of those interactions in particular molecular systems in terms of vibrational frequency shift (δ). Since the value of δ depends on the strength of NCIs present, sometimes it is very difficult to obtain the information concerning long-range interactions. Hence, the modern density functionals which incorporate long-range dispersion-corrected repulsion terms have been used to calculate the optimized structural properties, vibrational frequencies and stabilization energies due to charge delocalization within the molecule. Moreover, in the present study, quantum topological atoms in molecule (QTAIM) analysis have been performed to explore the interactions ranging from short to long range which occurs between the non-bonded atoms. The reduced density gradient and isodensity surface plots have been used to distinguish the interactions taking place within the molecule.
Chlorambucil is a well-known nitrogen mustard alkylating agent that is being used for treating chronic myelogenous leukemia (CML) (Kufe et al., 2003). The electronic structure of chlorambucil and its tentative vibrational assignments have been investigated (Gunasekaran et al., 2008a, 2008b). Also, the important functional group (−CH2CH2COOH) which is expected to be involved in the formation of dimer was considered as a single mass point group and its assignments were not proposed. In the same year, tentative vibrational assignments of all the functional groups of chlorambucil were also proposed (Gunasekaran et al., 2008a, 2008b). Unfortunately, the proposed tentative assignments were not supported by the experimental especially the modes of —COOH group. Hence, there were lots of discrepancies between the experimental and theoretical results assignments proposed by Gunasekaran et al. (Gunasekaran et al., 2008a, 2008b). This clearly indicates the dimer form of chlorambucil through —COOH group.
Charak et al. studied the interactions of chlorambucil with DNA with the help of spectroscopic and molecular docking approaches (Charak et al., 2012). The non-covalent interactions of chlorambucil with calf thymus DNA have been investigated by multi-spectroscopic techniques and molecular docking study (Rehman et al., 2015).
However, the spatial conformations along the flexible bonds, dimer structure and the details of NCIs of chlorambucil were not yet revealed. As it has nine flexible bonds, there are numerous conformers possibly to exist at the room temperature. Hence, an attempt has been made to find the feasibly existing stable conformers of chlorambucil along the flexible bonds without any steric hindrance. The expected vibrational frequency shifts due to the existence of NCIs have been explained with the detailed potential energy distribution (PED) results. Since the molecule that contains carboxylic acid group exhibit as dimer form in the solid state (Balachandran et al., 2011; Balachandran et al., 2012; Karthick et al., 2015), the electronic structure of chlorambucil dimer has been proposed and the spectral features of dimer has been compared with the experimental spectra. In addition to NCIs, the interaction energies due to charge delocalization between the donor and acceptor species are also taking part in the chemical stability of a molecule. Hence, the natural population and second order perturbation theory analysis have been carried out to find the donor-acceptor species which are more predominant to the chemical stability.
2 Experimental
A pure sample of chlorambucil was purchased from Sigma Aldrich chemicals Ltd. For FT-IR measurement, the mixer of chlorambucil and potassium bromide in the ratio 100:1 was used to make pellet which readily usable for recording the spectra. The FT-IR spectrum of chlorambucil was recorded in a mid-IR region on a Bruker Tensor 27 spectrophotometer which outfitted with mid-IR source (4000 to 400 cm−1), a KBr beam splitter and a room temperature DTGS detector. The multi-tasking OPUS software was used for the observed peak averaging, peak enhancement, baseline correction and other spectral manipulations. A spectral resolution of 1.0 cm−1 was used and the source was allowed to incident on the pellet surfaces for about 32 scans during the measurement. The FT-Raman spectrum of chlorambucil was recorded in the region 3500–100 cm−1 on Bruker RFS multi RAM standalone spectrophotometer. The spectral resolution of 2 cm−1 bandwidth of 1.5 nm was fixed and a couple of silicon diode detectors are used to record the spectrum during the process. The Nd:YAG laser of 1064 nm wavelength is used as excitation source. 2D-HSQC NMR spectrum of chlorambucil has been performed using Bruker AVANCE III 500 MHz (AV 500) multi nuclei solution NMR Spectrometer.
3 Computational details
Initially, a careful attempt has been made to find the stable conformers of chlorambucil. Since chlorambucil has a large number of flexible bonds, the high-performance conformational analysis tool “CONFLEX” (Goto et al., 2012; Goto and Osawa, 1989,1993) has been used. From the list of 1420 conformers, only three conformers are taken from the lowest energy order for discussion. The list of conformers and their energy profile along with the percentage of population is depicted in Table S1. The conformers obtained from CONFLEX are further optimized by dispersion corrected modern density functional B97D, wB97XD and M06-2X (plus D3 empirical dispersion function) with 6-311++G(d,p) basis set combinations using Gaussian 09W software (Frisch et al., 2003). An implementation of 6-311++G(d,p) basis set is adequate for the analysis of molecules like chlorambucil as it has been successful utilized in earlier reports (Brovarets and Hovorun, 2011, 2014). All the harmonic frequencies obtained in the calculations of both monomer and dimer are positive which proved that the geometries were converged to energy minima and holds stable normal modes. The assignments of vibrational normal modes were attained by Gar2Ped program (Martin and Van Alsenoy, 1995). With the intention of finding non-covalent interactions, the electron density based topological analysis has been carried out using AIMALL package (Keith, 2013; Biegler-Konig et al., 2001). Plots of a sign(λ2)ρ (i.e. the sign second greatest Eigen value of the electron density Hessian matrix multiplied by the electron density) versus reduced density gradient (s) reveals the signature of NCIs. The cube files for both the functions were generated using Multiwfn software (Lu and Chen, 2012) and isodensity surface plot for the non-covalent interaction studies have been done using VMD visualization software (Humphrey et al., 1996). The second order perturbation interaction energies between the donor and acceptor species within the molecule has been found by natural bond orbital (NBO) analysis using wB97XD method with 6-311++G(d,p) basis set combinations.
4 Results and discussion
4.1 Geometry optimization
The spatial conformations around all the possible rotational bonds have been searched using MMFF94s force field and full-matrix Newton-Raphson optimization method available in CONFLEX (Goto et al., 2012; Goto and Osawa, 1989,1993). Further optimization has been done using wB97XD and B97D functional with Grimme's D2 (GD2) dispersion correction and M06-2x functional with Grimme's D3 (GD3) dispersion correction so as to consider the long-range interactions into account. The optimized structural bond lengths of monomer and dimer structures of chlorambucil are given in Table 1 and their bond angles and dihedral angles are depicted in Table S2 (Supplementary Material). The lowest computed energy obtained for the conformer A shows its dominant stability than the conformers B and C. The relative energies of the conformers with respect to the most stable conformer A at different levels of theories are depicted in Table 2. The optimized molecular structures of these three conformers and the overlapping of one over another are shown in Figs. 1 and 2, respectively. It is clear from the Fig. 2 that all the studied conformers are structurally independent. Hence, we may expect distinct physicochemical properties, intramolecular bond patterns and spectroscopic fingerprints from each conformer. When we look at the Tables 1 and S2 (Supplementary Material), we can easily observe the noticeable changes where the non-covalent interactions taking place. For example, due to the diverse orientations, the bond length of C9C16 of conformer A and B is deviated from the conformer C by 0.0120 Å. Likewise, the significant changes were also identified in conformer A with respect to other bonds C16H32 and C16H33. Furthermore, the bond angle parameters C16C9H22, C16C9H23, and C9C16C19 of each conformer are found to be deviated by an angle 2–4° from one another because of different orientations and intramolecular bonding patterns. To avoid the complexity in understanding the three-dimensional structures for the readers, the Cartesian coordinates of all the conformers are depicted in Tables S3 (Supplementary Material).
