Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Journal of the Indian Chemical Society 100 (2023) 101030

Contents lists available at ScienceDirect

Journal of the Indian Chemical Society


journal homepage: www.journals.elsevier.com/journal-of-the-indian-chemical-society

Nitrogen as a probable problematic factor of computational chemistry: A


benchmarking study
Mert Metin a, *, Tomonori Kawano a, Tadashi Okobira b
a
Graduate School of Environmental Engineering, The University of Kitakyushu, 1-1 Hibikino, Wakamatsu, Kitakyushu, Fukuoka, 808 0135, Japan
b
Department of Creative Engineering, National Institute of Technology, Ariake College, 150 Higashihagio, Omuta, Fukuoka, 836-8585, Japan

A R T I C L E I N F O A B S T R A C T

Keywords: In the present study, theoretical IR frequencies and geometric parameters of triamterene, furosemide, and
Amine triamterene-furosemide salt were investigated. ORCA software was handled for commonly used BLYP, B3LYP,
Benchmark and HF (combined with various basis sets including 6-31G**, def2-SVP, cc-pVDZ, and pcseg-1 basis sets) cal­
Computational chemistry failure
culations. Theoretical IR frequencies and geometric parameters of triamterene and furosemide were compared
IR
Molecular geometry
with the corresponding experimental data. In addition, theoretical geometric parameters of triamterene-
Nitrogen furosemide salt were, as well, compared with the corresponding experimental data. Based on these compari­
sons, the performances of computational methods/basis sets were measured. As a result of this study, it was
discovered that the studied computational approaches were generally successful in the prediction of molecular
geometries and IR frequencies of the target molecules. When IR and geometry data were more deeply investi­
gated, it was observed that there were problems specific to nitrogen atoms. The computed geometry and IR of
amine and sulphonamide nitrogen atoms’ bonds strongly disagreed with the experimental data. Similar dis­
agreements were also observed for sulphur, oxygen, and cyclic hydrogen atoms but the level of disagreements
was lower.

1. Introduction structures of large chemical systems [19,20]. Results from such in­
vestigations are required to be supported by theoretical studies. Quan­
Computational chemistry is a field of science that uses computational tum chemical studies are generally highly reliable for a meaningful
methods to study chemical phenomena [1]. It is a powerful tool that comparison with the related experimental data [21]. The requirement of
complements experimental approaches in many areas of chemistry [2], huge computational resources renders quantum chemical studies on
such as drug discovery [3–8], materials science [9–12], and environ­ large chemical systems highly problematic. Especially, the computation
mental chemistry [9,13–16]. Computational methods can provide in­ of the IR/Raman spectra of large chemicals by the employment of
sights into the behaviour of molecules, atoms, and ions at a microscopic correlated ab initio approaches is a herculean mission [22].
level, which can be difficult to obtain experimentally [1]. One of the The choice of method/basis set pair can critically act on the accuracy
main advantages of computational chemistry is that it allows re­ and the speed of QM calculations. In general, stand-alone DFT methods
searchers to explore a wide range of chemical systems and conditions, are computationally very cheap and can provide reasonable accuracy.
including those that are difficult or impossible to study experimentally Differently, MP2, MP3, and MP4 methods (some enhanced versions of
[2]. Additionally, computational chemistry can be used to simulate HF) are computationally expensive and known to provide high-level
complex systems, such as biological membranes, and study their accuracy. On the other hand, hybrid functionals like B3LYP [3,23–26]
behaviour under different conditions [17]. As computational methods and M06-2X [27–29] are very popular for their good performance in
continue to improve, their use in science is likely to become even more terms of accuracy and their reasonable computational cost. HF is a very
widespread [1]. old method with known deficiencies, however, it is even in use and can
Advancements in spectroscopy rendered the detection of structures provide desired results depending on the case [30–33].
and properties of chemicals possible [18]. Particularly, IR and Raman Similar to method choice, basis set preference can change the results
spectroscopy are frequently utilized approaches for the investigation of significantly. Various basis set families are available in many QM

* Corresponding author.
E-mail addresses: mertmetin3335@gmail.com (M. Metin), kawanotom@gmail.com (T. Kawano), okobira@ariake-nct.ac.jp (T. Okobira).

https://doi.org/10.1016/j.jics.2023.101030
Received 22 January 2023; Received in revised form 19 April 2023; Accepted 20 May 2023
Available online 20 May 2023
0019-4522/© 2023 Indian Chemical Society. Published by Elsevier B.V. All rights reserved.
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

programs. Some basis sets are parameterized for some special functions. All DFT calculations were dispersion corrected by D3BJ [86,87].
For instance, Jensen basis sets (i.e. pcseg-1) are specifically designed for Firstly, the structures were processed by BLYP/def2-SVP method/
the computation of NMR spectra [34]. Some basis sets are limited to basis set. The BLYP/def2-SVP optimized structures were verified to be
some elements such as Pople basis sets (i.e. 6-31G*) [35,36]. Pople basis minima (apparent by the absence of negative frequencies). Thus, the
sets are very old basis sets but they are even now frequently referred to BLYP/def2-SVP optimized structures were handled as inputs for B3LYP/
Refs. [37–40]. In addition to the family of basis sets, polarization and def2-SVP and other DFT calculations. The B3LYP/def2-SVP optimized
diffusion/augmentation functions can own significant influences on the structures were verified to be minima (apparent by the absence of
computations [41]. negative frequencies) and so they were used as inputs for HF/def2-SVP
Another notable but less discussed issue of quantum chemistry could and for other B3LYP calculations. The HF/def2-SVP optimized struc­
be the choice of software. There is various QM software, like ORCA [42], tures were verified to be minima (apparent by the absence of negative
MOPAC [43], Psi4 [44–46], MOLPRO [47,48], Gamess US [49], frequencies) and so they were used as inputs for other HF calculations.
Gaussian [50], CP2K [51], QUICK [52], Quantum Espresso [53–55], Some calculations ended up with negative frequencies, which signs
Turbomole [56,57], Q-Chem [58], VASP [59–61], and NWChem [62]. that the optimized structures were not minima. However, some of these
Each software has its own advantageous sides. While some are negative frequencies were not lower than − 100 cm− 1 [88]. Thus, it was
open-source (such as CP2K and Psi4), some (ORCA) are freely available concluded that such structures were not minima but they were very
for academic purposes as others (Turbomole and Q-Chem) can provide close to minima. This means that they were sufficiently well-optimized.
some discounts & promotions for academic users. MOPAC software is However, some calculations ended up with negative frequencies which
limited to semi-empirical methods while ORCA can handle both DFT and were lower than − 100 cm− 1. This drove us to use a different input
semi-empirical calculations. Similar to ORCA, Gaussian can handle a structure for these calculations. Instead of def2-SVP optimized struc­
wide range of calculation methods. Gaussian owns its special graphical tures, the starting pdb files were used as inputs for such calculations.
user interface (GUI) called GaussView [63]. GaussView has some prac­ Unfortunately, BLYP/6-31++G** calculation for furosemide again had
ticalities. This GUI has its database where the user can access the initial remarkable negative frequencies. So, BLYP/6-31++G** calculation for
atomic coordinates of some compounds, like furosemide [64]. furosemide was removed from the study.
Furosemide is a diuretic drug with low bioavailability due to its poor When ORCA starts a calculation, it starts with the calculation of the
solubility and poor membrane permeability. Triamterene is a potassium- bond lengths and angles of starting geometry. So, in the output file of
sparing diuretic with low bioavailability due to its poor solubility. Peng BLYP/def2-SVP calculation, geometrical parameters of the pdb file were
et al. successfully synthesized a novel drug-drug salt composed of tri­ also available. In this way, the authors had easy access to the experi­
amterene and furosemide. The apparent equilibrium solubility (in pH mentally detected bond lengths and angles of triamterene and furose­
2.0 buffer) of triamterene–furosemide salt was improved compared with mide. Theoretical bond lengths and triangular angles were also
furosemide and a physical mixture of triamterene with equimolar extracted from the output files. Later, theoretical bond lengths and an­
furosemide by 15.3-fold and 12.2-fold enhancements, respectively [65]. gles (by various methods/basis sets) were compared to experimental
In this study, the authors computationally investigated furosemide, bond lengths and angles by the calculation of correlation coefficients
triamterene, and furosemide-triamterene salt. The study benchmarks (R2) and Root Mean Square Error (RMSE). Based on R2 (Tables 2 and 3)
commonly handled computational approaches on two chemicals and RMSE (Tables 5 and 6) values, the performances of methods/basis
together with their salt. It is important to note that, in general, the sets were assigned and compared.
molecular sizes of the benchmark molecules are not large. Three of the
studied molecules, especially the salt, are very large chemicals to run a 2.2. Triamterene–furosemide salt
benchmark study on them. Therefore, the large sizes of the target
chemicals separate this study from the canonical benchmark studies of Triamterene–furosemide salt crystal structure was obtained in cif file
computational chemistry. format [65]. The cif file was opened in PyMOL and the molecule was
As a result of our investigation, we observed that computed data saved as a pdb file. The pdb file was used by Gabedit to create input files
related to amine and amide nitrogen atoms strongly disagree with the for ORCA quantum mechanics software. Shortly, the computational
experimental data (both geometries and IR frequencies), regardless of chemical calculations were done by the following methods/basis sets;
the method/basis set choice. We evaluated these results by referring to
our previous study on iminodiacetic acid (IDA) [41] and to studies from * BLYP method with 6-31G, 6-31G*, 6-31G**, 6-31+G**, 6-31++G**,
the literature [66–69]. 6-311G**, def2-SVP, cc-pVDZ, pcseg-1, def2-SVPD and aug-pcseg-1
basis sets
2. Methodology * HF method with def2-SVP basis set
* B3LYP method with def2-SVP basis set
2.1. Triamterene and furosemide
All DFT calculations were dispersion corrected by D3BJ.
Triamterene and furosemide crystal structures were obtained in cif Firstly, the structure was processed by BLYP/def2-SVP approach.
file format [70,71]. The cif file was opened in PyMOL [72–74] and the The BLYP/def2-SVP optimized structure was verified to be a minimum
molecule was saved as a pdb file. The pdb file was used by Gabedit [75] (apparent by the absence of negative frequencies). Thus, the BLYP/def2-
to create input files for ORCA quantum mechanics software [76,77]. SVP optimized structure was handled as input for B3LYP/def2-SVP and
Shortly, the computational chemical calculations were done by the other DFT calculations. The B3LYP/def2-SVP optimized structure was
following methods/basis sets; verified to be a minimum (apparent by the absence of negative fre­
quencies) and so it was used as an input for HF/def2-SVP calculation.
* BLYP method with 6-31G [35,36,78], 6-31* [35,36,78], 6-31G** Among BLYP calculations, BLYP/6-31G** and BLYP/6-311G** cal­
[35,36,78], 6-31+G** [35,36], 6–31++G** [35,36], 6-311G** culations ended up with negligible negative frequencies (one for each
[78–82], def2-SVP [83], cc-pVDZ [78,84,85], and pcseg-1 [34,78] calculation, nearly − 4 cm− 1). Because the values of negative frequencies
basis sets were very close to 0 cm− 1, these structures were thought to be very close
* HF method with 6-31G, 6-31*, 6-31G**, 6-31+G**, 6-31++G**, 6- to their minimum. Thus, such structures were counted as well optimized.
311G**, def2-SVP, cc-pVDZ, and pcseg-1 basis sets Differently, BLYP/6-31+G** and BLYP/6-31++G** calculations ended
* B3LYP method with 6-31G, 6-31*, 6-31G**, 6-31+G**, 6-31++G**, up with so many negative frequencies (some of them with high negative
6-311G**, def2-SVP, cc-pVDZ, and pcseg-1 basis sets values). This drove authors to use a different input structure instead of

