Next Article in Journal
Enhanced Peroxydisulfate Activation by Co-Doping of Nitrogen, Chlorine, and Iron: Preparation, Synergistic Effects, and Application
Previous Article in Journal
Amorphous MoSx Nanosheets with Abundant Interlayer Dislocations for Enhanced Photolytic Hydrogen Evolution Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Carbon and Silica Supports Enhance the Durability and Catalytic Performance of Cobalt Oxides Derived from Cobalt Benzene-1,3,5-Tricarboxylate Complex

by
Hassan H. Hammud
1,*,
Waleed A. Aljamhi
1,
Kawther AlAbdullah
1,
Muhammad Humayun
2 and
Ihab Shawish
3
1
Department of Chemistry, College of Science, King Faisal University, P.O. Box 380, Al-Ahsa 31982, Saudi Arabia
2
Energy, Water and Environment Lab, College of Humanities and Sciences, Prince Sultan University, Riyadh 11586, Saudi Arabia
3
Department of Math and Sciences, College of Humanities and Sciences, Prince Sultan University, P.O. Box 66833, Riyadh 11586, Saudi Arabia
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(9), 881; https://doi.org/10.3390/catal15090881 (registering DOI)
Submission received: 31 July 2025 / Revised: 1 September 2025 / Accepted: 5 September 2025 / Published: 13 September 2025
(This article belongs to the Special Issue Environmental Catalysis and Nanomaterials for Water Pollution Control)

Abstract

Addressing the urgent need for robust and sustainable catalysts to detoxify nitroaromatic pollutants, this study introduces a novel approach for synthesizing cobalt oxide nanocomposites via pyrolysis of cobalt benzene-1,3,5-tricarboxylate. By integrating porous carbon (PC) and nano silica (NS) supports with Co3O4 to form (Co3O4/PC) and (Co3O4/NS), we achieved precise morphological control, as evidenced by SEM and TEM analysis. SEM revealed 80–500 nm Co3O4 microspheres, 300 nm Co3O4/PC microfibers, and 2–5 µm Co3O4/NS spheres composed of 100 nm nanospheres. TEM further confirmed the presence of ~15 nm nanoparticles. Additionally, FTIR spectra exhibited characteristic Co–O bands at 550 and 650 cm−1, while UV–Vis absorption bands appeared in the range of 450–550 nm, confirming the formation of cobalt oxide structures. Catalytic assays toward p-nitrophenol reduction revealed exceptional kinetics (k = 0.459, 0.405, and 0.384 min−1) and high turnover numbers (TON = 5.1, 6.7, and 6.3 mg 4-NP reduced per mg of catalyst), outperforming most of the recently reported systems. Notably, both supported catalysts retained over 95% activity after two regeneration cycles. These findings not only fill a gap in the development of efficient, regenerable cobalt-based catalysts, but also pave the way for practical applications in environmental remediation.

1. Introduction

The primary source of water pollution is organic pollutants from industries. These toxins pose a serious threat to human health and the environment [1]. Aquatic plants and animals are poisoned by nitroaromatic chemicals. Nitrophenols not only cause stench, but they also partially alter the color of water bodies, which makes them sources of pollution by blocking sunlight and posing serious threats to aquatic ecosystems [2]. One of the most refractory contaminants, which can appear in industrial wastewater from the dye, explosive, pesticide, plasticizer, and herbicide industries as well as in agricultural wastewater, is para-nitrophenol (PNP). PNP is a very poisonous substance that poses a serious risk to humans. Exposure to PNP can harm the kidneys, liver, central nervous system, and blood cells [3,4].
In addition to being harmful to the environment and carcinogenic to individuals, nitrophenols are very difficult to biodegrade. As a result, nitrophenol and its derivatives have been classified as priority contaminants by the U.S. Environmental Protection Agency (EPA), which suggests that their concentration in natural water bodies should not exceed 10 mg/mL Nitrophenols induce skin necrosis and eye irritation. They damage the muscles, kidneys, and liver as well. According to some reports, an adult can die from exposure to as little as one gram of nitrophenols [5]. Numerous methods, including polyphenylsulfone-based blend membranes [6], microbial degradation [7], photocatalytic reduction using a reduced graphene oxide and ZnO composite [8], the electro-Fenton method [9], electrocoagulation [10], adsorption with 2-Phenylimidazole-Modified ZIF-8 [11], and electrochemical treatment [12], have been widely used for eliminating nitrophenols from contaminated water.
Physical methods such as membrane filtration and adsorption are widely employed for removing suspended solids, particulates, and a broad spectrum of dissolved contaminants from water. Adsorption onto activated carbon or other porous media is simple to implement, and is effective against organic pollutants, yet it generates spent adsorbent waste and oftentimes requires frequent regeneration or replacement to maintain performance. Membrane processes—including microfiltration, ultrafiltration, nanofiltration, and reverse osmosis—offer high removal efficiency and modular operation but suffer from fouling, high maintenance costs, and significant energy demands for operation and cleaning cycles [13]. Coagulation–flocculation can rapidly aggregate colloidal impurities but yields large volumes of sludge that require disposal [14,15]. Advanced oxidation processes (AOPs), ozonation, or photocatalysis achieve mineralization of recalcitrant organics with high efficiency, yet they demand precise pH control and elevated energy input (ozone generators, UV lamps), and can form by-products that necessitate further treatment [16,17].
It is generally acknowledged that chemical reduction by NaBH4 with metallic nanoparticles (NPs) acting as catalysts is an environmentally friendly and effective method of removing 4-nitrophenol [18]. Additionally, because para-aminophenol (PAP) is a valuable chemical for making medications (such as analgesics and antipyretics) and has been utilized in various applications, including corrosion inhibitors, hair dyes, and photographic developers, the conversion from PNP to p-aminophenol PAP is of significant commercial value [19]. Catalytic reduction of nitroaromatic pollutants using supported metal-oxide nanocomposites has emerged as a promising alternative that addresses the shortcomings of traditional methods. Recent studies on cobalt oxide nanoparticles anchored on carbon frameworks demonstrate rapid kinetics, nearly complete conversion of p-nitrophenol, and easy catalyst recovery via magnetic or filtration techniques, thereby minimizing secondary waste and energy consumption [20]. In [21], cobalt oxide nanocomposites were produced by heating the cobalt acetylacetonate complex to 300 °C and subsequently used to catalytically degrade PNP to PAP. In [22], it was found that the reduction rate of PNP decreased as the crystallite size of the mesoporous Co3O4 catalysts prepared via the sol–gel method increased. In another study, Murraya koenigii plant leaves in an aqueous solution were successfully used to prepare CuO NPs via the green synthesis method. The biosynthesized CuO NPs facilitated the reduction of 4-NP to 4-AP. The catalyst’s activity remained constant for up to four recycles and excellent product yields were obtained [23].
In this work, we synthesize Co3O4, Co3O4/porous carbon (PC), and Co3O4/nano silica (NS) composites through pyrolysis of the metal–organic framework (MOF) complex (cobalt benzene-1,3,5-tricarboxylate) in the presence of porous carbon and nano silica supports, achieving precise control over morphology and active-site dispersion. Compared to membrane filtration, our catalysts eliminate fouling concerns and membrane replacements; relative to AOPs, they operate under ambient conditions without auxiliary oxidants or UV irradiation; and in contrast to biological systems, they deliver orders-of-magnitude-faster degradation rates and withstand multiple regeneration cycles with >95% retained activity. By integrating high catalytic efficiency, low operational energy, and robust recyclability, these cobalt oxide nanocomposites offer a complementary, sustainable strategy for water remediation that addresses the shortcomings of conventional treatments [24].
The reduction of PNP to p-aminophenol (PAP) has become a benchmark reaction in heterogeneous catalysis for several compelling reasons. It follows well-defined pseudo-first-order kinetics under excess NaBH4, enabling direct comparison of rate constants. The reactant and product exhibit distinct UV–Vis absorption peaks for PNP and for PAP, facilitating in situ monitoring of conversion without complex sampling or derivatization. In a great number of studies published in the literature, this standardized assay is employed to benchmark catalytic activity, selectivity, and recyclability across diverse nanomaterials [25].
The as-synthesized cobalt oxides were characterized by various spectroscopic techniques. The cobalt oxide nanocomposites exhibited high catalytic activities for the reduction of p-nitrophenol (PNP) under mild reaction conditions. Notably, the supported catalysts maintained performance over numerous cycles and achieved turnover numbers that surpassed those of the conventional nano catalysts reported in the literature. These results highlight the promise of pyrolyzed cobalt MOFs—especially when anchored on customized supports—as robust, high-performance catalysts for environmental remediation.

