research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logo STRUCTURAL
CHEMISTRY
ISSN: 2053-2296

Exploration of the structure and inter­actions of 4-(di­methyl­amino)-3-methyl­phenyl N-methyl­car­bam­ate (Aminocarb)

crossmark logo

aJan Boeyens Structural Chemistry Laboratory, Molecular Sciences Institute, School of Chemistry, University of the Witwatersrand, Private Bag 3, Johannesburg, 2050, South Africa
*Correspondence e-mail: [email protected], [email protected]

Edited by R. I. Cooper, University of Oxford, United Kingdom (Received 24 February 2025; accepted 26 April 2025; online 13 May 2025)

Amino­carb [4-(di­methyl­amino)-3-methyl­phenyl N-methyl­car­bam­ate, C11H16N2O2], a synthetic pesticide, was crystallized and characterized by single-crystal and powder X-ray diffraction. In the solid state, the mol­ecules have a strong chain N—H⋯O hy­dro­gen bond with a strength of −29.37 kJ mol−1 and a short C—H⋯π contact that build a wave-like three-dimensional structure. The structural stability and inter­molecular inter­action of Amino­carb were investigated using differential scanning calorimetry (DSC) and density functional theory (DFT). The results show that the com­pound is chemically stable, and the two dominating inter­actions are electrostatic and dispersion energies. An electrostatic potential map reveals the binding sites of the mol­ecules for reactivity. The understanding of the structural stability and interactions in Aminocarb provided in this study could be used to design new compounds with improved solubility and bioavailability. The dimethylamino group and the methyl group on the carbamate could be modified with other alkyl groups, which might reduce the Aminocarb toxicity, thereby leading to the development of safe, efficient and cost-effective compounds.

1. Introduction

Amino­carb is a synthetic pesticide in the class of car­bam­ates (Scheme 1[link]). It is one of the most common classes of car­bam­ate that are commonly used for pest control of lepidopterous larvae, aphids, soil mollusks and other types of chewing insects. However, it is highly toxic to humans, animals and the environment (Moreira et al., 2024[Moreira, S., Martins, A. D., Alves, M. G., Pastor, L. M., Seco-Rovira, V., Oliveira, P. F. & Pereira, M. L. (2024). Biology, 13, 721.]; Dias et al., 2014[Dias, E., Morais, S., Ramalheira, E. & Pereira, M. L. (2014). J. Toxicol. Environ. Health A, 77, 849-855.]).

[Scheme 1]

The structure of a mol­ecule and the inter­actions in the solid state are related to the properties of the com­pound. Therefore, insight into the structural characteristics of a com­pound could facilitate the modification of such a com­pound for optimal functionality and the elimination of toxicity. In fact, an understanding of the structural properties of a com­pound is important and has been used in different fields, including materials science, pharmaceuticals and agrochemicals (Choi, 2024[Choi, K. (2024). Mini Rev. Med. Chem. 24, 330-340.]; Biswas et al., 2022[Biswas, K., Gurao, N. P., Maiti, T. & Mishra, R. S. (2022). Structural Properties, in High Entropy Materials: Processing, Properties, and Applications, pp. 195-257. Singapore: Springer Nature Singapore.]), to improve the performance, solubility, bioavailability and stability properties of the material.

Studies have analysed the crystal structure of car­bam­ates to understand the chemical structures and properties (Xu et al., 2005[Xu, L.-Z., Yu, G.-P. & Yang, S.-H. (2005). Acta Cryst. E61, o1924-o1926.]; Wu et al., 2009[Wu, J., Xie, M.-H., Luo, S.-N., Zou, P. & He, Y.-J. (2009). Acta Cryst. E65, o381.]; Xia, 2010[Xia, C.-C. (2010). Acta Cryst. E66, o2580.]), for instance, analysis of the ethyl car­bam­ate crystal structure provides insight into the geometry, noncovalent inter­actions and packing arrangements in the solid state (Bamigboye et al., 2021[Bamigboye, C., Abbo, H. S., Kwong, H. C., Tan, S. L., Tiekink, E. R., Kamounah, F. S. & Titinchi, S. J. (2021). Z. Kristallogr. Cryst. Mater. 236, 187-199.]). Similarly, the crystal structure of the inner salt of 2-[(amino­imino­meth­yl)amino]­ethyl­carbamic acid and its analysis highlighted the geometry and inter­actions in the com­pound (Matulková et al., 2017[Matulková, I., Charvátová, H., Císařová, I. & Štěpnička, P. (2017). Z. Kristallogr. New Cryst. Struct. 232, 685-687.]). However, there is no study on the analysis of the crystal structure and properties of Amino­carb in the literature.

In this article, we will analyse the crystal structure of Amino­carb and its inter­actions to provide insight into the geometry, bonding, packing arrangement and stability of the structure in the solid state. Insecticides with better chemical stability last longer and are more effective in controlling pests and disease vectors. The insight from this study could facilitate the development of effective, safe and less toxic analogues of Amino­carb, thereby reducing the environmental pollution of the com­pound.

2. Materials and methods

2.1. Experimental

2.1.1. Crystal growth experiment

High-purity Amino­carb was purchased from Sigma–Aldrich and utilized as received. Amino­carb (10 mg) was dissolved in acetone (2 ml). The solution was heated and stirred gently on a hotplate at a low tem­per­a­ture of about 30 °C until the com­pound dissolved com­pletely. The heated solution was cooled to room tem­per­a­ture, before the vial was covered with a perforated Parafilm sheet to allow for the slow evaporation of the solvent. Prismatic crystals formed after 3 d and a suitably sized crystal was carefully selected for single-crystal X-ray diffraction (SCXRD) to ascertain the crystal structure.

2.1.2. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 1[link]. All atoms were refined anisotropically before the inclusion of H atoms. C-bound H atoms on aromatic rings were placed in calculated positions, and the sp3-hybridized C-bound H atoms were placed in calculated positions that match the electron-density map. The N-bound H atom was located in the difference Fourier map, and its fractional coordinates and isotropic displacement parameter refined freely.