Bond length parameters
Theoretical bond length (Å)
M062x-D3/6-311++G(d,p)
wB97XD/6-311++G(d,p)
B97D/6-311++G(d,p)
Conf. A
Monomer
Conf. A
Monomer
Conf. A
Monomer
Monomer
Dimer
Conf. B
Conf. C
Monomer
Dimer
Conf. B
Conf. C
Monomer
Dimer
Conf. B
Conf. C
Cl1C17
1.7947
1.7949
1.7945
1.7945
1.7988
1.7986
1.7989
1.7988
1.8266
1.8259
1.8270
1.8267
Cl2C18
1.7946
1.7910
1.7946
1.7949
1.7988
1.8024
1.7988
1.7991
1.8267
1.8320
1.8265
1.8275
O3C19
1.3487
1.3803
1.3485
1.3498
1.3481
1.3161
1.3480
1.3483
1.3706
1.3313
1.3707
1.3707
O3H38
0.9667
0.9390
0.9667
0.9667
0.9646
0.9923
0.9646
0.9647
0.9729
1.0053
0.9728
0.9728
O4C19
1.2016
1.1831
1.2016
1.2016
1.2033
1.2218
1.2034
1.2042
1.2134
1.2343
1.2133
1.2145
N5C6
1.3938
1.3912
1.3938
1.3937
1.3940
1.3966
1.3938
1.3942
1.4011
1.4022
1.4013
1.4014
N5C10
1.4470
1.4458
1.4468
1.4472
1.4454
1.4466
1.4458
1.4461
1.4575
1.4573
1.4576
1.4579
N5C11
1.4470
1.4431
1.4471
1.4467
1.4457
1.4496
1.4457
1.4453
1.4577
1.4607
1.4574
1.4573
C6C12
1.4052
1.4067
1.4070
1.4055
1.4030
1.4015
1.4053
1.4035
1.4163
1.4130
1.4180
1.4168
C6C13
1.4071
1.4088
1.4051
1.4065
1.4052
1.4035
1.4032
1.4046
1.4182
1.4167
1.4166
1.4176
C7C8
1.5073
1.5074
1.5075
1.5088
1.5077
1.5076
1.5079
1.5086
1.5127
1.5155
1.5126
1.5135
C7C9
1.5406
1.5397
1.5403
1.5372
1.5385
1.5394
1.5384
1.5367
1.5497
1.5516
1.5502
1.5491
C7H20
1.0910
1.0916
1.0937
1.0954
1.0923
1.0917
1.0946
1.0949
1.0974
1.0958
1.1001
1.0999
C7H21
1.0937
1.0936
1.0911
1.0934
1.0945
1.0946
1.0923
1.0944
1.1004
1.1018
1.0976
1.1000
C8C14
1.3930
1.3924
1.3941
1.3939
1.3914
1.3920
1.3930
1.3925
1.4031
1.4037
1.4043
1.4045
C8C15
1.3941
1.3944
1.3930
1.3936
1.3930
1.3927
1.3915
1.3922
1.4045
1.4051
1.4033
1.4038
C9C16
1.5254
1.5256
1.5253
1.5355
1.5255
1.5253
1.5250
1.5376
1.5360
1.5375
1.5357
1.5498
C9H22
1.0929
1.0929
1.0940
1.0927
1.0935
1.0935
1.0950
1.0940
1.0986
1.0986
1.0998
1.0991
C9H23
1.0940
1.0940
1.0929
1.0935
1.0950
1.0950
1.0935
1.0944
1.0999
1.0998
1.0986
1.0997
C10C17
1.5260
1.5248
1.5261
1.5260
1.5267
1.5279
1.5266
1.5264
1.5356
1.5367
1.5357
1.5356
C10H24
1.0920
1.0929
1.0919
1.0918
1.0924
1.0915
1.0924
1.0926
1.0974
1.0966
1.0977
1.0975
C10H25
1.0925
1.0929
1.0925
1.0926
1.0933
1.0929
1.0931
1.0933
1.0975
1.0980
1.0975
1.0976
C11C18
1.5262
1.5271
1.5260
1.5261
1.5267
1.5258
1.5264
1.5268
1.5356
1.5353
1.5359
1.5354
C11H26
1.0919
1.0916
1.0920
1.0920
1.0923
1.0926
1.0924
1.0923
1.0976
1.0985
1.0975
1.0975
C11H27
1.0924
1.0928
1.0926
1.0926
1.0933
1.0929
1.0931
1.0932
1.0976
1.0985
1.0974
1.0975
C12C14
1.3893
1.3908
1.3877
1.3888
1.3882
1.3867
1.3863
1.3876
1.3982
1.3968
1.3970
1.3977
C12H28
1.0820
1.0831
1.0821
1.0820
1.0824
1.0813
1.0825
1.0824
1.0866
1.0846
1.0868
1.0867
C13C15
1.3878
1.3870
1.3894
1.3879
1.3863
1.3871
1.3882
1.3868
1.3969
1.3977
1.3983
1.3975
C13H29
1.0821
1.0822
1.0820
1.0821
1.0825
1.0824
1.0825
1.0824
1.0867
1.0869
1.0866
1.0868
C14H30
1.0863
1.0849
1.0862
1.0857
1.0862
1.0876
1.0865
1.0856
1.0905
1.0905
1.0909
1.0892
C15H31
1.0862
1.0862
1.0863
1.0863
1.0864
1.0864
1.0862
1.0866
1.0909
1.0906
1.0905
1.0911
C16C19
1.5058
1.5073
1.5059
1.5052
1.5058
1.5043
1.5057
1.5044
1.5175
1.5151
1.5174
1.5145
C16H32
1.0934
1.0932
1.0946
1.0900
1.0939
1.0941
1.0952
1.0907
1.0988
1.0982
1.1011
1.0953
C16H33
1.0946
1.0948
1.0935
1.0927
1.0953
1.0951
1.0941
1.0925
1.1012
1.1012
1.0988
1.0982
C17H34
1.0882
1.0884
1.0882
1.0881
1.0887
1.0885
1.0886
1.0887
1.0938
1.0935
1.0937
1.0938
C17H35
1.0887
1.0887
1.0887
1.0888
1.0894
1.0894
1.0894
1.0894
1.0950
1.0950
1.0950
1.0950
C18H36
1.0882
1.0886
1.0882
1.0881
1.0887
1.0883
1.0887
1.0886
1.0937
1.0938
1.0936
1.0938
C18H37
1.0887
1.0891
1.0887
1.0887
1.0894
1.0890
1.0894
1.0894
1.0950
1.0947
1.0951
1.0949
O42H38
–
1.6759
–
–
–
1.6788
–
–
–
1.6684
–
–
O4H76
–
1.6828
–
–
–
1.6836
–
–
–
1.6755
–
–
Conformers
Theoretical
MP2/6-311++G(d,p)
B97D/6-311++G(d,p)
wB97XD/6-311++G(d,p)
M062x-D3/6–311++G(d,p)
Energy(E) (kJ/mol)
*ΔE (kJ/mol)
†ΔG (kJ/mol)
Energy(E) (kJ/mol)
*ΔE (kJ/mol)
†ΔG (kJ/mol)
Energy(E) (kJ/mol)
*ΔE (kJ/mol)
†ΔG (kJ/mol)
Energy(E) (kJ/mol)
*ΔE (kJ/mol)
†ΔG (kJ/mol)
A
−4377994.41
0.00
0.00
−4385432.16
0.00
0.00
−4386023.39
0.00
0.00
−4385753.28
0.00
0.00
B
−4377994.22
0.17
1.93
−4385432.01
0.15
1.81
−4386023.36
0.03
1.29
−4385753.11
0.17
0.62
C
−4377992.59
0.37
2.14
−4385431.57
0.59
3.57
−4386022.04
1.35
0.36
−4385751.73
1.38
2.56
Dimer of A
−8756060.73
–
–
−8770957.83
–
–
−8772148.01
–
–
−8771604.56
–
–
Optimized structure of chlorambucil.
Overlapping structure of one conformer over another (Conformer A, B and C are shown in Red, Yellow and Green, respectively.)
4.1.1 Dimer geometry
In the recorded IR and Raman spectra of chlorambucil, the peak corresponding to the free O—H stretching is not present and the carbonyl stretching peak is very strong and appears to be red shifted. Hence, one can come into the conclusion that the C⚌O and O—H bonds in the acid group may participate in making either strong intramolecular or weak intermolecular hydrogen bond. To account for intermolecular hydrogen bond, the theoretical anharmonic coupling model of chlorambucil dimer is proposed via carboxylic acid group. The proposed chlorambucil dimer consisting of two hydrogen bonds viz. O42⋯H38 and O4⋯H76 whose hydrogen bond distances are estimated to be 1.6788 and 1.6836 Å, respectively at wB97XD/6-311++G(d,p) level of theory while at M06-2X/6-311++G(d,p) level with GD3 correction, the same were estimated to be 1.6759 and 1.6828 Å, respectively. The optimized geometry of chlorambucil dimer is shown in Fig. S1 (Supplementary Material). The optimized geometrical parameters of respective bonds in monomer and dimer structures are equal expect the bonds where dimer is formed. Upon dimerization, the O3C19 bond is contracted by 0.0316, 0.0320 and 0.0393 Å as calculated by M06-2X (with GD3 correction), wB97XD and B97D (GD2 correction) methods with 6-311++G(d,p), respectively. In contrast, the O3H38 and O4C19 bond lengths are elongated by 0.0277 and 0.0185 Å at wB97XD/6-311++G(d,p) level of theory and the same trend is also identified at M06-2X/6-311++G(d,p) level of theory. The contraction and elongation of bonds clearly implies the sharing of charges from electron donor to acceptor species.
