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

Journal logoSTRUCTURAL
CHEMISTRY
ISSN: 2053-2296

The synthesis and crystal structure of (E)-2-{[(4-meth­­oxy­naph­tha­len-1-yl)methyl­­idene]amino}-4-methyl­phenol: Hirshfeld surface analysis, DFT calculations and anti­corrosion studies

crossmark logo

aHigher Normal School of Technological Education of Skikda (ENSET), Algeria, bUnit of Research CHEMS, Chemistry Department, University of Mentouri Brothers, Constantine 1, Algeria, cLaboratory of Electrochemistry, Molecular Engineering and Redox Catalysis (LEIMCR), Department of Basic Education in Technology, Faculty of Technology, University Ferhat Abbas, Setif-1, Algeria, dDepartment of Chemistry, Faculty of Sciences, University of Setif-1, Setif, Algeria, eFaculty of Exact Sciences and Computer Science, University of Jijel, BP 98, Jijel 18000, Algeria, fLaboratory of Analytical Physicochemistry and Crystallochemistry of Organometallic and Biomolecular Materials, University of Constantine 1, 25000, Algeria, gSuperior Normal School of Constantine, University of Constantine 3, 25000, Algeria, hChemistry Department, University of Fribourg, Chemin du Musée 9, CH-1700 Fribourg, Switzerland, and iInstitute of Physics, University of Neuchâtel, rue Emile-Argand 11, 2000 Neuchâtel, Switzerland
*Correspondence e-mail: nadirghichi@yahoo.com

Edited by M. Rosales-Hoz, Cinvestav, Mexico (Received 21 May 2023; accepted 3 July 2023; online 10 July 2023)

This article is part of a collection of articles to commemorate the founding of the African Crystallographic Association and the 75th anniversary of the IUCr.

The title Schiff base com­pound, (E)-2-{[(4-meth­oxy­naph­tha­len-1-yl)methyl­idene]amino}-4-methyl­phenol, C19H17NO2 (I), was synthesized via the reaction of 2-amino-4-methyl­phenol with 4-meth­oxy­naph­tha­lene-1-carbaldehyde. The structure of I was characterized by NMR, IR and UV–Vis spectroscopies in different solvents. The inter­atomic contacts in the crystal structure were explored using Hirshfeld surface analysis, which, together with the two-dimensional fingerprint plots, confirm the predominance of dispersion forces in the crystal structure. The mol­ecule of I has a twisted conformation, with the mean plane of the naph­tha­lene ring system being inclined to the plane of the phenol ring by 33.41 (4)°. In the crystal, mol­ecules are linked by C—H⋯O hydrogen bonds to form inversion dimers. There are parallel-displaced ππ inter­actions present, together with C—H⋯π inter­actions, resulting in the formation of a three-dimensional structure. The anti­corrosion potential of I was also investigated using density functional theory (DFT) in the gas phase and in various solvents. The com­pound was shown to exhibit significant anti­corrosion properties for iron and copper. The mol­ecular structure of I was determined by DFT calculations at the M062X/6-311+g(d) level of theory and com­pared with the crystallographically determined structure. Local and global reactivity descriptors were com­puted to predict the reactivity of I. Excellent agreement was observed between the calculated results and the experimental data.

1. Introduction

Schiff bases, also known as imines, are organic com­pounds containing an azomethine functional group (>C=N–). They are formed by the condensation reaction of derivatives of a primary amine (NH2) and a carbonyl com­pound (C=O), such as an aldehyde or a ketone (Yakan et al., 2023[Yakan, H., Omer, H. S., Buruk, O., Çakmak, Ş., Marah, S., Veyisoğlu, A., Muğlu, H., Ozen, T. & Kütük, H. (2023). J. Mol. Struct. 1277, 134799.]). The first Schiff base com­pounds were synthesized by the Italian scientist Hugo Schiff in 1864 (Qin et al., 2013[Qin, W., Long, S., Panunzio, M. & Biondi, S. (2013). Molecules, 18, 12264-12289.]). They are characterized by the chemical diversity of their structures, i.e. R1R2C=NR3, with R1 and R2 indicating organic side chains, while R3, bound to the N atom, may be an aryl or alkyl group. Due to their inter­esting and important properties, many researchers have explored ways to obtain general, simple, efficient and inexpensive methods for synthesizing Schiff bases. At present, several chemical procedures exist for preparing these com­pounds, including gentle synthetic methods, such as reflux (Brown & Granneman, 1975[Brown, E. V. & Granneman, G. R. (1975). J. Am. Chem. Soc. 97, 621-627.]; Li et al., 2016[Li, C., Yang, W., Zhou, W., Zhang, M., Xue, R., Li, M. & Cheng, Z. (2016). New J. Chem. 40, 8837-8845.]; Olar et al., 2017[Olar, R., Badea, M., Ferbinteanu, M., Stanica, N. & Alan, I. (2017). J. Therm. Anal. Calorim. 127, 709-719.]) or microwave-assisted synthesis (Segura et al., 2016[Segura, J. L., Mancheño, M. J. & Zamora, F. (2016). Chem. Soc. Rev. 45, 5635-5671.]; Sk et al., 2020[Sk, I., Khan, M. A., Haque, A., Ghosh, S., Roy, D., Homechuadhuri, S. & Alam, M. A. (2020). Curr. Opin. Green Sustain. Chem. 3, 100006.]; Mishra et al., 2020[Mishra, N., Yadav, R., Kumar, K., Pandey, H. & Pandey, R. (2020). J. Phys. Conf. Ser. 1504, 012002.]). Schiff bases are of great importance in organic chemistry due to their chemical properties and they have wide-ranging applications. They are also used as ligands in coordination chemistry (Mohamed, 2006[Mohamed, G. G. (2006). Spectrochim. Acta A Mol. Biomol. Spectrosc. 64, 188-195.]; Tarafder et al., 2000[Tarafder, M. T. H., Ali, M. A., Saravanan, N., Weng, W. Y., Kumar, S., Umar-Tsafe, N. & Crouse, K. A. (2000). Transition Met. Chem. 25, 295-298.]; Chandra et al., 2009[Chandra, S., Jain, D., Sharma, A. K. & Sharma, P. (2009). Molecules, 14, 174-190.]), as inter­mediates in organic synthesis (Matsumoto et al., 2020[Matsumoto, Y., Sawamura, J., Murata, Y., Nishikata, T., Yazaki, R. & Ohshima, T. (2020). J. Am. Chem. Soc. 142, 8498-8505.]) and as colorimetric indicators in mineral analysis.

Recent research has been focused on studying the effect of corrosion on metals and alloys, and developing materials and techniques to prevent it (John et al., 2023[John, M., Ralls, M. A., Kuruveri, U. B. & Menezes, P. M. (2023). Metals, 13, 397.]). Corrosion pre­vention relies mainly on the physical and chemical properties of the inhibitor mol­ecule, the π-orbital properties of the electron donor and the electronic structure of the inhibitor (Obot et al., 2009[Obot, I. B., Obi-Egbedi, N. O. & Umoren, S. A. (2009). Corros. Sci. 51, 276-282.]). One effective way of preventing corrosion is the use of organic inhibitors (Zheludkevich et al., 2005[Zheludkevich, M. L., Yasakau, K. A., Poznyak, S. K. & Ferreira, M. G. S. (2005). Corros. Sci. 47, 3368-3383.]; Aljourani et al., 2009[Aljourani, J., Raeissi, K. & Golozar, M. A. (2009). Corros. Sci. 51, 1836-1843.]). Schiff bases can be used to reduce corrosion and minimize negative environmental effects, based on the ease of their synthesis from relatively inexpensive starting materials and their environmentally friendly properties (Lashgari et al., 2010[Lashgari, M., Arshadi, M. R. & Miandari, S. (2010). Electrochim. Acta, 55, 6058-6063.]; Küstü et al., 2007[Küstü, C., Emregül, K. C. & Atakol, O. (2007). Corros. Sci. 49, 2800-2814.]). It is well known that the behaviour of the inhibitors depends on the inter­action between the functional groups in the mol­ecule and the surfaces of the metal. The ability of Schiff bases to form closely packed stable com­plexes in the field of metal ion coordination classifies them as com­pounds capable of pre­ven­ting corrosion. The Schiff base is adsorbed onto the metal surface due to the presence of lone-pair electrons on the N atom in the –CH=N— group and the distribution of π double-bond electrons in the structure. This adsorption behaviour leads to the formation of a very thin layer that covers the surface of the metal, thus preventing corrosion (Afshari et al., 2023[Afshari, F., Ghomi, E. R., Dinari, M. & Ramakrishna, S. (2023). Chem. Select, 8, e202203231.]; Zhang et al., 2023[Zhang, Z., Song, Q., Jin, Y., Feng, Y., Li, J. & Zhang, K. (2023). Metals, 13, 386.]; Jafari et al., 2022[Jafari, H., Ameri, E., Rezaeivala, M. & Berisha, A. (2022). J. Indian Chem. Soc. 99, 100665.]).

[Scheme 1]

In view of this inter­est, we report herein on the synthesis of the title com­pound, namely, (E)-2-{[(4-meth­oxy­naph­tha­len-1-yl)methyl­idene]amino}-4-methyl­phenol (I). The structure of I was fully characterized by crystallographic and spectroscopic techniques. The anti­corrosion efficiency of I was evaluated. Quantum chemical calculations were also carried out for the estimation of the geometrical parameters and reactivity indices of I.

2. Experimental

2.1. Measurements and materials

The reagents used for the synthesis of com­pound I were obtained commercially and were used without further purification. The NMR spectra were recorded on an RMN 400 MHz Bruker Ascend for 1H. IR spectra were recorded on a Shimadzu IRAffinity-1S spectrometer. The UV–Vis spectra were recorded from 200 to 600 nm using a Unicam Helios Alpha spectrophotometer, with a quartz cell having a path length of 1 cm.