Table 1
Experimental details

Crystal data
Chemical formula C11H16N2O2
Mr 208.26
Crystal system, space group Monoclinic, P21/c
Temperature (K) 173
a, b, c (Å) 9.2445 (3), 12.4193 (4), 9.9910 (4)
β (°) 98.929 (1)
V3) 1133.17 (7)
Z 4
Radiation type Mo Kα
μ (mm−1) 0.09
Crystal size (mm) 0.29 × 0.28 × 0.15
 
Data collection
Diffractometer Bruker APEXII CCD
Absorption correction Multi-scan (SADABS; Bruker, 2016[Bruker (2016). APEX4 and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.]; Krause et al., 2015[Krause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3-10.])
Tmin, Tmax 0.595, 0.746
No. of measured, independent and observed [I > 2σ(I)] reflections 20029, 2719, 2468
Rint 0.056
(sin θ/λ)max−1) 0.660
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.041, 0.118, 1.06
No. of reflections 2719
No. of parameters 145
H-atom treatment H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.34, −0.27
Computer programs: APEX4 (Bruker, 2016[Bruker (2016). APEX4 and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.]), SAINT (Bruker, 2016[Bruker (2016). APEX4 and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXT2018 (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL2018 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]), OLEX2 (Dolomanov et al., 2009[Dolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339-341.]) and Mercury (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]).
2.1.3. Powder X-ray diffraction (PXRD)

A diffractogram of a powdered crystalline sample of Amino­carb was measured at 293 K using a Bruker D2 phaser powder X-ray diffractometer. The instrument was equipped with a sealed-tube Co Kα1 X-ray source (λ = 1.78896 Å) and a LynxEye PSD detector in Bragg–Brentano geometry, and operated at 30 kV and 10 mA. The data collection was carried out with a scanning inter­val ranging from 2θ = 5.0015 to 40° at a scan speed of 0.5 s per step (with an increment step size of 0.028445°). The X'Pert High Score Plus (Degen & van den Oever, 2009[Degen, T. & van den Oever, J. (2009). Powder Diffr. 24, 163.]) software program was used to analyze the PXRD pattern, and the result was com­pared with the simulated pattern obtained from the SCXRD data to further establish the formation of the crystal structure.

2.1.4. Thermal analysis

Differential scanning calorimetry (DSC) detects the thermal changes in a material relative to a reference by measuring the change in the heat flow of a sample as tem­per­a­ture or time changes (Newman & Wenslow, 2018[Newman, A. & Wenslow, R. (2018). Solid-state Characterization Techniques, in Pharmaceutical Crystals: Science and Engineering, edited by T. Li & A. Mattei, pp. 89-121. Hoboken, New Jersey: John Wiley & Sons Inc.]). Samples are heated and cooled to determine the melting points and enthalpies, and to detect any phase transitions. A Mettler Toledo DSC 3 was used to collect the DSC data, and aluminium pans were placed under nitro­gen gas at a flow rate of 10 ml min−1. The tem­per­a­ture and energy calibrations were performed using pure indium (purity 99.9%, m.p. 156.6 °C, heat of fusion 28.45 J g−1) and pure zinc (purity 99.9%, m.p. 419.5 °C, heat of fusion 112 J g−1). At a heating or cooling rate of 10 °C min−1, the samples were heated from 25 to 155 °C and then cooled to 25 °C.

2.1.5. Cambridge Structural Database analysis

The CSD (Version 5.44; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) was used to analyse the structure of Amino­carb.

2.2. Theoretical studies (using density functional theory, DFT)

2.2.1. Geometry optimization

The GAUSSIAN16 suite of programs (Frisch et al., 2016[Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Mennucci, B., Petersson, G. A., Nakatsuji, H., Caricato, M., Li, X., Hratchian, H. P., Izmaylov, A. F., Bloino, J., Zheng, G., Sonnenberg, J. L., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Montgomery, J. A., Peralta, J. E., Ogliaro, F., Bearpark, M., Heyd, J. J., Brothers, E., Kudin, K. N., Staroverov, V. N., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Rega, N., Millam, J. M., Klene, M., Know, J. E., Cross, J. B., Bakken, V., Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R. E., Yazyev, O. A., Austin, J., Cammi, R., Pomelli, C., Ochterski, J. O., Martin, R. L., Morokuma, K., Zakrzweski, V. G., Voth, G. A., Salvador, P., Dannenberg, J. J., Dapprich, S., Daniels, A. D., Farkas, O., Foresman, J. B., Ortiz, J. V., Cioslowski, J. & Fox, D. J. (2016). GAUSSIAN16. Revision C.01. Gaussian Inc., Wallingford, CT, USA. https://gaussian.com/.]) was used for the minimization of Amino­carb with the default Berny algorithm (Li & Frisch, 2006[Li, X. & Frisch, M. J. (2006). J. Chem. Theory Comput. 2, 835-839.]). Geometry optimization of the Amino­carb structure obtained from the solved crystal structure was calculated in the gas phase at the M06l functional (Wang et al., 2017[Wang, Y., Jin, X., Yu, H. S., Truhlar, D. G. & He, X. (2017). Proc. Natl Acad. Sci. USA, 114, 8487-8492.]; Zhao & Truhlar, 2008a[Zhao, Y. & Truhlar, D. G. (2008a). Acc. Chem. Res. 41, 157-167.]) with the def2-TZVP basis set (Zhao & Truhlar, 2008b[Zhao, Y. & Truhlar, D. G. (2008b). Theor. Chem. Acc. 120, 215-241.]) and incorporate Grimme's D3 dispersion correction (Grimme et al., 2010[Grimme, S., Antony, J., Ehrlich, S. & Krieg, H. (2010). J. Chem. Phys. 132, 154104.]) for a proper description of the dispersion inter­actions. The frequency com­putation was additionally com­puted at the same theoretical level (M06l-D3/def2-TZVP) to ensure no imaginary frequency and to ascertain that the structure that was optimized is a universal minimal structure with the lowest energy. Chemcraft (Zhurko & Zhurko, 2015[Zhurko, G. & Zhurko, D. (2015). Chemcraft. http://www.chemcraftprog.com/.]) software was used in this study to analyze and visualize the output generated from Gaussian calculations. The program extracts the relevant information and presents it in a clear way for easy understanding of the output.

