Next Article in Journal
Photo-Oxidation of Ammonia to Molecular Nitrogen in Water under UV, Vis and Sunlight Irradiation
Next Article in Special Issue
A Novel Machine Learning Model to Predict the Photo-Degradation Performance of Different Photocatalysts on a Variety of Water Contaminants
Previous Article in Journal
Chiral Synthesis of 3-Amino-1-phenylbutane by a Multi-Enzymatic Cascade System
Previous Article in Special Issue
Fabrication of Six Manganese Containing Polyoxometalate Modified Graphite C3N4 Nanosheets Catalysts Used to Catalyze Water Decomposition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Degradation Efficiency and Kinetics Analysis of an Advanced Oxidation Process Utilizing Ozone, Hydrogen Peroxide and Persulfate to Degrade the Dye Rhodamine B

Department of Water Protection, Central Mining Institute, Plac Gwarków 1, 40-166 Katowice, Poland
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(8), 974; https://doi.org/10.3390/catal11080974
Submission received: 17 July 2021 / Revised: 13 August 2021 / Accepted: 13 August 2021 / Published: 14 August 2021
(This article belongs to the Special Issue 10th Anniversary of Catalysts—Feature Papers in Photocatalysis)

Abstract

:
In this study, the effectiveness of a rhodamine B (RhB) dye degradation process at a concentration of 20 mg/L in different advanced oxidation processes—H2O2/UV, O3/UV and PDS/UV—has been studied. The use of UV in a photo-assisted ozonation process (O3/UV) proved to be the most effective method of RhB decolorization (90% after 30 min at dye concentration of 100 mg/L). The addition of sulfate radical precursors (sodium persulfate, PDS) to the reaction environment did not give satisfactory effects (17% after 30 min), compared to the PDS/UV system (70% after 30 min). No rhodamine B decolorization was observed using hydrogen peroxide as a sole reagent, whereas an effect on the degree of RhB degradation was observed when UV rays strike the sample with H2O2 (33% after 30 min). The rhodamine B degradation process followed the pseudo-first-order kinetics model. The combined PDS/O3/UV process has shown 60% color removal after 30 min of reaction time at an initial dye concentration of 100 mg/L. A similar effectiveness was obtained by only applying ozone or UV-activated persulfate, but at a concentration 2–5 times lower (20 mg/L). The results indicated that the combined PDS/O3/UV process is a promising method for high RhB concentrations (50–100 mg/L) comparing to other alternative advanced oxidation processes.

1. Introduction

Synthetic dyes are used in different industrial sectors, in e.g., the textile, cosmetic, pharmaceutical, plastics, tanning, paper and food industries. Azo, anthraquinone, phthalocyanine and triphenylmethane dyes are some of the most commonly utilized dye groups in the textile industry [1], and the progression of the industrial revolution has led to an increase in their use. Moreover, as much as 50% of dyes are lost to the process wastewater as dyestuff residues during the dyeing and finishing process [2], causing a significant environmental pollution problem.
One of the dyes utilized in the chemical, textile, paper and paint industries is rhodamine B (RhB). The physicochemical properties of RhB are summarized in Table 1. Rhodamine B is part of the rhodamine dyes—organic chemical compounds of the triphenylmethane xanthene dye group—classified as micropollutants. These compounds exhibit toxicity after short-term exposure (acute toxicity), but the main research focus is concentrated on analyzing the long-term health effects resulting from their presence in the environment (chronic toxicity) [3]. RhB is irritating to the skin, eyes and upper respiratory tract. It also exhibits potential mutagenic and carcinogenic effects [4].
Due to high concentrations of dyes in process wastewater and their low biodegradability, it seems reasonable to develop new solutions of water and wastewater treatment, particularly from the perspective of removing colored pollutants that may pose significant environmental problems.
Advanced oxidation processes (AOPs) are proposed as an alternative solution to traditional water and wastewater treatment methods. AOPs utilizing physicochemical processes enable the total or partial elimination of hazardous substances through their mineralization into environmentally neutral carbon dioxide and water, as well as other simple chemical compounds [6]. AOPs generate highly reactive hydroxyl radicals OH with a reduction potential of E0 = 2.80 V. Currently, there are a number of AOPs used for degradation of dyes such as rhodamine B, e.g., O3/UV, TiO2/UV (photocatalysis) or H2O2/Fe2+ (Fenton’s reagent). In studies on the degradation of the rhodamine B dye from aqueous solutions using photo-Fenton and heterogeneous photocatalytic processes with an innovative Ni-Cu@MWCNT photocatalyst, Tariq et al. [7] achieved RhB dye degradation at levels of 86% and 98%, respectively. The high degradation percentage was achieved through the generation of large amounts of hydroxyl radicals under the influence of UV radiation. Similar results were also obtained by Dhahir et al. in their research [8].
Recent literature data [9,10,11,12,13] indicates that current research is focused on a new class of oxidizers—peroxyacids, including inorganic peroxyacid salts (peroxymonosulfate, PMS; and persulfate, PS). The primary target of using peroxyacids is to generate sulfate radicals. The sulfate radical ( SO 4 ) is characterized by an oxidizing potential E0 = 2.50–3.10 V and is produced as a result of the interaction of various energy sources (e.g., heat, UV radiation) with the SO 4 sulfate radical precursor (e.g., sodium persulfate Na2S2O8) or as a result of the reactions of transition metal ions (e.g., Fe2+, Co2+) with the SO 4 radical precursor. In sulfate radicals-based processes, it is possible to achieve organic pollutant degradation under the influence of both sulfate and hydroxyl radicals. Activation of PS or PMS results in the generation of sulfate radicals, or both sulfate and hydroxyl radicals, for PS and PMS respectively [14,15,16].
Zawadzki [17] investigated the degradation of RhB in the presence of ultrasound (US) and sodium persulfate (Na2S2O8) activated by visible light (Vis). The performed study proved that a hybrid sono-visible light-sodium persulfate system is an interesting alternative for treatment of solutions containing rhodamine B at varying dye concentrations (up to 20 mg/L). Heidarpour et al. [18] improved the RhB decolorization process by increasing the oxidative capacity of the reaction system through simultaneous photocatalysis and heterogeneous persulfate activation, thereby confirming the significant role of sulfate radicals in the RhB degradation process.
Considering the broad application of rhodamine B in various branches of industry, and therefore the presence of this substance in process wastewater and the problems related to its removal, the aim of this study was to assess the effectiveness of rhodamine B degradation in an advanced oxidation process: H2O2/UV, O3/UV, PDS/UV. The concentration of dyes in textile effluents depends on several parameters such as the type of dyeing process, dyes used, the nature of the effluent. Dye concentrations in industrial wastewater have been determined at varying levels, usually from about 10 to 200 mg/L, however concentrations between 600 and 7000 mg/L was also reported [19]. Therefore, the combined PDS/O3/UV process was investigated to determine the ability to remove RhB over a wide concentration range (from 20 to 100 mg/L). The results of the combined PDS/O3/UV process were compared with other literature data in order to show the advantage of the proposed method.