4.1.2 Two dimensional NMR spectrum of chlorambucil
1H–13C HSQC spectra provides correlation between the aliphatic and aromatic carbon in the chlorambucil and its attached protons. To make the precise assignments, the experimental contour plots were assigned on the basis of predicted chemical shifts of 13C and 1H NMR. In the spectrum shown in Fig. 3, proton chemical shifts are recorded along x-axis and carbon chemical shifts are recorded along y-axis in order to correlate the protein with their directly attached carbon nuclei. The 1H signals at 6.61 and 6.63 ppm correlates with 13C signal at 112 ppm which is ascribed to C14-H30 and C15-H31 bonds. Another two 1H signals at 7.06 and 7.08 ppm correlates with 13C signals at 130 and 129 ppm, respectively are assigned to C13-H29 and C12-H28 bonds. Multiple 1H signals near 3.7 and 3.6 ppm correlates with 13C signals at 54 and 41 ppm, respectively are attributed to the hydrogen atoms attached to carbon nuclei of —CH2CH2Cl group. The downfield 1H signals in the spectrum coincides with the downfield 13C signals are found to be the signals of C—H bonds of CH2 groups linked with —COOH group.2D-HSQC NMR spectrum of chlorambucil.
4.2 Intramolecular charge delocalization
In case of electronic structure calculations, the intramolecular charge delocalization plays a vital role in the chemical stability. The stabilization energies obtained in the NBO calculations clearly demonstrate the patterns of charge delocalization from various lone pair atomic orbitals (Lewis base) to antibonding orbitals (Lewis acid). Recent contributions from Saeed and his co-workers proposed the importance of “remote” donor-acceptor interactions for explaining the conformational properties of molecular systems via the resonance interactions between lone pair (LP) electrons and anti-bonding (BD∗) orbitals (Saeed et al., 2014; Saeed et al., 2015; Saeed et al., 2016). In chlorambucil conformers, the LP → BD∗, BD → BD∗ transitions exists and the strength of these interactions in LP → BD∗ is more significant than the BD → BD∗. It is seen from the Table 3 that the charge delocalization patterns of the conformers A and C are almost equal qualitatively and conformer B has different patterns of transitions. In both the conformers A and C, the major stabilization energies of about 52.68, and 131.29 kcal/mol, respectively drawn from the transition LP(1)O3 → BD∗(1)C18-H37. The charges transferred from O3 to the remote part (C18-H37) in this case may cause a geometrical rearrangement in the molecule. When rearrangement of molecule occurs, one can expect the shortening of X-H bond lengths (Coates, 2006). Since the C18 is attached to highly electronegative Cl2 and its natural charge is just half time that of the O3, the possibility of an improper hydrogen bond is limited. While in conformer B, the major stabilization energy of about 36.2 kcal/mol is obtained for LP(2)O3 → BD∗(2)O4-C19. Besides, the charge delocalization from the bonding to anti-bonding orbitals (BD → BD∗) shows a significant contribution to the stability of a molecule. The charge transfers from BD(1)O3-H38 to BD∗(1)C18-H37 stabilizes the chlorambucil conformers A and C by 39.41 and 86.31 kcal/mol, respectively. In conformer B, the delocalization charge within the aromatic ring stabilizes the molecule by an energy 17.06 kcal/mol.
Donor (i)
Acceptor (j)
Stabilization energy (kcal/mol)
Conformer A
Conformer B
Conformer C
BD → BD*
BD(1)O3-H38
BD*(1)C18-H37
39.41
–
86.31
BD(1)O3-C19
BD*(1)C18-H37
6.78
–
38.31
BD(2)C8-C14
BD*(2)C13-C15
17.08
–
17.58
BD(2)C8-C15
BD*(2)C12-C14
–
17.06
–
BD(2)C6-C12
BD*(2)C8-C14
15.96
–
16.28
BD(2)C12-C14
BD*(2)C6-C13
–
16.05
–
BD(2)C13-C15
BD*(2)C6-C12
16.01
–
16.02
BD(2)C6-C13
BD*(2)C8-C15
–
15.97
–
BD(2)C8-C15
BD*(2)C6-C13
–
14.16
–
BD(2)C8-C14
BD*(2)C6-C12
14.15
–
14.07
BD(2)C12-C14
BD*(2)C8-C15
–
12.89
–
BD(2)C13-C15
BD*(2)C8-C14
12.89
–
12.61
BD(2)C6-C12
BD*(2)C13-C15
12.41
–
12.49
BD(2)C6-C13
BD*(2)C12-C14
–
12.41
–
BD(1)O3-H38
BD*(1)C18-H36
3.85
–
12.24
BD(1)O3-H38
BD*(1)C17-H35
4.17
–
9.3
BD(1)C16-H33
BD*(2)O4-C19
5.14
3.4
2.73
BD(1)C16-H32
BD*(2)O4-C19
3.33
5.12
–
LP → BD*
LP(1)O3
BD*(1)C18-H37
52.68
–
131.29
LP(1)N5
BD*(2)C6-C12
35.53
–
42.71
LP(1)N5
BD*(1)C18-H37
8.03
–
41.22
LP(1)O4
BD*(1)C18-H37
37.04
–
26.89
LP(2)O3
BD*(2)O4-C19
35.82
36.2
36.8
LP(2)O4
BD*(1)O3-C19
29.88
30.59
34.03
LP(1)N5
BD*(2)C6-C13
–
34.1
–
LP(2)O4
BD*(1)C18-H37
4.33
–
24.19
LP(1)O3
BD*(1)C18-H36
5.54
–
20.71
LP(2)O4
BD*(1)C16-C19
15.96
15.69
15.03
LP(3)Cl1
BD*(1)C18-H37
–
–
15.88
LP(1)O3
BD*(1)C17-H35
5.99
–
15.83
LP(1)Cl2
BD*(1)C18-H37
–
–
12.26
LP(1)N5
BD*(1)C18-H36
0.71
–
10.51
LP(1)N5
BD*(1)C11-C18
6.61
6.64
7.29
LP(1)N5
BD*(1)C10-C17
6.75
6.48
7.15
LP(1)O3
BD*(1)O4-C19
6.38
6.13
6.56
The predicted stabilization energies of interacting intermolecular species from unit 1 to unit 2 and vice versa for the chlorambucil dimer of conformer A are given in Table S4 (Supplementary Material). The charge delocalization from the lone pair oxygen (O4 and O41) and chlorine (Cl2 and Cl39) atoms of both the units to the anti-bonding species in the remote parts stabilizes the dimer to a greater extent. The stabilization energy of about 21.45 kcal/mol predicted due to an interaction between the lone pair LP(2)O42 and BD∗(1)O3-H38 indicates the hydrogen bonding of O42⋯O3-H38 as shown in Fig. S1 (Supplementary Material).
4.3 Electron density topological and isodensity surface analysis
The theory of atoms in molecules enables us to determine the volume of electron density of atoms in a system and the changes in these volumes caused by a chemical reaction. Atoms in molecule (AIM) theory is a convenient method used to analyze the H-bonding and other interactions in a various molecular system in terms of critical points. The formation of a hydrogen bond is the result of an interaction between two closed-shell systems, and this is reflected in the properties of the charge distribution, particularly in its Laplacian, the quantity ∇2ρ(r) (Bader and Essen, 1984). The Laplacian of the electron density measures locally concentrated ∇2ρ < 0 and depleted ∇2ρ > 0 within a molecular system. QTAIM study the concept of chemical bond and bond strength in terms of topological features, electron density distribution function and energetic parameters. Geometrical as well as topological parameters are a useful tool to characterize the strength of hydrogen bond. The geometrical and topological parameters corresponding to the critical points which are connecting two non-bonded atoms in all three conformers of chlorambucil are given in Table 4. The geometrical parameters for the hydrogen bonds in all conformers are given in Table 5. On the basis of these parameters, there exist three non-conventional bonds in chlorambucil, two H H bonds which are different from dihydrogen bond (Reyes-Márquez et al., 2008; Hernandez-Trujillo and Matta, 2007) and the other O H bond such as H29 H24, H26 H28, H20 O4 in conformer A and H28 H24, H29 H26, H21 O4 in conformer B and H29 H26, H28 H24, H30 O4 in conformer C as shown in Fig. 4. The obtained results indicate that the strength of the hydrogen bonds are medium in nature as they satisfy the criteria ∇2ρBCP>0, HBCP < 0 proposed (Rozas et al., 2000). The relation between H-bond energy (E) and potential energy density (VBCP) at H O/N contact: E = 1/2(VBCP) is proposed (Espinosa et al., 1998). According to this equation, the interaction energy of H20 O4, H21 O4 and H30 O4 is calculated as −7.85, −7.98 and −5.6446 kJ/mol in conformer A, B and C which shows that hydrogen bonds (O H) are stronger in conformer A and B with comparison to conformer C. The bond ellipticity (ε) is a measure of nature and strength of the bond between the non-bonded atoms. The predicted values of ε in atomic units confirm that the strength of the O—H bonds such as H20⋯O4 (0.8691) in conformer A and H21⋯O4 (0.7389) in conformer B are stronger than the H30⋯O4 (0.0205) in conformer C. The ellipticity ε (0.7–0.8 a.u) at the (3,−1) line critical points (LCPs) of the non-bonded atoms shows the dynamical stability of a molecule (Brovarets et al., 2017). To validate the AIM topology results, the isosurface plot has been illustrated (see Fig. 5). The cube files for the functions λ2ρ (the product of second largest eigen value of Hessian matrix and electron density) and RDG were generated and used them to plot colored isosurface which directly indicates the strength of the non-bonded interactions present in the molecule. In all the three conformers, the space between the hydrogen atoms in the alkyl chain and aromatic hydrogens are filled with green colored isosurface which illustrates the weak H—H bonding between them. The slight red shaded green isosurfaces found between the Oxygen in the carboxylic acid group and the hydrogen atom in the alkyl chain (conformer A and B) or in the aromatic ring (conformer C) indicates the C—O⋯H hydrogen bonds.