2
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Table 1
Correlation Coefficients for The Geometric Parameters of Triamterene-Furosemide Salt.
In this table, the green colour was used for highlighting the best-performing approach
while the red colour was utilized for the worst-performing approach.

BLYP optimized one. So, it was tried to handle the starting pdb file as an The computationally calculated IR peaks were compared to the
input for BLYP/6-31+G** and BLYP/6-31++G** calculations. Howev­ experimental peaks by the calculation of correlation coefficients (Ta­
er, those calculations did not appear normal to proceed (they had bles 13 and 14). Correlation coefficients calculated both with Nujol mull
excessive iterations), therefore, they were terminated. Following that, it and KBr disk IR frequencies were handled to decide which approaches
was decided to give a try on other families of basis sets to have at least were the best and the worst performers. Here, RMSE values were not
one diffusion function-including calculation. Thus, additional BLYP/ computed because the trend similarity of the compared data was the
def2-SVPD and BLYP/aug-pcseg-1 calculations were carried out. Un­ interest, not numerical closeness.
fortunately, they also had so many negative frequencies (some of them For triamterene-furosemide salt, no vibrational frequency data were
with high negative values), similar to other diffusion function-including available to us. So, it was not possible to measure the performances of
calculations. So, a diffusion function-including calculation for tri­ methods/basis sets in terms of IR frequency prediction. As a solution, it
amterene–furosemide salt could not be included in the study. was tried to index all computations to one of the calculations. BLYP/
When ORCA starts a calculation, it starts with the calculation of the def2-SVP calculation was chosen as the target calculation to index all
bond lengths and angles of starting geometry. So, in the output file of others. Correlation coefficients were calculated to measure the similar­
BLYP/def2-SVP calculation, geometrical parameters of the pdb file were ity between BLYP/def2-SVP and other calculations in terms of IR peaks
also available. In this way, there was easy access to the experimentally (Table 15). By the comparison of their relative similarity to BLYP/def2-
detected bond lengths and angles of triamterene–furosemide salt. SVP computed IR frequencies, methods/basis sets were compared to
Theoretical bond lengths and triangular angles were also extracted from each other.
the output files. Later, theoretical bond lengths and angles (by various
methods/basis sets) were compared to experimental bond lengths and 2.4. NPA charges, reactivity indices, and dipole moment and
angles by R2 and RMSE values. Based on R2 (Table 1) and RMSE thermodynamic data
(Table 4), the performances of methods/basis sets were assigned and
compared. NPA charges were calculated via NBO analysis with the help of
In addition, the bond lengths and angles from B3LYP/def2-SVP cal­ JANPA software [93,94]. Only the charges of heteroatoms were detailly
culations were individually investigated for the details about disagree­ investigated. Heteroatom charges from different calculations were
ment sites between experimental and theoretical data (Tables 7, 8, 9, 10, compared (Table_NPA.docx).
11, and 12). Reactivity indices (eHOMO, eLUMO, electronegativity, and hard­
ness) were extracted from the result files of IR calculations, with the help
2.3. IR frequencies of Gabedit software. Reactivity indices from different calculations were
compared (Table_RI.docx).
In this study, theoretical IR peaks of triamterene and furosemide Dipole moment and thermodynamic data (entropy, inner energy,
were investigated and compared to corresponding experimental data. enthalpy, and Gibbs free energy) were easily available in the result files
All calculations were carried out by analytical Hessians. Basis sets do not of IR calculations. Dipole moment and thermodynamic data from
have a strong impact on scaling factors. Therefore, in all of the calcu­ different calculations were compared (Table_TD.docx).
lations, the use of scaling factors for the 6-311G* basis set was preferred.
HF calculations were scaled by 0.904, BLYP calculations were scaled by 2.5. Supplementary materials
0.998, and B3LYP calculations were scaled by 0.966 [89,90].
Experimentally detected IR peaks of triamterene and furosemide Some data was very long to present in this report. So, just the sum­
were obtained from Spectral Database for Organic Compounds (SDBS) mary for some parts, like the NPA charges section, was presented. As a
[91]. For IR peaks, the authors had two options, peaks measured by the result, there was a rich supplementary material and the authors needed
Nujol mull or KBr disk method. Both of the options were selected. It was to assign a dictionary for it. There are various supplementary files,
afforded to match experimental IR peaks with the computationally
produced ones with the help of the Vibrational Mode Automatic Rele­ • Supplementary_Material.xlsx includes nearly all the raw and pro­
vance Determination (VMARD) outputs from vibAnalysis software (Ta­ cessed data of this study.
bles 16 and 17) [92]. B3LYP/def2-SVP results were used for VMARD • Table_VMARD.docx includes VMARD analysis of B3LYP/def2-SVP
analysis (Table_VMARD.docx). frequencies together with the experimental data comparison.

3
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Table 2
Correlation Coefficients for The Geometric Parameters of Triamterene. In this table, the orange colour was
utilized to highlight HF calculations, the blue colour was used for BLYP calculations, and the purple colour was
utilized for B3LYP calculations, on the left side of the table. On the right side of the table, it was recognized that
6-31G and def2-SVP calculations were grouped and isolated from others. Thus, the red colour was utilized to
highlight grouped 6-31G calculations while the brown colour was used for def2-SVP calculations.

• Table_RI.docx includes reactivity indices data together with the • TFS_Geo_Difference includes the error calculation for the B3LYP/
comments on the issue. SVP geometry data of salt molecule.
• Table_NPA.docx includes heteroatom NPA charges data together • TFS_Geo_RMSE includes RMSE calculations for the geometry data of
with the comments on the issue. the salt.
• Table_TD.docx includes the dipole moment and thermodynamic • TFS_NPA includes the raw and processed data for the NPA charges of
data together with the comments on the issue. triamterene-furosemide salt.
• TFS_RI includes the raw and processed data for the reactivity indices
Supplementary_Material.xlsx is a crowded file which needs a sepa­ of triamterene-furosemide salt.
rate legend and here it is, • TFS_TD includes the raw and processed data for the dipole moment
and thermodynamic parameters of the salt.
• Atom_Labels_1 includes the atom number labels for the atoms of the • TFS_IR includes the raw and processed data for the IR frequencies of
salt molecule. the salt molecule.
• TFS_Geo includes the raw and processed data for the geometry pa­ • Atom_Labels_2 includes the atom number labels for the atoms of
rameters of triamterene-furosemide salt. triamterene.

4
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Table 3
Correlation Coefficients for The Geometric Parameters of Furosemide. In this table, the green colour was
used for highlighting the best-performing approach while the red colour was utilized for the worst-
performing approach.

Table 4
RMSE Values for The Geometry Data of Triamterene-Furosemide Salt. In this table, the orange colour
was utilized to highlight HF calculations, the blue colour was used for BLYP calculations, and the purple
colour was utilized for B3LYP calculations.

5
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Table 5
RMSE Values for The Geometry Data of Triamterene. In this table, the orange colour was utilized to highlight HF
calculations, the blue colour was used for BLYP calculations, and the purple colour was utilized for B3LYP calculations.

