Next Article in Journal
Reactivity of Antibodies Immobilized on Gold Nanoparticles: Fluorescence Quenching Study
Previous Article in Journal
Poly(levodopa)-Modified β-(1 → 3)-D-Glucan Hydrogel Enriched with Triangle-Shaped Nanoparticles as a Biosafe Matrix with Enhanced Antibacterial Potential
Previous Article in Special Issue
Highly Engineered Cr-In/H-SSZ-39 Catalyst for Enhanced Performance in CH4-SCR of NOx
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Cobalt Hydroxychloride and Its Application as a Catalyst in the Condensation of Perimidines

by
Cássio Siqueira
1,†,
Gabriela R. Borges
2,†,
Fernanda S. Portela
1,
Maria E. Miks
1,
Felipe F. Marques
3,
Gleison A. Casagrande
2,
Sumbal Saba
3,*,
Rafael Marangoni
1,*,
Jamal Rafique
2,3,* and
Giancarlo V. Botteselle
1
1
Laboratory of Organic Synthesis and Catalysis (LabSOC), Midwestern Parana State University-UNICENTRO, Guarapuava 85040-167, PR, Brazil
2
Instituto de Química (INQUI), Universidade Federal do Mato Grosso do Sul-UFMS, Campo Grande 79074-460, MS, Brazil
3
Laboratory of Sustainable Synthesis and Organochalcogen (LabSO), Instituto de Química (IQ), Universidade Federal de Goiás–UFG, Goiânia 74690-900, GO, Brazil
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2026, 31(1), 182; https://doi.org/10.3390/molecules31010182
Submission received: 16 August 2025 / Revised: 13 September 2025 / Accepted: 23 September 2025 / Published: 4 January 2026
(This article belongs to the Special Issue Heterogeneous Catalysis for Sustainability and Carbon-Neutrality)

Abstract

Herein, we report the synthesis, characterization, and catalytic evaluation of cobalt hydroxide chloride [Co2(OH)3Cl] in the solvent-free synthesis of 2-substituted 2,3-dihydroperimidines. The presented method aligns with several green chemistry principles, offering operational simplicity, purification by recrystallization, no by-product formation, high yields (64–99%), and short reaction times. A total of 16 dihydroperimidines were synthesized to demonstrate substrate scope versatility. Additionally, the catalyst was successfully recycled and reused in multiple cycles without significant loss. Its robustness was further confirmed by gram-scale synthesis, achieving an 89% yield.

Graphical Abstract

1. Introduction

Cobalt-based hydroxychlorides are an important class of inorganic materials that have attracted significant interest due to their unique structural features and various technological uses [1]. Among these, cobalt hydroxychloride Co2(OH)3Cl is particularly notable for its structural resemblance to minerals of the atacamite group, such as paratacamite [Cu3(Cu,M)Cl2(OH)6, where M = Zn, Mg, Ni] and their synthetic analogues [2,3].
The atacamite group is characterized by complex three-dimensional framework structures made of interconnected metal octahedra [2,4]. These structures feature sheets of edge-sharing MCl2(OH)4 octahedra connected through interlayer metal sites coordinated by six OH ligands, forming M-OH-M bridges that create a strong three-dimensional network rather than a true layered structure [2,5]. The structural similarity between Co2(OH)3Cl and paratacamite-type compounds suggests that cobalt hydroxychloride likely adopts a similar three-dimensional framework, with Co2+ ions occupying both the sheet-forming and interlayer octahedral sites.
Metal hydroxysalts, whether layered or framework-structured, have found applications in environmental sustainability [6], energy storage [7], water treatment [8], drug delivery [9,10], photocatalysis [11,12] and catalysis [12,13] due to their unique properties such as layered structure, memory effect, selective ion exchange, high surface area, tuneable bandgap, and the presence of Brønsted–Lowry and Lewis acid/base sites [6,14,15]. Furthermore, these materials offer advantages such as high stability, low cost, low toxicity, facile synthesis, and reusability [6].
Among layered metal hydroxy salts, layered double hydroxides (LDHs) are well-established catalysts for diverse organic transformations. These include the synthesis of xanthenes, 1,4-dihydropyrimidines [16], fused pyrimidines [17], β-nitroalcohols [18], 2-aryl benzimidazoles, benzothiazoles, benzoxazoles [19], and oxidative amidation reactions [20]. In contrast, the catalytic applications of layered hydroxy salts (LHSs) in organic synthesis remain scarce, limited primarily to biodiesel esterification [21,22] and click reactions for synthesizing 1,2,3-triazoles [23]. Notably, the use of specific metal hydroxysalts, such as Co2(OH)3Cl, is, to our knowledge, absent from the literature.
Perimidines are a class of N-heterocyclic compounds with notable pharmacological potential, exhibiting antioxidants [24], antimicrobial [25,26], anticancer [27,28] and anti-inflammatory properties [29,30]. Their synthesis usually involves a condensation reaction between 1,8-diaminonaphthalene and aldehydes, often catalyzed by Brønsted–Lowry [31] or Lewis’s acid catalysts [32,33].
In this context, cobalt-based catalysts have demonstrated efficiency in organic synthesis [34,35,36,37], and some have been explored for perimidines synthesis. However, existing methods frequently require calcination [38] or ligand complexation followed by pyrolysis [39], highlighting the need for simpler, more sustainable catalytic systems. Therefore, the screening for simpler cobalt catalysts remains an area of interest. Furthermore, in recent years, cobalt-catalyzed solvent-free reactions have gained significant attention in the scientific community due to their alignment with the principles of sustainable chemistry [40,41,42,43].
To the best of our knowledge, the application of cobalt as a catalyst in the synthesis of 2-substituted 2,3-dihydroperimidines via condensation reaction has not yet been explored. Thus, in connection with our continuing interest in designing new synthetic methodologies, transition metal-catalysis, as well as biologically active N-heterocyclic compounds are explored [44,45,46,47,48,49]; herein, we present a comprehensive investigation of Co2(OH)3Cl synthesis via urea-mediated hydrolysis, focusing on the structural characterization, phase identification, and property evaluation of the resulting material as a potential catalyst for 2,3-dihydro-1H-peridimines condensation reaction. Through detailed X-ray diffraction analysis, we establish the structural relationship between Co2(OH)3Cl and the paratacamite family, providing insights into the 3D framework that underpins its unique properties.

2. Results and Discussion

2.1. Cobalt Hydroxychloride [Co2(OH)3Cl] Characterization

2.1.1. Structural and Phase Formation Characterization

The material’s structure was determined using the crystallographic chart COD2310848, corresponding to a hydroxy salt with the general formula Co2(OH)3Cl. It crystallizes in a hexagonal system (space group R-3m), with unit cell parameters a = b = 6.84 Å, c = 14.50 Å, and angles α = β = 90°, γ = 120°.
The XRD pattern in Figure 1A shows an intense reflection at the (101) plane at 16° (2θ), deviating from the typical (00l) pattern of classical brucite-like layered hydroxide salts (LHS), instead suggesting a paratacamite-type structure. This unique diffraction arises from monoatomic layers linked by octahedra, featuring 25% vacancies at alternating cobalt ion sites. These cations move to interlayer positions and are accompanied by a diagonal shift (a/√3) along the (120), which enhances the {101} planes [50].
Paratacamite is a polymorph of atacamite, and both display intense (101) reflections. However, they are distinguishable by their crystal symmetry: atacamite-like structures crystallize in an orthorhombic system (Pnma), which cannot transform into a hexagonal framework [51]. Conversely, as previously noted, paratacamite adopts a rhombohedral lattice (R-3m), compatible with the hexagonal system, as observed in the Co2(OH)3Cl material [51]. This structural arrangement further confirms that Cl ions are incorporated within the crystal lattice rather than in the interlayer region, distinguishing it from traditional layered hydroxy salt [50].
Furthermore, the XRD pattern shown in Figure 1B (after catalytic use) demonstrates that the crystalline framework remains largely intact after the perimidine condensation reaction. However, notable changes in the (101) peak profile, including both reduced intensity and increased peak broadening (full width at half maximum (FWHM), increasing from 0.124° to 0.301°), indicate structural modifications within the material. While several factors could contribute to these observations (including preferential orientation effects, surface modifications, adsorption of organic molecules, or localized disorder at interlayer sites), the (101) plane’s sensitivity to interlayer arrangements suggests that structural perturbations may have occurred at Co–OH–Co bridging regions. The combination of intensity reduction and peak broadening is consistent with the introduction of structural irregularities at these interlayer sites, which represent the most accessible regions for organic molecule interactions during catalysis.

