Next Article in Journal
Coordinated Frequency Control of an Energy Storage System with a Generator for Frequency Regulation in a Power Plant
Previous Article in Journal
Assessment of the Chemical Reactivity of Brazilian Stone Cutting Plant Waste into Cementitious Matrices
Previous Article in Special Issue
Photoelectrochemical Water Splitting and H2 Generation Enhancement Using an Effective Surface Modification of W-Doped TiO2 Nanotubes (WT) with Co-Deposition of Transition Metal Ions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mn2+ Doped Cobalt Oxide and Its Composite with Carbon Nanotubes for Adsorption-Assisted Photocatalytic Applications

1
Department of Chemistry, COMSATS University Islamabad (CUI), Abbottabad Campus, Khyber Pakhtunkhwa 22010, Pakistan
2
Department of Chemistry, School of Natural Sciences (SNS), National University of Science and Technology (NUST), H-12, Islamabad 46000, Pakistan
3
Department of Chemistry, Abbottabad University of Science and Technology, Abbottabad 22010, Pakistan
4
Department of Chemistry, Quaid-i-Azam University, Islamabad 15320, Pakistan
5
Department of Chemistry, College of Science, Princess Nourah Bint Abdulrahman University, P.O. Box 84428, Riyadh 11671, Saudi Arabia
6
Department of Pharmaceutical Sciences, College of Pharmacy, AlMaarefa University, Riyadh 13713, Saudi Arabia
7
Department of Chemistry, Faculty of Applied Science, Umm Al-Qura University, Makkah 24230, Saudi Arabia
8
Department of Chemistry, College of Science, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
9
Department of Biotechnology, College of Science, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Sustainability 2022, 14(24), 16932; https://doi.org/10.3390/su142416932
Submission received: 3 October 2022 / Revised: 7 December 2022 / Accepted: 14 December 2022 / Published: 16 December 2022
(This article belongs to the Special Issue Sustainable Photocatalytic Water Treatment and Energy Production)

Abstract

:
In this study, cobalt oxide (Co3O4), Mn-doped Co3O4 (MDCO), and Mn-doped Co3O4-functionalized carbon nanotube (MDCO-CNTs) were synthesized via the co-precipitation method using cobalt nitrate and manganese nitrate as a cobalt and manganese precursor, respectively. Synthesized materials were assessed using different characterization techniques like scanning electron microscopy, X-ray diffraction, and UV-visible spectroscopy. Congo red in an aqueous solution was adopted as the model dye to estimate the adsorption-assisted photocatalytic efficiency of the synthesized materials. The samples studied for adsorpsstion-assisted photocatalysis were found to be highly effective and among all the samples, the best removal performance (80%) was obtained by treating the MDCO-CNTs composite for 50 min at 50 °C. Mathematical modeling shows that all of the samples followed a pseudo-second-order kinetic model and data best fitted to a Langmuir isotherm, implying that the process involved in the removal of Congo red dye is chemisorption.

1. Introduction

Water pollution, a global concern threatening the survival of humans, has been remarkably accelerated with the boom of industrialization and the resultant effluents [1,2,3] In response, a variety of strategies including sedimentation, adsorption, and biodegradation have been employed for wastewater treatment [4]. However, addressing this problem is complex due to the diverse compositions and physicochemical behavior of the contaminants which calls for efficient, low-energy consumption and eco-friendly remediation for industrial effluents. Photocatalysis has been a hot topic across the globe for a variety of applications, particularly environmental remediation. In a typical heterogeneous photocatalytic system, semiconductors act upon light irradiation, absorbing photons to stimulate the electrons to jump to the conduction band from the valence band, leaving holes behind. These excited electrons form superoxide (O2) radical ions by reacting with oxygen; hydroxyl radicals (HO) are also produced in the presence of H2O and OH-(electron donor species). Both radical ions, superoxide (O2) and hydroxyl (HO), are strong oxidizing agents and have the potential to degrade water pollutants [5,6,7].
The most common heterogeneous photocatalytic systems are typically based on semiconductors, such as TiO2, a key contributor in the field of photocatalysis with over 2,240,000 photocatalytic reports [8,9,10,11]. Although efforts have been dedicated to discovering low-cost potential materials to operate as sole photocatalysts, currently there is also an option to develop strategies for fabricating composite catalysts by incorporating a semiconductor with different functional components and/or metals [1]. To design efficient catalysts, the exploitation of metal oxide-carbon nanotubes (CNTs) composites is commonly observed [12]. CNTs, owing to the tunable structural, physical, chemical, and electrical properties are inspiring enough to be used in this regard [13], e.g., TiO2-CNTs [14,15,16,17,18,19,20,21,22], ZnO-CNTs [23,24], Cu2O-CNTs [25], SnO2-CNTs [26], Fe2O3-CNTs [27], CoS2-CNTs [28], NiFe2O4-CNTs [29], CdS-CNTs [30], and in many other binary and ternary compounds [31,32,33,34,35,36], etc.
Cobalt oxide (Co3O4), is widely applied and the most versatile transition-metal oxide, having stable spinel geometry with Co2+ and Co3+ ions at tetrahedral and octahedral coordination, respectively [37,38]. Furthermore, different morphologies of Co3O4 nanoparticles are reported to degrade water pollutants very effectively [39]. Although considerable development in photocatalyst designing has been achieved, there are still a few aspects, that need to be considered, mainly: (i) relatively complex synthesis methods, (ii) high synthesis cost, and (iii) their harmful impact on the environment. Herein, we aimed to develop the new composite photocatalysts i.e., manganese (Mn) doped Co3O4 and Mn-doped Co3O4-functionalized CNTs using a simple and cost-effective co-precipitation method for wastewater treatment by opting for Congo red as a model dye.