• Tri_Geo includes the raw and processed data for the geometry pa­ • Fur_Geo_RMSE includes RMSE calculations for the geometry data of
rameters of triamterene. furosemide.
• Tri_Geo_Difference includes the error calculation for the B3LYP/ • Fur_NPA includes the raw and processed data for the NPA charges of
SVP geometry data of triamterene. furosemide.
• Tri_Geo_RMSE includes RMSE calculations for the geometry data of • Fur_RI includes the raw and processed data for the reactivity indices
triamterene. of furosemide.
• Tri_NPA includes the raw and processed data for the NPA charges of • Fur_TD includes the raw and processed data for the dipole moment
triamterene. and thermodynamic parameters of furosemide.
• Tri_RI includes the raw and processed data for the reactivity indices • Fur_IR includes the raw and processed data for the IR frequencies of
of triamterene. furosemide.
• Tri_TD includes the raw and processed data for the dipole moment
and thermodynamic parameters of triamterene. 2.6. Atom labels
• Tri_IR includes the raw and processed data for the IR frequencies of
furosemide. In the article, some tables and descriptions need information about
• Atom_Labels_3 includes the atom number labels for the atoms of atom numbers/labels of the studied compounds. So, just below, atom
furosemide. number labels for triamterene, furosemide, and their salt were provided.
• Fur_Geo includes the raw and processed data for the geometry pa­ Figures were obtained with the help of Avogadro and Gabedit soft­
rameters of furosemide. ware. Readers should note that Avogadro and Gabedit programs differ
• Fur_Geo_Difference includes the error calculation for the B3LYP/ from ORCA in their way of labelling atoms of target molecules. They
SVP geometry data of furosemide. start counting atoms from 1 while ORCA starts from 0. JANPA and
vibAnalysis handle the same numbering mechanism. Please investigate

6
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Table 6
RMSE Values for The Geometry Data of Furosemide. In this table, the orange colour was utilized to highlight HF
calculations, the blue colour was used for BLYP calculations, and the purple colour was utilized for B3LYP
calculations.

this document according to this knowledge. prediction of bond lengths of triamterene (Table 2, left side), all
methods/basis sets performed well. The best-performing approach was
3. Results BLYP/6-31G while the performance of HF/def2-SVP was lower than
others. On the left side of Table 2, BLYP calculations dominate the up­
3.1. Bond lengths and triangular angles side of the list, B3LYP dominates in the middle, and HF dominates the
downside. This means that the choice of the method matters for the
3.1.1. Comparison of the geometries via R2 computation of triamterene bond lengths. Additionally, HF/6-31G per­
According to correlation coefficients presented in Tables 1 and in the formed better than other HF calculations, BLYP/6-31G performed better
prediction of bond lengths of triamterene-furosemide salt (Table 1, left than other BLYP calculations, and B3LYP/6-31G performed better than
side), BLYP/6-31G** performed the worst. Oppositely, HF/def2-SVP other B3LYP calculations. HF/def2-SVP performed worse than other HF
performed the best. Except for BLYP/6-31G** and BLYP/6-31G, all calculations, BLYP/def2-SVP performed worse than other BLYP calcu­
other methods/basis sets performed somehow well. The performances of lations, and B3LYP/def2-SVP performed worse than other B3LYP cal­
HF/def2-SVP, B3LYP/def2-SVP, and BLYP/def2-SVP were similar. culations. This shows that the computation of bond lengths of
In the case of triangular angles (Table 1, right side), all methods triamterene was affected by the polarization function and by the use of
seem to fail in the prediction of the bond angles of triamterene- different basis set families.
furosemide salt. HF/def2-SVP was more successful than other In the prediction of triangular angles of triamterene (Table 2, right
methods/basis sets and BLYP/6-31G** performed the worst. HF/def2- side), all methods/basis sets had low-level success. However, BLYP/6-
SVP and B3LYP/def2-SVP outperformed BLYP/def2-SVP. 31G* performed better than others. It is easily recognizable that, on
Correlation coefficients suggest that HF/def2-SVP was the most the right side of Table 2, def2-SVP and 6-31G calculations were grouped
successful method/basis set pair and BLYP/6-31G** performed the at the downside of the list. So, def2-SVP and 6-31G calculations were the
worst, in the computation of the geometry parameters of triamterene- least successful approaches. This demonstrates that the computation of
furosemide salt. triangular angles of triamterene was strongly affected by the polariza­
According to correlation coefficients presented in Tables 2 and in the tion function and by the use of different basis set families.

7
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Table 7
Triamterene-Furosemide Salt Bond Lengths Displaying Remarkable Differences Between the B3LYP/def2-SVP
and Experimental Results (≥2%, in Ǻ). In the table, blue colour was handled for N bonds and yellow for S
bonds. For atom labels of triamterene-furosemide salt, see Fig. 1.

Correlation coefficients suggest that the computation of triamterene Correlation coefficients suggest that the prediction of furosemide
geometry was strongly affected by the polarization function and the use geometry was best performed by B3LYP/6-31++G** approach and
of different basis set families. BLYP/6-31G* and/or BLYP/6-31G** can worst performed by BLYP/6-31G as both bond lengths and triangular
be evaluated as the most successful methods/basis sets among studied angles were investigated.
approaches, in the prediction of triamterene geometry.
According to correlation coefficients presented in Tables 3 and in the 3.1.2. Comparison of the geometries via RMSE
prediction of bond lengths of furosemide (Table 3, left side), all ap­ According to Table 4, all RMSE values for the bond lengths were
proaches except HF/6-31G and BLYP/6-31G performed somehow well. greater than 0.05 and all RMSE values for the bond angles were less than
HF/6-31+G** and HF/6-31++G** performed the best while BLYP/6- 5. This could be interpreted as a sign of the better functioning of all
31G performed the worst. calculations in the case of the bond angles, as compared to the bond
In the prediction of triangular angles of furosemide (Table 3, right lengths of triamterene-furosemide salt. Additionally, the method choice
side), all of the studied approaches failed. HF/6-31G performed the best seems to strongly act on the RMSE values, and so on the success ratio, for
while BLYP/6-31G performed the worst. the bond lengths and angles of triamterene-furosemide salt.

8
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Table 8
Triamterene-Furosemide Salt Triangular Angles Displaying Remarkable Differences Between the B3LYP/
def2-SVP and Experimental Results (≥2%, in Ǻ). In the table, blue colour was handled for N bonds,
yellow for S bonds, and red for O bonds. For atom labels of triamterene-furosemide salt, see Fig. 1.

Table 9
Triamterene Bond Lengths Displaying Remarkable Differences Between the B3LYP/def2-SVP and
Experimental Results (≥2%, in Ǻ). In the table, blue colour was handled for N bonds. For atom labels of
triamterene, see Fig. 2.

Table 10
Triamterene Triangular Angles Displaying Remarkable Differences Between the B3LYP/def2-SVP and
Experimental Results (≥2%, in Ǻ). In the table, blue colour was handled for N bonds. For atom labels of
triamterene, see Fig. 2.

RMSE values for the bond lengths and angles propose that HF/def2- calculations in the case of the bond angles, as compared to the bond
SVP was the best-performing approach while BLYP/6-31G** was the lengths of triamterene. Additionally, the method choice seems to
worst performer for the molecular geometry of triamterene-furosemide strongly act on the RMSE values for the bond lengths of triamterene as
salt. the absence of polarization function and the utilization of def2-SVP basis
According to Table 5, all RMSE values for bond lengths were greater set caused worse performance in RMSE values for the bond angles.
than 0.05 and all RMSE values for the bond angles were highly less than RMSE values for the bond lengths and angles propose that HF/6-
5. This could be interpreted as a sign of the better functioning of all 311G** was the best-performing approach while BLYP/6-31G was the

9
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Table 11
Furosemide Bond Lengths Displaying Remarkable Differences Between the B3LYP/def2-SVP and
Experimental Results (≥2%, in Ǻ). In the table, blue colour was handled for N bonds, yellow for S bonds,
and red for O bonds. For atom labels of furosemide, see Fig. 3.

Table 12
Furosemide Triangular Angles Displaying Remarkable Differences Between the B3LYP/def2-SVP and
Experimental Results (≥2%, in Ǻ). In the table, blue colour was handled for N bonds, yellow for S bonds,
and red for O bonds. For atom labels of furosemide, see Fig. 3.

worst performer for the molecular geometry of triamterene. bonds of O9-H40 and O2-H53. These are the bonds binding triamterene
According to Table 6, all RMSE values for bond lengths were greater with furosemide. So, the hydrogen bond lengths are truly predicted.
than 0.05 and all RMSE values for bond angles were less than 5. This Similarly, the bond angles associated with the O9-H40 and O2-H53
could be interpreted as a sign of the better functioning of all calculations bonds were not in Table 8. Thus the bond angles for O9-H40 and O2-
in the case of the bond angles, as compared to the bond lengths of H53 bonds were also well-predicted. These observations are remark­
furosemide. Additionally, the method choice seems to strongly act on able and positive regarding the performance of the B3LYP/def2-SVP
the RMSE values for bond lengths and angles of furosemide. approach.
RMSE values for the bond lengths and angles propose that HF/6- Table 8 demonstrates that the sulphonamide N-H bond angle and
31+G** was the best-performing approach while BLYP/6-31G* was the pentagonal bond angles of triamterene-furosemide salt have relatively
worst performer for the molecular geometry of furosemide.. large errors in comparison with the other bonds. The error was larger for
Table 7 indicates that B3LYP/def2-SVP bond lengths for triamterene- the N-H bond angle. The errors in bond angles (less than 10%) were
furosemide salt strongly disagree with the experimental data in the case smaller and less abundant in comparison with the errors in bond lengths
of amine and sulphonamide N-H bonds and cyclic C-H bonds. The level (which can easily exceed 10%)..
of disagreement was numerically very larger for amine N-H bonds as Table 9 indicates that B3LYP/def2-SVP bond lengths for triamterene
compared to the C-H bonds. The table does not include the hydrogen strongly disagree with the experimental data in the case of amine N-H

10
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Table 13
Correlation Coefficients for IR Frequencies of Triamterene-Furosemide Salt. Here, the red
colour was utilized to sign HF/def2-SVP calculation which was least similar to BLYP/def2-
SVP.