2.2.2. Inter­action energy of units in Amino­carb crystallized structures

The H-atom positions of two mol­ecular structures related by an inversion symmetry operation, obtained from the crystal structures, were optimized separately in the gas phase at the same theoretical (M06l-D3/def2-TZVP) method because the H-atom positions from crystal structures are not accurately determined by X-ray crystallography. This was done to obtain the inter­action energy between the units in the crystal structure of Amino­carb, and it was calculated with the same theoretical method, taking the basis set superposition error (BSSE) into account with the counterpoise correction (Simon et al., 1996[Simon, S., Duran, M. & Dannenberg, J. J. (1996). J. Chem. Phys. 105, 11024-11031.]; Boys & Bernardi, 1970[Boys, S. F. & Bernardi, F. J. M. P. (1970). Mol. Phys. 19, 553-566.]; Ransil, 1961[Ransil, B. J. (1961). J. Chem. Phys. 34, 2109-2118.]). The expression for the inter­action energy is given by

[\Delta E_{\rm cp} = E \left( A \right)_{ab} + E \left( B \right)_{ab} - \bigl( E \left( A \right)_{a} + E \left( B \right)_{b} \bigr)]

where ΔEcp is the energy of com­plexation, E(A)ab is the energy of A calculated in the presence of B ghost basis functions, similar to E(B)ab, E(A)a is the energy of A and E(A)b is the energy of B. The inter­action energy is

[\Delta E = E \left( AB \right)_{ab} - \bigl( E \left( A \right)_{a} + E \left( B \right)_{b} \bigr) + \Delta E_{\rm cp}]

2.2.3. Hirshfeld surfaces and inter­molecular inter­action

CrystalExplorer (Spackman et al., 2021[Spackman, P. R., Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Jayatilaka, D. & Spackman, M. A. (2021). J. Appl. Cryst. 54, 1006-1011.]) was used in this study to generate the Hirshfeld surfaces (HS) at high standard resolution using the CIF as the input file. CrystalExplorer creates colour-coded and HS surface maps that help to visualize the important regions of the inter­molecular inter­actions on the surface. The standard normalized contact (dnorm) of Hirshfeld surface analysis is given as follows:

[d_{\rm norm} = {{d_{\rm i} - r_{\rm i}^{\rm VDW}}\over{r_{\rm i}^{\rm VDW}}} + {{d_{\rm e} - r_{\rm e}^{\rm VDW}}\over{r_{\rm e}^{\rm VDW}}}]

where di is the distance that represents nearest core inside the surface and de is the distance from the HS to the nearest core outside the surface (Dege et al., 2022[Dege, N., Gökce, H., Doğan, O. E., Alpaslan, G., Ağar, T., Muthu, S. & Sert, Y. (2022). Colloids Surf. A Physicochem. Eng. Asp. 638, 128311.]).

The dnorm is the distances range between the surface and the nearest atomic external surfaces (de) and inter­nal surfaces (di). Red contacts are those that are shorter than van der Waals radii (vdW), indicating that the atoms that form inter­molecular bonds are closer than the sum of their radii (Spackman et al., 2008[Spackman, M. A., McKinnon, J. J. & Jayatilaka, D. (2008). CrystEngComm, 10, 377-388.]). Contacts with distances equal to the sum of the vdW radii are shown on the white surface. A blue colour indicates inter­actions that are more distinct – that is, contacts that are longer than the sum of the vdW radii (Dege et al., 2022[Dege, N., Gökce, H., Doğan, O. E., Alpaslan, G., Ağar, T., Muthu, S. & Sert, Y. (2022). Colloids Surf. A Physicochem. Eng. Asp. 638, 128311.]; Zeng et al., 2023[Zeng, W., Wang, X., Kong, X., Li, Y. & Zhang, Y. (2023). J. Mol. Struct. 1279, 135017.]; Garg & Azim, 2022[Garg, U. & Azim, Y. (2022). J. Mol. Struct. 1269, 133820.]; Garg et al., 2021[Garg, U., Azim, Y. & Alam, M. (2021). RSC Adv. 11, 21463-21474.], 2022[Garg, U., Azim, Y., Alam, M., Kar, A. & Pradeep, C. P. (2022). Cryst. Growth Des. 22, 4316-4331.]).

CrystalExplorer was also used in this study to calculate the inter­molecular inter­action energies of the crystal structure. The wavefunctions of the mol­ecular system were calculated with the built-in TONTO program at the CE-B3LYP/6-31G(d,p) theoretical level (Jayatilaka & Grimwood, 2003[Jayatilaka, D. & Grimwood, D. J. (2003). TONTO: a FORTRAN-based object-oriented system for quantum chemistry and crystallography, in International Conference on Computational Science, pp. 142-151. Berlin, Heidelberg: Springer.]; Mackenzie et al., 2017[Mackenzie, C. F., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). IUCrJ, 4, 575-587.]; Turner et al., 2015[Turner, M. J., Thomas, S. P., Shi, M. W., Jayatilaka, D. & Spackman, M. A. (2015). Chem. Commun. 51, 3735-3738.]). All the energies of inter­action between the selected mol­ecule (at the centre of the cluster) and its neighbouring mol­ecules were com­puted; the model then separated the total energies into different com­ponents such as electrostatic, polarization, dispersion and repulsion energy com­ponents.

3. Results and discussion

3.1. Experimental results

3.1.1. Single-crystal X-ray diffraction (SCXRD) results

Amino­carb crystallized in the monoclinic space group P21/c with Z′ = 1 (Fig. 1[link]). Mol­ecules of Amino­carb related by the glide-plane operation form a hy­dro­gen-bonded chain with a C(4) motif (Bernstein et al., 1995[Bernstein, J., Davis, R. E., Shimoni, L. & Chang, N. L. (1995). Angew. Chem. Int. Ed. Engl. 34, 1555-1573.]), using the N1—H1⋯O1i hy­dro­gen bond, as shown in Fig. 2[link](a) and Table 2[link]. Adjacent chains are connected via a short contact from H10B to the ring centroid of an adjacent mol­ecule [C10—H10Bπ, with H10Bπ = 2.90 Å; the centroid (Cg) is at (x, −y + [{1\over 2}], z − [{1\over 2}])]. The Amino­carb mol­ecules form layers that stack on top of one another, giving a wave-like shape in the overall packing when viewed along the a axis, as shown Figs. 2[link](b) and 2(c).

Table 2
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
N1—H1⋯O1i 0.825 (17) 2.083 (17) 2.8174 (12) 148.1 (15)
Symmetry code: (i) [x, -y+{\script{3\over 2}}, z+{\script{1\over 2}}].
[Figure 1]
Figure 1
The mol­ecular structure and asymmetric unit of Amino­carb, with displacement ellipsoids drawn at the 50% probability level and H atoms shown as small spheres of arbitrary radii.
[Figure 2]
Figure 2
(a) The chain hy­dro­gen bond C(4); (b) the C—H⋯π contact joining adjacent chains [symmetry code: (i) x, −y + [{1\over 2}], z − [{1\over 2}]]; (c) mol­ecules packed in opposite directions in a wave-like manner in the overall packing.
3.1.2. PXRD

PXRD is an effective detection tool that can be used to ascertain phase purity in a bulk polycrystalline sample. The overlay of the powder pattern from the PXRD experiment and that simulated from the SCXRD experiment exhibit a shift in peak positions due to the different tem­per­a­tures of the experiments (293 and 173 K, respectively). The PXRD data are not of sufficient quality to conclusively rule out minor impurities, but the strong diffraction peaks are observed at related d-spacings in both patterns. The PXRD result is presented in the supporting information.