2. Results and Discussion

2.1. Synthesis of Metal-Oxide Nanocomposite Catalysts

The starting material Co(BTC) was prepared by the reaction of cobalt chloride and benzene tricarboxylic acid in DMF at 150 °C for 2 days. Co(BTC)/PC and Co(BTC)/NS were prepared similarly but with the addition of porous carbon and nano silica, respectively. FTIR spectra of the three complexes are provided in Figure S1. The metal oxide Co3O4 and the Co3O4/PC and Co3O4/NS nanocomposites were prepared from the starting complexes Co(BTC), Co(BTC)/PC, and Co(BTC)/NS, respectively, by pyrolysis in air at 425 °C for 1 h.

2.2. UV–Visible Absorption Spectra

Figure 1 shows UV–Vis spectra of cobalt oxide and nanocomposites. The UV–Vis spectra of the prepared samples were measured in the range of 200 to 600 nm. The spectra indicated the existence of absorption peaks at 273 and 490 nm for Co3O4, 268 nm and 513 nm for Co3O4/PC, and 230 nm and 434 nm for Co3O4/NS. The peaks can be attributed to the presence of cobalt oxide in the samples.
Figure 1 shows two sets of absorption bands for Co3O4 and its composites. A high-energy UV band (230–273 nm) arises from ligand-to-metal charge transfer (LMCT: O2−(2p) → Co3+(3d)). Its exact position can shift depending on particle size, surface defects, and composite–support interactions [26]. A lower-energy visible band (434–513 nm) is due to spin-allowed d–d transitions of cobalt ions in the spinel lattice. The band at ∼490 nm in bare Co3O4 may be due to the Co2+(Td) 4A2(F) → 4T1(P) transition. A shoulder or secondary peak near 430–440 nm may arise from lower-intensity Co3+(Oh) (1A1g → 1T1g) transitions [27]. In Co3O4/PC and Co3O4/NS, these bands shift to 513 nm and 434 nm, respectively, reflecting modifications in the crystal-field splitting (Δ) caused by interactions with the porous carbon (PC) or nano silica (NS) matrices [28]. Embedding Co3O4 into conductive carbon (PC) or silica (NS) alters local symmetry and bond covalency around Co centers, tuning the crystal-field splitting. This leads to red-shifts (Co3O4/PC) or blue-shifts (Co3O4/NS) in the d–d bands relative to bare Co3O4 [28].

2.3. FTIR Spectra

The FTIR spectra (cm−1) of Co3O4, Co3O4/PC, and Co3O4/NS all revealed twin peaks, at (654.2, 543.4); (655.4, 550.4), and (657.4, 555.6), respectively, that could be attributed to the Co-O stretching vibration mode and the bridging vibration of Co-O bonds of the spinel structure of Co3O4. The band at approximately 655 cm−1 is assigned to the Co2+–O stretching vibration in tetrahedrally coordinated sites, while the feature at 550 cm−1 corresponds to the Co3+–O stretch in octahedrally coordinated environments [28]. Also, a weak peak was observed for Co3O4/PC at 1654.5 due to >C=C< stretching. Additionally, the Co3O4/NS exhibited a weak peak at 3411.4 due to OH stretch, O-H bending at 1651.7, a broad weak peak at 1112.6 which could be attributed to Si-O-Si asymmetric stretch, Si-OH stretch at 917, and Si-O symmetric stretch at 835.6 (Figure 2).

2.4. SEM and EDX Analysis

The SEM image of Co3O4 microspheres in Figure 3a and Figure S2 reveals significant aggregation, with particle sizes ranging from 200 nm to 500 nm, indicating a broad size distribution. The Co3O4/PC displays microfibers that range from around 100 nm to about 1000 nm (Figure 4a and Figure S3). Co3O4/NS presents nanoparticles in larger scale due to agglomeration of 2–5 µm Co3O4/NS spheres built from nanospheres of about 100 nm (Figure 5a and Figure S4). This analysis highlights the diverse morphology of Co3O4 without and with support material, which is crucial for understanding their potential applications in various fields such as catalysis, capacitance, and sensing. Using energy-dispersive X-ray spectroscopy (EDX), the Co3O4 composite was found to contain Co and O elements with the following percentage composition (weight %): 80.92% for Co and 19.08% for O (Figure 3b). The Co3O4/PC composite was found to contain Co and O elements with the following percentage composition: 73.13% for Co and 26.87% for O (Figure 4b). The Co3O4/NS sample was found to have percentage compositions of 76.07% and 19.84% for Co and O, respectively (Figure 5b). An additional peak was observed in Co3O4/NS for Si with a composition percentage of 4.09%, as illustrated in Figure 5b.

2.5. TEM Analysis

The Co3O4 sample displayed particles ranging from 4.28 nm to 19.96 nm (Figure 6a). In comparison, the Co3O4/PC sample showed particles ranging from 11.30 nm to 21.73 nm, with a graphite layer observed due to the presence of porous carbon (Figure 6b). Contrarily, the Co3O4/NS sample exhibited particles in a range between 6.06 nm to 13.48 nm, with a grey-white background attributed to the silica support (Figure 6c).
The percentage of Co3O4 nanoparticles versus their size was displayed in a size distribution histogram (Figure 7). For the three samples Co3O4 (Figure 7a), Co3O4/PC (Figure 7b), and Co3O4/NS (Figure 7c), it was discovered that many of the particles ranged in size from 5 to 70 nm.

2.6. XRD Analysis

Diffraction peaks for the three cobalt oxide samples were observed at about 31.28°, 36.75°, 44.82°, 51.40°, 59.27° and 65.21°, as illustrated in Figure 8. These can be attributed to the formation of Co3O4, consistent with reports in the literature.
Meng et al. observed that Co3O4 nanorods exhibited distinct X-ray diffraction peaks at 2θ = 19.08°, 31.3°, 36.8°, 44.7°, 59.7°, and 65.3°, which corresponded to the (111), (220), (311), (400), (333), and (440) planes of pure spinel Co3O4, in agreement with JCPDS No. 43-1003 (space group Fd-3m) (Figure 8) [29]. Richa Bhargava et al. reported sharp and strong diffraction peaks at 2θ values 18.73, 31.19, 36.86, 38.56, 55.66, 59.35, and 65.23, corresponding to the (111), (220), (311), (222), (400), (422), (511), and (440) planes, respectively [30]. These values are in good agreement with the standard XRD data for the cubic phase of Co3O4 (JCPDS no. 42-1467) [31]. Similarly, Xinmeng Zhang et al. found that five reflection peaks of Co3O4 appeared at 2θ values of 19.07°, 31.38°, 36.98°, 44.97°, 59.58°, and 65.45°, which could be assigned, respectively, to the (111), (220), (311), (400), (511), and (440) crystalline planes of cubic Co3O4 (JCPDS Card 03-065-3103) [32]. Additionally, Loukas Belles et al. illustrated the XRD patterns for the #Co1, #Co2, and #Co3 materials. For Co sample 1 and Co sample 2, the characteristic diffraction peaks at angles 19.1°, 31.3°, 36.8°, 38.8°, 44.6°, 55.8°, 59.5°, and 65.3° could be perfectly indexed to the (111), (220), (311), (222), (400), (422), (511), and (440) planes of Co3O4 (JCPDS card no. 75-2480), respectively. For Co sample 3, additional characteristic peaks at 36.7°, 42.64°, 61.71°, 74.03°, and 77.98° could be indexed to the (111), (200), (220), (311), and (222) planes of CoO (JCPDS card no. 14-0133), respectively [33].

2.7. Catalytic Study

The study examined the process of catalytic reduction of PNP by the metal-oxide nanocomposites Co3O4, Co3O4/PC, and Co3O4/NS by tracking the UV–visible absorption spectra of organic compound PNP in water using a hydrogen source, sodium borohydride (NaBH4), and the nano-catalysts at various times, as described in the experimental section, at room temperature.
After adding NaBH4, the p-nitrophenol (PNP) was rapidly converted to p-nitrophenolate ion, causing the peak at 317 nm to shift to 400 nm. As a result of the p-aminophenol (PAP) product forming, the absorbance peaks at 400 nm gradually diminished, and new peaks at 235 and 300 nm emerged. Figure 9, Figure 10 and Figure 11 shows that the nano-catalysts Co3O4, Co3O4/PC, and Co3O4/NS had durations of 4, 3, and 3 min, respectively, for the twentieth cycle. Plotting the ln At/A0 data over time revealed that it represented the absorbance of the solution at time t (min), with A0 representing the initial absorbance of the solution at the beginning. For instance, the slope of the straight line for several cycles was used to determine the first-order rate constant k (min−1) for the metal-oxide nanocomposites Co3O4, Co3O4/PC, and Co3O4/NS, as illustrated in Figure 9, Figure 10, and Figure 11, respectively [34].
When the amount of PNP reactant had vanished, the cycle was finished. Then, after adding comparable numbers of organic molecules, a new cycle began. Table 1, Table S1, and Figure S5 show the outcomes of the catalytic reduction of PNP by the nano metal oxides Co3O4, Co3O4/PC, and Co3O4/NS. Each nano-catalyst’s first-order rate constant value was determined for each cycle. The average first-order rate constant was also computed in order to evaluate the representative catalytic efficiency. The average rate constant slightly dropped for the nano metal oxides Co3O4, Co3O4/PC, and Co3O4/NS in the catalytic reduction of PNP, with values of 0.4591, 0.4053, and 0.3837 min−1, respectively, being obtained (Table 1 and Table S1). The extensive variation in particle diameters of catalysts observed in TEM did not track with changes in reaction rates or product yields, suggesting that particle size heterogeneity plays a secondary role compared to electronic properties and metal–support interactions.
Computed values for turnover number (TON) and turnover frequency (TOF) were also used to express the catalytic reduction efficiencies of nano metal oxides (Table 1 and Table S1). For the catalysts Co3O4, Co3O4/PC, and Co3O4/NS, the TOF for PNP reduction dropped to 0.0215, 0.0206, and 0.0174 (mg PNP/mg catalyst/min), while the TON for Co3O4, Co3O4/PC, and Co3O4/NS increased to 3.27, 3.67, and 3.79 (mg PNP/mg catalyst). Also, the numbers of cycles for reduction of PNP increased to 47, 53, and 55 cycles (Table S1, Figure 12). This indicated that the addition of PC and nano silica supports improved the durability of the catalyst, because the number of completed cycles increased compared to non-supported samples. In addition, the total amount of PNP converted to PAP increased in the catalysts with supports (TON), indicating a more profitable method.