2. Materials and Methods

For the synthesis of Co3O4, the co-precipitation method was adopted. First, a 0.1 molar cobalt nitrate (Co(NO3)2.6H2O) solution was prepared in double distilled water. Then, 0.2 M solutions of sodium hydroxide (NaOH) were added dropwise into the reaction medium to attain pH 10 and stirred for 2 h continuously at 60 °C. After that, precipitates of cobalt hydroxide (Co(OH)2) were filtered and washed well with double distilled water to eliminate nitrates and other contaminants. These precipitates were dehydrated in the oven at 105 °C for 5 h, and after crushing, annealing was performed at 400 °C for 2 h.
For the synthesis of Mn-doped Co3O4, equimolar, equi-volume solutions of manganese nitrate (Mn(NO3)2.6H2O) and Co(NO3)2.6H2O were mixed in a beaker and heated continuously at 60 °C with constant stirring. Then, a 0.2 M solution of NaOH was added dropwise into the reaction medium to attain pH 10. With the addition of NaOH, precipitates of metal hydroxide appeared in the reaction mixture. These precipitates were washed with double distilled water after filtration. They were dried at 60 °C for 5 h and crushed to make fine powders. The muffle furnace was used to anneal the solution at 400 °C for two hours.
For the synthesis of the composite material, CNTs were used as matrix and undoped, as well as doped metal oxides, were applied as reinforcements. CNTs were functionalized before the formation of the composite. For the functionalization, CNTs were treated with a mixture of concentrated H2SO4 and HNO3 in a ratio of 3 to 1. Then, 10 mg of unfunctionalized CNTs were mixed with an 8 ml mixture of acids. The mixture was refluxed at 70 °C for 4 h. After refluxing, the functionalized CNTs were filtered using vacuum filtration with 0.2 µm pore size filter paper. The CNTs were dried in the oven at 60 °C for 24 h.
For the synthesis of the MDCO-CNT composite, the mixture of ethanol and distilled water was used as a solvent in the ratio of 3:1, respectively. Next, 100 mL of MDCO (0.2 mg/mL) solution was mixed with 100 mL of CNT (0.4 mg/mL). This mixture was continuously sonicated for 3 h. After sonication, the mixture was filtered and cleaned with distilled water. The resultant composite was then dried in the oven at 60 °C for 5 h. Synthesized materials were subjected to different characterization techniques. The materials’ crystallization and phase studies were assessed using the Rigaku (MiniFlex 600) X-ray powder diffraction system. To ascertain the topographical and morphological structure of the materials, scanning electron microscopy (SEM) pictures were gathered using a VEGA3 (TESCAN) scanning electron microscope. Degradation efficiency for Congo red dye was studied by UV-vis. spectrophotometer BMS (UV-1602) (Shimadzu, Tokyo, Japan).

2.1. Adsorption Study

The adsorption study was carried out by using a 10 mg catalyst in a 25 mL solution of each concentration of 5 ppm to 50 ppm. Then, 25 mL of each solution with 10 mg of catalyst was continuously stirred in the dark for 60 min to achieve an adsorption-desorption equilibrium. Every 20 min, an aliquot was taken, filtered, and studied under a UV-visible spectrophotometer.

2.2. Photocatalytic Study

For the band gap analysis of the synthesized materials, a graph was plotted between absorbance and energy (eV) and the band gap value was assumed at the cutoff point of the graph.
Photocatalytic studies were carried out under a UV lamp and a 100-watt bulb in a specified photoreactor. For this study, dye solutions were irradiated under photo reactors. In the UV reactor, an Hg lamp was used and for visible light, a tungsten lamp (100 watts) was used at a distance of 15 cm. For this study, 25 mL of each concentration (5 ppm to 35 ppm) with a 10 mg catalyst was first kept in dark for adsorption and then under a tungsten lamp for 30 min. After irradiating, a solution was filtered and used for studies under a UV-visible spectrophotometer. The same process is used under UV radiation.

2.3. Effect of Different Parameters on Adsorption

By adjusting several factors, including the dye concentration, adsorbent dose, duration, temperature, and solution pH, the variation in dye adsorption on the surface of manufactured adsorbent was examined.

2.4. Effect of Dye Concentration

Solutions of dye at various concentrations, from 5 ppm to 19 ppm, were employed to study the impact of dye concentration on adsorption. Then, 25 mL of each concentration was stirred in the dark for 30 min with 10 mg adsorbent, filtered, and analyzed by UV-visible spectrophotometer.

2.5. Effect of Catalyst Dosage

To study the effect of adsorbent dosage, 25 mL of 15 ppm solution was put into a small beaker and various amounts of adsorbent were added: 10 mg, 20 mg, 30 mg, 40 mg, 50 mg, and 60 mg. Each solution was continuously stirred in the dark for 30 min, filtered, and analyzed by a UV-visible spectrophotometer.

2.6. Effect of Time

The effect of time studied was by varying the contact time for adsorption. In total, 50 mL of dye solution was used in this work, and it was constantly agitated in the presence of 20 mg of adsorbent. An aliquot was obtained every 5 min, and it was then examined using a UV-visible spectrophotometer. The duration ranged from 5 to 60 min.

2.7. Point of Zero Charges (PZC) of Photocatalyst

The pH value at which a material’s surface becomes neutral is known as the point of zero charges (PZC). Material surfaces change from being positively charged above the PZC value to being negatively charged below.
To find the charge on the surface, 15 mL of electrolyte solution (0.1 M NaCl) was taken, and the pH was adjusted using buffer solutions. After adjusting the pH, 10 mg of the catalyst was added to the electrolyte solution and continuously stirred for 1 h. The final pH of the solution was taken using the pH meter [40].

2.8. Effect of pH

For 25 mL of a 15-ppm dye solution, the impact of pH was investigated. By employing a buffer solution with a different pH, the pH of the solution was adjusted. After setting the pH of each solution, 10 mg of catalyst was added and stirred for 30 min.

2.9. Effect of Temperature

The temperature of the fluid was regulated for this investigation using a water bath with flowing water. Then, 50 mL of 15 ppm solution of Congo red dye was taken and continuously stirred in the presence of 20 mg of the catalyst. The effect was examined by varying the temperature of the solution from 20 °C to 60 °C.

2.10. Quantitative Analysis of Data

The effect of each parameter on adsorption and removal was quantized by using some mathematical relations. The following equation was used to compute the adsorption capacity, qt (mg/g):
q e = C ° C t V W
where “qe” is adsorbing capacity (mg/g), “ C ° ” is the initial concentration of dye in solution, Ct is the concentration of dye in solution after time t, V is the volume of the solution in liters (L), W is mass of adsorbent in gram (g).
The percentage (%) removal of dye from the solution was evaluated by using the following equation:
Φ = C ° C t C ° × 100
where Φ is removal efficiency, C ° is the initial concentration of dye, and Ct is the concentration of dye after time.

3. Results and Discussion

In this work, Co3O4 and MDCO were synthesized using the co-precipitation method. MDCO-CNT composite was synthesized by combining CNTs with MDCO using the solution method. The photocatalytic performance of the synthesized samples was assessed in a UV reactor using an Hg lamp, while visible light was produced using a tungsten lamp (100 watts). System temperature was maintained at 25 °C by a circulating water bath and the distance between light sources and reaction solution was 15 cm. The Congo red solutions and synthesized powders were agitated for three hours in the dark before illumination to achieve the adsorption equilibrium. The dye removal process was evaluated and monitored by the UV-visible absorption spectrum and the concentration of the solution was analyzed quantitatively by measuring the maximum absorption at 497 nm.
XRD behavior of the synthesized materials was evaluated by matching with the standard reference pattern of Co3O4 (ICSD code-00-009-0418) and is given in Figure 1. According to the standard pattern, the intense peak was indexed as (311) at 36°. The absence of Mn or any other impurity shows the purity of the sample with good homogeneous dispersion of Mn in the Co3O4 crystals. Co3O4 showed a cubic structure. CNTs showed a characteristic peak (002) at 26°. The calculated value of lattice parameters is given in Table 1. No distortion in crystal structure after the addition of doping is observed [41] while a decrease in crystallinity by the addition of doping is observed, evident by a decrease in peak intensity.
Unit cell parameters were calculated by using this formula:
1 d 2   =   h 2 + k 2 + l 2 a 2
Unit cell volume was determined from data of unit cell parameters as:
V c e l l   =   a 3
The average crystallite size, D was calculated using Scherrer’s equation:
D = k λ β cos θ B
X-ray density (gcm−3) was determined by the following mathematical equation:
ρ x r a y = Z M V c e l l N A
where (d) is the d-spacing of the XRD pattern, (hkl) are corresponding Miller indices, (β) is the full width at half maximum of intensity, (λ) is the wavelength of Cu Kα X-rays, equal to 1.542 Å, (θ) is the Bragg’s angle (k) is a constant value = 0. 0.9 for the cubic system, (Z) is the number of molecules per formula unit (for cubic crystal structures it is 8), (M) is the molar mass of the synthesized compound, and (NA): 6.02 × 1023/mole is the Avogadro’s number. Values of different XRD parameters are given in Table 1
In undoped Co3O4, Co2+ (0.65 Å) the tetrahedral sites reside, while Co3+ (0.61 Å) occupies the octahedral sites. When dopant content is incorporated, Mn2+ replaces Co3+ at the tetrahedral site making spinel with a random distribution of cations at tetrahedral/octahedral sites causing expansion of the lattice. The general formula for random spinel can be written as below:
(Mn2+Co2+ Co3+) tetra [Mn2+Co3+ Co2+] OctaO4
X-ray density of parent material is greater compared to its doped derivatives because of reduced molecular masses.
Table 1 discusses the crystallographic aspects of the selected synthesized samples. Samples subjected to XRD and further analysis were coded thus: Undoped Co3O4 as Co3O4, MDCO as Mn0.8Co2.2O4, and Mn0.8Co2.2O4-CNT carbon nanotubes as MDCO-CNT composite.
SEM images of Co3O4, MDCO, and MDCO-CNT composite are given in Figure 2 and results show that parent Co3O4 has a heterogeneous surface while MDCO has flower-like morphology. MDCO-CNT composite micrograph shows that beads like metal oxides are distributed on threads like carbon nanotubes. These metal oxides are homogeneously distributed on the CNTs’ surface with very few agglomerations.
Band gap analysis of the synthesized materials showed a decrease in the band gap of Co3O4 by incorporating Mn was observed due to the replacement of Co3+ ions by Mn2+. As low-charge Mn2+ replaces the highly charged Co3+, it forms an acceptor level above the valence band, hence shrinking the band gap and making the electronic transition easier. Low Mn2+ doped Co3O4 showed a band gap value of 2.0 eV while highly doped Mn2+ has 1.9 eV.
As can be seen in Figure 3, the MDCO-CNT composite’s chemical composition and the electronic states of each of its constituent elements were assessed by employing XPS analysis. As shown in Figure 3a, the high-resolution XPS curve fitting of Co 2p shows distinctive peaks at 778.72 and 793.73 eV that are comparable to the Co3+ state’s Co 2p3/2 and Co 2p1/2 spin orbits. The coexistence of Co2+ and Co3+ states is indicated by the peaks at 780.93 and 796.61 eV, which were also attributed to the unpaired electrons present in the 3D state and the shake-up satellite peaks. Peaks at 529.96 and 531.45 eV are visible in Figure 3b’s high-resolution XPS spectra of the O1s curve fitting and are due to potent interfacial interactions with Co-O and C-O/-C. Figure 3c illustrates the Mn 2p3/2 and Mn 2p1/2 pattern’s contribution to the two distinct peaks at 641.51 and 653.37 eV in the high-resolution XPS curve fitting of Mn 2p. Figure 3d illustrates how the high-resolution XPS spectra of the C1s curve fitting revealed unique peaks at 284.71, 286.11, 288.81, and 289.57 eV, attributed to C-C, C-O, and C = O (O), respectively. The XPS study also showed that the MDCO-CNTs had Co3O4, Mn2+, and CNTs, which gave more weight to the SEM results. On the other hand, the BET surface area was assessed to be 21.84, 31.72, and 89.93 m2/g for all formulations: Co3O4, MDCO, and MDCO-CNTs (Figure 4).
UV-vis. spectra of Congo red before and after adsorption on MDCO and MDCO-CNT composite at different Concentrations are given in Figure 5a–c. These spectra show that both the materials have the tendency to adsorb dye, but MDCO-CNT composite has more efficiency than bare MDCO due to its large surface area, owing to the incorporation of CNTs.
The removal efficiency of the synthesized samples for dye is shown in Figure 6a,b. The removal efficiency of MDCO-CNT composite is greater than bare MDCO due to the high surface area of the composite than bare MDCO. Additionally, adsorption capacity increases with an increase in concentration up to a certain level because, after occupying all vacant sites, the adsorption process slows. Functionalized CNTs have π-bonds, carboxylic and hydroxyl groups which interact with dye via Van der Waal’s forces and π-π stacking. An increase in adsorption capacity (mg/g) of both materials with the increase in dye concentration is because of the concentration gradient and this leads to a reduction in the resistance in the mass transfer of dye molecules between liquid and solid phase [42].
In addition to the effects of dye concentration, the influence of a sample dose on dye removal performance (percent removal and adsorption capacity) has also been examined. The results are shown in Figure 6c,d. The efficiency is shown to grow linearly with the sample dose, as anticipated, and reaches up to 80% as a result of the improvement in adsorption sites with greater dosage [43].
The adsorption process depends on the nature of the dye as well as the nature of the adsorbent. Another crucial factor for dye adsorption research is the sample’s pH. For this adsorption experiment, manufactured samples were utilized to assess pH selectivity. Both dye molecules and adsorbents behave differently in acidic and basic mediums. Congo red is an anionic dye while in the acidic medium it becomes cationic because its pKa value is 4.1. To find the charge on the surface, 15 mL of electrolyte solution (0.1 M NaCl) was used, and pH was adjusted via buffer solutions. After adjusting the pH, 10 mg of catalyst was added to the electrolyte solution and continuously stirred for 1 h and the final pH of the solution was measured. PZC values for bare MDCO and MDCO-CNT composites are 8.1 and 7.2, which indicates that, below these values, both adsorbent materials carried a positive charge while above this point, the surface of the material becomes negatively charged. This charge variation is dependent on the pH of the medium and the structure of the material. In an acidic medium, both materials show cationic behavior because the oxygen of metal oxide captures proton from the solution and becomes positively charged (M-OH2+) as shown in the equation below:
nCo3O4 + nH2O → (Co3O4)n[OH]n + H+ → (Co3O4)n(OH2+)n
At the same time, in a basic medium, the oxygen of cobalt oxide becomes negative after losing a proton [44].
nCo3O4 + nH2O → (Co3O4)n[OH]n + OH → (Co3O4)n[O]n + H2O
In the case of the composite, the hydroxyl group on CNTs captures protons from the solution, and oxygen become positively charged in form of C-OH2+. On the other hand, in the basic medium hydroxyl and carboxylic group present in CNTs loses the proton and appear negative in form of (C-O and C-COO). Additionally, due to this charge variation, bare MDCO and MDCO-CNT composite have interactions with anionic and cationic dyes [45].
Figure 7b,c shows that maximum adsorption of the Congo red dye was observed at pH 6 which is 88 and 92%, respectively. At lower pH, Congo red dye becomes cationic and its pKa value is 4.1 whereas both MDCO and MDCO-CNT composites are positively charged below their PZC value, which is 8.2 and 7.1, respectively. Below this point, adsorbent and adsorbate cannot attract each other due to the same surface charges. Above pH 6, adsorption starts decreasing because above PZC value, the surface of the adsorbent becomes negatively charged and it repels the anionic dye molecules [46,47]. The result shows that both adsorbents almost have the same tendency to adsorb pollutant dye in an acidic medium, but in the basic medium at pH 10, MDCO-CNT composite shows very low adsorption efficiency than bare MDCO. This irregularity shows that the negative charge density on the surface of bare MDCO is less than that of the MDCO-CNT composite. Additionally, this negative charge density on the surface of the material in the basic medium repels the acidic dye and retards the adsorption process [48].
Contact time is an important factor for the photocatalytic removal study of dye on photocatalysts. Calculated data show that the removal efficiency of Congo red dye on MDCO and MDCO-CNT composites increases with an increase in time, as with time, more and more radicals were generated because light interacts with the photocatalysts and degrades the dye molecules [49]. Figure 8a,b illustrates how contact time affects dye adsorption. Since more dye molecules may interact with the adsorbent over time, the adsorption of dye rises with time. Because there are more accessible unoccupied sites at first, the rate of adsorption is greater. Later, the rate of adsorption decreases because an adsorbate layer forms on the surface of the adsorbent, and there are fewer open sites for subsequent interactions [50].
In addition to contact duration, adsorption tests were conducted at five different temperatures to examine the impact of temperature on the adsorption process. The rate of adsorption rises with temperature owing to an increase in Congo red solubility, as seen in Figure 8c,d. Additionally, when the temperature rises, the number of active sites on the adsorbent’s surface and the mobility of the dye molecules in the solution both increase. Maximum adsorption was observed at 323 K. After 323 K, the further increase in temperature led to a decrease in removal efficiency because at higher temperatures dye molecules start to detach from the surface due to the breaking of weak bonds between dye and composite. MDCO-CNT composite is a more efficient adsorbent than bare MDCO because of the larger surface area [50,51].
Data fitted to Temkin isotherm show that MDCO-CNT and MDCO do not follow the pattern which is a clue that there is non-uniformity in the distribution of bonding energies on the surface of the material. The Freundlich isotherm is shown in Figure 9c,d for MDCO-CNT composite and bare MDCO, respectively. The various functional groups attached to the surfaces of the adsorbent are connected to the heterogeneity on those surfaces. Calculated values of regression coefficient R2 and n value shows that both MDCO and MDCO-CNT composite does not follow the Freundlich isotherm showing that the synthesized materials followed the homogenous adsorption.
According to Figure 9d,e, both bare MDCO and MDCO-CNT composites follow the Langmuir isotherm which gives an idea about the mechanism of adsorption. Dye molecules are chemically adsorbed on the surface of the material. Metal cations in MDCO interact with negatively charged sulfate ions of Congo red dye through electrostatic force of attraction and in the case of MDCO-CNT, both MDCO and CNT are involved in the adsorption of dye. Carbon nanotubes capture Congo red dye through π-π stacking. The calculated value of RL for bare MDCO and MDCO-CNT composites given in Table 2 are less than 1 and greater than 0 which is 0.1955 and 0.2364, respectively. Therefore, the adsorption process is favorable for both materials [52].
Both MDCO and MDCO-CNT composites follow the pseudo-second-order kinetic model as shown in Figure 9a,b. Parameters of pseudo-second order given in Table 3 and calculated data suggest that dye molecules are adsorbed through chemisorption. Both adsorbent and adsorbate show chemical interaction through the exchange or sharing of electrons. Positive charge metal ions in MDCO and Congo red comprehend sulfate ions with negative oxygen. During the process of adsorption adsorbent and adsorbate, dye interacts with each through the electrostatic force of attraction. On the other hand, MDCO-CNT composite contains metal oxide and functionalized CNTs. These CNT can adsorb Congo red dye through π-π stacking and p-type metal oxide captures dye molecules electrostatically [53].
Van’t Hoff plots for the elimination of Congo red dye on MDCO and MDCO-CNT composite are shown in Figure 10a,b. The values of Gibb’s free energy, enthalpy, and entropy that were determined using these plots are shown in Table 4 below. According to the data, the Gibbs free energy (ΔG°) and enthalpy (ΔH°) values for both materials are negative, suggesting an exothermic and spontaneous reaction. A drop in temperature favors accelerating the photodegradation of Congo red dye. In the case of MDCO value of entropy is positive indicating an increase in the degree of randomness and for the composite, its value is negative because randomness decreases with the degradation of the dye [54].
Thermodynamic parameters i.e., enthalpy (∆H°), Gibb’s free energy (∆G°), and entropy (∆S°) were calculated by using the Van’t Hoff’s plot which is given in Figure 8a,b. Values of these parameters mentioned in Table 4 express that the adsorption process is spontaneous and exothermic because the value of Gibb’s free energy and enthalpy is negative [55]. Therefore, by increasing the temperature up to a certain level, the rate of adsorption increases as the solubility of dye increases with increasing temperature but at higher temperatures, dye molecules start disengaging from the surface of the adsorbent. Thermodynamic data depict as well that dye molecules chemically adsorbed on MDCO and MDCO-CNT. In chemisorption a strong interaction between dye molecules and adsorbent is present and this interaction is stable at lower and higher temperatures as well.
Studies were conducted at five different temperatures, 293, 303, 313, 323, and 333 K, as shown in Figure 11, to investigate the impact of temperature (a,b). Results show that an increase in temperature reduces the removal capacity of photocatalysts because this process is exothermic and favorable at low temperatures. Maximum adsorption for both MDCO and MDCO-CNT composite is seen at 50 min owing to the formation of equilibrium after 60 min, at which point adsorption capacity becomes constant, as shown in Figure 11c,d. This is because the rate of adsorption rises with increasing contact time.

4. Conclusions

Cobalt oxide (Co3O4), Mn-doped cobalt oxide (MDCO), and Mn-doped cobalt oxide functionalized carbon nanotube (MDCO-CNTs) composites were successfully synthesized using soft chemistry and Congo red dye was chosen as a dye model to evaluate the removal efficiency of the synthesized materials. Adsorption-assisted photocatalytic performances of the synthesized samples for the removal of Congo red dye were evaluated by UV–Vis. spectrophotometry. The best removal performance (95%) was obtained by MDCO-CNTs (Mn0.8Co2.2O4-CNT) nanocomposites. This improved photocatalytic activity was attributed to the synergistic effect of CNTs and Mn incorporation in cobalt oxide, in which Co3O4 absorbs photons, whereas Mn incorporation contributed to enhancing the photon absorbance and CNTs assisted in the adsorption procedure. The experimental results showed correlations with models for traditional, Freundlich, and Temkin adsorption. Additionally, analysis was conducted utilizing both the pseudo-first-order and pseudo-second-order kinetic models. According to the correlation coefficient, the pseudo-first-order model is more likely to match the kinetic data. This study can be helpful for the synthesis of novel nanocomposites and facilitates their application by the synergistic (adsorption + catalysis) removal of dye in the field of water treatment.

Author Contributions

Conceptualization, writing-original draft Preparation, M.F. and S.I.; methodology, M.S.; software, writing-original draft Preparation, T.F.; validation, project administration, writing review and editing, B.I., H.u.R. and F.F.A.-F.; writing review and editing, E.B.E.; resources, R.A.P.; data curation, E.A.; writing-original draft preparation, A.-E.F. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to thank the Deanship of Scientific Research at Umm Al-Qura University for supporting this work by Grant Code: (22UQU4320141DSR43). This research was funded by Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2022R156), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data will be available on request.

Acknowledgments

The authors would like to thank the Deanship of Scientific Research at Umm Al-Qura University for supporting this work by Grant Code: (22UQU4320141DSR43). This research was funded by Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2022R156), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Morin-Crini, N.; Lichtfouse, E.; Liu, G.; Balaram, V.; Ribeiro, A.R.L.; Lu, Z.; Stock, F.; Carmona, E.; Teixeira, M.R.; Picos-Corrales, L.A.; et al. Worldwide cases of water pollution by emerging contaminants: A review. Environ. Chem. Lett. 2022, 20, 2311–2338. [Google Scholar] [CrossRef]
  2. Qamar, M.A.; Javed, M.; Shahid, S.; Iqbal, S.; Abubshait, S.A.; Abubshait, H.A.; Ramay, S.M.; Mahmood, A.; Ghaithan, H.M. Designing of highly active g-C3N4/Co@ZnO ternary nanocomposites for the disinfection of pathogens and degradation of the organic pollutants from wastewater under visible light. J. Environ. Chem. Eng. 2021, 9, 105534. [Google Scholar] [CrossRef]
  3. Hussain, W.; Malik, H.; Hussain, R.A.; Hussain, H.; Green, I.R.; Marwat, S.; Bahadur, A.; Iqbal, S.; Farooq, M.U.; Li, H.; et al. Synthesis of MnS from single-and multi-source precursors for photocatalytic and battery applications. J. Electron. Mater. 2019, 48, 2278–2288. [Google Scholar] [CrossRef]
  4. Athanasekou, C.P.; Likodimos, V.; Falaras, P. Recent developments of TiO2 photocatalysis involving advanced oxidation and reduction reactions in water. J. Environ. Chem. Eng. 2018, 6, 7386–7394. [Google Scholar] [CrossRef]
  5. Kim, S.H.; An, H.-R.; Lee, M.; Hong, Y.; Shin, Y.; Kim, H.; Kim, C.; Park, J.-I.; Son, B.; Jeong, Y.; et al. High removal efficiency of industrial toxic compounds through stable catalytic reactivity in water treatment system. Chemosphere 2021, 287, 132204. [Google Scholar] [CrossRef]
  6. Sher, M.; Javed, M.; Shahid, S.; Hakami, O.; Qamar, M.A.; Iqbal, S.; AL-Anazy, M.M.; Baghdadi, H.B. Designing of highly active g-C3N4/Sn doped ZnO heterostructure as a photocatalyst for the disinfection and degradation of the organic pollutants under visible light irradiation. J. Photochem. Photobiol. A Chem. 2021, 418, 113393. [Google Scholar] [CrossRef]
  7. Irfan, R.M.; Tahir, M.H.; Nadeem, M.; Maqsood, M.; Bashir, T.; Iqbal, S.; Zhao, J.; Gao, L. Fe3C/CdS as noble-metal-free composite photocatalyst for highly enhanced photocatalytic H2 production under visible light. Appl. Catal. A Gen. 2020, 603, 117768. [Google Scholar] [CrossRef]
  8. Morais, E.; O’Modhrain, C.; Thampi, K.R.; Sullivan, J.A. RuO2/TiO2 photocatalysts prepared via a hydrothermal route: Influence of the presence of TiO2 on the reactivity of RuO2 in the artificial photosynthesis reaction. J. Catal. 2021, 401, 288–296. [Google Scholar] [CrossRef]
  9. Yang, Y.; Zhao, S.; Bi, F.; Chen, J.; Li, Y.; Cui, L.; Xu, J.; Zhang, X. Oxygen-vacancy-induced O2 activation and electron-hole migration enhance photothermal catalytic toluene oxidation. Cell Rep. Phys. Sci. 2022, 3, 101011. [Google Scholar] [CrossRef]
  10. Mansurov, R.; Zverev, V.; Safronov, A. Dynamics of diffusion-limited photocatalytic degradation of dye by polymeric hydrogel with embedded TiO2 nanoparticles J. Catal. 2022, 406, 9–18. [Google Scholar] [CrossRef]
  11. Dubsok, A.; Khamdahsag, P.; Kittipongvises, S. Life cycle environmental impact assessment of cyanate removal in mine tailings wastewater by nano-TiO2/FeCl3 photocatalysis. J. Clean. Prod. 2022, 366, 132928. [Google Scholar] [CrossRef]
  12. Niyitanga, T.; Kim, H. Hematite-nickel oxide/carbon nanotube composite catalyst for oxygen evolution reaction. Mater. Chem. Phys. 2021, 275, 125266. [Google Scholar] [CrossRef]
  13. Keller, N.; Rebmann, G.; Barraud, E.; Zahraa, O.; Keller, V. Macroscopic carbon nanofibers for use as photocatalyst support. Catal. Today 2005, 101, 323–329. [Google Scholar] [CrossRef]
  14. Wang, W.; Serp, P.; Kalck, P.; Faria, J.L. Photocatalytic degradation of phenol on MWNT and titania composite catalysts prepared by a modified sol–gel method. Appl. Catal. B Environ. 2005, 56, 305–312. [Google Scholar] [CrossRef]
  15. Wang, W.; Serp, P.; Kalck, P.; Faria, J.L. Visible light photodegradation of phenol on MWNT-TiO2 composite catalysts prepared by a modified sol–gel method. J. Mol. Catal. A Chem. 2005, 235, 194–199. [Google Scholar] [CrossRef]
  16. Yu, Y.; Yu, J.C.; Chan, C.-Y.; Che, Y.-K.; Zhao, J.-C.; Ding, L.; Ge, W.-K.; Wong, P.-K. Enhancement of adsorption and photocatalytic activity of TiO2 by using carbon nanotubes for the treatment of azo dye. Appl. Catal. B Environ. 2005, 61, 1–11. [Google Scholar] [CrossRef]
  17. Yu, Y.; Yu, J.C.; Yu, J.-G.; Kwok, Y.-C.; Che, Y.-K.; Zhao, J.-C.; Ding, L.; Ge, W.-K.; Wong, P.-K. Enhancement of photocatalytic activity of mesoporous TiO2 by using carbon nanotubes. Appl. Catal. A Gen. 2005, 289, 186–196. [Google Scholar] [CrossRef]
  18. Krishna, V.; Pumprueg, S.; Lee, S.-H.; Zhao, J.; Sigmund, W.; Koopman, B.; Moudgil, B. Photocatalytic Disinfection with Titanium Dioxide Coated Multi-Wall Carbon Nanotubes. Process Saf. Environ. Prot. 2005, 83, 393–397. [Google Scholar] [CrossRef]
  19. Gao, B.; Peng, C.; Chen, G.; Puma, G.L. Photo-electro-catalysis enhancement on carbon nanotubes/titanium dioxide (CNTs/TiO2) composite prepared by a novel surfactant wrapping sol–gel method. Appl. Catal. B Environ. 2008, 85, 17–23. [Google Scholar] [CrossRef]
  20. Kuo, C.-Y. Prevenient dye-degradation mechanisms using UV/TiO2/carbon nanotubes process. J. Hazard. Mater. 2009, 163, 239–244. [Google Scholar] [CrossRef]
  21. Wang, H.; Wang, H.L.; Jiang, W.F. Solar photocatalytic degradation of 2,6-dinitro-p-cresol (DNPC) using multi-walled carbon nanotubes (MWCNTs)–TiO2 composite photocatalysts. Chemosphere 2009, 75, 1105–1111. [Google Scholar] [CrossRef] [PubMed]
  22. Gao, B.; Chen, G.; Puma, G.L. Carbon nanotubes/titanium dioxide (CNTs/TiO2) nanocomposites prepared by conventional and novel surfactant wrapping sol–gel methods exhibiting enhanced photocatalytic activity. Appl. Catal. B Environ. 2009, 89, 503–509. [Google Scholar] [CrossRef]
  23. Jiang, L.; Gao, L. Fabrication and characterization of ZnO-coated multi-walled carbon nanotubes with enhanced photocatalytic activity. Mater. Chem. Phys. 2005, 91, 313–316. [Google Scholar] [CrossRef]
  24. Zhu, L.-P.; Liao, G.-H.; Huang, W.-Y.; Ma, L.-L.; Yang, Y.; Yu, Y.; Fu, S.-Y. Preparation, characterization and photocatalytic properties of ZnO-coated multi-walled carbon nanotubes. Mater. Sci. Eng. B 2009, 163, 194–198. [Google Scholar] [CrossRef]
  25. Yu, Y.; Ma, L.-L.; Huang, W.-Y.; Du, F.-P.; Yu, J.C.; Yu, J.-G.; Wang, J.-B.; Wong, P.-K. Sonication assisted deposition of Cu2O nanoparticles on multiwall carbon nanotubes with polyol process. Carbon 2005, 43, 670–673. [Google Scholar] [CrossRef]
  26. Wang, N.; Xu, J.; Guan, L. Synthesis and enhanced photocatalytic activity of tin oxide nanoparticles coated on multi-walled carbon nanotube. Mater. Res. Bull. 2011, 46, 1372–1376. [Google Scholar] [CrossRef]
  27. Yu, F.; Chen, J.; Chen, L.; Huai, J.; Gong, W.; Yuan, Z.; Wang, J.; Ma, J. Magnetic carbon nanotubes synthesis by Fenton’s reagent method and their potential application for removal of azo dye from aqueous solution. J. Colloid Interface Sci. 2012, 378, 175–183. [Google Scholar] [CrossRef]
  28. Meng, Z.; Oh, W. Photodegradation of Organic Dye by CoS2 and Carbon(C60, Graphene, CNT)/TiO2 Composite Sensitizer. Chin. J. Catal. 2012, 33, 1495–1501. [Google Scholar] [CrossRef]
  29. Xiong, P.; Fu, Y.; Wang, L.; Wang, X. Multi-walled carbon nanotubes supported nickel ferrite: A magnetically recyclable photocatalyst with high photocatalytic activity on degradation of phenols. Chem. Eng. J. 2012, 195-196, 149–157. [Google Scholar] [CrossRef]
  30. Zhu, L.; Meng, Z.-D.; Cho, K.-Y.; Oh, W.-C. Synthesis of CdS/CNT-TiO2 with a high photocatalytic activity in photodegradation of methylene blue. New Carbon Mater. 2012, 27, 166–174. [Google Scholar] [CrossRef]
  31. Di Paola, A.; García-López, E.; Marcì, G.; Palmisano, L. A survey of photocatalytic materials for environmental remediation. J. Hazard. Mater. 2012, 211–212, 3–29. [Google Scholar] [CrossRef] [PubMed]
  32. Irfan, R.M.; Tahir, M.H.; Iqbal, S.; Nadeem, M.; Bashir, T.; Maqsood, M.; Zhao, J.; Gao, L. Co3C as a promising cocatalyst for superior photocatalytic H2 production based on swift electron transfer processes. J. Mater. Chem. C 2021, 9, 3145–3154. [Google Scholar] [CrossRef]
  33. Abubshait, S.A.; Iqbal, S.; Abubshait, H.A.; AlObaid, A.A.; Al-Muhimeed, T.I.; Abd-Rabboh, H.S.; Bahadur, A.; Li, W. Effective heterointerface combination of 1D/2D Co-NiS/S-g-C3N4 heterojunction for boosting spatial charge separation with enhanced photocatalytic degradation of organic pollutants and disinfection of pathogens. Colloids Surfaces A Physicochem. Eng. Asp. 2021, 628, 127390. [Google Scholar] [CrossRef]
  34. Qian, H.; Anwer, S.; Bharath, G.; Iqbal, S.; Chen, L. Nanoporous Ag-Au Bimetallic triangular Nano-prisms Synthesized by Galvanic Replacement for Plasmonic Applications. J. Nanomater. 2018, 2018, 1263942. [Google Scholar]
  35. Iqbal, S.; Javed, M.; Hassan, S.S.; Nadeem, S.; Akbar, A.; Alotaibi, M.T.; Alzhrani, R.M.; Awwad, N.S.; Ibrahium, H.A.; Mohyuddin, A. Binary Co@ZF/S@GCN S-scheme heterojunction enriching spatial charge carrier separation for efficient removal of organic pollutants under sunlight irradiation. Colloids Surfaces A Physicochem. Eng. Asp. 2022, 636, 128177. [Google Scholar] [CrossRef]
  36. Iqbal, S. Spatial charge separation and transfer in L-cysteine capped NiCoP/CdS nano-heterojunction activated with intimate covalent bonding for high-quantum-yield photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2020, 274, 119097. [Google Scholar] [CrossRef]
  37. Tang, C.; Liu, E.; Wan, J.; Hu, X.; Fan, J. Co3O4 nanoparticles decorated Ag3PO4 tetrapods as an efficient visible-light-driven heterojunction photocatalyst. Appl. Catal. B Environ. 2016, 181, 707–715. [Google Scholar] [CrossRef]
  38. Lee, D.-S.; Chen, H.-J.; Chen, Y.-W. Photocatalytic reduction of carbon dioxide with water using InNbO4 catalyst with NiO and Co3O4 cocatalysts. J. Phys. Chem. Solids 2012, 73, 661–669. [Google Scholar] [CrossRef]
  39. Huang, J.; Ren, H.; Chen, K.; Shim, J.-J. Controlled synthesis of porous Co3O4 micro/nanostructures and their photocatalysis property. Superlattices Microstruct. 2014, 75, 843–856. [Google Scholar] [CrossRef]
  40. Reymond, J.; Kolenda, F. Estimation of the point of zero charge of simple and mixed oxides by mass titration. Powder Technol. 1999, 103, 30–36. [Google Scholar] [CrossRef]
  41. Patil, D.; Patil, P.; Subramanian, V.; Joy, P.A.; Potdar, H.S. Highly sensitive and fast responding CO sensor based on Co3O4 nanorods. Talanta 2010, 81, 37–43. [Google Scholar] [CrossRef]
  42. Wanyonyi, W.C.; Onyari, J.M.; Shiundu, P.M. Adsorption of Congo Red Dye from aqueous solutions using roots of Eichhornia crassipes: Kinetic and Equilibrium studies. Energy Procedia 2014, 50, 862–869. [Google Scholar] [CrossRef] [Green Version]
  43. Nimkar, D.; Chavan, S. Removal of Congo red dye from aqueous solution by using saw dust as an adsorbent. Int. J. Eng. Res. Appl. 2014, 4, 47–51. [Google Scholar]
  44. Akpan, U.; Hameed, B.H. Parameters affecting the photocatalytic degradation of dyes using TiO2-based photocatalysts: A review. J. Hazard. Mater. 2009, 170, 520–529. [Google Scholar] [CrossRef] [PubMed]
  45. Zhu, H.-Y.; Jiang, R.; Xiao, L. Adsorption of an anionic azo dye by chitosan/kaolin/γ-Fe2O3 composites. Appl. Clay Sci. 2010, 48, 522–526. [Google Scholar] [CrossRef]
  46. Purkait, M.; Maiti, A.; DasGupta, S.; De, S. Removal of congo red using activated carbon and its regeneration. J. Hazard. Mater. 2007, 145, 287–295. [Google Scholar] [CrossRef]
  47. Kamil, A.M.; Mohammed, H.T.; Alkaim, A.F.; Hussein, F.H. Adsorption of Congo red on multiwall carbon nanotubes: Effect of operational parameters. J. Chem. Pharm. Sci. 2016, 9, 1128–1133. [Google Scholar]
  48. Rezaei, H.; Razavi, A.; Shahbazi, A. Removal of Congo red from aqueous solutions using nano-Chitosan. Environ. Resour. Res. 2017, 5, 25–34. [Google Scholar]
  49. Erdemoğlu, S.; Aksu, S.K.; Sayılkan, F.; Izgi, B.; Asiltürk, M.; Sayılkan, H.; Frimmel, F.; Güçer, Ş. Photocatalytic degradation of Congo Red by hydrothermally synthesized nanocrystalline TiO2 and identification of degradation products by LC–MS. J. Hazard. Mater. 2008, 155, 469–476. [Google Scholar] [CrossRef]
  50. Bhoi, S.K. Adsorption Characteristics of Congo Red Dye onto PAC and GAC based on S/N ratio: A Taguchi Approach. Ph.D. Thesis, The National Institute of Technology, Rourkela, Odisha, 2010. [Google Scholar]
  51. Yagub, M.T.; Sen, T.K.; Afroze, S.; Ang, H.M. Dye and its removal from aqueous solution by adsorption: A review. Adv. Colloid Interface Sci. 2014, 209, 172–184. [Google Scholar] [CrossRef]
  52. Shan, R.-R.; Yan, L.-G.; Yang, Y.-M.; Yang, K.; Yu, S.-J.; Yu, H.-Q.; Zhu, B.-C.; Du, B. Highly efficient removal of three red dyes by adsorption onto Mg–Al-layered double hydroxide. J. Ind. Eng. Chem. 2015, 21, 561–568. [Google Scholar] [CrossRef]