3.1.3. Differential scanning calorimetry (DSC)

The DSC thermograms of Amino­carb were obtained to determine the melting point and enthalpy of fusion, and to ascertain the presence of phase changes. The thermal data as given in Fig. 3[link] show an onset of melting point of 94.88 °C, and the enthalpy of melting is −26.46 kJ mol−1. The results indicate that Amino­carb is relatively thermally and chemically stable up to its melting point. However, the com­pound does not show any phase transition.

[Figure 3]
Figure 3
DSC scan of the Amino­carb structure.

3.2. Theoretical results

3.2.1. Full geometry optimization of the Amino­carb structure

The optimized structure of Amino­carb (Fig. 4[link]) calculated with density functional theory (DFT) has a minimum energy value of −1.81 × 106 kJ mol−1. The geometry shows excellent agreement with the X-ray data of Amino­carb.

[Figure 4]
Figure 4
The geometry-optimized structure of Amino­carb. The experimental bond length is shown in sky blue, while the DFT bond length is in green (the stated values are in Ångstroms).
3.2.2. Energetic properties of the structure of Amino­carb

The inter­action energy between two Amino­carb structures related by a "glide plane operation is −7.02 kcal mol−1 (−29.37 kJ mol−1) in the gas phase. This suggests that the chain hy­dro­gen bond between two mol­ecules of Amino­carb, as shown in Fig. 2[link](a), is quite strong and is contributing to the stability of the crystal structure.

3.2.3. Hirshfeld surface (HS) analysis of the Amino­carb structure

To visualize the mol­ecular packing and inter­actions in the crystal structure, the structure HS analysis was generated using CrystalExplorer (Version 21) (Dege et al., 2022[Dege, N., Gökce, H., Doğan, O. E., Alpaslan, G., Ağar, T., Muthu, S. & Sert, Y. (2022). Colloids Surf. A Physicochem. Eng. Asp. 638, 128311.]; Spackman & Jayatilaka, 2009[Spackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19-32.]).

3.2.4. Fingerprint plots (FP)

The percentage contributions of the different atoms in close contact (inter­action) in the crystal structure packing of Amino­carb are given in Fig. 5[link]. The di and de distances – the former representing the distance from the HS to the nearest atom outside and the latter representing the distance from the HS to the nearest atom in the inter­ior – are used to construct the FP (Dege et al., 2022[Dege, N., Gökce, H., Doğan, O. E., Alpaslan, G., Ağar, T., Muthu, S. & Sert, Y. (2022). Colloids Surf. A Physicochem. Eng. Asp. 638, 128311.]).

[Figure 5]
Figure 5
2D fingerprint plots of the Amino­carb structure.

The 2D fingerprint plots of the structure, as well as split into individual elements, are displayed in Fig. 5[link]. The H⋯H inter­actions account for the largest portion of the HS region in the 2D fingerprint maps, representing 62.4%. The O⋯H/H⋯O interactions account for 17.9% of the two sharp spikes. The other contacts are C⋯H/H⋯C (16.7%) and N⋯H/H⋯N (3.1%). These contacts suggest that electrostatic and dispersion inter­actions may be the dominant noncovalent inter­actions in the crystal structure.

3.2.5. Surfaces: dnorm, curvedness, shape index and electrostatic potential

The colour-coded distances representing the different inter­molecular inter­actions of the structure were mapped onto the Hirshfeld surfaces.

(1) The HS is plotted over dnorm in Fig. 6[link]. The intensity of the red spots is a qualitative indicator of the strength of the contacts, as seen in the inter­molecular inter­actions presented. The N—H⋯O hy­dro­gen bonds produce two intense red patches, while the C—H contacts are indicated by a light-red region.

[Figure 6]
Figure 6
Three-dimensional Hirshfeld surface view along the c axis of Amino­carb mapped on dnorm between −0.5196 and 1.3010 a.u.

(2) A deeper understanding of mol­ecular packing was ob­tained by mapping the HS over shape index and curvedness. This was used to gain insight into the weak inter­actions in the crystal packing configuration. There is no indication of stacking inter­actions between mol­ecules of Amino­carb, as shown by the curvedness plot (Fig. 7[link]), which does not display any flat surface area. This is in correlation with the fingerprint plot.

[Figure 7]
Figure 7
Three-dimensional Hirshfeld surface view along the c axis of Amino­carb mapped on curvedness between −4.0000 and 0.4000 a.u.

(3) The absence of red and blue triangles on the shape-index surface of Amino­carb shown in Fig. 8[link] demonstrated that the structure did not have a ππ stacking inter­action, which supported the absence of C⋯C contacts in the fingerprint plot (Spackman & Jayatilaka, 2009[Spackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19-32.]; Garg et al., 2022[Garg, U., Azim, Y., Alam, M., Kar, A. & Pradeep, C. P. (2022). Cryst. Growth Des. 22, 4316-4331.]; Akhileshwari et al., 2022[Akhileshwari, P., Sharanya, K. & Sridhar, M. A. (2022). J. Chem. Crystallogr. 52, 324-336.]).

[Figure 8]
Figure 8
Three-dimensional Hirshfeld surface view along the c axis of Amino­carb mapped on shape index between −1.0000 and 1.0000 a.u.

(4) The preferred binding sites and the electron donors and acceptors were identified and visualized with the use of an electrostatic potential (ESP) map. With the B3LYP/6-311G(d,p) theoretical method, the ESP property was com­puted on the HS surface at high standard resolution, and the results were mapped over the com­puted ESP (Fig. 9[link]) (Mackenzie et al., 2017[Mackenzie, C. F., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). IUCrJ, 4, 575-587.]; Turner et al., 2017[Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). CrystalExplorer17. University of Western Australia. https://crystalexplorer.net/.]). The areas around the atoms that correspond to electropositive and electronegative potentials are shown by blue and red, respectively, as hy­dro­gen-bond donors and acceptors. On the ESP surface, the red region represents electrophilic sites and are electron deficient, while the blue region represents nucleophilic sites and are electron rich. In the Amino­carb structure, the negative potential is around C2=O1, O2 and the C atoms in the arene ring, while the positive potential is concentrated over N1—H1 and N2.

[Figure 9]
Figure 9
Three-dimensional Hirshfeld surface view of Amino­carb mapped on electrostatic potential between −0.0997 and 0.1443 a.u.
3.2.6. Inter­action energy of the crystal structure

The addition of inter­action energy calculations in CrystalExplorer allows for the precise calculation of the intensity of inter­actions, which may be directly com­pared to the outcomes obtained from HS analysis.