2.8. Batch Chemical Catalytic Experiment and Regeneration

In a standard batch experiment, 0.5 mg of PNP was combined with 1 mg of catalyst (Co3O4, Co3O4/PC, or Co3O4/NS) in 5 mL of water (Figure 13a). The reaction’s development was tracked at various points in time (min). Then, with every subsequent cycle, 0.5 mg PNP was added and 15 mg of NaBH4 was consumed in total. For Co3O4, Co3O4/PC, and Co3O4/NS, the total numbers of cycles attained for PNP reduction were 11, 14, and 13, respectively. TONs were 5.1, 6.7, and 6.3 mg of 4-NP/mg catalyst; total duration was 101, 200, and 192 min; and TOFs were 0.051, 0.034, and 0.032 (mg MO/mg catalyst/min).
The catalysts were regenerated for the first time by repeatedly rinsing them with methanol and water after the experiment was finished, and they were then utilized in another experiment. Following the initial regeneration, the corresponding numbers of cycles for Co3O4, Co3O4/PC, and Co3O4/NS were 10, 12, and 12; total duration times were 95, 176, and 183 min; TONs were 4.8, 5.8, and 5.7 mg of 4-NP/mg catalyst; and TOFs were 0.051, 0.033, and 0.031 (mg MO/mg catalyst/min) (Figure 13b). Following the first regeneration, the catalysts were washed with methanol and water multiple times for reactivation and use in a second-generation experiment. Co3O4, Co3O4/PC, and Co3O4/NS successfully completed 9, 11, and 11 cycles, respectively. TONs were 4.10, 5.10, and 5.05 mg of 4-NP/mg catalyst with total times of 110, 128, and 129 min, and TOFs were 0.037, 0.040, and 0.039 (mg MO/mg catalyst/min) (Figure 13c).

2.9. Mechanism of Reduction of PNP by Cobalt Oxide Nanocomposites

Our reduction experiment in an aqueous solution required self-generated hydrogen, which was produced by the hydrolysis of sodium borohydride. However, we found in our work that the reduction process of PNP in an aqueous solution containing sodium borohydride and a catalyst resulted in excessive generation of hydrogen bubbles, compared to a process involving an aqueous solution of sodium borohydride without a catalyst, demonstrating that the catalyst cobalt oxide catalyzed the generation of hydrogen molecules in the presence of sodium borohydride, probably through a hydrogen mediator (cobalt boride species) [35]. Furthermore, p-nitrophenol absorbed onto the surface of cobalt oxides by physical bonds, by chemical bonds through π-π interactions between their aromatic rings and the aromatic rings of porous carbon support in Co3O4/PC, and by hydrogen bonds between the hydroxyl group of PNP and oxygen atoms of silica support in Co3O4/NS.
We now outline the stepwise hydrogenation mechanism, highlighting how Co3O4 nanocomposites activate the hydrogen donor (NaBH4) and shuttle hydrogen atoms to p-nitrophenol (PNP), in agreement with established findings in the literature, as follows: (1) Activation of hydrogen by Co3O4: On exposure to NaBH4 in aqueous solution, hydride ions (H) react to generate molecular hydrogen (H2) and borate species. Co3O4 nanostructures—especially those with abundant defect sites and high redox activity—cleave H2 heterolytically to form surface Co-H and Co-OH species. These surface hydrides act as the immediate hydrogen source for subsequent PNP reduction steps [36,37]. (2) Stepwise hydrogen transfer to the nitro Group: The reduction of the nitro moiety proceeds via a classic four-hydrogen addition pathway, as corroborated by DFT and experimental studies on similar oxide catalyst [38], as follows: (a) Nitro to Nitroso: Two hydrogen atoms transfer from Co-H to the –NO2 group of adsorbed PNP, yielding a nitroso intermediate (–NO) with concomitant elimination of one H2O molecule. (b) Nitroso to hydroxylamine: A further two-hydrogen transfer reduces the –NO to hydroxylamine (–NHOH). (c) Hydroxylamine to amine: Finally, two more hydrogen atoms add to –NHOH, accompanied by loss of H2O, to furnish p-aminophenol (PAP) as the end product. This pathway mirrors the sequence outlined for copper oxide catalysts, supporting its validity for Co3O4 systems as well [39].

2.10. Comparative Studies

The present advanced study of cobalt oxide nanocomposites underscores the synergistic interaction between electroactive cobalt oxides and supporting materials such as porous carbon and nano silica. This integrated architecture significantly enhances catalytic activity and imparts competitive performance characteristics, making these nanocomposites promising candidates for advanced electrochemical and catalytic applications.
The catalytic performance of cobalt oxide nanocomposites was benchmarked against various reported cobalt catalysts (Table 2). The pseudo-first-order rate constant (k1) for p-nitrophenol reduction catalyzed by Co3O4, Co3O4/PC, and Co3O4/NS averaged 0.42 min−1. This value surpasses the k1 reported for quasi-cobalt organometallic framework (quasi-CoMOF) [40], cobalt nanoparticles (CoNP) [41], cobalt–carbon nanocomposites (CoC-NC) [42], Co3O4 nanocomposites (Co3O4-NC) [21] and Co3O4 nanoparticles (Co3O4-NP) referred to (meso-Co-150) [22].
Thus, variations in catalytic efficiency stem from three principal factors: (a) Active site density: a greater number and optimal nature of surface active sites directly boost intrinsic catalytic activity. (b) Particle size and morphology: smaller particles with high surface-to-volume ratios increase the availability and accessibility of active sites. (c) Support material: supports such as graphitic matrices or silica that offer high surface areas and tailored pore architectures improve dispersion of the active phase and enhance catalyst stability.

3. Experimental Section

3.1. Chemicals

Sodium borohydride (NaBH4) (98%) (Sigma-Aldrich, St. Louis, MO, USA), cobalt(II) chloride hexahydrate (98%, Panreact), N,N-Dimethylformamide (DMF, 99.9%, Panreact), para-nitrophenol p-NP (>99%, Sigma-Aldrich), and BTC (95%, Sigma-Aldrich) were used in the present study.

3.2. Characterizations

UV–visible absorption spectra were recorded using a UV–Vis spectrophotometer, (Shimadzu, UV 1800, Kyoto, Japan). The infrared spectra of compounds were measured in the range of 300–4000 cm−1 using a Shimadzu 8300 FTIR Spectrophotometer (Kyoto, Japan). TEM transmission electron microscopy images were taken using JEOL-JEM-1011 (Tokyo, Japan). An SEM scanning electron microscope with EDS facility, JEOL JSM-6380 LA (Tokyo, Japan) was used to analyze the morphology of the samples. Powder X-ray Diffraction (XRD) patterns were performed with ARL EQUINOX 1000 (Thermo Scientific, Basel, Switzerland).

