Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Chemosphere 303 (2022) 135123

Contents lists available at ScienceDirect

Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere

Controlled carbonization of microplastics loaded nano zero-valent iron for


catalytic degradation of tetracycline
Ruirui Sun, Jiapeng Yang, Rong Huang, Chongqing Wang *
School of Chemical Engineering, Zhengzhou University, Zhengzhou, 450001, China

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Carbon substrate was fabricated by


controlled carbonization of
microplastic.
• Nano zero-valent iron loaded porous
carbon was used as heterogeneous Fen­
ton catalyst.
• Effectively degrading tetracycline was
proved by high rate constant and low
activation energy.
• Degradation of POPs including Rhoda­
mine B, p-nitrophenol, and butylxan­
thate was verified.
• Catalytic mechanism was revealed
associated with multiple characteriza­
tions and analysis.

A R T I C L E I N F O A B S T R A C T

Handling Editor: Yasser Vasseghian Nano zero-valent iron loaded porous carbon derived from microplastics was designed as heterogeneous catalyst
for degradation of persistent organic pollutants. Controlled carbonization of microplastics with molten salt was
Keywords: conducted to tune the morphology of carbon product. Controlled carbonization induces higher carbon yield
Advanced oxidation processes (from 17.73% to 52.24%) and larger surface area (from 403.72 m2/g to 601.82 m2/g). The catalyst (Fe/MMPC)
Carbonization
was characterized by Raman, Fourier transform infrared spectroscopy, X-ray diffraction, X-ray photoelectron
Persistent organic pollutants
spectroscopy, and scanning electron microscope. Loading nano zero-valent iron onto porous carbon are verified
Microplastics
Nano zero-valent iron in the catalyst. The process factors including Fe/MMPC dosage, H2O2, pH, anions, and temperature were studied
to estimate the catalytic performance. Tetracycline degradation (81.8% within 10 min) is effectively obtained in
the Fe/MMPC and H2O2 system. The apparent rate constant is 0.1311–0.2999 min− 1 under different tempera­
ture, and the activation energy of catalytic process is 22 kJ/mol. Pollutants including rhodamine B, p-nitro­
phenol, and butylxanthate are efficiently degraded in the catalytic system. The predominant species of catalytic
reactions are hydroxyl radicals, which are mainly produced from H2O2 activation enhanced by zero-valent iron
in Fe/MMPC. This work offers an innovative strategy for microplastic management and wastewater treatment.

1. Introduction various environments and attract increasing global attention (Saeed


et al., 2020). The concept of microplastics was first proposed by
Microplastics, as emerging pollutants, have been widely detected in Thompson et al. (2004). Plastic fibers, particles, or films with a particle

* Corresponding author.
E-mail address: zilangwang@126.com (C. Wang).

https://doi.org/10.1016/j.chemosphere.2022.135123
Received 1 March 2022; Received in revised form 24 April 2022; Accepted 23 May 2022
Available online 25 May 2022
0045-6535/© 2022 Elsevier Ltd. All rights reserved.
R. Sun et al. Chemosphere 303 (2022) 135123