2.2. Anti­corrosion experiment

The synthesized com­pound was subjected to an evaluation of its corrosion-inhibiting efficiency of a steel in sulfuric acid medium, using polarization curves and electrochemical im­pe­dance spectroscopy (EIS).

For the experimental conditions, the electrochemical tests were performed in a conventional three-electrode Pyrex glass cell with a capacity of 20 ml using a Gamry Instruments potentiostat/galvanostat/ZRA (Reference 3000), controlled by Gamry framework analysis software (Gamry Echem Analyst, https://www.gamry.com/support-2/software/). The exposed surface area of the working electrode (steel) was 0.18 cm2. Before each test, the working electrode was polished mechanically using emery paper of decreasing grain size (400, 600, 1200 and 2000) under a water jet, then degreased with ethanol, washed with double-distilled water and air-dried. A platinum electrode was used as a counter-electrode (auxiliary electrode), its role being to ensure the passage of electric current in the electrochemical cell. A saturated calomel electrode [Hg/Hg2Cl2/KCl(sat)/ECS] was used as the reference electrode.

Polarization curves are plotted in a potential range from −800 to −200 mV at a scan rate of 1 mV s−1. The impedance measurements were carried out at open-circuit potential (Eocp) using ac signals of 10 mV amplitude (peak to peak) at frequencies between 100 kHz and 0.01 Hz (Mezhoud et al., 2016[Mezhoud, B., Bouchouit, M., Said, M. E., Messaadia, L., Belfaitah, A., Merazig, H., Chibani, A., Bouacida, S. & Bouraiou, A. (2016). Res. Chem. Intermed. 42, 7451.]; Sakki et al., 2022[Sakki, B., Said, M. E., Mezhoud, B., Allal, H., Larbah, L., Kherrouba, A., Chibani, A. & Bouraiou, A. (2022). J. Adh. Sci. Tech. 36, 2245-2268.]).

2.3. Computational details

Density functional theory (DFT) calculations were per­formed for I in the gas phase and in various solvents [n-hex­ane, chloro­form, acetone, methanol, di­chloro­methane (DCM) and water]. All calculations were carried out using GAUSSIAN16 software (Frisch et al., 2019[Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Petersson, G. A., Nakatsuji, H., Li, X., Caricato, M., Marenich, A. V., Bloino, J., Janesko, B. G., Gomperts, R., Mennucci, B., Hratchian, H. P., Ortiz, J. V., Izmaylov, A. F., Sonnenberg, J. L., Williams-Young, D., Ding, F., Lipparini, F., Egidi, F., Goings, J., Peng, B., Petrone, A., Henderson, T., Ranasinghe, D., Zakrzewski, V. G., Gao, J., Rega, N., Zheng, G., Liang, W., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Throssell, K., Montgomery, J. A. Jr, Peralta, J. E., Ogliaro, F., Bearpark, M. J., Heyd, J. J., Brothers, E. N., Kudin, K. N., Staroverov, V. N., Keith, T. A., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A. P., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Millam, J. M., Klene, M., Adamo, C., Cammi, R., Ochterski, J. W., Martin, R. L., Morokuma, K., Farkas, O., Foresman, J. B. & Fox, D. J. (2019). GAUSSIAN16. Revision C.01. Gaussian Inc., Wallingford, CT, USA. https://gaussian.com/.]) without any restriction on the symmetry. The geometry optimization, harmonic frequency calculations and reactivity descriptors were carried out and generated at the M06-2X/6-311+G(d) level of theory. For the solvent-phase calculations, the default self-consistent (solvent) reaction field in GAUSSIAN16 was used, which is the polarizable continuum model (PCM) using the integral equation formalism variant (IEFPCM).

Electronic chemical potential (μ) is the tendency of an atom or mol­ecule to attract electrons; it is given by Equation (1)[link] (see the supporting information for all equations and relations).

The chemical hardness (η) expresses the resistance of a system to change its number of electrons; it is calculated with relation 2.

The global electrophilicity index (ω), introduced by Parr et al. (1999[Parr, R. G., von Szentpály, L. & Liu, S. (1999). J. Am. Chem. Soc. 121, 1922-1924.]), is calculated from the hardness and chemical potential [see Equation (3)].

During the inter­action between two mol­ecular systems, the electrons flow from the lower electronegativity (nucleophile, Nu) to the higher electronegativity (electrophile, E) until the chemical potential becomes equalized. The fraction of the transferred electron, ΔN [see Equation (4)], was estimated according to Pearson (Parr & Pearson, 1983[Parr, R. G. & Pearson, R. G. (1983). J. Am. Chem. Soc. 105, 7512-7516.]; Hannachi et al., 2015[Hannachi, D., Ouddai, N., Arotçaréna, M. & Chermette, H. (2015). Mol. Phys. 113, 1541-1550.]; Fellahi et al., 2021[Fellahi, Z., Chenaf-Ait youcef, H., Hannachi, D., Djedouani, A., Ouksel, L., François, M., Fleutot, S. & Bourzami, R. (2021). J. Mol. Struct. 1244, 130955.]).

On the other hand, the Fukui function f(r) and the dual descriptor Δf(r) are local reactivity descriptors which reflect the ability of a com­pound site to donate or accept electrons. The Fukui function proposed by Parr & Yang (1984[Parr, R. G. & Yang, W. (1984). J. Am. Chem. Soc. 106, 4049-4050.]) can be evaluated for nucleophilic attack [f_{\rm k}^+], electrophilic attack [f_{\rm k}^{-}] and radical (neutral) attack [f_{\rm k}^0] (see Equations 5, 6 and 7).

Furthermore, the local philicity index ([\omega_{\rm k}^{\alpha}]) can easily be evaluated from Equations (8) and (9).

The dual descriptor Δf(r) (Morell et al., 2005[Morell, C., Grand, A. & Toro-Labbé, A. (2005). J. Phys. Chem. A, 109, 205-212.]; Roy et al., 1998[Roy, R. K., Krishnamurti, S., Geerlings, P. & Pal, S. (1998). J. Phys. Chem. A, 102, 3746-3755.]) is more convenient to use than the Fukui function (Chen et al., 2022[Chen, X. M., Li, H. R., Feng, X. L., Wang, H. T. & Sun, X. H. (2022). ACS Omega, 7, 24942-24950.]).

2.4. Synthesis and crystallization

A mixture of 2-amino-4-methyl­phenol (0.246 g) and 4-me­th­oxy­naph­tha­lene-1-carbaldehyde (0.372 g) in methanol (25 ml) was stirred for 1 h. At the end of the reaction, the solvent was evaporated in vacuo. The resulting residue was recrystallized from methanol to give small orange block-like crystals (m.p. 147–149 °C).

2.5. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 1[link]. The OH H atom of I was located in a difference Fourier map and refined freely. The C-bound H atoms were positioned geometrically (C—H = 0.94–0.97 Å) and refined as riding, with Uiso(H) = 1.5Ueq(C) for methyl H atoms and 1.2Ueq(C) otherwise.

Table 1
Experimental details

Crystal data
Chemical formula C19H17NO2
Mr 291.33
Crystal system, space group Triclinic, P[\overline{1}]
Temperature (K) 250
a, b, c (Å) 7.7740 (4), 7.9955 (4), 12.4836 (6)
α, β, γ (°) 106.383 (4), 98.191 (4), 93.279 (4)
V3) 732.96 (7)
Z 2
Radiation type Mo Kα
μ (mm−1) 0.09
Crystal size (mm) 0.61 × 0.39 × 0.18
 
Data collection
Diffractometer STOE IPDS II
Absorption correction Multi-scan (MULABS; Spek, 2020[Spek, A. L. (2020). Acta Cryst. E76, 1-11.])
Tmin, Tmax 0.875, 1.000
No. of measured, independent and observed [I > 2σ(I)] reflections 11018, 3090, 2742
Rint 0.018
(sin θ/λ)max−1) 0.634
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.038, 0.113, 1.04
No. of reflections 3090
No. of parameters 205
H-atom treatment H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.20, −0.14
Computer programs: WinXpose, Recipe, Integrate and LANA in X-AREA (Stoe & Cie, 2022[Stoe & Cie (2022). WinXpose, Recipe, Integrate and LANA in X-AREA. Stoe & Cie GmbH, Darmstadt, Germany.]), SHELXS97 (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]), PLATON (Spek, 2020[Spek, A. L. (2020). Acta Cryst. E76, 1-11.]), 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.]), SHELXL2018 (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]) and publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]).

3. Results and discussion

3.1. Analytical data

3.1.1. IR spectroscopy

In the experimental IR spectrum of I, the C—Har (ar is aromatic) band at 2931–2846 cm−1, the C=C elongation bands at 1512 cm−1 and the deformation domain in the C—Har plane (1000 to 700 cm−1) were of particularly high intensity. The C—Har region can be divided into three subregions: adjacent 1H (1022 to 906 cm−1), adjacent 2H (centred on 819 cm−1) and 4H (centred on 750 cm−1). The IR spectra show the main characteristic imine (C=N) band at 1568 cm−1. A band located at 3348 cm−1 can be attributed to the stretching of the hy­droxy function (–OH) of the phenol group. The peak obtained at an average intensity of 1512 cm−1 can be attributed to the elongation vibration of the C=C aromatic bond. The observed strong peak at 1217 cm−1 can be attributed to the stretching of the C—O—C bond of the meth­oxy ether group (–OCH3) present in the mol­ecule, and the strong peak at 1093 cm−1 can be attributed to the stretching of the C—OH bond in the ring. The band observed around 1350 cm−1 was assigned to CH3 bending. The IR spectra of I are consistent with the structure determined by X-ray diffraction analysis, indicating a high degree of agreement between the two methods.