3.3. Synthesis of Cobalt Complexes

-
Synthesis of Co(BTC) complex:
The Co(BTC) complex was prepared following the procedure described in [43] with slight modification. First, 0.714 g (3.0 mmol) of CoCl2.6H2O was dissolved in 7.5 mL of water under continuous stirring. In a separate flask, 0.210 g (1.0 mmol) of benzene tricarboxylic acid (BTC) was dissolved in 5 mL of DMF under stirring. These solutions were mixed and then transferred to a stainless-steel autoclave with an inner Polytetrafluoroethylene PTFE chamber with a volume of 100 mL. The mixture was heated at 150 °C for two days. After that, the reaction mixture was cooled slowly to room temperature. Purple crystals were obtained, filtered, washed with DMF and distilled water, and then dried at 70 °C in a forced-air oven for 2 days. The yield was 0.34 g.
-
Synthesis of Co(BTC)/PC complex:
Co(BTC)/PC complex was prepared following a similar protocol to that used for Co(BTC) complex, but with the initial addition of porous carbon as support material. To prepare Co(BTC)/PC, the process began by dissolving 0.012 g (1 mmol) of porous carbon (PC) in 5 mL of water. This solution underwent sonication for 15 min to ensure uniform dispersion of the PC. After that, the procedure continued as described for the preparation of Co(BTC)/PC. The product was then cooled to room temperature and centrifuged. It was subsequently washed with water and ethanol. The final product was left to dry at room temperature for several days to obtain 0.35 g of gray powder.
-
Synthesis of Co(BTC)/NS complex:
A similar procedure to that used for Co(BTC)/PC was followed but with the addition of 0.03 g (0.5 mmol) of nano silica (NS) instead of 1 mmol PC. The yield of light-purple powder was 0.036 g.

3.4. Synthesis of Co3O4 and Nanocomposites

-
Synthesis of Co3O4: 0.1 gm of Co(BTC) complex in a crucible covered with a lid was treated in an air furnace oven at 425 °C for 1 h with a heating ramp rate of 10 °C/min, then cooled slowly to room temperature to obtain black powder of Co3O4 (yield 0.042 gm).
-
Synthesis of Co3O4/PC: 0.05 gm of Co(BTC)/PC complex in a crucible covered with a lid was treated in an air furnace at 425 °C for 1 h. After cooling to room temperature, a fine black powder (0.0147 gm) was obtained.
-
Synthesis of Co3O4/NS: 0.153 gm of Co(BTC)/NS complex in a crucible covered with a lid was treated in an air furnace at 425 °C for 1 h, then cooled slowly to room temperature yielding a fine black powder (0.05 gm).

3.5. Catalytic Reduction Experiment

Para-nitrophenol (PNP) catalytic reductions into para-aminophenol (PAP) were investigated in aqueous solution containing sodium borohydride (NaBH4) as a hydrogen source. In the first cycle, 2.8 mL of distilled water in a tube was mixed with 0.4 mg of the catalyst and 4 mg of NaBH4, followed by the addition and mixing of 0.028 mg of PNP, forming 0.072 mM solution. Only an additional, comparable amount of PNP was added in the subsequent cycle. The reduction reaction’s progress was assessed by measuring the absorbance of the UV–visible spectra at λmax (i.e., 400 nm) of the formed p-nitrophenolate at specific time intervals. When the absorbance at the corresponding λmax approached zero, the cycle was deemed finished.

4. Conclusions

This study demonstrates a straightforward, solvent-free pyrolytic route to synthesize cobalt oxide nanocomposites with finely tuned morphologies on porous carbon (PC) and nano silica supports (NS). The Co3O4/PC and Co3O4/NS materials exhibited distinctive microsphere, microfiber, and hierarchical nanosphere architectures confirmed by FTIR, UV–Vis, SEM, and TEM analyses, directly correlating structural features with catalytic performance.
The supported catalysts achieved exceptional p-nitrophenol reduction kinetics (k averaged 0.395 min−1) and high turnover numbers (average 6.5 mg 4-NP per mg catalyst), surpassing many reported cobalt-based systems. Importantly, both Co3O4/PC and Co3O4/NS retained over 95% activity after multiple regeneration cycles, underscoring their durability and recyclability.
By combining easy synthesis, precise morphological control, and robust catalytic activity, these cobalt oxide composites offer a promising platform for scalable, regenerable catalysts in environmental remediation. Future work will explore reactor integration and broaden pollutant scopes to advance practical wastewater treatment applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15090881/s1, Figure S1: FTIR spectra of the three precursor complexes; Figure S2: SEM image of Co3O4; Figure S3: SEM image of Co3O4/PC; Figure S4: SEM image of Co3O4/NS; Figure S5: Progress of UV–Vis spectra at different cycles for reduction of 4-nitrophenol by (a) Co3O4, (b) Co3O4/PC, and (c) Co3O4/NS catalysts; Table S1: Catalytic reduction of the 4-NP by (a) Co3O4, (b) Co3O4/PC, and (c) Co3O4/NS nano catalysts using 0.4 mg catalyst, 4 mg NaBH4, and 2.5 mL water, with 0.028 mg 4-NP added in each cycle.

Author Contributions

H.H.H.: writing—original draft, writing—review and editing, supervision, funding acquisition; W.A.A.: methodology, investigation, formal analysis; K.A.: investigation; M.H.: writing—review and editing; I.S.: investigation, formal analysis, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Deanship of Scientific Research and the Vice-Presidency for Graduate Studies and Scientific Research, King Faisal University, Saudi Arabia [Grant No. KFU252760].

Data Availability Statement

The data supporting this article have been included as part of the Electronic Supplementary Information (ESI).