The CE-B3LYP/6-31G(d,p) energy model, which is accessible in CrystalExplorer21 (Dege et al., 2022[Dege, N., Gökce, H., Doğan, O. E., Alpaslan, G., Ağar, T., Muthu, S. & Sert, Y. (2022). Colloids Surf. A Physicochem. Eng. Asp. 638, 128311.]), is used to com­pute the inter­molecular inter­action energies. By default, a cluster of mol­ecules is created by applying crystallographic symmetry operations to a chosen central mol­ecule within a radius of 3.8 Å (Frisch et al., 1984[Frisch, M. J., Pople, J. A. & Binkley, J. S. (1984). J. Chem. Phys. 80, 3265-3269.]). The total energies (Etot) are separated into different com­ponents, such as electrostatic (Eele), polarization (Epol), dispersion (Edis) and repulsion (Erep) energies (Hirshfeld, 1977[Hirshfeld, F. L. (1977). Theor. Chim. Acta, 44, 129-138.]), with their respective scale factors being 1.057, 0.740, 0.871 and 0.618 (Fig. 10[link]).

[Figure 10]
Figure 10
The Coulombic inter­action, dispersion and total inter­action energies of the Amino­carb mol­ecule are shown in red, blue, purple, green, yellow and pink along the b axis.

The summation of energy of inter­actions (in kJ mol−1) for the Amino­carb structure (Table 3[link]) are −54.75 (Eele), −11.91 (Epol), −108.53 (Edis), 57.17 (Erep) and −118.03 (Etot) for N—H⋯O. The evaluation of the energy com­ponents shows that the dispersion energy is the highest contributor to the stability of the structure. Electrostatic energy is also strong and contributes about half of the dispersion energy to the stability of the structure.

Table 3
The theoretical inter­action energy of the Amino­carb mol­ecule (in kJ mol−1)

The total of the four energy com­ponents – polarization (Epol), dispersion (Edis), repulsion (Erep) and electrostatic (Eele) energies – is the inter­action energy (Etot). R is the atomic position (distance in Å) between mol­ecular centres.

[Scheme 2]
No. N Symop R Electron density Eele Epol Edis Erep Etot
1 2 x, y, z 9.24 B3LYP/6-31G(d,p) 0.2 −0.1 −1.7 0.0 −1.3
2 2 x, −y + [{1\over 2}], z + [{1\over 2}] 8.20 B3LYP/6-31G(d,p) −5.3 −0.9 −24.2 15.0 −18.1
3 2 x, y + [{1\over 2}], −z + [{1\over 2}] 7.50 B3LYP/6-31G(d,p) −3.0 −1.3 −26.2 14.6 −17.9
4 1 x, −y, −z 6.81 B3LYP/6-31G(d,p) −4.5 −1.8 −23.7 8.6 −21.4
5 2 x, y + [{1\over 2}], −z + [{1\over 2}] 8.14 B3LYP/6-31G(d,p) −6.7 −2.9 −16.5 10.9 −16.9
6 2 x, −y + [{1\over 2}], z + [{1\over 2}] 7.74 B3LYP/6-31G(d,p) −34.2 −8.9 −16.5 37.6 −33.8
7 1 x, −y, −z 6.25 B3LYP/6-31G(d,p) −2.8 −1.0 −14.5 3.9 −13.9
8 1 x, −y, −z 5.90 B3LYP/6-31G(d,p) −1.2 −0.9 −29.9 16.1 −18.1
9 2 x, −y + [{1\over 2}], z + [{1\over 2}] 11.45 B3LYP/6-31G(d,p) 0.3 −0.0 −1.4 0.1 −0.9
10 1 x, −y, −z 13.48 B3LYP/6-31G(d,p) 2.3 −0.3 −8.1 0.0 −4.9

4. Conclusions

Amino­carb crystallized in a monoclinic crystal system and the structure packed with a characteristic wave-like pattern. The geometry of the Amino­carb structure obtained experimentally is in excellent agreement with the theoretically optimized geometry. The chain hy­dro­gen bond in Amino­carb has an inter­action energy of −29.37 kJ mol−1, supporting the importance of this inter­action. DSC analysis shows that the structure is chemically stable, which is in correlation with the calculated inter­action energy. Reactive sites in Amino­carb are identified using a mol­ecular electrostatic potential map. The dominant contacts in the fingerprint plot are O⋯H, C⋯H and H⋯H, confirming the presence of electrostatic and dispersion energy as the dominant inter­action energies. The detailed information of the structural analysis, bonding, stability property and inter­actions energy in Amino­carb provided in this study could be used to modify the structure through functional-group manipulation to improve the properties and perhaps aid the development of safe and effective Amino­carb.