2.1.2. Vibrational Spectroscopy Analysis

Attenuated total reflectance Fourier transform infrared spectroscopy (ATR-FTIR) analysis (Figure 2) reveals a sharp absorption band at 3550 cm−1 corresponding to the stretching vibration of non-hydrogen-bonded OH, characteristic of brucite-like layered material [52,53]. The bands observed at 838 and 697 cm−1 are assigned to Co-O stretching vibrations, and the signal at 416 cm−1 is attributed to the bending mode of Co-OH bonds, respectively [53,54].
A minor absorption band at 2369 cm−1 suggests the presence of adsorbed atmospheric CO2 [53], and the band at 1350 cm−1 may be related to residual urea from the synthesis, usually indicating the presence of carbonate species (CO32−) [52,55] interacting or incorporated into the LHS structure. Notably, no signals corresponding to H2O bound to the material were detected, suggesting enhanced availability of Brønsted–Lowry acidic and basic active sites due to the absence of competing water interactions.

2.1.3. Electronic Behaviour

The diffuse reflectance UV-Vis spectrum (Figure 3) exhibits an absorption band at 416 nm, attributed to a ligand-to-metal charge transfer (LMCT) from O2− (hydroxyl groups) to Co2+, consistent with observations in related transition metal LDH systems, where bands below 480 nm typically indicate surface octahedral M–OH sites [56]. In addition, the bands between 510 and 548 nm are assigned to the 4T1g4T1g(P) and 4A2g(F) transitions of octahedral coordinated Co2+ (Oh) [54,55]. Moreover, the absence of a band around 700 nm confirms that no Co3+ is present on octahedral sites [56,57,58], and similarly, the lack of bands near 650 nm indicates that Co2+ in tetrahedral sites is not present in the structure [55,56,57].
The material’s pink coloration is characteristic of β-Co(OH)2 brucite-like octahedral systems. Colorimetric analysis (CIE L*a*b*) yielded values of L* = 63.93, a* = 18.79, and b* = −10.97, with a chroma (C = 21.76)* and hue angle (h = 329.72°)* further defining its optical properties [58,59].

2.1.4. Morphological, Surface, and Chemical Analysis

SEM micrographs presented in Figure 4 reveal morphological changes in the Co2(OH)3Cl samples before (a) and after (b) perimidine catalysis. The post-catalytic sample exhibits a notable fragmentation pattern, with the original agglomerates breaking down into smaller plate-shaped particles, suggesting mechanical stress or structural reorganization during the catalytic process.
Complementary TEM analysis (Figure 4c,d) confirms the preservation of crystalline order at the nanoscale level, with well-defined atomic plane coherence maintained throughout the material. The inset images showing individual crystallites demonstrate that despite the morphological fragmentation observed by SEM, the particles retain their crystalline structure with well-organized lattice fringes, indicating that the structural modifications are localized rather than involving complete crystalline breakdown.
This behaviour is beneficial for catalyst use because it provides more contact area, which offers more active sites. Conversely, the results from the B.E.T. analysis on the surface area and porosity of the material show that Co2(OH)3Cl has a low surface area of 3.717 m2/g. Regarding porosity, it has a total pore volume of 0.0056 cm3/g and a pore size of 1.809 nm, classified as microporous based on IUPAC standards. Elemental analysis by EDS showed 55% Co, 28% O, and 17% Cl, aligning with the composition of Co2(OH)3Cl with a paratacamite-type structure [50].

2.2. Synthesis of 2-Substituted 2,3-dihydro-1H-perimidines

The model reaction for optimizing reaction conditions was conducted between 1,8-diaminonaphthalene (1) and benzaldehyde (2a), with variations in catalyst type, temperature, and catalyst loading (Table 1). Initially, a study was carried out to determine the best catalyst, from a variety of cobalt hydroxide salts, to be used in the reaction for the synthesis of perimidine 3a, under solvent-free conditions, 1 mol% of catalyst, at 70 °C for 5 min (Table 1, entries 1–3). Thus, it was observed that the counter ion used in the salt had a significant influence on the reaction yield, since when sulphate and nitrate were used, the product 3a was obtained in 67 and 86% yield (Table 1, entries 1 and 2), respectively, while when chloride [Co2(OH)3Cl] was used the product was obtained in 96% yield (Table 1, entry 3). The influence of temperature was highly significant. Lowering the temperature to 60 °C drastically reduced the yield of 3a to 30% (Table 1, entry 4). This is likely due to the partial fusion of 1,8-diaminonaphthalene, which has a reported melting point of 65 °C, hindering reactant mobility. Finally, an increase in catalyst load did not have a significant influence on product yield (Table 1, entry 5).
With optimized reaction conditions in hand, next we explored the generality and scope of this catalytic system. For this purpose, a wide range of aldehydes was employed, totaling 16 examples to evaluate the scope of the synthetic route, as presented in Table 2.
The reaction of unsubstituted benzaldehyde (2a) afforded the desired perimidine product (3a) in 96% yield (Table 2, entry 1). The introduction of methyl substituents showed interesting effects: ortho-tolualdehyde (2b) gave a quantitative yield (99%) of perimidine (3b), while para-tolualdehyde (2c) provided a slightly lower yield of 84% (Table 2, entries 2–3). Notably, cinnamaldehyde (2d), featuring a vinylic group attached to the aromatic ring, proved to be an excellent substrate, yielding perimidine (3d) in 97% (Table 2, entry 4). The catalyst system showed good compatibility with halogenated substrates. Bromo-substituted aldehydes at both meta- (2e) and para- (2f) positions gave the corresponding perimidines (3e–3f) in 88% and 84% yields, respectively (Table 2, entries 5–6). Similarly, chloro-substituted analogues at ortho (2g) and para (2h) positions afforded products (3g–3h) in comparable yields of 84% and 85% (Table 2, entries 7–8). Oxygen-containing substituents were particularly effective, with para-hydroxybenzaldehyde (2i) and para-anisaldehyde (2j) yielding perimidines (3i–3j) in excellent yields of 94% and 85%, respectively (Table 2, entries 9–10). The methodology extended successfully to heteroaromatic aldehydes. Pyridine-carboxaldehydes with varying nitrogen positions (ortho 2k, meta 2l, para 2m) gave products (3k–3m) in 85%, 93%, and 88% yields (Table 2, entries 11–13). Other heterocyclic substrates, including thiophene-2-carboxaldehyde (2n) and furfural (2o), provided the expected perimidines (3n–3o), albeit with somewhat lower yields of 80% and 64% (Table 2, entries 14–15). The system also accommodated more complex substrates, as evidenced by the 88% yield obtained for the indole-containing perimidine (3p) (Table 2, entry 16).
The recyclability of the Co2(OH)3Cl catalyst was evaluated using the model reaction between 1,8-diaminonaphthalene (1) and benzaldehyde (2a) to produce perimidine (3a) (Table 3). Initial attempts to recover the catalyst by simple filtration proved ineffective due to the small mass of catalyst employed (1 mol%). To overcome this challenge, we implemented a centrifugation protocol prior to recrystallization, washing the organic phase with ethanol (3 × 5 mL) at 3000 rpm for 2 min. This approach successfully recovered the catalyst while maintaining good catalytic performance. The recycled catalyst demonstrated consistent efficiency for two cycles, yielding perimidine (3a) in approximately 90% yield for each run. However, a noticeable decrease in activity was observed in the third cycle, with the yield dropping to 71% (Table 3). This reduction in efficiency suggests potential catalyst degradation or loss during the recovery process.
To evaluate the practical applicability of this methodology, we performed a gram-scale synthesis using 1,8-diaminonaphthalene (1) (5.0 mmol, 791 mg) and benzaldehyde (2a) (5.0 mmol, 510 μL) in the presence of 5 mol% Co2(OH)3Cl catalyst (0.05 mmol, 51 mg). As illustrated in Figure 5, this scaled-up reaction successfully afforded perimidine 3a in 89% yield (1.09 g), confirming the robustness and scalability of the catalytic system. The maintained high yield at increased scale underscores the method’s potential for practical synthetic applications.
As previously discussed, condensation reactions with perimidine commonly employ acid catalysts [32,33]. In this context, both water [60] and glycerol [61] have been employed as reaction media for perimidine synthesis, where solvent hydroxyl groups facilitate aldehyde activation. Building on these observations, we propose that the surface hydroxyl groups of Co2(OH)3Cl serve as the primary catalytic sites in our system. These hydroxyl groups are also proposed as bifunctional centres, due to the amphoteric behaviour in hydroxylated oxides [62]. As illustrated in Figure 6, these sites likely coordinate with 1,8-diaminonaphthalene, thereby promoting nucleophilic addition to the activated aldehyde. Furthermore, surface Co2+ centres may stabilize the imine intermediate, consistent with previous reports on cobalt-based catalysts [63,64]. Supporting this mechanism, the UV-Vis spectrum reveals a ligand-to-metal charge transfer (LMCT) band at 416 nm, confirming the presence of strongly basic O2− sites [65]. These sites are expected to preferentially coordinate with 1,8-diaminonaphthalene, enhancing reactant interaction and facilitating the condensation process.
The catalytic cycle begins with aldehyde activation through coordination to acidic hydroxyl sites on the Co2(OH)3Cl surface (I), followed by proton transfer from the nucleophilic amine to the carbonyl oxygen (II). The activated 1,8-diaminonaphthalene then attacks the electrophilic carbonyl carbon, forming a Schiff base intermediate with concomitant water elimination (III). T he liberated water molecule remains coordinated to the catalyst structure. The resulting imine intermediate is stabilized through two potential pathways: (a) coordination to surface Co2+ Lewis acid sites or (b) hydrogen bonding with surface hydroxyl groups, which generates a partial positive charge resembling an iminium ion (IV). This enhanced electrophilicity facilitates intramolecular cyclization via nucleophilic attack by the second amine group, ultimately yielding the perimidine product (V), Figure 6.