3.1.2. 1H NMR spectroscopy

Compound I was characterized by its 1H NMR spectrum, which exhibited several diagnostic signals. 1H NMR (400 MHz, CDCl3): δ 2.34 (3H, s, CH3), 4.06 (3H, s, OCH3), 6.93–6.88 (2H, m, H17, H3), 6.99 (1H, dd, J = 8.2, 1.5 Hz, H16), 7.11 (1H, s, H14), 7.55 (1H, ddd, J = 8.1, 6.9, 1.0 Hz, H6), 7.64 (1H, ddd, J = 8.4, 6.9, 1.3 Hz, H7), 8.07 (1H, d, J = 8.2 Hz, H8), 8.35 (1H, d, J = 7.7 Hz, H2), 8.96 (1H, d, J = 8.5 Hz, H5), 9.17 (1H, s, CHN).

For the 1H NMR spectrum of I (see Figs. S9 and 10 in the supporting information), two singlets were observed at δ = 2.34 and 4.06 ppm, corresponding to the protons of methyl groups 19 and 11, respectively. Protons H17 and H3 were observed in the 6.93–6.88 ppm region and exhibit a multiplet pattern. There is a 1H integration doublet of doublets at δ = 6.99 ppm, with coupling constants J = 8.2 and 1.5 Hz, which is characteristic for proton H16. The singlet with a 1H integration at δ = 7.11 ppm was assigned to proton H14. Two 1H integrations as doublet of doublet of doublets were detected at δ = 7.55 and 7.64 ppm, with coupling constants J = 8.1/6.9/1.0 and 8.4/6.9/1.3 Hz, respectively, which is characteristic for protons H6 and H7, respectively. Three 1H integration doublets were detected at δ = 8.07, 8.35 and 8.69 ppm, with coupling constants J = 8.2, 7.7 and 8.5 Hz, respectively, which is characteristic of an ortho coupling for protons H8, H2 and H5, respectively. Finally, a 1H integration singlet was detected at δ = 9.17 ppm, corresponding to the proton of the CH=N group.

3.2. Crystallographic studies

3.2.1. Mol­ecular and crystal structure

The mol­ecular structure of I is illustrated in Fig. 1[link]. A search of the Cambridge Structural Database (CSD, Version 5.43, last update No­vem­ber 2022; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) returned only one similar com­pound, viz. 5-methyl-2-[(naph­tha­len-1-yl­methyl­idene)amino]­phenol (CSD refcode EXIPIL; Orona et al., 2011[Orona, G., Molinar, V., Fronczek, F. R. & Isovitsch, R. (2011). Acta Cryst. E67, o2505-o2506.]).

[Figure 1]
Figure 1
A view of the mol­ecular structure of I, with the atom labelling. The displacement ellipsoids are drawn at the 50% probability level and the intra­molecular O—H⋯N hydrogen bond is shown as a dashed line.

In I, the mean plane of the naph­tha­lene ring system (atoms C1–C10; r.m.s. deviation 0.015 Å) is inclined to the plane of the phenol ring (C13–C18) by 33.41 (4)°. In EXIPIL, the corresponding dihedral angle is only 7.96 (9)°. The meth­oxy group in I lies slightly out of the plane of the ring to which it is attached, with the dihedral angle of CH3—O—Car with respect to the naph­tha­lene ring plane (C1–C4/C9/C10) being 6.08 (12)°. There is an intra­molecular hydrogen bond (O2—H2O⋯N1) involving the phenol –OH group and the amino N atom (Table 2[link]), forming an S(5) ring motif. The con­figuration about the C=N double bond is E and the C12=N1 bond length is 1.2741 (14) Å. This con­figuration is the same as in EXIPIL, where the C=N bond length is 1.273 (3) Å.

Table 2
Hydrogen-bond geometry (Å, °)

Cg1 is the centroid of the phenol ring (atoms C13–C18).

D—H⋯A D—H H⋯A DA D—H⋯A
O2—H2O⋯N1 0.84 (2) 2.069 (18) 2.6436 (13) 126 (2)
C11—H11B⋯O2i 0.97 2.51 3.4725 (19) 170
C19—H19ACg1ii 0.97 2.79 3.7362 (14) 165
Symmetry codes: (i) [-x, -y+2, -z+1]; (ii) [-x-1, -y+1, -z].

In the crystal of I, mol­ecules are linked by C—H⋯O hydrogen bonds to form inversion dimers (Fig. 2[link] and Table 2[link]). As shown in Fig. 2[link], the dimers are linked by parallel-displaced ππ inter­actions involving inversion-related naph­tha­lene rings [blue; the centroid–centroid distance is 3.967 (1) Å, the inter­planar distance is 3.4707 (4) Å and the slippage is 1.922 Å] and inversion-related phenol rings [red; the centroid–centroid distance is 3.732 (1) Å, the inter­planar distance is 3.445 (1) Å and the slippage is 1.437 Å], forming slabs lying parallel to the bc plane. The slabs are linked by a C—H⋯π inter­action (Table 2[link]) and a second ππ inter­action involving inversion-related naph­tha­lene rings [blue; the centroid–centroid distance is 3.819 (1) Å, the inter­planar distance is 3.5302 (4) Å and the slippage is 1.458 Å] to form a three-dimensional (3D) framework.

[Figure 2]
Figure 2
A view along the b axis of the crystal packing of I. The intra­molecular O—H⋯N and inter­molecular C—H⋯O hydrogen bonds are shown as dashed lines, and the C—H⋯π inter­actions as blue dashed arrows (see Table 2[link]). The coloured double-headed arrows indicate the presence of ππ inter­actions (naph­tha­lene rings are shown in blue and phenol rings in red). Only the H atoms involved in these inter­actions have been included.
3.2.2. Hirshfeld surface analysis and two-dimensional (2D) fingerprint plots

The Hirshfeld surface analysis (Spackman & Jayatilaka, 2009[Spackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19-32.]) and the associated 2D fingerprint plots (McKinnon et al., 2007[McKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2007). Chem. Commun. pp. 3814-3816.]) were performed and created with CrystalExplorer21 (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.]), following the protocol of Tiekink and collaborators (Tan et al., 2019[Tan, S. L., Jotani, M. M. & Tiekink, E. R. T. (2019). Acta Cryst. E75, 308-318.]).

The various Hirshfeld surfaces (HS) of I are shown in Fig. 3[link]. There are important contacts present in the crystal; the stronger hydrogen bonds are indicated by the small and large red zones in Fig. 3[link](a). The presence of ππ inter­actions is confirmed by the blue and red triangular shapes in Fig. 3[link](b), and by the flat regions around the naph­tha­lene and phenol rings in Fig. 3[link](c).

[Figure 3]
Figure 3
The Hirshfeld surface of com­pound I mapped over (a) dnorm in the colour range from −0.1552 to 1.2739 a.u., (b) the shape-index property and (c) the curvedness.

The 2D fingerprint plots for I are given in Fig. 4[link]. They reveal that the principal contributions to the overall HS surface involve H⋯H contacts at 54.9% and C⋯H/H⋯C contacts at 19.1%. The latter are the result of the C—H⋯π inter­actions in the crystal. These are followed by the O⋯H/H⋯O and C⋯C contacts at 11.1% each. These are related to the C—H⋯O hydrogen bonds and the various ππ inter­actions present in the crystal (see Table 2[link] and Fig. 3[link]). The N⋯H/H⋯N contacts at 2.7% reflect the presence of an intra­molecular O—H⋯N hydrogen bond. The O⋯C/C⋯O and N⋯C/C⋯N contacts amount to only 0.6 and 0.5%, respectively.

[Figure 4]
Figure 4
The full 2D fingerprint plots for com­pound I and those delineated into H⋯H, C⋯H/H⋯C, N⋯H/H⋯N, O⋯H/H⋯O, C⋯C, N⋯C/C⋯N and O⋯C/C⋯O contacts.
3.2.3. Energy frameworks

Fig. 5[link] shows a com­parison of the energy frameworks calculated for I, including the electrostatic potential forces (Eele), the dispersion forces (Edis) and the total energy diagrams (Etot). The energies were obtained using the wavefunction at the HF/3-2IG level of theory. The cylin­dri­cal radii are proportional to the relative strengths of the corresponding energies (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.]; Tan et al., 2019[Tan, S. L., Jotani, M. M. & Tiekink, E. R. T. (2019). Acta Cryst. E75, 308-318.]). They have been adjusted to the same scale factor of 90 with a cutoff value of 6 kJ mol−1 within a radius of 6 Å of a central reference mol­ecule. As can be seen in Fig. S1 (see supporting information), the inter­atomic inter­actions in the crystal are dominated by dispersion forces (Edis), reflecting the absence of classical inter­molecular hydrogen bonds in the crystal. The colour-coded inter­action mappings within a radius of 6 Å of a central reference mol­ecule and the various contributions to the total energy (Etot) for I are given in Fig. S1.

[Figure 5]
Figure 5
The energy frameworks calculated for I viewed along the b-axis direction, showing the electrostatic potential forces (Eele, red), the dispersion forces (Edis, green) and the total energy diagrams (Etot, blue).

3.3. Anti­corrosion study

The polarization curves and electrochemical impedance spectroscopy (EIS) plots of steel in 0.5 M H2SO4 medium in the absence and presence of the inhibitor I are presented in Figs. 6[link] and 7[link], respectively. It can be seen that in the presence of 5 × 10−3M of I, the anodic and cathodic curve shifts towards the lower values of the current densities, which means that the addition of this com­pound to the corrosive medium reduces the anodic dissolution of the steel and also delays the reduction of hydrogen ions, which may be due to adsorption on the steel surface (Li et al., 2011[Li, X., Deng, S. & Fu, H. (2011). Corros. Sci. 53, 302-309.]; Deng et al., 2011[Deng, S., Li, X. & Fu, H. (2011). Corros. Sci. 53, 3596-3602.]). The diameter of the capacitive loop increases considerably after addition of the inhibitor, indicating the formation of an inhibition film on the steel surface.