Supporting information


Computing details top

4-(Dimethylamino)-3-methylphenyl methylcarbamate top
Crystal data top
C11H16N2O2 F(000) = 448
Mr = 208.26 Dx = 1.221 Mg m3
Monoclinic, P21/c Mo Kα radiation, λ = 0.71073 Å
a = 9.2445 (3) Å Cell parameters from 9968 reflections
b = 12.4193 (4) Å θ = 2.6–28.3°
c = 9.9910 (4) Å µ = 0.09 mm1
β = 98.929 (1)° T = 173 K
V = 1133.17 (7) Å3 Prism, colourless
Z = 4 0.29 × 0.28 × 0.15 mm
Data collection top
Bruker APEXII CCD
diffractometer
2468 reflections with I > 2σ(I)
φ and ω scans Rint = 0.056
Absorption correction: multi-scan
(SADABS; Bruker, 2016; Krause et al., 2015)
θmax = 28.0°, θmin = 2.2°
Tmin = 0.595, Tmax = 0.746 h = 1212
20029 measured reflections k = 1616
2719 independent reflections l = 1313
Refinement top
Refinement on F2 Hydrogen site location: mixed
Least-squares matrix: full H atoms treated by a mixture of independent and constrained refinement
R[F2 > 2σ(F2)] = 0.041 w = 1/[σ2(Fo2) + (0.0534P)2 + 0.3083P]
where P = (Fo2 + 2Fc2)/3
wR(F2) = 0.118 (Δ/σ)max < 0.001
S = 1.06 Δρmax = 0.34 e Å3
2719 reflections Δρmin = 0.27 e Å3
145 parameters Extinction correction: SHELXL2018 (Sheldrick, 2015b), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4
0 restraints Extinction coefficient: 0.017 (4)
Primary atom site location: dual
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Refinement. Single-crystal X-ray diffraction data were obtained at 173 (2) K on a Bruker D8 Venture Bio PHOTON III 28-pixel array area detector (208 × 128 mm2) diffractometer with a Mo Kα (λ= 0.71073 Å) IµS DIAMOND source (50 kV, 1.4 mA).The APEX4 software package integrated in the instrument was used to control the experiment and analyze the diffraction data. The Bruker SAINT V8.40B (Bruker AXS Inc. Madison, WI, USA)software package in APEX4 was used to refine, reduce, and integrate the crystal data. The multi-scan method implemented in SADABS-2016 (Krause et al., 2015) was used for empirical absorption corrections and correction of other systematic errors. The crystal structure was solved using intrinsic phasing (SHELXT-2018/2) (Sheldrick, 2015) and refined using SHELXL-2018/3 (Sheldrick, 2015) within the Olex2 (Dolomanov et al., 2009) graphical user interface.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
x y z Uiso*/Ueq
C1 0.94691 (15) 0.87750 (9) 0.54104 (12) 0.0383 (3)
H1A 0.925393 0.903606 0.447504 0.057*
H1B 1.053129 0.871112 0.567792 0.057*
H1C 0.907894 0.928398 0.601360 0.057*
C2 0.85428 (11) 0.70693 (8) 0.44563 (10) 0.0242 (2)
C3 0.75368 (12) 0.53609 (8) 0.38075 (10) 0.0276 (2)
C4 0.85833 (11) 0.46956 (9) 0.34131 (10) 0.0288 (2)
H4 0.959144 0.480141 0.375204 0.035*
C5 0.81382 (11) 0.38662 (8) 0.25100 (11) 0.0283 (2)
H5 0.885518 0.340819 0.222498 0.034*
C6 0.66600 (11) 0.36902 (8) 0.20112 (10) 0.0274 (2)
C7 0.56008 (12) 0.43636 (9) 0.24538 (11) 0.0307 (2)
C8 0.60681 (12) 0.52047 (9) 0.33361 (11) 0.0310 (2)
H8 0.536614 0.567796 0.361778 0.037*
C9 0.39819 (13) 0.41537 (12) 0.20829 (15) 0.0462 (3)
H9A 0.361023 0.452308 0.123207 0.069*
H9B 0.347108 0.442386 0.280403 0.069*
H9C 0.381220 0.337770 0.197136 0.069*
C10 0.72656 (15) 0.19696 (10) 0.10624 (15) 0.0433 (3)
H10A 0.806421 0.224555 0.061890 0.065*
H10B 0.679616 0.136030 0.054304 0.065*
H10C 0.765914 0.173188 0.198231 0.065*
C11 0.55224 (18) 0.31373 (13) 0.02376 (14) 0.0530 (4)
H11A 0.484083 0.373363 0.017912 0.080*
H11B 0.498969 0.252460 0.069487 0.080*
H11C 0.628724 0.336801 0.075314 0.080*
N1 0.87965 (11) 0.77349 (8) 0.55057 (9) 0.0310 (2)
N2 0.61929 (11) 0.28174 (8) 0.11263 (10) 0.0340 (2)
O1 0.88204 (9) 0.72368 (6) 0.33288 (7) 0.0330 (2)
O2 0.79138 (9) 0.61438 (6) 0.48184 (7) 0.0331 (2)
H1 0.8600 (16) 0.7530 (13) 0.6242 (17) 0.045 (4)*
Atomic displacement parameters (Å2) top
U11 U22 U33 U12 U13 U23
C1 0.0516 (7) 0.0303 (6) 0.0332 (6) 0.0074 (5) 0.0077 (5) 0.0074 (4)
C2 0.0263 (5) 0.0260 (5) 0.0200 (4) 0.0017 (3) 0.0027 (3) 0.0002 (3)
C3 0.0369 (5) 0.0254 (5) 0.0208 (5) 0.0052 (4) 0.0054 (4) 0.0007 (4)
C4 0.0282 (5) 0.0334 (5) 0.0247 (5) 0.0047 (4) 0.0036 (4) 0.0006 (4)