3. Materials and Methods

3.1. Reagents and Apparatus

3.1.1. Reagents and Characterization for [Co2(OH)3Cl]

All reagents were commercially obtained to synthesize cobalt hydroxychloride [Co2(OH)3Cl]. The material’s structure and phase were studied using X-ray diffraction measurements performed on a Bruker diffractometer XRD-D2 Phaser, with a copper source (λ = 1.5418 Å), covering 3° to 70° (2θ), with 0.2°/s increments. Infrared spectra recorded in attenuated total reflectance (ATR) mode were obtained with a PerkinElmer Frontier spectrophotometer, in the 4000–650 cm−1 range. The optical diffuse absorbance was measured on the powder form of the compound using a Varian UV-VIS-NIR spectrophotometer CARYb5G in the 400–800 nm range. Colorimetric parameters were analyzed with a portable colorimeter (NR60CP-3nh), equipped with a D65 light source and an 8 mm measuring aperture. The surface area, pore size, and volume were determined through nitrogen adsorption analysis (B.E.T.), using an Anton Paar Nova 800 instrument. The sample was analyzed in a 9 mm cell and vacuum degassed for 4 h at 180 °C. Transmission Electron Microscope (TEM) images were obtained in a JEOL JEM-2100, Tokyo, Japan, equipped with an Energy-Dispersive X-ray Spectrometer (EDS).

3.1.2. Reagents and Characterization for 2,3-Dihydro-1H-perimidines

The reagents 1,8-diaminonaphthalene, aldehydes, and solvents were commercially acquired. The Nuclear Magnetic Resonance (NMR) spectra were obtained on a Bruker AVANCE NEO-500 and Bruker DPX-300 spectrometers using CDCl3 as the solvent for Hydrogen and Carbon (1H 500 and 300 MHz and 13C 125 and 75 MHz NMR). All chemical shifts were reported in parts per million (ppm), with the 1H spectrum referenced to 0.00 ppm (TMS) and the 13C spectrum referenced to 77.16 ppm (CDCl3). The thin-layer chromatography (TLC) plates utilized were GF254 silica with a thickness of 0.20 mm, from the brand Macherey-Nagel, and the visualization methods used were iodine chamber and ultraviolet light (254 nm).

3.2. Experimental Procedures

3.2.1. Cobalt Hydroxichloride [Co2(OH)3Cl] Solid State Synthesis

The cobalt hydroxichloride [Co2(OH)3Cl] was produced through a solid-state reaction. The decomposition of urea is necessary to supply the hydroxide ions for synthesis, as explained by Rajamathi (2001) [59]. Therefore, 9.0 g of cobalt (II) chloride hexahydrate (CoCl2·6H2O) and 2.0 g of urea were dissolved in 2 mL of distilled water in a hermetically sealed reaction flask, then heated at 110 °C for four hours in an oven. After cooling to room temperature, the resulting material was washed with distilled water and centrifuged five times to remove impurities. The solid was then dried in an oven at 40 °C for five days. Afterwards, the sample was characterized using XRD, FTIR, TEM/EDS, and UV-Vis.