[Figure 6]
Figure 6
Polarization curves for steel in 0.5 M H2SO4 solution with and without 5 × 10−3M of inhibitor I at 298 K.
[Figure 7]
Figure 7
EIS for mild steel in 0.5 M H2SO4 solution with and without 5 × 10−3M of inhibitor I at 298 K.

The values of the electrochemical parameters and of the inhibitory efficiency (IE, %) in the absence and in the presence of 5 × 10−3M of the inhibitor obtained by the polarization curves and the electrochemical impedance spectroscopy (EIS) are given in Table 3[link].

Table 3
Polarization curves and electrochemical impedance spectroscopy (EIS)

Method Polarization curves EIS
  Concentration (M) Ecorr (mV) Icorr (µA cm−2) IE (%) Rct (Ω cm2) IE (%)
H2SO4 0.5 432 2794.44 8.08
Inhibitor I 5 × 10−3 498 48.94 98.24 266.22 96.96

The IE was calculated using the following equations:

[{\rm IE} \ \left( \% \right) = I_{\rm corr\left( 0 \right)} - I_{\rm corr}/I_{\rm corr\left( 0 \right)} \times 100 \eqno(1)]

[{\rm IE} \ \left( \% \right) = R_{\rm ct} - R_{\rm ct\left(0 \right)}/R_{\rm ct} \times 100 \eqno(2)]

where Icorr and Icorr(0) are the corrosion current densities in the absence and presence of the inhibitor, respectively, and Rct and Rct(0) are the charge-transfer resistance values with and without inhibitor.

From the parameters shown in Table 3[link], it can be seen that the addition of 5 × 10−3M of com­pound I to the corrosive medium decreases significantly the value of the corrosion cur­rent density (Icorr) and increases the charge-transfer re­sis­­tance. The decrease in Icorr and the increase in load-transfer resistance can be attributed to the adsorption of inhibitor mol­ecules onto the steel surface, thus forming a pro­tective layer. On the other hand, the value of the corrosion potential (Ecorr) moves in a negative direction. This shift relative to H2SO4 alone is less than 85 mV/SCE (about 66 mV). This suggests that I acts as a mixed inhibitor (Musa et al., 2010[Musa, A. Y., Kadhum, A. A. H., Mohamad, A. B. & Takriff, M. S. (2010). Corros. Sci. 52, 3331-3340.]; Döner et al., 2011[Döner, A., Solmaz, R., Özcan, M. & Kardaş, G. (2011). Corros. Sci. 53, 2902-2913.]). The IE values calculated by both methods show good protection of this com­pound against the corrosion of API 5L Grade B steel in 0.5 M sulfuric acid.

3.4. Experimental absorption spectral analysis

The absorption spectra of com­pound I in solvents of different polarities were measured in order to study the effects of solvent polarities on the electronic absorption spectra of the com­pound (see Fig. S2). A series of absorption spectra were recorded in n-hexane, chloro­form, di­chloro­methane (DCM), acetone, methanol and water solutions (at concentrations of the order of 10−4 mol l−1) to study the role of solvent polarity in modifying the electronic states of the com­pound (Fig. S2). Two main absorption bands characterize the absorption spectra of the studied com­pound; the first is attributed to the electronic transitions ππ* and the second to the transition n→π*.

From the absorption spectra, we can see that the use of solvents of different polarities caused differences in the electronic transitions, which led to differences in the absorption spectra. Two main absorption bands characterize the different ab­sorption spectra of our mol­ecule. The first bands around 240 nm were very close to each other for all the solvents except acetone, where we observed a bathochromic shift.

In general, the presence of an enol–keto equilibrium, the nature of the substitutions, the solvent environment, hydrogen bonding, temperature, pH and changes in the dipole moment of the mol­ecules are the key factors in determining the solvatochromism of com­pounds (Zakerhamidi et al., 2012[Zakerhamidi, M. S., Ghanadzadeh, A. & Moghadam, M. (2012). Chem. Sci. Trans. 1, 1-8.]).

3.5. DFT and TD–DFT studies

3.5.1. DFT-optimized geometry

The geometry optimization of I was carried out using the X-ray coordinates and DFT calculations at the M062x/6-311+g(d) level in the gas phase and various solvent environments (see Fig. S3 in the sup­porting information). Table 4[link] includes the equilibrium bond distances and angles, as well as the r.m.s. error (RMSE) values.

Table 4
Structural parameters (Å, °) calculated in the gas phase and different solvents using the M062x/6-311+g(d) level of theory

  Exp. Gas (ɛ = 1) n-Hexane (ɛ = 1.8819) Chloro­form (ɛ = 4.711) DCM (ɛ = 8.93) Acetone (ɛ = 20.493) Methanol (ɛ = 32.613) Water (ɛ = 78.355)
C27—N4 1.410 1.406 1.406 1.406 1.407 1.407 1.407 1.407
N4—C25 1.274 1.275 1.276 1.277 1.277 1.277 1.277 1.277
C25—C5 1.460 1.466 1.466 1.466 1.466 1.466 1.466 1.465
C35—O2 1.370 1.352 1.354 1.356 1.357 1.357 1.358 1.358
O1—C10 1.360 1.350 1.348 1.347 1.346 1.346 1.346 1.346
O2—H3 2.066 2.124 2.125 2.124 2.124 2.123 2.122 2.122
C27—N4—C25 120 120 120 120 120 120 120 120
N4—C25—C5 122 122 122 122 122 122 120 120
C27—N4—C25—C5 178 175 175 175 175 176 176 176
R.m.s. error (RMSE) (Å) 0.368 0.353 0.340 0.335 0.330 0.329 0.328

The optimized structure of com­pound I com­pares well with the experimental data. In particular, the calculated C27—N4, C35—O2 and C10—O1 bond lengths in the gas phase and solvents are smaller than the experimental values (by ap­prox­i­mate­ly −0.003, −0.014 and −0.012 Å, respectively), whereas the N4—C25 and C25—C5 bonds are longer than the experimental values by ca 0.002 and 0.005 Å. Additionally, Fig. S5 presents an overlay of the X-ray crystallographic structure and the optimized geometry of I in water, and the RMSE values were calculated as 0.368, 0.353, 0.340, 0.329, 0.330 and 0.328 Å in the gas phase, n-hexane, chloro­form, methanol, acetone and water, respectively. These results indicate that the RMSE values are lower in polar solvents (water, acetone and methanol) in com­parison to nonpolar solvents. Based on these findings, we can deduce that the calculated geometries, which include bond lengths and angles, are in excellent agreement with the experimental data. On the other hand, it should be noted that the evaluation of the com­puted thermodynamic parameters at 298.15 K in Table 5[link] showed a minimal effect of the solvents on the enthalpy (Ho), Gibbs free energy (Go) and the molar heat capacity (Cv) of the title com­pound.

Table 5
Thermodynamic parameters calculated in the gas phase and different solvents at room temperature using the M062x/6-311+g(d) level of theory

  Gas (ɛ = 0) n-Hexane (ɛ = 1.8819) Chloro­form (ɛ = 4.711) Acetone (ɛ = 20.493) DCM (ɛ = 8.93) Methanol (ɛ = 32.613) Water (ɛ = 78.355)
H (a.u.) −938.949911 −938.955536 −938.959863 −938.962279 −938.962556 −938.962832
G (a.u.) −939.017972 −939.023693 −939.028201 −939.030818 −939.031122 −939.031430
Cv (cal mol−1 K−1) 73.813 73.916 74.012 74.076 74.084 74.092
S (cal mol−1 K−1) 143.245 143.448 143.828 144.252 144.309 144.376
3.5.2. Inhibition mechanism

To investigate the inter­action between the inhibitor mol­ecule and the bulk metal surface (Fe and Cu), we com­puted the global and local reactivity indices; they are listed in Table 6[link], and in Table S1 in the supporting information. Our quantum chemical calculations revealed that the gas phase exhibited the lowest values of chemical hardness, suggesting the weaker stability of I. Conversely, the solvent demonstrated the largest hardness value of I (greater stability). The order of increasing hardness in the studied com­pound was as follows: η(water) = η(acetone) = η(methanol) > η(DCM) > η(chloro­form) > η(n-hexa­ne). On the other hand, the chemical potential of I was found to be greater in the gas phase com­pared to that in the solvent (gas > n-hexane > chloro­form > DCM > acetone > methanol > water). These outcomes suggest that the tendency of the electron to depart from the equilibrium inhibitor com­pound increases with a decrease in the dielectric constant (which increases from the gas to nonpolar and polar environments). In simpler terms, the increase in the μ values signifies that I has a greater inclination to donate electrons in n-hexane than in any other solvent. Based on the global electrophilicity scale (Hannachi et al., 2015[Hannachi, D., Ouddai, N., Arotçaréna, M. & Chermette, H. (2015). Mol. Phys. 113, 1541-1550.], 2021[Hannachi, D., El Houda Amrane, N., Merzoud, L. & Chermette, H. (2021). New J. Chem. 45, 13451-13462.]; Domingo et al., 2002[Domingo, L. R., Aurell, M. J., Pérez, P. & Contreras, R. (2002). Tetrahedron, 58, 4417-4423.]), inhibitor mol­ecule I can be categorized as a strong electrophile, with values ranging from 1.446 to 1.55 eV.

