Next Article in Journal
Metformin Loaded Zein Polymeric Nanoparticles to Augment Antitumor Activity against Ehrlich Carcinoma via Activation of AMPK Pathway: D-Optimal Design Optimization, In Vitro Characterization, and In Vivo Study
Previous Article in Journal
Sn(IV)porphyrin-Incorporated TiO2 Nanotubes for Visible Light-Active Photocatalysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Studies of the Functionalized α-Hydroxy-p-Quinone Imine Derivatives Stabilized by Intramolecular Hydrogen Bond

1
Institute of Chemistry and Chemical Technology, Faculty of Natural Sciences and Technology, Riga Technical University, P. Valdena Str. 3, LV-1048 Riga, Latvia
2
Latvian Institute of Organic Chemistry, Aizkraukles Str. 21, LV-1006 Riga, Latvia
3
Institute of Materials and Surface Engineering, Faculty of Natural Sciences and Technology, Riga Technical University, P. Valdena Str. 3, LV-1048 Riga, Latvia
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(7), 1613; https://doi.org/10.3390/molecules29071613
Submission received: 6 March 2024 / Revised: 25 March 2024 / Accepted: 29 March 2024 / Published: 3 April 2024
(This article belongs to the Section Organic Chemistry)

Abstract

:
In this work, reactions between 6,7-dichloropyrido[1,2-a]benzimidazole-8,9-diones with different benzohydrazides were studied. Nucleophilic substitution at C(6) was followed by isomerization and led to α-hydroxy-p-quinone imine derivatives. Synthesized compounds represent a combination of several structural motifs: a benzimidazole core fused with α-hydroxy-p-quinone imine, which contains a benzamide fragment. X-ray crystallography analysis revealed the formation of dimers linked through OH···O interactions and stabilization of the imine form by strong intramolecular NH···N hydrogen bonds. The protonation/deprotonation processes were investigated in a solution using UV–Vis spectroscopy and a 1H NMR titration experiment. Additionally, the electrochemical properties of 6,7-dichloropyrido[1,2-a]benzimidazole-8,9-dione and its α-hydroxy-p-quinone imine derivative as cathode materials were investigated in acidic and neutral environments using cyclic voltammetry measurements. Cathode material based on 6,7-dichloropyrido[1,2-a]benzimidazole-8,9-dione could act as a potentially effective active electrode in aqueous electrolyte batteries; however, further optimization is required.

1. Introduction

Quinones and quinone derivatives are well known due to the redox activity that is important in a wide range of biological processes, such as photosynthesis [1] and cellular respiration [2]. Quinones represent a class of biologically active compounds with both cytotoxic and cytoprotective effects [3]. Considerable attention to redox-active compounds in general and quinones in particular [4,5] can be explained by growing demands on energy storage devices for portable electronics and renewable energy-powered vehicles [6]. Quinones can potentially be used in different applications connected with energy storage due to their remarkable redox activity: as organic cathode materials for different kinds of rechargeable batteries [7], including redox flow batteries [8] and Zn-ion batteries [9], or as redox mediators in lithium–sulfur batteries [10]. Physical properties of quinones can be modulated by the introduction of heteroaromatics fused with quinone cores [8] and different substituents that affect solubility [11] or affect the form of quinone fragment [12] or can facilitate binding with metal cations [13]. Additionally, redox properties may be tuned to some extent by intra- and intermolecular hydrogen bonding [14].
Redox potentials, solubility, and stability in the case of small quinones can be affected by modification with electron-donating or -withdrawing functional groups or in combination with a side chain that can form hydrogen bonds. Additionally, substituted o-quinones, besides their “classical” form, can also exist in various forms [15], like quinone methides, quinone imines, and zwitterions. This ability to adopt different structures allows modulation of their properties for different applications. Despite progress in the design of quinone derivatives and their wide application as redox-active materials, limited information on molecular-level insights into the bulk properties (e.g., solubility, stability, redox activity, etc.) is available. The investigation of quinone structure at the molecular level, self-assembly in solid state, and behavior in solution will help tune the performance of quinone-functionalized materials.
An approach to modulate the redox properties of quinones is an introduction of nitrogen-containing redox-active groups (e.g., C≡N, C=N, and N=N) or incorporation of unsaturated carbon–nitrogen bonds and π-conjugated aromatic fragments [16].
This work aimed to gain an understanding of the molecular structure and behavior of unsymmetrical heterocyclic o-quinones and their derivatives bearing imine moiety. Also, electrochemical studies of selected compounds were conducted to assess their potential applications.

2. Results and Discussion

2.1. Synthesis and Structural Studies of Quinone Derivatives 3ag

6,7-Dichloropyrido[1,2-a]benzimidazole-8,9-dione (1a) is a representative of unsymmetrical heterocyclic o-quinones that contains a combination of two structural motifs, an o-quinone fragment and imidazo[1,2-a]pyridine core, that possess C=N bonds (Scheme 1). It can be obtained in one-step synthesis from commercially available tetrachloro-1,4-benzoquinone and 2-aminopyridine [17]. During earlier studies [17,18,19], it was proved that quinone 1a and some of its derivatives are electrochemically active compounds. Investigation of the reactivity of heterocyclic quinones 1 with C- and N-nucleophiles (primary amines) indicated that the attack of the nucleophile proceeds selectively at the C(6)-position of quinone. Interestingly, obtained compounds containing different acceptor groups at the C(6)-position can exist in an o-quinone form or as p-quinone methides, depending on the introduced substituent.
To expand the scope of redox-active heterocyclic o-quinone derivatives, the modification of quinone 1a with different benzohydrazides was carried out, and structural studies in solid state and solution were conducted.

2.1.1. Synthesis of Quinone Derivatives 3ag

Quinone derivatives 3ag (Scheme 1a, atoms are numbered according to ORTEP diagram, vide infra) were obtained by the nucleophilic substitution of a chlorine atom of quinone 1a,b by benzohydrazides 2ad. Isolated compounds 3ag have a red-orange color in the solid state. Interestingly, in the case of aminoderivatives of quinone 1 (a merocyanine on the base of the o-quinone form), deep-blue colored crystals were obtained [17,19]. In general, derivatives containing aroyl hydrazine fragments were expected as a result of such substitution [15,20], and a few tautomeric structures can be supposed for the products [21,22]. For the compounds obtained (3ag), the structure determination of the quinone/substituent fragments (Scheme 1b) can explain the observed difference in color of crystals 3ag in comparison to aminosubstituted derivatives of quinone 1a.

2.1.2. Single-Crystal X-ray Analysis of Quinone Derivative 3a

The use of routine identification procedures (such as 1H-NMR and FTIR) to establish the molecular structure of compounds 3ag left some room for doubt. To clarify the situation, crystals of compound 3a were grown from dichloromethane (DCM) solution, and their molecular structure was established using single-crystal X-ray crystallography. Crystal data and refinement details for the studied crystal are presented in Table 1.
Figure 1a shows a perspective view of molecule 3a with thermal ellipsoids and the atom-numbering scheme. The p-quinone imine form was confirmed by the inspection of the bond length: bonds between C(9)=O(23) and C(6)=N(11) have double-bond character, and O(22)-C(8) is a single bond (typical bond distances: d (Å) (C-O) = 1.34, (C=O) = 1.21, (C-N) = 1.38, (C=N) = 1.28 [23]). Also, an analysis of bond lengths shows that the structure of compound 3a can be represented as a superposition of mesomeric forms. The main forms are shown in Figure 1b; at that, the non-ionized form has the highest specific weight.
The molecules of compound 3a are characterized by a flattened conformation; only the phenyl group is slightly out of the plane of the heterocyclic system (the angle between planes is 5.19°). In the structure of compound 3a, intramolecular hydrogen bonds NH···N and OH···O were found (Figure 2). The hydroxy group of compound 3a forms bifurcated hydrogen bonds where O(22)-H···O(23) is an intramolecular bond (Figure 1a) and O(22)-H···O’(23) is an intermolecular one (Figure 2a). By means of these intermolecular H-bonds, the centrosymmetric R 2 2 ( 10 ) molecular dimers are formed in the crystal structure. Additional intermolecular interactions were found: stacking interaction between the planes of the molecules and a short intermolecular contact between heterocycle (C(3)-H) and the amide group of the substituent (d C(3)···O(14) = 3.112 Å, d C(3)-H···O(14) = 2.344 Å) (Figure 2b,c).
Hirshfeld surfaces and energy framework calculations (Figures S14 and S15) were obtained in a whole-of-molecule approach using the B3LYP/6-31G(d,p) energy model implemented in CrystalExplorer 21.5 software [24]. The Hirshfeld surface (shown in Figure S14, mapped with dnorm) investigation is a useful approach to establishing close contacts in a crystal. Areas of close contact are highlighted with different colors (red, white, and blue spots) to show intermolecular contacts with distances less than, equal to, and larger than van der Waal radii. As seen from the Hirshfeld surface analysis of compound 3a, bright red areas of the same size indicate strong interactions, specifically hydrogen bonds OH···O, between neighboring molecules. Fainter red or white areas suggest relatively weak interactions involving hydrogen atoms from heterocyclic and phenyl rings and chlorine atoms. Examples include C-H bonds interacting with oxygen (C(3)-H···O(14)), nitrogen (C(3)-H···N(11)), and chlorine (C(4)-H··· Cl and C(phenyl)-H⋯Cl).
Energy frameworks provide an opportunity to explore the cooperative effects of intermolecular interactions in the crystal, admitting the electrostatic, dispersion, and total energy between pairs of molecules [25]. In the case of the crystal of compound 3a (Figure S15), a strong stabilizing interlayer electrostatic interaction was found between molecules involved in the formation of hydrogen-bonded (O–H⋯O) dimers. On the other hand, dispersion energy was more dominant for the intercolumn stacking motif. Overall, the energy framework analysis of the crystal revealed two distinct patterns of electrostatic and dispersion energies, with each contributing similarly.