bonds and benzene C-H bonds. The level of disagreement was numeri­ frontier factor influencing the IR frequencies of the studied molecules.
cally a little bit larger for the C-H bonds as compared to the amine N-H
bonds. 3.2.1. VMARD analysis
Table 10 shows that amine N-H bond angles of triamterene have Table 16 shows that B3LYP/def2-SVP IR frequencies strongly
relatively large errors in comparison with the other bonds. The errors in disagree with the experimental data in the case of amine N-H bonds,
bond angles (less than 6%) were smaller and less abundant in compar­ benzene C-H bonds, cyclic N-H bonds, and cyclic C-N bonds. The level of
ison with the errors in bond lengths (which can easily exceed 10%). disagreement was numerically very large for amine N-H bonds as
Table 11 shows that B3LYP/def2-SVP bond lengths for furosemide compared to others.
strongly disagree with the experimental data in the case of amine and Table 17 shows that B3LYP/def2-SVP IR frequencies strongly
sulphonamide N-H bonds, carboxylic acid oxygen bonds, and cyclic C-H disagree with the experimental data in the case of amine N-H bonds,
bonds. The level of disagreement was numerically larger for amine N-H benzene C-H bonds, sulphonamide N-S bonds, and S-O bonds of sul­
bonds as compared to the C-H bonds. phonamide. The level of disagreement was numerically very large for
Table 12 demonstrates that carboxylic acid oxygen bond angles of amine N-H bonds, as compared to others.
furosemide have relatively large errors in comparison with the other
bonds. The errors in bond angles (less than 10%) were smaller and less 3.2. NPA charges, reactivity indices, and dipole moment and
abundant in comparison with the errors in bond lengths (which can thermodynamic data
exceed easily exceeds 10%)..
The method preference was the primary factor acting on the
3.2. IR frequencies computation of NPA charges of heteroatoms, reactivity indices, ther­
modynamic data, and dipole moment of three of the studied molecules.
According to Tables 13 and in IR frequencies, all calculations were For details, see Table_NPA.docx, Table_RI.docx, and Table_TD.
similar to BLYP/def2-SVP and so to each other. Among all approaches,
HF/def2-SVP was the least similar to BLYP/def2-SVP. Thus, HF differed 4. Discussion
also from other approaches. Therefore, the preference of method can
influence the IR results for the salt molecule. Here, in this study, the performances of computational chemical
In the prediction of IR frequencies of triamterene (Table 14), all approaches on the prediction of molecular properties of triamterene,
methods/basis sets performed well. The best-performing approach was furosemide, and triamterene-furosemide salt were investigated. Exper­
B3LYP/6-31+G** while the performance of BLYP/6-31G was lower than imental data for molecular geometries and IR frequencies were
others. compared to the theoretical ones. Both R2 and RMSE can be utilized as
Here, on the left side of Table 14, BLYP calculations accumulated at tools to measure the similarity between two data sets. However, RMSE
the bottom of the table. This could be a sign of the remarkable impact of concentrates on the numerical closeness of the compared data sets while
the method preference. However, no such order was observed for the R2 reflects the similarity in terms of the trend. For IR frequencies the
right side of the table. Nevertheless, the method choice could be the trend of the data matters more than the numerical closeness so R2 is
primary factor acting on the computation of IR frequencies of handled to compare IR frequencies. Differently, it was observed that
triamterene. both R2 [64] and RMSE [66] have been handled by chemists to compare
In the prediction of IR frequencies of furosemide (Table 15), all molecular geometry data (bond lengths and angles). Here, both R2 and
methods/basis sets performed well. The best-performing approach was RMSE calculations were applied to the molecular geometry data.
B3LYP/6-31G** while the performance of HF/6-31G was lower than According to R2 data, in the prediction of bond lengths of triamterene
others. Moreover, HF calculations accumulated at the bottom of and furosemide, all methods/basis sets performed well. The method
Table 15 (both right and left sides). Therefore, the method choice ap­ preference was the major factor acting on the accuracy of the compu­
pears as the major factor influencing the computation of IR frequencies tations. In the prediction of bond lengths of triamterene-furosemide salt,
of furosemide. except BLYP/6-31G** and BLYP/6-31G, all other methods/basis sets
In case all the observations done for IR frequencies of three of the performed somehow well. R2 data suggests that, in the case of triangular
studied chemicals were summed up, it can be suggested that BLYP/6- angles, all methods failed in the successful prediction of the experi­
31G** could be the optimal approach, in terms of computational cost mental angles of three of the studied molecules. Based on the R2 data of
and satisfactory accuracy. However, B3LYP/6-31G* appears to be the three of the studied chemicals, it can be suggested that BLYP/6-31G*
most accurate one. Furthermore, the method preference seems to be the could be chosen as the optimal approach for the prediction of the

11
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Table 14
Correlation Coefficients for IR Frequencies of Triamterene. In this table, the green colour was used for high­
lighting the best-performing approach while the red colour was utilized for the worst-performing approach.

molecular geometries of the studied compounds, in terms of computa­ be interpreted as a sign of the better functioning of all calculations in the
tional cost and satisfactory accuracy. However, B3LYP/6-31++G** was case of the bond angles, as compared to the bond lengths of the inves­
the one correlating best with the experimental data. tigated compounds. Such a proposal is the just opposite of what it was
According to RMSE data, in the prediction of bond lengths of tri­ said for R2 data which suggests that the performance of the utilized
amterene and furosemide, all methods/basis sets performed well. The methods/basis sets were very bad for bond angles and they were very
method preference appeared to be the major factor acting on the accu­ good for the bond lengths. It was found the answers for the conflict, as
racy of the computations. In general, both HF was superior to B3LYP and bonds and bond angles from B3LYP/def2-SVP calculation were indi­
BLYP and HF/6-31G* emerged as the most accurate approach. All RMSE vidually investigated (Tables 7, 8, 9, 10, 11, and 12). The errors in bond
values for the bond lengths of the studied compounds were greater than angles were smaller (less than 10%) and less abundant in comparison
0.05 and all RMSE values for the bond angles were less than 5. This could with the errors in bond lengths (which can exceed easily exceeds 10%) of

12
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Table 15
Correlation Coefficients for IR Frequencies of Furosemide. In this table, the green colour was used for high­
lighting the best-performing approach while the red colour was utilized for the worst-performing approach.

the three of the studied molecules. So, this indicates that the predictions experimental frequencies.
of the bond angles were more successful than those of bond lengths. In this report, significant sites of the VMARD analysis for B3LYP/
Therefore, it was concluded that R2 is not a good tool for the comparison def2-SVP frequencies were provided. R2 comparison of IR frequencies
of molecular geometry data and its use should be avoided in the case of suggested that nearly all methods performed well on the IR prediction
molecular geometry data. for the studied molecules. However, the VMARD analysis demonstrated
In the prediction of IR frequencies of triamterene and furosemide, all that there are significant site-specific problems with the computational
methods/basis sets seem to perform well. Based on our observations of IR frequencies, in terms of their association with the experimental data.
IR frequencies of three of the studied chemicals, it can be suggested that Especially, amine/amide N-H bonds both from triamterene and furose­
BLYP/6-31G** could be the optimal approach among the studied ap­ mide were problematic in this sense. These problems were, as well,
proaches, in terms of computational cost and satisfactory accuracy. shared by other methods/basis sets (for details, please visit Tri_IR and
However, B3LYP/6-31G* provided the best correlation with the Fur_IR Sheets in Supplementary_Material.xlsx), in addition to

13
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Table 16
Triamterene Frequencies Displaying Remarkable Differences Between the B3LYP/def2-SVP and Experimental Results (Frequencies in cm− 1). For the full version of the
table, see Table_VMARD.docx. For atom labels of triamterene, see Fig. 2.
KBr Disk Frequencies Nujol Mull Frequencies B3LYP/def2-SVP Frequencies Scaled Frequencies (0.966) Composition

3473 3473 3730.98 3604 -0.6952 (54.2%) BOND N11H26


+0.5728 (44.6%) BOND N11H25
3390 3389 3692.87 3567 -0.6898 (54.2%) BOND N13H29
+0.5828 (45.8%) BOND N13H30
3370 3370 3614.23 3491 +0.8199 (47.5%) BOND N12H27
+0.7902 (45.8%) BOND N12H28
3291 3286 3588.62 3467 -0.8702 (49.4%) BOND N11H25
-0.7280 (41.3%) BOND N11H26
3137 3127 3565.54 3444 +0.8522 (49.3%) BOND N13H30
+0.7264 (42.0%) BOND N13H29
2955 3201.44 3093 -0.8397 (46.8%) BOND C15H20
-0.3927 (21.9%) BOND C16H21
+0.2252 (12.5%) BOND C19H24
-0.1678 (9.3%) BOND C17H22
2925 3191.08 3083 +0.7286 (34.4%) BOND C17H22
-0.3671 (17.3%) BOND C15H20
+0.3349 (15.8%) BOND C18H23
+0.3134 (14.8%) BOND C16H21
-0.2441 (11.5%) BOND C19H24
2855 3179.05 3071 -0.6758 (35.8%) BOND C16H21
+0.6451 (34.2%) BOND C18H23
+0.2400 (12.7%) BOND C15H20
-0.1891 (10.0%) BOND C19H24
1679 1680 1682.45 1625 -0.4795 (17.1%) BOND C1 N2
+0.3159 (11.3%) BOND C1 N11
-0.2670 (9.5%) BOND N4 C5
1495 1495 1589.36 1535 -0.2394 (6.5%) BOND C3 N4
-0.2284 (6.2%) ANGLE H29 N13H30
-0.2233 (6.0%) BOND C1 C10
+0.2118 (5.7%) BOND C1 N2
-0.2111 (5.7%) BOND C3 N12
1287 1287 1364.48 1318 -0.3047 (8.7%) BOND C14C19
+0.2983 (8.5%) BOND C16C17
-0.2803 (8.0%) BOND C17C18
-0.2778 (7.9%) BOND C15C16
+0.2671 (7.6%) BOND C14C15
+0.2537 (7.2%) BOND C18C19