3.2.2. Synthesis of 2-Substituted 2,3-dihydro-1H-permidine

In a 5 mL test tube, 79.1 mg (0.5 mmol) of 1,8-diaminonaphthalene 1 and Co2(OH)3Cl (1 mol%, 1.0 mg) were added, heated until 70 °C, and left under stirring until the reaction mixture had fused. Subsequently, the respective aldehyde 2a-p (0.5 mmol) was added. The reaction progress was monitored by TLC until complete consumption of the starting materials.
Following the reaction completion, the organic phase was solubilized in ethanol (10 mL) and filtered to separate the Co2(OH)3Cl catalyst. The product in the organic phase was then recrystallized in a cold-water bath (15 mL). Finally, the precipitate was filtered and dried at an ambient temperature. The described spectra data are presented below, and the spectra are found in the Supplementary Materials (Figures S1–S32).
2-phenyl-2,3-dihydro-1H-perimidine (3a) [66]: 118 mg (96%), 1H NMR (300 MHz, CDCl3) δ 7.66–7.27 (m, 9H), 6.54 (d, J = 6.1 Hz, 2H), 5.47 (s, 1H), 4.54 (s, 1H). 13C NMR (75 MHz, CDCl3) δ 142.23, 140.21, 135.02, 129.72, 128.96, 128.03, 127.00, 117.99, 113.57, 105.95, 68.50.
2-(o-tolyl)-2,3-dihydro-1H-perimidine (3b) [66]: 129 mg (99%), 1H NMR (300 MHz, CDCl3) δ 7.74(s, 1H), 7.23 (m, 7H), 6.51 (s, 2H), 5.71 (s, 1H), 4.43 (s, 2H), 2.50 (s, 3H). 13C NMR (75 MHz, CDCl3) δ 142.54, 137.66, 136.66, 135.11, 131.06, 129.20, 128.21, 126.96, 126.76, 117.97, 113.71, 106.09, 65.27, 19.25.
2-(p-tolyl)-2,3-dihydro-1H-perimidine (3c) [66]: 109 mg (84%), 1H NMR (300 MHz, C6D6) δ 7.50 (d, J = 6.9, 2H), 7.24–7.22 (m, 6H), 6.49 (d, J = 6.3, 2H), 5.41 (s, 1H), 4.49 (s, 2H), 2.39 (s, 3H). 13C NMR (75 MHz, CDCl3) δ 142.35, 139.61, 137.34, 135.04, 129.59, 127.90, 126.99, 117.92, 113.59, 105.90, 68.31, 21.41.
(E)-2-styryl-2,3-dihydro-1H-perimidine (3d) [67]: 132 mg (97%), 1H NMR (300 MHz, CDCl3) δ 7.35–7.19 (m, 9H), 6.69 (d, J = 15.7 Hz, 1H), 6.48 (d, J = 7.3, 2H), 6.34 (dd, J = 15.6, 7.6 Hz, 1H), 5.00 (d, J = 7.3, 2H), 4.42 (s, 2H). 13C NMR (75 MHz, CDCl3) δ 141.19, 135.74, 134.86, 134.67, 128.81, 128.58, 128.06, 127.61, 127.02, 117.87, 113.54, 106.16, 67.08.
2-(3-bromophenyl)-2,3-dihydro-1H-perimidine (3e) [56]: 142 mg (88%), 1H NMR (300 MHz, CDCl3) δ 7.82 (s, 1H), 7.55 (s, 1H), 7.25 (m, 6H), 6.54 (s, 2H), 5.42 (s, 1H), 4.43 (s, 2H). 13C NMR (75 MHz, CDCl3) δ 142.48, 141.78, 134.97, 132.82, 131.27, 130.53, 127.03, 126.70, 122.97, 118.29, 113.52, 106.17, 67.80.
2-(4-bromophenyl)-2,3-dihydro-1H-perimidine (3f) [66]: 136 mg (84%), 1H NMR (500 MHz, CDCl3) δ 7.52 (s, 2H), 7.44 (s, 2H), 7.21 (m, 4H), 6.48 (s, 2H), 5.34 (s, 1H), 4.42 (s, 2H). 13C NMR (126 MHz, CDCl3) δ 141.84, 139.21, 134.93, 132.07, 129.71, 127.00, 123.66, 118.19, 113.47, 106.08, 67.78.
2-(2-chlorophenyl)-2,3-dihydro-1H-perimidine (3g) [66]: 118 mg (84%), 1H NMR (300 MHz, CDCl3) δ 7.82 (s, 1H), 7,41–7.24 (m, 7H), 6.57 (s, 2H), 5.98 (s, 1H), 4.60 (s, 2H). 13C NMR (75 MHz, CDCl3) δ 141.69, 137.78, 134.95, 133.33, 130.29, 129.84, 129.17, 127.77, 127.06, 118.16, 113.42, 106.24, 63.94.
2-(4-chlorophenyl)-2,3-dihydro-1H-perimidine (3h) [66]: 119 mg (85%),1H NMR (300 MHz, CDCl3) δ 7.54 (d, J = 8.2 Hz, 1H), 7.40 (d, J = 8.2 Hz, 1H), 7.28–7.23 (m, 4H), 6.50 (d, J = 6.6 Hz, 2H), 5.40 (s, 1h), 4.45 (s, 2H). 13C NMR (75 MHz, CDCl3) δ 141.91, 138.74, 135.50, 134.96, 129.44, 129.15, 127.01, 118.21, 113.51, 106.09, 67.78.
4-(2,3-dihydro-1H-perimidin-2-yl)phenol (3i) [67]: 123 mg (94%), 1H NMR (300 MHz, CDCl3) δ 7.50 (d, J = 7.9 Hz, 1H), 7.26–7.23 (m, 4H), 6.88 (d, J = 7.9 Hz, 2H), 6.52 (d, J = 6.4 Hz, 2H), 5.40 (s, 1H), 4.49 (s, 2H). 13C NMR (75 MHz, CDCl3) δ 156.96, 142.41, 135.07, 132.42, 129.51, 127.02, 117.99, 115.73, 113.62, 105.93, 68.11.
2-(4-methoxyphenyl)-2,3-dihydro-1H-perimidine (3j) [66]: 117 mg (85%), 1H NMR (300 MHz, CDCl3) δ 7.54 (d, J = 8.4 Hz, 2H), 7.27–7.18 (m, 4H), 6.94 (d, J = 8.4 Hz, 2H), 6.49 (d, J = 6.3 Hz, 2H) 5.39 (s, 1H), 4.48 (s, 2H), 3.83 (s, 3H). 13C NMR (75 MHz, CDCl3) δ 160.67, 142.42, 135.05, 132.43, 129.28, 127.00, 117.91, 114.24, 113.57, 105.85, 68.06, 55.52.
2-(pyridin-2-yl)-2,3-dihydro-1H-perimidine (3k) [68]: 105 mg (85%), 1H NMR (500 MHz, CDCl3) δ 8.54 (d, J = 4.8 Hz, 1H), 7.58 (td, J = 7.7, 1.4 Hz, 1H)., 7.50 (d, J = 7.9 Hz, 1H), 7.23–7.15 (m, 5H), 6.53 (d, J = 7.1 Hz, 2H), 5.51 (s, 1H), 4.98 (s, 2H). 13C NMR (126 MHz, CDCl3) δ 159.30, 149.27, 137.25, 134.75, 126.96, 123.59, 120.80, 118.00, 113.96, 106.62, 67.65.
3-(pyridin-2-yl)-2,3-dihydro-1H-perimidine (3l): 115 mg (93%), 1H NMR (500 MHz, CDCl3) δ 8.70 (s, 1H), 8.62 (d, J = 4.0 Hz, 1H), 7.95 (dd, J = 9.7, 1.8 Hz, 1H), 7.33 (dd, J = 7.8, 4.9 Hz, 1H), 7.24–7.21 (m, 4H), 6.51 (dd, J = 6.6, 1.6 Hz, 2H), 5.42 (s, 1H), 4.58 (s, 2H). 13C NMR (126 MHz, CDCl3) δ 150.94, 149.49, 141.68, 135.97, 134.87, 126.99, 124.03, 118.53, 118.32, 113.44, 106.20, 66.13.
4-(pyridin-2-yl)-2,3-dihydro-1H-perimidine (3m): 109 mg (88%), 1H NMR (500 MHz, CDCl3) δ 8.69 (d, J = 5.9 Hz, 1H), 7.55 (d, J = 5.9 Hz, 1H), 7.28–7.25 (m, 3H), 6.57 (dd, J = 6.4, 1.8 Hz, 2H), 5.48 (s 1H), 4.53 (s, 2H). 13C NMR (126 MHz, CDCl3) δ 150.66, 148.89, 141.19, 134.92, 127.07, 122.72, 118.60, 113.62, 106.52, 67.19.
2-(thiophen-2-yl)-2,3-dihydro-1H-perimidine (3n) [68,69]: 101 mg (80%),1H NMR (500 MHz, CDCl3) δ 7.38 (s, 1H), 7.25–7.23 (m, 5H), 7.02 (s, 1H), 6.53 (s, 2H), 5.80 (s, 1H), 4.65 (s, 2H). 13C NMR (126 MHz, CDCl3) δ 144.15, 141.47, 134.94, 127.24, 127.02, 126.56, 126.45, 118.33, 113.82, 106.29, 63.90.
2-(furan-2-yl)-2,3-dihydro-1H-perimidine (3o) [70,71]: 76 mg (64%),1H NMR (500 MHz, CDCl3) δ 7.42 (s, 1H), 7.26–7.22 (m, 4H), 6.59 (d, J = 6.6 Hz, 2H), 6.42 (s, 1H), 6.36 (s, 1H), 5.65 (s, 1H), 4.70 (s, 2H). 13C NMR (126 MHz, CDCl3) δ 153.54, 142.81, 140.83, 134.88, 127.01, 118.44, 114.04, 110.66, 107.79, 106.79, 61.68.
3-(1H-indol-3-yl)-2,3-dihydro-1H-perimidine (3p): 125 mg (88%), 1H NMR (500 MHz, CDCl3) δ 8.25 (s, 1H), 7.90 (d, J = 7.9 Hz, 1H), 7.32 (d, J = 8.2 Hz, 1H), 7.26–7.08 (m, 7H), 6.47 (d, J = 7.9 Hz, 1H), 5.79 (s, 1H), 4.60 (s, 2H). 13C NMR (126 MHz, CDCl3) δ 142.78, 136.32, 135.06, 127.02, 125.62, 123.76, 122.64, 120.07, 119.89, 117.69, 115.11, 113.85, 111.61, 105.91, 62.17.

3.2.3. Co2(OH)3Cl Recycling

In a 50 mL round-bottom flask, 1,8-diaminonaphthalene 1 (0.5 mmol) and 1 mol% of catalyst (1.0 mg) were added. The mixture was melted while stirring, and then benzaldehyde 2a (0.5 mmol) was added. TLC confirmed the reaction finished after 5 min. The crude product was then dissolved in ethanol (5 mL), transferred to a Falcon tube, centrifuged for 2 min at 3000 rpm, and the catalyst was dried either in an oven at 40 °C for 2 h or under vacuum using a rotary evaporator. The organic phase was recrystallized with water (20 mL) in an ice bath, and this process was repeated three times.

3.2.4. Scale-Up Synthesis

In a 50 mL round-bottom flask, a magnetic stir bar was added, followed by 1,8-diaminonaphthalene 1 (5.0 mmol, 792 mg) and Co2(OH)3Cl (5 mol%, 51 mg). Then, after the medium was melted and under stirring, benzaldehyde 2a (5.0 mmol, 510 µL) was added.
Afterwards, after 5 min, the reaction was complete, and the round-bottom flask was removed from the temperature; completion was confirmed through TLC. The crude product was dissolved in 10 mL of ethanol, filtered to remove the catalyst, and finally recrystallized in 15 mL of cold water to produce product 3a.

4. Conclusions

A facile solid-state synthesis was developed for the preparation of the layered hydroxy salt Co2(OH)3Cl, utilizing urea and minimal water (2 mL). Comprehensive characterization by XRD, FTIR, UV-Vis spectroscopy, and SEM/EDS confirmed the formation of a paratacamite-type structure with well-defined crystallinity and morphology. The catalytic potential of Co2(OH)3Cl was successfully demonstrated in the efficient synthesis of perimidines, which are pharmaceutically relevant heterocycles. Under optimized conditions (70 °C, solvent-free, 1 mol% catalyst, 5 min), the reaction exhibited a broad substrate scope, accommodating 16 diverse aryl aldehydes with yields ranging from 64% to 99%. Notably, the catalyst retained high activity (~90% yield) over two consecutive cycles, confirming its reusability and structural stability. Furthermore, the methodology proved scalable, delivering 1.09 g (89% yield) of perimidine in a gram-scale reaction, highlighting its potential for practical applications.
This work presents Co2(OH)3Cl as a sustainable, efficient, and reusable catalyst for the rapid synthesis of perimidines under mild conditions, offering a promising alternative to conventional acid-catalyzed approaches.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules31010182/s1. NMR Spectra (Figures S1–S32).

Author Contributions

Conceptualization, S.S., R.M., J.R. and G.V.B.; Methodology, C.S., G.R.B., F.F.M., F.S.P. and M.E.M.; Formal analysis, G.A.C., R.M., F.F.M. and S.S.; Investigation, C.S., F.S.P., J.R. and G.V.B.; Resources, S.S., R.M. and G.V.B.; Data curation, G.R.B., F.S.P. and M.E.M.; Writing—original draft preparation, C.S. and G.V.B.; Writing—review and editing, J.R. and R.M.; Visualization, G.A.C. and S.S.; Supervision, J.R. and G.V.B.; Project administration, J.R. and G.V.B.; Funding acquisition, S.S., R.M., J.R. and G.V.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES, Finance Code 001), and Fundação Araucária for financial support and fellowships. C.S., G.R.B., F.S.P., and M.EM., would like to thank the scientific research scholarship of CAPES. This study was financed in part by the Universidade Federal de Mato Grosso do Sul- Brasil (UFMS)—Finance Code 001. G.V.B. is grateful to CNPq for funding this research project (Grant number 04172/2023-7). S.S., and J.R. would like to acknowledge CNPq (401355/2025-0, 316687/2023-5, 309975/2022-0, 404172/2023-7, and 405655/2023-1). S.S. also acknowledges the following FAPEG public calls: Chamada Pública FAPEG/SES Nº 18/2025 (ARB2025191000003), and Chamada Pública FAPEG Nº 05/2025 (PVE2025041000055).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are provided in the article.