C5 0.0299 (5) 0.0287 (5) 0.0274 (5) 0.0008 (4) 0.0078 (4) 0.0000 (4)
C6 0.0328 (5) 0.0252 (5) 0.0245 (5) 0.0045 (4) 0.0054 (4) 0.0008 (4)
C7 0.0289 (5) 0.0309 (5) 0.0315 (5) 0.0013 (4) 0.0026 (4) 0.0009 (4)
C8 0.0327 (5) 0.0289 (5) 0.0318 (5) 0.0021 (4) 0.0065 (4) 0.0018 (4)
C9 0.0293 (6) 0.0519 (8) 0.0559 (8) 0.0024 (5) 0.0016 (5) 0.0111 (6)
C10 0.0481 (7) 0.0323 (6) 0.0521 (8) 0.0059 (5) 0.0165 (6) 0.0140 (5)
C11 0.0635 (9) 0.0574 (9) 0.0348 (7) 0.0054 (7) 0.0029 (6) 0.0123 (6)
N1 0.0444 (5) 0.0300 (5) 0.0195 (4) 0.0050 (4) 0.0077 (4) 0.0029 (3)
N2 0.0386 (5) 0.0309 (5) 0.0327 (5) 0.0067 (4) 0.0063 (4) 0.0078 (4)
O1 0.0489 (5) 0.0319 (4) 0.0196 (4) 0.0049 (3) 0.0094 (3) 0.0014 (3)
O2 0.0477 (5) 0.0318 (4) 0.0210 (4) 0.0106 (3) 0.0092 (3) 0.0039 (3)
Geometric parameters (Å, º) top
C1—H1A 0.9800 C7—C8 1.3915 (15)
C1—H1B 0.9800 C7—C9 1.5076 (15)
C1—H1C 0.9800 C8—H8 0.9500
C1—N1 1.4431 (15) C9—H9A 0.9800
C2—N1 1.3271 (13) C9—H9B 0.9800
C2—O1 1.2116 (12) C9—H9C 0.9800
C2—O2 1.3621 (12) C10—H10A 0.9800
C3—C4 1.3755 (15) C10—H10B 0.9800
C3—C8 1.3801 (16) C10—H10C 0.9800
C3—O2 1.4063 (12) C10—N2 1.4546 (16)
C4—H4 0.9500 C11—H11A 0.9800
C4—C5 1.3890 (15) C11—H11B 0.9800
C5—H5 0.9500 C11—H11C 0.9800
C5—C6 1.3972 (15) C11—N2 1.4617 (17)
C6—C7 1.4101 (15) N1—H1 0.825 (17)
C6—N2 1.4228 (13)
H1A—C1—H1B 109.5 C7—C8—H8 119.5
H1A—C1—H1C 109.5 C7—C9—H9A 109.5
H1B—C1—H1C 109.5 C7—C9—H9B 109.5
N1—C1—H1A 109.5 C7—C9—H9C 109.5
N1—C1—H1B 109.5 H9A—C9—H9B 109.5
N1—C1—H1C 109.5 H9A—C9—H9C 109.5
N1—C2—O2 110.25 (9) H9B—C9—H9C 109.5
O1—C2—N1 126.20 (10) H10A—C10—H10B 109.5
O1—C2—O2 123.55 (9) H10A—C10—H10C 109.5
C4—C3—C8 121.08 (10) H10B—C10—H10C 109.5
C4—C3—O2 120.94 (9) N2—C10—H10A 109.5
C8—C3—O2 117.60 (9) N2—C10—H10B 109.5
C3—C4—H4 120.6 N2—C10—H10C 109.5
C3—C4—C5 118.75 (10) H11A—C11—H11B 109.5
C5—C4—H4 120.6 H11A—C11—H11C 109.5
C4—C5—H5 119.3 H11B—C11—H11C 109.5
C4—C5—C6 121.50 (10) N2—C11—H11A 109.5
C6—C5—H5 119.3 N2—C11—H11B 109.5
C5—C6—C7 118.93 (9) N2—C11—H11C 109.5
C5—C6—N2 121.77 (10) C1—N1—H1 119.5 (11)
C7—C6—N2 119.22 (9) C2—N1—C1 122.04 (9)
C6—C7—C9 122.11 (10) C2—N1—H1 118.4 (11)
C8—C7—C6 118.78 (10) C6—N2—C10 115.82 (9)
C8—C7—C9 118.96 (10) C6—N2—C11 114.59 (10)
C3—C8—C7 120.93 (10) C10—N2—C11 110.43 (11)
C3—C8—H8 119.5 C2—O2—C3 117.46 (8)
C3—C4—C5—C6 0.69 (16) C8—C3—C4—C5 0.93 (16)
C4—C3—C8—C7 0.46 (17) C8—C3—O2—C2 108.71 (11)
C4—C3—O2—C2 78.28 (12) C9—C7—C8—C3 173.61 (11)
C4—C5—C6—C7 0.90 (16) N1—C2—O2—C3 178.34 (9)
C4—C5—C6—N2 177.57 (10) N2—C6—C7—C8 178.99 (10)
C5—C6—C7—C8 2.24 (16) N2—C6—C7—C9 3.48 (17)
C5—C6—C7—C9 173.27 (11) O1—C2—N1—C1 0.36 (18)
C5—C6—N2—C10 15.87 (15) O1—C2—O2—C3 1.78 (15)
C5—C6—N2—C11 114.51 (13) O2—C2—N1—C1 179.51 (10)
C6—C7—C8—C3 2.05 (16) O2—C3—C4—C5 173.69 (9)
C7—C6—N2—C10 160.78 (11) O2—C3—C8—C7 172.54 (9)
C7—C6—N2—C11 68.84 (14)
Hydrogen-bond geometry (Å, º) top
D—H···A D—H H···A D···A D—H···A
N1—H1···O1i 0.825 (17) 2.083 (17) 2.8174 (12) 148.1 (15)
Symmetry code: (i) x, y+3/2, z+1/2.
The theoretical interaction energy of the Aminocarb molecule is expressed in kJ mol-1. The total of the four energy components – polarization (Epol), dispersion (Edisp), repulsion (Erep) and electrostatic (Eele) energies – is the interaction energy (Etot). R is the atomic position (distance in Å) between molecular centres. top
No. N Symop R Electron density Eele Epol Edisp Erep Etot
1 2 x, y, z 9.24 B3LYP/6-31G(d,p) 0.2 -0.1 -1.7 0.0 -1.3
2 2 x, -y+1/2, z+1/2 8.20 B3LYP/6-31G(d,p) -5.3 -0.9 -24.2 15.0 -18.1
3 2 -x, y+1/2, -z+1/2 7.50 B3LYP/6-31G(d,p) -3.0 -1.3 -26.2 14.6 -17.9
4 1 -x, -y, -z 6.81 B3LYP/6-31G(d,p) -4.5 -1.8 -23.7 8.6 -21.4
5 2 -x, y+1/2, -z+1/2 8.14 B3LYP/6-31G(d,p) -6.7 -2.9 -16.5 10.9 -16.9
6 2 x, -y+1/2, z+1/2 7.74 B3LYP/6-31G(d,p) -34.2 -8.9 -16.5 37.6 -33.8
7 1 -x, -y, -z 6.25 B3LYP/6-31G(d,p) -2.8 -1.0 -14.5 3.9 -13.9
8 1 -x, -y, -z 5.90 B3LYP/6-31G(d,p) -1.2 -0.9 -29.9 16.1 -18.1
9 2 x, -y+1/2, z+1/2 11.45 B3LYP/6-31G(d,p) 0.3 -0.0 -1.4 0.1 -0.9
10 1 -x, -y, -z 13.48 B3LYP/6-31G(d,p) 2.3 -0.3 -8.1 0.0 -4.9
 