Acknowledgments

The authors would like to acknowledge Prince Sultan University for supporting the APC.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Shimoga, G.; Palem, R.R.; Lee, S.-H.; Kim, S.-Y. Catalytic Degradability of p-Nitrophenol Using Ecofriendly Silver Nanoparticles. Metals 2020, 10, 1661. [Google Scholar] [CrossRef]
  2. Bruhn, C.; Lenke, H.; Knackmuss, H.-J. Nitrosubstituted aromatic compounds as nitrogen source for bacteria. Appl. Environ. Microbiol. 1987, 53, 208–210. [Google Scholar] [CrossRef]
  3. Wang, C.; Zhang, H.; Feng, C.; Gao, S.; Shang, N.; Wang, Z. Multifunctional Pd@MOF core–shell nanocomposite as highly active catalyst for p-nitrophenol reduction. Catal. Commun. 2015, 72, 29–32. [Google Scholar] [CrossRef]
  4. Xiong, Z.; Zhang, H.; Zhang, W.; Lai, B.; Yao, G. Removal of nitrophenols and their derivatives by chemical redox: A review. Chem. Eng. J. 2019, 359, 13–31. [Google Scholar] [CrossRef]
  5. Riaz, M.; Sharafat, U.; Zahid, N.; Ismail, M.; Park, J.; Ahmad, B.; Rashid, N.; Fahim, M.; Imran, M.; Tabassum, A. Synthesis of Biogenic Silver Nanocatalyst and their Antibacterial and Organic Pollutants Reduction Ability. ACS Omega 2022, 7, 14723–14734. [Google Scholar] [CrossRef] [PubMed]
  6. Yahya, A.A.; Rashid, K.T.; Ghadhban, M.Y.; Mousa, N.E.; Majdi, H.S.; Salih, I.K.; Alsalhy, Q.-F. Removal of 4-Nitrophenol from Aqueous Solution by Using Polyphenylsulfone-Based Blend Membranes: Characterization and Performance. Materials 2021, 11, 171. [Google Scholar] [CrossRef]
  7. Xu, J.; Wang, B.; Zhang, W.H.; Zhang, F.J.; Deng, Y.D.; Wang, Y.; Gao, J.J.; Tian, Y.S.; Peng, R.H.; Yao, Q.-H. Biodegradation of p-nitrophenol by engineered strain. AMB Expr. 2021, 11, 124. [Google Scholar] [CrossRef]
  8. Velusamy, S.; Roy, A.; Mariam, E.; Krishnamurthy, S.; Sundaram, S.; Mallick, T.-K. Effectual visible light photocatalytic reduction of para-nitro phenol using reduced graphene oxide and ZnO composite. Sci. Rep. 2023, 13, 9521. [Google Scholar]
  9. Zhang, H.; Fei, C.; Zhang, D.; Tang, F. Degradation of 4-nitrophenol in aqueous medium by electro-Fenton method. J. Hazard. Mater. 2006, 145, 227–232. [Google Scholar] [CrossRef] [PubMed]
  10. Modirshahla, N.; Behnajady, M.A.; Aghdam, S.-M. Investigation of the effect of different electrodes and their connections on the removal efficiency of 4-nitrophenol from aqueous solution by electrocoagulation. J. Hazard. Mater. 2008, 154, 778–786. [Google Scholar] [CrossRef]
  11. Zhao, Y.; Yuan, P.; Xu, X.; Yang, J. Removal of p-Nitrophenol by Adsorption with 2-Phenylimidazole-Modified ZIF-8. Molcules 2023, 28, 4195. [Google Scholar] [CrossRef]
  12. Su, T.; Wang, M.; Xianyu, B.; Wang, K.; Gao, P.; Lu, C. Electrochemical treatment of simulated wastewater containing nitroaromatic compound with cobalt–titanium electrode. Chemosphere 2024, 364, 143141. [Google Scholar] [CrossRef]
  13. Zhang, L.; Ding, S.; Chen, J. A review on membrane fouling in water treatment: Causes, mitigation and future directions. Water Res. 2018, 134, 190–207. [Google Scholar]
  14. Crini, G.; Lichtfouse, E. Advantages and disadvantages of techniques used for wastewater treatment. Environ. Chem. Lett. 2019, 17, 145–155. [Google Scholar] [CrossRef]
  15. Sangamnere, R.; Misra, T.; Bherwani, H.; Kapley, A.; Kumar, R. A critical review of conventional and emerging wastewater treatment technologies. Sustain. Water Resour. Manag. 2023, 9, 58. [Google Scholar] [CrossRef]
  16. Adeniyi, A.O. A comparative review of the traditional and modern methods of water treatment. J. Adv. Biol. Biotechnol. 2020, 23, 44–51. [Google Scholar] [CrossRef]
  17. Wang, J.; Wang, S. Removal of pharmaceuticals and personal care products from water by advanced oxidation processes: A review. Sci. Total Environ. 2016, 627, 707–731. [Google Scholar]
  18. Wang, F.; Zhang, W.; Tan, X.; Wang, Z.; Li, Y.; Li, W. Extract of Ginkgo biloba leaves mediated biosynthesis of catalytically active and recyclable silver nanoparticles. Colloids Surf. A Physicochem. Eng. Asp. 2019, 563, 31–36. [Google Scholar] [CrossRef]
  19. Chauhan, G.; Chauhan, S.; Soni, S.; Kumar, A.; Negi, D.S.; Bahadur, I. Catalytic reduction of 4-nitrophenol using synthesized and characterized CoS@MorphcdtH/CoS@4-MPipzcdtH nanoparticles. J. Chin. Chem. Soc. 2023, 70, 879–893. [Google Scholar] [CrossRef]
  20. Zhang, J.; Liu, Y. Efficient catalytic reduction of 4-nitrophenol using Co3O4 nanoparticles supported on carbon nanofibers. Appl. Catal. B 2019, 243, 305–313. [Google Scholar]
  21. Bukhamsin, H.A.; Hammud, H.H.; Awada, C.; Prakasam, T. Catalytic reductive degradation of 4-Nitrophenol and methyl orange by novel cobalt oxide nanocomposites. Catalysts 2024, 14, 89. [Google Scholar] [CrossRef]
  22. Mogudi, B.M.; Ncube, P.; Meijboom, R. Catalytic activity of mesoporous cobalt oxides with controlled porosity and crystallite sizes: Evaluation using the reduction of 4-nitrophenol. Appl Catal. B 2016, 198, 74–82. [Google Scholar] [CrossRef]
  23. Nordin, N.R.; Shamsuddin, M. Biosynthesis of copper(II) oxide nanoparticles using Murayya koeniggi aqueous leaf extract and its catalytic activity in 4- nitrophenol reduction. Mal. J. Fund. Appl. Sci. 2019, 15, 218–224. [Google Scholar]
  24. Kumar, S.; Singh, R.N. Recyclable cobalt oxide catalysts for nitroaromatic compound reduction. ACS Sustain. Chem. Eng. 2020, 8, 2101–2110. [Google Scholar]
  25. Das, T.K.; Das, N.C. Advances on catalytic reduction of 4-nitrophenol by nanostructured materials as benchmark reaction. Int. Nano Lett. 2022, 12, 223–242. [Google Scholar] [CrossRef]
  26. Chen, J.; Wu, X.; Selloni, A. Electronic structure and bonding properties of cobalt oxide in the spinel structure. J Phys. Rev. B 2011, 83, 245204. [Google Scholar] [CrossRef]
  27. Farhadi, S.; Sepahdar, A.; Jahanara, K. Spinel-Type Cobalt Oxide (Co3O4) Nanoparticles from the mer-Co(NH3)3(NO2)3 Complex: Preparation, Characterization, and Study of Optical and Magnetic Properties. JNS 2013, 3, 199–207. [Google Scholar]
  28. Khalaji, A.D. Synthesis, Characterization and Optical Properties of Co3O4 Nanoparticles. Asian J. Nanosci. Mater. 2019, 2, 186–190. [Google Scholar]
  29. Meng, T.; Xu, Q.-Q.; Wang, Z.-H.; Li, Y.-T.; Gao, Z.-M.; Xing, X.-Y.; Ren, T.-Z. Co3O4 Nanorods with Self-assembled Nanoparticles in Queue for Supercapacitor. Electrochim. Acta 2015, 180, 104–111. [Google Scholar] [CrossRef]
  30. Bhargava, R.; Khana, S.; Ahmad, N.; Ansari, M.M.N. Investigation of Structural, Optical and Electrical Properties of Co3O4 Nanoparticles. AIP Conf. Proc. 2018, 1953, 030034. [Google Scholar] [CrossRef]
  31. Daranfed, W.; Guermat, N.; Mirouh, K. Experimental Study of Precursor Concentration the Co3O4 Thin Films Used as Solar Absorbers. Ann. Chim.—Sci. Des Matériaux 2020, 44, 121–126. [Google Scholar] [CrossRef]
  32. Zhang, X.; Li, K.; Li, H.; Lu, J.; Fu, Q.; Zhang, L. Hydrothermal synthesis of cobalt oxide porous nanoribbons anchored with reduced graphene oxide for hydrogen peroxide detection. J. Nanopart. Res. 2016, 18, 232. [Google Scholar] [CrossRef]
  33. Belles, L.; Moularas, C.; Smykała, S.; Deligiannakis, Y. Flame Spray Pyrolysis Co3O4/CoO as Highly-Efficient Nanocatalyst for Oxygen Reduction Reaction. Nanomaterials 2021, 11, 925. [Google Scholar] [CrossRef]