Acknowledgments

The authors thank the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES), and Fundação Universidade Federal de Mato Grosso do Sul (UFMS) for the support offered in this research. Authors also acknowledge CRTI-Universidade Federal de Goiás (UFG) for the TEM analysis.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Park, G.D.; Ko, Y.N.; Kang, Y.C. Electrochemical Properties of Cobalt Hydroxychloride Microspheres as a New Anode Material for Li-Ion Batteries. Sci. Rep. 2014, 4, 5785. [Google Scholar] [CrossRef] [PubMed]
  2. Kampf, A.R.; Sciberras, M.J.; Leverett, P.; Williams, P.A.; Malcherek, T.; Schlüter, J.; Welch, M.D.; Dini, M.; Molina Donoso, A.A. Paratacamite-(Mg), Cu3 (Mg,Cu)Cl2 (OH)6; a New Substituted Basic Copper Chloride Mineral from Camerones, Chile. Mineral. Mag. 2013, 77, 3113–3124. [Google Scholar] [CrossRef]
  3. Kampf, A.R.; Sciberras, M.J.; Williams, P.A.; Dini, M.; Molina Donoso, A.A. Leverettite from the Torrecillas Mine, Iquique Provence, Chile: The Co-Analogue of Herbertsmithite. Mineral. Mag. 2013, 77, 3047–3054. [Google Scholar] [CrossRef]
  4. Welch, M.D.; Sciberras, M.J.; Williams, P.A.; Leverett, P.; Schlüter, J.; Malcherek, T. A Temperature-Induced Reversible Transformation between Paratacamite and Herbertsmithite. Phys. Chem. Miner. 2014, 41, 33–48. [Google Scholar] [CrossRef]
  5. Braithwaite, R.S.W.; Mereiter, K.; Paar, W.H.; Clark, A.M. Herbertsmithite, Cu3 Zn(OH)6 Cl2, a New Species, and the Definition of Paratacamite. Mineral. Mag. 2004, 68, 527–539. [Google Scholar] [CrossRef]
  6. Kumari, S.; Sharma, A.; Kumar, S.; Thakur, A.; Thakur, R.; Bhatia, S.K.; Sharma, A.K. Multifaceted Potential Applicability of Hydrotalcite-Type Anionic Clays from Green Chemistry to Environmental Sustainability. Chemosphere 2022, 306, 135464. [Google Scholar] [CrossRef]
  7. Hu, J.; Tang, X.; Dai, Q.; Liu, Z.; Zhang, H.; Zheng, A.; Yuan, Z.; Li, X. Layered Double Hydroxide Membrane with High Hydroxide Conductivity and Ion Selectivity for Energy Storage Device. Nat. Commun. 2021, 12, 3409. [Google Scholar] [CrossRef]
  8. Morais de Faria, J.; Alkimin Muniz, L.; Netto, J.F.Z.; Scheres Firak, D.; B. De Sousa, F.; da Silva Lisboa, F. Application of a Hybrid Material Formed by Layered Zinc Hydroxide Chloride Modified with Spiropyran in the Adsorption of Ca2+ from Water. Colloids Surf. A Physicochem. Eng. Asp. 2021, 631, 127738. [Google Scholar] [CrossRef]
  9. Kaassis, A.Y.A.; Al-Jamal, W.T.; Strimaite, M.; Severic, M.; Williams, G.R. Biocompatible Hydroxy Double Salts as Delivery Matrices for Non-Steroidal Anti-Inflammatory and Anti-Epileptic Drugs. Appl. Clay. Sci. 2022, 221, 106456. [Google Scholar] [CrossRef]
  10. Kaassis, A.Y.A.; Xu, S.-M.; Guan, S.; Evans, D.G.; Wei, M.; Williams, G.R. Hydroxy Double Salts Loaded with Bioactive Ions: Synthesis, Intercalation Mechanisms, and Functional Performance. J. Solid State Chem. 2016, 238, 129–138. [Google Scholar] [CrossRef]
  11. Ramalho, R.R., Jr.; Oliveira, A.F.; de Andrade, S.J.; da Silva Lisboa, F. Congo Red Dye Degradation by the Composite Zn(II)/Co(II) Layered Double Hydroxy Salt/Peanut Shell Biochar as Photocatalyst. Mater. Today Commun. 2025, 44, 111924. [Google Scholar] [CrossRef]
  12. Chuaicham, C.; Sekar, K.; Balakumar, V.; Zhang, L.; Trakulmututa, J.; Smith, S.M.; Sasaki, K. Fabrication of Hydrotalcite-like Copper Hydroxyl Salts as a Photocatalyst and Adsorbent for Hexavalent Chromium Removal. Minerals 2022, 12, 182. [Google Scholar] [CrossRef]
  13. Centi, G.; Perathoner, S. Catalysis by Layered Materials: A Review. Microporous Mesoporous Mater. 2008, 107, 3–15. [Google Scholar] [CrossRef]
  14. Arizaga, G.; Satyanarayana, K.; Wypych, F. Layered Hydroxide Salts: Synthesis, Properties and Potential Applications. Solid State Ion. 2007, 178, 1143–1162. [Google Scholar] [CrossRef]
  15. Jijoe, P.S.; Yashas, S.R.; Shivaraju, H.P. Fundamentals, Synthesis, Characterization and Environmental Applications of Layered Double Hydroxides: A Review. Environ. Chem. Lett. 2021, 19, 2643–2661. [Google Scholar] [CrossRef]
  16. Rathee, G.; Kohli, S.; Singh, N.; Awasthi, A.; Chandra, R. Calcined Layered Double Hydroxides: Catalysts for Xanthene, 1,4-Dihydropyridine, and Polyhydroquinoline Derivative Synthesis. ACS Omega 2020, 5, 15673–15680. [Google Scholar] [CrossRef]
  17. Sahu, P.K. Eco-Friendly Grinding Synthesis of a Double-Layered Nanomaterial and the Correlation between Its Basicity, Calcination and Catalytic Activity in the Green Synthesis of Novel Fused Pyrimidines. RSC Adv. 2016, 6, 78409–78423. [Google Scholar] [CrossRef]
  18. Bharali, D.; Devi, R.; Bharali, P.; Deka, R.C. Synthesis of High Surface Area Mixed Metal Oxide from the NiMgAl LDH Precursor for Nitro-Aldol Condensation Reaction. New J. Chem. 2015, 39, 172–178. [Google Scholar] [CrossRef]
  19. Sahu, P.K. A Green Approach to the Synthesis of a Nano Catalyst and the Role of Basicity, Calcination, Catalytic Activity and Aging in the Green Synthesis of 2-Aryl Bezimidazoles, Benzothiazoles and Benzoxazoles. RSC Adv. 2017, 7, 42000–42012. [Google Scholar] [CrossRef]
  20. Gupta, S.S.R.; Nakhate, A.V.; Rasal, K.B.; Deshmukh, G.P.; Mannepalli, L.K. Oxidative Amidation of Benzaldehydes and Benzylamines with N -Substituted Formamides over a Co/Al Hydrotalcite-Derived Catalyst. New J. Chem. 2017, 41, 15268–15276. [Google Scholar] [CrossRef]
  21. Cordeiro, C.S.; da Silva, F.R.; Marangoni, R.; Wypych, F.; Ramos, L.P. LDHs Instability in Esterification Reactions and Their Conversion to Catalytically Active Layered Carboxylates. Catal. Lett. 2012, 142, 763–770. [Google Scholar] [CrossRef]
  22. Nakagaki, S.; Machado, G.S.; Stival, J.F.; Henrique dos Santos, E.; Silva, G.M.; Wypych, F. Natural and Synthetic Layered Hydroxide Salts (LHS): Recent Advances and Application Perspectives Emphasizing Catalysis. Prog. Solid State Chem. 2021, 64, 100335. [Google Scholar] [CrossRef]
  23. Marangoni, R.; Carvalho, R.E.; Machado, M.V.; Dos Santos, V.B.; Saba, S.; Botteselle, G.V.; Rafique, J. Layered Copper Hydroxide Salts as Catalyst for the “Click” Reaction and Their Application in Methyl Orange Photocatalytic Discoloration. Catalysts 2023, 13, 426. [Google Scholar] [CrossRef]
  24. Azam, M.; Warad, I.; Al-Resayes, S.; Zahin, M.; Ahmad, I.; Shakir, M. Syntheses, Physico-Chemical Studies and Antioxidant Activities of Transition Metal Complexes with a Perimidine Ligand. Z Anorg. Allg. Chem. 2012, 638, 881–886. [Google Scholar] [CrossRef]
  25. Farghaly, T.A.; Abdallah, M.A.; Muhammad, Z.A. New 2-Heterocyclic Perimidines: Synthesis and Antimicrobial Activity. Res. Chem. Intermed. 2015, 41, 3937–3947. [Google Scholar] [CrossRef]
  26. Abu-Melha, S. Confirmed Mechanism for 1,8-Diaminonaphthalene and Ethyl Aroylpyrovate Derivatives Reaction, DFT/B3LYP, and Antimicrobial Activity of the Products. J. Chem. 2018, 2018, 4086824. [Google Scholar] [CrossRef]
  27. Kumar, A.; Banerjee, S.; Roy, P.; Sondhi, S.M.; Sharma, A. Solvent-Free Synthesis and Anticancer Activity Evaluation of Benzimidazole and Perimidine Derivatives. Mol. Divers. 2018, 22, 113–127. [Google Scholar] [CrossRef] [PubMed]
  28. Eldeab, H.A.; Eweas, A.F. A Greener Approach Synthesis and Docking Studies of Perimidine Derivatives as Potential Anticancer Agents. J. Heterocycl. Chem. 2018, 55, 431–439. [Google Scholar] [CrossRef]
  29. Bassyouni, F.A.; Abu-Bakr, S.M.; Hegab, K.H.; El-Eraky, W.; El Beih, A.A.; Rehim, M.E.A. Synthesis of New Transition Metal Complexes of 1H-Perimidine Derivatives Having Antimicrobial and Anti-Inflammatory Activities. Res. Chem. Intermed. 2012, 38, 1527–1550. [Google Scholar] [CrossRef]
  30. Zhang, H.-J.; Wang, X.-Z.; Cao, Q.; Gong, G.-H.; Quan, Z.-S. Design, Synthesis, Anti-Inflammatory Activity, and Molecular Docking Studies of Perimidine Derivatives Containing Triazole. Bioorg. Med. Chem. Lett. 2017, 27, 4409–4414. [Google Scholar] [CrossRef]
  31. Khopkar, S.; Shankarling, G. Squaric Acid: An Impressive Organocatalyst for the Synthesis of Biologically Relevant 2,3-Dihydro-1H-Perimidines in Water. J. Chem. Sci. 2020, 132, 31. [Google Scholar] [CrossRef]
  32. Mobinikhaledi, A.; Steel, P.J. Synthesis of Perimidines Using Copper Nitrate as an Efficient Catalyst. Synth. React. Inorg. Met.-Org. Nano-Met. Chem. 2009, 39, 133–135. [Google Scholar] [CrossRef]
  33. Behbahani, F.K.; Golchin, F.M. A New Catalyst for the Synthesis of 2-Substituted Perimidines Catalysed by FePO4. J. Taibah Univ. Sci. 2017, 11, 85–89. [Google Scholar] [CrossRef]
  34. Michiyuki, T.; Komeyama, K. Recent Advances in Four-Coordinated Planar Cobalt Catalysis in Organic Synthesis. Asian J. Org. Chem. 2020, 9, 343–358. [Google Scholar] [CrossRef]
  35. Tohidi, M.M.; Paymard, B.; Vasquez-García, S.R.; Fernández-Quiroz, D. Recent Progress in Applications of Cobalt Catalysts in Organic Reactions. Tetrahedron 2023, 136, 133352. [Google Scholar] [CrossRef]
  36. Liandi, A.R.; Cahyana, A.H.; Kusumah, A.J.F.; Lupitasari, A.; Alfariza, D.N.; Nuraini, R.; Sari, R.W.; Kusumasari, F.C. Recent Trends of Spinel Ferrites (MFe2O4: Mn, Co, Ni, Cu, Zn) Applications as an Environmentally Friendly Catalyst in Multicomponent Reactions: A Review. Case Stud. Chem. Environ. Eng. 2023, 7, 100303. [Google Scholar] [CrossRef]
  37. Grigolo, T.A.; de Campos, S.D.; Manarin, F.; Botteselle, G.V.; Brandão, P.; Amaral, A.A.; de Campos, E.A. Catalytic Properties of a Cobalt Metal–Organic Framework with a Zwitterionic Ligand Synthesized in Situ. Dalton Trans. 2017, 46, 15698–15703. [Google Scholar] [CrossRef] [PubMed]
  38. Retizlaf, A.; de Souza Sikora, M.; Ivashita, F.F.; Schneider, R.; Botteselle, G.V.; Junior, H.E.Z. CoFe2O4 Magnetic Nanoparticles: Synthesis by Thermal Decomposition of 8-Hydroxyquinolinates, Characterization, and Application in Catalysis. MRS Commun. 2023, 13, 567–573. [Google Scholar] [CrossRef]
  39. Schwob, T.; Ade, M.; Kempe, R. A Cobalt Catalyst Permits the Direct Hydrogenative Synthesis of 1 H-Perimidines from a Dinitroarene and an Aldehyde. ChemSusChem 2019, 12, 3013–3017. [Google Scholar] [CrossRef]
  40. Zong, Y.; Yang, L.; Tang, S.; Li, L.; Wang, W.; Yuan, B.; Yang, G. Highly Efficient Acetalization and Ketalization Catalyzed by Cobaloxime under Solvent-Free Condition. Catalyst 2018, 8, 48. [Google Scholar] [CrossRef]
  41. Dobras, G.; Kasperczyk, K.; Jurczhyk, S.; Orlinska, B. N-Hydroxyphthalimide Supported on Silica Coated with Ionic Liquids Containing CoCl2 (SCILLs) as New Catalytic System for Solvent-Free Ethylbenzene Oxidation. Catalysts 2020, 10, 252. [Google Scholar] [CrossRef]