2. Results and Discussion

2.1. Advanced Oxidation Study

The rhodamine B degradation using the following processes: H2O2, H2O2/UV, O3, O3/UV as well as PDS and PDS/UV was investigated. The results are presented in Figure 1. It was observed that the degree of rhodamine B degradation increased simultaneously over time, independently of the treatment system. For example, the H2O2/UV system degrade the RhB at a level of 14% and 33%, after 15 and 30 min respectively. On the other hand, the application of sodium persulfate (PDS) enabled to degrade the RhB at a level of 6% and 12% after 10 and 20 min.
Addition of oxidizers (O3, H2O2, PDS) into the reaction environment results in the initiation of reactions generating oxidative radicals (hydroxyl OH or sulfate SO 4 ). There are two methods of pollutant elimination in the case of the sole ozonation process [20]. The first, the direct oxidation route, consists in a reaction of the ozone with the pollutant. Direct oxidation is more selective and the most appropriate for acidic conditions. In neutral reaction environment (pH = 6.0–7.0), the available methods include both direct as well as indirect oxidation, which consists in oxidation by means of hydroxyl radicals. For this reason, the reaction environment pH after the addition of the rhodamine B standard was maintained at a level of about 6.0. Due to the certain limitations characteristic of ozonation in water and wastewater technologies (e.g., low ozone transfer rate, ozone production costs) [21], one of the methods used to assist the ozonation reaction is the application of UV radiation. Ozone has a molar absorptivity (ε) of 2952 cm/M at λ = 254 nm [22]. For this reason, UV radiation is used in order to accelerate the decomposition of ozone and hydroxyl radical generation, as shown in Equations (1) and (2). This effect was documented by authors [23,24]. In this study, the use of UV irradiation-assisted ozonation made it possible to increase the degree of rhodamine B decolorization from 75% to 90% after 30 min of reaction time:
O 3 hv   ( λ   <   310   nm ) O 2 + O
O + H 2 O   2 OH  
The next studied processes were PDS and PDS/UV systems involving sulfate radical generation. Literature data indicates that sulfate radicals find application not only in dye decolorization [25], but also in the removal of a number of other organic substances, e.g., pharmaceutical substances [26] and pesticides [27] as well as liquid media, such as landfill leachate [28].
The effects of sole sulfate radical precursors (sodium persulfate, PDS) did not bring about satisfactory results of rhodamine B decolorization effects, as can be observed in Figure 1. Only 17% of dye degradation was achieved after 30 min of reaction time. Nevertheless, it should be noted that the effectiveness of sole PDS was only 5% lower compared to the H2O2 process catalyzed by UV radiation (H2O2/UV). As indicated in the literature [29,30], persulfates are capable of reacting via direct reactions with organic compounds without the presence of catalysts, but the degree of pollutants elimination is lower compared to the one obtained through the use of activated persulfate [31]. The reason for this is that the reaction in the presence of radicals generated as a result of persulfate activation is faster by a factor of 103–105 compared to the sole presence of the persulfate without activation [32].
The energy transferred to the persulfate anion results in the breaking of the peroxide bond and the generation of two sulfate radicals (Equation (3)). Sodium persulfate activation by UV irradiation results in reactions according to Equations (4) and (5) [2]. The application of UV radiation for PDS activation made it possible to shorten the time necessary to achieve rhodamine B degradation compared e.g., to the H2O2, H2O2/UV and PDS processes. For example, after 10 min of the PDS/UV process, the obtained effectiveness was similar to the PDS system after 30 min of reaction time. A reaction time of 30 min in the PDS/UV system made it possible to achieve twice the effectiveness of the H2O2/UV process:
S 2 O 8 2 + energy 2 SO 4
S 2 O 8 2 UV 2 SO 4
S 2 O 8 2 + 2 H 2 O   2 SO 4 + H 2 O 2  
No rhodamine B decolorization was observed during the process of hydrogen peroxide as the sole source of radicals, as can be seen in Figure 1. Although hydrogen peroxide finds wide application as a low cost and commercially available oxidant, it shows certain practical limitations, e.g., it does not oxidize some types of organic pollutants. For example, the effectiveness of anthraquinone dye decolorization in literature [33] did not depend on the initial dose of H2O2. A double increase in the dose of hydrogen peroxide resulted in an increase of the decolorization effectiveness from 5.8% to 7.8% after 45 min of reaction time. However, the obtained results indicate that UV irradiation increases the RhB degradation effectiveness from <1% to 33% after 30 min of reaction time. The RhB degradation in the H2O2/UV process is achieved through the reactions of hydroxyl radicals formed through irradiation of H2O2 (Equation (6)). Literature data [34,35] indicates that the RhB degradation in the H2O2/UV system takes places primarily through radical mechanisms, however conversion via UV radiation absorption cannot be excluded as well (Equation (7)):
H 2 O 2 hv 2 OH
RhB h v O H degradation   compound ( s )   *
where * indicates that identification of degradation products was beyond the scope of this study

2.2. Kinetics Studies

The linear relationships [ln(C/C0)] and [1/(C/C0)] as functions of reaction time (min) are presented in Figure 2. As shown in the studies [36,37], the pseudo-first-order model is not always the best model to describe the reaction kinetics. For example, Haddad et al. [36] showed that the pseudo-second-order model showed a better fit in the Fenton process compared to the pseudo-first-order model. The above observations were also confirmed by Aljuboury and Palaniandy [37]. It is therefore justified to study both pseudo-first and pseudo-second order kinetics.
For each treatment systems, kinetic reaction parameters were determined based on the analyzed kinetic models (Table 2). As can be observed in Figure 2 and Table 2, the R2 values obtained for the pseudo-first-order model exhibit better fitting compared to the pseudo-second-order model. There is a very good relationship between ln(C/C0) and t (except for the H2O2 process), therefore the rhodamine B degradation followed a pseudo-first-order model. Satisfactory fitting (60–80%) was obtained only for the H2O2 process, as a result of the very low RhB degradation effectiveness (under 1%).
A phenomenon characteristic of advanced oxidation processes is the occurrence of (typically) two time intervals for the conducted oxidation reactions [38]. Other authors [39,40] observed a similar phenomenon as well, resulting most likely from the indirect reactions of oxidizing components with the pollutant degradation products. Based on the linear relationship analysis, it was found that this phenomenon does not occur in the studied AOP methods, as it is primarily characteristic of heterogeneous oxidation processes (e.g., TiO2 photocatalysis) or other processes conducted in the presence of solid catalysts (e.g., the Cu(II)-polyampholyte heterogeneous catalyst) [41]. In these types of processes, the first, “rapid” stage occurs due to the high availability of active centers in the catalyst and the simultaneous availability of a high number of oxidative radicals. As the reaction proceeds, the number of active centers becomes less available due to blockage of catalyst pores and lower radicals generation, which results in a transition to the second, “slow” stage of the reaction.
The obtained results demonstrate that each of the applied treatment methods is characterized by different kinetic parameter values. The O3/UV system exhibited the greatest steady reaction rate, which was a result of the ozone oxidation path, involving direct oxidation with ozone and radical oxidation with the OH radical.