B3LYP/def2-SVP. While this study was being reviewed, our group studied more on IDA.
Actually, our group have heard rumours (Acknowledgements) As a result, we suggested that the O-H and N-H bonds of an IDA molecule
about the failure of computational chemistry approaches in the case of could be linked to other IDA molecules’ N and O atoms via hydrogen
nitrogen (specifically its bond with hydrogen). Moreover, we have bonds, in the solid phase. Although probably single molecule compu­
experienced problems with this issue. In our previous study on IDA, HF, tation significantly affected the results for IDA vibrational frequencies,
B3LYP, and semi-empirical calculations failed in the prediction of IR the error cannot be attributed only to this factor. Nevertheless, the failed
frequencies. IDA is a small molecule with a secondary amine structure. predictions for IDA appear to strongly associate with the performance of
The disagreement between experimental and computational data was computational chemistry approaches (unpublished, Benchmarking
very strong specifically in the frequencies higher than 2200 cm− 1. Now, computational chemistry approaches on iminodiacetic acid, Part
we guess that amine nitrogen (specifically its bond with hydrogen) II).
played a role in the problem there. In addition, we have observed some N-H bonds are known to be strongly affected by hydrogen bonds
problems with the N-H bond in the case of proton NMR chemical shifts [98]. A similar situation could be observed also in the case of O-H bonds
(unpublished, Study of benchmarking methodologies for 1H NMR and C-H bonds of benzene in case there is hydrogen bonding. Let’s as­
chemical shifts of thiazolidin-4-one derivative; A novel drug sume that N-H bond lengths in the triamterene side of the salt molecule
candidate). In our new study, we studied 2-(Furan-2-ylmethylenehy­ were wrongly predicted because of hydrogen bonding with neighbour­
drazono)-5-(3-methylbenzyl)thiazolidin-4-one with HF/6-31G, HF/6- ing salt molecules. Then, how the case of the N2-H41 bond (Fig. 1)
31G**, HF/def2-SVP, HF/def2-SVPD, HF/def2-TZVP, B3LYP/def2-SVP, should be described? In triamterene-furosemide salt, N2-H41 and
BLYP/def2-SVP, PBE0/def2-SVP, PBE/def2-SVP, TPSSh/def2-SVP, N6-H54 bonds (Fig. 1) connect two sides of the molecule with hydrogen
TPSS/def2-SVP, and BP86/def2-SVP. However, none of the studied ap­ bonds. According to Table 7, the N2-H41 bond was the most problematic
proaches could remediate the strong disagreement between the theo­ bond to be predicted. The error rate was 31.42%. N2-H41 bond already
retical and experimental proton NMR chemical shift for the N-H site owns a hydrogen bond with the neighbouring oxygens. Authors don’t
(unpublished, Study of benchmarking methodologies for 1H NMR know how salt molecules interact with each other and the way they are
chemical shifts of thiazolidin-4-one derivative; A novel drug organized in the solid phase. Although such bonding, in the cif file of the
candidate). As opposed to such observations by our group, Karakaya salt, was not observed, possibly, N2 may link to hydrogen from other salt
et al. successfully reached a good agreement between theoretical and molecules. However, such bonding should not lead to an extreme change
experimental IR and geometry data for the nitrogenous sites of the gli­ causing a 31.42% prediction error. Such errors should be also related to
clazide molecule, by B3LYP and HF calculations [64]. However, our the performance of computations. In all B3LYP, HF, and BLYP calcula­
group was not the single group that experiences problems with nitrogen tions of the salt molecule, the N2-H41 bond length was predicted
atoms [66,67,97]. wrongly (TFS_Geo Sheet in Supplementary_Material.xlsx).

14
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Table 17
Furosemide Frequencies Displaying Remarkable Differences Between the B3LYP/def2-SVP and Experimental Results (Frequencies in cm− 1). For the full version of the
table, see Table_VMARD.docx. For atom labels of furosemide, see Fig. 3.
KBr Disk Frequencies Nujol Mull Frequencies B3LYP/def2-SVP Frequencies Scaled Frequencies (0.966) Composition

3401 3401 3671.76 3547 -0.6433 (50.2%) BOND N11H31


+0.6372 (49.8%) BOND N11H32
3351 3352 3535.62 3415 +0.8091 (46.4%) BOND N11H32
+0.8038 (46.1%) BOND N11H31
3284 3287 3490.31 3372 +0.9841 (77.6%) BOND N2 H22
2955 3240.07 3130 +0.9776 (85.3%) BOND C5 H23
-0.1002 (8.7%) BOND C8 H24
2924 3091.4 2986 +0.7270 (53.5%) BOND C9 H30
-0.5824 (42.9%) BOND C9 H29
2854 3041.67 2938 -0.8344 (53.4%) BOND C9 H29
-0.6815 (43.6%) BOND C9 H30
1674 1674 1786.38 1726 +0.8360 (49.7%) BOND C18 O20
-0.1099 (6.5%) BOND C4 C18
+0.0980 (5.8%) BOND C1 N2
1509 1504 1624.69 1569 +0.3693 (13.1%) BOND C1 C4
-0.2981 (10.6%) BOND C5 C6
-0.2854 (10.1%) BOND C1 N2
+0.1856 (6.6%) ANGLE C1 N2 H22
-0.1618 (5.8%) BOND C1 C8
1353 1368.77 1322 +0.6001 (45.0%) BOND S3 O12
-0.5917 (44.4%) BOND S3 O14
1078 1149.33 1110 -0.7400 (32.4%) BOND S3 O14
-0.7366 (32.3%) BOND S3 O12
540 588.58 569 +0.5366 (10.5%) TORSION C4 C18 O19H28
+0.5355 (10.5%) OUT N2 C1 C4 C8
+0.4904 (9.6%) TORSION C5 C4 C18 O19
+0.3160 (6.2%) TORSION N2 C1 C4 C5
514 577.74 558 -0.4968 (8.8%) ANGLE C6 S3 O12
-0.4920 (8.7%) ANGLE C6 S3 O14
-0.4242 (7.5%) ANGLE O19C18 O20
+0.3445 (6.1%) BOND C4 C18

Although we heard some rumours, a report specifically emphasizing well-functioning of the approach of interest on the studied chemicals
the problems with nitrogen atoms discussed here could not be found. In and their investigated sites, the computation will be safer.
case there is, this study will be a good supplementation to that for all 5
molecules (the three being from this study + IDA + the thiazolidin-4-one 4.1. What are the possible solutions?
derivative) our group have studied in our past and current studies,
B3LYP, like other methods, did not end up with a trustable level of ac­ With inspiration from the work by Spinn et al. [67], we propose that
curacy for the nitrogenous sites. Amines cover a huge portion of organic the addition of an extra point charge nearby nitrogen atoms could be a
and biological chemistry and they play very significant roles in biolog­ good solution to the problems regarding N-H bonds. We did not have the
ical and environmental processes. So, in the case where N-H bonds play opportunity to try such a solution thus our group encourage readers to
significant roles and computations on nitrogen are problematic, there do it. If any of the readers go for it and achieves success, the authors will
could be wrong conclusions on biological and chemistry issues. In such a be happy to know it. In a such situation please communicate with our
situation, the use of computational chemistry could be a problem, rather group through the corresponding author.
than a remedy. Authors don’t know where such problems with nitrogen
originate from.
4.2. Other approaches and techniques
In the literature, some cases where computations for N-H bonds
failed can be found [68,69,99–104]. For example, the N-H bond
Authors could only study with HF, BLYP, and B3LYP methods, in this
stretching frequency for N-(2,5-dimethyl-4-nitrophenyl)-4-methylben
study. But, there is a need to measure the success of higher-level ap­
zenesulfonamide was predicted with a 5.2% error rate by
proaches like B2LYP, a double hybrid, and MP2. However, such ap­
B3LYP/6-311++G** approach while the error rate was around 39%
proaches can also fail like the methods studied, here. They may provide
when NMR shift for the nitrogen-bonded hydrogen was computed by
just fine-tuned versions of the results provided by HF and B3LYP. It is
B3LYP/6-311+G(2d,p) [103]. In support of this, Karrouchi et al. pre­
also a good idea to go for another benchmark study with three of the
dicted (by B3LYP/6-311++G**) NMR shifts of the nitrogen-linked hy­
studied molecules to test the performances of some new methods, such
drogens of (E)-N’-(4-methoxybenzylidene)-5-methyl-1H-pyrazole-
as GFN2-xTB, also known as XTB2 [106]. Further, some other ami­
3-carbohydrazide with 29.2% and 29.6% error rates [100]. Especially,
ne/amide molecules should be investigated. However, vibrational
the problem becomes more noticeable in the case of proton NMR shifts
spectrometry data for large molecules are crowded and complex. As
[97]. However, it is significant to emphasize that failure is not the case
opposed to this, computational and experimental NMR data is easier to
for every computation [105].
study. Therefore, this study could be followed by an NMR study, rather
Based on the above discussion, the authors propose that before deep
than an IR investigation.
investigations of the molecules by computational chemistry approaches,
it could be a good option to first investigate if the selected computational
5. Conclusion
approach is functioning sufficiently well on the chemical of interest. For
that, a comparison with any available experimental data, like crystal­
In this study, various commonly used methods/basis sets were
lography, vibrational spectrometry, bond lengths, bond angles, and/or
benchmarked mainly on molecular geometry and IR frequencies of tri­
NMR chemical shifts, could be useful. In the case there seems general
amterene, furosemide, and their salt. The method preference was the

15
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Fig. 1. Atom labels for triamterene-furosemide salt. The B3LYP/def2-SVP optimized structure was visualized by Avogadro software [95,96].

primary factor influencing the fate of the results for molecular geome­
tries, IR frequencies, NPA charges of heteroatoms, reactivity indices,
thermodynamic data, and dipole moment of three of the studied mole­
cules. The studied computational approaches generally performed well
in the IR frequency and molecular geometry prediction missions.
However, some site-specific problems were recognized. In the study,
amine and amide sites of the studied molecules were unsuccessfully
predicted, by the studied approaches, in terms of molecular geometry
and IR spectroscopy.
Although it was not mentioned in the article, there is one more
probability behind the observed disagreements between the experi­
mental and theoretical data. The experimental data could be the prob­
lematic side, rather than the theoretical results. Even though
experiments are commonly accepted as a reference, various natural and/
or human-sourced factors can affect the experimental results. Although
this is not the case mostly, this scenario should be also mentioned and
kept in mind.
Fig. 2. Atom labels for triamterene. The B3LYP/def2-SVP optimized structure We wish that this study will guide the scientists who computationally
was visualized by Avogadro software. investigate amines/amides and computational chemistry techniques.

Funding

Our group thanks The Ministry of Education, Culture, Sports, Sci­


ence, and Technology (MEXT) of Japan for their grant of a PhD schol­
arship to author M.M. while this study is being done.

Availability of data and material

All data generated or analysed during this study are included in this
published article and its supplementary files.

Code availability

Here, in this study, authors handled ORCA (version 4.2.1), Gabedit


(version 2.5.2), Avogadro (version 1.2.0), JANPA (version 2.02), vibA­
nalysis (version 1.2.2), and PyMOL (version 2.5.0). All of these programs
were freely and easily available from their website except ORCA. ORCA
was freely available to M.M. for academic purposes. In the enrichment
process of the introduction section ChatGPT [107] was utilized while
Grammarly [108] was handled for the improvement of the article’s
Fig. 3. Atom labels for furosemide. The B3LYP/def2-SVP optimized structure
was visualized by Avogadro software.
grammar and fluency.

16
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Authors’ contributions [19] H. Kim, M. Cho, Infrared probes for studying the structure and dynamics of
biomolecules, Chem. Rev. 113 (8) (2013) 5817–5847.
[20] Y. Fan, J. Ho, R.P. Bettens, Approximating Coupled Cluster level vibrational
The conception and design of the study: M.M. and T.K. frequencies with composite methods, J. Phys. Chem. 110 (8) (2006) 2796–2800.
Acquisition of data, analysis, and interpretation of data: M.M. [21] R.A. Friesner, Ab initio quantum chemistry: methodology and applications, Proc.
Drafting the article: M.M. Natl. Acad. Sci. USA 102 (19) (2005) 6648–6653.
[22] S.S. Khire, N. Sahu, S.R. Gadre, Harnessing desktop computers for ab initio
Proofreading: T.O. calculation of vibrational IR/Raman spectra of large molecules, J. Chem. Sci. 130
(11) (2018) 1–7.
Declaration of competing interest [23] T.M.A. Al-Shboul, et al., Synthesis, characterization, computational and
biological activity of some schiff bases and their Fe, Cu and Zn complexes,
INORGA 10 (8) (Aug. 2022), https://doi.org/10.3390/inorganics10080112. Art.
The authors declare that they have no conflict of interest. no. 8.
[24] I.B. Isik, N. Tekin, S.G. Sagdinc, The analyses of solvent effects on infrared spectra
and thermodynamic parameters, Hirshfeld surface, reduced density gradient and
Acknowledgements molecular docking of ketoprofen as a member of nonsteroidal anti-inflammatory
drugs, J. Mol. Struct. 1250 (Feb. 2022), 131861, https://doi.org/10.1016/j.
The authors also want to thank Uezu Kazuya from The University of molstruc.2021.131861.
[25] D. Rajaraman, L.A. Anthony, P. Nethaji, R. Vallangi, One-pot synthesis, NMR,
Kitakyushu for sharing his experiences with computational chemistry. quantum chemical approach, molecular docking studies, drug-likeness and in-
silico ADMET prediction of novel 1-(2,3-dihydrobenzo[b][1,4]dioxin-6-yl)-2-
Appendix A. Supplementary data (furan-2-yl)-4,5-diphenyl-1H-imidazole derivatives, J. Mol. Struct. 1273 (Feb.
2023), 134314, https://doi.org/10.1016/j.molstruc.2022.134314.
[26] A. Mili, et al., Tautomerization of diazene<=>hydrazine via single proton
Supplementary data to this article can be found online at https://doi. tautomerization, spectral, XRD/HSA-interactions, optical and DFT/TD-DFT of
org/10.1016/j.jics.2023.101030. new hydrazine ligand, J. Mol. Struct. 1272 (Jan. 2023), 134113, https://doi.org/
10.1016/j.molstruc.2022.134113.
[27] H. Zhang, H. Zhou, H. Qin, J. Liu, W. Fan, The mechanism for the oxidation of
References benzoquinoline: a theoretical study, Combust. Flame 248 (Feb. 2023), 112553,
https://doi.org/10.1016/j.combustflame.2022.112553.
[1] C.J. Cramer, Essentials of Computational Chemistry: Theories and Models, John [28] A. Briš, Y. Murata, Y. Hashikawa, D. Margetić, Utilization of sym-tetrazines as
Wiley & Sons, 2013. guanidine delivery cycloaddition reagents. An experimental and computational
[2] F. Jensen, Introduction to Computational Chemistry, John wiley & sons, 2017. study, J. Mol. Struct. 1272 (Jan. 2023), 134207, https://doi.org/10.1016/j.
[3] H. Boulebd, Theoretical insights into the antioxidant activity of moracin T, Free molstruc.2022.134207.
Radic. Res. 54 (4) (2020) 221–230. [29] M.Z. Sadvakassova, et al., Rotational barriers in N-benzhydrylformamides: an
[4] G. Palermo, M. De Vivo, Computational chemistry for drug discovery, in: NMR and DFT study, Molecules 28 (2) (Jan. 2023), https://doi.org/10.3390/
B. Bhushan (Ed.), Encyclopedia of Nanotechnology, Springer Netherlands, molecules28020535. Art. no. 2.
Dordrecht, 2014, pp. 1–15, https://doi.org/10.1007/978-94-007-6178-0_ [30] J.E.M. McGeoch, M.W. McGeoch, Chiral 480 nm absorption in the hemoglycin
100975-1. space polymer: a possible link to replication, Sci. Rep. 12 (1) (Sep. 2022), https://
[5] G. Sliwoski, S. Kothiwale, J. Meiler, E.W. Lowe, Computational methods in drug doi.org/10.1038/s41598-022-21043-4. Art. no. 1.
discovery, Pharmacol. Rev. 66 (1) (Jan. 2014) 334–395, https://doi.org/ [31] S. K. Mishra, A. Parikh, K. Rangan, and A. K. Sah, “Crystal structure of N-(2-
10.1124/pr.112.007336. Hydroxynapthylidene)-L-isoleucinyl-4,6-O-ethylidene-β-D-glucopyranosylamine
[6] P.A. Townsend, M.N. Grayson, Density functional theory in the prediction of and an insight from experimental and theoretical calculations,” Cryst. Res.
mutagenicity: a perspective, Chem. Res. Toxicol. 34 (2) (Feb. 2021) 179–188, Technol., vol. n/a, no. n/a, p. 2200209, doi: 10.1002/crat.202200209.
https://doi.org/10.1021/acs.chemrestox.0c00113. [32] C. Sudhakar, et al., Pharmacological and quantum chemical studies of 2-amino­
[7] E. Köse, M. Erkan Köse, S. Güneşdoğdu Sağdınç, Principal component analysis of benzo[d]thiazol-3-ium 4-chlorobenzenesulphonate: synthesis, spectral, thermal
quantum mechanical descriptors data to reveal the pharmacological activities of analysis and structural elucidation, Results in Chemistry 4 (Jan. 2022), 100442,
oxindole derivatives, Results in Chemistry 5 (Jan. 2023), 100905, https://doi. https://doi.org/10.1016/j.rechem.2022.100442.
org/10.1016/j.rechem.2023.100905. [33] P.S. Idante, G.C. Apebende, H. Louis, I. Benjamin, U.J. Undiandeye, I.J. Ikot,
[8] I.Y. Joel, et al., Insights into features and lead optimization of novel type 1½ Spectroscopic, DFT study, and molecular docking investigation of N-(3-
inhibitors of p38α mitogen-activated protein kinase using QSAR, quantum methylcyclohexyl)-2-phenylcyclopropane-1-carbohydrazide as a potential
mechanics, bioisostere replacement and ADMET studies, Results in Chemistry 2 antimicrobial drug, J. Indian Chem. Soc. 100 (2) (Feb. 2023), 100806, https://
(Jan. 2020), 100044, https://doi.org/10.1016/j.rechem.2020.100044. doi.org/10.1016/j.jics.2022.100806.
[9] J. Bhawsar, P.K. Jain, P. Jain, Experimental and computational studies of [34] F. Jensen, Unifying general and segmented contracted basis sets. Segmented
Nicotiana tabacum leaves extract as green corrosion inhibitor for mild steel in polarization consistent basis sets, J. Chem. Theor. Comput. 10 (3) (2014)
acidic medium, Alex. Eng. J. 54 (3) (Sep. 2015) 769–775, https://doi.org/ 1074–1085.
10.1016/j.aej.2015.03.022. [35] M.M. Francl, et al., Self-consistent molecular orbital methods. XXIII. A
[10] G. Gece, The use of quantum chemical methods in corrosion inhibitor studies, polarization-type basis set for second-row elements, J. Chem. Phys. 77 (7) (1982)
Corrosion Sci. 50 (11) (Nov. 2008) 2981–2992, https://doi.org/10.1016/j. 3654–3665.
corsci.2008.08.043. [36] W.J. Hehre, R. Ditchfield, J.A. Pople, Self—consistent molecular orbital methods.
[11] A. Irfan, A. Mahmood, Designing of efficient acceptors for organic solar cells: XII. Further extensions of Gaussian—type basis sets for use in molecular orbital
molecular modelling at DFT level, J. Cluster Sci. 29 (2) (Mar. 2018) 359–365, studies of organic molecules, J. Chem. Phys. 56 (5) (1972) 2257–2261.
https://doi.org/10.1007/s10876-018-1338-x. [37] A.M. Alsubaiyel, et al., Adsorption and electronic properties of pristine and Al-
[12] A. Irfan, A. Mahmood, Computational designing of low energy gap small doped C60 fullerenes using N2O molecule: a theoretical study, J. Mol. Liq. 369
molecule acceptors for organic solar cells, Journal of the Mexican Chemical (Jan. 2023), 120855, https://doi.org/10.1016/j.molliq.2022.120855.
Society 61 (4) (2017), https://doi.org/10.29356/jmcs.v61i4.461. [38] V. Sathya, V. Srinivasadesikan, L. Ming-Chang, V. Padmini, Highly sensitive and
[13] J. Stevens, Virtually going green: the role of quantum computational chemistry in selective detection of tryptophan by antipyrine based fluorimetric sensor, J. Mol.
reducing pollution and toxicity in chemistry, Physical Sciences Reviews 2 (7) (Jul. Struct. 1272 (Jan. 2023), 134241, https://doi.org/10.1016/j.
2017), https://doi.org/10.1515/psr-2017-0005. molstruc.2022.134241.
[14] J.J. Villaverde, C. López-Goti, M. Alcamí, A.M. Lamsabhi, J.L. Alonso-Prados, [39] I.A. Udofia, et al., Experimental and theoretical calculation of pKa values of
P. Sandín-España, Quantum chemistry in environmental pesticide risk substituted-2,4,6-trinitrodiphenylamines, J. Mol. Liq. 371 (Feb. 2023), 120926,
assessment, Pest Manag. Sci. 73 (11) (2017) 2199–2202, https://doi.org/ https://doi.org/10.1016/j.molliq.2022.120926.
10.1002/ps.4641. [40] R. kerkour, N. Chafai, O. Moumeni, S. Chafaa, Novel α-aminophosphonate
[15] D. Xia, et al., Potential application of machine-learning-based quantum chemical derivates synthesis, theoretical calculation, Molecular docking, and in silico
methods in environmental chemistry, Environ. Sci. Technol. 56 (4) (Feb. 2022) prediction of potential inhibition of SARS-CoV-2, J. Mol. Struct. 1272 (Jan.
2115–2123, https://doi.org/10.1021/acs.est.1c05970. 2023), 134196, https://doi.org/10.1016/j.molstruc.2022.134196.
[16] R. Kumar Mandal, S. Ghosh, T. Pal Majumder, Comparative study between [41] M. Metin, T. Kawano, T. Okobira, Benchmarking computational chemistry
degradation of dyes (MB, MO) in monotonous and binary solution employing approaches on iminodiacetic acid, J. Indian Chem. Soc. 100 (2) (Jan. 2023),
synthesized bimetallic (Fe-CdO) NPs having antioxidant property, Results in 100895, https://doi.org/10.1016/j.jics.2023.100895.
Chemistry 5 (Jan. 2023), 100788, https://doi.org/10.1016/j. [42] F. Neese, Software update: the ORCA program system—version 5.0, Wiley
rechem.2023.100788. Interdiscip. Rev. Comput. Mol. Sci. 12 (5) (2022) e1606.
[17] M. Tuckerman, Statistical Mechanics: Theory and Molecular Simulation, Oxford [43] J.J. Stewart, MOPAC: a semiempirical molecular orbital program, J. Comput.
university press, 2010. Aided Mol. Des. 4 (1) (1990) 1–103.
[18] M. Hochlaf, Advances in spectroscopy and dynamics of small and medium sized [44] J.M. Turney, et al., Psi4: an open-source ab initio electronic structure program,
molecules and clusters, Phys. Chem. Chem. Phys. 19 (32) (2017) 21236–21261. WIREs Computational Molecular Science 2 (4) (2012) 556–565, https://doi.org/
10.1002/wcms.93.

17
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

[45] R.M. Parrish, et al., Psi4 1.1: an open-source electronic structure program [76] F. Neese, The ORCA program system, WIREs Computational Molecular Science 2
emphasizing automation, advanced libraries, and interoperability, J. Chem. (1) (2012) 73–78, https://doi.org/10.1002/wcms.81.
Theor. Comput. 13 (7) (Jul. 2017) 3185–3197, https://doi.org/10.1021/acs. [77] F. Neese, Software update: the ORCA program system, version 4.0, WIREs
jctc.7b00174. Computational Molecular Science 8 (1) (2018) e1327, https://doi.org/10.1002/
[46] D.G.A. Smith, et al., PSI4 1.4: open-source software for high-throughput quantum wcms.1327.
chemistry, J. Chem. Phys. 152 (18) (May 2020), 184108, https://doi.org/ [78] F. Weigend, Accurate Coulomb-fitting basis sets for H to Rn, Phys. Chem. Chem.
10.1063/5.0006002. Phys. 8 (9) (2006) 1057–1065.
[47] H.-J. Werner, et al., The Molpro quantum chemistry package, J. Chem. Phys. 152 [79] T. Clark, J. Chandrasekhar, G.W. Spitznagel, P.V.R. Schleyer, Efficient diffuse
(14) (Apr. 2020), 144107, https://doi.org/10.1063/5.0005081. function-augmented basis sets for anion calculations. III. The 3-21+ G basis set
[48] H.-J. Werner, P.J. Knowles, G. Knizia, F.R. Manby, M. Schütz, Molpro: a general- for first-row elements, Li–F, J. Comput. Chem. 4 (3) (1983) 294–301.
purpose quantum chemistry program package, WIREs Computational Molecular [80] M.J. Frisch, J.A. Pople, J.S. Binkley, Self-consistent molecular orbital methods 25.
Science 2 (2) (2012) 242–253, https://doi.org/10.1002/wcms.82. Supplementary functions for Gaussian basis sets, J. Chem. Phys. 80 (7) (1984)
[49] G.M.J. Barca, et al., Recent developments in the general atomic and molecular 3265–3269.
electronic structure system, J. Chem. Phys. 152 (15) (Apr. 2020), 154102, [81] R. Krishnan, J.S. Binkley, R. Seeger, J.A. Pople, Self-consistent molecular orbital
https://doi.org/10.1063/5.0005188. methods. XX. A basis set for correlated wave functions, J. Chem. Phys. 72 (1)
[50] M.J. Frisch, et al., Gaussian 16, Rev. C.01, Wallingford, CT, 2016. (1980) 650–654.
[51] T.D. Kühne, et al., CP2K: an electronic structure and molecular dynamics software [82] A. McLean, G. Chandler, Contracted Gaussian basis sets for molecular
package - quickstep: Efficient and accurate electronic structure calculations, calculations. I. Second row atoms, Z= 11–18, J. Chem. Phys. 72 (10) (1980)
J. Chem. Phys. 152 (19) (May 2020), 194103, https://doi.org/10.1063/ 5639–5648.
5.0007045. [83] F. Weigend, R. Ahlrichs, Balanced basis sets of split valence, triple zeta valence
[52] A. Shajan, M. Manathunga, A. Goetz, K. Merz, Geometry Optimization: A and quadruple zeta valence quality for H to Rn: design and assessment of
Comparison of Different Open-Source Geometry Optimizers, 2023. accuracy, Phys. Chem. Chem. Phys. 7 (18) (2005) 3297–3305.
[53] P. Giannozzi, et al., Quantum espresso: a modular and open-source software [84] T.H. Dunning Jr., Gaussian basis sets for use in correlated molecular calculations.
project for quantum simulations of materials, J. Phys. Condens. Matter 21 (39) I. The atoms boron through neon and hydrogen, J. Chem. Phys. 90 (2) (1989)
(Sep. 2009), 395502, https://doi.org/10.1088/0953-8984/21/39/395502. 1007–1023.
[54] P. Giannozzi, et al., Advanced capabilities for materials modelling with Quantum [85] D.E. Woon, T.H. Dunning Jr., Gaussian basis sets for use in correlated molecular
ESPRESSO, J. Phys. Condens. Matter 29 (46) (Oct. 2017), 465901, https://doi. calculations. III. The atoms aluminum through argon, J. Chem. Phys. 98 (2)
org/10.1088/1361-648X/aa8f79. (1993) 1358–1371.
[55] P. Giannozzi, et al., Quantum ESPRESSO toward the exascale, J. Chem. Phys. 152 [86] S. Grimme, J. Antony, S. Ehrlich, H. Krieg, A consistent and accurate ab initio
(15) (Apr. 2020), 154105, https://doi.org/10.1063/5.0005082. parametrization of density functional dispersion correction (DFT-D) for the 94
[56] R. Ahlrichs, M. Bär, M. Häser, H. Horn, C. Kölmel, Electronic structure elements H-Pu, J. Chem. Phys. 132 (15) (2010), 154104.
calculations on workstation computers: the program system turbomole, Chem. [87] S. Grimme, S. Ehrlich, L. Goerigk, Effect of the damping function in dispersion
Phys. Lett. 162 (3) (Oct. 1989) 165–169, https://doi.org/10.1016/0009-2614 corrected density functional theory, J. Comput. Chem. 32 (7) (2011) 1456–1465,
(89)85118-8. https://doi.org/10.1002/jcc.21759.
[57] F. Furche, R. Ahlrichs, C. Hättig, W. Klopper, M. Sierka, F. Weigend, Turbomole,” [88] ORCA input library - ORCA common errors and problems. https://sites.google.
WIREs Computational Molecular Science 4 (2) (2014) 91–100, https://doi.org/ com/site/orcainputlibrary/orca-common-errors-and-problems. (Accessed 24
10.1002/wcms.1162. December 2022).
[58] Y. Shao, et al., Advances in molecular quantum chemistry contained in the Q- [89] CCCBDB Vibrational frequency scaling factors. https://cccbdb.nist.gov/vibnotes.
Chem 4 program package, Mol. Phys. 113 (2) (Jan. 2015) 184–215, https://doi. asp. (Accessed 16 January 2023).
org/10.1080/00268976.2014.952696. [90] CCCBDB vibrational frequency scaling factors. https://cccbdb.nist.gov/vsfx.asp.
[59] G. Kresse, J. Hafner, Ab initio molecular dynamics for liquid metals, Phys. Rev. B (Accessed 16 January 2023).
47 (1) (Jan. 1993) 558–561, https://doi.org/10.1103/PhysRevB.47.558. [91] AIST:Spectral database for organic compounds,SDBS. https://sdbs.db.aist.go.jp
[60] G. Kresse, J. Furthmüller, Efficiency of ab-initio total energy calculations for /sdbs/cgi-bin/cre_index.cgi. (Accessed 24 December 2022).
metals and semiconductors using a plane-wave basis set, Comput. Mater. Sci. 6 [92] F. Teixeira, M.N.D.S. Cordeiro, “Improving vibrational Mode interpretation using
(1) (Jul. 1996) 15–50, https://doi.org/10.1016/0927-0256(96)00008-0. bayesian regression,”, J. Chem. Theor. Comput. 15 (1) (Jan. 2019) 456–470,
[61] G. Kresse, J. Furthmüller, Efficient iterative schemes for ab initio total-energy https://doi.org/10.1021/acs.jctc.8b00439.
calculations using a plane-wave basis set, Phys. Rev. B 54 (16) (Oct. 1996) [93] T.Y. Nikolaienko, L.A. Bulavin, D.M. Hovorun, JANPA: an open source cross-
11169–11186, https://doi.org/10.1103/PhysRevB.54.11169. platform implementation of the Natural Population Analysis on the Java
[62] E. Aprà, et al., NWChem: past, present, and future, J. Chem. Phys. 152 (18) (May platform, Computational and Theoretical Chemistry 1050 (2014) 15–22.
2020), 184102, https://doi.org/10.1063/5.0004997. [94] T.Y. Nikolaienko, L.A. Bulavin, Localized orbitals for optimal decomposition of
[63] R. Dennington, T. Keith, J. Millam, GaussView, version 5 (2009). molecular properties, Int. J. Quant. Chem. 119 (3) (2019), e25798.
[64] M. Karakaya, M. Kürekçi, B. Eskiyurt, Y. Sert, Ç. Çırak, Experimental and [95] M.D. Hanwell, D.E. Curtis, D.C. Lonie, T. Vandermeersch, E. Zurek, G.
computational study on molecular structure and vibrational analysis of an R. Hutchison, Avogadro: an advanced semantic chemical editor, visualization,
antihyperglycemic biomolecule: gliclazide, Spectrochim. Acta Mol. Biomol. and analysis platform, J. Cheminf. 4 (1) (2012) 1–17.
Spectrosc. 135 (Jan. 2015) 137–146, https://doi.org/10.1016/j.saa.2014.06.152. [96] “Avogadro: an open-source molecular builder and visualization tool.” [Online].
[65] B. Peng, J.-R. Wang, X. Mei, Triamterene–furosemide salt: structural aspects and Available: http://avogadro.cc/.
physicochemical evaluation, Acta Crystallogr. B 74 (6) (Dec. 2018), https://doi. [97] H.C. Da Silva, W.B. De Almeida, Theoretical calculations of 1H NMR chemical
org/10.1107/S2052520618013185. Art. no. 6. shifts for nitrogenated compounds in chloroform solution, Chem. Phys. 528 (Jan.
[66] C. Kramer, A. Spinn, K.R. Liedl, Charge anisotropy: where atomic multipoles 2020), 110479, https://doi.org/10.1016/j.chemphys.2019.110479.
matter most, J. Chem. Theor. Comput. 10 (10) (2014) 4488–4496, Oct, https:// [98] G. Żuchowski, K. Zborowski, The influence of solvent molecules on NMR
doi.org/10.1021/ct5005565. spectrum of barbituric acid in the DMSO solution, Cent. Eur. J. Chem. 4 (2006)
[67] A. Spinn, P.H. Handle, J. Kraml, T.S. Hofer, K.R. Liedl, Charge anisotropy of 523–532.
nitrogen: where chemical intuition fails, J. Chem. Theor. Comput. 16 (7) (Jul. [99] Z. Afroz, M. Faizan, M.J. Alam, V.H.N. Rodrigues, S. Ahmad, A. Ahmad,
2020) 4443–4453, https://doi.org/10.1021/acs.jctc.0c00204. Synthesis, structural, hirshfeld surface, spectroscopic studies and quantum
[68] M. Macit, H. Tanak, M. Orbay, N. Özdemir, “Synthesis, crystal structure, chemical calculation of the proton transfer complex between 2-amino-4-hydroxy-
spectroscopic characterization and DFT studies of bis[(1Z,2E)-N-(2,6- 6-methylpyrimidine and salicylic acid, J. Mol. Struct. 1171 (Nov. 2018) 438–448,
diethylphenyl)-N′ -hydroxy-2-(hydroxyimino)acetimidamidato]nickel(II),”, Inorg. https://doi.org/10.1016/j.molstruc.2018.06.020.
Chim. Acta. 459 (Apr. 2017) 36–44, https://doi.org/10.1016/j.ica.2017.01.013. [100] K. Karrouchi, et al., Synthesis, structural, molecular docking and spectroscopic
[69] A. Fatima, et al., Quantum computational, spectroscopic, Hirshfeld surface, studies of (E)-N’-(4-methoxybenzylidene)-5-methyl-1H-pyrazole-3-
electronic state and molecular docking studies on sulfanilic acid: an anti-bacterial carbohydrazide, J. Mol. Struct. 1225 (Feb. 2021), 129072, https://doi.org/
drug, J. Mol. Liq. 346 (Jan. 2022), 117150, https://doi.org/10.1016/j. 10.1016/j.molstruc.2020.129072.
molliq.2021.117150. [101] N. Dege, et al., Quantum computational, spectroscopic investigations on N-(2-((2-
[70] B.K. Sarojini, H.S. Yathirajan, B. Narayana, K. Sunil, M. Bolte, CCDC 657920: chloro-4,5-dicyanophenyl)amino)ethyl)-4-methylbenzenesulfonamide by DFT/
Experimental Crystal Structure Determination, Cambridge Crystallographic Data TD-DFT with different solvents, molecular docking and drug-likeness researches,
Centre, 2008, https://doi.org/10.5517/CCQ2M7M. Colloids Surf. A Physicochem. Eng. Asp. 638 (Apr. 2022), 128311, https://doi.
[71] M. Tutughamiarso, M. Bolte, CCDC 668589: Experimental Crystal Structure org/10.1016/j.colsurfa.2022.128311.
Determination, Cambridge Crystallographic Data Centre, 2008, https://doi.org/ [102] K. Karrouchi, et al., Synthesis, X-ray structure, vibrational spectroscopy, DFT,
10.5517/CCQFQD7. biological evaluation and molecular docking studies of (E)-N’-(4-(dimethylamino)
[72] L.L.C. Schrödinger, “The PyMOL molecular graphics system, version 1.8, https:// benzylidene)-5-methyl-1H-pyrazole-3-carbohydrazide, J. Mol. Struct. 1219 (Nov.
pymol.org, November, 2015. 2020), 128541, https://doi.org/10.1016/j.molstruc.2020.128541.
[73] L.L.C. Schrödinger, The JyMOL Molecular Graphics Development Component, [103] P.K. Murthy, et al., Synthesis, characterization and computational study of the
Nov, 2015, Version 1.8. newly synthetized sulfonamide molecule, J. Mol. Struct. 1153 (Feb. 2018)
[74] L. Schrodinger, The AxPyMOL Molecular Graphics Plugin for Microsoft 212–229, https://doi.org/10.1016/j.molstruc.2017.10.028.
PowerPoint, Schrödinger, LLC, New York, NY, 2015 version 1.8. [104] N. Dege, et al., The synthesis, characterization and theoretical study on nicotinic
[75] A.-R. Allouche, Gabedit—a graphical user interface for computational chemistry acid [1-(2,3-dihydroxyphenyl)methylidene]hydrazide, Spectrochim. Acta Mol.
softwares, J. Comput. Chem. 32 (1) (2011) 174–182.

18
M. Metin et al. Journal of the Indian Chemical Society 100 (2023) 101030

Biomol. Spectrosc. 120 (Feb. 2014) 323–331, https://doi.org/10.1016/j. multipole electrostatics and density-dependent dispersion contributions, J. Chem.
saa.2013.10.030. Theor. Comput. 15 (3) (2019) 1652–1671.
[105] N. Boukabcha, et al., Synthesis, crystal structure, spectroscopic characterization [107] T. Brown, et al., Language models are few-shot learners, Adv. Neural Inf. Process.
and nonlinear optical properties of (Z)-N’-(2,4-dinitrobenzylidene)-2-(quinolin-8- Syst. 33 (2020) 1877–1901.
yloxy) acetohydrazide, J. Mol. Struct. 1194 (Oct. 2019) 112–123, https://doi. [108] My Grammarly - Grammarly. https://app.grammarly.com/. (Accessed 9 April
org/10.1016/j.molstruc.2019.05.074. 2023).
[106] C. Bannwarth, S. Ehlert, S. Grimme, GFN2-xTB—an accurate and broadly
parametrized self-consistent tight-binding quantum chemical method with

19

You might also like