  42. Alamier, W.M.; El-Telbani, E.M.; Syed, I.S.; Bakry, A.M. Cobalt Ferrite Nanoparticles: Highly Efficient Catalysts for the Biginelli Reaction. Ceramics 2025, 8, 102. [Google Scholar] [CrossRef]
  43. Javad, S.; Hossein, B.S.; Shiva, D.K. Cobalt Nanoparticles Promoted Highly Efficient One Pot Four-Component Synthesis of 1,4-Dihydropyridines under Solvent-Free Conditions. Chin. J. Catal. 2011, 32, 1850–1855. [Google Scholar] [CrossRef]
  44. Tornquist, B.L.; Bueno, G.P.; Manzano Willing, J.C.; de Oliveira, I.M.; Stefani, H.A.; Rafique, J.; Saba, S.; Almeida Iglesias, B.; Botteselle, G.V.; Manarin, F. Ytterbium (III) triflate/sodium dodecyl sulfate: A versatile recyclable and water-tolerant catalyst for the synthesis of bis(indolyl)methanes (BIMs). ChemistrySelect 2018, 3, 6358–6363. [Google Scholar] [CrossRef]
  45. Rocha, M.S.T.; Rafique, J.; Saba, S.; Azeredo, J.B.; Back, D.; Godoi, M.; Braga, A.L. Regioselective hydrothiolation of terminal acetylene catalyzed by magnetite (Fe3O4) nanoparticles. Synth. Commun. 2017, 47, 291–298. [Google Scholar] [CrossRef]
  46. Doerner, C.V.; Scheide, M.R.; Nicoleti, C.R.; Durigon, D.C.; Idiarte, V.D.; Sousa, M.J.A.; Mendes, S.R.; Saba, S.; Neto, J.S.S.; Martins, G.M.; et al. Versatile electrochemical synthesis of selenylbenzo[b]furan derivatives through the cyclization of 2-alkynylphenols. Front. Chem. 2022, 10, 880099. [Google Scholar] [CrossRef]
  47. Godoi, M.; Botteselle, G.V.; Rafique, J.; Rocha, M.S.T.; Pena, J.M.; Braga, A.L. Solvent-free Fmoc protection of amines under microwave irradiation. Asian J. Org. Chem. 2013, 2, 746–749. [Google Scholar] [CrossRef]
  48. Botteselle, G.V.; Elias, W.C.; Bettanin, L.; Canto, R.F.S.; Salin, D.N.O.; Barbosa, F.A.R.; Saba, S.; Gallardo, H.; Ciancaleoni, G.; Domingos, J.B.; et al. Catalytic Antioxidant Activity of Bis-Aniline-Derived Diselenides as GPx Mimics. Molecules 2021, 26, 4446. [Google Scholar] [CrossRef]
  49. Rafique, J.; Farias, G.; Saba, S.; Zapp, E.; Bellettini, I.C.; Momoli Salla, C.A.; Bechtold, I.H.; Scheide, M.R.; Santos Neto, J.S.; Monteiro de Souza Junior, D.; et al. Selenylated-oxadiazoles as promising DNA intercalators: Synthesis, electronic structure, DNA interaction and cleavage. Dyes Pigm. 2020, 180, 108519. [Google Scholar] [CrossRef] [PubMed]
  50. de Wolff, P.M. The Crystal Structure of Co2 (OH)3 Cl. Acta Crystallogr. 1953, 6, 359–360. [Google Scholar] [CrossRef]
  51. Malcherek, T.; Welch, M.D.; Williams, P.A. The Atacamite Family of Minerals—A Testbed for Quantum Spin Liquids. Acta Crystallogr. B Struct. Sci. Cryst. Eng. Mater. 2018, 74, 519–526. [Google Scholar] [CrossRef]
  52. Thomas, N.; Rajamathi, M. High Selectivity in Anion Exchange Reactions of the Anionic Clay, Cobalt Hydroxynitrate. J. Mater. Chem. 2011, 21, 18077. [Google Scholar] [CrossRef]
  53. Shinde, V.; Uthayakumar, M.; Karthick, R. Self-Assembled Cobalt Hydroxide Micro Flowers from Nanopetals: Structural, Fractal Analysis and Molecular Docking Study. Surf. Interfaces 2022, 32, 102163. [Google Scholar] [CrossRef]
  54. Diouane, Y.; Seijas-Da Silva, A.; Oestreicher, V.; Abellán, G. Exploring Pseudohalide Substitution in α-Cobalt-Based Layered Hydroxides. Dalton Trans. 2025, 54, 6538–6549. [Google Scholar] [CrossRef]
  55. Sanchis-Gual, R.; Hunt, D.; Jaramillo-Hernández, C.; Seijas-Da Silva, A.; Mizrahi, M.; Marini, C.; Oestreicher, V.; Abellán, G. Crystallographic and Geometrical Dependence of Water Oxidation Activity in Co-Based Layered Hydroxides. ACS Catal. 2023, 13, 10351–10363. [Google Scholar] [CrossRef]
  56. Velu, S.; Suzuki, K.; Hashimoto, S.; Satoh, N.; Ohashi, F.; Tomura, S. The Effect of Cobalt on the Structural Properties and Reducibility of CuCoZnAl Layered Double Hydroxides and Their Thermally Derived Mixed Oxides. J. Mater. Chem. 2001, 11, 2049–2060. [Google Scholar] [CrossRef]
  57. Ma, R.; Liu, Z.; Takada, K.; Fukuda, K.; Ebina, Y.; Bando, Y.; Sasaki, T. Tetrahedral Co(II) Coordination in α-Type Cobalt Hydroxide: Rietveld Refinement and X-Ray Absorption Spectroscopy. Inorg. Chem. 2006, 45, 3964–3969. [Google Scholar] [CrossRef]
  58. Neilson, J.R.; Schwenzer, B.; Seshadri, R.; Morse, D.E. Kinetic Control of Intralayer Cobalt Coordination in Layered Hydroxides: Co1−0.5 x oct Co x tet (OH)2 (Cl) x (H2O) n. Inorg. Chem. 2009, 48, 11017–11023. [Google Scholar] [CrossRef] [PubMed]
  59. Rajamathi, M.; Kamath, P.V. Urea Hydrolysis of Cobalt (II) Nitrate Melts: Synthesis of Novel Hydroxides and Hydroxynitrates. Int. J. Inorg. Mater. 2001, 3, 901–906. [Google Scholar] [CrossRef]
  60. Ann Harry, N.; Mary Cherian, R.; Radhika, S.; Anilkumar, G. A Novel Catalyst-Free, Eco-Friendly, on Water Protocol for the Synthesis of 2,3-Dihydro-1H-Perimidines. Tetrahedron. Lett. 2019, 60, 150946. [Google Scholar] [CrossRef]
  61. Siqueira, C.; Dos Santos, T.J.C.; Theisen, R.; Saba, S.; Rafique, J.; Gallina, A.L.; Botteselle, G.V. Direct Utilization of Waste Glycerol from Biodiesel Production for the Synthesis of Perimidines. Lett. Org. Chem. 2025, 22, 723–729. [Google Scholar] [CrossRef]
  62. Boehm, H.P. Acidic and Basic Properties of Hydroxylated Metal Oxide Surfaces. Discuss. Faraday. Soc. 1971, 52, 264. [Google Scholar] [CrossRef]
  63. Lee, P.-S.; Fujita, T.; Yoshikai, N. Cobalt-Catalyzed, Room-Temperature Addition of Aromatic Imines to Alkynes via Directed C–H Bond Activation. J. Am. Chem. Soc. 2011, 133, 17283–17295. [Google Scholar] [CrossRef]
  64. Harrowfield, J.M.; Sargeson, A.M.; Springborg, J.; Snow, M.R.; Taylor, D. Imine Formation and Stability and Interligand Condensation with Cobalt(III) 1,2-Ethanediamine Complexes. Inorg. Chem. 1983, 22, 186–193. [Google Scholar] [CrossRef]
  65. Cavani, F.; Trifirò, F.; Vaccari, A. Hydrotalcite-Type Anionic Clays: Preparation, Properties and Applications. Catal. Today 1991, 11, 173–301. [Google Scholar] [CrossRef]
  66. Zhang, B.; Li, J.; Zhu, H.; Xia, X.F.; Wang, D. Novel Recyclable Catalysts for Selective Synthesis of Substituted Perimidines and Aminopyrimidines. Catal. Lett. 2022, 153, 2388–2397. [Google Scholar] [CrossRef]
  67. Sahiba, N.; Sethiya, A.; Soni, J.; Agarwal, S. Metal Free Sulfonic Acid Functionalized Carbon Catalyst for Green and Mechanochemical Synthesis of Perimidines. ChemistrySelect 2020, 5, 13076–13080. [Google Scholar] [CrossRef]
  68. Das, K.; Mondal, A.; Pal, D.; Srivastava, H.K.; Srimani, D. Phosphine-Free Well-Defined Mn(I) Complex-Catalyzed Synthesis of Amine, Imine, and 2,3-Dihydro-1 H -Perimidine via Hydrogen Autotransfer or Acceptorless Dehydrogenative Coupling of Amine and Alcohol. Organometallics 2019, 38, 1815–1825. [Google Scholar] [CrossRef]
  69. Alama, M.; Lee, D.-U. Synthesis, spectroscopic and computational studies of 2-(thiophen-2-yl)-2,3-dihydro-1H-perimidine: An enzymes inhibition study. Comput. Biol. Chem. 2016, 64, 184–201. [Google Scholar] [CrossRef] [PubMed]
  70. Harry, N.A.; Radhika, S.; Neetha, M.; Anilkumar, G. A novel catalyst-free mechanochemical protocol for the synthesis of 2,3-dihydro-1H-perimidines. J. Heterocyclic. Chem. 2020, 57, 2037–2043. [Google Scholar] [CrossRef]
  71. Sadri, Z.; Behdbhani, F.K.; Keshmirizahe, E. Synthesis and Characterization of a Novel and Reusable Adenine Based Acidic Nanomagnetic Catalyst and Its Application in the Preparation of 2-Substituted-2,3-dihydro -1H-perimidines under Ultrasonic Irradiation and Solvent-Free Condition. Polycycl. Aromat. Comp. 2022, 43, 1898–1913. [Google Scholar] [CrossRef]
Figure 1. Diffractogram of Co2(OH)3Cl (A) as prepared (B) after application.
Figure 1. Diffractogram of Co2(OH)3Cl (A) as prepared (B) after application.
Molecules 31 00182 g001
Figure 2. Co2(OH)3Cl ATR-FTIR spectrum.
Figure 2. Co2(OH)3Cl ATR-FTIR spectrum.
Molecules 31 00182 g002
Figure 3. Electronic spectrum of Co2(OH)3Cl.
Figure 3. Electronic spectrum of Co2(OH)3Cl.
Molecules 31 00182 g003
Figure 4. Scanning electron microscopy (SEM), and transmission electron microscopy (TEM) of Co2(OH)3Cl, before (a,c), and after the catalytic process (b,d). Scale bars: 2 μm (SEM) and 5 nm (TEM main images) and 200 nm (TEM insets).
Figure 4. Scanning electron microscopy (SEM), and transmission electron microscopy (TEM) of Co2(OH)3Cl, before (a,c), and after the catalytic process (b,d). Scale bars: 2 μm (SEM) and 5 nm (TEM main images) and 200 nm (TEM insets).
Molecules 31 00182 g004
Figure 5. Gram-scale synthesis for perimidine 3a.
Figure 5. Gram-scale synthesis for perimidine 3a.
Molecules 31 00182 g005
Figure 6. Co2(OH)3Cl compound structure (A) and proposed mechanism of catalytic effect in 2,3-dihydro-1H-perimidine condensation (B).
Figure 6. Co2(OH)3Cl compound structure (A) and proposed mechanism of catalytic effect in 2,3-dihydro-1H-perimidine condensation (B).
Molecules 31 00182 g006
Table 1. Reaction parameters optimization a.
Table 1. Reaction parameters optimization a.
Molecules 31 00182 i001
EntryCatalyst (mol%)T (°C)Yield (%) b
1Co2(OH)3SO4 (1)7067
2Co2(OH)3NO3 (1)7086
3Co2(OH)3Cl (1)7096
4Co2(OH)3Cl (1)6030
5Co2(OH)3Cl (5)7094
a Reaction conditions: 1 (1 mmol), 2a (1 mmol), catalyst, temperature, 5 min. b Isolated yield.
Table 2. Synthesis of Co2(OH)3Cl-catalyzed perimidines 3a-p a.
Table 2. Synthesis of Co2(OH)3Cl-catalyzed perimidines 3a-p a.
Molecules 31 00182 i002
EntryAldehydeProductYield (%) b
1Molecules 31 00182 i003  2aMolecules 31 00182 i004  3a96
2Molecules 31 00182 i005  2bMolecules 31 00182 i006  3b99
3Molecules 31 00182 i007  2cMolecules 31 00182 i008  3c84
4Molecules 31 00182 i009  2dMolecules 31 00182 i010  3d97
5Molecules 31 00182 i011  2eMolecules 31 00182 i012  3e88
6Molecules 31 00182 i013  2fMolecules 31 00182 i014  3f84
7Molecules 31 00182 i015  2gMolecules 31 00182 i016  3g84
8Molecules 31 00182 i017  2hMolecules 31 00182 i018  3h85
9Molecules 31 00182 i019  2iMolecules 31 00182 i020  3i94
10Molecules 31 00182 i021  2jMolecules 31 00182 i022  3j85
11Molecules 31 00182 i023  2kMolecules 31 00182 i024  3k85
12Molecules 31 00182 i025  2lMolecules 31 00182 i026  3l93
13Molecules 31 00182 i027  2mMolecules 31 00182 i028  3m88
14Molecules 31 00182 i029  2nMolecules 31 00182 i030  3n80
15Molecules 31 00182 i031  2oMolecules 31 00182 i032  3o64
16Molecules 31 00182 i033  2pMolecules 31 00182 i034  3p88
a Reaction conditions: 1 (0.5 mmol), aldehyde (0.5 mmol), Co2(OH)3Cl (1 mol%), 70 °C, 5 min. b Isolated yield.
Table 3. Catalyst recycling a.
Table 3. Catalyst recycling a.
EntryCycleYield (%) b
196
290
371
a Reactions conditions: 1 (1 mmol), 2a (1 mmol), Co2(OH)3Cl (1 mol%), 70 °C, 5 min. b Isolated yield.
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