Atoms involved in interactions
ρBCP (a.u.)
∇2ρBCP (a.u.)
GBCP (a.u.)
VBCP (a.u.)
HBCP (a.u.)
Eint (kJ/mol)
Ellipticity (ε) (a.u)
Conformer A
H29⋯H24
0.0142
0.0564
−0.0024
−0.0093
−0.0117
−12.1558
0.8265
H26⋯H28
0.0140
0.0558
−0.0024
−0.0091
−0.0115
−11.9985
0.8678
H20⋯O4
0.0092
0.0327
−0.0011
−0.0060
−0.0071
−7.8500
0.8691
Conformer B
H28⋯H24
0.0143
0.0560
−0.0024
−0.0092
−0.0116
−12.0508
0.7620
H29⋯H26
0.0143
0.0562
−0.0024
−0.0092
−0.0116
−12.1035
0.7674
H21⋯O4
0.0094
0.0329
−0.0011
−0.0061
−0.0072
−7.9814
0.7389
Conformer C
H29⋯H26
0.0143
0.0563
−0.0024
−0.0093
−0.0117
−12.1558
0.7918
H28⋯H24
0.0141
0.0560
−0.0024
−0.0092
−0.0116
−12.0771
0.8497
H30⋯O4
0.0073
0.0224
−0.0007
−0.0043
−0.0050
−5.6446
0.0205
Interactions (D-H…A)
Bond length (Å)
Angle D-H…A(°)
(rH + rA)(Å)
dD-H
dH-A
dD-A
Conformer A
C13-H29…H24
1.0825
2.0732
2.6099
107.4
2.4
C11-H26…H28
1.0923
2.0772
2.5612
103.4
2.4
C7-H20….O4
1.0923
2.5832
3.1937
114.4
2.7
Conformer B
C12-H28…H24
1.0825
2.0582
2.6186
109.0
2.4
C13-H29…H26
1.0825
2.0607
2.6140
108.5
2.4
C7-H21…O4
1.0923
2.5733
3.1883
114.7
2.7
Conformer C
C13-H29…H26
1.0824
2.0642
2.6164
108.4
2.4
C12-H28…H24
1.0824
2.0771
2.6144
107.5
2.4
C14-H30…O4
1.0856
2.6197
3.4834
136.1
2.7
Molecular graph illustrating the non-bonded interactions of chlorambucil.
Isosurface density plot illustrating the non-bonded interactions of chlorambucil.
4.4 Vibrational assignments
In this section, the detailed vibrational assignments of normal modes of chlorambucil conformers and the vibrational frequency shift due to the non-covalent interactions upon orientation changes have been explained. Moreover, the intermolecular bond formation enthalpy changes (also known as H-bond energy) between dimer and monomer structure of chlorambucil have been estimated using following Iogansen’s relationship (Iogansen, 1999).
Modes
Experimental wavenumbers (cm–1)
Scaled wavenumbers (cm–1)
Assignments with PED contribution†
IR
Raman
Monomer
Normalized Intensity
Dimer
1
–
–
15
0.0007
15
τ(CC)(36)+ω(NCC)(33)+R[τas](9)
2
–
–
24
0.0003
21,18
R[τ(C6-N5)](40)+τ(CN)(33)+τ(CC)(15)
3
–
–
29
0.0017
31,29
ω(NCC)(47)+τ(CC)(24)
4
–
–
34
0.0047
36
τ(CN)(34)+τ(CC)(33)+R[τ(C6-N5)](15)+β(NCC)(6)
5
–
–
39
0.0031
41,37
τ(CC)(73)+α(CCC)(9)
6
–
–
49
0.0042
55,53
τ(CC)(52)+τ(CN)(16)+ρ(NCC)(8)
7
–
–
66
0.0012
70
R[τ(C6-N5)](47)+τ(CC)(25)+τ(CN)(19)
8
–
–
80
0.0024
78,81
R[τas](24)+γ(CC)(12)+τ(CC)(20)+ω(NCC)(8)+α(CCC)(7)
9
–
–
121
0.0018
130
τ(CN)(36)+τ(CC)(18)+α(CCC)(9)+γ(CC)(9)
10
–
152 w
149
0.0058
146
τ(CC)(31)+τ(CN)(18)+β(CC)(17)+α(CCC)(6)
11
–
–
152
0.0059
154,152
τ(CN)(55)+τ(CC)(17)+β(NCC)(6)
12
–
–
177
0.0039
177,172
R[τas](25)+α(CCC)(18)+γ(CC)(7)+τ(CC)(6)+τ(CN)(5)
13
222 vw
217 vw
227
0.0055
229,227
ρ(CH2Cl)(55)+α(CCC)(8)+βs(CH2Cl)(8)+R[τ(C6-N5)](6)
14
240 vw
235 vw
235
0.0004
240
α(CCC)(54)+τ(C6-N5)(8)
15
258 w
251 vw
254
0.0196
245
ρ(CH2Cl)(37)+α(CCC)(20)+R[β(C6-N5)](12)
16
–
–
263
0.0287
263
τ(CN)(22)+ρ(CH2Cl)(16)+R[β(C6-N5)](12)+α(CCC)(9)+ρ(NCC)(7)+β(COOH)(5)
17
–
–
265
0.0051
265
τ(CC)(34)+α(CCC)(28)+β(COOH)(10)+γ(CC)(5)
18
–
315 w
322
0.0082
324
α(CCC)(37)+R[β(C6-N5)](10)+β(COOH)(8)+R[φ](5)
19
340 vw
344 vw
348
0.0180
359
α(CCC)(26)+ρ(NCC)(11)+R[τas](8)+β(CC)(6)+τ(CC)(6)+γ(CC)(5)
20
374 w
375 vw
374
0.0087
376
β(NCC)(20)+R[βas](17)+R[τ'as](8)+α(CCC)(7)+τ(CN)(5)
21
392 w
395 vw
402
0.0031
402
β(CC)(35)+α(CCC)(17)+τ(CC)(10)+R[β(C6-N5)](9)+R[β'αs](9)
22
–
–
432
0.0019
432,428
R[τ'as](76)+R[γ(CH)](14)
23
–
448 vw
444
0.0245
439,435
R[φ](14)+α(CCC)(11)+ρ(COOH)(9)+R[γ(C6-N5)](8)+β(NCC)(7)+τ(CC)(6)+R[τ'as](5)
24
496 w
–
498
0.0646
456,449
ρ(COOH)(30)+α(CCC)(18)+β(NCC)(9) +ν(CC)(9)+ρ(CH2)(6)
25
–
–
529
0.0302
522,515
ω(COOH)(19)+R[γ(C6-N5)](15)+R[τas](13)+τCOOH(10)+γ(CC)(7)+ρ(CH2)(7)
26
544 m
545 vw
533
0.1016
541,538
ω(COOH)(17)+R[βas](13)+β(NCC)(11)+τCOOH(11)+ρ(CH2)(10)+R[γ(C6-N5)](6)+τ(CC)(6)+R[τas](6)
27
561 m
–
564
0.0794
569,565
ρ(NCC)(24)+R[β(C6-N5)](13)+ν(C18-Cl2)(7)+α(CCC)(10)
28
–
–
574
0.0262
573,572
R[γ(C6-N5)](17)+R[τas](12)+γ(CC)(10)+α(CCC)(9)
29
640 w
640 w
636
0.0599
591,586
β(COOH)(29)+α(CCC)(16)+ρ(CH2)(11)+ρ(COOH)(8)+ν(CO)(5)
30
–
–
657
0.0019
660,658
R[β'as](78)
31
680 w
681 w
665
0.3056
666
τCOOH(60)+ω(COOH)(18)+ρ(CH2)(9)
32
725 m
725 m
708
0.0134
713,708
ν(CC)(23)+R[δ](13)+R[ν(C6-N5)](10)+R[φ](10)
33
747 m
748 m
746
0.0140
753,748
R[φ](59)+R[γ(C6-N5)](13)+γ(CC)(10)
34
767 sh
–
766
0.2237
767,764
ν(C17-Cl1)(34)+ν(C18-Cl2)(23)+ρ(CH2)(7)+ρ'(CH2Cl)(6)
35
–
772 m
781
0.0690
781,777
ρ(CH2)(35)+ρ'(CH2Cl)(30)+ν(C17-Cl1)(5)
36
–
–
785
0.2163
788,787
ν(C18-Cl2)(35)+ν(C17-Cl1)(22)+α(CCC)(16)+α(CCC)(7)+ρ(CH2Cl)(7)
37
794 sh
793 vw
790
0.0515
799,797
ρ(CH2)(25)+ρ'(CH2Cl)(13)+R[φ](8)+ν(CC)(5)+ν(C18-Cl2)(5)
38
806 m
807 m
801
0.0067
806,804
ρ'(CH2Cl)(22)+ρ(CH2)(16)+ν(C17-Cl1)(5)
39
825 m
–
823
0.0407
824,822
R[ν(CC)](17)+R[βαs](14)+R[γ(CH)](17)+ν(CC)(12)+R[γ(C6-N5)](5)+R[ν(C6-N5)](5)
40
–
829 sh
835
0.0001
834,829
R[γ(CH)](97)
41
–
840 vw
846
0.1395
852,847
R[γ(CH)](54)+R[γ(C6-N5)](9)+R[τas](8)+ν(CC)(6)+γ(CC)(5)
42
862 vw
–
871
0.0481
879,876
ρ(CH2)(30)+ν(CC)(21)+ω(COOH)(11)+α(CCC)(6)+τ(CH2)(5)
43
–
–
910
0.0045
921,918
ν(CC)(52)+ρ(CH2)(15)
44
964 w
964 sh
964
0.0064
966,959
R[γ(CH)](74)+R[φ](17)
45
981 sh
982 w
981
0.0023
982,980
R[γ(CH)](88)+R[τ'as](8)
46
–
1001 sh
997
0.0735
1002,995
ν(CN)(55)+ρ(CH2)(9)+R[δ](6)
47
1014 vw
1012 w
1011
0.0123
1016,1013
ρ(CH2)(50)+R[δ](11)+ω(CH2)(5)+ν(CC)(5)
48
–
–
1018
0.0146
1025,1019
ρ(CH2)(27)+R[δ](18)+R[ν(C8-C15)](11)+α(CCC)(8)
49
–
1027 w
1029
0.0095
1033,1031
R[ν(CC)](22)+R[δ](19)+ρ(CH2)(17)+α(CCC)(6)+ν(CC)(6)
50
–
1037 w
1033
0.0156
1035,1034
ρ(CH2)(38)+βas(CH2Cl)(36)+ν(CN)(7)
51
1062 vw
1060 w
1046
0.0065
1039
ν(CC)(81)+α(CCC)(6)
52
1062 vw
1060 w
1046
0.0041
1055,1053
ν(CC)(82)+α(CCC)(6)
53
–
–
1084
0.0006
1084
βas(CH2Cl)(41)+ρ(CH2)(19)+ν(CN)(7)+R[ν(CC)](6)+β'as(CH2Cl)(5)+β(NCC)(5)
54
–
–
1095
0.0541
1099,1097
ν(CC)(60)+ρ(CH2)(7)+ω(CH2)(5)+ν(CO)(5)+α(CCC)(5)
55
1139 m
1139 vw
1143
0.3292
1147
τ(CH2)(24)+ν(CO)(20)+ω(CH2)(11)+β(COH)(10)+R[ν(CC)](7)
56
1152 m
1153 vw
1160
0.2158
1161
ν(CO)(11)+R[β(CH)](29)+R[ν(CC)](17)+β(COH)(6)+ν(CN)(6)
57
1178 m
1178 m
1172
0.0754
1176,1171
τ(CH2)(24)+βas(CH2Cl)(19)+ρ'(CH2Cl)(18)+ν(CN)(13)
58
1190 sh
1190 w
1199
0.2103
1187,1181
τ(CH2)(38)+ν(CO)(8)+ω(CH2)(6)+β(COH)(6)+R[ν(CC)](5)
59
–
1202 w
1202
0.1680
1207,1201
τ(CH2)(30)+ρ'(CH2Cl)(10)+R[ν(C6-N5)](9)
60
1213 sh
–
1209
0.3739
1211,1209
τ(CH2)(35)+β(COH)(9)+ν(C—O)(9)
61
–
1220 vw
1221
0.0077
1224,1216
R[β(CH)](59)+R[ν(CC)](8)
62
1234 m
–
1241
0.0018
1249,1239
ν(CC)(26)+R[ν(CC)](22)+R[β(CH)](20)+ω(CH2)(13)+R[δ](9)
63
1250 w
1251 w
1253
0.1386
1255,1251
ν(CN)(35)+R[ν(CC)](19)+ρ'(CH2Cl)(11)+ρ(CH2)(8)+R[β(CH)](5)
64
1275 m
1277 vw
1282
0.1006
1279,1278
βs(CH2Cl)(46)+ω(CH2)(21)+ρ(CH2Cl)(11)+β'αs(CH2Cl)(10)
65
–
–
1288
0.0633
1289,1284
βs(CH2Cl)(49)+βαs(CH2Cl)(16)+ω(CH2)(14)+ρ(CH2Cl)(12)
66
–
–
1292
0.0145
1296,1290
τ(CH2)(53)+ω(CH2)(12)
67
–
–
1301
0.0387
1301,1296
β(COH)(28)+ω(CH2)(19)+τ(CH2)(12)+ν(CO)(6)
68
–
–
1303
0.1405
1306,1303
τ(CH2)(34)+ρ'(CH2Cl)(22)+R[ν(C6-N5)](8)+ρ(CH2)(6)
69
1308 m
1306 w
1306
0.0361
1314,1309
R[ν(CC)](25)+τ(CH2)(21)+ρ'(CH2Cl)(12)+β(COH)(7)
70
1333 w
1334 w
1332
0.0031
1340,1331
R[ν(CC)](38)+R[β(CH)](17)+τ(CH2)(6)+ω(CH2)(9)+τ(CH2)(8)
71
1355 sh
1357 w
1347
0.0339
1350,1345
R[β(CH)](29)+τ(CH2)(10)+ω(CH2)(9)+R[ν(CC)](7)+ρ(NCC)(6)+R[β(C6-N5)](5)
72
1365 m
1365 w
1365
0.0690
1366,1359
ω(CH2)(45)+τ(CH2)(13)+ν(CC)(12)+R[β(CH)](6)
73
–
–
1373
0.0507
1379,1374
ω(CH2)(62)+τ(CH2)(11)+ν(CC)(11)
74
1384 w
1384 w
1383
0.0304
1386,1382
ω(CH2)(52)+ν(CN)(10)+ν(CC)(5)
75
–
–
1393
0.1070
1391,1389
ω(CH2)(64)+ν(CN)(8)+ν(CC)(5)+βas(CH2Cl)(5)
76
1408 w
1408 w
1414
0.2085
1416
ω(CH2)(37)+ν(CC)(25)+ν(C19-O3)(10)+β(COOH)(7)+α(CH2)(6)+τ(CH2)(6)
77
1429 sh
–
1418
0.3282
1422
R[ν(C6-N5)](26)+τ(CH2)(25)+β(NCC)(6)+α(CH2)(5)+ν(CN)(8)
78
1446 m
1448 m
1443
0.0425
1437,1436
α(CH2)(85)
79
1456 m
–
1456
0.0409
1457,1452
R[ν(CC)(30)+R[β(CH)](25)+α(CH2)(12)+R[β(C6-N5)](8)
80
–
–
1462
0.0327
1469,1463
α(CH2)(91)
81
–
–
1468
0.0537
1472,1470
βas(CH2Cl)(59)+βs(CH2Cl)(22)
82
1477 sh
–
1477
0.0440
1479,1474
α(CH2)(42)+β'as(CH2Cl)(20)+βas(CH2Cl)(11)+βs(CH2Cl)(10)
83
–
1478 w
1482
0.0591
1482
β'as(CH2Cl)(31)+α(CH2)(31)+βs(CH2Cl)(14)+βas(CH2Cl)(11)
84
–
1493 vw
1496
0.0294
1486
α(CH2)(92)
85
1517 s
1518 w
1503
0.0211
1505,1504
α(CH2)(88)
86
–
1552 vw
1547
1.0000
1553,1544
R[β(CH)](41)+R[ν(CC)](31)+R[ν(C6-N5)](15)+ν(CC)(5)
87
1613 s
1613 s
1609
0.0270
1613,1610
R[ν(CC)](69)+R[β'as](9)+R[β(C6-N5)](6)
88
–
–
1664
0.3926
1666,1662
R[ν(CC)(65)+R[β(CH)](20)+R[βas](11)
89
1697 s
–
1811
0.9842
1760,1715
ν(C19 = O4)(79)+ρ(COOH)(6)+ν(C16-C19)(6)
90
2932 s
2932 vs
2932
0.0574
2931,2928
νs(C9-H23)(52)+νs(C9-H22)(25)+ν(C16-H33)(12)+ν(C7-H21)(8)
91
–
–
2940
0.1044
2941
νs(C16-H33)(38)+νs(C7-H21)(35)+νs(C16-H32)(16)+νs(C7-H20)(8)
92
–
–
2944
0.0669
2945,2943
νs(C10-H25)(31)+νs(C11-H27)(28)+νs(C10-H24)(19)+νs(C11-H26)(15)
93
–
–
2944
0.0925
2946
νs(C7-H21)(35)+νs(C16-H33)(20)+νs(C9-H23)(13)+νs(C7-H20)(11) +νs(C16-H32)(10)+νs(C9-H22)(5)
94
2953 sh
–
2952
0.0595
2956,2951
νs(C11-H27)(29)+νs(C10-H25)(24)+νs(C11-H26)(23)+νs(C10-H24)(22)
95
–
2961 s
2973
0.0349
2972
νas(C9-H22)(43)+νas(C9-H23)(23)+νas(C16-H32)(16)+νas(C16-H33)(11)
96
–
–
2983
0.0455
2986,2981
νs(C18-H37)(58)+νs(C18-H36)(31)+ν(C11-H27)(6)
97
–
–
2984
0.0407
2984,2974
νs(C17-H35)(56)+νs(C17-H34)(29)+νas(C10-H25)(8)+νas(C10-H24)(4)
98
–
–
2984
0.0651
2984,2980
νas(C16-H32)(54)+ν(C9-H22)(19)+νas(C16-H33)(18)
99
–
–
2989
0.0645
2990
νas(C10-H24)(35)+νas(C11-H26)(19)+νas(C10-H25)(19)+ν(C17-H34)(9)+νas(C11-H27)(8)
100
–
–
2992
0.0237
2996,2994
νas(C11-H26)(37)+νas(C11-H27)(25)+νas(C10-H24)(16)+νas(C10-H25)(14)+ν(C18-H36)(5)
101
3002 vw
3003 sh
2996
0.0553
3009,3001
νas(C7-H20)(73)+νas(C7-H21)(15)+νas(C9-H23)(6)+νas(C9-H22)(4)
102
3033vw
–
3036
0.0531
3046,3044
R[νas(C15-H31)](93)+R[νas(C13-H29)](5)
103
–
–
3039
0.0544
3050, 3033
R[νas(C14-H30)(93)+R[νas(C12-H28)](5)
104
–
3043 s
3043
0.0109
3042,3039
νas(C18-H36)(55)+νas (C18-H37)(37)
105
–
–
3044
0.0254
3044
νas(C17-H34)(55)+νas(C17-H35)(37)
106
–
–
3071
0.0427
3088,3083
R[νs(C13-H29)](92)+R[νs(C15-H31)](5)
107
3079 w
3080 s
3077
0.0353
3112,3096
R[νs(C12-H28)](93)+R[νs(C14-H30)(5)
108
–
–
3635
0.2699
3194, 3119
ν(O3-H38)(1 0 0)
Comparison of Experimental and simulated (wB97XD/6-311++G(d,p) level of theory) IR spectrum of chlorambucil.
Comparison of Experimental and simulated (wB97XD/6-311++G(d,p) level of theory) Raman spectrum of chlorambucil.
4.4.1 Benzene ring vibrations
The C—H stretching modes of the disubstituted benzene derivatives are expected to appear in the region 3010–3120 cm−1 (Coates, 2006). In the present study, the peaks at 3095, 3080 cm−1 in FT-Raman and 3079 cm−1 in FT-IR spectra are assigned to symmetric C—H stretching of the benzene ring, respectively. The observed and the calculated dimer wavenumbers of conformer A confirms that the aromatic hydrogen atoms H28 and H29 are involved in making non-bonded interactions with the alkyl chain. The spectroscopic result validates the predicted non-bonded interactions of H29⋯H24, H26⋯H28 by AIM analysis. The weak intensity peak at 3033 cm−1 in FT-IR is attributed to C—H asymmetric stretching. The scaled theoretical values of conformer A obtained at wB97XD/76-311++G(d,p) method are found to be in good agreement with the experimental observations. The peaks corresponding to the ring C—C stretching vibrations of the disubstituted benzene appear as a medium to strong intensity in the fingerprint region. In the present study, the peaks identified at 1613, 1456, 1333, 1308 cm−1 in FT-IR and the peaks at 1613, 1334, 1306 and 1027 cm−1 in FT-Raman are ascribed to ring C—C stretching modes. The C—C stretching of ring carbon (C8) and a carbon atom in the alkyl group (C7) is identified at 1234 cm−1. The in-plane deformation vibrations of the ring C—H bonds are strongly coupled with the ring C—C stretching and they present at 1335 and 1152 cm−1 in FT-IR and at 1552, 1357, 1220 and 1153 cm−1 in FT-Raman. The weak intensity peaks found at 981, 964 cm−1 in FT-IR and the peaks at 982, 964, 840 and 829 cm−1 in FT-Raman are attributed to C—H out-of-plane bending deformation modes. Since the ring deformation modes such as puckering (R[ϕ]), asymmetric in-plane (R[βas], R[β’as]) and torsion (R[τas], R[τ’as]) are in strong association with the other modes, these modes are not appeared in both the IR and Raman spectra. Thus, the scaled values of conformer A at 746 cm−1 (R[ϕ]), 657 cm−1 (R[β’as]), 432 cm−1 (R[τ’as]) and 177 cm−1 (R[τas]) are assigned as given in Table 6. The mixed mode of trigonal deformation (R[δ]) of the ring is found to be red shifted (theoretical) by ∼5 cm−1 with respect to the experimental and dimer wavenumbers clearly indicates the contribution of ring hydrogen atoms (H28 and H29) in the formation of H⋯H bonding with the hydrogen atoms (H26 and H24) in the alkyl chain, respectively.
4.4.2 Chloromethyl group vibrations
The highly electronegative chlorine atoms (Cl1, Cl2) that attached to the CH2 groups generate vibrations identical to the CH3 groups. Due to the identical environment nearer to both CH2Cl groups, the degeneracy of modes likely to be possible. As expected the difference between the vibrational wavenumbers of the C—H asymmetric stretching modes is not significant, in fact, they are nearly equal. The degeneracy also has been found in the case of symmetric stretching of C—H. In the present study, the weak intensity peak observed at 3043 cm−1 in the FT-Raman spectrum is assigned to asymmetric C—H stretching of the CH2Cl group. The PED results indicate that these modes are pure. The symmetric stretching of C—H bonds is not identified in both the spectra. Hence, the theoretically predicted values of 2984 and 2983 cm−1 at wB97XD/6-311++G(d,p) are assigned to this mode. The stretching modes of C—Cl bond in the alkyl chain usually appear in the region 700–800 cm−1 (Coates, 2006). The scaled value of 785 cm−1 and the 767 cm−1 are assigned to the stretching modes C18-Cl2 and C17-Cl1, respectively. The asymmetric and symmetric deformation modes of this group are found in association with the deformation vibrations of the CH2 groups in the region 1480–1080 cm−1. In the experimental FT-Raman, the weak intense peak identified at 1478 cm−1 is endorsed as CH2Cl asymmetric deformation (β’as(CH2Cl)) and the peaks observed at 1275 cm−1 (FT-IR) and 1277 (FT-Raman) are assigned to symmetric deformation modes of CH2Cl (βs(CH2Cl). The scaled values assigned to these modes at 1482, 1468, 1288, 1282 and 1084 cm−1 are found in good agreement with the experimental values.
4.4.3 Methylene group vibrations
The five methylene groups in the title molecule generate peaks at 3002 cm−1 in FT-IR and 3003, 2961 cm−1 in FT-Raman which corresponds to C—H asymmetric stretching mode. The C—H symmetric stretching of the methylene group appears at 2953, 2932 cm−1 in FT-IR and 2932 cm−1 in FT-Raman. The scissoring deformation modes of the methylene group are identified at 1517, 1477 and 1446 cm−1 in FT-IR and 1518, 1493 and 1448 cm−1 in FT-Raman. The red shift of ∼15 cm−1 found in conformer A with respect to the experimental (both IR and Raman) also confirms the H⋯H bonding between H24 and H29. Since the PED contributions of scissoring modes are predominant, they are very pure. In the wagging modes of the methylene, a considerable amount of vibrational contributions of CC and CN stretching modes are involved and they produce peaks 1408, 1384 and 1365 cm−1 in the experimental (IR and Raman) spectra. The calculated values of this mode are in good agreement with the experimental values. The weak intensity peaks at 1213 cm−1 (FT-IR), 1190 cm−1 (FT-IR, FT-Raman) and medium strong intensity peak at 1202 cm−1 are recognized as twisting modes.
4.4.4 CN vibrations
The modes associated with a group that contains a nitrogen atom which is attached to the ring and the chlorinated alkyl group are recognized as mixed modes in the fingerprint region. At 1429 cm−1 in the FT-IR spectrum, the stretching of a C—N bond which is in connection with the ring has been identified. In this mode, the equal vibrational contribution of τCH2 also has been noted. The red shift of 11 cm−1 is noticed with respect to the experimental at this mode, which may due to the participation of hydrogen atoms (H24, H26) in H⋯H bonding. The stretching peaks associated with the bonds between a nitrogen atom and the chlorinated alkyl group are identified at 1251 and 1001 cm−1 in FT-Raman. The out-of-plane bending mode of the bond C6-N5 is not identified in the whole region of the spectra, hence theoretically value of 574 cm−1 at wB97XD/6-311++G(d,p) method is assigned to R[γ(C6-N5)]. Since the vibrational contribution of the in-plane bending of the C6-N5 bond is highly coupled with the deformation modes of N5C10C11 such as ρ(NCC) and τCN, the assignment of the peak is quite uncertain. Hence the peak observed at 561 cm−1 in FT-IR or the calculated value of 263 cm−1 may recognize as R[β(C6-N5)].
4.4.5 CC vibrations
The CC bonds in the alkyl chains on both sides of the benzene ring produce peaks in the fingerprint region at 1234 and 1062 cm−1 in FT-IR and 1060 cm−1 in FT-Raman which belongs to CC stretching. In the assignment made, we have noticed a red shift of ∼15 cm−1 for the degenerate modes (modes 51 and 52 in Table 6). This may due to the formation of H⋯H bond (H24⋯H29 and H26⋯H28) in conformer A. The scaled dimer wavenumbers at 1055 and 1053 cm−1 match with the experimental which also validates the formation of H⋯H. The in-plane bending mode of the bond C8–C7 appear as weak intensity peak at 392 cm−1 in FT-IR and 395 cm−1 in FT-Raman. As the torsion modes (τCC) of CC bonds between the CH2 groups are not found in both the spectra, the theoretically predicted values at 265, 149, 49, 39 and 15 cm−1 is assigned to τCC.
4.4.6 Carboxylic acid group vibrations
The chances of homo-dimer formation are more probable in chlorambucil because of the carboxylic acid group. In the monomer conformers, the influences of the neighbouring molecule have not been accounted. In order to make the precise discussion, the dimer of most stable conformer A has been considered. Since a peak corresponding to the O—H stretching is not observed in both the IR and Raman, the stretching of free OH and C ⚌ O calculated at 3635 and 1811 cm−1 calculated for the monomer structure are assigned. The computed wavenumbers of dimer at 3194 and 3119 cm−1 are belongs to hydrogen bonded O—H stretching. The weak intensity peak at 3360 cm−1 (FT-IR) and 3375 cm−1 (FT-Raman) assigned by Gunasekaran et al. (Gunasekaran et al., 2008a, 2008b) also confirms the hydrogen bonded dimer in a condensed phase of chlorambucil. According to Iogansen’s relationship, the estimated H-bond energy of dimer of chlorambucil is about 27.6 kJ/mol. Hence, these interactions have the strongest impact on the stabilization of the molecular confirmation. The observed peak at 1697 cm−1 in FT-IR and the calculated dimer wavenumber at 1715 cm−1 reflects that C ⚌ O is involved in the formation of dimer via. C ⚌ O⋯H as shown in Fig. S1 (Supplementary Material). The vibrational mode of C—O stretching is strongly coupled deformation modes of the CH2 groups. Hence, the mixed mode present at 1139 cm−1 is endorsed as C—O stretching although its PED contribution is less than the τCH2. The deformation modes of —COOH group are found active in the finger print region at 680, 640, 544 and 496 cm−1 in FT-IR and 681, 640 and 545 cm−1. In the assignments made, the theoretical wavenumbers corresponding to the torsion and wagging modes (modes 30 and 25 in Table 5) are red shifted which demonstrates the intramolecular hydrogen bond between the electron donor in the carboxylic acid group and the neighbouring CH2 groups as confirmed by AIM analysis.
4.5 Interactive sites of chlorambucil
The feasible interactive sites which are appropriate for the electrophilic and nucleophilic attack have been determined using Fukui function analysis. In order to find the site selectivity precisely, the values of relative electrophilicity (sk+/sk-) and nucleophilicity (sk−/sk+) that derived from the hirshfeld charges corresponding to anion (N+1), cation (N−1) and neutral radical species have been computed by wB97XD and B97D methods with 6-311++G(d,p) as proposed (Roy et al., 1998). The atomic center or site having the highest sk+/sk− is the most feasible site to be attacked by a nucleophile, and having the highest sk−/sk+ ratio is the most feasible site to be attacked by an electrophile (Karthick and Tandon, 2016). The results obtained from both the functional (wB97XD and B97D) show that the atomic centers N5 and C7 which are attached to the ring carbon atoms are the most suitable sites to be attacked by an electrophile in all the conformers (Refer the Table S10 (Supplementary Material)). In the calculations obtained for the relative electrophilicity of all the conformers using the functional wB97XD are quantitatively not comparable as we acquire the negative value of fk− in conformer A and B. Fortunately, the positive Fukui functions (fk+, fk−) obtained using the functional B97D show that the carbon atom C19 is the most prominent to be attacked by nucleophile as it has higher value of sk+/sk− 11.7019 e, while in conformer B and C, the atoms O3 and C19 are found to be more probable sites to be attacked by nucleophile. In many chemical systems, the interactions between the electrophiles and nucleophiles play an important role in the chemical stability. In charge transfer process, the electrophiles act as a Lewis acid and the nucleophile act as a Lewis base (Cedillo and Contreras, 2012). In order to find the exact picture of charge distribution, the normalized Parr functions P− and P+ (Domingo et al., 2013) based on respective hirshfeld atomic spin densities (ASDs) of cation ρsrc (k) and anion ρsra (k) radicals also have been considered (depicted in Table S11 (Supplementary Material)). In results of both the functional, the N5 and C8 are found to be the most electrophilic center. This validates the results of relative nucleophilicity as suggested by Roy et al. (Roy et al., 1998). From the values of Parr function of cation radical (P−) obtained by wB97XD/6-311++G(d,p), the atoms O3, C14 and C9 are identified as the most nucleophilic centers of conformer A, B and C respectively. The results obtained in this case are in good correlation with the Fukui functions (sk+/sk−) of the conformers except in conformer B where the carbon atom C14 is identified as the most nucleophilic center rather than C16.
4.6 Automated docking analysis for binding sites
Since the chlorambucil is already a well-known drug, the probable active binding sites on the unexamined biological targets were found using the automated docking tool. The selection of targets has been done by cross-validation probabilities based on shaping the interaction landscape of molecules using Swiss target prediction (Gfeller et al., 2013; Gfeller et al., 2014). The chlorambucil is found to be more active against the targets in the family of Aldo-keto reductase family 1 proteins (AKR1B1, AKR1B10) and FAD-linked sulfhydryl oxidase ALR (GFER). The docking results obtained from Autodock 4.2 are depicted in Table S12 (Supplementary Material) and are summarized as follows.
4.6.1 Docked state parameters of chlorambucil and AKR1B1
The target AKR1B1 is found to interact with carboxylic acid derivatives which involve in oxidative stress diseases, cell signal transduction and cell proliferation process (Maccari and Ottanà, 2015). In the previous literature, β-glucogallin inhibited AKR1B1 with an IC50 value of 58 ± 3 μM (Puppala et al., 2012). The estimated free binding energy of about −5.8 kcal/mol and inhibition constant of 55.72 μM in this study shows that chlorambucil is a potential inhibitor of AKRB1 (PDB ID: 5H7A). In the docked state as shown in Fig. 8a, it is found that the conventional hydrogen bond (O⋯H—N) of 2.0122 Å present between the Oxygen O4 atom in the ligand and —N—H species in the target residue LEU 618 in chain B. Also, the strong interactions are also identified between the —OH in the carboxylic acid of the ligand and oxygen donors in the target residues SER580, THR582 with bond distances of 1.9725 and 1.6225 Å, respectively.Docked state conformers of chlorambucil with (a) 5H7A, (b) 1ZUA (c) 4LDK.
4.6.2 Docked state parameters of chlorambucil and AKR1B10
The target AKR1B10 (PDB id: 1ZUA) is involved in the production of cancer of the liver through modulation of fatty acid and lipid synthesis (Jin et al., 2016). The inhibition of AKR1B10 results in apoptosis of tumor cells whose phospholipids were decreased by half suggesting the involvement of phospholipids in AKR1B10’s oncogenic function. Generally, the potential inhibitors of AKRB1 reported earlier were not found to have a negligible inhibition towards AKR1B10 (Zhang et al., 2013). In our present study, it is also confirmed that AKR1B10 has a negligible inhibition constant (4.36 mM) and the target residue LYS92 binds with the chlorambucil through the hydrogen bond (1.7502 Å) as shown in Fig. 8b.
4.6.3 Docked state parameters of chlorambucil and GFER
The estimated free binding energy of about −6.5 kcal/mol and inhibition constant of 17.18 μM in this study shows that chlorambucil is a potential inhibitor of GFER (PDB id: 4ldk). In the docked state as shown in Fig. 8c, we have identified two conventional hydrogen bonds (O⋯H—N) of 2.0626, 2.3534 Å present between the Oxygen O4 atom in the ligand and —N—H species in the target residue ARG104 and the —OH species in the carboxylic acid of the ligand and Oxygen donors in the target residue LYS 183 with bond distance of 1.7199 Å.