2.3. Combined PDS/O3/UV

In the final stage of the studies, the model solution containing the rhodamine B dye standard at an initial concentration of 20, 50 and 100 mg/L was subjected to treatment in a combined PDS/O3/UV process. The process was carried out for 30 min. The results are presented in Figure 3a. The use of hydrogen peroxide was omitted in order to minimize the risk of oxidation process inhibition by an excessive quantity of H2O2, which may result in OH radical recombination and the auto-decomposition of H2O2 into H2O and O2, as evidenced [42,43,44].
Based on the obtained results, it was concluded that the decolorization degree depends on the initial dye concentration. For example, after 15 min, at concentrations of 20, 50 and 100 mg/L, the degree of RhB degradation was 42%, 59% and 99% respectively. The experimental conditions were the same for each of the analyzed systems (temperature, reaction, atmospheric pressure, reaction vessel volume, reaction vessel shape), so, presumably the quantity of the generated oxidative radicals was stable and decreased simultaneously with the increase in RhB concentration. This phenomenon was also confirmed in other publications [45,46], and it is influenced by e.g., dye particle density, which blocks UV penetration into the treated solution. Another reason for the decreased decolorization effect is the reduced likelihood of oxidative radical and dye particle collisions, or the fact that the radicals are first used for oxidizing the by-products of the oxidation process [47]. Also high concentration of dye molecules causes greater consumption of radicals resulting in an insufficient amount of radicals to decompose the excess of RhB. Moreover, intermediates may consume active species, which results in a poor dye decolorization (competitiveness effect). However, a detailed study of the possible formation of oxidation by-products should be carried out. Despite the high efficiency of AOPs for degradation of many pollutants, particular attention should be paid to the by-products of the oxidation process, which are often more toxic and dangerous than the original compounds [48].
The effectiveness of RhB degradation may also be influenced by the composition of wastewater, e.g., solid particles suspended in the treated wastewater, or the inorganic ions, such as Cl ions; CO32−; HCO3− or NO3 and NO2. Solid particles can increase the consumption of reagents needed for the oxidation of pollutants and absorb UV radiation. In turn, the ions can inhibit RhB degradation as a result of scavenging SO 4 and/or OH radicals [49].
Although the advanced oxidation process was conducted at high dye concentrations (50–100 mg/L), the combined PDS/O3/UV process, RhB removal of 69–79% after 30 min of reaction time was achieved. A similar effectiveness was obtained by only applying ozone or UV-activated persulfate, but at a concentration 2–5 times lower (20 mg/L).
A kinetic study was also performed in order to investigate the PDS/O3/UV process dynamics. Since it was demonstrated that the rhodamine B degradation followed a pseudo-first-order model, the analysis of RhB decolorization dynamics was performed based on this model. The relationship between [ln(C/C0)] and the process time t is presented in Figure 3b. As expected, the process performed at the lowest dye concentration showed the highest value of reaction rate constant k (Table 3). The high RhB concentrations resulted in an increase of the half-life (t/2) from 2.8 min (C0[RhB] = 20 mg/L) to 13.3 min and 20.3 min at concentrations of 50 and 100 mg/L respectively.
Table 4 presents the chemical reaction rate classification based on the mathematically calculated half-life t/2. Reactions of RhB decolorization in the concentration range of 20–50 mg/L were classified as moderate (1–103 s), whereas at a concentration of 100 mg/L were classified as slow (106–103 s), whereas the mathematically calculated half-life, expressed in seconds, was 1218 s.

3. Comparison with Other AOPs

This study showed that the combined PDS/O3/UV is a promising method for high rhodamine B concentrations (50–100 mg/L). The high concentration of the dye did not make it difficult to obtain over 50% RhB decomposition within 30 min. In this study, also ozone or UV-activated persulfate were also examined but at a concentration of 20 mg/L and similar efficiencies were obtained. As presented in Table 5, combined PDS/O3/UV is an interesting alternative to eliminate RhB in comparison with other studies. As is shown, the PDS/O3/UV process has great potential for the degradation of rhodamine B, particularly where the dye concentrations are high (up to 100 mg/L). Relatively short reaction time (30 min) is benefit of using this method. The reaction time needed to achieve 60% decolorization is two times shorter compared with the research performed by Rokesh et al. [51]. In turn, in a study by Jinisha et al. [52], it took 180 min to decompose approximately 98% of the RhB. The result obtained was found even better than those reported by Natarajan et al. [53], although the authors investigated the degradation of RhB at a lower concentration (about 50 mg/L).
This proposed process is characterized also by simplicity and low consumption of chemical reagents (excluding O3 and UV, only PDS is needed). Although not all of the dye was removed, this study was designed to use lower ozone flow (10 mL/min) comparing to 30 mL/min in the study by Cuiping et al. [55]. It can be assumed that the use of higher ozone flow would increase the efficiency of the process. Similar conclusions can be considered when analyzing higher lamp power. Another essential advantage of the combined PDS/O3/UV (under the experimental conditions) is the energy-saving power lamp. Although a lower energy lamp (10 W) was used in this study, satisfactory decolorization effects were obtained without the need for a higher wattage lamp (e.g., 16 W in the study by Cuiping et al. [55] or 300 W in the study by Rokesh et al. [51]). Hence, RhB need less energy consumption for degradation.
What distinguishes the use of PDS/O3/UV from other solutions presented in the table is the neutral reaction environment (pH = 6.0). It can be seen that most of the experiments shown in the table were performed in the very low pH range (2 to 4). The technology proposed in this article is based on carrying out the reaction at a pH close to neutral (pH = 6.0). Therefore, the costs of a potential pH correction are negligible.
It is worth noting that the PDS/O3/UV combination did not use solid catalysts (e.g., TiO2). Recent literature reports [58,59,60] identify titanium (Ti) as a contaminant due to the drastic increase in the use of Ti in personal care products (cosmetics, sunscreens), environmental applications (photocatalysts), and drug delivery systems. Sodium persulfate (PDS) is not expected to bioaccumulate (it decomposes into ions ubiquitous in the environment) and will not be absorbed into soil or sediments due to its dissociation and high water solubility [61], whereas ozone and UV radiation are commonly used in water disinfection. In addition, despite the possibility of regeneration of catalysts, such as titanium dioxide (washing with solvents, hydrogen peroxide and thermal treatment), their lifetime is limited, and the problem of their disposal after the process still remains. PDS readily dissolves in water, as mentioned above.
In conclusion, the use of various sources of free radicals (ozone, PDS, UV) can be useful to reduce the costs necessary to remove rhodamine B (60% at the initial concentration of 100 mg/L). As can be deduced from the above discussion, proposed technology is compatible with the principles of green chemistry, including: use of catalysts (PDS), design for degradation (degradation of RhB), waste prevention and less hazardous chemical (no chemical residues, lack of nanomaterials), energy efficiency (e.g., low energy-consumption lamp).

4. Materials and Methods

4.1. Materials

Aqueous solutions were prepared by dissolving the dye standard at a concentration of 20 mg/L (purity of >97%) in distilled water. This dye concentration (20 mg/L) can be identified in the raw wastewater from the textile industry as reported in Section 1. The dye was obtained from Chempur (Tarnowskie Góry, Poland). The pH of the model solution before and after the addition of the dye solution was pH = 6.0. Any deviations from this value, before and after the addition of any reagents, were corrected with 0.1 mol/L HCl or 0.1 mol/L NaOH from Sigma-Aldrich (Poznań, Poland). The aim of the study was to analyze the oxidation at initial pH = 6.0. The change in pH during the oxidation process was not controlled. The RhB degradation was carried out using persulfate and hydroxyl radicals generated from sodium persulfate (Na2S2O8), ozone and hydrogen peroxide. The chemical reagents used during the experiment were obtained from Sigma-Aldrich.

4.2. Advanced Oxidation Experiments

The advanced oxidation processes were conducted in glass vessels. The sodium persulfate oxidation (PDS) was assisted by a CUV-510 ultraviolet lamp (Grech, Maao, China) with a power of 10 W (PDS/UV). The dose of sodium persulfate (Na2S2O8) was set to 20 mM. Higher doses of sodium persulfate did not cause a significant increase in RhB degradation. The use of lower doses (2–10 mg/L) resulted in the removal efficiency of RhB ranging from 2–47 times less. A ZY-H103 ozone generator (Eltom, Warsaw, Poland) with a power of 20 W (ozone flow 10 mL/min) was used in the O3, O3/UV experiments. The dose of ozone was set to 100 mg/L. The O3 dose was similar to [62], but it should be noted, that ozone dose may vary depending on the specific application. Ozone was continuously fed to the reactor. In each experiment, the decolorization studies were conducted in glass vessels containing the RhB dye at a concentration of 20 mg/L (model solution). The dose of hydrogen peroxide was set to 40 mg/L (H2O2, H2O2/UV). The H2O2 dosage was similar to a dose range (30–2250 mg/L) tested by [63]. The hydrogen peroxide and PDS were added each time before the start of the experiment. A magnetic stirrer was used to ensure the optimal mixing of the model solutions. Reagents—radical precursors (sodium persulfate, ozone, hydrogen peroxide), were added into the model solutions. The PDS, O3 and H2O2 doses were determined by preliminary studies, in which various AOPs at different catalysts dose, pH, reaction vessel volume, temperature, and different dye concentrations were tested. The decolorization process was conducted continuously for 30 min. Samples for analysis were collected at 5, 10, 15, 20, 25 and 30 min intervals. Experimental conditions, i.e., the temperature (T = 298.0 K), reaction (pH = 6.0), atmospheric pressure (p = 1013.0 hPa), reaction vessel volume (V = 0.5 L) and reaction vessel shape (borosilicate glass cylinder beakers), remained unchanged throughout each experiment.

4.3. Instruments and Analytical Methods

At specific time intervals, the analytes were collected and analyzed with V-730 spectrophotometer (Jasco, Kraków, Poland). Figure 4 presents the obtained RhB absorption spectrum. The UV-Vis absorption spectrum was covered a range from approximately 400 to 700 nm. The compound absorption spectrum exhibits a peak at a wavelength λmax = 554.0 nm. The solutions were tested at the same wavelength (λmax = 554.0 nm). In solutions, in which the absorbance was more than 1, the suitable dilution was made to increase the sensitivity of quantitative analysis.

4.4. Kinetics Studies

Although studies indicate the use of pseudo-first-order kinetic model (Equation (8)) [64] in order to assess the fit of rhodamine B degradation rate to the appropriate model, in this study the pseudo-second-order model (Equation (9)) was also studied. The integrated dye mass balance in batch reactor are presented in Table 6.

5. Conclusions

The complexity and relevance of the problem of the presence of colored compounds in wastewater creates the urgent necessity to develop new and effective technologies for their elimination. Therefore, in this study the possibility of utilizing advanced oxidation processes (AOPs) for the degradation of the rhodamine B (RhB) dye was carried out. The advantage of AOPs for wastewater decolorization is the possibility of radical generation in various ways, which allow to select a more suitable method meeting the requirements for the treatment of colored solutions. The use of UV-assisted ozonation proved to be the most effective for RhB decolorization, i.e., at a concentration of 20 mg/L 90% of the dye was removed after 30 min of reaction time. Satisfactory results were also obtained in a combined PDS/O3/UV system (nearly 60% RhB removal at a concentration of 100 mg/L after 30 min). The results indicated that combined PDS/O3/UV method is promising for dyeing of rhodamine B aqueous solutions at a concentration up to 100 mg/L compared to other advanced oxidation processes.

Author Contributions

Conceptualization, Investigation, Methodology, Writing—original draft, P.Z.; Data curation, Formal analysis, Writing—review & editing, M.D. All authors have read and agreed to the published version of the manuscript.

Funding

The presented study was performed in the framework of the research work in the Central Mining Institute in Poland, financially supported by the Polish Ministry of Science and Higher Education (No. 11131041-340).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Syafiuddin, A.; Fulazzaky, M.A. Decolorization kinetics and mass transfer mechanisms of Remazol brilliant blue R dye mediated by different fungi. Biotechnol. Rep. 2020, 29, e00573. [Google Scholar] [CrossRef]
  2. Wang, C.; Yuan, Z.; Wang, A.; Qu, J.; Fang, Z.; Wen, Y. Ultraviolet light enhanced sodium persulfate oxida-tion of cellulose to facilitate the preparation of cellulose nanofibers. Cellulose 2020, 27, 2041–2051. [Google Scholar] [CrossRef]
  3. Zawadzki, P.; Kudlek, E.; Dudziak, M. Titanium (IV) oxide modified with activated carbon and ultrasounds for caffeine photodegradation: Adsorption isotherm and kinetics study. J. Ecol. Eng. 2020, 21, 137–145. [Google Scholar] [CrossRef]
  4. Jain, R.; Mathur, M.; Sikarwar, S.; Mittal, D.A. Removal of the hazardous dye rhodamine B through photocatalytic and adsorption treatments. J. Environ. Manag. 2007, 85, 956–964. [Google Scholar] [CrossRef] [PubMed]
  5. Yahia, I.S.; Rammah, Y.S.; Khaled, K.F. Fabrication of an electrochemical cell based on Rhodamine B dye for low power applications. J. Mater. Environ. Sci. 2013, 4, 442–447. [Google Scholar]
  6. Garrido-Cardenas, J.A.; Esteban-García, B.; Agüera, A.; Sánchez-Pérez, J.A.; Manzano-Agugliaro, F. Wastewater treatment by advanced oxidation process and their worldwide research trends. Int. J. Environ. Res. Public Health 2019, 17, 170. [Google Scholar] [CrossRef] [Green Version]
  7. Tariqa, M.; Muhammada, M.; Khana, J.; Raziqa, A.; KashifUddinbc, M.; Niazd, A.; Ahmedef, S.; Rahimg, A. Removal of Rhodamine B dye from aqueous solutions using photo-fenton processes and novel Ni-Cu@MWCNTs photocatalyst. J. Mol. Liq. 2020, 312, 113399. [Google Scholar] [CrossRef]
  8. Dhahir, S.A.; Al-saade, K.A.; Al-Jobouri, I.S. Degradation studies of rhodamine b in the presence of UV/H2O2/Fe2+. Int. J. Tech. Res. Appl. 2014, 2, 123–127. [Google Scholar]
  9. Devi, P.; Das, U.; Dalai, A.K. In-situ chemical oxidation: Principle and applications of peroxide and persulfate treatments in wastewater systems. Sci. Total Environ. 2016, 571, 643–657. [Google Scholar] [CrossRef]
  10. Ike, I.; Linden, K.G.; Orbell, J.D.; Duke, M. Critical review of the science and sustainability of persulphate advanced oxidation processes. Chem. Eng. J. 2018, 338, 651–669. [Google Scholar] [CrossRef]
  11. Matzek, L.W.; Carter, K.E. Activated persulfate for organic chemical degradation: A review. Chemosphere 2016, 151, 178–188. [Google Scholar] [CrossRef]
  12. Wacławek, S.; Lutze, H.; Grübel, K.; Padil, V.V.T.; Černík, M.; Dionysiou, D. Chemistry of persulfates in water and wastewater treatment: A review. Chem. Eng. J. 2017, 330, 44–62. [Google Scholar] [CrossRef]
  13. Wang, J.; Wang, S. Activation of persulfate (PS) and peroxymonosulfate (PMS) and application for the degradation of emerging contaminants. Chem. Eng. J. 2018, 334, 1502–1517. [Google Scholar] [CrossRef]
  14. Giannakis, S.; Lin, K.-Y.A.; Ghanbari, F. A review of the recent advances on the treatment of industrial wastewaters by sulfate radical-based advanced oxidation processes (SR-AOPs). Chem. Eng. J. 2020, 406, 127083. [Google Scholar] [CrossRef]
  15. Oh, W.-D.; Dong, Z.; Lim, T.-T. Generation of sulfate radical through heterogeneous catalysis for organic contaminants removal: Current development, challenges and prospects. Appl. Catal. B Environ. 2016, 194, 169–201. [Google Scholar] [CrossRef]
  16. Yang, J.; Zhu, M.; Dionysiou, D.D. What is the role of light in persulfate-based advanced oxidation for water treatment? Water Res. 2020, 189, 116627. [Google Scholar] [CrossRef]
  17. Zawadzki, P. Comparative studies of Rhodamine B decolorization in the combined process Na2S2O8/visible light/ultrasound. Desalination Water Treat. 2021, 213, 269–278. [Google Scholar] [CrossRef]
  18. Heidarpour, H.; Padervand, M.; Soltanieh, M.; Vossoughi, M. Enhanced decolorization of rhodamine B solution through simultaneous photocatalysis and persulfate activation over Fe/C3N4 photocatalyst. Chem. Eng. Res. Des. 2019, 153, 709–720. [Google Scholar] [CrossRef]
  19. Yaseen, D.A.; Scholz, M. Textile dye wastewater characteristics and constituents of synthetic effluents: A critical review. Int. J. Environ. Sci. Technol. 2018, 16, 1193–1226. [Google Scholar] [CrossRef] [Green Version]
  20. Mathon, B.; Coquery, M.; Miege, C.; Penru, Y.; Choubert, J.-M. Removal efficiencies and kinetic rate constants of xenobiotics by ozonation in tertiary treatment. Water Sci. Technol. 2017, 75, 2737–2746. [Google Scholar] [CrossRef] [PubMed]
  21. Zhang, H.; Duan, L.; Zhang, D. Decolorization of methyl orange by ozonation in combination with ultrasonic irradiation. J. Hazard. Mater. 2006, 138, 53–59. [Google Scholar] [CrossRef] [PubMed]
  22. Atkinson, R.; Baulch, D.L.; Cox, R.A.; Hampson, R.F.; Kerr, J.A.; Rossi, M.J.; Troe, J. Evaluated kinetic and photochemical data for atmospheric chemistry: Supplement VI. IUPAC subcommittee on gas kinetic data evaluation for atmospheric chemistry. J. Phys. Chem. Ref. Data 1997, 26, 1329. [Google Scholar] [CrossRef] [Green Version]
  23. Faryadi, M.; Rahimi, M.; Akbari, M. Process modeling and optimization of Rhodamine B dye ozonation in a novel microreactor equipped with high frequency ultrasound wave. Korean J. Chem. Eng. 2015, 33, 922–933. [Google Scholar] [CrossRef]
  24. Sierra, B.J.; Cabrera, A.M.; Arias, A.N.H.; Eguiluz, R.P.; Callejas, R.L.; Méndez, B.G.R.; Alvarado, R.V. Methylene blue degradation assessment by advance oxidation methods. J. Appl. Res. Technol. 2019, 17. [Google Scholar] [CrossRef] [Green Version]
  25. Ding, X.; Gutierrez, L.; Croue, J.-P.; Li, M.; Wang, L.; Wang, Y. Hydroxyl and sulfate radical-based oxidation of RhB dye in UV/H2O2 and UV/persulfate systems: Kinetics, mechanisms, and comparison. Chemosphere 2020, 253, 126655. [Google Scholar] [CrossRef] [PubMed]
  26. Smaali, A.; Berkani, M.; Merouane, F.; Le, V.T.; Vasseghian, Y.; Rahim, N.; Kouachi, M. Photocatalytic persulfate oxidation for diclofenac removal from aqueous solutions: Modeling, optimization and biotoxicity test assessment. Chemosphere 2020, 266, 129158. [Google Scholar] [CrossRef]
  27. Jiang, Z.; Li, J.; Jiang, D.; Gao, Y.; Chen, Y.; Wang, W.; Cao, B.; Tao, Y.; Wang, L.; Zhang, Y. Removal of atrazine by biochar-supported zero-valent iron catalyzed persulfate oxidation: Reactivity, radical production and transformation pathway. Environ. Res. 2020, 184, 109260. [Google Scholar] [CrossRef]
  28. Usman, M.; Alam Cheema, S.; Farooq, M. Heterogeneous fenton and persulfate oxidation for treatment of landfill leachate: A review supplement. J. Clean. Prod. 2020, 256, 120448. [Google Scholar] [CrossRef]
  29. Ahmadi, S.; Igwegbe, C.; Rahdar, S. The application of thermally activated persulfate for degradation of acid blue 92 in aqueous solution. Int. J. Ind. Chem. 2019, 10, 249–260. [Google Scholar] [CrossRef] [Green Version]
  30. Piotr, Z. Decolorisation of methylene blue with sodium persulfate activated with visible light in the presence of glucose and sucrose. Water Air Soil Pollut. 2019, 230, 313. [Google Scholar] [CrossRef]
  31. Liu, C.S.; Higgins, C.P.; Wang, F.; Shih, K. Effect of temperature on oxidative transformation of perfluorooctanoic acid (PFOA) by persulfate activation in water. Sep. Purif. Technol. 2012, 91, 46–51. [Google Scholar] [CrossRef]
  32. Tsitonaki, A.; Petri, B.; Crimi, M.; Mosbæk, H.; Siegrist, R.L.; Bjerg, P.L. In situ chemical oxidation of contaminated soil and groundwater using persulfate: A review. Crit. Rev. Environ. Sci. Technol. 2010, 40, 55–91. [Google Scholar] [CrossRef]
  33. Perkowski, J.; Ledakowicz, S. Decomposition of anthraquinone dye in the aqueous solution by ozone, hydrogen peroxide or UV radiation. Fibres Text. East. Eur. 2002, 10, 72–77. [Google Scholar]
  34. Simonenko, E.; Gomonov, A.; Rolle, N.; Molodkina, L. Modeling of H2O2 and UV oxidation of organic pollutants at wastewater post-treatment. Procedia Eng. 2015, 117, 337–344. [Google Scholar] [CrossRef] [Green Version]
  35. Hamedi, F.H.; Rauf, M.; Ashraf, S. Degradation studies of Rhodamine B in the presence of UV/H2O2. Desalination 2009, 239, 159–166. [Google Scholar] [CrossRef]
  36. Haddad, M.E.; Regti, A.; Laamari, M.R.; Mamouni, R.; Saffaj, N. Use of fenton reagent as advanced oxidative process for removing textile dyes from aqueous solutions. J. Mater. Environ. Sci. 2014, 5, 667–674. [Google Scholar]
  37. Aljuboury, D.A.D.; Palaniandy, P. Kinetic study of inorganic carbon (IC) removal and COD removal from refinery wastewater by solar photo-fenton. Glob. NEST J. 2017, 19, 641–649. [Google Scholar]
  38. Zawadzki, P.; Kudlek, E.; Dudziak, M. Kinetics of the photocatalytic decomposition of bisphenol a on modified photocatalysts. J. Ecol. Eng. 2018, 19, 260–268. [Google Scholar] [CrossRef]
  39. Covinich, L.G.; Felissia, F.; Massa, P.; Fenoglio, R.; Area, M.C. Kinetic modeling of a heterogeneous fenton-type oxidative treatment of complex industrial effluent. Int. J. Ind. Chem. 2018, 9, 215–229. [Google Scholar] [CrossRef] [Green Version]
  40. Idrissi, M.; Miyah, Y.; Benjelloun, Y.; Chaouch, M. Degradation of crystal violet by heterogeneous fenton-like reaction using Fe/Clay catalyst with H2O2. J. Mater. Environ. Sci. 2016, 7, 50–58. [Google Scholar]
  41. Martínez, J.M.L.; Denis, M.F.L.; Piehl, L.L.; de Celis, E.R.; Buldain, G.Y.; Orto, V.C.D. Studies on the activation of hydrogen peroxide for color removal in the presence of a new Cu(II)-polyampholyte heterogeneous catalyst. Appl. Catal. B Environ. 2008, 82, 273–283. [Google Scholar] [CrossRef]
  42. Ziembowicz, S.; Kida, M.; Koszelnik, P. Sonochemical formation of hydrogen peroxide. Proceedings 2017, 2, 188. [Google Scholar] [CrossRef] [Green Version]
  43. Mandal, T.; Maity, S.; Dasgupta, D.; Datta, S. Advanced oxidation process and biotreatment: Their roles in combined industrial wastewater treatment. Desalination 2010, 250, 87–94. [Google Scholar] [CrossRef]
  44. Ku, Y.; Tu, Y.-H.; Ma, C.-M. Effect of hydrogen peroxide on the decomposition of monochlorophenols by sonolysis in aqueous solution. Water Res. 2005, 39, 1093–1098. [Google Scholar] [CrossRef]
  45. Thangavadivel, K.; Megharaj, M.; Mudhoo, A.; Naidu, R. Degradation of organic pollutants using ultra-sound. In Handbook on Application of Ultrasound: Sonochemistry for Sustainability; Sharma, S.K., Mudhoo, A., Eds.; Taylor and Francis Group: Abingdon-on-Thames, UK, 2011. [Google Scholar]
  46. Daneshvar, N.; Behnajady, M.A.; Mohammadi, M.K.A.; Dorraji, M.S.S. UV/H2O2 treatment of Rhodamine B in aqueous solution: Influence of operational parameters and kinetic modeling. Desalination 2008, 230, 16–26. [Google Scholar] [CrossRef]
  47. Cai, C.; Zhang, H.; Zhong, X.; Hou, L. Ultrasound enhanced heterogeneous activation of peroxymonosulfate by a bimetallic Fe–Co/SBA-15 catalyst for the degradation of orange II in water. J. Hazard. Mater. 2015, 283, 70–79. [Google Scholar] [CrossRef] [PubMed]
  48. Lellis, B.; Fávaro-Polonio, C.Z.; Pamphile, J.A.; Polonio, J.C. Effects of textile dyes on health and the environment and bioremediation potential of living organisms. Biotechnol. Res. Innov. 2019, 3, 275–290. [Google Scholar] [CrossRef]
  49. Ahmed, N.; Vione, D.; Rivoira, L.; Carena, L.; Castiglioni, M.; Bruzzoniti, M. A review on the degradation of pollutants by fenton-like systems based on zero-valent iron and persulfate: Effects of reduction potentials, pH, and anions occurring in waste waters. Molecules 2021, 26, 4584. [Google Scholar] [CrossRef] [PubMed]
  50. Wright, M.R. An Introduction to Chemical Kinetics; John Wiley and Sons: Hoboken, NJ, USA, 2004. [Google Scholar]
  51. Rokesh, K.; Mohan, S.C.; Karuppuchamy, S.; Jothivenkatachalam, K. Photo-assisted advanced oxidation processes for Rhodamine B degradation using ZnO-Ag nanocomposite materials. J. Environ. Chem. Eng. 2018, 6, 3610–3620. [Google Scholar] [CrossRef]
  52. Jinisha, R.; Gandhimathi, R.; Ramesh, S.; Nidheesh, P.; Velmathi, S. Removal of rhodamine B dye from aqueous solution by electro-fenton process using iron-doped mesoporous silica as a heterogeneous catalyst. Chemosphere 2018, 200, 446–454. [Google Scholar] [CrossRef] [PubMed]
  53. Jiang, L.; Zhang, Y.; Zhou, M.; Liang, L.; Li, K. Oxidation of Rhodamine B by persulfate activated with porous carbon aerogel through a non-radical mechanism. J. Hazard. Mater. 2018, 358, 53–61. [Google Scholar] [CrossRef]
  54. Natarajan, T.S.; Thomas, M.; Natarajan, K.; Bajaj, H.C.; Tayade, R.J. Study on UV-LED/TiO2 process for degradation of Rhodamine B dye. Chem. Eng. J. 2011, 169, 126–134. [Google Scholar] [CrossRef]
  55. Cuiping, B.; Xianfeng, X.; Wenqi, G.; Dexin, F.; Mo, X.; Zhongxue, G.; Nian, X. Removal of rhodamine B by ozone-based advanced oxidation process. Desalination 2011, 278, 84–90. [Google Scholar] [CrossRef]
  56. Maharana, D.; Niu, J.; Gao, D.; Xu, Z.; Shi, J. Electrochemical degradation of rhodamine B over Ti/SnO2-sb electrode. Water Environ. Res. 2015, 87, 304–311. [Google Scholar] [CrossRef]
  57. Hou, M.-F.; Liao, L.; Zhang, W.-D.; Tang, X.; Wan, H.-F.; Yin, G.-C. Degradation of rhodamine B by Fe(0)-based Fenton process with H2O2. Chemosphere 2011, 83, 1279–1283. [Google Scholar] [CrossRef]
  58. Hauser-Davis, R.A.; Monteiro, F.; Chavez da Rocha, R.C.; Lemos, L.; Cardoso, M.D.; Siciliano, S. Titanium as a contaminant of emerging concern in the aquatic environment and the current knowledge gap regarding sea-bird contamination. Ornithologia 2020, 11, 7–15. [Google Scholar]
  59. Asztemborska, M.; Jakubiak, M.; Stęborowski, R.; Chajduk, E.; Bystrzejewska-Piotrowska, G. Titanium dioxide nanoparticle circulation in an aquatic ecosystem. Water Air Soil Pollut. 2018, 229, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Borghesi, F.; Migani, F.; Andreotti, A.; Baccetti, N.; Bianchi, N.; Birke, M.; Dinelli, E. Metals and trace elements in feathers: A geochemical approach to avoid misinterpretation of analytical responses. Sci. Total Environ. 2016, 544, 476–494. [Google Scholar] [CrossRef] [PubMed]
  61. Sodium Persulfate—The Dossier. Available online: https://www.santos.com/wp-content/uploads/2021/04/Sodium-persulfate-March-2021.pdf (accessed on 12 August 2021).
  62. Yasar, A.; Ahmad, N.; Chaudhry, M.N.; Rehman, M.S.U.; Khan, A.A.A. Ozone for color and COD removal of raw and anaerobically biotreated combined industrial wastewater. Pol. J. Environ. Stud. 2007, 16, 289–294. [Google Scholar]
  63. Collivignarelli, M.C.; Pedrazzani, R.; Sorlini, S.; Abbà, A.; Bertanza, G. H2O2 based oxidation processes for the treatment of real high strength aqueous wastes. Sustainability 2017, 9, 244. [Google Scholar] [CrossRef] [Green Version]
  64. Asenjo, N.G.; Santamaría, R.; Blanco, C.; Granda, M.; Álvarez, P.; Menéndez, R. Correct use of the Langmuir-Hinshelwood equation for proving the absence of a synergy effect in the photocatalytic degradation of phenol on a suspended mixture of titania and activated carbon. Carbon 2013, 55, 62–69. [Google Scholar] [CrossRef]
  65. Armenise, S.; Garcia-Bordeje, E.; Valverde, J.L.; Romeo, E.; Monzon, A. A Langmuir–Hinshelwood approach to the kinetic modelling of catalytic ammonia decomposition in an integral reactor. Phys. Chem. Chem. Phys. 2013, 15, 12104. [Google Scholar] [CrossRef] [PubMed]
  66. Marcinkowski, D.; Wałęsa-Chorab, M.; Patroniak, V.; Kubicki, M.; Kądziołka, G.; Michalkiewicz, B. A new polymeric complex of silver(i) with a hybrid pyrazine-bipyridine ligand-synthesis, crystal structure and its photocatalytic activity. New J. Chem. 2013, 38, 604–610. [Google Scholar] [CrossRef]
Figure 1. Photodegradation of rhodamine B during different AOP. Conditions: C0[RhB] = 20.0 mg/L; V = 0.5 L; T = 298.0 K; p = 1013.0 hPa; pH = 6.0; t = 0–30 min.
Figure 1. Photodegradation of rhodamine B during different AOP. Conditions: C0[RhB] = 20.0 mg/L; V = 0.5 L; T = 298.0 K; p = 1013.0 hPa; pH = 6.0; t = 0–30 min.
Catalysts 11 00974 g001
Figure 2. Kinetics of rhodamine B degradation (a—pseudo-first-order; b—pseudo-second-order). Conditions: C0[RhB] = 20.0 mg/L; V = 0.5 L; T = 298.0 K; p = 1013.0 hPa; pH = 6.0; t = 0–30 min.
Figure 2. Kinetics of rhodamine B degradation (a—pseudo-first-order; b—pseudo-second-order). Conditions: C0[RhB] = 20.0 mg/L; V = 0.5 L; T = 298.0 K; p = 1013.0 hPa; pH = 6.0; t = 0–30 min.
Catalysts 11 00974 g002
Figure 3. Photodegradation of rhodamine B during different AOP (a) and reaction kinetics (b). Conditions: C0[RhB] = 20–100 mg/L; V = 0.5 L; T = 298.0 K; p = 1013.0 hPa; pH = 6.0; t = 0–30 min.
Figure 3. Photodegradation of rhodamine B during different AOP (a) and reaction kinetics (b). Conditions: C0[RhB] = 20–100 mg/L; V = 0.5 L; T = 298.0 K; p = 1013.0 hPa; pH = 6.0; t = 0–30 min.
Catalysts 11 00974 g003
Figure 4. Absorption spectrum of rhodamine B standard (20 mg/L).
Figure 4. Absorption spectrum of rhodamine B standard (20 mg/L).
Catalysts 11 00974 g004
Table 1. Molecular structure and some physicochemical properties of RhB [5].
Table 1. Molecular structure and some physicochemical properties of RhB [5].
Chemical StructurePhysicochemical Properties
Catalysts 11 00974 i001Chemical formulaC28H31CIN2O3
Molecular weight [g/mol]479.00
CAS number81-88-9
Water solubility at 20 °C [g/L]15.0
Dissociation constant (pKa) [–]3.7
logKOW [–]1.9–2.0
λmax in water [nm]554.0
Table 2. Kinetic parameters of the advanced oxidation process.
Table 2. Kinetic parameters of the advanced oxidation process.
ModelParameterProcess
H2O2H2O2/UVPDSPDS/UVO3O3/UV
pseudo-first-orderk, 1/min0.000.00650.01260.04080.04590.0726
R20.620.990.960.990.990.99
pseudo-second-orderk, 1/min0.000.00700.01500.07680.10220.2625
R20.610.990.940.950.980.87
Table 3. Kinetic parameters of the combined advanced oxidation process (PDS/O3/UV) in a different dye concentrations.
Table 3. Kinetic parameters of the combined advanced oxidation process (PDS/O3/UV) in a different dye concentrations.
ParameterDye Concentration, mg/L
2050100
k, 1/min0.25130.05230.0341
R20.990.980.96
t/22.813.320.3
Table 4. Classification of reaction rates based on half-lives [50].
Table 4. Classification of reaction rates based on half-lives [50].
Half-Life, sTime Span for Near-CompletionRate Classification
10−15–10−12ps or lessultra fast rate
10−12–10−6µs or lessvery fast rate
10−6–1secondsfast rate
1–103minutes or hoursmoderate rate
106–103weeksslow rate
>106weeks or yearsvery slow rate
Table 5. The summary of different AOPs applied for RhB degradation.
Table 5. The summary of different AOPs applied for RhB degradation.
AOPConditionsRemoval Efficiency, %References
PDS/O3/UVC0[RhB] = 100 mg/L
pH = 6.0
time = 30 min
O3 flow = 10 mL/min
T = 295.0 K
Lamp = 10 W
H2O2 dose = 40 mg/L
PDS dose = 20 mM
60.0This study
PDS/Vis/USC0[RhB] = 10 mg/L
pH = 6.0
time = 60 min
T = 295.0 K
US = 40 kHz, 60 W
PDS dose = 20 mM
Lamp = 10 W
85.0[17]
Peroxide assisted photocatalytic degradation in the presence of ZnOC0[RhB] = 5 mg/L
time = 90 min
Catalyst dose = 500 mg/L
lamp (300 W)
60.0[51]
Electro-FentonC0[RhB] = 10 mg/L
pH = 2.0
time = 180 min
Electrode dose = 15 mg/L
Voltage = 8 V
97.7[52]
Carbon aerogel/persulfateC0[RhB] = 10 mg/L
pH = 7.0
time = 60 min
Persulfate dose = 1.0 mM
Carbon aerogel dose = 100 mg/L
80.0[53]
UV-LED/TiO2C0[RhB] = 49 mg/L
time = 180 min
pH = 3.05
TiO2 = 1.6 g/L
Lamp = LED (5)
luminous intensity = 350 mcd
radiant flux = 10–12 mW at 20 mA
51.0[54]
O3/UVC0[RhB] = 100 mg/L
pH = 3.0
time = 15 min
ozone flow = 30 mL/min
Lamp = 16 W
99.8[55]
Ti/SnO2-Sb electrodeC0[RhB] = 50 mg/L
pH = 3.0
time = 20 min
at a constant current density of 20 mA/cm2 with a 10 mmol/L Na2SO4 supporting electrolyte solution
97.8[56]
Fe(0)-based Fenton process with H2O2C0[RhB] = 49 mg/L
Time = 30 min
pH = 4.0
Fe(0) = 1.0 g/L
H2O2 = 2.0 mM
100.0[57]
Table 6. The equation of kinetics.
Table 6. The equation of kinetics.
ModelPlotReferences
Pseudo-first-order ln ( C t C 0 ) VS   t (8)[65]
Pseudo-second-order 1 C t / C 0 VS   t (9)[66]
where: Ct—concentration at time t (mg/L); C0—initial concentration (mg/L); k1—pseudo-first-order rate constant; k2—pseudo-second-order rate constant.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zawadzki, P.; Deska, M. Degradation Efficiency and Kinetics Analysis of an Advanced Oxidation Process Utilizing Ozone, Hydrogen Peroxide and Persulfate to Degrade the Dye Rhodamine B. Catalysts 2021, 11, 974. https://doi.org/10.3390/catal11080974

AMA Style

Zawadzki P, Deska M. Degradation Efficiency and Kinetics Analysis of an Advanced Oxidation Process Utilizing Ozone, Hydrogen Peroxide and Persulfate to Degrade the Dye Rhodamine B. Catalysts. 2021; 11(8):974. https://doi.org/10.3390/catal11080974

Chicago/Turabian Style

Zawadzki, Piotr, and Małgorzata Deska. 2021. "Degradation Efficiency and Kinetics Analysis of an Advanced Oxidation Process Utilizing Ozone, Hydrogen Peroxide and Persulfate to Degrade the Dye Rhodamine B" Catalysts 11, no. 8: 974. https://doi.org/10.3390/catal11080974

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