Table 6
Highest occupied mol­ecular orbital (HOMO) energy (ɛHOMO, eV), chemical potential (μ, eV), chemical hardness (η, eV), global electrophilicity index (ω, eV) and fraction of the transferred electron (ΔN, eV)

  Gas n-Hexane Chloro­form DCM Acetone Methanol Water
ɛHOMO −6.806 −6.883 −6.941 −6.959 −6.971 −6.975 −6.978
μ 5.578 5.602 5.612 5.614 5.615 5.615 5.615
η −4.016 −4.082 −4.134 −4.152 −4.164 −4.167 −4.170
ω 1.446 1.487 1.522 1.535 1.543 1.546 1.549
ΔN, I/Fe 0.267 0.260 0.255 0.253 0.252 0.252 0.251
ΔN, I/Cu 0.086 0.080 0.075 0.073 0.072 0.072 0.072

Electronegativity (χ) is an excellent parameter that helps to determine the direction of electron flow between the metal surface and the inhibitor com­pound until a balance in chemical potential is achieved. When the inhibitor mol­ecule is adsorbed on the metal surface, particularly on iron and copper with electronegativities of 7 and 4.9 eV (Michaelson, 1977[Michaelson, H. B. (1977). J. Appl. Phys. 48, 4729-4733.]; Lemoui et al., 2023[Lemoui, R., Allal, H., Hannachi, D., Djedouani, A., Ramli, I., Mohamed el hadi, S., Habila, I., Zabat, M., Merazig, H., Stoeckli-Evans, H. & Ghichi, N. (2023). J. Mol. Struct. 1286, 135569.]; Lesar & Milošev, 2009[Lesar, A. & Milošev, I. (2009). Chem. Phys. Lett. 483, 198-203.]), respectively, electrons should move in the system from the less electronegative to the more electronegative element. Our calculations reveal that I exhibits less electronegativity than iron and copper, indicating that it is capable of transferring electrons to the metal surface and the inhibitor will thus exhibit better kinetic inter­action with the iron surface com­pared to the cop­per surface (Lemoui et al., 2023[Lemoui, R., Allal, H., Hannachi, D., Djedouani, A., Ramli, I., Mohamed el hadi, S., Habila, I., Zabat, M., Merazig, H., Stoeckli-Evans, H. & Ghichi, N. (2023). J. Mol. Struct. 1286, 135569.]).

ΔN, the number of transferred electrons, is a useful quantum chemical descriptor for investigating and studying the inter­actions between a metal and an inhibitor. This de­scrip­tor can provide valuable insights into the mechanism of inhibition and help predict the inhibitory activity of new com­pounds. The results given in Table 6[link] revealed that the ΔN value of I/Fe was three times greater than that of I/Cu, suggesting a stronger inter­action between the corrosion inhibitor and the iron surface in both the gas and the solvent phases com­pared to the inter­action between the inhibitor and copper. Additionally, in gas and nonpolar media (n-hexane, chloro­form and DCM), com­pound I exhibited a slightly higher ΔN value among the tested solvents, indicating a slightly greater potential for releasing electrons into the low-lying vacant d orbitals of the metal (Lemoui et al., 2023[Lemoui, R., Allal, H., Hannachi, D., Djedouani, A., Ramli, I., Mohamed el hadi, S., Habila, I., Zabat, M., Merazig, H., Stoeckli-Evans, H. & Ghichi, N. (2023). J. Mol. Struct. 1286, 135569.]; Alaoui Mrani et al., 2021[Alaoui Mrani, S., Ech-chihbi, E., Arrousse, N., Rais, Z., El Hajjaji, F., El Abiad, C., Radi, S., Mabrouki, J., Taleb, M. & Jodeh, S. (2021). Arab. J. Sci. Eng. 46, 5691-5707.]) relative to acetone, methanol and water. This observation suggests that com­pound I may be a promising inhibitor for preventing corrosion in non-aqueous environments and provides insight into the factors that contribute to its inhibitory activity. By understanding the electronic structure and the inter­action of the inhibitor with the metal surface, we can develop new and more effective corrosion inhibitors for various industrial applications.

On the other hand, a corrosion inhibitor with weak electrophilicity (ω) exhibits a good potential for releasing electrons, which leads to efficient inter­action and stabilization on metal surfaces. The trend of electrophilicity (ω) is as follows: water > methanol > acetone > DCM > chloro­form > n-hexane > gas. The results indicate that gas, n-hexane, chloro­form and DCM (nonpolar media) showed a smaller value of ω than the polar solvent, suggesting that these media contribute significantly to the corrosion inhibition potential.

The sites that exhibit the highest values of condensed Fukui functions (FFs) f and f+ correspond to those suitable for electrophilic and nucleophilic attacks, respectively. Table 7[link] shows the values of the Fukui function and a dual descriptor for I after nucleophilic and electrophilic attack. DFT calculations show that the local reactivity descriptor has a larger value in polar solvents than in nonpolar solvents. Furthermore, an analysis of these results shows that atoms N4 and C25 have the highest values of [f_{\rm k}^+], suggesting that these sites act as electron acceptors. On the other hand, atoms O2 and C5 have large values of [f_{\rm k}^{-}], indicating that they act as electron donors. The local philicity index and dual descriptor Δf(r) (see Fig. S6) exhibit a remarkable agreement with the FF, and it is found that atom C5 is the pre­ferred site for electrophilic attack.

Table 7
Local reactivity descriptors calculated for the corrosion inhibitor I in the gas phase and in various solvents [values in parentheses represent Δf(r) > 0 and Δf(r) < 0]

  Gas n-Hexane Chloro­form DCM Acetone Methanol Water
  f Δf f Δf f Δf f Δf f Δf f Δf f Δf
O1 0.0443 −0.0204 0.0447 −0.0199 0.0484 −0.0236 0.0532 −0.0285 0.0579 −0.0334 0.0594 −0.035 0.0609 −0.0365
O2 0.0754 −0.0469 0.0697 −0.0449 0.0589 −0.037 0.0501 −0.0291 0.0424 −0.022 0.0401 −0.0199 0.0379 −0.0179
N4 0.0353 0.0859 0.0485 0.0781 0.0652 0.0672 0.075 0.0598 0.0821 0.0544 0.084 0.0529 0.0856 0.0518
C5 0.0624 −0.0487 0.073 −0.0597 0.0938 (−0.0816) 0.1119 −0.1001 0.1284 (−0.117) 0.1344 −0.1242 0.1392 −0.1292
C25 0.0116 0.0977 0.0084 0.1178 −0.0009 0.1416 −0.01 0.1387 −0.018 (0.135) −0.0199 0.134 −0.0221 0.1332
C27 0.0613 −0.0242 0.0588 −0.027 0.048 −0.021 0.0364 −0.0113 0.0258 −0.0019 0.0228 0.0008 0.0199 0.0033
3.5.3. TD–DFT calculations of absorption spectra

The electronic spectral analysis of the title com­pound was per­formed using the time-dependent DFT (TD–DFT) method and the IEFPCM model in gas and various polar and nonpolar solvent phases. The calculations were carried out at the TD-M062x/6-311+g(d) level of theory on the ground-state-optimized geometry. The com­puted absorption spectrum, wavelength (nm), excitation energy (eV) and oscillator strengths (fos), as well as the absorption properties of com­pound I, are presented in Table 7[link] and Figs. S3 and S4 of the supporting information.

The simulated UV–Vis spectrum of I (Fig. S7) contains two regions of high oscillator strengths falling in the ranges 180–280 and 280–450 nm. From the TD–DFT outcomes it is apparent that the solvent does not have a significant impact on the UV–Vis spectrum. Conversely, when com­paring the com­puted and experimental spectra, a shift towards higher wavelength can be observed for the experimental results, and the discrepancy between them is minimal. On the other hand, we can observe that the most intense band centred at about 351 and 345 nm (f = 0.8 and 0.6) for the solvent and gas phases, respectively, is caused by a HOMO→LUMO electronic transition (see Table 7[link]). Analysis of the frontier orbitals (Fig. S8) shows that this transition has a mixed character of ππ* and n→π* (n in N and O atoms). Furthermore, a moderate absorption band was observed at ∼231 nm for the solvent and at 226 nm (f = 0.44) in the gas phase, which cor­responds to the electronic transition from HOMO-3 to LUMO having a ππ* character, where π is localized on fragment I (see Fig. S3 in the supporting information). The absorption bands observed at 296 nm in acetone and methanol are a result of the HOMO-1→LUMO transitions with n→π*/ππ* character, and these transitions exhibit lower oscillator strengths with a value of f = 0.004, whereas the band at 199 nm is due to a HOMO-1→LUMO+2 transition with n→π*/ππ* character (where the π* is localized in fragment I).

4. Conclusions

The title Schiff base, (E)-2-{[(4-meth­oxy­naph­tha­len-1-yl)methyl­idene]amino}-4-methyl­phenol (I) was synthesized and its structure characterized by X-ray diffraction analysis and by IR, 1H NMR and UV–Vis spectroscopies. All the major inter­atomic and inter­molecular inter­actions have been dis­cussed and explained. The Hirshfeld surface analysis was car­ried out and indicated the dominance of the H⋯H (54.9%) and C⋯H/H⋯C (19.1%) inter­atomic contacts in the crystal lattice, which could be helpful for future drug design. The high chemical hardness and chemical potential as an anti­corrosive agent emphasizes the possible use of I in the metals industry, and it has been shown to be a promising anti­­corrosive agent. The results obtained by both methods show that com­pound I could serve as an effective corrosion inhibitor of API 5L Grade B steel in 0.5 M H2SO4. The corrosion inhibition potentials of I were investigated using quantum chemical calculations at the ω B97XD level with the 6-311+g(d) basis set in the gas phase and in various solvents.