size of less than 5 mm are regarded as microplastics (González-Soto 2.1. Chemicals and materials
et al., 2019). Microplastics ranging micrometer to nanometer level are
invisible to the naked eye. Microplastics include primary microplastics Microplastics samples (plastic pieces less than 5 mm) were obtained
and secondary microplastics. The former refers to microscopical plastic from waste polyethylene terephthalate (PET) originated from used
particles produced for special applications. The latter represents plastic beverage bottles (as shown in Fig. S1). Potassium borohydride (NaBH4,
particles that are fragmented or smashed from macroscopic plastics 98%) and ferrous chloride tetrahydrate (FeCl2⋅4H2O, 99%) were ob­
through physical, chemical, or biological processes (Auta et al., 2017). tained from the Sinopharm Chemical Reagent (China). Calcein
The inherent polymers and additives in microplastics are toxic, and (C30H26N2O13; CA, > 90%), isopropyl alcohol (C3H8O; IPA, > 99%),
microplastics enriched pollutants due to large surface area increase the tert-butanol (C4H10O; TBA, > 99%), and o-phenylenediamine (C6H8N2;
threats. Microplastics can be easily taken in or swallowed by fish and OPD, 98%) were provided by the Aladdin Bio-Chem Technology
aquatic organisms, and eventually accumulated in the human body (China). Rhodamine B (C28H31ClN2O3; RhB, > 95%), p-nitrophenol
through the food chain (Schirinzi et al., 2017). Microplastic removal is (C6H5NO3; PNP, 98%), and butylxanthate (C5H11OS2; BXT, > 95%),
imperative from environment, and various physical techniques have were provided by the Tianjin Dehua Chemical Reagents (China).
been proposed to address the problem (Zhang et al., 2021). How to Tetracycline (C22H24N2O8; TC, 96%) was achieved from the Rhawn
dispose of the collected or separated microplastics becomes an emerging Chemicals (China). Basic information of RhB, PNP, BXT, and TC is
problem. Conversion of microplastics into value-added products is shown in Table S1. All chemicals are analytically pure.
promising for microplastic management.
Persistent organic pollutants (POPs) in aquatic environment result in 2.2. Catalyst synthesis
significant environmental pollution. Advanced oxidation processes
(AOPs) are regarded as promising technique for removing POPs due to Porous carbon derived from microplastics was fabricated using the
complete mineralization, high efficiency, and simple operation (Rana modified carbonization process reported by Zhang et al. (2019). Briefly,
et al., 2022). To overcome the drawbacks of conventional Fenton pro­ microplastic sample was mixed with molten salt (ZnCl2/NaCl mass ratio:
cess, heterogeneous AOPs receive more and more attention (Wang et al., 1:1), and then carbonized in a tubular resistance furnace. Carbonization
2021a). Heterogeneous AOPs are currently limited by the development was conducted at 280 ◦ C for 8 min and then 550 ◦ C for 10 min. Adding
of effective catalysts. Zero-valent iron (ZVI) is cheap, reactive, and NaCl to ZnCl2 reduces the melting temperature of ZnCl2 from 290 ◦ C to
environmentally benign, and it has been proven to be effect catalyst for ~250 ◦ C, which is close to the melting temperature of PET plastic
degradation of different POPs (Minella et al., 2019; Jiang et al., 2020). (Zhang et al., 2019). Heating the mixtures at 280 ◦ C induce the ho­
However, the oxidation and agglomeration of ZVI particles because of mogenous phase and uniform mixing of plastics and molten salt. The
high surface energy and magnetic interaction significantly limit the product was rinsed using 1 M HCl to remove the molten salt. After
potential application. Supporting ZVI onto porous carbonaceous mate­ washing with ultrapure water and dried at 60 ◦ C for 12 h, the product
rials is an effective strategy for improving the catalytic performance. ZVI was labeled as MMPC. Without the addition of molten salt, carbon
loaded on activated carbon cloth (Raji et al., 2021), biochar (Deng et al., product was fabricated using the same process for comparison, and it
2018), and graphene (Yang et al., 2019) has been proven for effective was labeled as MPC.
degradation of POPs. Due to high carbon content, microplastics are To improve the catalytic activity, loading ZVI onto the MMPC was
promising carbon source for fabricating carbon materials. Trans­ performed by liquid phase reduction method (Xiang et al., 2021). The
formation of microplastics into carbon-based catalyst realizes efficient synthesis procedure can be shown as Fig. S2. FeCl2⋅4H2O (0.355 g) was
management of separated microplastics, and also favors high-value dissolved in 100 mL ultrapure water in a three-necked flask (250 mL).
recycling. To the best of our knowledge, ZVI loaded on carbon support After adding MMPC (0.4 g) to the solution, degassing using N2 was
derived from microplastics is rarely reported. conducted for 30 min. 0.1 mM NaBH4 was added under N2 and stirred
It is hypothesized that ZVI loaded on carbon support derived from for 30 min for iron reduction. Vacuum filtration was performed to
microplastics can be catalyst for efficient degradation of POPs. separate the reaction product. After washing several times using ethanol
Controlled carbonization of microplastics was innovatively conducted to and ultrapure water and drying for 10 h in a vacuum oven at 60 ◦ C, the
tune the morphology of carbon product. ZVI supporting on porous car­ as-prepared composite was used as catalyst (Fe/MMPC).
bon was designed for as heterogeneous catalyst. Antibiotics receive
special attention due to potential threats, such as acute and chronic 2.3. Degradation experiments
inhibitory impacts on microbial activities and ecosystems (Liu et al.,
2020a). Tetracycline (TC), one of the representative antibiotics, was Degradation experiments were conducted in an Erlenmeyer flask
chosen as target pollutant. The aims of this work are (i) to fabricate with 50 mL solution. Fe/MMPC (40 mg/L) was put into TC solution (40
porous carbon by controlled carbonization of microplastics, (ii) to syn­ mg/L). The solution was oscillated in a thermostatic oscillator at 150
thesize novel carbon-based catalyst of ZVI supported by porous carbon, rpm. The following factors including H2O2, Fe/MMPC dosage, initial pH,
(iii) to investigate the physicochemical properties of catalyst, (iv) to anions, and temperature were studied by single factor experiment
assess the catalytic ability for pollutant degradation, and (v) to uncover (Wang et al., 2021c). At regular intervals, solution sample was taken out,
the reaction mechanism. centrifuged for 1 min, and then the concentration was determined
through UV–Vis spectrometer (Ashraf et al., 2019). The change of
2. Materials and methods UV–Vis spectra of different pollutants (RhB, PNP, BXT, and TC) were
recorded (Huang et al., 2022). IPA and TBA were used as radical scav­
The detailed information of chemical and materials used in this study engers for quenching experiments (Wang et al., 2021c). CA and BPY
are described in this section. Controlled carbonization of microplastic were employed as chelating agents to reveal the role of iron ions (Wang
samples is depicted to fabricate porous carbon support. ZVI loaded on et al., 2022).
carbon support is designed as catalyst for catalytic degradation of pol­
lutants. The degradation experiment to evaluate catalyst performance is 2.4. Characterizations
demonstrated. Multiple characterizations are conducted to examine the
characteristics of as-prepared catalyst. The morphology was examined using a FEI Quanta FEG 250 scanning
electron microscope (SEM, America), and the elemental distribution was
determined by an energy dispersive spectrometer (EDS). The functional
groups were characterized by a Thermo Scientific Nicolet iS10 Fourier