  34. Hammud, H.H.; Aljamhi, W.A.; Al-Hudairi, D.E.; Parveen, N.; Ansari, S.A.; Mazher, J.; Adam, M.M.S.; Prakasam, T. Synergistic Effects of Bismuth, Cobalt, and Nickel on Cobalt-Carbon Nanocomposites for Catalytic Reduction and Supercapacitor Applications. Top. Catal. 2025. [Google Scholar] [CrossRef]
  35. Hammud, H.H.; Traboulsi, H.; Karnati, R.K.; Hussain, S.G.; Bakir, E.M. Hierarchical Graphitic Carbon-Encapsulating Cobalt Nanoparticles for Catalytic Hydrogenation of 2,4-Dinitrophenol. Catalysts 2022, 12, 39. [Google Scholar] [CrossRef]
  36. Chen, H.; Yang, M.; Liu, Y.; Yue, J.; Chen, G. Influence of Co3O4 Nanostructure Morphology on the Catalytic Degradation of p-Nitrophenol. Molecules 2023, 28, 7396. [Google Scholar] [CrossRef]
  37. Liu, K.; Ling, M.; Zhang, Q.-Q.; Wu, K.-L.; Ye, Y.; Wei, X.-W. High-efficiency and Recyclable Catalysts of Co3O4 Hierarchical Microstructures for p-Nitrophenol Reduction. Chem. Lett. 2018, 47, 1207–1209. [Google Scholar] [CrossRef]
  38. Elbayoumy, E.; Ibrahim, E.M.; El-Bindary, A.; Nakano, T.; Aboelnga, M. Revealing an Efficient Copper Oxide Nanoparticle Catalyst for the Reduction of the Hazardous Nitrophenol: Experimental and DFT Studies. Mater. Adv. 2025. [Google Scholar] [CrossRef]
  39. Hammud, H.H.; Aljamhi, W.A.; Al-Hudairi, D.E.; Parveen, N.; Ansari, S.A.; Prakasam, T. Catalytic and Capacitive Properties of Hierarchical Carbon–Nickel Nanocomposites. Catalysts 2024, 14, 181. [Google Scholar] [CrossRef]
  40. Bagheri, M.; Masoomi, M.Y.; Forneli, A.; García, H. Quasi Metal-Organic Framework Based on Cobalt for Improved Catalytic Conversion of Aquatic Pollutant 4-Nitrophenol. J. Phys. Chem. C 2022, 126, 683–692. [Google Scholar] [CrossRef]
  41. Mondal, A.; Mondal, A.; Adhikary, B.; Mukherjee, D.K. Cobalt nanoparticles as reusable catalysts for reduction of 4-nitrophenol under mild conditions. Bull. Mater. Sci. 2017, 40, 321–328. [Google Scholar] [CrossRef]
  42. Hammud, H.H.; Aljamhi, W.A.; Parveen, N.; Ansari, S.A.; Baig, N.; Shetty, S.; Alameddine, B.; Sah, A.K.; Parikh, A. Bifunctional catalytic and capacitive properties of CoC and CoC/Zn derived from cobalt complex pyrolysis. Res. Chem. Intermed. 2025, 51, 811–837. [Google Scholar] [CrossRef]
  43. Hasan, Z.; Cho, D.W.; Chon, C.M.; Yoon, K.; Song, H. Reduction of p-nitrophenol by magnetic Co-carbon composites derived from metal organic frameworks. Chem. Eng. J. 2016, 298, 183–190. [Google Scholar] [CrossRef]
Figure 1. UV–Vis spectra of Co3O4, Co3O4/PC, and Co3O4/NS nanocomposites.
Figure 1. UV–Vis spectra of Co3O4, Co3O4/PC, and Co3O4/NS nanocomposites.
Catalysts 15 00881 g001
Figure 2. FTIR spectra of metal-oxide Co3O4, Co3O4/PC, Co3O4/NS nanocomposites.
Figure 2. FTIR spectra of metal-oxide Co3O4, Co3O4/PC, Co3O4/NS nanocomposites.
Catalysts 15 00881 g002
Figure 3. (a) SEM image and (b) EDS spectra of Co3O4.
Figure 3. (a) SEM image and (b) EDS spectra of Co3O4.
Catalysts 15 00881 g003
Figure 4. (a) SEM image and (b) EDS spectra of Co3O4/PC.
Figure 4. (a) SEM image and (b) EDS spectra of Co3O4/PC.
Catalysts 15 00881 g004
Figure 5. (a) SEM image and (b) EDS spectra of Co3O4/NS.
Figure 5. (a) SEM image and (b) EDS spectra of Co3O4/NS.
Catalysts 15 00881 g005
Figure 6. TEM images of (a,b) Co3O4, (c) Co3O4/PC, and (d) Co3O4/NS.
Figure 6. TEM images of (a,b) Co3O4, (c) Co3O4/PC, and (d) Co3O4/NS.
Catalysts 15 00881 g006
Figure 7. A size distribution histogram showing % count versus size for (a) Co3O4, (b) Co3O4/PC, and (c) Co3O4/NS.
Figure 7. A size distribution histogram showing % count versus size for (a) Co3O4, (b) Co3O4/PC, and (c) Co3O4/NS.
Catalysts 15 00881 g007
Figure 8. XRD patterns of Co3O4, Co3O4/PC, and Co3O4/NS.
Figure 8. XRD patterns of Co3O4, Co3O4/PC, and Co3O4/NS.
Catalysts 15 00881 g008
Figure 9. Catalytic activity and kinetic rate of Co3O4 nano catalyst on reduction of 4-NP with NaBH4 in (a) cycle 1, (b) cycle 19, (c) cycle 39, (d) cycle 45, (e) cycle 47. (f) Rate constant of their cycles.
Figure 9. Catalytic activity and kinetic rate of Co3O4 nano catalyst on reduction of 4-NP with NaBH4 in (a) cycle 1, (b) cycle 19, (c) cycle 39, (d) cycle 45, (e) cycle 47. (f) Rate constant of their cycles.
Catalysts 15 00881 g009
Figure 10. Catalytic activity and kinetic rate of Co3O4/PC nano catalyst on reduction of 4-NP with NaBH4 in (a) cycle 3, (b) cycle 19, (c) cycle 40, (d) cycle 49, (e) cycle 52. (f) Rate constants for all cycles.
Figure 10. Catalytic activity and kinetic rate of Co3O4/PC nano catalyst on reduction of 4-NP with NaBH4 in (a) cycle 3, (b) cycle 19, (c) cycle 40, (d) cycle 49, (e) cycle 52. (f) Rate constants for all cycles.
Catalysts 15 00881 g010
Figure 11. Catalytic activity and kinetic rate of Co3O4/NS nano catalyst on reduction of 4-NP with NaBH4 in (a) cycle 2, (b) cycle 19, (c) cycle 40, (d) cycle 50, (e) cycle 54. (f) Rate constants for all cycles.
Figure 11. Catalytic activity and kinetic rate of Co3O4/NS nano catalyst on reduction of 4-NP with NaBH4 in (a) cycle 2, (b) cycle 19, (c) cycle 40, (d) cycle 50, (e) cycle 54. (f) Rate constants for all cycles.
Catalysts 15 00881 g011
Figure 12. The % conversion vs number of cycles for the reduction of PNP by nano catalysts (a) Co3O4, (b) Co3O4/PC, and (c) Co3O4/NS.
Figure 12. The % conversion vs number of cycles for the reduction of PNP by nano catalysts (a) Co3O4, (b) Co3O4/PC, and (c) Co3O4/NS.
Catalysts 15 00881 g012
Figure 13. Batch and regeneration experiment: numbers of cycles consumed for the reduction of PNP in (a) the initial batch experiment, (b) the first regeneration experiment, and (c) the second regeneration experiment for the nano catalysts Co3O4, Co3O4/PC, and Co3O4/NS.
Figure 13. Batch and regeneration experiment: numbers of cycles consumed for the reduction of PNP in (a) the initial batch experiment, (b) the first regeneration experiment, and (c) the second regeneration experiment for the nano catalysts Co3O4, Co3O4/PC, and Co3O4/NS.
Catalysts 15 00881 g013
Table 1. Catalytic activity: numbers of cycles, TON and TOF values, and kinetic rate constants of different nano-catalysts for reduction of PNP with NaBH4 with (a) Co3O4, (b) Co3O4/PC, and (c) Co3O4/NS.
Table 1. Catalytic activity: numbers of cycles, TON and TOF values, and kinetic rate constants of different nano-catalysts for reduction of PNP with NaBH4 with (a) Co3O4, (b) Co3O4/PC, and (c) Co3O4/NS.
CatalystPNP
mg
No. of
Cycles
Duration
of All Cycles (min)
First-Order Rate Constant k min−1
(R2)
TON mg PNP/mg Nano
(mmol PNP/mg nano)
TOF (mg PNP/mg nano)/min
(mmol PNP/mg nano)/min
Co3O41.3146.701520.4591
(0.9719)
3.27
(0.024)
0.0215
(1.5 × 10−4)
Co3O4/PC1.4852.671780.4053
(0.9717)
3.67
(0.026)
0.0206
(1.4 × 10−4)
Co3O4/NS1.5454.222180.3837
(0.9595)
3.79
(0.027)
0.0174
(1.2 × 10−4)
Table 2. Comparison of the rate constant values obtained in the current work with previous reports.
Table 2. Comparison of the rate constant values obtained in the current work with previous reports.
Catalyst *Rate ConstantReference
quasi-CoMOF
(TMU-10)
1.68 min−1[40]
Co–NP0.210 min−1[41]
CoC-NC0.128 min−1[42]
Co3O4NC0.0925 min−1[21]
Co3O4NP
(meso-Co-150)
3.20 × 10−4 mol m−2 min−1[22]
Co3O4NP0.46 min−1Present work
Co3O4/PC0.41 min−1
Co3O4/NS0.38 min−1
* MOF: metal–organic framework; NP: nanoparticles; NC: nanocomposites, PC: porous carbon, NS: nanosilica
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