Supporting information


Computing details top

Data collection: WinXpose in X-AREA (Stoe & Cie, 2022); cell refinement: Recipe in X-AREA (Stoe & Cie, 2022); data reduction: Integrate and LANA in X-AREA (Stoe & Cie, 2022); program(s) used to solve structure: SHELXS97 (Sheldrick, 2008); program(s) used to refine structure: SHELXL2018 (Sheldrick, 2015); molecular graphics: PLATON (Spek, 2020) and Mercury (Macrae et al., 2020); software used to prepare material for publication: SHELXL2018 (Sheldrick, 2015), PLATON (Spek, 2020) and publCIF (Westrip, 2010).

(E)-2-{[(4-Methoxynaphthalen-1-yl)methylidene]amino}-4-methylphenol top
Crystal data top
C19H17NO2Z = 2
Mr = 291.33F(000) = 308
Triclinic, P1Dx = 1.320 Mg m3
a = 7.7740 (4) ÅMo Kα radiation, λ = 0.71073 Å
b = 7.9955 (4) ÅCell parameters from 11095 reflections
c = 12.4836 (6) Åθ = 1.7–27.3°
α = 106.383 (4)°µ = 0.09 mm1
β = 98.191 (4)°T = 250 K
γ = 93.279 (4)°Block, orange
V = 732.96 (7) Å30.61 × 0.39 × 0.18 mm
Data collection top
STOE IPDS II
diffractometer
3090 independent reflections
Radiation source: sealed X-ray tube, 12 x 0.4 mm long-fine focus2742 reflections with I > 2σ(I)
Plane graphite monochromatorRint = 0.018
Detector resolution: 6.67 pixels mm-1θmax = 26.8°, θmin = 1.7°
rotation method, ω scansh = 99
Absorption correction: multi-scan
(MULABS; Spek, 2020)
k = 109
Tmin = 0.875, Tmax = 1.000l = 1515
11018 measured reflections
Refinement top
Refinement on F2Primary atom site location: structure-invariant direct methods
Least-squares matrix: fullSecondary atom site location: difference Fourier map
R[F2 > 2σ(F2)] = 0.038Hydrogen site location: mixed
wR(F2) = 0.113H atoms treated by a mixture of independent and constrained refinement
S = 1.04 w = 1/[σ2(Fo2) + (0.0629P)2 + 0.1103P]
where P = (Fo2 + 2Fc2)/3
3090 reflections(Δ/σ)max = 0.005
205 parametersΔρmax = 0.20 e Å3
0 restraintsΔρmin = 0.14 e Å3
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.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
O10.37640 (11)0.70357 (10)0.74315 (6)0.0478 (2)
O20.00186 (12)0.95474 (11)0.15984 (8)0.0533 (2)
H2O0.031 (2)0.935 (2)0.2219 (16)0.084 (6)*
N10.00350 (12)0.69499 (12)0.25196 (8)0.0417 (2)
C10.13306 (13)0.60555 (14)0.40839 (9)0.0361 (2)
C20.13789 (14)0.76848 (14)0.48464 (9)0.0402 (2)
H20.0850810.8567810.4597970.048*
C30.21806 (15)0.80813 (14)0.59750 (9)0.0415 (2)
H30.2192260.9212330.6471900.050*
C40.29521 (13)0.68039 (14)0.63528 (9)0.0379 (2)
C50.37175 (15)0.37629 (15)0.59984 (10)0.0431 (3)
H50.4237520.4017510.6761410.052*
C60.37224 (16)0.21274 (15)0.52838 (11)0.0478 (3)
H60.4230910.1255440.5555050.057*
C70.29682 (16)0.17444 (14)0.41423 (10)0.0465 (3)
H70.2990560.0617740.3648240.056*
C80.22010 (14)0.29895 (14)0.37397 (9)0.0413 (2)
H80.1703680.2705570.2970800.050*
C90.21401 (12)0.47005 (13)0.44580 (8)0.0349 (2)
C100.29405 (13)0.50852 (13)0.56089 (9)0.0360 (2)
C110.3970 (2)0.87468 (17)0.81915 (11)0.0625 (4)
H11C0.4639710.8747520.8909430.094*
H11B0.2830630.9126360.8308270.094*
H11A0.4582150.9541180.7878350.094*
C120.04229 (13)0.57172 (14)0.29245 (9)0.0387 (2)
H120.0168170.4552040.2461660.046*
C130.09569 (13)0.65748 (14)0.14067 (9)0.0383 (2)
C140.18898 (14)0.49935 (14)0.07476 (9)0.0414 (3)
H140.1947060.4042290.1048540.050*
C150.27369 (14)0.47931 (15)0.03449 (9)0.0414 (3)
C160.26515 (15)0.62281 (16)0.07652 (9)0.0447 (3)
H160.3223710.6114320.1501970.054*
C170.17448 (16)0.78190 (15)0.01252 (10)0.0459 (3)
H170.1704580.8773610.0423770.055*
C180.09002 (14)0.79882 (14)0.09561 (9)0.0407 (2)
C190.37175 (17)0.30697 (16)0.10504 (10)0.0523 (3)
H19A0.4953640.3108290.1015430.078*
H19B0.3291540.2137800.0761310.078*
H19C0.3540110.2850870.1829620.078*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
O10.0577 (5)0.0459 (4)0.0346 (4)0.0057 (4)0.0041 (3)0.0091 (3)
O20.0640 (6)0.0426 (5)0.0495 (5)0.0021 (4)0.0053 (4)0.0161 (4)
N10.0428 (5)0.0459 (5)0.0368 (5)0.0026 (4)0.0009 (4)0.0157 (4)
C10.0341 (5)0.0402 (5)0.0353 (5)0.0018 (4)0.0042 (4)0.0143 (4)
C20.0427 (6)0.0391 (5)0.0410 (6)0.0070 (4)0.0045 (4)0.0158 (4)
C30.0470 (6)0.0367 (5)0.0388 (5)0.0047 (4)0.0051 (4)0.0088 (4)
C40.0370 (5)0.0415 (5)0.0340 (5)0.0002 (4)0.0022 (4)0.0118 (4)
C50.0430 (6)0.0456 (6)0.0419 (6)0.0045 (5)0.0002 (4)0.0180 (5)
C60.0490 (6)0.0416 (6)0.0560 (7)0.0091 (5)0.0039 (5)0.0211 (5)
C70.0507 (6)0.0353 (5)0.0511 (6)0.0053 (5)0.0068 (5)0.0096 (5)
C80.0428 (6)0.0405 (5)0.0386 (5)0.0012 (4)0.0037 (4)0.0106 (4)
C90.0322 (5)0.0364 (5)0.0368 (5)0.0004 (4)0.0049 (4)0.0130 (4)
C100.0330 (5)0.0379 (5)0.0373 (5)0.0005 (4)0.0037 (4)0.0131 (4)
C110.0786 (9)0.0528 (7)0.0423 (6)0.0112 (6)0.0096 (6)0.0007 (5)
C120.0374 (5)0.0426 (6)0.0364 (5)0.0039 (4)0.0041 (4)0.0130 (4)
C130.0369 (5)0.0440 (6)0.0350 (5)0.0064 (4)0.0037 (4)0.0142 (4)
C140.0411 (6)0.0429 (6)0.0415 (6)0.0040 (4)0.0030 (4)0.0165 (4)
C150.0377 (5)0.0460 (6)0.0393 (5)0.0068 (4)0.0044 (4)0.0108 (4)
C160.0452 (6)0.0546 (7)0.0345 (5)0.0108 (5)0.0021 (4)0.0146 (5)
C170.0519 (6)0.0470 (6)0.0435 (6)0.0098 (5)0.0046 (5)0.0214 (5)
C180.0413 (5)0.0402 (5)0.0408 (6)0.0060 (4)0.0041 (4)0.0133 (4)
C190.0507 (7)0.0514 (7)0.0477 (6)0.0010 (5)0.0011 (5)0.0087 (5)
Geometric parameters (Å, º) top
O1—C41.3601 (12)C8—C91.4161 (15)
O1—C111.4131 (14)C8—H80.9400
O2—C181.3705 (14)C9—C101.4212 (14)
O2—H2O0.835 (18)C11—H11C0.9700
N1—C121.2741 (14)C11—H11B0.9700
N1—C131.4109 (13)C11—H11A0.9700
C1—C21.3745 (15)C12—H120.9400
C1—C91.4379 (14)C13—C141.3927 (15)
C1—C121.4603 (14)C13—C181.3983 (15)
C2—C31.3961 (15)C14—C151.3886 (15)
C2—H20.9400C14—H140.9400
C3—C41.3735 (15)C15—C161.3917 (16)
C3—H30.9400C15—C191.5031 (16)
C4—C101.4238 (15)C16—C171.3845 (16)
C5—C61.3599 (17)C16—H160.9400
C5—C101.4137 (15)C17—C181.3805 (15)
C5—H50.9400C17—H170.9400
C6—C71.4012 (16)C19—H19A0.9700
C6—H60.9400C19—H19B0.9700
C7—C81.3653 (16)C19—H19C0.9700
C7—H70.9400
C4—O1—C11117.78 (9)O1—C11—H11C109.5
C18—O2—H2O102.0 (12)O1—C11—H11B109.5
C12—N1—C13120.69 (9)H11C—C11—H11B109.5
C2—C1—C9118.94 (9)O1—C11—H11A109.5
C2—C1—C12119.73 (9)H11C—C11—H11A109.5
C9—C1—C12121.31 (9)H11B—C11—H11A109.5
C1—C2—C3122.66 (10)N1—C12—C1122.15 (10)
C1—C2—H2118.7N1—C12—H12118.9
C3—C2—H2118.7C1—C12—H12118.9
C4—C3—C2119.40 (10)C14—C13—C18118.84 (10)
C4—C3—H3120.3C14—C13—N1127.09 (9)
C2—C3—H3120.3C18—C13—N1114.07 (9)
O1—C4—C3124.60 (10)C15—C14—C13121.35 (10)
O1—C4—C10114.58 (9)C15—C14—H14119.3
C3—C4—C10120.82 (9)C13—C14—H14119.3
C6—C5—C10120.86 (10)C14—C15—C16118.20 (10)
C6—C5—H5119.6C14—C15—C19120.66 (10)
C10—C5—H5119.6C16—C15—C19121.14 (10)
C5—C6—C7119.88 (10)C17—C16—C15121.64 (10)
C5—C6—H6120.1C17—C16—H16119.2
C7—C6—H6120.1C15—C16—H16119.2
C8—C7—C6120.73 (10)C18—C17—C16119.29 (10)
C8—C7—H7119.6C18—C17—H17120.4
C6—C7—H7119.6C16—C17—H17120.4
C7—C8—C9121.27 (10)O2—C18—C17120.14 (10)
C7—C8—H8119.4O2—C18—C13119.18 (10)
C9—C8—H8119.4C17—C18—C13120.67 (10)
C8—C9—C10117.52 (9)C15—C19—H19A109.5
C8—C9—C1123.66 (9)C15—C19—H19B109.5
C10—C9—C1118.82 (9)H19A—C19—H19B109.5
C5—C10—C9119.71 (10)C15—C19—H19C109.5
C5—C10—C4120.94 (10)H19A—C19—H19C109.5
C9—C10—C4119.35 (9)H19B—C19—H19C109.5
C9—C1—C2—C30.17 (16)O1—C4—C10—C50.35 (15)
C12—C1—C2—C3178.14 (9)C3—C4—C10—C5179.33 (10)
C1—C2—C3—C40.26 (17)O1—C4—C10—C9179.44 (8)
C11—O1—C4—C36.08 (17)C3—C4—C10—C90.89 (16)
C11—O1—C4—C10174.26 (11)C13—N1—C12—C1178.08 (9)
C2—C3—C4—O1179.75 (10)C2—C1—C12—N114.20 (16)
C2—C3—C4—C100.11 (17)C9—C1—C12—N1167.54 (10)
C10—C5—C6—C70.76 (18)C12—N1—C13—C1419.32 (17)
C5—C6—C7—C81.01 (18)C12—N1—C13—C18161.18 (10)
C6—C7—C8—C90.06 (18)C18—C13—C14—C150.86 (16)
C7—C8—C9—C101.32 (16)N1—C13—C14—C15179.66 (10)
C7—C8—C9—C1179.17 (10)C13—C14—C15—C160.84 (16)
C2—C1—C9—C8178.57 (10)C13—C14—C15—C19179.16 (10)
C12—C1—C9—C83.15 (16)C14—C15—C16—C170.33 (17)
C2—C1—C9—C100.94 (15)C19—C15—C16—C17179.66 (11)
C12—C1—C9—C10177.34 (9)C15—C16—C17—C180.14 (18)
C6—C5—C10—C90.53 (17)C16—C17—C18—O2179.35 (10)
C6—C5—C10—C4179.26 (10)C16—C17—C18—C130.12 (17)
C8—C9—C10—C51.54 (15)C14—C13—C18—O2179.84 (10)
C1—C9—C10—C5178.92 (9)N1—C13—C18—O20.61 (15)
C8—C9—C10—C4178.25 (9)C14—C13—C18—C170.37 (17)
C1—C9—C10—C41.29 (14)N1—C13—C18—C17179.92 (10)
Hydrogen-bond geometry (Å, º) top
Cg1 is the centroid of the phenol ring, C13-C18.
D—H···AD—HH···AD···AD—H···A
O2—H2O···N10.84 (2)2.069 (18)2.6436 (13)126 (2)
C11—H11B···O2i0.972.513.4725 (19)170
C19—H19A···Cg1ii0.972.793.7362 (14)165
Symmetry codes: (i) x, y+2, z+1; (ii) x1, y+1, z.
Polarization curves and electrochemical impedance spectroscopy (EIS) top
MethodPolarization curvesEIS
Compound IConcentration (M)-Ecorr (mV)Icorr (µA cm-2)IE (%)Rct (Ω cm-2)IE (%)
H2SO40.54322794.448.08
Inhibitor5 × 10-349848.9498.24266.2296.96
Structural parameters (Å, °) calculated in gas and different solvents using the M062x/6-311+g(d) level of theory top
Exp.Gas (ε = 1)n-Hexane (ε = 1.8819)Chloroform (ε = 4.711)DCM (ε = 8.93)Acetone (ε = 20.493)Methanol (ε = 32.613)Water (ε = 78.355)
C27—N41.4101.4061.4061.4061.4071.4071.4071.407
N4—C251.2741.2751.2761.2771.2771.2771.2771.277
C25—C51.4601.4661.4661.4661.4661.4661.4661.465
C35—O21.3701.3521.3541.3561.3571.3571.3581.358
O1—C101.3601.3501.3481.3471.3461.3461.3461.346
O2—H32.0662.1242.1252.1242.1242.1232.1222.122
C27—N4—C25120120120120120120120120
N4—C25—C5122122122122122122120120
C27—N4—C25—C5178175175175175176176176
R.m.s. error (Å)-0.3680.3530.3400.3350.3300.3290.328
Thermodynamic parameters calculated in gas and different solvents at room temperature using the M062x/6-311+g(d) level of theory top
Gas (ε = 0)n-Hexane (ε = 1.8819)Chloroform (ε = 4.711)Acetone (ε = 20.493)DCM (ε = 8.93)Methanol (ε = 32.613)Water (ε = 78.355)
H (a.u.)-938.949911-938.955536-938.959863-938.962279-938.962556-938.962832
G (a.u.)-939.017972-939.023693-939.028201-939.030818-939.031122-939.031430
Cv (cal mol-1 K-1)73.81373.91674.01274.07674.08474.092
S (cal mol-1 K-1)143.245143.448143.828144.252144.309144.376
Highest occupied molecular orbital (HOMO) Energy (ε, eV), chemical potential (µ, eV), chemical hardness (η, eV), global electrophilicity index (ω, eV) and fraction of the transferred electron ΔN, eV) top
Gasn-HexaneChloroformDCMAcetoneMethanolWater
εH-6.806-6.883-6.941-6.959-6.971-6.975-6.978
µ5.5785.6025.6125.6145.6155.6155.615
η-4.016-4.082-4.134-4.152-4.164-4.167-4.170
ω1.4461.4871.5221.5351.5431.5461.549
ΔN/I Fe0.2670.2600.2550.2530.2520.2520.251
ΔN/I Cu0.0860.0800.0750.0730.0720.0720.072
Local reactivity descriptors calculated for the corrosion inhibitor I in the gas phase and in various solvents. top
Gasn-HexaneChloroformDCMAcetoneMethanolWater
f-Δff-Δff-Δff-Δff-Δff-Δff-Δf
O10.0443-0.02040.0447-0.01990.0484-0.02360.0532-0.02850.0579-0.03340.0594-0.0350.0609-0.0365
O20.0754-0.04690.0697-0.04490.0589-0.0370.0501-0.02910.0424-0.0220.0401-0.01990.0379-0.0179
N40.03530.08590.04850.07810.06520.06720.0750.05980.08210.05440.0840.05290.08560.0518
C50.0624-0.04870.073-0.05970.0938(-0.0816)0.1119-0.10010.1284(-0.117)0.1344-0.12420.1392-0.1292
C250.01160.09770.00840.1178-0.00090.1416-0.010.1387-0.018(0.135)-0.01990.134-0.02210.1332
C270.0613-0.02420.0588-0.0270.048-0.0210.0364-0.01130.0258-0.00190.02280.00080.01990.0033
Comparison between the total energy of molecular docking of ciprofloxacin and I into the binding site of PDB entry 2XCT. PLP is ????? top
Energy overviewCiprofloxacinI
Total energy-102.177-97.198
I. External Ligand interactions-86.388-81.785
A. Protein–ligand PLP interactions-12.886-9.597
Nonpolar-11.316-8.713
Hydrogen bond-1.5110
Metal00
Buried-0.06-0.884
B. Cofactor–ligand PLP interactions-73.502-72.188
Nonpolar-67.436-67.81
Hydrogen bond-4.069-7.316
Metal00
Repulsive0.060.076
Buried-2.0572.862
II. Internal ligand interactions4.2114.587
Torsional potential1.2174.587
Clash potential2.9940
 

Acknowledgements

The authors acknowledge the University of Mentouri Brothers, Constantine 1, for constant support. The authors thank the Institute of Analytical Sciences, University Claude Bernard Lyon 1, France, for use of their computing facilities. HSE is grateful to the University of Neuchâtel for their support over the years. Funding for this research was provided by the Department of Higher Scientific Research and CHEMS Research Unit, University of Constantine 1, CRBT research center, Algeria.

References

First citationAfshari, F., Ghomi, E. R., Dinari, M. & Ramakrishna, S. (2023). Chem. Select, 8, e202203231.  Google Scholar
First citationAlaoui Mrani, S., Ech-chihbi, E., Arrousse, N., Rais, Z., El Hajjaji, F., El Abiad, C., Radi, S., Mabrouki, J., Taleb, M. & Jodeh, S. (2021). Arab. J. Sci. Eng. 46, 5691–5707.  Web of Science CrossRef CAS Google Scholar
First citationAljourani, J., Raeissi, K. & Golozar, M. A. (2009). Corros. Sci. 51, 1836–1843.  Web of Science CrossRef CAS Google Scholar
First citationBrown, E. V. & Granneman, G. R. (1975). J. Am. Chem. Soc. 97, 621–627.  CrossRef CAS Web of Science Google Scholar
First citationChandra, S., Jain, D., Sharma, A. K. & Sharma, P. (2009). Molecules, 14, 174–190.  CrossRef PubMed CAS Google Scholar
First citationChen, X. M., Li, H. R., Feng, X. L., Wang, H. T. & Sun, X. H. (2022). ACS Omega, 7, 24942–24950.  Web of Science CrossRef CAS PubMed Google Scholar
First citationDeng, S., Li, X. & Fu, H. (2011). Corros. Sci. 53, 3596–3602.  CrossRef CAS Google Scholar
First citationDomingo, L. R., Aurell, M. J., Pérez, P. & Contreras, R. (2002). Tetrahedron, 58, 4417–4423.  Web of Science CrossRef CAS Google Scholar
First citationDöner, A., Solmaz, R., Özcan, M. & Kardaş, G. (2011). Corros. Sci. 53, 2902–2913.  Google Scholar
First citationFellahi, Z., Chenaf-Ait youcef, H., Hannachi, D., Djedouani, A., Ouksel, L., François, M., Fleutot, S. & Bourzami, R. (2021). J. Mol. Struct. 1244, 130955.  Web of Science CSD CrossRef 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., Petersson, G. A., Nakatsuji, H., Li, X., Caricato, M., Marenich, A. V., Bloino, J., Janesko, B. G., Gomperts, R., Mennucci, B., Hratchian, H. P., Ortiz, J. V., Izmaylov, A. F., Sonnenberg, J. L., Williams-Young, D., Ding, F., Lipparini, F., Egidi, F., Goings, J., Peng, B., Petrone, A., Henderson, T., Ranasinghe, D., Zakrzewski, V. G., Gao, J., Rega, N., Zheng, G., Liang, W., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Throssell, K., Montgomery, J. A. Jr, Peralta, J. E., Ogliaro, F., Bearpark, M. J., Heyd, J. J., Brothers, E. N., Kudin, K. N., Staroverov, V. N., Keith, T. A., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A. P., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Millam, J. M., Klene, M., Adamo, C., Cammi, R., Ochterski, J. W., Martin, R. L., Morokuma, K., Farkas, O., Foresman, J. B. & Fox, D. J. (2019). GAUSSIAN16. Revision C.01. Gaussian Inc., Wallingford, CT, USA. https://gaussian.com/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 citationHannachi, D., El Houda Amrane, N., Merzoud, L. & Chermette, H. (2021). New J. Chem. 45, 13451–13462.  Web of Science CrossRef CAS Google Scholar
First citationHannachi, D., Ouddai, N., Arotçaréna, M. & Chermette, H. (2015). Mol. Phys. 113, 1541–1550.  Web of Science CrossRef CAS Google Scholar
First citationJafari, H., Ameri, E., Rezaeivala, M. & Berisha, A. (2022). J. Indian Chem. Soc. 99, 100665.  CrossRef Google Scholar
First citationJohn, M., Ralls, M. A., Kuruveri, U. B. & Menezes, P. M. (2023). Metals, 13, 397.  CrossRef Google Scholar
First citationKüstü, C., Emregül, K. C. & Atakol, O. (2007). Corros. Sci. 49, 2800–2814.  Google Scholar
First citationLashgari, M., Arshadi, M. R. & Miandari, S. (2010). Electrochim. Acta, 55, 6058–6063.  CrossRef CAS Google Scholar
First citationLemoui, R., Allal, H., Hannachi, D., Djedouani, A., Ramli, I., Mohamed el hadi, S., Habila, I., Zabat, M., Merazig, H., Stoeckli-Evans, H. & Ghichi, N. (2023). J. Mol. Struct. 1286, 135569.  CrossRef Google Scholar
First citationLesar, A. & Milošev, I. (2009). Chem. Phys. Lett. 483, 198–203.  Web of Science CrossRef CAS Google Scholar
First citationLi, C., Yang, W., Zhou, W., Zhang, M., Xue, R., Li, M. & Cheng, Z. (2016). New J. Chem. 40, 8837–8845.  Web of Science CSD CrossRef CAS Google Scholar
First citationLi, X., Deng, S. & Fu, H. (2011). Corros. Sci. 53, 302–309.  CrossRef CAS 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 citationMatsumoto, Y., Sawamura, J., Murata, Y., Nishikata, T., Yazaki, R. & Ohshima, T. (2020). J. Am. Chem. Soc. 142, 8498–8505.  CrossRef CAS PubMed Google Scholar
First citationMcKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2007). Chem. Commun. pp. 3814–3816.  Web of Science CrossRef Google Scholar
First citationMezhoud, B., Bouchouit, M., Said, M. E., Messaadia, L., Belfaitah, A., Merazig, H., Chibani, A., Bouacida, S. & Bouraiou, A. (2016). Res. Chem. Intermed. 42, 7451.  Google Scholar
First citationMichaelson, H. B. (1977). J. Appl. Phys. 48, 4729–4733.  CrossRef CAS Web of Science Google Scholar
First citationMishra, N., Yadav, R., Kumar, K., Pandey, H. & Pandey, R. (2020). J. Phys. Conf. Ser. 1504, 012002.  CrossRef Google Scholar
First citationMohamed, G. G. (2006). Spectrochim. Acta A Mol. Biomol. Spectrosc. 64, 188–195.  CrossRef PubMed Google Scholar
First citationMorell, C., Grand, A. & Toro-Labbé, A. (2005). J. Phys. Chem. A, 109, 205–212.  Web of Science CrossRef PubMed CAS Google Scholar
First citationMusa, A. Y., Kadhum, A. A. H., Mohamad, A. B. & Takriff, M. S. (2010). Corros. Sci. 52, 3331–3340.  CrossRef CAS Google Scholar
First citationObot, I. B., Obi-Egbedi, N. O. & Umoren, S. A. (2009). Corros. Sci. 51, 276–282.  CrossRef CAS Google Scholar
First citationOlar, R., Badea, M., Ferbinteanu, M., Stanica, N. & Alan, I. (2017). J. Therm. Anal. Calorim. 127, 709–719.  CrossRef CAS Google Scholar
First citationOrona, G., Molinar, V., Fronczek, F. R. & Isovitsch, R. (2011). Acta Cryst. E67, o2505–o2506.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationParr, R. G. & Pearson, R. G. (1983). J. Am. Chem. Soc. 105, 7512–7516.  CrossRef CAS Web of Science Google Scholar
First citationParr, R. G., von Szentpály, L. & Liu, S. (1999). J. Am. Chem. Soc. 121, 1922–1924.  Web of Science CrossRef CAS Google Scholar
First citationParr, R. G. & Yang, W. (1984). J. Am. Chem. Soc. 106, 4049–4050.  CrossRef CAS Web of Science Google Scholar
First citationQin, W., Long, S., Panunzio, M. & Biondi, S. (2013). Molecules, 18, 12264–12289.  Web of Science CrossRef CAS PubMed Google Scholar
First citationRoy, R. K., Krishnamurti, S., Geerlings, P. & Pal, S. (1998). J. Phys. Chem. A, 102, 3746–3755.  Web of Science CrossRef CAS Google Scholar
First citationSakki, B., Said, M. E., Mezhoud, B., Allal, H., Larbah, L., Kherrouba, A., Chibani, A. & Bouraiou, A. (2022). J. Adh. Sci. Tech. 36, 2245–2268.  CrossRef CAS Google Scholar
First citationSegura, J. L., Mancheño, M. J. & Zamora, F. (2016). Chem. Soc. Rev. 45, 5635–5671.  Web of Science CrossRef CAS PubMed Google Scholar
First citationSheldrick, G. M. (2008). Acta Cryst. A64, 112–122.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSheldrick, G. M. (2015). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSk, I., Khan, M. A., Haque, A., Ghosh, S., Roy, D., Homechuadhuri, S. & Alam, M. A. (2020). Curr. Opin. Green Sustain. Chem. 3, 100006.  CrossRef Google Scholar
First citationSpackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19–32.  Web of Science CrossRef 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 citationSpek, A. L. (2020). Acta Cryst. E76, 1–11.  Web of Science CrossRef IUCr Journals Google Scholar
First citationStoe & Cie (2022). WinXpose, Recipe, Integrate and LANA in X-AREA. Stoe & Cie GmbH, Darmstadt, Germany.  Google Scholar
First citationTan, S. L., Jotani, M. M. & Tiekink, E. R. T. (2019). Acta Cryst. E75, 308–318.  Web of Science CrossRef IUCr Journals Google Scholar
First citationTarafder, M. T. H., Ali, M. A., Saravanan, N., Weng, W. Y., Kumar, S., Umar-Tsafe, N. & Crouse, K. A. (2000). Transition Met. Chem. 25, 295–298.  Web of Science CrossRef CAS Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationYakan, H., Omer, H. S., Buruk, O., Çakmak, Ş., Marah, S., Veyisoğlu, A., Muğlu, H., Ozen, T. & Kütük, H. (2023). J. Mol. Struct. 1277, 134799.  CrossRef Google Scholar
First citationZakerhamidi, M. S., Ghanadzadeh, A. & Moghadam, M. (2012). Chem. Sci. Trans. 1, 1–8.  CAS Google Scholar
First citationZhang, Z., Song, Q., Jin, Y., Feng, Y., Li, J. & Zhang, K. (2023). Metals, 13, 386.  CrossRef Google Scholar
First citationZheludkevich, M. L., Yasakau, K. A., Poznyak, S. K. & Ferreira, M. G. S. (2005). Corros. Sci. 47, 3368–3383.  CrossRef CAS Google Scholar

This article is published by the International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.

Journal logoSTRUCTURAL
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