Next Article in Journal
Surface Treatment of Metals
Previous Article in Journal
Tribological Properties of the 40Cr/GCr15 Tribo-Pair under Unidirectional Rotary and Reciprocating Dry Sliding
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of a Heterogeneous Catalyst CuO-Fe2O3/CTS-ATP and Degradation of Methylene Blue and Ciprofloxacin

School of Petrochemical Engineering, Lanzhou University of Technology, 287 Langongping Road, Lanzhou 730050, China
*
Author to whom correspondence should be addressed.
Coatings 2022, 12(5), 559; https://doi.org/10.3390/coatings12050559
Submission received: 23 March 2022 / Revised: 16 April 2022 / Accepted: 18 April 2022 / Published: 20 April 2022
(This article belongs to the Topic Multiple Application for Novel and Advanced Materials)

Abstract

:
A heterogeneous particle catalyst (CuO-Fe2O3/CTS-ATP) was synthesized via injection molding and ultrasonic immersion method, which is fast and effective. The particle catalyst applied attapulgite (ATP) wrapped by chitosan (CTS) as support, which was loaded dual metal oxides CuO and Fe2O3 as active components. After a series of characterizations of catalysts, it was found that CuO and Fe2O3 were successfully and evenly loaded on the surface of the CTS-ATP support. The catalyst was used to degrade methylene blue (MB) and ciprofloxacin (CIP), and the experimental results showed that the degradation ratios of MB and CIP can reach 99.29% and 86.2%, respectively, in the optimal conditions. The degradation mechanism of as-prepared catalyst was analyzed according to its synthesis process and ∙OH production, and the double-cycle catalytic mechanism was proposed. The intermediate products of MB and CIP degradation were also identified by HPLC-MS, and the possible degradation pathways were put forward.

Graphical Abstract

1. Introduction

Refractory organics (ROS), such as dyes, antibiotics, pesticides, etc., always threaten the survival and health of animals and human beings [1,2,3,4]. Methylene blue (MB) is a common dye in the printing and dyeing industry, and it is also used as a chemical indicator and biological dye [5]. Ciprofloxacin (CIP), as one of the third generation of antibacterial medicines, is widely used for fighting against various organ infections due to its broad-spectrum antibacterial activity. Due to a high percentage (60%–90%) of CIP discharge into the natural environment, nowadays CIP is considered to be one of the most important emerging pollutants due to their inhibition of bacterial activities [6,7,8,9,10]. Ciprofloxacin is generally difficult to biodegrade and enters the environment in the form of prototypes or metabolites [11,12], which poses a serious threat to the health of humans and aquatic organisms.
As the most promising methods, advanced oxidation processes, such as Fenton reactions, photocatalysis, electrocatalysis, ozone oxidation, etc., are often used to deal with refractory organic compounds in water which can produce a lot of hydroxyl radicals with strong oxidizing activity [13,14,15,16,17,18]. Fenton reactions, including all kinds of Fenton-like reactions, have become a research hotspot in the field of water treatment due to its high degradation efficiency, mild reaction conditions, low energy consumption, and simple operation [19,20,21,22]. In particular, the heterogeneous Fenton method has the advantages of wide pH range, low H2O2 and catalysts consumption, easy separation, and no secondary pollution with high removal ratios [23,24,25,26].
Attapulgite clay is a crystalline hydrated magnesium aluminium silicate mineral with unique layer-chain structure. It is made up of two layers of silicon-oxygen tetrahedrons that are sandwiched with a layer of magnesium (aluminium) oxygen octahedron. Attapulgite has various excellent properties: good adsorption and catalytic properties because it is more porous, has a high specific surface area, and is rich in active groups [27,28,29,30]. Chitosan is the product of natural polysaccharide chitin by removing some acetyl groups, and it has many physiological functions, such as biodegradability, biocompatibility, anti-bacteria, anti-cancer, lipid-lowering, immunity enhancement, etc. [31,32,33]. Chitosan molecules have alkaline aminopolysaccharides with positive charge, which make them have good adsorption properties, and chitosan is widely used in textile, environmental protection, agriculture and other fields as a sorbent [31,32,33]. Introducing chitosan into attapulgite can effectively prevent the agglomeration of attapulgite, and attapulgite-chitosan composites are often used to treat wastewater [34,35,36,37,38]. For examples, Liang [35] et al. coated magnetic Fe3O4/APT nanoparticles modified by aminopropyltrimethoxy silane (APTS) with chitosan gel beads, and found that the adsorption capacity of this catalyst was much higher than that of attapulgite alone. Pan et al. [36] prepared low-cost chitosan/attapulgite composites by the self-assembly method for the removal of uranium in aqueous solution. Similarly, Liao et al. [37] prepared a novel three-dimensional porous polydopamine functionalized attapulgite/chitosan (AT@PDA/CS) aerogel for the adsorption of uranium, and when the pH of the solution was 5, it reached adsorption equilibrium in 40 min. Shi et al. [38] prepared divalent copper ion imprinted chitosan/attapulgite polymer, which showed good selectivity to copper ions, and the maximum experimental adsorption value reached 35.20 mg/g.
Most of the recent research about chitosan/attapulgite composites focused on adsorption materials, and few of those concentrated on catalytic materials. In this study, in order to study the catalytic abilities of attapulgite/chitosan composites, a catalyst CuO-Fe2O3/CTS-ATP was prepared by combining attapulgite with chitosan as a carrier and loading metal oxides on CTS-ATP support with an ultrasonic impregnation method. Dyes and antibiotics are refractory organics with complex and stable structures, for investigating the catalytic performance on ROS of the prepared catalyst, methylene blue (MB) from dyes and ciprofloxacin (CIP) from antibiotics were selected as target pollutants for their hard-to-degrade properties and harm to humans and the environment [39,40], and the catalytic abilities of prepared catalyst were evaluated when using it to degrade MB and CIP.

2. Materials and Methods

2.1. Materials

Methylene blue (C16H18ClN3S), ciprofloxacin (C17H18FN3O3), chitosan, hydrochloric acid, concentrated sulfuric acid (98%), hydrogen peroxide (30% w/w), anhydrous ethanol, sodium hydroxide, Fe(NO3)3·9H2O, and Cu (NO3)2·3H2O were all supplied by manufactures and analytically pure. Attapulgite clay (92%–98%) was obtained from Jiangsu province in China. The water used in experiments is all deionized water.

2.2. Preparation of CuO-Fe2O3/CTS-ATP

Catalyst CuO-Fe2O3/CTS-ATP was prepared by combining attapulgite with chitosan as a carrier and loading metal oxides on CTS-ATP support with the ultrasonic impregnation method. Firstly, 1 mL acetic acid was added into 99 mL deionized water to prepare 1% wt acetic acid solution, and 2 g attapulgite clay and 2 g chitosan were added into the prepared acetic acid solution, with magnetic stirring at room temperature until the two were fully mixed. Then, the mixed CTS-ATP gel solution was evenly poured into a 50 mL syringe and was dripped into 200 mL NaOH (1 mol/L) solution dropwise, and small CTS-ATP gel particles were dispersed in the solution. Then, 5 mL of 2.5% wt glutaraldehyde solution were added into the above solution. After an hour’s reaction, the prepared CTS-ATP gel particles were filtered and put into the drying oven at 100 °C for an hour, then put into the tube furnace and roasted with nitrogen atmosphere (15 mL/min) at 400 °C for 2 h, so black particles of CTS-ATP carrier with a diameter of 1 mm were obtained. Add the particles into a certain concentration of ferric nitrate and copper nitrate (1:1) mixing solutions (5% wt for MB degradation and 15% for CIP degradation, Supplementary Materials Figure S1) and impregnated with ultrasonic for 50 min (Figure S2). Finally, the particles were taken out and rinsed with deionized water, and roasted for 1 h (temperature 200 °C, nitrogen inlet rate 15 mL/min) in the tube furnace to obtain heterogeneous catalyst CuO-Fe2O3/CTS-ATP.

2.3. Characterization

The crystalline structure and stability of ATP, Fe2O3/CTS-ATP, and CuO-Fe2O3/CTS-ATP were identified by the Palytical X ‘PERT PRO X-ray diffraction (XRD, Almelo City?, The Netherlands) analyzer, operating voltage and current were 40 kV and 150 mA, the diffraction Angle was 5°–80°, and the scanning step was 0.02°/s. All samples above were tested by a Nicoletavtar 360 FT-IR spectrometer (Almelo, Madison, WI, USA), and the bonding mode between the compounds and elements was analyzed. The scanning wave range is 4000–400 cm−1.The surface morphology of samples were observed by a JEOL JSM-6701F scanning electron microscopy (SEM, JEOL, Tokyo, Japan). The acceleration voltage was 20 kV, and the samples’ surface needs Pt sprayed for good electrical conductivity. An energy dispersive spectrometer (EDS QUANTAX) (produced by Bruker Corporation, Rheinstetten, Germany) was used to analyze the content and element mapping of the main elements in the samples. The N2 adsorption and desorption isotherms of the samples were measured at −197 °C by the ASAP2010 specific surface area tester manufactured by Micromeritics, Norcross, GA, USA. The specific surface area, pore volume, and pore size were calculated by the BET equation. The qualitative and quantitative analysis of elements on the surface of materials, as well as the analysis of chemical valence states and valence electron states, were tested by the PHI5702 X-Ray photoelectron spectrometer (XPS) (produced by the American Physical Electronics Company, Chanhassen, MN, USA). The X-ray emission sources are Mg, Al double-anode target, and Al monochromator target.

2.4. Degradation Experiments of MB and CIP by the Catalysts

In addition, 1 g/L methylene blue and 250 mg/L ciprofloxacin stock solutions were prepared for later use. An appropriate amount of MB or CIP stock solutions and a certain amount of deionized water were added into a 250 mL flask to obtain MB or CIP solutions with different concentrations. Add catalyst and 30% w/w hydrogen peroxide into the flask, stirring it at a certain temperature. Take water samples every 10 min, and measure the absorbance of the solution by UV-1900 UV-visible spectrophotometer (Aoyi Instrument Co., Ltd., Shanghai, China). In order to reduce experimental error, one blank and three same samples were tested at the same conditions in each experiment group. The removal ratios of MB or CIP solutions under different conditions can be calculated by Equation (1):
Removal ratio (%) = (C0 − Ct)/C0 × 100%
C0 is the initial concentration of MB or CIP solutions before treatment, and Ct is the concentration at reaction time t.

2.5. OH Measurement and Intermediate Product Determination

·OH can react with salicylic acid and produce a chemical substance called 2, 3-dihydroxybenzoic acid with maximum absorption peak at the wavelength of 510 nm. Thus, the ·OH concentration produced by CuO-Fe2O3/CTS-ATP/H2O2 system can be measured by a UV1900 type ultraviolet-visible spectrophotometer with ·OH spectrophotometry at λ of 510 nm.
The intermediate products of MB and CIP degraded by CuO-Fe2O3/CTS-ATP/H2O2 system were determined by high performance liquid chromatography-mass spectrometry (HPLC-MS). Chromatography and mass spectrometry were performed by Agilent 1200 high performance liquid chromatography (Agilent, Santa Clara, CA, USA) and a 6130 quadruple mass spectrometer (Agilent, Santa Clara, CA, USA), respectively. The separation column is ZORBAX Eclipse Plus C18 (Analytical 4.6 mm × 150 mm 5-Micron). The operating conditions for the analysis of MB and CIP intermediate products were as follows: the mobile phase A was 0.2% formic acid high pure water, mobile phase B was acetonitrile, the temperature of separation column was 30 °C, the injection volume was 10 μL, the Electron Spray Ionization (ESI) in negative ion mode was adopted as the mass spectrometer detector, the capillary current was 5 nA, the flow rate of dry gas was 10 L/min, and the dry gas temperature was 350 °C.

3. Results and Discussion

3.1. Characterization of Samples

To investigate the crystalline structure and stability of ATP, Fe2O3/CTS-ATP and CuO-Fe2O3/CTS-ATP composites and powder X-ray diffraction patterns were recorded by an X-ray diffractometer. It can be seen from Figure 1a that all samples showed diffraction peaks at 2θ of 8.5°, 19.5°, 27.8°, 34.7°, and 35.4°, corresponding to the diffraction peaks of attapulgite crystal structure, indicating that ATP’s crystal structure was not damaged when CTS and metal oxides were introduced into ATP while the disorder degree of ATP crystal increased due to the changes of basal peaks. The diffraction peaks at 2θ of 24.2°, 33.2°, 40.9°, 49.6°, and 54.2° belong to Fe2O3 [41], and the diffraction peaks of 35.6°, 38.7°, 49.1°, 53.5°, 58.3°, and 65.9° are characteristic peaks of CuO [41,42]. These indicate that the metal oxides are successfully loaded on the surface of CTS-ATP. The infrared spectrums of the samples were shown in Figure 1b. The strong absorption peak near 3430 cm−1 is a O–H bending vibration peak of attapulgite. The peak near 1630 cm−1 is H–O–H stretching vibration peak of water. The peak around 1482 cm−1 is C=O stretching vibration peak. The bending vibration peak of Si–O appears at 1040 cm−1. The wave number peaks at 740 cm−1 and 476 cm−1 are stretching vibration peaks of metal oxides Cu–O bond and Fe–O bond on the surface of Fe2O3/CTS-ATP and CuO-Fe2O3/CTS-ATP composites, indicating that both iron and copper are successfully loaded on the surface of carrier in the form of oxides. The peaks of Fe-O and Cu-O may not be too obvious due to low loading percentage.
To compare the structure and morphology characteristics of ATP, CTS-ATP, Fe2O3/CTS-ATP, and CuO-Fe2O3/CTS-ATP, scanning electron microscopy (SEM), Mapping, and EDS were used to observe and analyze the element content and distribution on the sample surface. The results are shown in Figure 2. Figure 2a–d are SEM images of the samples, which showed that ATP has rod crystal structure (a), and ATP was wrapped by CTS when ATP mixed with CTS (b), and further loading Fe2O3 or Fe2O3/CuO and calcined to form active sites and holes (c,d). The purpose of introducing CTS into ATP is to regulate the surface morphology of the catalyst and make the loaded active materials more uniform on support, which can be proved by the Mapping results of Figure 2e,f, showing that the loaded Fe or Fe/Cu on the surfaces of the catalysts are relatively uniform. Figure 2g–I are DES results of the samples, and, from Figure 2h,I, it is also clearly shown that Fe or Fe/Cu were successfully loaded on the catalysts.
Table 1 shows the specific surface area, pore volume, and pore size of ATP, Fe2O3/CTS-ATP and CuO-Fe2O3/CTS-ATP. The BET specific surface area and pore volume of Fe2O3/CTS-ATP (21.977 m2/g and 0.061 cm3/g) and CuO-Fe2O3/CTS-ATP (23.23 m2/g and 0.095 cm3/g) decreased dramatically compared to those of ATP (93.620 m2/g and 21.509 cm3/g) due to the wrapping of CTS on ATP, while the pore size of Fe2O3/CTS-ATP (11.013 nm) and CuO-Fe2O3/CTS-ATP (14.271 nm) increased greatly compared to that of ATP (0.227 nm), which meant that mesopores are formed on the surface of the catalyst during the roasting process [43].
Figure 3 is nitrogen adsorption and desorption (A-D) curves as well as pore size distribution diagrams of Fe2O3/CTS-ATP and CuO-Fe2O3/CTS-ATP. The A-D curves of the two samples belong to type III isotherms and H3 magnetic loop according to IUPAC classification, representing their mesoporous structure, which is in accordance with the characteristic result of SEM images. The attached pore size distribution diagrams showed that the two samples had the same pore size distribution, while CuO-Fe2O3/CTS-ATP had more pores with the size ranging from 50–100 nm, and the pore structure belongs to the slit pore.
In order to analyze the chemical state of various elements of the samples, XPS tests of the catalysts were performed. The total spectra and characteristic spectra of the elements are shown in Figure 4. The survey spectra of the samples confirmed that Fe and Cu were present in these two samples besides C, O, and Si, while the peaks of Fe and Cu were very weak due to the low loading deposition level (Figure 4a). In Figure 4b, the peaks at 284.8, 286.2, and 286.6 eV of the samples were C1s bonded to H, N, and O, which might be formed during the roasting of the samples. In Figure 4c, the peaks at 531 eV of the samples were relevant to O1s bonded to Fe, Cu, and Si, whereas the peaks at higher binding energy of 532.2 eV were ascribed to the presence of OH group [44,45] on the catalysts’ surface. High resolution XPS spectrum of Fe2p in Figure 4d located at the binding energies of 713.2 (713.6) eV and 726.8 (726.4) eV were indicative of Fe2p3/2 and Fe2p1/2 of the oxidation state Fe3+ in samples [46]. The characteristic peaks of Cu2p3/2 and Cu2p1/2 at 933.1 and 952.4 in the spectrum of Cu2p in Figure 4e, accompanied by a series of satellite vibration peaks, indicated the presence of Cu2+ on the surface of the catalyst, which also agreed with the mapping results.

3.2. MB and CIP Degradation by Prepared Catalysts

Figure 5 shows the effects on the MB and CIP removal ratios in five different systems (only H2O2 existing, only Fe2O3/CTS-ATP or CuO-Fe2O3/CTS-ATP existing, Fe2O3/CTS-ATP/H2O2 system or CuO-Fe2O3/CTS-ATP/H2O2 system), respectively. From Figure 5a,b, it can be seen that MB can obtain higher removal ratios than CIP at the same conditions. This means that oxidation of MB is easier than that of CIP, and the structure of CIP is more stable than that of MB. When only hydrogen peroxide exists, MB and CIP can obtain about 40% removal ratios after 60 min reaction because H2O2 has a certain oxidation ability but not too much (Figure 5a,b). When only Fe2O3/CTS-ATP or CuO-Fe2O3/CTS-ATP exists in the system, MB and CIP removal ratios were 60% or 50% separately due to the adsorption of catalysts and O2 oxidation (Figure 5a,b). In the Fe2O3/CTS-ATP/H2O2 system, MB and CIP can obtain over 95% and 80% removal ratios after 60 min reaction because Fe2O3 catalyzed H2O2 to decompose ·OH, which had higher oxidation ability than O2 and H2O2, whether in acidic or alkaline conditions, as can be seen in Figure 5c. In the CuO-Fe2O3/CTS-ATP/H2O2 system, MB and CIP can obtain over 99% and 90% removal ratios after 60 min reaction, which were higher and faster than those in the Fe2O3/CTS-ATP/H2O2 system, which also can be seen from the catalytic reaction rate constant k (Figure 5a,b). The above results can be explained by Figure 5d. In the Fe2O3/CTS-ATP/H2O2 catalytic system, Fe(III) and Fe(II) formed a cycle during the process of catalyzing H2O2 to generate ·OH, while, in the CuO-Fe2O3/CTS-ATP/H2O2 catalytic system, Fe(III) and Fe(II), and Cu(I) and Cu(II), formed a dual catalytic cycle to catalyze more H2O2 and generate more ·OH, which led to higher degradation ratios and faster degradation rates.
A series of single-factor (such as reaction temperatures, pH values, catalyst dosages, H2O2 concentrations, recycle times of the catalysts, et al.) experiments were performed to optimize the degradation conditions of a CuO-Fe2O3/CTS-ATP/H2O2 system (Figures S3). From 40 to 80 °C, as shown in Figure S3, MB, and CIP, can obtain high removal ratios, and, with the temperature increasing, their removal ratios were increasing due to the boosting of ·OH radicals at a higher temperature and a faster catalysis to decompose H2O2. As can be seen from Figure S4, MB and CIP obtained the highest removal ratios at pH values of 2 attributed to a high redox potential of ·OH in an acidic condition. In addition, they all reached over 60% removal ratios under all experimental pH values. Figures S5 and S6 indicated that degradation ratios of MB and CIP ascended with increasing catalyst dosage and H2O2 concentration, suggesting that the more catalysts and H2O2 there are, the more active sites OH radicals there are, resulting in high removal efficiencies. However, too many catalysts and H2O2 will lead to high costs, and too many H2O2 will react with ·OH radicals and release oxygen. Recycle degradation experiments of MB and CIP by CuO-Fe2O3/CTS-ATP catalysts were performed five times to test their durability (Figure S7). The catalyst was only washed by distilled water after every run, and the reaction conditions maintained the same in every run. After five runs, MB and CIP also can obtain 98% and 86% degradation ratios, separately.

3.3. OH Concentrations Measurement

·OH concentrations in two catalytic systems were also detected, and the results were shown in Figure 6. It can be seen that the concentration of ·OH in CuO-Fe2O3/CTS-ATP/H2O2 system was more than that in the Fe2O3/CTS-ATP/H2O2 system, and the production rate of ·OH in former system was also higher than that in the latter. The results confirmed that these two systems could produce enough ·OH to degrade ROS, and CuO could be a good assistant catalytic component to enhance the ·OH production rate dramatically, which was coincident with the degradation mechanisms in Figure 5d.

3.4. The Pathways of MB and CIP Degradation by CuO-Fe2O3/CTS-ATP

As can be seen in Scheme 1a, attapulgite and chitosan combined via hydrogen bond; thus, a core–shell structure with attapulgite as the core and chitosan as the shell was formed. After calcination, chitosan was carbonized and formed a net structure to wrap attapulgite. When loading active components, iron oxide and copper oxide first formed an alloy [47,48,49] and then combined with the carrier (CTS-ATP) in the form of a covalent bond, in order to obtain an efficient surface catalytic system.
In order to explore the intermediate products of MB and CIP in CuO-Fe2O3/CTS-ATP/H2O2, HPLC-MS was used to analyze the degradation products of MB and CIP at reaction times of 30 and 60 min, respectively. The treated-water samples were filtered and then sent into the separation column through the autosampler. When MB solution was catalyzed for 30 min, the signal peak of MB (m/z = 284.19) at 14.6 min disappeared, and many small mass fragments appeared. The possible degradation path of MB molecule is shown in Scheme 1b, where the substance with m/z = 242.82 is the product after MB demethylation under the attacking of ·OH [50]; then, the –S– bond and –N= were broken, and MB was broken into two pieces. After a series of degradations, the main identified products were benzoquinone (m/z = 107, RT = 3.161 min), hydroquinone (m/z = 109, RT = 3.098 min), catechol (m/z = 109, RT = 3.308 min), and resorcinol (m/z = 109, RT 5.098 min).
The characteristic signal peak of the CIP was m/z = 332.20 at 2.9 min, and it disappeared completely after the 30 min degradation of CIP. According to the identity results of CIP intermediate products by HPLC-MS, the possible degradation path is shown in Scheme 1c. Under the attacking of ·OH, CIP lost its carboxyl (–COOH) firstly, and the piperazine ring in CIP molecule underwent oxidation and ring opening; then, it was totally removed. After losing –C3H5N [51], another nitrogen-containing heterocycle of CIP was also opened. The oxidation process continued, and the carbon chain containing aldehyde and carboxylic acid on the benzene ring was broken to form fluorophenol. Finally, the benzene ring was opened.

4. Conclusions

In this study, a heterogeneous catalyst with dual support components (CTS-ATP) and dual active components (Fe2O3-CuO) was prepared for the degradation of MB and CIP. The catalysts were characterized by XRD, FI-IR, SEM, BET, and XPS, which showed that both Fe2O3 and CuO were successfully loaded on the surface of the carrier, and CTS as a functional material wrapped ATP would make the loaded active components more uniform, which greatly improves the catalytic degradation performance of CuO-Fe2O3/CTS-ATP. The preparation conditions and degradation conditions were optimized. Under the optimal preparation and reaction conditions, the degradation ratios of MB and CIP can reach 99.29% and 86.20%, respectively. The new degradation mechanisms and possible pathways of MB and CIP were proposed. The authors believed that Fe2O3 and CuO first formed an alloy and then combined with the carrier (CTS-ATP) in the form of covalent bonds, in order to obtain an efficient surface catalytic system. Fe(III) and Fe(II) and Cu(I) and Cu(II) formed a dual catalytic cycle to catalyze more H2O2 and generate more ·OH, which led to higher degradation ratios and faster degradation rates.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/coatings12050559/s1, Figure S1: Influences of precursor concentration (during catalyst preparation) on MB (a) and CIP (b) removal.; Figure S2: Influences of ultrasonic impregnation time (during catalyst preparation) on MB (a) and CIP (b) removal.; Figure S3: Influences of reaction temperature on MB (a) and CIP (b) removal.; Figure S4: Influences of pH values on MB (a) and CIP (b) removal.; Figure S5: Influences of catalyst dosages on MB (a) and CIP (b) removal.; Figure S6: Influences of H2O2 concentrations on MB (a) and CIP (b) removal.; Figure S7: Influences of recycle times of prepared catalyst on MB (a) and CIP (b) removal.

Author Contributions

Conceptualization, T.Z. and Q.G.; Data curation, W.L. and Q.G.; Formal analysis, Q.G.; Funding acquisition, T.Z.; Investigation, W.L.; Methodology, T.Z. and Y.W.; Project administration, C.L.; Resources, Y.W. and C.L.; Supervision, C.L.; Validation, Y.W.; Writing—original draft, T.Z. and Q.G.; Writing—review & editing, T.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by Major Science and Technology Projects in Gansu Province of China (Grant No. 21ZD2JA001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or Supplementary Material.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ATPattapulgite
CTSchitosan
ROSrefractory organics
MBmethylene blue
CIPciprofloxacin

References

  1. Sun, Y.; Yu, I.K.; Tsang, D.C.; Fan, J.; Clark, J.H.; Luo, G.; Zhang, S.; Khan, E.; Graham, N.J. Tailored design of graphitic biochar for high-efficiency and chemical-free microwave-assisted removal of refractory organic contaminants. Chem. Eng. J. 2020, 398, 125505. [Google Scholar] [CrossRef]
  2. Ren, H.; Jin, X.; Li, C.; Li, T.; Liu, Y.; Zhou, R. Rosmarinic acid enhanced Fe(III)-mediated Fenton oxidation removal of organic pollutants at near neutral pH. Sci. Total Environ. 2020, 736, 139528. [Google Scholar] [CrossRef] [PubMed]
  3. Chen, M.; Zhang, Z.; Zhu, L.; Wang, N.; Tang, H. Bisulfite-induced drastic enhancement of bisphenol A degradation in Fe3+-H2O2 Fenton system. Chem. Eng. J. 2018, 361, 1190–1197. [Google Scholar] [CrossRef]
  4. Zhang, J.; Li, L.; Zheng, J.; Yang, P.; Wu, X.; Cheng, C.; Li, J.; Tian, Y.; Wang, F. Improved organic pollutants removal and simultaneous electricity production via integrating Fenton process and dual rotating disk photocatalytic fuel cell system using bamboo charcoal cathode. Chem. Eng. J. 2019, 361, 1198–1206. [Google Scholar] [CrossRef]
  5. Kumar, N.; Mittal, H.; Parashar, V.; Ray, S.S.; Ngila, J.C. Efficient removal of rhodamine 6G dye from aqueous solution using nickel sulphide incorporated polyacrylamide grafted gum karaya bionanocomposite hydrogel. RSC Adv. 2016, 6, 21929–21939. [Google Scholar] [CrossRef]
  6. Zhang, G.-F.; Liu, X.; Zhang, S.; Pan, B.; Liu, M.-L. Ciprofloxacin derivatives and their antibacterial activities. Eur. J. Med. Chem. 2018, 146, 599–612. [Google Scholar] [CrossRef]
  7. Hassani, A.; Khataee, A.; Karaca, S.; Karaca, C.; Gholami, P. Sonocatalytic degradation of ciprofloxacin using synthesized TiO2 nanoparticles on montmorillonite. Ultrason. Sonochem. 2017, 35, 251–262. [Google Scholar] [CrossRef]
  8. Salma, A.; Thoröe-Boveleth, S.; Schmidt, T.C.; Tuerk, J. Dependence of transformation product formation on pH during photolytic and photocatalytic degradation of ciprofloxacin. J. Hazard. Mater. 2016, 313, 49–59. [Google Scholar] [CrossRef]
  9. Xiong, J.-Q.; Kurade, M.B.; Kim, J.R.; Roh, H.-S.; Jeon, B.-H. Ciprofloxacin toxicity and its co-metabolic removal by a freshwater microalga Chlamydomonas mexicana. J. Hazard. Mater. 2017, 323, 212–219. [Google Scholar] [CrossRef]
  10. Luo, K.; Yang, Q.; Pang, Y.; Wang, D.; Li, X.; Lei, M.; Huang, Q. Unveiling the mechanism of biochar-activated hydrogen peroxide on the degradation of ciprofloxacin. Chem. Eng. J. 2019, 374, 520–530. [Google Scholar] [CrossRef]
  11. Li, S.; Zhang, X.; Huang, Y. Zeolitic imidazolate framework-8 derived nanoporous carbon as an effective and recyclable adsorbent for removal of ciprofloxacin antibiotics from water. J. Hazard. Mater. 2017, 321, 711–719. [Google Scholar] [CrossRef]
  12. Sun, Y.; Yue, Q.; Gao, B.; Gao, Y.; Xu, X.; Li, Q.; Wang, Y. Adsorption and cosorption of ciprofloxacin and Ni(II) on activated carbon-mechanism study. J. Taiwan Inst. Chem. Eng. 2014, 45, 681–688. [Google Scholar] [CrossRef]
  13. Kanakaraju, D.; Glass, B.D.; Oelgemoeller, M. Advanced oxidation process-mediated removal of pharmaceuticals from water: A review. J. Environ. Manag. 2018, 219, 189–207. [Google Scholar] [CrossRef]
  14. Zhang, Y.; Zhuang, Y.; Geng, J.; Ren, H.; Xu, K.; Ding, L. Reduction of antibiotic resistance genes in municipal wastewater effluent by advanced oxidation processes. Sci. Total Environ. 2016, 550, 184–191. [Google Scholar] [CrossRef]
  15. Anjali, R.; Shanthakumar, S. Insights on the current status of occurrence and removal of antibiotics in wastewater by advanced oxidation processes. J. Environ. Manag. 2019, 246, 51–62. [Google Scholar] [CrossRef]
  16. Sharma, A.; Ahmad, J.; Flora, S. Application of advanced oxidation processes and toxicity assessment of transformation products. Environ. Res. 2018, 167, 223–233. [Google Scholar] [CrossRef]
  17. Liang, D.; Li, N.; An, J.; Ma, J.; Wu, Y.; Liu, H. Fenton-based technologies as efficient advanced oxidation processes for microcystin-LR degradation. Sci. Total Environ. 2020, 753, 141809. [Google Scholar] [CrossRef]
  18. Oluwole, A.O.; Omotola, E.O.; Olatunji, O.S. Pharmaceuticals and personal care products in water and wastewater: A review of treatment processes and use of photocatalyst immobilized on functionalized carbon in AOP degradation. BMC Chem. 2020, 14, 32. [Google Scholar] [CrossRef]
  19. Chaturvedi, N.K.; Katoch, S.S. Evaluation and comparison of Fenton-like oxidation with Fenton’s oxidation for hazardous methoxyanilines in aqueous solution. J. Ind. Eng. Chem. 2020, 92, 101–108. [Google Scholar] [CrossRef]
  20. Casado, C.; Moreno-SanSegundo, J.; De la Obra, I.; García, B.E.; Pérez, J.A.S.; Marugán, J. Mechanistic modelling of wastewater disinfection by the photo-Fenton process at circumneutral pH. Chem. Eng. J. 2020, 403, 126335. [Google Scholar] [CrossRef]
  21. Prakash, K.; Kumar, J.V.; Latha, P.; Kumar, P.S.; Karuthapandian, S. Fruitful fabrication of CDs on GO/g-C3N4 sheets layers: A carbon amalgamation for the remediation of carcinogenic pollutants. J. Photochem. Photobiol. A Chem. 2018, 370, 94–104. [Google Scholar] [CrossRef]
  22. Wu, Q.; Siddique, M.S.; Yu, W. Iron-nickel bimetallic metal-organic frameworks as bifunctional Fenton-like catalysts for enhanced adsorption and degradation of organic contaminants under visible light: Kinetics and mechanistic studies. J. Hazard. Mater. 2020, 401, 123261. [Google Scholar] [CrossRef]
  23. Xue, C.; Peng, Y.; Chen, A.; Peng, L.; Luo, S. Drastically inhibited nZVI-Fenton oxidation of organic pollutants by cysteine: Multiple roles in the nZVI/O2/hv system. J. Colloid Interf. Sci. 2020, 582, 22–29. [Google Scholar] [CrossRef]
  24. Gao, C.; Chen, S.; Quan, X.; Yu, H.; Zhang, Y. Enhanced Fenton-like catalysis by iron-based metal organic frameworks for degradation of organic pollutants. J. Catal. 2017, 356, 125–132. [Google Scholar] [CrossRef]
  25. Guo, H.; Li, Z.; Zhang, Y.; Jiang, N.; Wang, H.; Li, J. Degradation of chloramphenicol by pulsed discharge plasma with heterogeneous Fenton process using Fe3O4 nanocomposites. Sep. Purif. Technol. 2020, 253, 117540. [Google Scholar] [CrossRef]
  26. Yang, J.; Zeng, D.; Zhang, Q.; Cui, R.; Hassan, M.; Dong, L.; Li, J.; He, Y. Single Mn atom anchored on N-doped porous carbon as highly efficient Fenton-like catalyst for the degradation of organic contaminants. Appl. Catal. B Environ. 2020, 279, 119363. [Google Scholar] [CrossRef]
  27. Jia, C.; Mi, Y.; Liu, Z.; Zhou, W.; Gao, H.; Zhang, S.; Lu, R. Attapulgite modified with covalent organic frameworks as the sorbent in dispersive solid phase extraction for the determination of pyrethroids in environmental water samples. Microchem. J. 2019, 153, 104522. [Google Scholar] [CrossRef]
  28. Yao, D.; Shi, Y.; Pan, H.; Zhong, D.; Hou, H.; Wu, X.; Chen, J.; Wang, L.; Hu, Y.; Crittenden, J.C. Promotion mechanism of natural clay colloids in the adsorption of arsenite on iron oxide particles in water. Chem. Eng. J. 2019, 392, 123637. [Google Scholar] [CrossRef]
  29. Chen, Y.; Lin, Z.; Hao, R.; Xu, H.; Huang, C. Rapid adsorption and reductive degradation of Naphthol Green B from aqueous solution by Polypyrrole/Attapulgite composites supported nanoscale zero-valent iron. J. Hazard. Mater. 2019, 371, 8–17. [Google Scholar] [CrossRef]
  30. Guo, H.; Zhang, H.; Peng, F.; Yang, H.; Xiong, L.; Wang, C.; Huang, C.; Chen, X.; Ma, L. Effects of Cu/Fe ratio on structure and performance of attapulgite supported CuFeCo-based catalyst for mixed alcohols synthesis from syngas. Appl. Catal. A Gen. 2015, 503, 51–61. [Google Scholar] [CrossRef]
  31. Han, D.; Zhao, H.; Gao, L.; Qin, Z.; Ma, J.; Han, Y.; Jiao, T. Preparation of carboxymethyl chitosan/phytic acid composite hydrogels for rapid dye adsorption in wastewater treatment. Colloids Surf. A Physicochem. Eng. Asp. 2021, 628, 127355. [Google Scholar] [CrossRef]
  32. Iber, B.T.; Okomoda, V.T.; Rozaimah, S.A.; Kasan, N.A. Eco-friendly approaches to aquaculture wastewater treatment: Assessment of natural coagulants vis-a-vis chitosan. Bioresour. Technol. Rep. 2021, 15, 100702. [Google Scholar] [CrossRef]
  33. Vedula, S.S.; Yadav, G.D. Wastewater treatment containing methylene blue dye as pollutant using adsorption by chitosan lignin membrane: Development of membrane, characterization and kinetics of adsorption. J. Indian Chem. Soc. 2021, 99, 100263. [Google Scholar] [CrossRef]
  34. Sun, N.; Zhang, Y.; Ma, L.; Yu, S.; Li, J. Preparation and characterization of chitosan/purified attapulgite composite for sharp adsorption of humic acid from aqueous solution at low temperature. J. Taiwan Inst. Chem. Eng. 2017, 78, 96–103. [Google Scholar] [CrossRef]
  35. Liang, X.X.; Ouyanga, X.K.; Wanga, S.; Yanga, L.Y.; Huanga, F.; Jia, C.; Chenb, X. Efficient adsorption of Pb(II) from aqueous solutions using aminopropyltriethoxysilane-modified magnetic attapulgite@chitosan (APTS-Fe3O4/APT@CS) composite hydrogel beads. Int. J. Biol. Macromol. 2019, 137, 741–750. [Google Scholar] [CrossRef]
  36. Pan, D.; Fan, Q.; Fan, F.; Tang, Y.; Zhang, Y.; Wu, W. Removal of uranium contaminant from aqueous solution by chitosan@attapulgite composite. Sep. Purif. Technol. 2017, 177, 86–93. [Google Scholar] [CrossRef]
  37. Liao, Y.; Wang, M.; Chen, D. Production of three-dimensional porous polydopamine-functionalized attapulgite/chitosan aerogel for uranium(VI) adsorption. J. Radioanal. Nucl. Chem. Artic. 2018, 316, 635–647. [Google Scholar] [CrossRef]
  38. Shi, Y.; Zhang, Q.; Feng, L.; Xiong, Q.; Chen, J. Preparation and adsorption characters of Cu(II)-imprinted chitosan/attapulgite polymer. Korean J. Chem. Eng. 2014, 31, 821–827. [Google Scholar] [CrossRef]
  39. Krishnan, A.; Vishwanathan, P.V.; Mohan, A.C.; Panchami, R.; Viswanath, S.; Krishnan, A.V. Tuning of photocatalytic performance of CeO2-Fe2O3 composite by Sn-doping for the effective degradation of methlene blue (MB) and methyl orange (MO) dyes. Surf. Interfaces 2020, 22, 100808. [Google Scholar] [CrossRef]
  40. Li, Y.; Wang, Z.; Xie, X.; Zhu, J.; Li, R.; Qin, T. Removal of Norfloxacin from aqueous solution by clay-biochar composite prepared from potato stem and natural attapulgite. Colloids Surf. A Physicochem. Eng. Asp. 2017, 514, 126–136. [Google Scholar] [CrossRef]
  41. Zhang, T.; Dong, L.; Du, J.; Qian, C.; Wang, Y. CuO and CeO2 assisted Fe2O3/attapulgite catalyst for heterogeneous Fenton-like oxidation of methylene blue. RSC Adv. 2020, 10, 23431–23439. [Google Scholar] [CrossRef]
  42. Lei, Y.; Luo, Y.; Li, X.; Lu, J.; Mei, Z.; Peng, W.; Chen, R.; Chen, K.; Chen, D.; He, D. The role of samarium on Cu/Al2O3 catalyst in the methanol steam reforming for hydrogen production. Catal. Today 2018, 307, 162–168. [Google Scholar] [CrossRef]
  43. Wang, Y.; Zhou, R.; Wang, C.; Zhou, G.; Hua, C.; Cao, Y.; Song, Z. Novel environmental-friendly nano-composite magnetic attapulgite functionalized by chitosan and EDTA for cadmium (II) removal. J. Alloys Compd. 2019, 817, 153286. [Google Scholar] [CrossRef]
  44. Ismail, I.; Abdallah, B.; Abou-Kharroub, M.; Mrad, O. XPS and RBS investigation of TiNxOy films prepared by vacuum arc discharge. Nucl. Instrum. Methods Phys. Res. B Beam Interact. Mater. At. 2012, 271, 102–106. [Google Scholar] [CrossRef]
  45. Guo, H.-J.; Li, Q.-L.; Zhang, H.-R.; Xiong, L.; Peng, F.; Yao, S.-M.; Chen, X.-D. Attapulgite supported Cu-Fe-Co based catalyst combination system for CO hydrogenation to lower alcohols. J. Fuel Chem. Technol. 2019, 47, 1346–1356. [Google Scholar] [CrossRef]
  46. Qu, J.; Che, T.; Shi, L.; Lu, Q.; Qi, S. A novel magnetic silica supported spinel ferrites NiFe2O4 catalyst for heterogeneous Fenton-like oxidation of rhodamine B. Chin. Chem. Lett. 2019, 30, 1198–1203. [Google Scholar] [CrossRef]
  47. Qin, X.; Wang, Z.; Guo, C.; Guo, R.; Lv, Y.; Li, M. Fulvic acid degradation in Fenton-like system with bimetallic magnetic carbon aerogel Cu-Fe@CS as catalyst: Response surface optimization, kinetic and mechanism. J. Environ. Manag. 2022, 306, 114500. [Google Scholar] [CrossRef]
  48. Xia, Q.; Zhang, D.; Yao, Z.; Jiang, Z. Revealing the enhancing mechanisms of Fe–Cu bimetallic catalysts for the Fenton-like degradation of phenol. Chemosphere 2022, 289, 133195. [Google Scholar] [CrossRef]
  49. Qi, H.; Pan, G.; Shi, X.; Sun, Z. Cu–Fe–FeC3@nitrogen-doped biochar microsphere catalyst derived from CuFe2O4@chitosan for the efficient removal of amoxicillin through the heterogeneous electro-Fenton process. Chem. Eng. J. 2022, 434, 134675. [Google Scholar] [CrossRef]
  50. Moztahida, M.; Lee, D.S. Photocatalytic degradation of methylene blue with P25/graphene/polyacrylamide hydrogels: Optimization using response surface methodology. J. Hazard. Mater. 2020, 400, 123314. [Google Scholar] [CrossRef]
  51. Zhang, T.; Qian, C.; Guo, P.; Gan, S.; Dong, L.; Bai, G.; Guo, Q. Guo A Novel Reduced Graphene Oxide-Attapulgite (RGO-ATP) Supported Fe2O3 Catalyst for Heterogeneous Fenton-like Oxidation of Ciprofloxacin: Degradation Mechanism and Pathway. Catalysts 2020, 10, 189. [Google Scholar] [CrossRef] [Green Version]
Figure 1. XRD (a) and FT-IR (b) patterns of the samples.
Figure 1. XRD (a) and FT-IR (b) patterns of the samples.
Coatings 12 00559 g001
Figure 2. SEM (ad); Mapping (e,f); and EDS (gi) images of the samples.
Figure 2. SEM (ad); Mapping (e,f); and EDS (gi) images of the samples.
Coatings 12 00559 g002aCoatings 12 00559 g002b
Figure 3. Nitrogen adsorption–desorption curves and pore size distribution diagrams of Fe2O3/CTS-ATP and CuO-Fe2O3/CTS-ATP.
Figure 3. Nitrogen adsorption–desorption curves and pore size distribution diagrams of Fe2O3/CTS-ATP and CuO-Fe2O3/CTS-ATP.
Coatings 12 00559 g003
Figure 4. XPS full spectra (a), C1s spectra (b), O2p spectra (c), Fe2p spectra (d), and Cu2p spectra (e) of prepared catalysts.
Figure 4. XPS full spectra (a), C1s spectra (b), O2p spectra (c), Fe2p spectra (d), and Cu2p spectra (e) of prepared catalysts.
Coatings 12 00559 g004aCoatings 12 00559 g004b
Figure 5. The effects on the MB and CIP removal ratios in different systems (a,b) and degradation mechanisms (c,d).
Figure 5. The effects on the MB and CIP removal ratios in different systems (a,b) and degradation mechanisms (c,d).
Coatings 12 00559 g005
Figure 6. The trend of ·OH concentration in two systems with time.
Figure 6. The trend of ·OH concentration in two systems with time.
Coatings 12 00559 g006
Scheme 1. Formation of catalytic system (a) and possible degradation pathways of MB (b) and CIP (c) molecules by the CuO-Fe2O3/CTS-ATP/H2O2 system.
Scheme 1. Formation of catalytic system (a) and possible degradation pathways of MB (b) and CIP (c) molecules by the CuO-Fe2O3/CTS-ATP/H2O2 system.
Coatings 12 00559 sch001
Table 1. BET measurement results of samples.
Table 1. BET measurement results of samples.
SamplesBET Surface Area/m2·g−1Pore Volume/cm3·g−1Pore Size/nm
ATP93.62021.5090.227
Fe2O3/CTS-ATP21.9970.06111.013
CuO-Fe2O3/CTS-ATP23.230.09514.271
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, T.; Li, W.; Guo, Q.; Wang, Y.; Li, C. Preparation of a Heterogeneous Catalyst CuO-Fe2O3/CTS-ATP and Degradation of Methylene Blue and Ciprofloxacin. Coatings 2022, 12, 559. https://doi.org/10.3390/coatings12050559

AMA Style

Zhang T, Li W, Guo Q, Wang Y, Li C. Preparation of a Heterogeneous Catalyst CuO-Fe2O3/CTS-ATP and Degradation of Methylene Blue and Ciprofloxacin. Coatings. 2022; 12(5):559. https://doi.org/10.3390/coatings12050559

Chicago/Turabian Style

Zhang, Ting, Wenhui Li, Qiyang Guo, Yi Wang, and Chunlei Li. 2022. "Preparation of a Heterogeneous Catalyst CuO-Fe2O3/CTS-ATP and Degradation of Methylene Blue and Ciprofloxacin" Coatings 12, no. 5: 559. https://doi.org/10.3390/coatings12050559

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