Siqueira, C.; Borges, G.R.; Portela, F.S.; Miks, M.E.; Marques, F.F.; Casagrande, G.A.; Saba, S.; Marangoni, R.; Rafique, J.; Botteselle, G.V. Synthesis of Cobalt Hydroxychloride and Its Application as a Catalyst in the Condensation of Perimidines. Molecules 2026, 31, 182. https://doi.org/10.3390/molecules31010182

AMA Style

Siqueira C, Borges GR, Portela FS, Miks ME, Marques FF, Casagrande GA, Saba S, Marangoni R, Rafique J, Botteselle GV. Synthesis of Cobalt Hydroxychloride and Its Application as a Catalyst in the Condensation of Perimidines. Molecules. 2026; 31(1):182. https://doi.org/10.3390/molecules31010182

Chicago/Turabian Style

Siqueira, Cássio, Gabriela R. Borges, Fernanda S. Portela, Maria E. Miks, Felipe F. Marques, Gleison A. Casagrande, Sumbal Saba, Rafael Marangoni, Jamal Rafique, and Giancarlo V. Botteselle. 2026. "Synthesis of Cobalt Hydroxychloride and Its Application as a Catalyst in the Condensation of Perimidines" Molecules 31, no. 1: 182. https://doi.org/10.3390/molecules31010182

APA Style

Siqueira, C., Borges, G. R., Portela, F. S., Miks, M. E., Marques, F. F., Casagrande, G. A., Saba, S., Marangoni, R., Rafique, J., & Botteselle, G. V. (2026). Synthesis of Cobalt Hydroxychloride and Its Application as a Catalyst in the Condensation of Perimidines. Molecules, 31(1), 182. https://doi.org/10.3390/molecules31010182

Article Metrics

Back to TopTop