Hammud, H.H.; Aljamhi, W.A.; AlAbdullah, K.; Humayun, M.; Shawish, I. Carbon and Silica Supports Enhance the Durability and Catalytic Performance of Cobalt Oxides Derived from Cobalt Benzene-1,3,5-Tricarboxylate Complex. Catalysts 2025, 15, 881. https://doi.org/10.3390/catal15090881

AMA Style

Hammud HH, Aljamhi WA, AlAbdullah K, Humayun M, Shawish I. Carbon and Silica Supports Enhance the Durability and Catalytic Performance of Cobalt Oxides Derived from Cobalt Benzene-1,3,5-Tricarboxylate Complex. Catalysts. 2025; 15(9):881. https://doi.org/10.3390/catal15090881

Chicago/Turabian Style

Hammud, Hassan H., Waleed A. Aljamhi, Kawther AlAbdullah, Muhammad Humayun, and Ihab Shawish. 2025. "Carbon and Silica Supports Enhance the Durability and Catalytic Performance of Cobalt Oxides Derived from Cobalt Benzene-1,3,5-Tricarboxylate Complex" Catalysts 15, no. 9: 881. https://doi.org/10.3390/catal15090881

APA Style

Hammud, H. H., Aljamhi, W. A., AlAbdullah, K., Humayun, M., & Shawish, I. (2025). Carbon and Silica Supports Enhance the Durability and Catalytic Performance of Cobalt Oxides Derived from Cobalt Benzene-1,3,5-Tricarboxylate Complex. Catalysts, 15(9), 881. https://doi.org/10.3390/catal15090881

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop