Next Article in Journal
L-Proline-Based Natural Deep Eutectic Solvents as Efficient Solvents and Catalysts for the Ultrasound-Assisted Synthesis of Aurones via Knoevenagel Condensation
Previous Article in Journal
Improvement in Structure and Visible Light Catalytic Performance of g-C3N4 Fabricated at a Higher Temperature
Previous Article in Special Issue
Highly Active Mo2C@WS2 Hybrid Electrode for Enhanced Hydrogen Evolution Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photodegradation of Congo Red by Modified P25-Titanium Dioxide with Cobalt-Carbon Supported on SiO2 Matrix, DFT Studies of Chemical Reactivity

by
Hassan H. Hammud
1,*,
Hassan Traboulsi
1,
Ranjith Kumar Karnati
1 and
Esam M. Bakir
1,2,*
1
Department of Chemistry, College of Science, King Faisal University, P.O. Box 400, Al-Ahsa 31982, Saudi Arabia
2
Department of Chemistry, Faculty of Science, Ain shams University, Cairo 11566, Egypt
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(3), 248; https://doi.org/10.3390/catal12030248
Submission received: 30 December 2021 / Revised: 16 February 2022 / Accepted: 16 February 2022 / Published: 22 February 2022

Abstract

:
Congo red is a hazardous material in the environment. The present work describes the synthesis of TiO2/CoC@SiO2-bipy (1) and TiO2/CoC@SiO2-phen (2) nanocomposites for the photodegradation of azo-dye Congo red (CR) dye in aqueous solution, by combining pure TiO2 with CoC@SiO2-bipy (s1) and CoC@SiO2-phen (s2) nanoparticles. The prepared nanocomposites were evaluated in term of photocatalytic activity rates in aqueous solution using CR. The nanocomposites TiO2/CoC@SiO2-bipy (1) and TiO2/CoC@SiO2-phen (2) were prepared from TiO2 (75%) and CoC@SiO2-bipy (s1) or CoC@SiO2-phen (s2) (25%) (weight ratio). Ultra-sonication and milling were used to prepare the heterogeneous nano catalysts. The pH, initial dye concentration, and catalyst dosage appeared to have a significant impact on the photocatalytic degradation performance. Molecular oxygen and other active species played a significant role in the photocatalyst degradation of CR with sunlight energy (UV-index 5.0). The chemical reactions were accelerated depending upon electrophilicity (ω) and energy gap (Eg) of azo dye species CR-N=N, CR-N=NH and CR=N-NH species which were calculated by density function theory (DFT). It can be concluded that the rate of electron–hole recombination of the TiO2 catalyst, when adding CoC@SiO2-bipy (s1) or CoC@SiO2-phen (s2), not only enhances the degradation but also effectively removes toxic dye molecules and their by-products. The newly prepared TiO2/CoC@SiO2-bipy (1) nanocomposites showed increased photocatalytic efficiency at low catalyst dose and faster rate of degradation of Congo red compared to TiO2/CoC@SiO2-phen (2) and TiO2 catalysts. The novel catalysts (1) and (2) can be easily separated by centrifugation and filtration, from the reaction mixture compared to TiO2.

1. Introduction

Textile dyes and other industrial dyestuffs constitute one of the largest groups of organic compounds that represent an increasing environmental danger. Azo dye is the main class of synthetic dyes which represents about 90% of all organic colorant [1]. One of the most common azo dyes used is Congo-red dye (CR, sodium salt of benzidine diazo bis-1-naphthylamine-4-sulfonic acid). Large amounts of Congo-red dye are reportedly used in textile, rubber, cosmetics, and plastic industries and consequently discharged into water resources [2]. The presence of Congo-red dye in water is reported to be non-degradable and causes several environmental and health related problems, such as cancer, gene alteration, and lung- and kidney diseases [2]. Although several methods have been used for the removal of dyes from wastewater, they had serious disadvantages or limitations such as consuming a lot of electricity or forming various byproducts [3].
TiO2 is commonly served as the photocatalyst for dye degradation in the presence of ultraviolet (UV) radiation. TiO2 nanoparticles possess high chemical stability and resistance in acidic and alkaline medium, can be prepared easily, and are safe due to their non-poisonous property [4]. Under visible light illumination, TiO2 is inactive owing to its wide band gap energy between the intervals 3.0–3.2 eV [5]. High energy is required to excite the electrons from the valence band to the conduction band. Thus, wide band gap energy limits the application of pure TiO2 nanoparticles. Therefore, to overcome this issue, the band structure of TiO2 should be altered. For this purpose several methodologies such as metal/non-metal elements doping have been developed to enhance the TiO2 photoactivity in the visible light region [6]. Promotion by different types of transition metals (e.g., Pd, Ag, Pt, Au, Rh, and Ru) [7] and transition metal oxides (e.g., CoO, Co3O4, Fe2O3, and CuO) [8] have been considered to enhance a visible light photoactivity in TiO2.
An increase in the surface area of TiO2 particles causes an increase in the generation of hydroxyl radicals. Thus, TiO2-immobilized on cellulose grains showed improved antibacterial, photocatalytic, electrical, and magnetic properties [9]. On the other hand, carbon is considered as the most widely applied support for heterogeneous catalytic activities due to its important electrical conductivity, affordability, easy accessibility, and high stability [10].
For example, in order to enhance the photocatalytic activity under both UV and Visible light, graphene oxide and carbon nanotubes TiO2 composites were prepared and investigated. These easy to make materials were tested, for the first time, in the photo degradation of RhB, as organic dye pollutant model, under UV and visible lights. Their photocatalytic behavior was evaluated and compared with the commercial TiO2 material, Degussa P25 [11]. New photocatalytic materials make use of visible or near infrared light which correspond roughly to 40–45% and 50–55% of the solar light intensity at sea level. For hydrogen photo production, the quantum efficiency values under a solar-type illumination can have a significant contribution from the UV part in several highly-active TiO2-based composite photocatalysts [12]. In spite of the fact that the UV part of the sunlight intensity is only ca. 4%. This would indicate that using the full range of wavelengths efficiently requires the development of visible or near infrared efficient materials but without dismissing UV activity. This short discussion points out that titanium dioxide can be the base of most active materials. In fact, a survey of literature summarized in review articles indicates that this is the general case in photocatalysis [13]. Pausova group [14] prepared an efficient active carbon (AC)–TiO2 composite, based on the mixing method, for photocatalytic degradation of azo-dye Acid Orange 7 (AO7). At the optimum ratio of AC/TiO2, the rate of photocatalytic oxidation AO7 increased twice as much as P25-TiO2 itself. Hoseini group [15] synthesized Co-TiO2 nano-catalyst containing different amounts of cobalt for photocatalytic degradation of 2,4-dichorophenol based on the sol-gel method. The photocatalytic degradation rate increased at low contents of cobalt in the Co-TiO2 catalyst. Karthikeyan et al. demonstrated that CO3O4/TiO2 doped with amine functionalized graphitic oxide promoted the photocatalytic degradation of CR and improved the photoinduced charge-separation [16]
Recently, we showed that in situ-prepared cobalt hierarchical graphitic carbon nanoparticles exhibited important catalytic activities in the hydrogenation of 2,4-dinitrophenol [17]. The aim of the present work is to examine the effect of addition of CoC@SiO2 to TiO2 on the photocatalytic degradation of CR dye. The objectives of this study are (i) preparation of the nanocomposites TiO2/CoC@SiO2-bipy (1) and TiO2/CoC@SiO2-phen (2) from TiO2 (75%) and CoC@SiO2-bipy (s1) or CoC@SiO2-phen (s2) (25%) (weight ratio); (ii) compare the photocatalytic activity of TiO2, (1) and (2) nanocomposites towards CR degradation under irradiation by sunlight; (iii) determine the effects of size diameter of (1) and (2) nanocomposites; (iii) assess the repeatability of the use of the immobilized (1) and (2) nanocomposites system for CR degradation, and (iv) relate the photocatalytic degradation of CR by (1) and (2) nanocomposites to the structure of CR supported by DFT studies.

2. Results and Discussion

2.1. Characterization of CoC@SiO2-Bipy (s1) and CoC@SiO2-Phen (s2)

The average particle size was calculated by counting at least 500 particles by using scanning electron microscope analysis. The average particle sizes of CoC@SiO2-bipy (s1) and CoC@SiO2-phen (s2) powder depend on the type of pyrolyzed starting complex. It is cobalt 2,2’-bipyridine chloride for (s1) and cobalt 1,10-phenanthroline chloride for (s2). The CoC@SiO2-bipy (s1) powder shows porous structure (Figure 1a). A significant increase in particle size occurred with CoC@SiO2-phen (s2) (Figure 1b). Particle size of the powder substantially increased to much larger grain size, approximately 200 nm. The Co nanoparticles were also supported on carbon-SiO2 monoliths shaped in CoC@SiO2-phen (s2). Cobalt nanoparticles agglomerate and form porous structure. These types of structures can have high specific surface areas [18].
Co2+ and Co3+ ions appeared in XPS analysis of CoC@SiO2-bipy (s1) prepared from pyrolysis of cobalt bipyridine chloride [19] and XPS of CoHGC@SiO2 prepared from cobalt phenanthroline sulfate complex [20] indicating that some cobalt atoms were oxidized to Co3O4. However, PXRD of these nanocomposites showed the presence of metallic cobalt (0). TEM of CoC@SiO2 show a graphitized structure around the cobalt nanoparticle embedded in the silica matrix. This was observed in CoC@SiO2-bipy (s1) [19] and CoHGC@SiO2 [20]. CoC@SiO2-bipy (s1) showed a moderate BET surface area of 98 m2/g [19], while CoC@SiO2-phen should have a higher surface area, since CoHGC@SiO2 showed 144.8 m2/g for BET surface area [20]. It should be noted that the addition of SiO2 in the preparation of (s1) and (s2) increased its surface area compared to CoC nanomaterials without silica matrix [19,20]. As the SEM indicated, the cobalt nanoparticles pointed out from the surface of the silica and were surrounded by tinny layer of graphitic carbon. This facilitated the photo degradation since cobalt species improve adsorption of Congo red into the surface and also the charge transfer. On the other hand, SEM-EDX revealed the elemental composition of CoC@SiO2 nanocomposites. CoC@SiO2-bipy (s1) [19] showed weight %C = 34.5, %N = 4.1, %O = 21.0, %Si = 30.8, and %Co = 10.5, while CoHGC@SiO2 nanocomposite prepared from pyrolysis of cobalt phenanthroline sulfate complex had similar weight % distribution, but with higher cobalt and lower silicon content: %C = 36.4, %O = 20.4, %Si = 11.4, and %Co = 25.8 [20].

2.2. Characterization of P25-TiO2, TiO2/CoC@SiO2-Bipy (1) and TiO2/CoC@SiO2-Phen (2) Nanocomposites

Figure 2 shows SEM images of pure TiO2 and TiO2/CoC@SiO2-bipy (1) and TiO2/CoC@SiO2-phen (2). The pure P25-TiO2 is spherical in shape with uniform dispersion (Figure 2a), with average size particle diameter of 103.70 ± 05 nm, while the particle sizes of TiO2/CoC@SiO2-phen (2) nanocomposite increased to become more than 200 nm. Thus TiO2/CoC@SiO2-bipy (1) with a smaller particle size can show improved adsorption of Congo red onto the catalyst (1) and increased photocatalytic efficiency. Larger size particles appeared after doping TiO2 with CoC@SiO2 (Figure 2a–c).

2.3. Effects of Parameters on the Degradation of CR

2.3.1. Effect of Amounts of Catalyst

To find the optimum catalyst dose that can be used, the effect of amount of catalysts was studied for P25-TiO2, TiO2/CoC@SiO2-bipy (1) or TiO2/CoC@SiO2-phen (2) in the range 11.25–56.25 mg L−1 with a fixed concentration of 10 µmol L−1 of CR solution. The mixture was ultrasonicated to give homogenous solutions. At low concentration ranges of CR, as its concentration increased the degradation increased. At this range photocatalysis was feasible, while below 11.25 mg/L of catalyst, there was minimum activity. The pH of TiO2, TiO2/CoC@SiO2-bipy (1) or TiO2/CoC@SiO2-phen (2) catalysts solution was around 4.0 [21]. After soaking in the dark for 30 min (Figure 3a), the absorbance of the solution was recorded for 60 min under sunlight (UV-index 5.0) [22]. CR was weakly adsorbed on the catalyst surface due to the large steric hindrance of large aromatic ensembles [23].
Figure 3a shows the absorbance spectra of photo degraded CR in the presence of TiO2/CoC@SiO2 catalysts under solar power (UV index 5.0). The absorption band of CR at 349 nm refers to π-π* transition (aromatic ring) and the band at 529 nm is related to the n–pi* transition (lone pair of N atom) in –N=N– azo moiety [24]. The degradation increased from 68% to 94.50% by increasing the amount of P25-TiO2 from 45.0 to 56.25 mg L−1, while the degradation increased from 93.10 to 95.80% by increasing the amounts of TiO2/CoC@SiO2-bipy (1) from 33.75 to 45.00 mg L−1. The degradation increased from 82.50% to 91.50% by increasing the amount of TiO2/CoC@SiO2-phen (2) catalyst from 11.25 to 22.50 mg L−1 (see Figure 3b). Figure 4 shows the vibrational band of azo group (-N=N-) of CR at (1430, 1665 cm−1); the vibration bands of SO3H observed at 1019 and 1110 cm−1, OH-stretching at 3286 and 3337 cm−1, C-H group stretching of the aromatic ring appeared at 2869 cm−1. The azo bands centered at 1430 to 1665 cm−1 disappeared after the photocatalysis process [25]. The C-H vibration band of phenyl groups also disappeared in FTIR spectrum after 60 min irradiation, thus no organic products remained. The expected degradation products are thus CO2, water, sulfate, and NOx. However, the degradation drastically increased by increasing the concentration of P25-TiO2 loading. For TiO2/CoC@SiO2-phen (2) catalyst, the % degradation of CR increased more at low loads and then decreased after adding more catalysts. This can be explained in terms of availability of active sites on the surface of catalysts TiO2/CoC@SiO2-bipy (1) and TiO2/CoC@SiO2-phen (2) and also on the penetration of photoactivating light into the mixture. The availability of active sites increased with increasing the suspension of catalyst nanoparticles, and the light penetration of the colloidal solution. At a higher concentration range, after 22.5 mg/L the catalyst (2) surface became saturated with degradation products. This hindered further interaction between the active catalyst sites and CR. The catalyst surface became unavailable for photon absorption and dye absorption, and this brought little stimulation to the catalytic reaction.
On the other hand, SEM image (Figure 2) shows that the size of the nanoparticles of the catalyst (1) were smaller than those of the catalyst (2) and thus (1) showed higher photocatalytic activity in the visible region, as expected [15]. The agglomeration increased more in TiO2/CoC@SiO2-phen (2) than in TiO2/CoC@SiO2-bipy (1), since (2) has larger particle size and thus the photocatalytic efficiency of (2) decreased.

2.3.2. Effect of Irradiation Time

The photo degradation reaction followed pseudo first order kinetics. The rate constant was determined from the slope of the plot of ln At/A0 versus irradiation time t (min) at pH 4, (Figure 5 and Table 1). The reaction rate decreased with irradiation time [26]. Additionally, a competition for degradation may have taken place between the reactant and the intermediate products.
The kinetics of the dyes’ degradation after a certain time decreased owing to:
  • The active sites of the catalyst being deactivated by the deposition of degradation products and the lifetime of the photocatalyst became short [27]. The intermediate and products of photodegradation that caused catalyst deactivation were aniline derivative and nitroso-compound resulting from cleavage of azo of CR as mentioned in [28] since these compounds can adsorb onto the catalyst surface.
  • The oxidation of N-atom of dye into N=O compounds became difficult [29].
  • The slow reaction was observed due to a small quantity of •OH radicals [30].

2.3.3. Effect of Acidic Medium

The absorption spectra of CR before (pH = 4.0) and after adding 10 µmol L−1 hydrochloric (HCl) (pH = 2.0) acid is shown in Figure 6. A bathochromic shift 80 nm was observed for the 529 nm band of CR due to the protonation of N-atoms of CR. A red shift of CR monomer bands was observed due to the partial self-association of CR monomers as anionic dimers [31]
Therefore, further experiments were carried out at the original (pH 5.0) of the P25-TiO2 and CR medium. In the acidic and neutral solutions, CR anions are easily adsorbed onto TiO2 particles with positive surface charge. These CR anions can be oxidized directly by oxygen under visible radiation. That is why high degradation ratios were achieved in acidic and neutral pH regions. However, at higher pH values, CR anions are generally excluded away from the negatively charged surface of (i) P25-TiO2 (ii) TiO2/CoC@SiO2-bipy (1) and (iii) TiO2/CoC@SiO2-phen (2). The % degradation efficiency and kinetics are shown in Figure 7a,b. Thus, degradation decreased at higher pH. A higher degradation rate at acidic pH has also been reported for Visible /TiO2 experiments due to the efficient electron transfer process due to strong surface complex bond formation. Titanium dioxide was recorded as having high oxidizing activity at lower pH [23].

2.4. Computational Details

All DFT calculations were performed using the Gaussian G09W program. Geometry optimizations were conducted using density functional theory (DFT) with Becker’s three parameter exchange functional [32], the Lee–Yang–Parr correlation functional (B3LYP), and the 6.31G(d) split-valence double zeta basis set were used [33]. After completing optimization, the theoretical properties of the compound studied such as electron affinity (EA), ionization energy (IE), hardness and softness, and the HOMO and LUMO were recorded.
The reactivity of organic compounds towards the catalytic process can be illustrated by DFT studies. DFT calculations propose the interactions between dyes and catalyst surfaces [34]. The theoretical reactivity parameters were calculated from CR-N=N. CR-N=NH and CR=N-NH species, which includes EHOMO (energy of the highest occupied molecular orbital) and ELUMO (energy of the lowest unoccupied molecular orbital), chemical hardness (η), softness(S), the electronic chemical potential (μ), and electrophilicity (ω) (see Table 2 and Figure 8.)
The theoretical reactivity parameters are represented in Equations (1)–(3) [35,36]:
η = 1 2 E H O M O E L U M O
μ = 1 2 E H O M O + E L U M O
ω = μ 2 2 η
In general, molecules are more reactive when chemical hardness is lower [36]. The ability of exchange electron density between chemical species depends on the electronic chemical potential (μ) [35] where strong electron–acceptor molecules have a higher electronic potential. Electrophilicity power ( ω ) reflects the ability of compounds to gain electron load, so CR-N=N gains more electrons from radicals HO and O2. The chemical hardness (η) of CR-N=N specie is the smallest, so CR-N=N is more available for catalysis (host–guest) interactions. The (Eg) energy gap [37] of CR-N=N is the smallest, the photocatalysis is accelerated due to recovery and recombination of hole–electron of the catalyst [38].
CR-N=NH anion present at low pH 2 with small electrophilicity (3.49) is more easily oxidized than CR-N=N and CR=N-NH of higher electrophilicity 5.36 and 4.11; by hydroxyl radical OH and oxo radical O produced under sun light irradiation (Figure 9). TiO2 leads to a higher percentage of degradation of CR-N=N. In addition, TiOH2+ formed at low pH can adsorb more anionic CR-N=NH and CR=N-NH species, thus enhances its catalysis [39]. TiO2 hole of HOMO orbital abstracts an electron from OH and forms an OH radical, and TiO2 donates an electron from its LUMO orbital to O2 species, which is responsible for the degradation of CR.
The photocatalytic performance of TiO2/CoC@SiO2-bipy (1) outperforms TiO2 at pH = 2 (for 10 mg of catalyst) and pH 4 (for 33 mg of catalyst); while (2) outperform TiO2 at pH = 4 (for 24 mg of catalyst) (Table 3, % degradation efficiency at fixed amount of catalyst).
The enhanced photocatalytic activity of TiO2/CoC@SiO2-bipy (1) and TiO2/CoC@SiO2-phen (2) composites is due to the charge transfer between Co(0) and Co3O4 species with TiO2. The graphene carbon matrix in the catalyst composites are electron carriers, which promotes additional photoinduced charge carrier separation (see Figure 9) [16]. A graphene carbon layer enhances the conductivity of electrons and can thus facilitate transport of an electron between active sites of the TiO2 catalyst and CR. Cobalt can enrich the conductive electrons TiO2. As silica has no effect on the exchange of electrons, it was used in the preparation in order to increase the surface area of cobalt nanoparticles species which formed a thin layer on top of the silica matrix [20].

3. Experimental Section

3.1. Materials and Reagents

Commercial P25 TiO2 powder (Evonik), CoCl2.6H2O, 2,2’-bipyridine and 1,10-phenanthroline and anthracene were purchased from sigma-Aldrich (Darmstadt, Germany). The stock solution of 1 mmol L−1 Congo Red (C32H24N6O6S2) was prepared by dissolving an accurate weight of Congo Red powder (sigma Aldrich) in distilled water. The absorbance of photocatalysis of Congo red was measured by UV-VIS spectrophotometer Model shimadzu (Tokyo, Japan) with UV probe software (USA). Fourier-transform infrared spectroscopy (FTIR) was measured using Alpha Bruker (Billerica, Massachusetts, USA) with OPUS software (USA). The images of scanning electron microscopy (SEM) were obtained from (FE-SEM, JEOL JSM-76700F), (JEOL, Tokyo, Japan). All the samples were analyzed with acceleration voltage of 15 kV, working distance of 8 mm, and sample vacuum of 1 × 10−5 Pa. The photo degradation studies were performed outdoor using Sunlight by recording the UV-index measurements according to a weather site reference and a solar-meter device [40].
Synthesis of cobalt–carbon silica nanocomposite CoC@SiO2-bipy (s1) and CoC@SiO2-phen (s2) were prepared similarly by mixing ethanolic solution of Co(2,2’-bipy)Cl2 (1.5 g) [19] or Co(1,10-phenanthroline)Cl2 (1.5 g), respectively, anthracene (1.5 g) and silica (1.5 g). Then, the mixture was heated under increased and constant temperature for different time intervals up to 850 °C under a vacuum/nitrogen atmosphere. After a slow cooling to room temperature, a black powder of (s1) or (s2) was obtained. The yield percentages of products by weight were 23.3% for CoC@SiO2-bipy (s1) and 40.7% for CoC@SiO2–phen (s2).

3.2. Synthesis of Titanium Oxide/Cobalt Silica Supporting Nanocomposites TiO2/CoC@SiO2-Bipy (1) and –Phen (2)

The nanocomposites TiO2/CoC@SiO2-bipy (1) and TiO2/CoC@SiO2-phen (2) were prepared by ultrasonication of TiO2 (75%) and CoC@SiO2-bipy (s1) or CoC@SiO2-phen (s2) (25%) (weight ratio) for 30 min. The mixture was then filtered, dried at 60 °C, and milled according to the methodology of Pausova and et al [14].

3.3. Photo Degradation of Congo Red

The amounts of catalysts in the photocatalytic degradation experiments were tested in the range 11.25–56.25 mg L−1 TiO2/CoC@SiO2-bipy (1) or TiO2/CoC@SiO2-phen (2) catalysts. The concentration of the CR solution was 10 µmol L−1. Before irradiation in the sunlight, mixtures of P25-TiO2, TiO2/CoC@SiO2-bipy (1), or TiO2/CoC@SiO2-phen (2) and CR were soaked in the dark for 30 min at room temperature to maintain adsorption of CR onto catalysts and ensure the adsorption/desorption equilibrium of CR aqueous solution with the catalyst [38]. Then, the sample absorbances were measured before light irradiation. The samples were immediately placed in the outdoor sunlight (UV-index = 5.0) for a constant time interval until complete degradation. Color removal of the samples was followed by measuring the absorbance peak of CR centered at 529 nm. In order to determine the optimum working condition for CR degradation, the effect of catalyst amounts of (P25-TiO2, TiO2/CoC@SiO2-bipy (1), or TiO2/CoC@SiO2-phen (2)), irradiation time and pH of the medium were tested. The degradation efficiency experiments were performed using a solution from batch studies.
% Degradation was calculated using the Equation (4),
%   D e g r a d a t i o n   efficiency   = A 0 A t A t   ×   100 = C 0 C t C t   ×   100
where A0 and At were the absorbance intensity of CR at initially and after time(t) of irradiation. C0 and Ct were the concentration of CR at initial and after time (t) of irradiation.
The relation of ln (A0/At) with irradiation time (t) was used to predict degradation kinetics [40], Equation (5)
l n A 0 A t = l n C 0 C t = k t

4. Conclusions

The nanocomposites TiO2/CoC@SiO2-bipy (1) and TiO2/CoC@SiO2-phen (2) were prepared from TiO2 (75%) and CoC@SiO2-bipy (s1) or CoC@SiO2-phen (s2) (25%) (weight ratio), ultrasonication of a mixture of TiO2 and CoC@SiO2-bipy (s1) or TiO2/CoC@SiO2-phen (s2), with the percent ratio (75:25), respectively. The nanocomposites were used for photodegradation of Congo red under sunlight irradiation (UV-index 5.0). The highest catalytic oxidation was recorded at pH 2 since TiO2, TiO2/CoC@SiO2-bipy (1) and TiO2/CoC@SiO2-phen (2), at low pH, were able to adsorb more anionic CR-N=NH and CR=N-NH species. The kinetic rate constants k (10−3 min−1) of Congo red oxidation were 55.0 for TiO2, 22.0 for TiO2/CoC@SiO2-bipy (1), and 51.0 for TiO2/CoC@SiO2-phen (2) at pH 2. The charge transfer between Co(0) or Co3O4 species and TiO2 increased the photodegradation of CR. Graphene carbon matrix in the catalyst composites, being electrons carriers, promoted additional photoinduced charge carrier separation. In DFT calculations, the CR-N=NH anion present at low pH 2 with small electrophilicity (3.49) was more easily oxidized than CR-N=N and CR=N-NH with higher electrophilicity of 5.36 and 4.11, respectively; by hydroxyl radical OH and oxo radical O produced under sunlight irradiation. The newly prepared TiO2/CoC@SiO2-bipy (1) nanocomposites showed increased photocatalytic efficiency at low catalyst doses and faster rates of degradation of Congo red compared to TiO2/CoC@SiO2-phen (2) and TiO2. The novel catalysts (1) and (2) were easily separated, by centrifugation and filtration, from the reaction mixture compared to TiO2. These results induced an efficient TiO2/CoC@SiO2 nano catalysts for the removal of hazardous materials.

Author Contributions

E.M.B. wrote the manuscript; H.H.H. assisted in writing the manuscript; E.M.B. and R.K.K. performed catalysis experiments: E.M.B. and H.H.H. performed nanomaterial analysis and characterizations; H.T. and R.K.K. assisted in discussing. H.H.H. supervised the work. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Deanship of Scientific Research at King Faisal University; grant number 1811029 And the APC was also funded by 1811029 Deanship of Scientific Research at King Faisal University.

Data Availability Statement

Not applicable.

Acknowledgments

The authors acknowledge the Deanship of Scientific Research at King Faisal University, Saudi Arabia, for financial support under the Research Group (Grant No. 1811029).

Conflicts of Interest

There is no conflict to declare.

References

  1. Carmen, Z.; Daniela, S. Textile organic dyes–characteristics, polluting effects and separation/elimination procedures from industrial effluents–a critical overview. In Organic Pollutants Ten Years After the Stockholm Convention; IntechOpen: London, UK, 2012; Volume 3. [Google Scholar] [CrossRef] [Green Version]
  2. Shaban, M.; Abukhadra, M.R.; Shahien, M.; Ibrahim, S.S. Novel bentonite/zeolite-NaP composite efficiently removes methylene blue and Congo red dyes. Environ. Chem. Lett. 2018, 16, 275–280. [Google Scholar] [CrossRef]
  3. Varga, M.; Kopecký, D.; Kopecká, J.; Křivka, I.; Hanuš, J.; Zhigunov, A.; Trchová, M.; Vrňata, M.; Prokeš, J. The ageing of polypyrrole nanotubes synthesized with methyl orange. Eur. Polym. J. 2017, 96, 176–189. [Google Scholar] [CrossRef]
  4. Nasabi, M.; Labbafi, M.; Khanmohammadi, M. Optimizing nano TiO2 assisted decoloration process for industrial date syrup utilizing response surface methodology. J. Food Process Eng. 2017, 40, e12537. [Google Scholar] [CrossRef]
  5. Kim, S.; Choi, W. Visible-light-induced photocatalytic degradation of 4-chlorophenol and phenolic compounds in aqueous suspension of pure titania: Demonstrating the existence of a surface-complex-mediated path. J. Phys. Chem. B 2005, 109, 5143–5149. [Google Scholar] [CrossRef]
  6. Ma, J.; Wang, C.; He, H. Enhanced photocatalytic oxidation of NO over g-C3N4-TiO2 under UV and visible light. Appl. Catal. B: Environ. 2016, 184, 28–34. [Google Scholar] [CrossRef] [Green Version]
  7. Iwase, A.; Ng, Y.H.; Ishiguro, Y.; Kudo, A.; Amal, R. Reduced Graphene Oxide as a Solid-State Electron Mediator in Z-Scheme Photocatalytic Water Splitting under Visible Light. J. Am. Chem. Soc. 2011, 133, 11054–11057. [Google Scholar] [CrossRef]
  8. Lee, J.; Jackson, D.H.; Li, T.; Winans, R.E.; Dumesic, J.A.; Kuech, T.F.; Huber, G.W. Enhanced stability of cobalt catalysts by atomic layer deposition for aqueous-phase reactions. Energy Environ. Sci. 2014, 7, 1657–1660. [Google Scholar] [CrossRef]
  9. Li, S.-M.; Dong, Y.-Y.; Ma, M.-G.; Fu, L.-H.; Sun, R.-C.; Xu, F. Hydrothermal synthesis, characterization, and bactericidal activities of hybrid from cellulose and TiO2. Carbohydr. Polym. 2013, 96, 15–20. [Google Scholar] [CrossRef]
  10. Serp, P.; Machado, B. Nanostructured Carbon Materials for Catalysis; Royal Society of Chemistry: London, UK, 2015. [Google Scholar] [CrossRef]
  11. Silva, M.R.; Lourenço, M.A.; Tobaldi, D.M.; da Silva, C.F.; Seabra, M.P.; Ferreira, P. Carbon-modified titanium oxide materials for photocatalytic water and air decontamination. Chem. Eng. J. 2020, 387, 124099. [Google Scholar] [CrossRef]
  12. Caudillo-Flores, U.; Kubacka, A.; Berestok, T.; Zhang, T.; Llorca, J.; Arbiol, J.; Cabot, A.; Fernández-García, M. Hydrogen photogeneration using ternary CuGaS2-TiO2-Pt nanocomposites. Int. J. Hydrog. Energy 2020, 45, 1510–1520. [Google Scholar] [CrossRef]
  13. Barba-Nieto, I.; Caudillo-Flores, U.; Fernández-García, M.; Kubacka, A. Sunlight-Operated TiO2-Based Photocatalysts. Molecules 2020, 25, 4008. [Google Scholar] [CrossRef] [PubMed]
  14. Šárka, P.; Riva, M.; Baudys, M.; Krýsa, J.; Barbieriková, Z.; Brezová, V. Composite materials based on active carbon/TiO2 for photocatalytic water purification. Catalysis Today 2019, 328, 178–182. [Google Scholar] [CrossRef]
  15. Hoseini, S.N.; Pirzaman, A.K.; Aroon, M.A.; Pirbazari, A.E. Photocatalytic degradation of 2, 4-dichlorophenol by Co-doped TiO2 (Co/TiO2) nanoparticles and Co/TiO2 containing mixed matrix membranes. J. Water Process Eng. 2017, 17, 124–134. [Google Scholar] [CrossRef]
  16. Jo, W.-K.; Kumar, S.; Isaacs, M.A.; Lee, A.F.; Karthikeyan, S. Cobalt promoted TiO2/GO for the photocatalytic degradation of oxytetracycline and Congo Red. Appl. Catal. B: Environ. 2017, 201, 159–168. [Google Scholar] [CrossRef] [Green Version]
  17. 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]
  18. Urkasame, K.; Yoshida, S.; Takanohashi, T.; Iwamura, S.; Ogino, I.; Mukai, S.R. Development of TiO2–SiO2 Photocatalysts Having a Microhoneycomb Structure by the Ice Templating Method. ACS Omega 2018, 3, 14274–14279. [Google Scholar] [CrossRef]
  19. Alotaibi, N.; Hammud, H.H.; Al Otaibi, N.; Hussain, S.G.; Prakasam, T. Novel cobalt–carbon@ silica adsorbent. Sci. Rep. 2020, 10, 1–17. [Google Scholar] [CrossRef]
  20. Hammud, H.H.; Karnati, R.K.; Alotaibi, N.; Hussain, S.G.; Prakasam, T. Cobalt–Carbon Nanoparticles with Silica Support for Uptake of Cationic and Anionic Dyes from Polluted Water. Molecules 2021, 26, 7489. [Google Scholar] [CrossRef]
  21. Ferretto, L.; Glisenti, A. Surface acidity and basicity of a rutile powder. Chem. Mater. 2003, 15, 1181–1188. [Google Scholar] [CrossRef]
  22. Borges, M.E.; Sierra, M.; Cuevas, E.; García, R.D.; Esparza, P. Photocatalysis with solar energy: Sunlight-responsive photocatalyst based on TiO2 loaded on a natural material for wastewater treatment. Solar Energy. 2016, 135, 527–535. [Google Scholar] [CrossRef]
  23. Lachheb, H.; Puzenat, E.; Houas, A.; Ksibi, M.; Elaloui, E.; Guillard, C.; Herrmann, J.-M. Photocatalytic degradation of various types of dyes (Alizarin S, Crocein Orange G, Methyl Red, Congo Red, Methylene Blue) in water by UV-irradiated titania. Appl. Catal. B Environ. 2002, 39, 75–90. [Google Scholar] [CrossRef]
  24. Costa, A.L.; Gomes, A.C.; Pillinger, M.; Gonçalves, I.S.; Pina, J.; Seixas de Melo, J.S. Insights into the Photophysics and Supramolecular Organization of Congo Red in Solution and the Solid State. ChemPhysChem 2017, 18, 564–575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Bartošová, A.; Blinová, L.; Sirotiak, M.; Michalíková, A. Usage of FTIR-ATR as non-destructive analysis of selected toxic dyes. Ved. Práce Mater. Available online: https://doi.org/10.1515/rput-2017-0012 (accessed on 29 December 2021). [CrossRef] [Green Version]
  26. Yang, J.; Luo, X. Ag-doped TiO2 immobilized cellulose-derived carbon beads: One-Pot preparation, photocatalytic degradation performance and mechanism of ceftriaxone sodium. Appl. Surf. Sci. 2021, 542, 148724. [Google Scholar] [CrossRef]
  27. Worstell, J. Chapter 3—Catalyst Deactivation. In Adiabatic Fixed-Bed Reactors; Worstell, J., Ed.; Butterworth-Heinemann: Boston, MA, USA, 2014; pp. 35–65. [Google Scholar]
  28. Argote-Fuentes, S.; Feria-Reyes, R.; Ramos-Ramírez, E.; Gutiérrez-Ortega, N.; Cruz-Jiménez, G. Photoelectrocatalytic Degradation of Congo Red Dye with Activated Hydrotalcites and Copper Anode. Catalysts 2021, 11, 211. [Google Scholar] [CrossRef]
  29. Rauf, M.A.; Meetani, M.A.; Khaleel, A.; Ahmed, A. Photocatalytic degradation of methylene blue using a mixed catalyst and product analysis by LC/MS. Chem. Eng. J. 2010, 157, 373–378. [Google Scholar] [CrossRef]
  30. Nosaka, Y.; Nosaka, A. Understanding hydroxyl radical (•OH) generation processes in photocatalysis. ACS Energy Lett. 2016, 1, 356–359. [Google Scholar] [CrossRef] [Green Version]
  31. Yaneva, Z.L.; Georgieva, N.V. Insights into Congo Red Adsorption on Agro-Industrial Materials-Spectral, Equilibrium, Kinetic, Thermodynamic, Dynamic and Desorption Studies. A Review. Int. Rev. Chem. Eng. 2012, 4, 127–146. [Google Scholar]
  32. Becke, A.D. A new mixing of Hartree–Fock and local density-functional theories. J. Chem. Phys. 1993, 98, 1372–1377. [Google Scholar] [CrossRef]
  33. Koch, W.; Holthausen, M.C. A Chemist’s Guide to Density Functional Theory; John Wiley & Sons: Hoboken, NJ, USA, 2015. [Google Scholar] [CrossRef]
  34. Domingo, L.R.; Ríos-Gutiérrez, M.; Pérez, P. Applications of the conceptual density functional theory indices to organic chemistry reactivity. Molecules 2016, 21, 748. [Google Scholar] [CrossRef] [Green Version]
  35. Parr, R.G.; Yang, W. Density functional approach to the frontier-electron theory of chemical reactivity. J. Am. Chem. Soc. 1984, 106, 4049–4050. [Google Scholar] [CrossRef]
  36. Parr, R.G.; Szentpály, L.v.; Liu, S. Electrophilicity index. J. Am. Chem. Soc. 1999, 121, 1922–1924. [Google Scholar] [CrossRef]
  37. Bredas, J.-L. Mind the gap! Mater. Horiz. 2014, 1, 17–19. [Google Scholar] [CrossRef]
  38. Valadi, F.M.; Ekramipooya, A.; Gholami, M.R. Selective separation of Congo Red from a mixture of anionic and cationic dyes using magnetic-MOF: Experimental and DFT study. J. Mol. Liq. 2020, 318, 114051. [Google Scholar] [CrossRef]
  39. Thomas, M.; Naikoo, G.A.; Sheikh, M.U.D.; Bano, M.; Khan, F. Effective photocatalytic degradation of Congo red dye using alginate/carboxymethyl cellulose/TiO2 nanocomposite hydrogel under direct sunlight irradiation. J. Photochem. Photobiol. A: Chem. 2016, 327, 33–43. [Google Scholar] [CrossRef]
  40. Soares, L.G.; de Oliveira Vaz, M.; Teixeira, S.R.; Alves, A.K. Absorbance determination and photocatalytic production of hydrogen using tungsten and TiO2 oxide nanostructures As catalyst. Clean. Eng. Technol. 2021, 5, 100268. [Google Scholar] [CrossRef]
Figure 1. SEM micrographs of the prepared (a) CoC@SiO2-bipy (s1) and (b) CoC@SiO2-phen (s2) which shows monoliths shape of SiO2-carbon decorated with Co nanoparticles.
Figure 1. SEM micrographs of the prepared (a) CoC@SiO2-bipy (s1) and (b) CoC@SiO2-phen (s2) which shows monoliths shape of SiO2-carbon decorated with Co nanoparticles.
Catalysts 12 00248 g001
Figure 2. SEM micrographs of (a) P25-TiO2 and the prepared (b) TiO2/CoC@SiO2-bipy (1) and (c) TiO2/CoC@SiO2-phen (2).
Figure 2. SEM micrographs of (a) P25-TiO2 and the prepared (b) TiO2/CoC@SiO2-bipy (1) and (c) TiO2/CoC@SiO2-phen (2).
Catalysts 12 00248 g002
Figure 3. (a) The absorption spectra of photo degradation of 10 µM CR after being mixed with TiO2/CoC@SiO2 (1) (a) 0 min and (b) 60 min under sunlight (UV-index 5.0). (b) Bar diagram of degradation efficiency of CR under solar irradiation, initial concentration of CR = 10µM; irradiation time = 60 min; volume = 2 mL; pH of the medium = 4.0.
Figure 3. (a) The absorption spectra of photo degradation of 10 µM CR after being mixed with TiO2/CoC@SiO2 (1) (a) 0 min and (b) 60 min under sunlight (UV-index 5.0). (b) Bar diagram of degradation efficiency of CR under solar irradiation, initial concentration of CR = 10µM; irradiation time = 60 min; volume = 2 mL; pH of the medium = 4.0.
Catalysts 12 00248 g003
Figure 4. The FTIR spectra of 10µmol L−1 of CR (a); 10µmol L−1 CR + 45.25 mg L−1 TiO2/CoC@SiO2-bipy (1) (b); and 10µmol L−1 CR + 45.25 mg L−1 TiO2/CoC@SiO2-bipy (1) (c) after irradiation for 60 min with UV-index 5.0.
Figure 4. The FTIR spectra of 10µmol L−1 of CR (a); 10µmol L−1 CR + 45.25 mg L−1 TiO2/CoC@SiO2-bipy (1) (b); and 10µmol L−1 CR + 45.25 mg L−1 TiO2/CoC@SiO2-bipy (1) (c) after irradiation for 60 min with UV-index 5.0.
Catalysts 12 00248 g004
Figure 5. The plotting of ln(A/A0) vs. time for degradation kinetics of CR. Initial concentration of CR = 10 µmol L−1; amount of P25-TiO2(■), TiO2/CoC@SiO2-bipy (1) (►) and TiO2/CoC@SiO2-phen (2) (●) were 56.25, 45.0 and 22.5 mg L−1, respectively; Volume = 2 mL; pH of the medium = 4.0.
Figure 5. The plotting of ln(A/A0) vs. time for degradation kinetics of CR. Initial concentration of CR = 10 µmol L−1; amount of P25-TiO2(■), TiO2/CoC@SiO2-bipy (1) (►) and TiO2/CoC@SiO2-phen (2) (●) were 56.25, 45.0 and 22.5 mg L−1, respectively; Volume = 2 mL; pH of the medium = 4.0.
Catalysts 12 00248 g005
Figure 6. The absorbance spectra of (a) 10 µmol L−1 CR before and after addition of (b) 10 µmol L−1 HCl.
Figure 6. The absorbance spectra of (a) 10 µmol L−1 CR before and after addition of (b) 10 µmol L−1 HCl.
Catalysts 12 00248 g006
Figure 7. (a) Effect of amount; (a) Bar diagram of degradation efficiency of CR under solar irradiation, initial concentration of CR = 10µM; irradiation time = 60 min; Volume = 2 mL; pH of the medium = 2.0. Figure 7b. The plotting ln(A/A0) vs. time for degradation kinetics of CR. initial concentration of CR = 10 µmol L−1; amount of (i) P25-TiO2(■), (ii) TiO2/CoC@SiO2-bipy (1) (►) and (iii) TiO2/CoC@SiO2-phen (2) (●) were 56.25, 45 and 22.5 mg L−1, respectively; Volume = 2 mL; pH of the medium = 2.0.
Figure 7. (a) Effect of amount; (a) Bar diagram of degradation efficiency of CR under solar irradiation, initial concentration of CR = 10µM; irradiation time = 60 min; Volume = 2 mL; pH of the medium = 2.0. Figure 7b. The plotting ln(A/A0) vs. time for degradation kinetics of CR. initial concentration of CR = 10 µmol L−1; amount of (i) P25-TiO2(■), (ii) TiO2/CoC@SiO2-bipy (1) (►) and (iii) TiO2/CoC@SiO2-phen (2) (●) were 56.25, 45 and 22.5 mg L−1, respectively; Volume = 2 mL; pH of the medium = 2.0.
Catalysts 12 00248 g007
Figure 8. (a) Electron density map; (b) Distribution of HOMO and LUMO Orbitals of the cationic and anionic CR dye.
Figure 8. (a) Electron density map; (b) Distribution of HOMO and LUMO Orbitals of the cationic and anionic CR dye.
Catalysts 12 00248 g008aCatalysts 12 00248 g008b
Figure 9. The proposed mechanism of photocatalysis CR by TiO2/CoC@SiO2-phen (2) nanocomposite.
Figure 9. The proposed mechanism of photocatalysis CR by TiO2/CoC@SiO2-phen (2) nanocomposite.
Catalysts 12 00248 g009
Table 1. Pseudo first order rate constant and catalyst amount for maximum % degradation of CR at pH 2 and pH 4.
Table 1. Pseudo first order rate constant and catalyst amount for maximum % degradation of CR at pH 2 and pH 4.
pH = 4pH = 2
CatalystsAmount of Catalyst (mg)k × 103 (min−1)% Degradationk × 103 (min−1)% Degradation
TiO256.2535.094.5%55.096.4%
(1)45.0031.095.8%22.086.8%
(2)22.5026.091.5%51.054.9%
Table 2. Represented HOMO and LUMO energies, hardness (η) and softness (S) of three forms of Congo red.
Table 2. Represented HOMO and LUMO energies, hardness (η) and softness (S) of three forms of Congo red.
MoleculesE(eV)
HOMO
E(eV)
LUMO
Eg
(eV)
𝜂
(Hardness)
S
(Softness)
µ
Chemical Potential
ω
Electrophilicity
CR-N=N−6.1−5.7−0.43.250.31−5.95.36
CR-N=NH−5.4−4.2−1.23.300.30−4.83.49
CR=N-NH−8.9−5.4−3.56.220.16−7.154.11
Table 3. % Degradation of CR at pH 2 and pH 4 at fixed amount of catalyst.
Table 3. % Degradation of CR at pH 2 and pH 4 at fixed amount of catalyst.
Catalyst% Degradation pH = 4% Degradation pH = 2
Amount24 mg33 mg10 mg33 mg
Formula
TiO275%65%63%91%
TiO2/CoC@SiO2-bipy (1)53%87%76%61%
TiO2/CoC@SiO2-phen (2)92%43%63%78%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hammud, H.H.; Traboulsi, H.; Karnati, R.K.; Bakir, E.M. Photodegradation of Congo Red by Modified P25-Titanium Dioxide with Cobalt-Carbon Supported on SiO2 Matrix, DFT Studies of Chemical Reactivity. Catalysts 2022, 12, 248. https://doi.org/10.3390/catal12030248

AMA Style

Hammud HH, Traboulsi H, Karnati RK, Bakir EM. Photodegradation of Congo Red by Modified P25-Titanium Dioxide with Cobalt-Carbon Supported on SiO2 Matrix, DFT Studies of Chemical Reactivity. Catalysts. 2022; 12(3):248. https://doi.org/10.3390/catal12030248

Chicago/Turabian Style

Hammud, Hassan H., Hassan Traboulsi, Ranjith Kumar Karnati, and Esam M. Bakir. 2022. "Photodegradation of Congo Red by Modified P25-Titanium Dioxide with Cobalt-Carbon Supported on SiO2 Matrix, DFT Studies of Chemical Reactivity" Catalysts 12, no. 3: 248. https://doi.org/10.3390/catal12030248

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