5 Conclusion
A careful attempt has been made to reveal the non-covalent interactions between the atoms of chlorambucil. The shifting of harmonic frequencies due to those interactions and charge delocalization has been identified using monomer and dimer structures and were explained in detail with the electron density topology and isosurface plot of monomer. The correlation between the carbon and hydrogen nuclei of chlorambucil has been obtained by 2D-HSQC NMR spectrum. In the AIM topology, we demonstrated the presence of weak hydrogen bonds and H—H in chlorambucil. From the NBO analysis, the transition from the lone pair to anti-bonding species located far from the lone pair has been identified due to the rearrangement of molecular geometry. However, the highly electronegative atoms in the near feature of remote part (antibonding species) have limited the formation of improper hydrogen bonds which prevents the molecule to attain lower stability. The active sites predicted from the Fukui and Parr functions were found to be involved in making the strong hydrogen bonding with the residues of target biomolecules in the family of Aldo-keto reductase.
Acknowledgements
The author T. Karthick, sincerely acknowledged University Grant Commission (UGC), New Delhi for providing financial assistance under D.S.Kothari fellowship scheme (Award letter No: F. 4-2/2006(BSR)/PH/14-15/0010 and Swapnil Singh and Karnica Srivastava also acknowledged the financial assistance provided under UGC-BSR Fellowship. Also, the instrument facility provided by SAIF, IIT Madras is greatly acknowledged.
References
- Spectroscopic studies, HOMO–LUMO and NBO calculations on monomer and dimer conformer of 5-nitrosalicylic acid. J. Mol. Struct.. 2011;1005:192-201.
- [Google Scholar]
- Vibrational spectroscopic studies, molecular orbital calculations and chemical reactivity of 6-nitro-m-toluic acid. Spectrochim. Acta. 2012;92A:137-147.
- [Google Scholar]
- AIM 2000 - program to analyze and visualize atoms in molecules. J. Comput. Chem.. 2001;22:545-559.
- [Google Scholar]
- IR Vibrational spectra of H-bonded complexes of adenine, 2-aminopurine and 2-aminopurine+ with cytosine and thymine: quantum-chemical study. Optics & Spectr.. 2011;111:750-757.
- [Google Scholar]
- DPT tautomerisation of the G·Asyn and A*·G*syn DNA mismatches: a QM/QTAIM combined atomistic investigation. Phys. Chem. Chem. Phys.. 2014;16:9074-9085.
- [Google Scholar]
- A QM/QTAIM research under the magnifying glass of the DPT tautomerisation of the wobble mispairs involving 2-aminopurine. New J. Chem.. 2017;41:7232-7243.
- [Google Scholar]
- A local extension of the electrophilicity index concept. J. Mex. Chem. Soc.. 2012;56(3):257-260.
- [Google Scholar]
- Spectroscopic and molecular docking studies on chlorambucil interaction with DNA. Int. J. Biol. Macromolec.. 2012;51:406-411.
- [Google Scholar]
- Interpretation of Infrared Spectra. Encyclopedia of Analytical Chemistry: A Practical Approach; 2006.
- Understanding the local reactivity in polar organic reactions through electrophilic and nucleophilic Parr functions. RSC Adv.. 2013;3:1486-1494.
- [Google Scholar]
- Hydrogen bond strengths revealed by topological analyses of experimentally observed electron densities. Chem. Phys. Lett.. 1998;285:170-173.
- [Google Scholar]
- Frisch, M.J., Trucks, G.W., Schlegel, H.B., Scuseria, G.E., Robb, M.A., Cheeseman, J.R., et al., 2013. Gaussian 09, Revision D.01, Gaussian, Inc., Wallingford CT.
- Shaping the interaction landscape of bioactive molecules. Bioinformatics. 2013;29(23):3073-3079.
- [Google Scholar]
- SwissTargetPrediction: a web server for target prediction of bioactive small molecules. Nucl. Acids Res.. 2014;42(W1):W32-W38.
- [Google Scholar]
- Corner flapping: a simple and fast algorithm for exhaustive generation of ring conformations. J Am Chem Soc.. 1989;111:8950-8951.
- [Google Scholar]
- An efficient algorithm for searching low-energy conformers of cyclic and acyclic molecules. J. Chem. Soc., (Perkin Trans. 2). 1993;2:187-198.
- [Google Scholar]
- CONFLEX 7. Tokyo, Japan: CONFLEX Corporation; 2012.
- Vibrational assignments and electronic structure calculations for chlorambucil. Asian J. Chem.. 2008;20:6149-6162.
- [Google Scholar]
- Vibrational spectra and normal coordinate analysis on structure of chlorambucil and thioguanine. Pramana – J. Phys.. 2008;41:1291-1300.
- [Google Scholar]
- Direct proportionality of the hydrogen bonding energy and the intensification of the stretching ν(XH) vibration in infrared spectra. Spectrochim. Acta, Part A. 1999;55:1585-1612.
- [Google Scholar]
- Aldo-keto reductase family 1 member B 10 mediates liver cancer cell proliferation through sphingosine-1-phosphate. Sci. Rep.. 2016;6(22746):1-11.
- [Google Scholar]
- Spectroscopic investigations, molecular interactions, and molecular docking studies on the potential inhibitor “thiophene-2-carboxylicacid”. Spectrochim. Acta. 2015;141A:104-112.
- [Google Scholar]
- Computational approaches to find the active binding sites of biological targets against busulfan. J. Mol. Model.. 2016;22(142):1-9.
- [Google Scholar]
- Evaluation of structural isomers, molecular interactions, reactivity descriptors, and vibrational analysis of tretinoin. Anal. Sci.. 2017;23:83-87.
- [Google Scholar]
- Characterization and intramolecular bonding patterns of busulfan: Experimental and quantum chemical approach. Spectrochim. Acta. 2017;173A:390-399.
- [Google Scholar]
- Keith, T.A., 2013. AIMALL (Version 13.05.06) TK Gristmill Software, Overland Park KS, USA.
- Holland-Frei Cancer Medicine (6th edition). Hamilton, Canada: BC Decker; 2003.
- Multiwfn: a multifunctional wavefunction analyzer. J. Comput. Chem.. 2012;33:580-592.
- [Google Scholar]
- Targeting aldose reductase for the treatment of diabetes complications and inflammatory diseases: new insights and future directions. J. Med. Chem.. 2015;58(5):2047-2067.
- [Google Scholar]
- GAR2PED: A Program to Obtain a Potential Energy Distribution from a Gaussian Archive Record. Antwerp, Belgium: University of Antwerp; 1995.
- The isolation and characterization of β-glucogallin as a novel aldose reductase inhibitor from Emblica officinalis. PLoS ONE. 2012;7(4):e31399. 1–9
- [Google Scholar]
- Deciphering the interactions between chlorambucil and calf thymus DNA: a multi-spectroscopic and molecular docking study. Arch. Biochem. Biophys.. 2015;566:7-14.
- [Google Scholar]
- Unconventional hydrogen and dihydrogen bonded supramolecular array of a 2,6-dioxa-9,16-diaza-1,3(1,2),4(1,4)-tribenzenacycloheptadecaphane-borane adduct. ARKIVOC. 2008;2008(v):115-123.
- [Google Scholar]
- Local softness and hardness based reactivity descriptors for predicting intra- and intermolecular reactivity sequences: carbonyl compounds. J. Phys. Chem. A.. 1998;102:3746-3755.
- [Google Scholar]
- Behavior of ylides containing N, O, and C atoms as hydrogen bond acceptors. J. Am. Chem. Soc.. 2000;122:11154-11161.
- [Google Scholar]
- Synthesis, X-ray crystal structure, thermal behavior and spectroscopic analysis of 1-(1-naphthoyl)-3-(halo-phenyl)-thioureas complemented with quantum chemical calculations. Spectrochim. Acta. 2015;150A:409-418.
- [Google Scholar]
- Competing intramolecular N-H O–C hydrogen bonds and extended intermolecular network in 1-(4-chlorobenzoyl)-3-(2-methyl-4-oxopentan-2-yl) thiourea analyzed by experimental and theoretical methods. Chem. Phys.. 2014;31–432:39-46.
- [Google Scholar]
- On the roles of close shell interactions in the structure of acyl-substituted hydrazones: an experimental and theoretical approach. Spectrochimica Acta. 2016;157A:138-145.
- [Google Scholar]
- Vibrational analysis and chemical activity of paracetamol–oxalic acid cocrystal based on monomer and dimer calculations: DFT and AIM approach. RSC Adv.. 2016;6:10024-10037.
- [Google Scholar]
- Inhibitor selectivity between aldo–keto reductase superfamily members AKR1B10 and AKR1B1: Role of Trp112 (Trp111) FEBS Letters. 2013;587:3681-3686.
- [Google Scholar]
Appendix A
Supplementary material
Supplementary data associated with this article can be found, in the online version, at https://doi.org/10.1016/j.arabjc.2017.10.012.
Appendix A
Supplementary material
Supplementary data 1
Supplementary data 1
Supplementary data 2
Supplementary data 2