2
R. Sun et al. Chemosphere 303 (2022) 135123

3. Results and discussion

3.1. Controlled carbonization of microplastic

Catalyst is designed by loading ZVI onto carbon support derived from


microplastic. The large surface and porous structure of carbon support
are imperative, and controlled carbonization of microplastic is con­
ducted to improve the properties. Fig. 1a shows that microplastic sam­
ples are transformed into black carbon material after carbonization.
Carbon product derived from the mixture of microplastic and ZnCl2/
NaCl molten salt is very fluffy compared with that obtained from
microplastic alone. The yield of MPC and MMPC is 17.73% and 52.24%,
respectively. This suggests that the presence of ZnCl2/NaCl molten salt
promotes carbon fixation in carbonization process.
According to IUPAC classification, MPC shows the type I adsorption
isotherm (Fig. 1a), indicating the microporous structure. A sharp in­
crease in N2 adsorption is observed at low relative pressure, while
saturation adsorption appears when it reaches higher relative pressure.
The adsorption isotherm of MMPC conforms to type II isotherm, sug­
gesting the existence of mesoporous structure. As displayed in Table S2,
the surface area of MPC and MMPC is determined to be 403.72 m2/g and
601.82 m2/g, respectively. The controlled carbonization results in the
notable increase of the surface area. Additionally, the total pore volume
Fig. 1. (a) Pictures and N2 adsorption isotherms of MPC and MMPC, (b) is increased from 1.82 cm3/g to 3.77 cm3/g.
schematic diagram of controlled carbonization using ZnCl2/NaCl molten salt as The addition of NaCl reduces the melting temperature of ZnCl2 from
a porogen, and (c) mechanism of the controlled carbonization of PET 290 ◦ C to 250 ◦ C, which is close to the melting temperature of PET
microplastic. plastic (Zhang et al., 2019). First step of carbonization at 280 ◦ C results
in simultaneously melting the mixtures and uniformly mixing. With
transform infrared spectrometer (FTIR, America). The specific surface increasing temperature to 550 ◦ C, the formation of carbon frame occurs,
area was examined using a Micromeritics ASAP 2020 surface area while the ZnCl2/NaCl molten salt as porogens produces porous struc­
analyzer (America) through N2 adsorption/desorption. Rigaku Ultimate ture. The mechanism of control carbonization is shown in Fig. 1b.
IV X-ray diffraction spectrometer (XRD, Japan) was used for XRD Molten salt promotes the decarboxylation and dehydration of micro­
analysis. The surface elements were detected by a Thermo ESCALAB plastics to produce vinyl terminal chain fragments and aromatic rings
250XI X-ray photoelectron spectroscopy (XPS, America). Raman spec­ (Fig. 1c). These intermediates build the carbon framework and further
trum was recorded employing a HORIBA LabRAM HR Evolution Raman crosslink and cyclize (Yu et al., 2016). The remaining molten salt as a
spectrometer (Japan). A zeta potential analyzer (DELSA-440SX, Coulter, porogen to produce more pore structure.
America) was used to analyze surface charge of catalyst. UV–vis spectra
of different pollutants were recorded using a Shimadzu UV-2600 spec­
3.2. Characteristics of catalyst
troscopy (Japan). The electron spin resonance (ESR) spectrum was
measured employing a JES FA200 spectrometer (JEOL, Japan) to iden­
The characteristics of carbon products derived from microplastics are
tify the reactive radicals in the catalytic system. An inductively coupled
characterized by multiple techniques. Fig. 2 exhibits the morphology of
plasma mass spectrometry (PerkinElmer Avio 500, America) was
Fe/MMPC, showing a porous lamellar structure. The clusters of small
employed to examine the leached iron.
particles on the surface may be the loaded nano iron species. EDS
mapping shows the uniform distribution of carbon and iron, implying
the efficient loading iron species on the carbon surface. MMPC was used

Fig. 2. SEM images of Fe/MMPC: (a) × 50 k magnifications, (b) × 100 k magnifications, and (c) EDS elemental mapping of iron, oxygen, and carbon.

3
R. Sun et al. Chemosphere 303 (2022) 135123

Fig. 3. (a) XRD patterns, (b) FTIR spectra, (c) Raman spectra of MMPC and Fe/MMPC, and high-resolution XPS spectra of Fe/MMPC: (d) C 1s, (e) O 1s, and (f) Fe 2p.

as the support to load nano zero-valent iron as heterogeneous catalyst. Card #65–3107) (Sahadevan et al., 2021). The appearance of Fe3O4 may
The successfully loaded iron onto carbon surface potentially provides be resulted from partial oxidation of Fe0 in the fabrication process. Iron
active sites for pollutant degradation. loading onto MMPC is verified by XRD analysis.
Fig. S3 shows the II type adsorption isotherm of Fe/MMPC. The Fig. 3b shows the FTIR spectrum of Fe/MMPC and MMPC. A wide
surface area is 472.85 m2/g and the total pore volume is 3.45 cm3/g. peak of about 3320 cm− 1 is derived from –OH vibration in adsorbed
Compared to MMPC, the loading zero-valent iron slightly lower the water molecules (Wang and Wang, 2018). The peak located at 1610
surface area and pore volume (Table S2). This can be attributed to pore cm− 1 is resulted from C– – O vibration (Wang et al., 2016), while the
plugging by the occupation of pore structure of carbon surface by iron absorbance peak at 1340 cm− 1 is assigned to the C–H stretching of
particles. It is speculated that the active metal iron species are success­ benzene structure. The slight shift of the peak at 1340 cm− 1 after iron
fully supported on the carbon support. loading indicates the variation of benzene structure, agreeing with the
XRD patterns of MMPC and Fe/MMPC are shown in Fig. 3a. The carbon fixation in the carbonization process. The new peak at 1080
broad peak at around 19.2◦ in XRD spectrum of MMPC is attributed to cm− 1 suggests the appearance of C–O group after iron loading (Wang
carbon. As to Fe/MMPC, the peak at 2θ of 44.7◦ is attributed to the plane et al., 2018).
of (110) crystal Fe0 (JCPDS PDF Card #65–4899) (Luo et al., 2020). The Raman spectra of MMPC and Fe/MMPC are shown in Fig. 3c. The
peaks at 2θ of 35.5◦ and 56.8◦ may be ascribed to Fe3O4 (JCPDS PDF broad peaks appeared at approximate 1350 cm− 1 and 1600 cm− 1 are

4
R. Sun et al. Chemosphere 303 (2022) 135123

Fig. 4. Effect of (a) catalyst dosage, (b) H2O2 concentration, (c) pH level, (d) anions, (e) temperature on TC degradation, and (f) kinetic fitting. (Experimental
conditions: H2O2 2.0 mM, Fe/MMPC 40 mg/L, TC 40 mg/L, pH level 4.0, and temperature 25 ◦ C).

assigned to D band and G band of carbon materials (Wang et al., 2020). oxygen, and lattice oxygen (Wei et al., 2019). Fig. 3f displays the fitted
D band represents the disorganized sp3 carbon, while G band reveals the peaks of iron. The presence of ferrous iron (Fe2+) is proved by the peaks
hybrid sp2 carbon in the graphite structure. The defect degree of carbon at binding energy of 711.0 and 724.7 eV, and the fitted peaks at 712.9
materials can be characterized by the D band and G band ratio (ID/IG) and 726.8 eV are attributed to ferric iron (Fe3+) (Wang et al., 2022). The
(Dissanayake et al., 2020). The calculated ID/IG value of MMPC and minor peak at 707.1 eV is ascribed to Fe0. XPS analysis is consistent with
Fe/MMPC is 0.81 and 0.85, suggesting that iron loading slightly in­ XRD results, proving the existence of Fe0 and Fe3O4 in Fe/MMPC.
creases defect degree. The defective carbon may be favorable for cata­
lytic reaction (Wang et al., 2021b).
3.3. TC degradation in different systems
Fig. 3d displays the signals of C 1s XPS peak. The fitted peaks at
284.8 eV, 286.6 eV, and 289.8 eV reveal the C–C/C– – C, C–O, and C– –O
The catalytic ability of Fe/MMPC is evaluated associated with H2O2
groups, respectively (Sun et al., 2021). This suggests the existence of
for degradation of target TC. Fig. S4 shows the comparison of TC
oxygen-containing groups on MMPC surface, which is consistent with
removal under Fe/MMPC alone, H2O2 alone, or Fe/MMPC and H2O2
FTIR analysis. The O 1s spectra manifest the fitted peaks at 532.9, 531.6,
system. The effect of Fe/MMPC or H2O2 alone on TC removal is negli­
and 530.2 eV (Fig. 3e), assigning to the hydroxyl oxygen, surface
gible, while significant decline of TC is seen in the presence of Fe/MMPC

5
R. Sun et al. Chemosphere 303 (2022) 135123

Fig. 5. Degradation of (a) TC, (b) PNP, (c) BXT, and (d) RhB in the catalytic system. (Experimental conditions: Fe/MMPC 40 mg/L, H2O2 2.0 mM, and tempera­
ture 25 ◦ C).

and H2O2. The H2O2 oxidation or adsorption effect by Fe/MMPC results


in minor TC removal, while Fe/MMPC effectively activates H2O2 for TC 2H2O2 → O2 + 2H2O (1)
degradation (81.8% within 10 min). Fig. S5 shows the zeta potential of
Fe/MMPC as a function of pH. The point of zero charge (pHpzc) of Fe/ H2O2 + ⋅OH → ⋅OOH + H2O (2)
MMPC is approximate 5.1, and Fe/MMPC is positively charged at pH ⋅OOH + ⋅OH → O2 + 2H2O (3)
below pHpzc. TC molecule in solution occurs as cations (pH < 3.3),
zwitterions (pH = 3.3–7.7) and anions (pH > 7.7) (Qi et al., 2018). Solution pH is an important factor in heterogeneous Fenton system.
Electrostatic repulsion may occur between Fe/MMPC and TC molecule Fig. 4c exhibits the impact of pH on TC degradation. TC degradation is
at low pH, agreeing well with low TC adsorption onto Fe/MMPC. efficiently gained at pH 3.0. Gradual decline of TC degradation occurs
Solid catalysts in heterogeneous Fenton system are indispensable for with increasing solution pH to 9.0. When pH is larger than 6.0, it is
H2O2 activation (Wang et al., 2019). It is necessary to determine the difficult to achieve TC degradation. The decline in TC degradation
suitable dosage since it affects the catalytic performance and operating associated with pH variation agrees with previous reports (Li et al.,
costs. Fig. 4a manifests the effect of Fe/MMPC on TC degradation. When 2020; Wang et al., 2021c). The optimal pH of the Fenton reaction be­
the dosage of Fe/MMPC increases from 0 to 80 mg/L, TC degradation is tween Fe2+ and H2O2 is 2.5–4.0. Low pH induces dissolved metal ions
improved significantly. The improved catalytic performance associated instead of forming hydroxide precipitation (He et al., 2020; Yu et al.,
with catalyst dosage are in accordance with previous studies (Chu et al., 2019). The high oxidation potential of active radicals at low pH also
2020; Xie et al., 2020). The addition of Fe/MMPC provides more cata­ contributes to high TC degradation. At higher pH, the weakened TC
lytic sites for H2O2 activation. When the dosage of Fe/MMPC increases degradation can be ascribed to H2O2 decomposition and the formation
from 40 to 80 mg/L, the improvement of degradation efficiency is of insoluble ferric hydroxide. Fig. 4d manifests the effect of co-existing
slightly weakened. Excessive catalyst may undesirably consume reactive ions. The presence of Cl− ion causes slight increase of TC degradation,
radicals. Fe/MMPC of 40 mg/L was chosen for subsequent experiments. while SO2−4 ion shows negligible effect on TC degradation. HCO3 ion

Fig. 4b shows TC degradation as a function of H2O2 concentration. significantly lowers TC degradation, and this can be attributed to the
TC removal is highly limited in absence of H2O2. TC degradation grad­ increased pH.
ually improves when the concentration of H2O2 increases from 0 to 4 TC degradation is enhanced with elevating temperature from 25 to
mM. The formation of reactive radicals such as ⋅OH depends on H2O2. 55 ◦ C (Fig. 4e). Higher temperature may accelerate the molecular
Increasing H2O2 concentration can improve the generation of ⋅OH in diffusion and thermal conversion of H2O2 into radicals. Increasing
heterogeneous Fenton system (Xin et al., 2021). TC degradation is temperature is not favorable for practical application due to higher
almost unchanged by increasing H2O2 from 2 to 4 mM. This can be operational costs. Pseudo-first-order (PFO) kinetic model (Eq. S1) are
ascribed to self-decomposition of hydrogen peroxide (Eq. (1)) and commonly used to study catalytic degradation of pollutants in hetero­
scavenging effect of excessive H2O2 (Eqs. 2-3) (Liu et al., 2020b). geneous Fenton system. The apparent rate constant increases from

6
R. Sun et al. Chemosphere 303 (2022) 135123

Fig. 6. (a) Effect of scavengers on TC degradation (b) spectra of OPD-trapped •OH, (c) ESR spectra of DMPO-trapped •OH and (d) effect of BPY and CA on TC
degradation. (Experimental conditions: Fe/MMPC 40 mg/L, H2O2 2.0 mM, pH 4.0, TC 40 mg/L, and temperature 25 ◦ C).

0.1311 to 0.2999 min− 1 when increasing temperature from 25 to 55 ◦ C system is promising for the treatment of wastewater containing POPs.
(Fig. 4f). Based on the linear format of Arrhenius equation (Eq. S2), the
activation energy of catalytic process is 22 kJ/mol, which is smaller than
that of the previous reports (20–56 kJ/mol) (Wang et al., 2021c). It is 3.4. Catalytic mechanism
significantly lower than that of H2O2 dissociation (214 kJ/mol). The
activation energy is slightly higher than that of diffusion limiting reac­ Quenching tests and ESR determination are carried out to determine
tion (10–13 kJ/mol) (Wang et al., 2019). It is clarified that Fe/MMPC is active radicals. Quenching tests are commonly utilized to elucidate
highly efficient for H2O2 activation toward TC degradation. radicals’ impact (Schneider et al., 2020). Hydroxyl radicals are
Fig. 5a shows the variation of the UV–vis spectra of TC solution. The commonly dominant oxidant in Fenton related processes for pollutants
characteristic peak at 357 nm is attributed to phenolic-diketone reso­ degradation (Dutta et al., 2021). IPA and TBA are widely employed as
nance groups (Wang et al., 2017). The peak obviously declines within scavengers of ⋅OH. Fig. 6a exhibits the obvious inhibition of TC degra­
10 min, implying the decomposition of the tetracene rings. Besides an­ dation after adding TBA and IPA. IPA traps ⋅OH in the solution, while
tibiotics, there exist various POPs in wastewater, such as dyes, phenolic TBA can quench ⋅OH both in solution and catalyst surface (Sun et al.,
compounds, and flotation reagents. Degradation of other pollutants is 2021). The results prove that ⋅OH exists on catalyst surface and in the
further examined by UV–vis spectra. PNP is an important industrial raw solution.
material for p-aminophenol, pesticide, and antipyretic production. Fig. 6b shows the UV–vis spectra of ⋅OH captured by OPD, and the
Fig. 5b displays the rapid decrease of the characteristic peak at 320 nm, peak at near 416 nm proves the existence of ⋅OH (Fang et al., 2009). ESR
indicating the destruction of the p-nitrophenolate in PNP molecule (Park determination was further conducted for radical identification. In
and Bae, 2019). BXT is one common flotation reagents for improving Fig. 6c, the ESR spectra of ⋅OH adducts captured by DMPO show strong
mineral separation, and the residual BXT in flotation wastewater causes four-fold 1:2:2:1 peak (Xin et al., 2021). The increase of peak intensity
environment pollution. From Fig. 5c, the maximum peak at 300 nm indicates the generation of more ⋅OH associated with increasing reaction
assigned to the xanthate group significantly declines within 8 min, time. Above analysis manifests that ⋅OH acts an important role in TC
manifesting the rapid BXT decomposition (Cui et al., 2016). Fig. 5d degradation.
demonstrates the UV–vis spectra of RhB with the characteristic peak of ZVI loads on porous carbon derived from microplastics in Fe/MMPC
N-ethyl group at 554 nm. The dramatic decline of the peak with minor catalyst. Fe0 reacts with H2O2 to form Fe2+ (Eq. (4)). Fe2+ reacts with
shift indicates the dominated pathway of chromophore cleavage for RhB H2O2 to generate ⋅OH (Eq. (5)). The reaction between Fe3+ and H2O2
degradation (Wang et al., 2021b). Figs. S6, S7, and S8 shows the UV–vis produce Fe2+ and ⋅OOH with lower oxidability compared with ⋅OH (Eq.
spectra of PNP, BXT and RhB in the presence of Fe/MMPC, and it is (6)). Fe2+/Fe3+ cycle is commonly limited due to the low reaction rate of
clearly observed the minor contribution of adsorption (Balaraman et al., Eq. (6). The presence of Fe0 promotes Fe2+/Fe3+ cycle through Eq. (7),
2022). Degradation of different pollutants provides that the catalytic thus enhancing catalytic reaction and TC degradation. CA and BPY can
be used to capture Fe3+ and Fe2+, respectively (Wang et al., 2021c).

7
R. Sun et al. Chemosphere 303 (2022) 135123

Fig. 7. Schematic catalytic mechanism in Fe/MMPC and H2O2 system.

Fig. 6d shows TC degradation with the addition of CA and BPY. CA Author contributions statement
causes the slight decline of TC degradation, while BPY induces almost
complete inhibition of TC degradation. The results suggest that Fe2+ is of Ruirui Sun led the experimental studies and the writing of the
high importance for TC degradation. The total Fe concentration is 0.312 manuscript; Jiapeng Yang and Rong Huang contributed to experimental
mg/L after reaction for 10 min, indicating iron leaching due to low so­ studies and analysis of results; Chongqing Wang led the research design
lution pH. Based on above analysis, the schematic illustration of cata­ and contributed to writing and revisions of the manuscript.
lytic mechanism can be shown in Fig. 7.

≡Fe0 + H2O2 + 2H+ → Fe2+ + 2H2O (4) Declaration of competing interest


≡Fe 2+
+ H2O2 → Fe 3+
+ ⋅OH + OH −
(5)
The authors declare that they have no known competing financial
≡Fe3+ + H2O2 → Fe2+ + ⋅OOH + H+ (6) interests or personal relationships that could have appeared to influence
the work reported in this paper.
≡Fe3+ + Fe0 → 2 Fe2+ (7)
Acknowledgements

4. Conclusions This work was supported by the Young Elite Scientist Sponsorship
Program by CAST (2019QNRC001), and the National Key Research and
(i) Controlled carbonization of microplastic improves the yield and Development Project (2020YFC1908802).
surface area of carbon product, making it proper catalyst support.
Loading nano ZVI onto carbon support is verified by Appendix A. Supplementary data
characterizations.
(ii) Fe/MMPC manifests significant improvement of catalyst degra­ Supplementary data to this article can be found online at https://doi.
dation of TC. TC degradation of 81.8% is obtained within 10 min. org/10.1016/j.chemosphere.2022.135123.
TC degradation is promoted with increasing Fe/MMPC, H2O2,
and temperature. Lower solution pH favors TC degradation, while References
co-existing ions shows different effects. The apparent rate con­
stant is 0.1311–0.2999 min− 1 under different temperature, and Ashraf, M.A., Peng, W.X., Fakhri, A., Hosseini, M., Kamyab, H., Chelliapan, S., 2019.
the calculated activation energy is 22 kJ/mol. Different POPs Manganese disulfide-silicon dioxide nano-material: synthesis, characterization,
photocatalytic, antioxidant and antimicrobial studies. J. Photochem. Photobiol. B
including PNP, BXT, and RhB are efficiently obtained in the Biol. 198, 111579.
catalytic system. Auta, H.S., Emenike, C.U., Fauziah, S.H., 2017. Distribution and importance of
(iii) The predominant role of •OH for TC degradation is proved by microplastics in the marine environment: a review of the sources, fate, effects, and
potential solutions. Environ. Int. 102, 165–176.
quenching experiments and ESR spectra. Fe0 in Fe/MMPC is Balaraman, P., Balasubramanian, B., Liu, W.C., Kaliannan, D., Durai, M., Kamyab, H.,
important active sites for enhancing H2O2 activation and Fe2+/ et al., 2022. Sargassum myriocystum-mediated TiO2-nanoparticles and their
Fe3+ cycle. The tentative mechanism is proposed based on antimicrobial, larvicidal activities and enhanced photocatalytic degradation of
various dyes. Environ. Res. 204, 112278.
experimental results and characterizations.
Chu, J.H., Kang, J.K., Park, S.J., Lee, C.G., 2020. Application of magnetic biochar derived
(iv) This work offers novel insights into microplastic management from food waste in heterogeneous Sono-Fenton-like process for removal of organic
and wastewater treatment. Future studies can be conducted dyes from aqueous solution. J. Water Proc. Eng. 37, 101455.
considering (1) real mixed microplastics as carbon sources, (2) Cui, K., He, Y., Jin, S., 2016. Enhanced UV–visible response of bismuth subcarbonate
nanowires for degradation of xanthate and photocatalytic reaction mechanism.
integrated process of microplastics removal and catalyst fabri­ Chemosphere 149, 245–253.
cation, and (3) treatment of real wastewater. Deng, J., Dong, H., Zhang, C., Jiang, Z., Cheng, Y., Hou, K., Fan, C., 2018. Nanoscale
zero-valent iron/biochar composite as an activator for Fenton-like removal of
sulfamethazine. Separ. Purif. Technol. 202, 130–137.
Dissanayake, P.D., Choi, S.W., Igalavithana, A.D., Yang, X., Tsang, D.C., Wang, C.H.,
Ok, Y.S., 2020. Sustainable gasification biochar as a high efficiency adsorbent for

8
R. Sun et al. Chemosphere 303 (2022) 135123

CO2 capture: a facile method to designer biochar fabrication. Renew. Sustain. Energy Thompson, R.C., Olsen, Y., Mitchell, R.P., Davis, A., Rowland, S.J., John, A.W.,
Rev. 124, 109785. Russell, A.E., 2004. Lost at sea: where is all the plastic? Science 304, 838, 838.
Dutta, V., Sharma, S., Raizada, P., Khan, A.A.P., Asiri, A.M., Nadda, A., Nguyen, V.H., Wang, C., Wang, H., Liu, Y., 2016. Purification of Pb (II) ions from aqueous solution by
2021. Recent advances and emerging trends in (BiO)2CO3 based photocatalysts for camphor leaf modified with succinic anhydride. Colloids Surf. A Physicochem. Eng.
environmental remediation: a review. Surface. Interfac. 25, 101273. Asp. 509, 80–85.
Fang, Y.F., Deng, A.P., Huang, Y.P., 2009. Determination of hydroxyl radical in Fenton Wang, X., Jia, J., Wang, Y., 2017. Combination of photocatalysis with hydrodynamic
system. Chin. Chem. Lett. 20, 1235–1240. cavitation for degradation of tetracycline. Chem. Eng. J. 315, 274–282.
González-Soto, N., Hatfield, J., Katsumiti, A., Duroudier, N., Lacave, J.M., Bilbao, E., Wang, C., Wang, H., 2018. Pb (II) sorption from aqueous solution by novel biochar
Cajaraville, M.P., 2019. Impacts of dietary exposure to different sized polystyrene loaded with nano-particles. Chemosphere 192, 1–4.
microplastics alone and with sorbed benzo [a] pyrene on biomarkers and whole Wang, H., Wang, J., Zou, Q., Liu, W., Wang, C., Huang, W., 2018. Surface treatment
organism responses in mussels Mytilus galloprovincialis. Sci. Total Environ. 684, using potassium ferrate for separation of polycarbonate and polystyrene waste
548–566. plastics by froth flotation. Appl. Surf. Sci. 448, 219–229.
He, L., Yang, S.S., Bai, S.W., Pang, J.W., Liu, G.S., Cao, G.L., Ren, N.Q., 2020. Fabrication Wang, C., Wang, H., Cao, Y., 2019. Waste printed circuit boards as novel potential
and environmental assessment of photo-assisted Fenton-like Fe/FBC catalyst engineered catalyst for catalytic degradation of orange II. J. Clean. Prod. 221,
utilizing mealworm frass waste. J. Clean. Prod. 256, 120259. 234–241.
Huang, R., Yang, J., Cao, Y., Dionysiou, D.D., Wang, C., 2022. Peroxymonosulfate Wang, C., Huang, R., Sun, R., 2020. Green one-spot synthesis of hydrochar supported
catalytic degradation of persistent organic pollutants by engineered catalyst of self- zero-valent iron for heterogeneous Fenton-like discoloration of dyes at neutral pH.
doped iron/carbon nanocomposite derived from waste toner powder. Separ. Purif. J. Mol. Liq. 320, 114421.
Technol. 291, 120963. Wang, C., Huang, R., Sun, R., Yang, J., Sillanpää, M., 2021a. A review on persulfates
Jiang, W., Dionysiou, D.D., Kong, M., Liu, Z., Sui, Q., Lyu, S., 2020. Utilization of formic activation by functional biochar for organic contaminants removal: synthesis,
acid in nanoscale zero valent iron-catalyzed Fenton system for carbon tetrachloride characterizations, radical determination, and mechanism. J. Environ. Chem. Eng.
degradation. Chem. Eng. J. 380, 122537. 106267.
Li, X., Cui, K., Guo, Z., Yang, T., Cao, Y., Xiang, Y., Xi, M., 2020. Heterogeneous Fenton- Wang, C., Sun, R., Huang, R., 2021b. Highly dispersed iron-doped biochar derived from
like degradation of tetracyclines using porous magnetic chitosan microspheres as an sawdust for Fenton-like degradation of toxic dyes. J. Clean. Prod. 297, 126681.
efficient catalyst compared with two preparation methods. Chem. Eng. J. 379, Wang, C., Sun, R., Huang, R., Wang, H., 2021c. Superior Fenton-like degradation of
122324. tetracycline by iron loaded graphitic carbon derived from microplastics: synthesis,
Liu, H., Qu, J., Zhang, T., Ren, M., Zhang, Z., Cheng, F., Zhang, Y.N., 2020a. Insights into catalytic performance, and mechanism. Separ. Purif. Technol. 270, 118773.
degradation pathways and toxicity changes during electro-catalytic degradation of Wang, C., Huang, R., Sun, R., Yang, J., Dionysiou, D.D., 2022. Microplastics separation
tetracycline hydrochloride. Environ. Pollut. 258, 113702. and subsequent carbonization: synthesis, characterization, and catalytic
Liu, G., Zhang, Y., Yu, H., Jin, R., Zhou, J., 2020b. Acceleration of goethite-catalyzed performance of iron/carbon nanocomposite. J. Clean. Prod. 330, 129901.
Fenton-like oxidation of ofloxacin by biochar. J. Hazard Mater. 397, 122783. Wei, Y., Wei, S., Liu, C., Chen, T., Tang, Y., Ma, J., Luo, S., 2019. Efficient removal of
Luo, T., Feng, H., Tang, L., Lu, Y., Tang, W., Chen, S., Chen, Z., 2020. Efficient arsenic from groundwater using iron oxide nanoneedle array-decorated biochar
degradation of tetracycline by heterogeneous electro-Fenton process using Cu-doped fibers with high Fe utilization and fast adsorption kinetics. Water Res. 167, 115107.
Fe@ Fe2O3: mechanism and degradation pathway. Chem. Eng. J. 382, 122970. Xiang, M., Huang, M., Li, H., Wang, W., Huang, Y., Lu, Z., Cao, W., 2021. Nanoscale zero-
Minella, M., Bertinetti, S., Hanna, K., Minero, C., Vione, D., 2019. Degradation of valent iron/cobaltmesoporous hydrated silica core–shell particles as a highly active
ibuprofen and phenol with a Fenton-like process triggered by zero-valent iron (ZVI- heterogeneous Fenton catalyst for the degradation of tetrabromobisphenol A. Chem.
Fenton). Environ. Res. 179, 108750. Eng. J. 417, 129208.
Park, J., Bae, S., 2019. Highly efficient and magnetically recyclable Pd catalyst supported Xie, Y., Wang, X., Tong, W., Hu, W., Li, P., Dai, L., Zhang, Y., 2020. FexP/biochar
by iron-rich fly ash@ fly ash-derived SiO2 for reduction of p-nitrophenol. J. Hazard composites induced oxygen-driven Fenton-like reaction for sulfamethoxazole
Mater. 371, 72–82. removal: performance and reaction mechanism. Chem. Eng. J. 396, 125321.
Qi, N., Wang, P., Wang, C., Ao, Y., 2018. Effect of a typical antibiotic (tetracycline) on Xin, S., Liu, G., Ma, X., Gong, J., Ma, B., Yan, Q., Xin, Y., 2021. High efficiency
the aggregation of TiO2 nanoparticles in an aquatic environment. J. Hazard Mater. heterogeneous Fenton-like catalyst biochar modified CuFeO2 for the degradation of
341, 187–197. tetracycline: economical synthesis, catalytic performance and mechanism. Appl.
Raji, M., Mirbagheri, S.A., Ye, F., Dutta, J., 2021. Nano zero-valent iron on activated Catal. B Environ. 280, 119386.
carbon cloth support as Fenton-like catalyst for efficient color and COD removal Yang, Y., Xu, L., Li, W., Fan, W., Song, S., Yang, J., 2019. Adsorption and degradation of
from melanoidin wastewater. Chemosphere 263, 127945. sulfadiazine over nanoscale zero-valent iron encapsulated in three-dimensional
Rana, A., Sudhaik, A., Raizada, P., Nguyen, V.H., Xia, C., Khan, A.A.P., Singh, P., 2022. graphene network through oxygen-driven heterogeneous Fenton-like reactions.
Graphitic carbon nitride based immobilized and non-immobilized floating Appl. Catal. B Environ. 259, 118057.
photocatalysts for environmental remediation. Chemosphere 297, 134229. Yu, Z.L., Li, G.C., Fechler, N., Yang, N., Ma, Z.Y., Wang, X., Yu, S.H., 2016.
Saeed, T., Al-Jandal, N., Al-Mutairi, A., Taqi, H., 2020. Microplastics in Kuwait marine Polymerization under hypersaline conditions: a robust route to phenolic polymer-
environment: results of first survey. Mar. Pollut. Bull. 152, 110880. derived carbon aerogels. Angew. Chem. 128, 14843–14847.
Schirinzi, G.F., Pérez-Pomeda, I., Sanchís, J., Rossini, C., Farré, M., Barceló, D., 2017. Yu, Y., Huang, F., He, Y., Liu, X., Song, C., Xu, Y., Zhang, Y., 2019. Heterogeneous
Cytotoxic effects of commonly used nanomaterials and microplastics on cerebral and Fenton-like degradation of ofloxacin over sludge derived carbon as catalysts:
epithelial human cells. Environ. Res. 159, 579–587. mechanism and performance. Sci. Total Environ. 654, 942–947.
Sahadevan, J., Sojiya, R., Padmanathan, N., Kulathuraan, K., Shalini, M.G., Zhang, B., Song, C., Liu, C., Min, J., Azadmanjiri, J., Ni, Y., Tang, T., 2019. Molten salts
Sivaprakash, P., Muthu, S.E., 2021. Magnetic property of Fe2O3 and Fe3O4 promoting the “controlled carbonization” of waste polyesters into hierarchically
nanoparticle prepared by solvothermal process. Mater. Today Proc. (in press). porous carbon for high-performance solar steam evaporation. J. Mater. Chem. 7,
Schneider, J.T., Firak, D.S., Ribeiro, R.R., Peralta-Zamora, P., 2020. Use of scavenger 22912–22923.
agents in heterogeneous photocatalysis: truths, half-truths, and misinterpretations. Zhang, Y., Jiang, H., Bian, K., Wang, H., Wang, C., 2021. A critical review of control and
Phys. Chem. Chem. Phys. 22, 15723–15733. removal strategies for microplastics from aquatic environments. J. Environ. Chem.
Sun, R., Zhang, X., Wang, C., Cao, Y., 2021. Co-carbonization of red mud and waste Eng. 9, 105463.
sawdust for functional application as Fenton catalyst: evaluation of catalytic activity
and mechanism. J. Environ. Chem. Eng. 9, 105368.

You might also like