  53. Ho, Y.-S. Review of second-order models for adsorption systems. J. Hazard. Mater. 2006, 136, 681–689. [Google Scholar] [CrossRef] [Green Version]
  54. Roy, T.K.; Mondal, N.K. Photocatalytic degradation of congo red dye on thermally activated zinc oxide. Int. J. Sci. Res. Environ. Sci. 2014, 2, 457–469. [Google Scholar] [CrossRef]
  55. Wang, L.; Li, J.; Wang, Y.; Zhao, L.; Jiang, Q. Adsorption capability for Congo red on nanocrystalline MFe2O4 (M = Mn, Fe, Co, Ni) spinel ferrites. Chem. Eng. J. 2012, 181, 72–79. [Google Scholar] [CrossRef]
Figure 1. Powder XRD patterns for Co3O4, MDCO, and MDCO-CNTs.
Figure 1. Powder XRD patterns for Co3O4, MDCO, and MDCO-CNTs.
Sustainability 14 16932 g001
Figure 2. Micrographs of (a) Co3O4, (b) MDCO, and (c) MDCO-CNT.
Figure 2. Micrographs of (a) Co3O4, (b) MDCO, and (c) MDCO-CNT.
Sustainability 14 16932 g002
Figure 3. High-resolution XPS spectra of MDCO-CNTs; (a) Co 2p, (b) O 1s, (c) Mn 2p and (d) C 1s.
Figure 3. High-resolution XPS spectra of MDCO-CNTs; (a) Co 2p, (b) O 1s, (c) Mn 2p and (d) C 1s.
Sustainability 14 16932 g003
Figure 4. The BET surface area isotherms estimated from N2 adsorption-desorption of Co3O4, MDCO, and MDCO-CNTs.
Figure 4. The BET surface area isotherms estimated from N2 adsorption-desorption of Co3O4, MDCO, and MDCO-CNTs.
Sustainability 14 16932 g004
Figure 5. UV-visible absorbance pattern of CR dye (a) at different concentrations, (b) after adsorption on MDCO, and (c) after adsorption on MDCO-CNT composite.
Figure 5. UV-visible absorbance pattern of CR dye (a) at different concentrations, (b) after adsorption on MDCO, and (c) after adsorption on MDCO-CNT composite.
Sustainability 14 16932 g005
Figure 6. (a) % removal of CR on MDCO and MDCO-CNT; (b) adsorption capacity at various concentrations of CR dye; (c) % removal of CR; (d) effect of adsorbent mass on the removal of CR.
Figure 6. (a) % removal of CR on MDCO and MDCO-CNT; (b) adsorption capacity at various concentrations of CR dye; (c) % removal of CR; (d) effect of adsorbent mass on the removal of CR.
Sustainability 14 16932 g006aSustainability 14 16932 g006b
Figure 7. (a) PZC values, (b) effect of pH on removal efficiency, and (c) on adsorption of CR dye on MDCO and MDCO-CNT composite.
Figure 7. (a) PZC values, (b) effect of pH on removal efficiency, and (c) on adsorption of CR dye on MDCO and MDCO-CNT composite.
Sustainability 14 16932 g007
Figure 8. (a) Effect of contact time on % removal of CR dye; (b) Effect of contact time on adsorption of CR on MDCO and MDCO-CNT composite; (c) Effect of temperature on % removal of CR dye; (d) Effect of temperature on adsorption of CR dye on MDCO and MDCO-CNT composite.
Figure 8. (a) Effect of contact time on % removal of CR dye; (b) Effect of contact time on adsorption of CR on MDCO and MDCO-CNT composite; (c) Effect of temperature on % removal of CR dye; (d) Effect of temperature on adsorption of CR dye on MDCO and MDCO-CNT composite.
Sustainability 14 16932 g008aSustainability 14 16932 g008b
Figure 9. (a) For adsorption of CR, data fitted to Temkin isotherm on MDCO-CNT composite; (b) Temkin isotherm on bare MDCO; (c) Freundlich isotherm on MDCO-CNT composite; (d) Freundlich isotherm on bare MDCO; (e) Langmuir isotherm MDCO-CNT composite; (f) Langmuir isotherm on bare MDCO.
Figure 9. (a) For adsorption of CR, data fitted to Temkin isotherm on MDCO-CNT composite; (b) Temkin isotherm on bare MDCO; (c) Freundlich isotherm on MDCO-CNT composite; (d) Freundlich isotherm on bare MDCO; (e) Langmuir isotherm MDCO-CNT composite; (f) Langmuir isotherm on bare MDCO.
Sustainability 14 16932 g009aSustainability 14 16932 g009b
Figure 10. For the adsorption of CR, Van’t Hoff plot of lnKdvs 1/T (a) on bare MDCO, (b) on MDCO-CNT composite.
Figure 10. For the adsorption of CR, Van’t Hoff plot of lnKdvs 1/T (a) on bare MDCO, (b) on MDCO-CNT composite.
Sustainability 14 16932 g010
Figure 11. (a) Effect of temperature on % removal; (b) effect of temperature on removal; (c) effect of contact time on removal by MDCO; (d) effect of contact time on removal by MDCO-CNT composite.
Figure 11. (a) Effect of temperature on % removal; (b) effect of temperature on removal; (c) effect of contact time on removal by MDCO; (d) effect of contact time on removal by MDCO-CNT composite.
Sustainability 14 16932 g011
Table 1. XRD analysis of Co3O4, Mn0.8Co2.2O4, and Mn0.8Co2.2O4-CNT.
Table 1. XRD analysis of Co3O4, Mn0.8Co2.2O4, and Mn0.8Co2.2O4-CNT.
SampleLattice Constant(Å)
8.084 *
Crystallite Size (nm)Vcell3) 528.30 *X-ray Density
(g/cm3)
6.05 *
Co3O48.1029532.266.11
MDCO8.1927549.105.92
MDCO-CNT8.0650523.305.93
*: Standard values matched with JCPDS (00-009-0418).
Table 2. Langmuir isotherm study.
Table 2. Langmuir isotherm study.
Material NameR2qm (mg/g)B (L/mg)RL
MDCO-CNT0.95742.1940.1560.236
MDCO0.90430.300.2090.195
Table 3. Kinetics study (pseudo 2nd order) for adsorption of CR on MDCO and MDCO-CNT.
Table 3. Kinetics study (pseudo 2nd order) for adsorption of CR on MDCO and MDCO-CNT.
MaterialR2K2
(g.mg−1.min−1)
qe
(g.mg−1)
MDCO0.9940.04421.92
MDCO-CNT0.9900.02119.88
Table 4. Thermodynamic analysis for adsorption of CR dye on bare MDCO and MDCO-CNT composite.
Table 4. Thermodynamic analysis for adsorption of CR dye on bare MDCO and MDCO-CNT composite.
Temperature (K)MDCOMDCO-CNT
∆G° (kJmol−1)∆H° (kJmol−1)∆S°
(Jmol−1 K−1)
∆G°
(kJmol−1)
∆H°
(kJmol−1)
∆S°
(Jmol−1 K−1)
293−9.116 −9.336
303−9.614 −10.10
313−10.724−15.34083.157−11.124−24.857119.42
323−11.830 −12.04
333−12.157 −11.81
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fida, M.; Iqbal, S.; Shah, M.; Fazal, T.; Ismail, B.; Rehman, H.u.; Al-Fawzan, F.F.; Elkaeed, E.B.; Pashameah, R.A.; Alzahrani, E.; et al. Mn2+ Doped Cobalt Oxide and Its Composite with Carbon Nanotubes for Adsorption-Assisted Photocatalytic Applications. Sustainability 2022, 14, 16932. https://doi.org/10.3390/su142416932

AMA Style

Fida M, Iqbal S, Shah M, Fazal T, Ismail B, Rehman Hu, Al-Fawzan FF, Elkaeed EB, Pashameah RA, Alzahrani E, et al. Mn2+ Doped Cobalt Oxide and Its Composite with Carbon Nanotubes for Adsorption-Assisted Photocatalytic Applications. Sustainability. 2022; 14(24):16932. https://doi.org/10.3390/su142416932

Chicago/Turabian Style

Fida, Muhammad, Shahid Iqbal, Mazloom Shah, Tanzeela Fazal, Bushra Ismail, Hafiz ur Rehman, Foziah F. Al-Fawzan, Eslam B. Elkaeed, Rami Adel Pashameah, Eman Alzahrani, and et al. 2022. "Mn2+ Doped Cobalt Oxide and Its Composite with Carbon Nanotubes for Adsorption-Assisted Photocatalytic Applications" Sustainability 14, no. 24: 16932. https://doi.org/10.3390/su142416932

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