2.1.3. 1H NMR Spectroscopy Analysis of Quinone Derivatives 3ag

To determine the structure of obtained products in solution compounds 3ag, 1H NMR spectroscopy data were analyzed, and a set of two broad signals corresponding to NH and OH protons was observed (Figures S1–S7). In the DMSO-d6 solution, signals appeared at 14.36–14.90 ppm and can be assigned to the NH proton, while signals of the OH group were observed at 10.89–11.41 ppm. Additionally, the 1H NMR spectrum of compound 3a was also recorded in CDCl3 solution (a solvent in which hydrogen-bonding interactions are expected to be weaker [26]) (Figure S8), where the NH proton was found at 14.68 ppm (versus 14.71 ppm in DMSO-d6 solution). It can be concluded that a strong intramolecular bond between the NH group proton of the substituent (benzamide group at imine bond) and the nitrogen of the heterocycle (N(12)-H···N(5)) can be found in solution as well as in a solid state (vide supra).
It is known [27] that in the case of α-hydroxyquinone derivatives, an intramolecular hydrogen bond was observed, and an OH proton signal appears at 7.30 ppm in the CDCl3 solution. For compound 3a, a distinguishable shift was observed for the OH proton signal in the CDCl3 solution (7.20 ppm) in comparison to the DMSO-d6 solution (10.99 ppm), which can indicate the formation of the additional intermolecular interactions between the OH group and a solvent (DMSO-d6) with hydrogen bond acceptor abilities [28].
Moreover, the introduction of various substituents in the phenyl ring (benzamide fragment) showed the linear correlation (Figure S9) between the type of substituent group (represented by the Hammett constant [29] (Table S1)) and OH signal shift in the sets of compounds 3ac (R1 = H) and 3eg (R1 = NO2). The shift of the OH signal correlates well (R2 = 0.99) with the increasing electron-withdrawing character of the substituent group for both sets of compounds (Figure S9). The presence of an electron-withdrawing group (EWG) at the heterocyclic fragment (R1 = NO2) led to a downfield shift in the OH signal compared to molecules without any substituent (R1 = H) on that ring (ΔδOH ≈ 0.3 ppm).
The most downfield signal of the NH proton (14.90 ppm in DMSO-d6 solution) in the 1H NMR spectra was observed for compound 3c with a NO2 group at the phenyl ring (R2 at benzamide fragment). The most upfield signal of the NH proton (14.36 ppm) was found for compound 3f with a NO2 group at the heterocyclic fragment (R1) and electron-donating group at the phenyl ring (R2 = OMe). Well-defined linear correlations between the Hammett substituent constant and the NH signal are presented in Figure S9. The variation in the electronic character of the substituent in the para-position of the phenyl ring showed the same trend for compounds 3ac (R2 = 0.96) and 3eg (R2 = 0.99), leading to the downfield shift of the NH signal going from an electron-donating to electron-withdrawing group. Interestingly, the presence of EWG (R₁ = NO₂) at the heterocyclic ring caused the NH signal upfield shift (ΔδNH ≈ 0.3 ppm) compared to a molecule where the same position has a hydrogen atom (R₁ = H). In general, the more downfield shifted the NH proton signal, the stronger the intramolecular H-bond [30]. Thus, EWG at the phenyl ring (R2 = NO2) increases the acidity of the NH proton and increases the intramolecular H-bond strength, but EWG at the heterocyclic fragment (R1 = NO2) influences the electron density at N(5), affecting intramolecular H-bond in turn.
In the case of compounds 3ag, the moiety at the C(6) position can be described as a structural analog of aroyl hydrazone (a different approach [22] to the naming of quinone imine derivatives was observed). A well-known characteristic of compounds containing the carbon–nitrogen double bond is the ability to undergo E/Z isomerization in the solution activated by light and/or chemical inputs [31,32].
Compounds 3a and 3b were chosen for the investigation of the isomerization process. In general, in the case of E/Z isomerization, an additional set of signals [33,34] is expected to appear. No signals of the second form (isomerization products) were observed in the 1H NMR spectra of compound 3a either in DMSO-d6 or in CDCl3 solution. Also, in the case of compound 3b, configurational switching was not induced by the addition of excess trifluoracetic acid (TFA) and following irradiation by UV light (365 nm; irradiation by a high-pressure mercury lamp at room temperature) judging from the 1H NMR spectra of compound 3b in DMSO-d6 solution (Figure S10). The formation of a strong intramolecular hydrogen bond N(12)-H···N(5) can explain the existence of a single configuration of substituted imine that agrees with the stabilization of only one form in the presence of an intramolecular hydrogen bond. Additional stabilization of the molecule may be explained by excitation energy dissipation caused by zwitterionic structure.
It is known [35] that for redox properties tests (fabrication of electrodes), a mixture of an organic compound, a conductive additive, and a binder is often prepared using N-methyl-2-pyrrolidone (NMP) [36] as a solvent (strongly basic solvent) [37]. This fact prompted us to investigate the influence of the base on the structure of the products 3ag.
During preliminary solubility tests of compounds 3ag, the color change (from yellow to green or blue) was observed in NMP solution or in the presence of a base. For a better understanding of the effect, a few 1H NMR experiments were carried out. Upon addition of an excess of the base (1,8-diazabicyclo(5.4.0)undec-7-ene, DBU), the 1H NMR spectrum of compound 3b in DMSO-d6 solution showed some changes (Figure 3): the signal of OH proton completely disappeared; the sharpened signal of the NH proton (the sharp line can indicate a dynamically stable state) shifted upfield. Additionally, a new minor proton signal at 13.44 ppm was observed. Simultaneously, the yellow-colored solution of compound 3b became dark blue. It should be noted that the same deprotonation behavior was observed for compound 3a in CDCl3 solution (Figure S11). After the excess TFA was added, the solution became yellow again, a minor signal at 13.44 ppm disappeared, and the signal of the OH proton was restored. It can be concluded that deprotonation provides the formation of an anionic polymethine dye structure [38] (blue), and protonation restores quinone imine form (yellow); consequently, the equilibrium between the two forms is reversible.
To gain more information about the deprotonation process of compound 3b, a 1H NMR titration experiment was carried out. As shown in Figure 4, in the case of compound 3b deprotonation upon sequential addition of the base (DBU), the 1H NMR spectra reveal several features. The broad signal of the OH proton vanished upon the addition of only 0.15 equivalents of the base, which can be explained by the dynamical process as well; color changes were immediate (Figure 4, highlighted in yellow). The signal of the NH proton (benzamide fragment) sharpens and undergoes an upfield shift from 14.70 to 14.51 ppm (Figure 4, highlighted in red).
A new signal appeared at 9.58 ppm (+0.15 eqv. of DBU) (Figure 4, highlighted in green), which can be explained by the formation of a hydrogen-bonded complex of protonated DBU (1H NMR spectrum of DBU and TFA mixture in DMSO-d6 solution was recorded; the NH+ signal appears at 9.69 ppm (Figure S12)) with the deprotonated compound 3b. Upon further addition of the base, this signal was broadened and shifted downfield (9.86 ppm) due to the interaction of protonated DBU (DBUH+) with the anionic compound 3b through the N–H bond [39].
Upon addition of more than 1.05 eqv. of the base, a second minor form of the compound appears (Figure 4, highlighted in blue); the ratio between major and minor forms is 0.95:0.05, taking into consideration the signals of all protons. Moreover, the addition of an excess amount of the base (4 and 8 eqv.) did not result in the change of the 1H NMR spectra of compound 3b (additional processes as a function of the time and/or temperature in the solution should not be excluded as 1H NMR titration experiment was carried out within an hour after addition of DBU to the compound 3b at room temperature (T = 294 K)).
Also, the presence of two different species (one major and one minor) was detected from the changes in the 1H NMR spectra of compounds 3a and 3c upon deprotonation with DBU in DMSO-d6 solution (after mixed with more than 1 equivalent of DBU). For compound 3a, the 1H NMR spectrum was also recorded in the presence of NaOH. As a result, the acquired spectrum was identical to the one with an excess of DBU (Figure S13).
Unfortunately, the low solubility of compounds 3ag limited the possibility of obtaining qualitative 13C NMR spectra.

2.1.4. Electronic Absorption

The UV–Vis absorption spectra of compound 3a were investigated in solution using solvents of various polarities (DCM and DMSO). Two absorption maxima were found at 381 nm and at 446–449 nm in the absorption spectra of compound 3a in both solutions (Figure 5a), which can be attributed to the neutral form of the compound. Upon addition of the base (DBU) to the DCM solution of 3a, the solution instantaneously turned violet, and the absorption revealed a broad band centered at 556 nm. When a base was added to the DMSO solution of compound 3a, the bathochromic shift was observed with an absorption maxima at 607 nm accompanied by a blue coloration.
The effect of substituents (R1 and R2) on the longwave absorption maximum was examined when DBU was added to the initial solution of compounds 3ac (variable R2 while R1 = H) and 3eg (variable R2 while R1 = NO2) in CHCl3 (Figure S16, Table S2). It was noticed that the longwave absorption band of deprotonated species of derivatives 3eg was red-shifted (λ = 569–578 nm) in comparison to compounds 3ac (λ = 556–563 nm). This observation can be explained by the influence of the electron-withdrawing substituent at C(2) on the electron distribution in the heterocyclic fragment. Moreover, a hyperchromic effect was caused by the introduction of the nitro group in the benzamide fragment (compound 3c in Figure 5b; compounds 3c and 3g in Figure S16).

2.2. Electrochemistry/Redox Chemistry Studies of Quinone Derivatives 1a and 3a

Quinone imines contain structural fragments with potentially high redox activity. It was shown by Almeida, R. et al. [40] that p-quinone imines are known to undergo a redox cycle through aminophenols [41]. Compound 1a was previously proved to be electrochemically active [42] when dissolved in the MeCN solution. However, it has not been tested as a cathode material before. To analyze the redox properties of quinone imine 3a in comparison to initial o-quinone 1a, open-circuit potential (OCP) and cyclic voltammetry (CV) measurements in a solid state were carried out. Preliminary solubility tests showed limited solubility of compounds 1a and 3a in aqueous media; compound 1a was insoluble in water (whole pH range), and at the same time, compound 3a was insoluble in neutral and acidic environments.

2.2.1. Open-Circuit Potential Measurements

To analyze the electrochemical properties of the compounds, CV measurements and OCP measurements before and after CV were performed (Figure S17). Cathode materials CM-1a and CM-3a were prepared by combining compounds 1a and 3a with Vulcan XC72 CB, respectively (the detailed sample preparation is described in the Experimental section). The OCP of freshly assembled half-cells for cathode materials CM-1a and CM-3a in the acidic (0.5 M H2SO4) electrolyte was 0.44 V and 0.37 V vs. Ag/AgCl; however, in a neutral (0.5 M K2SO4) electrolyte, the potentials of both were 0.28 V vs. Ag/AgCl. After the CV measurements, the OCP of sample half-cells stabilized. For both samples in the acidic electrolyte, they were 0.41 V vs. Ag/AgCl, and in the neutral electrolyte, they were 0.37 V vs. Ag/AgCl. The OCP for both samples is approximately the same, and with decreasing pH, there is a visible shift to higher potential values going from neutral to the acidic electrolyte.

2.2.2. Cyclic Voltammetry Measurements

CV results for samples CM-1a, CM-3a, and a sample without active material (substrate) in neutral and acidic electrolytes at varying scanning speeds are shown in Figure 6. For the substrate in neutral electrolyte, no visible redox processes were observed. In the acidic electrolyte, a hydrogen evolution reaction can be observed around the potential of −0.3 V vs. Ag/AgCl and one insignificant redox process at faster scan rates around 0.3 V vs. Ag/AgCl. However, when scanning samples with active materials, this substrate process cannot be observed and, therefore, has no electrochemical significance other than providing electrical conductivity.
For sample CM-3a in a neutral electrolyte (Figure 6), no significant redox processes can be observed. Also, in an acidic electrolyte for sample CM-3a (Figure 6), there are no significant processes; however, upon closer inspection (Figure S18), two reversible oxidation (at 0.40 V and 0.07 V vs. Ag/AgCl) and reduction processes (at 0.23 V and −0.06 V vs. Ag/AgCl) can be seen.
Sample CM-1a has two redox maxima in the scanned potential window from −0.4 V to 1.0 V vs. Ag/AgCl electrode in both neutral and acidic electrolytes (Figure 6). In a neutral electrolyte (Figure S19), oxidation peaks are at 0.27 V and 0.10 V vs. Ag/AgCl; however, reduction peaks are at 0.14 V and −0.32 V vs. Ag/AgCl. In addition, the oxidation peaks are found at 0.48 V and 0.26 V vs. Ag/AgCl and reduction peaks at 0.40 V and 0.23 V vs. Ag/AgCl in an acidic electrolyte (Figure S20). This indicates a shift in reaction potential to more positive values by increasing H+ ion concentration and thus lowering the pH level of the electrolyte. Both redox processes for sample CM-1a correspond to the o-quinone fragment in the molecule. However, by comparing the electrochemical performance of both samples CM-1a and CM-3a, it can be concluded that by converting o-quinone 1a to α-hydroxy-p-quinone imine 3a accompanied by the additional stabilization by intra- and intermolecular hydrogen bonds, the electrochemical reactivity of the cathode material has been greatly suppressed.
All CV measurement result developments in time can be seen in Figure S21. At the start, samples were cycled at the potential window of −0.4 V to 1.0 V vs. Ag/AgCl reference electrode. Observations indicate that all samples go through the surface activation phase, where the sample-specific capacity increases with each subsequent cycle. During the measurements at different scanning speeds, the stabilization of the system is observed. However, sample CM-1a in neutral electrolyte goes through an irreversible oxidation process at 0.11 V vs. Ag/AgCl and a reduction process at −0.32 V vs. Ag/AgCl. This irreversible redox process can be observed during all scan speeds. At an increased potential window (from −1.0 V to 1.5 V), another irreversible process at a scan speed of 0.1 V/s for sample CM-1a in a neutral electrolyte can be observed during the oxidation at −0.31 V and reduction at −0.56 V vs. Ag/AgCl. A slight capacity decrease due to the possible dissolution of active materials can be observed for sample CM-3a in a neutral electrolyte and sample CM-1a in an acidic electrolyte.

2.2.3. Raman Measurements

Raman measurements were performed on pure compounds 1a and 3a, prepared cathodes (CM-1a and CM-3a), and after cycling them in acidic and neutral electrolytes (Figure 7). For sample CM-3a, the spectra for prepared and cycled cathodes remain as for pure compound 3a, where an amide band can be seen at 1600–1630 cm−1, as well as bands for aromatic/heteroaromatic rings at 1550 and 1470 cm−1 [43]. This indicates that compound 3a was preserved in the cathode-forming process and did not go through any chemical changes. Also, after CV measurements, the active material is unchanged and present in the samples. For sample CM-1a, the spectra for the prepared and cycled cathode in an acidic electrolyte remain as for pure compound 1a (bands for aromatic/heteroaromatic rings at 1570 and 1450 cm−1 and band for carbonyl groups at 1650–1690 cm−1) [43]. However, for cathode CM-1a cycled in neutral electrolyte, only C and D bands of carbon [44,45] can be seen. Since redox processes are visible for this sample in CV measurements (Figure 6), the active material could have dissolved from the electrode in the electrolyte and gone through the electrochemical reactions from the electrolyte.

2.2.4. Scanning Electron Microscopy Measurements

Scanning electron microscopy examinations of compounds 1a and 3a (Figure S22) were performed to assess the morphology of the products. Compound 1a consists of needle-like particles with sizes ranging from 1 µm to 10 µm in diameter and 3 µm to 30 µm in length. Compound 3a has smaller particles with an overall size of 1 µm in diameter and 5–20 µm in length.
Also, the prepared cathode surfaces with and without active materials before and after cyclic voltammetry are shown in Figure 8. The resemblance of the structures of compounds 1a and 3a (as shown in Figure S22) can be seen in the images of cathode disks (samples CM-1a and CM-3a in Figure 8) before CV measurements. For sample CM-3a, these structures can also be seen in images after CV in neutral and acidic electrolytes with some partial dissolution in an acidic electrolyte, as fewer structures can be seen. The formation of non-covalent interactions between compound 3a and substrate can probably explain the greater stability of CM-3a in comparison to CM-1a. However, for sample CM-1a cycled in neutral and acidic electrolytes, only a few original structures can be seen. This correlates with findings from Raman spectroscopy (Figure 7b) that the active material 1a dissolves in a neutral electrolyte and goes through electrochemical reactions from it.

3. Materials and Methods

3.1. Materials and Instrumentation

Polyvinylidene fluoride (PVDF) (MW ~530,000) and Dimethylformamide (DMF) were purchased from Merck; Vulcan XC72 Carbon Black (CB) was used, and 0.05 mm thick conductive graphite paper (RERAS, purchased from China and used as electrode substrate) was used to prepare cathode materials.
Melting points were measured on a Kruess KSP 11 Melting Point Analyzer. 1H NMR spectra were recorded on a Brucker Avance spectrometer at 300 or 500, respectively, in DMSO-d6 or CDCl3 solutions. Chemical shifts were expressed in parts per million (δ, ppm) relative to solvent signal (DMSO-d6: 2.50 ppm CDCl3: 7.26 ppm for 1H NMR) [46]. Compounds 3ag are too insoluble to record a qualitative 13C NMR spectrum. Elemental CHN analysis was carried out on a Euro Vector EA 3000 analyzer. FTIR spectra were recorded on a Perkin-Elmer Spectrum 100 FTIR spectrometer. The UV–Vis absorption spectra were acquired with a Perkin-Elmer 35 UV/Vis spectrometer using 1 cm length quartz cuvettes with a concentration of compound c = 2.5 × 10−5 M. Low-resolution mass spectra were acquired on a Waters EMD 1000MS mass detector (ESI + mode, voltage 30 V) with an Xterra MS C18 5 μm 2.1 × 100 mm column and gradient eluent mode using 0.1% HCOOH in deionized water and MeCN or MeOH.

3.2. X-ray Crystallography Analysis

For compound 3a, diffraction data were collected at a low temperature (T = 150.0(1) K) on Rigaku, XtaLAB Synergy, Dualflex, HyPix diffractometer using copper monochromated Cu-Kα radiation (λ = 1.54184 Å). The crystal structure was solved with the help of the ShelXT structure solution program [47] using the Intrinsic Phasing solution method. The model was refined with version of the program olex2.refine using Levenberg–Marquardt minimization [48]. All nonhydrogen atoms were refined in anisotropical approximation. For further details, see crystallographic data for compound 3a deposited at the Cambridge Crystallographic Data Centre as Supplementary Publications Numbers CCDC 2238663 (for compound 3a). These data can be obtained free of charge via https://www.ccdc.cam.ac.uk/structures/ (accessed on 1 April 2024) or from the CCDC, 12 Union Road, Cambridge CB2 1EZ, U.K.; Fax: +44 1223 336033; e-mail: [email protected]). For crystal packing visualization, the program Mercury [49] was used.

3.3. Cathode Material Preparation

Quinone derivatives 1a or 3a were combined with Vulcan XC72 CB at a mass ratio of 5:4. The resulting powder was dried overnight at 80 °C. Then, a binder solution of PVDF:DMF (mass ratio 1:9) was added so the quinone to PVDF mass ratio would be 5:1. Stirring and ultra-sonication were used to create the ink slurry. Extra DMF was added to the slurry to form homogeneous ink (the total weight ratio of DMF to quinone was approximately 12:1). A manual doctor blade coater (with a 25 µm gap size) was employed to apply the coating onto carbon paper that was pre-dried at 120 °C for an hour. Coated cathode substrates were then dried in air to evaporate DMF. For further material characterization, cathode disks were cut out using a hollow punch.

3.4. Cyclic Voltammetry

To analyze the electrochemical properties of the different sample compounds, cyclic voltammetry (CV) measurements were performed using a 3-electrode measuring cell “TSC Surface” (from rhd instruments). For different measurements, the prepared thin-layer electrodes on carbon paper (with or without compound 1a or 3a) were used as working electrodes placed in 1 mL of electrolyte with the platinum counter electrode and Ag/AgCl (3 M KCl) reference electrode. Two different pH electrolyte solutions were used for measurements: neutral 0.5 M K2SO4 and acidic 0.5 M H2SO4 solutions. The CV measurements were performed from −0.4 to +1.0 V with the following program: (1) open-circuit measurement (OCP) of freshly assembled half-cell before the CV measurements; (2) ten cycles with scan speed of 0.075 V/s to stabilize the half-cell; (3) 5 cycles of 5 scans with scan rates ranging from 0.005 V/s to 0.1 V/s; and (4) OCP measurement after CV.

3.5. Raman Spectroscopy

Raman measurements were performed using a Renishaw In-ViaV727 spectrometer in a backscattering geometry at room temperature at 100× magnification. For phonon excitation, a red laser (He-Ne, λ = 633 nm, grating—1200 mm−1, 125 µW) was used, and the sample exposure time was 10 s.

3.6. Scanning Electron Microscopy

A Hitachi TM3000 Tabletop scanning electron microscope (SEM) with an acceleration voltage of 5 kV was used to obtain surface information of the obtained electrode and sample materials. To characterize the obtained samples, different magnifications were used.

3.7. Synthesis of Quinone Derivatives 1a,b and 3ag

6,7-Dichloropyrido[1,2-a]benzimidazole-8,9-dione (1a) and 6,7-dichloro-2-nitropyrido[1,2-a]benzimidazole-8,9-dione (1b) were prepared according to the previously reported procedure [17,42].
General method for synthesis of compounds 3ag. To a stirring solution of 6,7-dichloropyrido[1,2-a]benzimidazole-8,9-dione (1a, 1 eq) or 6,7-dichloro-2-nitropyrido[1,2-a]benzimidazole-8,9-dione (1b, 1 eq) in dichloromethane (DCM) at room temperature, a solution of benzhydrazide derivative (2ad, 2 eq) in DCM or DMF was added. Triethylamine (1 eq) was added to the reaction mixture, which was then stirred at room temperature for 8 h. After completion of the reaction, the reaction mixture was filtered through a filter paper, and the solvent was distilled in vacuo to a residual volume of 20 mL. The resulting orange-colored precipitate was collected, recrystallized from DCM/n-hexane, washed with hot ethanol (20 mL), and dried at room temperature.
Compound 3a. Prepared using 6,7-dichloropyrido[1,2-a]benzimidazole-8,9-dione (150 mg, 0.56 mmol, 1 eq), benzohydrazide (153 mg, 1.12 mmol, 2 eq), and triethylamine (d = 0.73 g/mL, v = 78 μL, 0.56 mmol, 1 eq). Yield: 59%, orange powder. M.P.: 258–260 °C. MS: C18H11ClN4O3 requires [M + H]+ 367.1; found [M + H]+ 367.2. 1H NMR (500 MHz, DMSO-d6): 14.71 (br.s., 1H, exchange with D2O, NH), 10.98 (br.s., 1H, exchange with D2O, OH), 9.29 (d, J = 6.6 Hz, 1H, H-1), 8.16 (d, J = 9.0 Hz, 1H, H-4), 8.11 (d, J = 7.4, 2H, CHPh), 7.89 (t, J = 8.0, 1H, H-3), 7.73 (m, 3H, CHPh), 7.51 (d, J = 6.8 Hz, 1H, H-2). FTIR (KBr, cm−1): 3331, 3094, 3033, 1708, 1628, 1604, 1548, 1437, 1351, 1247. Anal. Calcd. for C18H11ClN4O3: C, 58.95; H, 3.02; N, 15.28; found C, 58.82; H, 3.02; N, 15.31.
Compound 3b. Prepared using 6,7-dichloropyrido[1,2-a]benzimidazole-8,9-dione (150 mg, 0.56 mmol, 1 eq), 4-methoxybenzohydrazide (187 mg, 1.12 mmol, 2 eq), and triethylamine (d = 0.73 g/mL, v = 78 μL, 0.56 mmol, 1 eq). Yield: 36%, orange solid. M.P.: >300 °C. MS: C19H13ClN4O4 requires [M + H]+ 397.1; found [M + H]+ 397.2. 1H NMR (500 MHz, DMSO-d6): 14.70 (br.s., 1H, exchange with D2O, NH), 10.95 (br.s., 1H, exchange with D2O, OH), 9.33 (d, J = 6.7 Hz, 1H, H-1), 8.24 (d, J = 9.0 Hz, 1H, H-4), 8.11 (d, J = 8.6 Hz, 2H, CHPh), 7.91 (m, 1H, H-3), 7.53 (t, J = 6.8 Hz, 1H, H-2), 7.25 (d, J = 8.5 Hz, 2H, CHPh), 3.91 (s, 3H, -OCH3). FTIR (KBr, cm−1): 3468, 3301, 3083, 3024, 2975, 2832, 1690, 1630, 1609, 1582, 1552, 1504, 1351, 1325, 1259. Anal. Calcd. for C19H13ClN4O4: C, 57.51; H, 3.30; N, 14.12; found C, 57.58; H, 3.35; N, 13.82.
Compound 3c. Prepared using 6,7-dichloropyrido[1,2-a]benzimidazole-8,9-dione (150 mg, 0.56 mmol, 1 eq), 4-nitrobenzohydrazide (204 mg, 1.12 mmol, 2 eq), and triethylamine (d = 0.73 g/mL, v = 78 μL, 0.56 mmol, 1 eq). Yield: 52%, orange solid. M.P.: 275–278 °C. MS: C18H10ClN5O5 requires [M + H]+ 412.1; found [M + H]+ 412.2. 1H NMR (300 MHz, DMSO-d6): 14.95 (br.s., 1H, exchange with D2O, NH), 11.07 (br.s., 1H, exchange with D2O, OH), 9.29 (d, J = 6.5, 1H, H-1), 8.53 (m, 2H, CHPh), 8.32 (d, J = 8.3, 3H, H-4 un CHPh), 7.91 (m, 1H, H-3), 7.53 (m, 1H, H-2). FTIR (KBr, cm−1): 3618, 3306, 3113, 3089, 3016, 1703, 1626, 1605, 1573, 1556, 1519, 1346, 1275. Anal. Calcd. for C18H10ClN5O5: C, 52.51; H, 2.45; N, 17.01; found C, 52.20; H, 2.58; N, 16.73.
Compound 3d. Prepared using 6,7-dichloropyrido[1,2-a]benzimidazole-8,9-dione (150 mg, 0.56 mmol, 1 eq), isonicotinohydrazide (154 mg, 1.12 mmol, 2 eq), and triethylamine (d = 0.73 g/mL, v = 78 μL, 0.56 mmol, 1 eq). Yield: 50%, orange crystals. M.P.: >250 °C (decomp.). MS: C17H10ClN5O3 requires [M + H]+ 368.1; found [M + H]+ 368.2. 1H NMR (300 MHz, DMSO-d6): 14.52 (br.s., 1H, exchange with D2O, NH), 10.90 (br.s., 1H, exchange with D2O, NH), 9.31 (d, J = 6.7, 1H, H-1), 8.91 (d, J = 3.8, 2H, CHPy), 8.18 (d, J = 8.9, 1H, H-4), 7.97 (d, J = 4.6, 2H, CHPy), 7.89 (t, J = 8.0, 1H, H-3), 7.52 (t, J = 6.7, 1H, H-2). FTIR (KBr, cm−1): 3420, 3055, 1709, 1647, 1568, 1555, 1331, 1280. Anal. Calcd. for C17H10ClN5O3 + 0.5H2O: C, 54.20; H, 2.94; N, 18.59; found C, 54.29; H, 2.88; N, 18.30.
Compound 3e. Prepared using 6,7-dichloro-2-nitropyrido[1,2-a]benzimidazole-8,9-dione (150 mg, 0.48 mmol, 1 eq), benzohydrazide (131 mg, 0.96 mmol, 2 eq), and triethylamine (d = 0.73 g/mL, v = 67 μL, 0.48 mmol, 1 eq). Yield: 61%, yellow solid. M.P.: >250 °C (decomp.). MS: C18H10ClN5O5 requires [M + H]+ 412.1; found [M + H]+ 412.4. 1H NMR (300 MHz, DMSO-d6): 14.42 (br.s., 1H, exchange with D2O, NH), 11.30 (br.s., 1H, exchange with D2O, OH), 10.06 (d, J = 2.3 Hz, 1H, H-1), 8.56 (dd, J = 9.8, 2.1 Hz 1H, H-3), 8.40 (d, J = 9.7 Hz, 1H, H-4), 8.12 (m, 2H, CHPh), 7.72 (d, J = 7.7, 3H, CHPh). FTIR (KBr, cm−1): 3397, 3091, 3033, 1690, 1642, 1552, 1525, 1351, 1311, 1268. Anal. Calcd. for C18H10ClN5O5: C, 52.51; H, 2.45; N, 17.01; found C, 52.21; H, 2.67; N, 16.76.
Compound 3f. Prepared using 6,7-dichloro-2-nitropyrido[1,2-a]benzimidazole-8,9-dione (150 mg, 0.48 mmol, 1 eq), 4-methoxybenzohydrazide (160 mg, 0.96 mmol, 2 eq), and triethylamine (d = 0.73 g/mL, v = 67 μL, 0.48 mmol, 1 eq). Yield: 42%, yellow crystals. M.P.: >250 °C (decomp.). MS: C19H12ClN5O6 requires [M + H]+ 442.1; found [M + H]+ 442.2. 1H NMR (300 MHz, DMSO-d6): 14.36 (br.s., 1H, exchange with D2O, NH), 11.23 (br.s., 1H, exchange with D2O, OH), 10.05 (s, 1H, H-1), 8.56 (d, J = 9.7 Hz, 1H, H-3), 8.42 (d, J = 10.8 Hz, 1H, H-4), 8.10 (d, J = 7.1 Hz, 2H, CHPh), 7.24 (d, J = 8.3 Hz, 2H, CHPh), 3.90 (s, 3H, OCH3). FTIR (KBr, cm−1): 3399, 3085, 3028, 1687, 1640, 1605, 1555, 1523, 1350, 1310, 1266. Anal. Calcd. for C19H12ClN5O6: C, 51.66; H, 2.74; N, 15.85; found C, 51.35; H, 2.68; N, 15.61.
Compound 3g. Prepared using 6,7-dichloro-2-nitropyrido[1,2-a]benzimidazole-8,9-dione (150 mg, 0.48 mmol, 1 eq), 4-nitrobenzohydrazide (174 mg, 0.96 mmol, 2 eq), and triethylamine (d = 0.73 g/mL, v = 67 μL, 0.48 mmol, 1 eq). Yield: 36%, orange solid. M.P.: >250 °C (decomp.). MS: C18H9ClN6O7 requires [M + H]+ 457.0; found [M + H]+ 457.1. 1H NMR (300 MHz, DMSO-d6): 14.61 (br.s., 1H, exchange with D2O, NH), 11.41 (br.s., 1H, exchange with D2O, OH), 10.03 (s, 1H, H-1), 8.58 (d, J = 10.8 Hz, 2H, H-4 and H-3), 8.51 (m, 2H, CHPh), 8.32 (d, J = 8.3 Hz, 2H, CHPh). FTIR (KBr, cm−1): 3340, 3114, 3081, 1707, 1626, 1602, 1547, 1518, 1344, 1306, 1266. Anal. Calcd. for C18H9ClN6O7: C, 47.33; H, 1.99; N, 18.40; found C, 47.24; H, 2.02; N, 18.14.

4. Conclusions

To summarize, a set of heterocyclic α-hydroxy-p-quinone imine derivatives was obtained via one-step nucleophilic substitution of 6,7-dichloropyrido[1,2-a]benzimidazole-8,9-diones 1a,b with different benzohydrazides 2ad followed by isomerization. The α-Hydroxy-p-quinone imine form of the synthesized products was proved by X-ray crystallography analysis of compound 3a. The formation of a strong intramolecular hydrogen bond N(12)-H···N(5) in a solid state and in solution can explain the stabilization of only one configuration of the C=N bond of the substituted imine. The 1H NMR acid-base titration experiment showed that deprotonation/protonation processes are reversible. Deprotonation led to the electronic delocalization in the molecule, which is accompanied by distinct changes in the UV–Vis spectra.
Sample CM-1a has a distinct maximum of two redox reactions. In an acidic environment, both peaks are stable, while in a neutral environment, only one of them is stable and remains unchanged after several CV cycles. Moreover, for the stable redox reactions, the potential difference is only up to 0.2 V. This indicates that the sample CM-1a could act as an effective active electrode in aqueous electrolyte batteries. However, the dissolution of the sample in the electrolyte was observed, so the potential application as cathode material for aqueous batteries would require compound 1a to be attached to the polymer backbone.
Attempts to modulate redox properties by incorporation of additional unsaturated carbon–nitrogen bonds to the heterocyclic quinone 1a structure led to changes in the redox-active fragment and the formation of p-quinone imine 3a. As a result, the electrochemical behavior is changed, as it is no longer possible to observe pronounced redox peaks in CV measurements. Structural changes of the quinone fragment (probably induced by the formation of intramolecular H-bond) decreased the redox activity of the derivative despite the introduction of an additional C=N bond; at the same time, the introduced substituent enhanced the stability of the cathode material, which was confirmed by Raman spectroscopy measurements. Therefore, further structure optimizations for the elaboration of effective cathode material should include the analysis of the possible tautomerization, the formation of intramolecular H-bonds, the effect of substituents and interaction with basic solvents versus redox properties, and the stability of the compound.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29071613/s1. Figures S1–S8: 1H NMR spectra (for compounds 3ag in DMSO-d6 solution and for compound 3a in CDCl3 solution). Figure S9 and Table S1: Correlations between the Hammett constants (σp) of substituent (R2) and the chemical shifts of NH or OH signals of compounds 3ac and 3eg. Figure S10: Additional 1H NMR spectra for compound 3b upon acid addition and irradiation. Figures S11 and S13: 1H NMR spectra for compounds 3ac upon base addition. Figure S12: 1H NMR spectrum of DBU and TFA mixture. Figures S14 and S15: Hirshfeld surfaces and energy frameworks calculated with CrystalExplorer software. Figure S16 and Table S2: UV–Vis spectroscopy data of compounds 3ac, 3eg. Figures S17–S22: CV curves of samples CM-1a and CM-3a in neutral and acidic electrolytes at various scan speeds (PDF).

Author Contributions

Conceptualization, N.B.; formal analysis, A.G.; investigation, A.G., S.B., R.D., Ņ.G. and A.Z.; writing—original draft preparation, A.G., S.B., R.D., A.Z. and N.B.; writing—review and editing, A.G. and N.B.; visualization, A.G., R.D., Ņ.G. and A.Z.; supervision, N.B.; funding acquisition, A.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been supported by the European Social Fund within project no 8.2.2.0/20/I/008, «Strengthening of PhD students and academic personnel of Riga Technical University and BA School of Business and Finance in the strategic fields of specialization» of the Specific Objective 8.2.2, «To Strengthen Academic Staff of Higher Education Institutions in Strategic Specialization Areas» of the Operational Programme «Growth and Employment». This research/publication was supported by Riga Technical University’s Doctoral Grant program (DOK.LĶI/23).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article or Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Cao, P.; Bracun, L.; Yamagata, A.; Christianson, B.M.; Negami, T.; Zou, B.; Terada, T.; Canniffe, D.P.; Shirouzu, M.; Li, M.; et al. Structural basis for the assembly and quinone transport mechanisms of the dimeric photosynthetic RC–LH1 supercomplex. Nat. Commun. 2022, 13, 1977. [Google Scholar] [CrossRef] [PubMed]
  2. Gutiérrez-Fernández, J.; Kaszuba, K.; Minhas, G.S.; Baradaran, R.; Tambalo, M.; Gallagher, D.T.; Sazanov, L.A. Key role of quinone in the mechanism of respiratory complex I. Nat. Commun. 2020, 11, 4135. [Google Scholar] [CrossRef] [PubMed]
  3. Bolton, J.L.; Dunlap, T. Formation and Biological Targets of Quinones: Cytotoxic versus Cytoprotective Effects. Chem. Res. Toxicol. 2017, 30, 13–37. [Google Scholar] [CrossRef]
  4. Mansha, M.; Anam, A.; Khan, S.A.; Alzahrani, A.S.; Khan, M.; Ahmad, A.; Arshad, M.; Ali, S. Recent Developments on Electroactive Organic Electrolytes for Non-Aqueous Redox Flow Batteries: Current Status, Challenges, and Prospects. Chem. Rec. 2023, 24, e202300233. [Google Scholar] [CrossRef] [PubMed]
  5. Go, C.Y.; Shin, J.; Choi, M.K.; Jung, I.H.; Kim, K.C. Switchable Design of Redox-Enhanced Nonaromatic Quinones Enabled by Conjugation Recovery. Adv. Mater. 2023, 2311155. [Google Scholar] [CrossRef] [PubMed]
  6. Simon, P.; Gogotsi, Y. Perspectives for electrochemical capacitors and related devices. Nat. Mater. 2020, 19, 1151–1163. [Google Scholar] [CrossRef] [PubMed]
  7. Shea, J.J.; Luo, C. Organic Electrode Materials for Metal Ion Batteries. ACS Appl. Mater. Interfaces 2020, 12, 5361–5380. [Google Scholar] [CrossRef]
  8. Jethwa, R.B.; Hey, D.; Kerber, R.N.; Bond, A.D.; Wright, D.S.; Grey, C.P. Exploring the Landscape of Heterocyclic Quinones for Redox Flow Batteries. ACS Appl. Energy Mater. 2023, 7, 414–426. [Google Scholar] [CrossRef] [PubMed]
  9. Blanc, L.E.; Kundu, D.; Nazar, L.F. Scientific Challenges for the Implementation of Zn-Ion Batteries. Joule 2020, 4, 771–799. [Google Scholar] [CrossRef]
  10. Tsao, Y.; Lee, M.; Miller, E.C.; Gao, G.; Park, J.; Chen, S.; Katsumata, T.; Tran, H.; Wang, L.-W.; Toney, M.F.; et al. Designing a Quinone-Based Redox Mediator to Facilitate Li2S Oxidation in Li-S Batteries. Joule 2019, 3, 872–884. [Google Scholar] [CrossRef]
  11. Sieuw, L.; Jouhara, A.; Quarez, É.; Auger, C.; Gohy, J.-F.; Poizot, P.; Vlad, A. A H-bond stabilized quinone electrode material for Li–organic batteries: The strength of weak bonds. Chem. Sci. 2018, 10, 418–426. [Google Scholar] [CrossRef] [PubMed]
  12. Sugumaran, M. Reactivities of Quinone Methides versus o-Quinones in Catecholamine Metabolism and Eumelanin Biosynthesis. Int. J. Mol. Sci. 2016, 17, 1576. [Google Scholar] [CrossRef] [PubMed]
  13. Poddel’Sky, A.I.; Druzhkov, N.O.; Fukin, G.K.; Cherkasov, V.K.; Abakumov, G.A. Bifunctional iminopyridino-catechol and its o-quinone: Synthesis and investigation of coordination abilities. Polyhedron 2017, 124, 41–50. [Google Scholar] [CrossRef]
  14. Astaf’eva, T.V.; Arsenyev, M.V.; Rumyantcev, R.V.; Fukin, G.K.; Cherkasov, V.K.; Poddel’sky, A.I. Imine-Based Catechols and o-Benzoquinones: Synthesis, Structure, and Features of Redox Behavior. ACS Omega 2020, 5, 22179–22191. [Google Scholar] [CrossRef]
  15. Ribeiro, R.C.B.; Ferreira, P.G.; Borges, A.D.A.; Forezi, L.D.S.M.; Silva, F.D.C.D.; Ferreira, V.F. 1,2-Naphthoquinone-4-sulfonic acid salts in organic synthesis. Beilstein J. Org. Chem. 2022, 18, 53–69. [Google Scholar] [CrossRef]
  16. Wu, Z.; Liu, Q.; Yang, P.; Chen, H.; Zhang, Q.; Li, S.; Tang, Y.; Zhang, S. Molecular and Morphological Engineering of Organic Electrode Materials for Electrochemical Energy Storage. Electrochem. Energy Rev. 2022, 5, 1–67. [Google Scholar] [CrossRef]
  17. Batenko, N.; Belyakov, S.; Kiselovs, G.; Valters, R. Synthesis of 6,7-dichloropyrido[1,2-a]benzimidazole-8,9-dione and its analogues and their reactions with nucleophiles. Tetrahedron Lett. 2013, 54, 4697–4699. [Google Scholar] [CrossRef]
  18. Gaile, A.; Belyakov, S.; Turovska, B.; Batenko, N. Synthesis of Asymmetric Coupled Polymethines Based on a 7-Chloropyrido[1,2-a]benzimidazole-8,9-dione Core. J. Org. Chem. 2022, 87, 2345–2355. [Google Scholar] [CrossRef] [PubMed]
  19. Gaile, A.; Belyakov, S.; Rjabovs, V.; Mihailovs, I.; Turovska, B.; Batenko, N. Investigation of Weak Noncovalent Interactions Directed by the Amino Substituent of Pyrido- and Pyrimido-[1,2-a]benzimidazole-8,9-diones. ACS Omega 2023, 8, 40960–40971. [Google Scholar] [CrossRef]
  20. Yamada, T.; Yamashita, T.; Nakamura, M.; Shimamura, H.; Yamaguchi, A.; Takaya, M. Synthesis and Hemostatic Activity of 1, 2-Naphthoquinones. Yakugaku Zasshi 1980, 100, 799–806. [Google Scholar] [CrossRef]
  21. Carroll, F.I.; Miller, H.W.; Meck, R. Thiosemicarbazone and amidinohydrazone derivatives of some 1,4-naphthoquinones. J. Chem. Soc. C Org. 1970, 3, 1993–1996. [Google Scholar] [CrossRef]
  22. Dudley, K.H.; Miller, H.W.; Schneider, P.W.; McKee, R.L. Potential naphthoquinone antimalarials. 2-Acylhydrazino-1,4-naphthoquinones and related compounds. J. Org. Chem. 1969, 34, 2750–2755. [Google Scholar] [CrossRef]
  23. Smith, M.B.; March, J. March’s Advanced Organic Chemistry: Reactions, Mechanisms, and Structure, 6th ed.; John Wiley & Sons: Hoboken, NJ, USA, 2006; Volume 9780471720. [Google Scholar]
  24. Spackman, P.R.; Turner, M.J.; McKinnon, J.J.; Wolff, S.K.; Grimwood, D.J.; Jayatilaka, D.; Spackman, M.A. CrystalExplorer: A program for Hirshfeld surface analysis, visualization and quantitative analysis of molecular crystals. J. Appl. Crystallogr. 2021, 54, 1006–1011. [Google Scholar] [CrossRef] [PubMed]
  25. Turner, M.J.; Thomas, S.P.; Shi, M.W.; Jayatilaka, D.; Spackman, M.A. Energy frameworks: Insights into interaction anisotropy and the mechanical properties of molecular crystals. Chem. Commun. 2014, 51, 3735–3738. [Google Scholar] [CrossRef] [PubMed]
  26. Catalán, J. Toward a Generalized Treatment of the Solvent Effect Based on Four Empirical Scales: Dipolarity (SdP, a New Scale), Polarizability (SP), Acidity (SA), and Basicity (SB) of the Medium. J. Phys. Chem. B 2009, 113, 5951–5960. [Google Scholar] [CrossRef] [PubMed]
  27. Shapet’Ko, N.N.; Shigorin, D.N. NMR study of intramolecular hydrogen bond protons in quinoid structures. J. Struct. Chem. 1968, 8, 474–476. [Google Scholar] [CrossRef]
  28. Kamlet, M.J.; Taft, R.W. The solvatochromic comparison method. I. The .beta.-scale of solvent hydrogen-bond acceptor (HBA) basicities. J. Am. Chem. Soc. 1976, 98, 377–383. [Google Scholar] [CrossRef]
  29. Hansch, C.; Leo, A.; Taft, R.W. A survey of Hammett substituent constants and resonance and field parameters. Chem. Rev. 1991, 91, 165–195. [Google Scholar] [CrossRef]
  30. Su, X.; Lõkov, M.; Kütt, A.; Leito, I.; Aprahamian, I. Unusual para-substituent effects on the intramolecular hydrogen-bond in hydrazone-based switches. Chem. Commun. 2012, 48, 10490–10492. [Google Scholar] [CrossRef]
  31. Su, X.; Aprahamian, I. Hydrazone-based switches, metallo-assemblies and sensors. Chem. Soc. Rev. 2014, 43, 1963–1981. [Google Scholar] [CrossRef]
  32. Johnson, J.E.; Morales, N.M.; Gorczyca, A.M.; Dolliver, D.D.; McAllister, M.A. Mechanisms of Acid-Catalyzed Z/E Isomerization of Imines. J. Org. Chem. 2001, 66, 7979–7985. [Google Scholar] [CrossRef] [PubMed]
  33. Su, X.; Aprahamian, I. Switching Around Two Axles: Controlling the Configuration and Conformation of a Hydrazone-Based Switch. Org. Lett. 2010, 13, 30–33. [Google Scholar] [CrossRef] [PubMed]
  34. Ryabchun, A.; Li, Q.; Lancia, F.; Aprahamian, I.; Katsonis, N. Shape-Persistent Actuators from Hydrazone Photoswitches. J. Am. Chem. Soc. 2019, 141, 1196–1200. [Google Scholar] [CrossRef]
  35. Kim, J.; Ling, J.; Lai, Y.; Milner, P.J. Redox-Active Organic Materials: From Energy Storage to Redox Catalysis. ACS Mater. Au 2024. [Google Scholar] [CrossRef]
  36. Wang, M.; Dong, X.; Escobar, I.C.; Cheng, Y.-T. Lithium Ion Battery Electrodes Made Using Dimethyl Sulfoxide (DMSO)—A Green Solvent. ACS Sustain. Chem. Eng. 2020, 8, 11046–11051. [Google Scholar] [CrossRef]
  37. Laurence, C.; Legros, J.; Chantzis, A.; Planchat, A.; Jacquemin, D. A Database of Dispersion-Induction DI, Electrostatic ES, and Hydrogen Bonding α1 and β1 Solvent Parameters and Some Applications to the Multiparameter Correlation Analysis of Solvent Effects. J. Phys. Chem. B 2015, 119, 3174–3184. [Google Scholar] [CrossRef]
  38. Matsui, M. Polymethine Dyes. In Progress in the Science of Functional Dyes; Ooyama, Y., Yagi, S., Eds.; Springer: Singapore, Singapore, 2021; pp. 3–19. [Google Scholar] [CrossRef]
  39. Miran, M.S.; Kinoshita, H.; Yasuda, T.; Susan, A.B.H.; Watanabe, M. Hydrogen bonds in protic ionic liquids and their correlation with physicochemical properties. Chem. Commun. 2011, 47, 12676–12678. [Google Scholar] [CrossRef] [PubMed]
  40. Almeida, R.G.; Valença, W.O.; Rosa, L.G.; de Simone, C.A.; de Castro, S.L.; Barbosa, J.M.C.; Pinheiro, D.P.; Paier, C.R.K.; de Carvalho, G.G.C.; Pessoa, C.; et al. Synthesis of quinone imine and sulphur-containing compounds with antitumor and trypanocidal activities: Redox and biological implications. RSC Med. Chem. 2020, 11, 1145–1160. [Google Scholar] [CrossRef] [PubMed]
  41. Klopčič, I.; Dolenc, M.S. Chemicals and Drugs Forming Reactive Quinone and Quinone Imine Metabolites. Chem. Res. Toxicol. 2018, 32, 1–34. [Google Scholar] [CrossRef]
  42. Batenko, N.; Kricka, A.; Belyakov, S.; Turovska, B.; Valters, R. A novel method for the synthesis of benzimidazole-based 1,4-quinone derivatives. Tetrahedron Lett. 2015, 57, 292–295. [Google Scholar] [CrossRef]
  43. Smith, E.; Dent, G. Modern Raman Spectroscopy—A Practical Approach; John Wiley & Sons: Chichester, UK, 2004. [Google Scholar] [CrossRef]
  44. Kudin, K.N.; Ozbas, B.; Schniepp, H.C.; Prud’Homme, R.K.; Aksay, I.A.; Car, R. Raman Spectra of Graphite Oxide and Functionalized Graphene Sheets. Nano Lett. 2007, 8, 36–41. [Google Scholar] [CrossRef] [PubMed]
  45. Saravanan, M.; Ganesan, M.; Ambalavanan, S. An in situ generated carbon as integrated conductive additive for hierarchical negative plate of lead-acid battery. J. Power Sources 2014, 251, 20–29. [Google Scholar] [CrossRef]
  46. Gottlieb, H.E.; Kotlyar, V.; Nudelman, A. NMR Chemical Shifts of Common Laboratory Solvents as Trace Impurities. J. Org. Chem. 1997, 62, 7512–7515. [Google Scholar] [CrossRef] [PubMed]
  47. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  48. Bourhis, L.J.; Dolomanov, O.V.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. The anatomy of a comprehensive constrained, restrained refinement program for the modern computing environment–Olex2dissected. Acta Crystallogr. Sect. A Found. Adv. 2015, 71, 59–75. [Google Scholar] [CrossRef]
  49. 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.; et al. Mercury 4.0: From visualization to analysis, design and prediction. J. Appl. Crystallogr. 2020, 53, 226–235. [Google Scholar] [CrossRef]
Scheme 1. (a) Synthesis of compounds 3ag, atom numbering of heterocyclic p-quinone imine fragment of compounds 3ag is shown in blue; (b) p-Quinone imine and benzamide fragments of obtained compounds 3ag.
Scheme 1. (a) Synthesis of compounds 3ag, atom numbering of heterocyclic p-quinone imine fragment of compounds 3ag is shown in blue; (b) p-Quinone imine and benzamide fragments of obtained compounds 3ag.
Molecules 29 01613 sch001
Figure 1. (a) ORTEP diagram of compound 3a showing thermal ellipsoids at the 50% probability level; (b) two mesomeric structures of compound 3a based on the bond distances in the crystal structure.
Figure 1. (a) ORTEP diagram of compound 3a showing thermal ellipsoids at the 50% probability level; (b) two mesomeric structures of compound 3a based on the bond distances in the crystal structure.
Molecules 29 01613 g001
Figure 2. (a) A dimer (highlighted) formed in the crystal structure of compound 3a. (b) Crystal packing of compound 3a along the a axis. (c) Crystal packing of compound 3a along the c axis. Color map: C, grey; N, blue; O, red; Cl, green; H, white.
Figure 2. (a) A dimer (highlighted) formed in the crystal structure of compound 3a. (b) Crystal packing of compound 3a along the a axis. (c) Crystal packing of compound 3a along the c axis. Color map: C, grey; N, blue; O, red; Cl, green; H, white.
Molecules 29 01613 g002
Figure 3. An expansion of 1H NMR spectra of compound 3b in DMSO-d6 solution upon sequential addition of base (DBU) and acid (TFA).
Figure 3. An expansion of 1H NMR spectra of compound 3b in DMSO-d6 solution upon sequential addition of base (DBU) and acid (TFA).
Molecules 29 01613 g003
Figure 4. 1H NMR titration of compound 3b in DMSO-d6 solution with DBU (0–8 equivalents).
Figure 4. 1H NMR titration of compound 3b in DMSO-d6 solution with DBU (0–8 equivalents).
Molecules 29 01613 g004
Figure 5. (a) UV–Vis absorption spectra of neutral (in DMSO or DCM solution) and deprotonated (upon addition base to the DMSO or DCM solution) forms of compound 3a. (b) UV–Vis absorption spectra neutral (in DMSO solution) and deprotonated (upon addition of a base to the DMSO solution) forms of compounds 3a and 3c.
Figure 5. (a) UV–Vis absorption spectra of neutral (in DMSO or DCM solution) and deprotonated (upon addition base to the DMSO or DCM solution) forms of compound 3a. (b) UV–Vis absorption spectra neutral (in DMSO solution) and deprotonated (upon addition of a base to the DMSO solution) forms of compounds 3a and 3c.
Molecules 29 01613 g005
Figure 6. CV results at varying scanning speeds for samples CM-1a, CM-3a, and substrate in neutral (0.5 M K2SO4) electrolyte and in acidic (0.5 M H2SO4) electrolyte.
Figure 6. CV results at varying scanning speeds for samples CM-1a, CM-3a, and substrate in neutral (0.5 M K2SO4) electrolyte and in acidic (0.5 M H2SO4) electrolyte.
Molecules 29 01613 g006
Figure 7. (a) Raman spectra of pure compound 3a (green line), prepared cathode CM-3a before (blue line) and after cycling in acidic (red line) and neutral (gray line) electrolytes. (b) Raman spectra of pure compound 1a (green line), prepared cathode CM-1a before (blue line) and after cycling in acidic (red line) and neutral (gray line) electrolytes.
Figure 7. (a) Raman spectra of pure compound 3a (green line), prepared cathode CM-3a before (blue line) and after cycling in acidic (red line) and neutral (gray line) electrolytes. (b) Raman spectra of pure compound 1a (green line), prepared cathode CM-1a before (blue line) and after cycling in acidic (red line) and neutral (gray line) electrolytes.
Molecules 29 01613 g007
Figure 8. Scanning electron microscopy images of samples CM-1a, CM-3a, and substrate (coating without active material) before and after CV measurements in neutral and acidic electrolytes (magnification of ×2500).
Figure 8. Scanning electron microscopy images of samples CM-1a, CM-3a, and substrate (coating without active material) before and after CV measurements in neutral and acidic electrolytes (magnification of ×2500).
Molecules 29 01613 g008
Table 1. Crystal data and structure refinement details for compound 3a.
Table 1. Crystal data and structure refinement details for compound 3a.
Crystal ParameterCompound 3a
Empirical formulaC18H11ClN4O3
Calculated density (g/cm3)1.556
Formula weight366.766
ColorRed
Size/mm30.18 × 0.03 × 0.01
Temperature/K150.0 (1)
Crystal systemmonoclinic
Space groupP21/n
a5.65994 (5)
b14.90948 (19)
c18.5815 (2)
α/°90
β/°93.1950 (9)
γ/°90
V31565.60 (3)
Wavelength/Å1.54184
Radiation typeCu Kα
Absorption coefficient (mm−1)2.419
θmin/°3.8
2θmax/°155.0
Measured reflections17875
Number of independent reflections3322
Reflections with I ≥ 2σ(I)3089
Rint0.0327
Number of refined parameters243
Restraints0
Largest peak0.3476
Deepest hole−0.3143
Goodness of fit1.0362
wR2(all data)0.0962
wR20.0945
R1(all data)0.0367
R10.0346
CCDC deposition number2238663
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gaile, A.; Belyakov, S.; Dūrena, R.; Griščenko, Ņ.; Zukuls, A.; Batenko, N. Studies of the Functionalized α-Hydroxy-p-Quinone Imine Derivatives Stabilized by Intramolecular Hydrogen Bond. Molecules 2024, 29, 1613. https://doi.org/10.3390/molecules29071613

AMA Style

Gaile A, Belyakov S, Dūrena R, Griščenko Ņ, Zukuls A, Batenko N. Studies of the Functionalized α-Hydroxy-p-Quinone Imine Derivatives Stabilized by Intramolecular Hydrogen Bond. Molecules. 2024; 29(7):1613. https://doi.org/10.3390/molecules29071613

Chicago/Turabian Style

Gaile, Anastasija, Sergey Belyakov, Ramona Dūrena, Ņikita Griščenko, Anzelms Zukuls, and Nelli Batenko. 2024. "Studies of the Functionalized α-Hydroxy-p-Quinone Imine Derivatives Stabilized by Intramolecular Hydrogen Bond" Molecules 29, no. 7: 1613. https://doi.org/10.3390/molecules29071613

Article Metrics

Back to TopTop