Acknowledgements

Funding for this research was provided by the School of Chemistry and a Postdoctoral grant by the URIC of the University of the Witwatersrand.

References

First citationAkhileshwari, P., Sharanya, K. & Sridhar, M. A. (2022). J. Chem. Crystallogr. 52, 324–336.  Web of Science CSD CrossRef CAS Google Scholar
First citationBamigboye, C., Abbo, H. S., Kwong, H. C., Tan, S. L., Tiekink, E. R., Kamounah, F. S. & Titinchi, S. J. (2021). Z. Kristallogr. Cryst. Mater. 236, 187–199.  Web of Science CSD CrossRef CAS Google Scholar
First citationBernstein, J., Davis, R. E., Shimoni, L. & Chang, N. L. (1995). Angew. Chem. Int. Ed. Engl. 34, 1555–1573.  CrossRef CAS Web of Science Google Scholar
First citationBiswas, K., Gurao, N. P., Maiti, T. & Mishra, R. S. (2022). Structural Properties, in High Entropy Materials: Processing, Properties, and Applications, pp. 195–257. Singapore: Springer Nature Singapore.  Google Scholar
First citationBoys, S. F. & Bernardi, F. J. M. P. (1970). Mol. Phys. 19, 553–566.  CrossRef CAS Web of Science Google Scholar
First citationBruker (2016). APEX4 and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationChoi, K. (2024). Mini Rev. Med. Chem. 24, 330–340.  Web of Science CrossRef CAS PubMed Google Scholar
First citationDege, N., Gökce, H., Doğan, O. E., Alpaslan, G., Ağar, T., Muthu, S. & Sert, Y. (2022). Colloids Surf. A Physicochem. Eng. Asp. 638, 128311.  Web of Science CrossRef Google Scholar
First citationDegen, T. & van den Oever, J. (2009). Powder Diffr. 24, 163.  CrossRef Google Scholar
First citationDias, E., Morais, S., Ramalheira, E. & Pereira, M. L. (2014). J. Toxicol. Environ. Health A, 77, 849–855.  Web of Science CrossRef CAS PubMed Google Scholar
First citationDolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339–341.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationFrisch, M. J., Pople, J. A. & Binkley, J. S. (1984). J. Chem. Phys. 80, 3265–3269.  CrossRef CAS Web of Science Google Scholar
First citationFrisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Mennucci, B., Petersson, G. A., Nakatsuji, H., Caricato, M., Li, X., Hratchian, H. P., Izmaylov, A. F., Bloino, J., Zheng, G., Sonnenberg, J. L., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Montgomery, J. A., Peralta, J. E., Ogliaro, F., Bearpark, M., Heyd, J. J., Brothers, E., Kudin, K. N., Staroverov, V. N., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Rega, N., Millam, J. M., Klene, M., Know, J. E., Cross, J. B., Bakken, V., Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R. E., Yazyev, O. A., Austin, J., Cammi, R., Pomelli, C., Ochterski, J. O., Martin, R. L., Morokuma, K., Zakrzweski, V. G., Voth, G. A., Salvador, P., Dannenberg, J. J., Dapprich, S., Daniels, A. D., Farkas, O., Foresman, J. B., Ortiz, J. V., Cioslowski, J. & Fox, D. J. (2016). GAUSSIAN16. Revision C.01. Gaussian Inc., Wallingford, CT, USA. https://gaussian.com/Google Scholar
First citationGarg, U. & Azim, Y. (2022). J. Mol. Struct. 1269, 133820.  Web of Science CSD CrossRef Google Scholar
First citationGarg, U., Azim, Y. & Alam, M. (2021). RSC Adv. 11, 21463–21474.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationGarg, U., Azim, Y., Alam, M., Kar, A. & Pradeep, C. P. (2022). Cryst. Growth Des. 22, 4316–4331.  Web of Science CSD CrossRef CAS Google Scholar
First citationGrimme, S., Antony, J., Ehrlich, S. & Krieg, H. (2010). J. Chem. Phys. 132, 154104.  Web of Science CrossRef PubMed Google Scholar
First citationGroom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171–179.  Web of Science CrossRef IUCr Journals Google Scholar
First citationHirshfeld, F. L. (1977). Theor. Chim. Acta, 44, 129–138.  CrossRef CAS Web of Science Google Scholar
First citationJayatilaka, D. & Grimwood, D. J. (2003). TONTO: a FORTRAN-based object-oriented system for quantum chemistry and crystallography, in International Conference on Computational Science, pp. 142–151. Berlin, Heidelberg: Springer.  Google Scholar
First citationKrause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3–10.  Web of Science CSD CrossRef ICSD CAS IUCr Journals Google Scholar
First citationLi, X. & Frisch, M. J. (2006). J. Chem. Theory Comput. 2, 835–839.  Web of Science CrossRef CAS PubMed Google Scholar
First citationMackenzie, C. F., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). IUCrJ, 4, 575–587.  Web of Science CrossRef CAS PubMed IUCr Journals Google Scholar
First citationMacrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226–235.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMatulková, I., Charvátová, H., Císařová, I. & Štěpnička, P. (2017). Z. Kristallogr. New Cryst. Struct. 232, 685–687.  Google Scholar
First citationMoreira, S., Martins, A. D., Alves, M. G., Pastor, L. M., Seco-Rovira, V., Oliveira, P. F. & Pereira, M. L. (2024). Biology, 13, 721.  Web of Science CrossRef PubMed Google Scholar
First citationNewman, A. & Wenslow, R. (2018). Solid-state Characterization Techniques, in Pharmaceutical Crystals: Science and Engineering, edited by T. Li & A. Mattei, pp. 89–121. Hoboken, New Jersey: John Wiley & Sons Inc.  Google Scholar
First citationRansil, B. J. (1961). J. Chem. Phys. 34, 2109–2118.  CrossRef CAS Web of Science Google Scholar
First citationSheldrick, G. M. (2015a). Acta Cryst. A71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSheldrick, G. M. (2015b). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSimon, S., Duran, M. & Dannenberg, J. J. (1996). J. Chem. Phys. 105, 11024–11031.  CrossRef CAS Web of Science Google Scholar
First citationSpackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19–32.  Web of Science CrossRef CAS Google Scholar
First citationSpackman, M. A., McKinnon, J. J. & Jayatilaka, D. (2008). CrystEngComm, 10, 377–388.  CAS Google Scholar
First citationSpackman, P. R., Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Jayatilaka, D. & Spackman, M. A. (2021). J. Appl. Cryst. 54, 1006–1011.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationTurner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). CrystalExplorer17. University of Western Australia. https://crystalexplorer.net/Google Scholar
First citationTurner, M. J., Thomas, S. P., Shi, M. W., Jayatilaka, D. & Spackman, M. A. (2015). Chem. Commun. 51, 3735–3738.  Web of Science CrossRef CAS Google Scholar
First citationWang, Y., Jin, X., Yu, H. S., Truhlar, D. G. & He, X. (2017). Proc. Natl Acad. Sci. USA, 114, 8487–8492.  Web of Science CrossRef CAS PubMed Google Scholar
First citationWu, J., Xie, M.-H., Luo, S.-N., Zou, P. & He, Y.-J. (2009). Acta Cryst. E65, o381.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationXia, C.-C. (2010). Acta Cryst. E66, o2580.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationXu, L.-Z., Yu, G.-P. & Yang, S.-H. (2005). Acta Cryst. E61, o1924–o1926.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationZeng, W., Wang, X., Kong, X., Li, Y. & Zhang, Y. (2023). J. Mol. Struct. 1279, 135017.  Web of Science CSD CrossRef Google Scholar
First citationZhao, Y. & Truhlar, D. G. (2008a). Acc. Chem. Res. 41, 157–167.  Web of Science CrossRef PubMed CAS Google Scholar
First citationZhao, Y. & Truhlar, D. G. (2008b). Theor. Chem. Acc. 120, 215–241.  Web of Science CrossRef CAS Google Scholar
First citationZhurko, G. & Zhurko, D. (2015). Chemcraft. http://www.chemcraftprog.com/Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logo STRUCTURAL
CHEMISTRY
ISSN: 2053-2296
Follow Acta Cryst. C
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds