J Cej 2018 07 072

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 42

Accepted Manuscript

Enhanced activation of persulfate by carbohydrate-derived carbon cryogels for


effective removal of organic pollutants

Liang Liang, Minghua Zhou, Weilu Yang, Lili Jiang

PII: S1385-8947(18)31315-9
DOI: https://doi.org/10.1016/j.cej.2018.07.072
Reference: CEJ 19470

To appear in: Chemical Engineering Journal

Received Date: 10 April 2018


Revised Date: 19 June 2018
Accepted Date: 10 July 2018

Please cite this article as: L. Liang, M. Zhou, W. Yang, L. Jiang, Enhanced activation of persulfate by carbohydrate-
derived carbon cryogels for effective removal of organic pollutants, Chemical Engineering Journal (2018), doi:
https://doi.org/10.1016/j.cej.2018.07.072

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Enhanced activation of persulfate by carbohydrate-derived carbon cryogels for effective

removal of organic pollutants

Liang Lianga,b,c, Minghua Zhoua,b,c,*, Weilu Yanga,b,c, Lili Jianga,b,c

a
Key Laboratory of Pollution Process and Environmental Criteria, Ministry of Education,

College of Environmental Science and Engineering, Nankai University, Tianjin 300350, P. R.

China.

b
Tianjin Key Laboratory of Urban Ecology Environmental Remediation and Pollution

Control, College of Environmental Science and Engineering, Nankai University, Tianjin

300350, P. R. China.

c
Tianjin Advanced Water Treatment Technology International Joint Research Center, College

of Environmental Science and Engineering, Nankai University, Tianjin 300350, P. R. China.

*
Corresponding author. E-mail address: zhoumh@nankai.edu.cn (M. Zhou).

1
Abstract

Nanocarbons are promising alternatives to metal-based catalysts in advanced oxidation

processes for wastewater treatment. Herein, carbon cryogels were prepared through the

hydrothermal carbonization (HTC) of glucose directed by polyaniline (PANI) and their

catalytic performances towards persulfate (PS) activation for the degradation of organic

pollutants were investigated. When the polymerization temperature of PANI was increased

from 0 to 80 °C, carbon cryogels exhibited increased total pore volume (0.23 to 0.33 cm3 g−1)

and decreased average particle diameter (132 to 63 nm), which further enhanced their

adsorption and catalytic activity for PS activation. The excellent catalytic activity of carbon

cryogels for PS activation was universal in the degradation of different organic pollutants,

comparable to various metal catalysts and reported metal-free catalyst, applied to a wide pH

range (pH 3–9), and easily regenerated through calcination, proving that it had potential

practical application for organic wastewater treatment. Radical quenching and electron

paramagnetic resonance (EPR) experiments demonstrated the nonradical pathway in catalytic

oxidation of orange acid 7 (AO7) on carbon cryogels. It was found that the reaction rate of

catalytic oxidation of AO7 was positively correlated with the meso-/macropore volume and

negatively with particle diameter of carbon cryogels. Finally, the catalytic process for organic

pollutants on carbon cryogels was suggested.

Keywords: Carbon cryogels; Polymerization temperature; Metal-free catalyst; Persulfate

activation; Nonradical mechanism; Organic pollutants

2
1. Introduction

Water pollution has become a serious issue that threatens ecological safety and human

safety. In the past few decades, various technologies have been developed for the removal of

organic pollutants in water [1-3]. Among them, adsorption is a simple and efficient method,

yet it cannot decompose organics and needs a proper post-treatment [4]. One of the most

efficient technologies for decomposing organic pollutants is advanced oxidation processes

(AOPs) based on hydroxyl radicals (•OH) or sulfate radicals (SO4•−) [5]. •OH-based AOPs

have some drawbacks such as pH limitation (pH = 3−4) and instability [6]. SO4•− is an

excellent alternative to •OH due to higher oxidation potential (2.5–3.1 V vs. NHE), wider pH

range (2-8) for application and longer half-life period (30-40 μs) [6]. Usually, SO4•− is

generated by activation of persulfate (PS) or peroxymonosulfate (PMS) using metal-based

catalysts [7-9]. However, the employment of these metal-based catalysts always suffers

from secondary contamination caused by the leaching of toxic metals into water [10, 11].

Therefore, it is highly desirable to develop metal-free catalysts for high-efficient degradation

of organic pollutants.

Recently, carbon materials such as carbon nanotubes, graphene, and nanodiamonds have

been employed for the catalytic oxidation of organic pollutants [11-14]. Like metal-based

catalysts, these carbon materials can also facilitate catalytic activation of PS/PMS to generate

radicals for degradation of organics. In addition, a new nonradical mechanism in the catalytic

oxidation of organics was proposed by Lee’s and Wang’s groups [15, 16]. For example, Duan

et al. reported the synthesis of nitrogen-doped single-walled carbon nanotubes and

demonstrated that graphitic N in carbon structure can act as catalytic sites for PS activation to

3
degrade phenol by a nonradical process [16]. The same results were also found on Prussian

blue analogues derived porous nitrogen-doped carbon microspheres [17]. Besides graphitic N,

defects can also enhance the catalytic oxidation of organic pollutants via a nonradical

pathway [18]. However, few studies have focused on the roles of the pore structure and

particle size of catalysts. On the other hand, these carbon materials used as metal-free

catalysts are derived mainly from nonrenewable fossil resources and they usually suffer from

high cost and complicated preparation technology [19].

The hydrothermal carbonization (HTC) of biomass or carbohydrates is a promising

approach to synthesize carbon materials due to its simple process and the use of low cost and

renewable resources [20-22]. In view of the poor pore structure of hydrochars prepared by the

HTC of pure carbohydrates, many functional carbon materials have been prepared via the

HTC process by introducing different additives such as metal salts, hard/soft templates and

polyionic liquid [23-26]. In particular, the HTC of nitrogen-containing compounds would

result in nitrogen-doped carbon materials which have exhibited excellent performances such

as oxygen reduction reaction, supercapacitors and adsorption [27-31]. Although many

applications have been explored on carbon materials from the HTC of biomass or

carbohydrates, very few studies have been focused on the catalytic oxidation of organic

pollutants in water.

Polyaniline (PANI) is one of the most intensively studied conducting polymers with a

wide application potential as a functional polymer [32]. PANI prepared at higher temperature

contains lower molecular weight chains and more defect sites, which can significantly

depreciate the mechanical and electrical properties [33, 34]. Because of these disadvantages,

4
the application of PANI with many defect sites was ignored. Our group recently introduced

PANI into the process of the HTC of glucose and synthesized monolithic nitrogen-doped

carbon cryogels with an excellent performance for CO2 capture [35]. However, the influence

of the PANI polymerization temperature on the structural properties and surface chemistry

were not explored. In this work, carbon cryogels were prepared at varying polymerization

temperature of PANI (0, 20, 40, 60 and 80 °C) and employed to activate PS for removal of

acid orange 7 (AO7) and other typical organic pollutants. The performance comparison of

carbon cryogels with other metal and metal-free catalysts was investigated. The applied pH

range and reusability of carbon cryogels were also examined for further practical application.

2. Experimental

2.1. Synthesis of carbon cryogels

Carbon cryogels were prepared according to our previous report [35]. Firstly, 10 g

D-glucose was dissolved in 25 mL 1 M HCl solution with a desired temperature. 1.225 g

ammonium persulfate (APS) and 0.5 mL aniline were added sequentially into the solution.

When the solutions turned dark and sticky, they were transferred to a quartz tube, then placed

in a Teflon-lined autoclave and sealed. The mixtures were then heated to 160 °C for 10 h

under self-generated pressure in this closed system. The obtained hydrogels were rinsed in

deionized water and in ethanol several times and then were freeze-dried for 3 days under

vacuum at −50 °C. Carbon cryogels were finally obtained by carbonization at 800 °C for 2 h

in N2 flow of 80 ml min−1 with a heating rate of 10 °C min−1 (Scheme S1). The as-prepared

samples were donated as CC-T, where T corresponds to the polymerization temperature of

5
PANI. The carbonized sample without PANI was also prepared and donated as CS.

2.2. Characterization

Scanning electron microscopy (SEM) was performed on a ZEISS MERLIN Compact

instrument. Transmission electron microscopy (TEM) was operated on a JEOL (JEM-2000

FX) instrument operating at 200 kV. Particle diameter was measured using Nano Measurer

software based on the SEM/TEM images. Nitrogen adsorption-desorption analysis was

carried out at −196 °C using Micromeritics ASAP 2460 Surface Area and Porosity Analyzer.

Raman spectra were recorded on a Renishaw inVia Raman microscope. X-ray photoelectron

spectra (XPS) were obtained on a Kratos AXIS Ultra DLD system under UHV conditions

with Al Kα X-ray. XPS spectra were analyzed with CasaXPS software. Zeta-potential

measurements were carried out on a Malvern Zetasizer Nano ZS instrument.

2.3. Adsorption and catalytic activity tests.

The catalytic activity of the materials was evaluated in the catalytic oxidation of organic

pollutants (AO7, rhodamine B (RhB), phenol and tetracycline). Typically, the reaction was

carried out in a beaker with 100 mL of 20 mg L−1 AO7 solution. 1 mM PS was firstly added

to the solution and then 0.2 g L−1 catalyst was added to start the reaction. At each time

interval, 1mL reaction solution was withdrawn by a syringe and filtered by a 0.45 μm

membrane filter and then immediately mixed with 0.5 mL methanol to quench the reaction.

AO7 adsorption experiments were carried out by the same procedures without the addition of

PS and methanol. For the stability tests, the catalysts were collected by vacuum filtration,

6
washed with deionized water several times, and then dried overnight at 80 °C in an oven. The

regeneration of the catalysts after the 2nd run were conducted by washing with water for

several times and then annealing in a muffle furnace at 350 °C under air atmosphere for 1 h.

2.4. Analytic methods

The absorbance (A) decay of AO7, RhB and tetracycline were determined by UV−vis

spectrophotometer (UV2600, Shimadzu) at the maximum adsorption wavelengths of 484, 554

and 357 nm, respectively. The removal efficiency was calculated according to Eq. (1):

Removal efficiency (%) = [(C0 − Ct)/C0] × 100% (1)

The pseudo-first-order rate constant (k) was determined by Eq. (2):

ln (Ct/C0) = − kt (2)

where C0, and Ct are the pollutants concentration at the initial time and time t, respectively.

Phenol were analyzed by an UltiMate 3000 ultrahigh performance liquid

chromatography (UPLC) with UV lamp at 270 nm with an Acclaim 120 C18 column (3 μm *

3.0 mm * 100 mm). The mobile phase was made of 60% methanol, 38% ultrapure water and

2% acetic acid and at a flow rate of 0.3 mL min−1.

The DMPO (5,5-dimethyl-1-pyrroline-N-oxide) trapped electron paramagnetic

resonance (EPR) spectra were recorded on a MiniScope MS400 spectrometer and operated in

the following conditions: center field: 320 mT; sweep width: 10 mT; microwave frequency:

10 GHz; scan time: 60 s.

3. Results and discussion

7
3.1. Characterization of materials

SEM images of several carbon materials are shown in Fig. 1. Fig. 1a shows that the

HTC of pure glucose followed by carbonization resulted in micrometer-sized carbon spheres

(CS), which was consistent with other literatures [26]. There was a huge difference on the

morphology after adding PANI. Besides, the particle sizes also changed (Table 1). Because of

the crosslinking effect of PANI, the average diameter dropped sharply from about 5 μm (CS)

to 132 nm (CC-0). The higher the polymerization temperature of PANI was, the smaller the

particle diameter of carbon cryogels was (Fig. 1b-f). When the polymerization temperature

increased to 80 °C, the particle diameter decreased to 63 nm. As a novel structure-directing

agent, the molecular weight of PANI decreases with the increase of the polymerization

temperature [33], which might result in the smaller particle size for carbon cryogels obtained

at the higher polymerization temperature. The smaller particle size provides materials with

more active sites, which are beneficial for catalytic activity [36]. Therefore, the smaller

particle size might be favorable to PS activation towards the degradation of organic pollutants.

TEM images show that the as-prepared carbon cryogels held a hierarchical porous structure

composed of interconnected coral-like fibers (Fig. 2a and c). The high-resolution TEM

images (HRTEM) revealed ribbon-like graphitic structure assigned to graphitic layers (002)

(Fig. 2b and d), suggesting the partial graphitization [37].

Fig. 1

Fig. 2

8
Nitrogen adsorption-desorption isotherms were conducted to further investigate the

textural properties of the samples (Fig. 3). CC-0 exhibited type-I isotherm with steep rise at

low relative pressures (< 0.1), suggesting that micropores were dominant [38]. For carbon

cryogels obtained at the higher polymerization temperature, the continuous increase of the

adsorption quantity at the pressure range of 0.1-0.4 indicated the formation of mesopores [38].

Besides, the adsorption quantity increased sharply when the pressure approached 1.0,

suggesting the existence of macropores [37]. The detail data were summarized in Table 1.

Without PANI, CS with the BET surface area of 546 m2 g−1 contained only micropores,

which was in agreement with the previous report [25]. As the polymerization temperature

increased, the total pore volume increased from 0.23 (CC-0) to 0.33 cm3 g−1 (CC-80) while

the volume of micropore almost kept stable, indicating that the higher polymerization

temperature could facilitate the formation of meso-/macropores. Such hierarchically porous

structure endows this materials with plentiful exposed catalytic sites and minimum diffusion

resistance [39], which might be beneficial for PS activation.

Fig. 3

Table 1

The structural properties of the samples were also investigated by Raman spectra as

shown in Fig. 4a. All samples displayed two characteristic peaks at 1355 cm−1 (D band) and

1585 cm−1 (G band) [17]. The broad D band indicated the presence of a large amount of

defects [12]. The intensity ratio of D band to G band (ID/IG) could be used to evaluate the

9
structural disorder of carbon materials. The ID/IG ratio of all samples were about 0.8−0.9,

indicating the similar defective degree for all carbon cryogels.

XPS were carried out to investigate the composition and chemical states of the samples.

Fig. 4b shows two main peaks at 284.5 and 532 eV corresponding to C1s and O1s,

respectively. High-resolution N1s XPS spectra reveals that nitrogen atoms have been

successfully incorporated into carbon cryogels. They could be fitted to three component

peaks corresponding to pyridinic N (397.49-397.7 ev), pyrrolic N (399.6-399.86 ev), and

graphitic N (401.35-401.67 ev) (Fig. 4c and Fig. S1) [12]. A quantitative analysis were

summarized in Table 2. It shows that the contents of carbon (94.17-95.49 at%), nitrogen

(1.56-1.79 at%) and oxygen (2.92-4.04 at%) had little change, indicating that the

polymerization temperature of PANI had little effect on the elemental composition of the

resulting material.

Fig. 4

Table 2

The surface charges of carbon materials can be characterized by zeta-potential

experiments [40]. Fig. 5 shows that zeta-potential experiments as a function of pH on carbon

cryogels. All carbon powders clearly showed positive zeta-potentials below about pH 3,

indicating the coexistence of positively charged amino groups at the particle’s surface with

more negatively charged oxygen groups (COOH and OH) (Fig. S2). At pH 7, all samples

exhibited negative zeta-potentials and the quantitative values had an increasing trend with the

10
polymerization temperature of PANI: −6.6 mV (CC-0), −15.1 mV (CC-20), −15.1 mV

(CC-40), −19.2 mV (CC-60), and −22.9 mV (CC-80). It suggests that negative charges

increased with the polymerization temperature of PANI increasing.

Fig. 5

3.2. Adsorption and catalytic oxidation of AO7 on carbon cryogels

Before the catalytic oxidation, the adsorption performance of AO7 on carbon cryogels

were tested. As shown in Fig. 6a, all samples reached adsorption equilibrium in 45 min. In 60

min, only 4% of AO7 was adsorbed on CC-0. As revealed above, micropores were dominant

for CC-0, which was disadvantage for the adsorption of AO7. With the increase of the

polymerization temperature, the as-prepared carbon cryogels exhibited an enhanced

adsorption ability. CC-80 provided 44% of AO7 adsorption, 11 times higher than that of

CC-0. This could be explained by the increased meso-/macropores volume with the

polymerization temperature increasing. Fig. 6b shows the AO7 removal by catalytic

activation of PS on carbon cryogels. It was noted that PS itself could hardly degrade AO7,

indicating that PS was not activated. After adding CC-0, 10% AO7 removal was achieved in

60 min, indicating that CC-0 had a relatively low catalytic activity for PS activation. The

removal efficiencies of AO7 were different with carbon cryogels obtained at different

polymerization temperature and the trend was agreement with the adsorption experiment. On

CC-80, almost 100% AO7 removal was achieved in 20 min and the corresponding reaction

rate constant was estimated to be 0.21 min−1, about 68 times higher than that of CC-0. To

11
investigate the catalytic effect of carbon cryogels, the adsorption efficiencies were subtracted

from Fig. 6b and the corresponding results are shown in Fig. S3. It was observed that the

AO7 degradation efficiencies by only catalytic oxidation were 6%, 23%, 42%, 54% and 61%

on CC-0, CC-20, CC-40, CC-60 and CC-80, respectively, which indicated that the catalytic

activity of carbon cryogels for PS activation was also enhanced with the increase of

polymerization temperature.

Fig. 6

3.3. Evaluation of catalytic activity of CC-80

Due to the best catalytic activity, CC-80 was chose in the following study. The influence

of different initial pH values on the degradation of AO7 in the CC-80/PS system was

investigated. As shown in Fig. 7a, the removal of AO7 increased with time at all pH values

without any significant difference over pH 3 to 9. At the reaction time of 30 min, the AO7

degradation at pH of 3, 5, 7 and 9 all reached 100%, while the removal of AO7 at pH 11

arrived at 75%. The results was inconsistent with the literature [14], where organics

degradation through a radical reaction was enhanced with the increase of pH value. This

suggests a nonradical mechanism in CC-80/PS system as depicted in Section 3.2. The

decreased catalysis effect at pH 11 might be due to the complexation between OH− and

N-containing functional groups on CC-80, which inhibited the ability of CC-80 to activate PS

[41]. Moreover, CC-80 remained high catalytic activity for PS activation in a wider pH range

(pH 3–9) compared to the conventional homogeneous AOPs (pH < 3). Considering the wide

12
pH range of real wastewater, CC-80 may be a promising heterogeneous catalyst for the real

wastewater treatment.

Classic metal and metal oxides were further employed to compare with CC-80. As

shown in Fig. 7b, Fe3O4, Co3O4, MnO2 and CuO exhibited poor catalytic activities and 2%,

6%, 12% and 18% AO7 removal were achieved in 120 min, respectively. Fe0 could efficiently

activate PS to generate SO4•−, providing 100% AO7 removal in 20 min. However, the main

drawback related to Fe0 is the leaching of iron ion into water, which could lead to secondary

contamination [10, 11]. CC-80 could be even comparable to Fe0, meanwhile completely

avoiding the leaching of metal ions into the water. Moreover, the other reported metal-free

catalysts were compared with CC-80 (Table S1). The pseudo-first-order reaction rate constant

for AO7 degradation in CC-80/PS system was estimated to be 0.21 min−1, 605-fold

enhancement over activated carbon [42]. For phenol degradation, the pseudo-first-order

reaction rate constant was 0.041 min−1, comparable to N-doped carbon nanotube [11],

N-doped single-walled carbon nanotube [43], reduced graphene oxide [44], annealed

nanodiamond [14].

As is known to all, catalyst’s reusability is crucial to the practical application. The tests

on the reusability of CC-80 were conducted and the results were shown in Fig. 7c. 100%

removal of AO7 was achieved in 30 min for the fresh CC-80. In the second run, 78% removal

were achieved in 120 min. The adsorption of AO7 or its intermediates could result in the

coverage of catalytic sites and the decrease of the pore volume (Fig. S4) [14], which might

lead to the decrease of the AO7 removal. Similar results were also reported on the

degradation of phenol in the PS activation with reduced graphene oxide [44]. The

13
regeneration of used CC-80 was conducted by thermal annealing. The activity of the

regenerated CC-80 was significantly recovered as indicated by the complete AO7 removal in

120 min and could be kept stable after the 5th regeneration. This might be due to the removal

of adsorbed AO7/intermediates and the recovery of the pore volume (0.21 cm3 g−1 to 0.32

cm3 g−1) (Fig. S4).

Fig. 7

The catalytic activity of CC-80 was also evaluated on the removal of several typical

organic pollutants including rhodamine B (RhB), phenol and tetracycline. Fig. 8 shows that

about 100% RhB and phenol could be decomposed in 120 min in CC-80/PS system. The

removal of tetracycline (88% in 120 min) was lower than that of the other targets, which

might be attributed to the different molecular structures of the contaminants. The results

strongly indicate that the CC-80/PS system is a promising method to remove organic

pollutants.

Fig. 8

3.4. Mechanism of PS activation and organics oxidation on carbon cryogels

PS can be activated by metal and metal-free catalysts to generate •OH and SO4•− for the

efficient removal of organic pollutants [44]. Ethanol is usually employed to scavenge both


OH (k•OH = 1.2−2.8 ×109 M−1 S−1) and SO4•− ( = 1.6−7.8 ×107 M−1 S−1) generated in

14
the activation process [16]. Therefore, radical quenching experiments were conducted to

investigate the role of radicals in catalytic oxidation of AO7 on CC-80. Fig. 9a shows effect

of ethanol on AO7 degradation in CC-80/PS system. After adding ethanol with a molar ratio

of ethanol to PS at 500 : 1, the AO7 removal efficiency is reduced from 98.5% to 94.2%

within 20 min. Even the ratio rose to 1000 : 1, the AO7 removal efficiency was still 93.9%

within 20 min, which indicated that the contributions of •OH and SO4•− were rather limited.

The Fe2+/PS system was also carried out for comparison (Fig. S5). Without ethanol, 100%

removal of AO7 was achieved in 20 min while only 22% was removed in 60 min after adding

ethanol (the molar ratio of ethanol and PS was 1000 : 1), confirming that the radicals played

an essential role on the AO7 degradation in Fe2+/PS system. These results suggests a

nonradical pathway in the CC-80/PS system, as reported in other literatures [15-17].

EPR technique using DMPO as a radical trapping agent was employed to further

confirm the nonradical mechanism in the CC-80/PS system. Fig. 9b shows the EPR spectra at

different time. In absence of CC-80, no obvious signals were observed. Different from the

signals of DMPO-•OH, DMPO-SO4•− and DMPO-O2•−, a clear signal pattern was obtained in

the presence of CC-80, which was assigned to 5,5-dimethylpyrrolidone-2-(oxy)-(1) or

5,5-dimethyl-1-pyrrolidone-2-oxyl (DMPOX) produced by the oxidation of DMPO.

According to the previous reports [15], the complexes with strong oxidative properties were

formed because of the combination of PS and CC-80.

Fig. 9

15
Generally, defects and graphitic N are both considered as active sites for enhancing the

catalytic oxidation of organic pollutants via a nonradical pathway [17, 18]. According to the

results of Raman and XPS, CC-20 had the highest contents of defects and graphitic N.

However, the catalytic performance of CC-20 was lower. These results suggests that there are

other factors influencing the catalytic performance of carbon cryogels besides defects and

graphitic N.

A large specific surface area can provide more active sites, which is beneficial for

catalytic oxidation of organic pollutants [12]. As shown in Table 1, the specific surface area

of carbon cryogels did not increase with the polymerization temperature of PANI, suggesting

that not all of surface area in carbon cryogels was available for enhancing activation of PS.

From the results above, meso-/macropore volume increased with the increasing

polymerization temperature of PANI. A nearly linear correlation between meso-/macropore

volume and reaction rate of catalytic oxidation of AO7 was further observed in Fig. 10,

demonstrating that reactions of PS activation mainly happened on the meso-/macroporous

surface in carbon cryogels.

In addition, particle size is an important parameter influencing catalytic activity of metal

nanoparticles. With the decreasing size of nanoparticle, there are more surface atoms and a

larger electron density, which are beneficial for heterogeneous catalysis [45]. Recently,

size-dependent electrocatalytic and photocatalytic activities were also reported on metal-free

catalysts [36, 46]. In this study, the particle diameter of carbon cryogels was decreased with

the increase of polymerization temperature of PANI and correlated negatively with the

reaction rate of catalytic oxidation of AO7 (Fig. 10), indicating the importance of particle size

16
on the catalytic activation of PS for AO7 degradation. According to the zeta-potential

experiments, at the high polymerization temperature of PANI, the surface of carbon cryogels

carried more negative charges that can facilitated electron transfer for PS activation [47].

Fig. 10

Based on the results above, the possible mechanism of the catalytic activation of PS on

carbon cryogels for the removal of organic pollutants (i.e. AO7) was illustrated in Fig. 11.

Carbon cryogels (i.e. CC-80) prepared at higher polymerization temperature of PANI had

larger meso-/macropore volume and smaller particle size, contributing to higher adsorption

capacity and more available catalytic sites. After adding CC-80 into the reaction system, PS

and AO7 were adsorbed onto CC-80, followed by in situ PS activation and AO7 oxidation via

a nonradical process, which made a strong decomposition for AO7. In contrast, carbon

cryogels (i.e. CC-0) prepared at lower polymerization temperature of PANI had smaller

meso-/macropore volume and larger particle size, leading to a weak adsorption capacity and

catalytic oxidation of AO7.

Fig. 11

4. Conclusions

In this work, carbon cryogels were prepared by the PANI-directed HTC of glucose at

different polymerization temperature of PANI. The polymerization temperature induced a

17
crucial impact on the structural properties. At a higher polymerization temperature, the

prepared carbon cryogels exhibited a larger total pore volume and a smaller particle size,

which further enhanced their adsorption and catalytic activity for PS activation. The good

catalytic activity of CC-80 for PS activation was universal in the degradation of different

organic pollutants, comparable to various metal catalysts and metal-free catalyst reported,

applied to a wide pH range (pH 3–9), and easily regenerated through calcination. Radical

quenching and EPR experiments demonstrated nonradical mechanism for AO7 oxidation in

CC-80/PS system. The reaction rate of catalytic oxidation of AO7 was examined as functions

of meso-/macropore volume and particle diameter in carbon cryogels and the nearly linear

correlations reflected that reactions of PS activation mainly happened on the

meso-/macroporous surface in carbon cryogels. We believed that this work could provide new

insights to fabricate high-efficient metal-free catalysts for environmental remediation.

Acknowledgements

This work was financially supported by Natural Science Foundation of China (no.

91545126 and 21273120), National Key Research and Development Program (no.

2016YFC0400706), Key Project of Natural Science Foundation of Tianjin (no.

16JCZDJC39300), Sino-Canada Cooperation Projects of Tianjin Binhai, China National

Water Project (nos. 2015ZX07203-11), and Fundamental Research Funds for the Central

Universities.

Appendix A. Supplementary data

18
Supplementary data associated with this article can be found, in the online version, at

References

[1] Z.W. Du, S.B. Deng, S.Y. Zhang, W. Wang, B. Wang, J. Huang, Y.J. Wang, G. Yu, B.S.

Xing, Selective and fast adsorption of perfluorooctanesulfonate from wastewater by magnetic

fluorinated vermiculite, Environ. Sci. Technol. 51 (2017) 8027-8035.

[2] S.J. Varjani, V.N. Upasani, A new look on factors affecting microbial degradation of

petroleum hydrocarbon pollutants, Int. Biodeter. Biodegr. 120 (2017) 71-83.

[3] L. Liang, F.K. Yu, Y.R. An, M.M. Liu, M.H. Zhou, Preparation of transition metal

composite graphite felt cathode for efficient heterogeneous electro-Fenton process, Environ.

Sci. Pollut. R. 24 (2017) 1122-1132.

[4] J. Gong, J. Liu, X.C. Chen, Z.W. Jiang, X. Wen, E. Mijowska, T. Tang, Converting

real-world mixed waste plastics into porous carbon nanosheets with excellent performance in

the adsorption of an organic dye from wastewater, J. Mater. Chem. A 3 (2015) 341-351.

[5] L. Liang, Y. An, M. Zhou, F. Yu, M. Liu, G. Ren, Novel rolling-made gas-diffusion

electrode loading trace transition metal for efficient heterogeneous electro-Fenton-like, J.

Environ. Chem. Eng. 4 (2016) 4400-4408.

[6] P.D. Hu, M.C. Long, Cobalt-catalyzed sulfate radical-based advanced oxidation: A review

on heterogeneous catalysts and applications, Appl. Catal. B: Environ. 181 (2016) 103-117.

[7] G.P. Anipsitakis, D.D. Dionysiou, Radical generation by the interaction of transition

metals with common oxidants, Environ. Sci. Technol. 38 (2004) 3705-3712.

[8] L.W. Matzek, K.E. Carter, Activated persulfate for organic chemical degradation: A

19
review, Chemosphere 151 (2016) 178-188.

[9] F. Ghanbari, M. Moradi, Application of peroxymonosulfate and its activation methods for

degradation of environmental organic pollutants: Review, Chem. Eng. J. 310 (2017) 41-62.

[10] Q.J. Yang, H. Choi, Y.J. Chen, D.D. Dionysiou, Heterogeneous activation of

peroxymonosulfate by supported cobalt catalysts for the degradation of 2,4-dichlorophenol in

water: The effect of support, cobalt precursor, and UV radiation, Appl. Catal. B: Environ. 77

(2008) 300-307.

[11] H.Q. Sun, C. Kwan, A. Suvorova, H.M. Ang, M.O. Tade, S.B. Wang, Catalytic oxidation

of organic pollutants on pristine and surface nitrogen-modified carbon nanotubes with sulfate

radicals, Appl. Catal. B: Environ. 154 (2014) 134-141.

[12] W.C. Peng, S.Z. Liu, H.Q. Sun, Y.J. Yao, L.J. Zhi, S.B. Wang, Synthesis of porous

reduced graphene oxide as metal-free carbon for adsorption and catalytic oxidation of

organics in water, J. Mater. Chem. A 1 (2013) 5854-5859.

[13] S. Indrawirawan, H.Q. Sun, X.G. Duan, S.B. Wang, Low temperature combustion

synthesis of nitrogen-doped graphene for metal-free catalytic oxidation, J. Mater. Chem. A 3

(2015) 3432-3440.

[14] X.G. Duan, C. Su, L. Zhou, H.Q. Sun, A. Suvorova, T. Odedairo, Z.H. Zhu, Z.P. Shao,

S.B. Wang, Surface controlled generation of reactive radicals from persulfate by

carbocatalysis on nanodiamonds, Appl. Catal. B: Environ. 194 (2016) 7-15.

[15] H. Lee, H.J. Lee, J. Jeong, J. Lee, N.B. Park, C. Lee, Activation of persulfates by carbon

nanotubes: Oxidation of organic compounds by nonradical mechanism, Chem. Eng. J. 266

(2015) 28-33.

20
[16] X.G. Duan, H.Q. Sun, Y.X. Wang, J. Kang, S.B. Wang, N-doping-induced nonradical

reaction on single-walled carbon nanotubes for catalytic phenol oxidation, ACS Catal. 5

(2015) 553-559.

[17] N. Wang, W.J. Ma, Z.Q. Ren, Y.C. Du, P. Xu, X.J. Han, Prussian blue analogues derived

porous nitrogen-doped carbon microspheres as high-performance metal-free

peroxymonosulfate activators for non-radical-dominated degradation of organic pollutants, J.

Mater. Chem. A 6 (2018) 884-895.

[18] X.G. Duan, Z.M. Ao, L. Zhou, H.Q. Sun, G.X. Wang, S.B. Wang, Occurrence of radical

and nonradical pathways from carbocatalysts for aqueous and nonaqueous catalytic oxidation,

Appl. Catal. B: Environ. 188 (2016) 98-105.

[19] Y. Si, X.Q. Wang, C.C. Yan, L. Yang, J.Y. Yu, B. Ding, Ultralight biomass-derived

carbonaceous nanofibrous aerogels with superelasticity and high pressure-sensitivity, Adv.

Mater. 28 (2016) 9512-9518.

[20] M.M. Titirici, M. Antonietti, Chemistry and materials options of sustainable carbon

materials made by hydrothermal carbonization, Chem. Soc. Rev. 39 (2010) 103-116.

[21] B. Hu, K. Wang, L.H. Wu, S.H. Yu, M. Antonietti, M.M. Titirici, Engineering carbon

materials from the hydrothermal carbonization process of biomass, Adv. Mater. 22 (2010)

813-828.

[22] M.M. Titirici, R.J. White, C. Falco, M. Sevilla, Black perspectives for a green future:

hydrothermal carbons for environment protection and energy storage, Energy Environ. Sci. 5

(2012) 6796-6822.

[23] Y.H. Ni, L.N. Jin, L. Zhang, J.M. Hong, Honeycomb-like Ni@C composite

21
nanostructures: synthesis, properties and applications in the detection of glucose and the

removal of heavy-metal ions, J. Mater. Chem. 20 (2010) 6430-6436.

[24] S. Ikeda, K. Tachi, Y.H. Ng, Y. Ikoma, T. Sakata, H. Mori, T. Harada, M. Matsumura,

Selective adsorption of glucose-derived carbon precursor on amino-functionalized porous

silica for fabrication of hollow carbon spheres with porous walls, Chem. Mater. 19 (2007)

4335-4340.

[25] S.P. Wang, R.H. Liu, C.L. Han, J. Wang, M.M. Li, J. Yao, H.R. Li, Y. Wang, A novel

strategy to synthesize hierarchical, porous carbohydrate-derived carbon with tunable

properties, Nanoscale 6 (2014) 13510-13517.

[26] P.F. Zhang, J.Y. Yuan, T.P. Fellinger, M. Antonietti, H.R. Li, Y. Wang, Improving

hydrothermal carbonization by using poly(ionic liquid)s, Angew. Chem., Int. Ed. 52 (2013)

6028-6032.

[27] C.L. Han, S.P. Wang, J. Wang, M.M. Li, J. Deng, H.R. Li, Y. Wang, Controlled synthesis

of sustainable N-doped hollow core-mesoporous shell carbonaceous nanospheres from

biomass, Nano Res. 7 (2014) 1809-1819.

[28] M. Qiao, C. Tang, G. He, K. Qiu, R. Binions, I.P. Parkin, Q. Zhang, Z. Guo, M.M.

Titirici, Graphene/nitrogen-doped porous carbon sandwiches for the metal-free oxygen

reduction reaction: conductivity versus active sites, J. Mater. Chem. A 4 (2016) 12658-12666.

[29] S.M. Alatalo, K.P. Qiu, K. Preuss, A. Marinovic, M. Sevilla, M. Sillanpaa, X. Guo, M.M.

Titirici, Soy protein directed hydrothermal synthesis of porous carbon aerogels for

electrocatalytic oxygen reduction, Carbon 96 (2016) 622-630.

[30] M. Sevilla, W. Gu, C. Falco, M.M. Titirici, A.B. Fuertes, G. Yushin, Hydrothermal

22
synthesis of microalgae-derived microporous carbons for electrochemical capacitors, J.

Power Sources 267 (2014) 26-32.

[31] L. Zhao, Z. Bacsik, N. Hedin, W. Wei, Y.H. Sun, M. Antonietti, M.M. Titirici, Carbon

dioxide capture on amine-rich carbonaceous materials derived from glucose, ChemSusChem

3 (2010) 840-845.

[32] S. Bhadra, D. Khastgir, N.K. Singha, J.H. Lee, Progress in preparation, processing and

applications of polyaniline, Prog. Polym. Sci. 34 (2009) 783-810.

[33] J. Stejskal, A. Riede, D. Hlavata, J. Prokes, M. Helmstedt, P. Holler, The effect of

polymerization temperature on molecular weight, crystallinity, and electrical conductivity of

polyaniline, Synthetic Met. 96 (1998) 55-61.

[34] M. Blaha, M. Varga, J. Prokes, A. Zhigunov, J. Vohlidal, Effects of the polymerization

temperature on the structure, morphology and conductivity of polyaniline prepared with

ammonium peroxodisulfate, Eur. Polym. J. 49 (2013) 3904-3911.

[35] L. Liang, M. Zhou, K. Li, L. Jiang, Facile and fast polyaniline-directed synthesis of

monolithic carbon cryogels from glucose, Micropor. Mesopor. Mat. 265 (2018) 26-34.

[36] S.Z. Liu, J. Ke, H.Q. Sun, P. Liu, M.O. Tade, S.B. Wang, Size dependence of uniformed

carbon spheres in promoting graphitic carbon nitride toward enhanced photocatalysis, Appl.

Catal. B: Environ. 204 (2017) 358-364.

[37] X.J. Wei, S.G. Wan, S.Y. Gao, Self-assembly-template engineering nitrogen-doped

carbon aerogels for high-rate supercapacitors, Nano Energy 28 (2016) 206-215.

[38] W.J. Tian, H.Y. Zhang, X.G. Duan, H.Q. Sun, M.O. Tade, H.M. Ang, S.B. Wang,

Nitrogen- and sulfur-codoped hierarchically porous carbon for adsorptive and oxidative

23
removal of pharmaceutical contaminants, ACS Appl. Mater. Inter. 8 (2016) 7184-7193.

[39] S. Dutta, A. Bhaumik, K.C.W. Wu, Hierarchically porous carbon derived from polymers

and biomass: effect of interconnected pores on energy applications, Energy Environ. Sci. 7

(2014) 3574-3592.

[40] N. Baccile, M. Antonietti, M.M. Titirici, One-step hydrothermal synthesis of

nitrogen-doped nanocarbons: Albumine directing the carbonization of glucose,

ChemSusChem 3 (2010) 246-253.

[41] H. Chen, K.C. Carroll, Metal-free catalysis of persulfate activation and organic-pollutant

degradation by nitrogen-doped graphene and aminated graphene, Environ. Pollut. 215 (2016)

96-102.

[42] S.Y. Yang, X. Yang, X.T. Shao, R. Niu, L.L. Wang, Activated carbon catalyzed persulfate

oxidation of Azo dye acid orange 7 at ambient temperature, J. Hazard. Mater. 186 (2011)

659-666.

[43] X.G. Duan, Z.M. Ao, H.Q. Sun, L. Zhou, G.X. Wang, S.B. Wang, Insights into N-doping

in single-walled carbon nanotubes for enhanced activation of superoxides: a mechanistic

study, Chem. Commun. 51 (2015) 15249-15252.

[44] X.G. Duan, H.Q. Sun, J. Kang, Y.X. Wang, S. Indrawirawan, S.B. Wang, Insights into

heterogeneous catalysis of persulfate activation on dimensional-structured nanocarbons, ACS

Catal. 5 (2015) 4629-4636.

[45] S.W. Cao, F. Tao, Y. Tang, Y.T. Li, J.G. Yu, Size- and shape-dependent catalytic

performances of oxidation and reduction reactions on nanocatalysts, Chem. Soc. Rev. 45

(2016) 4747-4765.

24
[46] Q.Q. Li, S. Zhang, L.M. Dai, L.S. Li, Nitrogen-doped colloidal graphene quantum dots

and their size-dependent electrocatalytic activity for the oxygen reduction reaction, J. Am.

Chem. Soc. 134 (2012) 18932-18935.

[47] L. Tang, Y.I. Liu, J.J. Wang, G.M. Zeng, Y.C. Deng, H.R. Dong, H.P. Feng, J.J. Wang, B.

Peng, Enhanced activation process of persulfate by mesoporous carbon for degradation of

aqueous organic pollutants: Electron transfer mechanism, Appl. Catal. B: Environ. 231 (2018)

1-10.

25
Figure Captions:

Fig. 1. SEM images of (a) CS, (b) CC-0, (c) CC-20, (d) CC-40, (e) CC-60, and (f) CC-80.

Fig. 2. (a) TEM and (b) HRTEM images of CC-0; (c) TEM and (d) HRTEM images of

CC-80.

Fig. 3. Nitrogen adsorption-desorption isotherms of carbon cryogels.

Fig. 4. (a) Raman and (b) XPS spectra of carbon cryogels; (c) high-resolution N1s spectra of

CC-80.

Fig. 5. Zeta-potential experiments on carbon cryogels.

Fig. 6. (a) Adsorption of AO7 on carbon cryogels ([AO7] = 20 mg L−1, [catalyst] = 0.2 g L−1,

initial pH = 6.5); (b) catalytic oxidation of AO7 on carbon cryogels ([AO7] = 20 mg L−1,

[catalyst] = 0.2 g L−1, [PS] = 1 mM, initial pH = 6.5).

Fig. 7. (a) Effect of initial solution pH value on the AO7 removal in CC-80/PS system ([AO7]

= 20 mg L−1, [CC-80] = 0.2 g L−1, [PS] = 1 mM); (b) comparison with different catalysts for

PS activation ([AO7] = 20 mg L−1, [catalyst] = 0.2 g L−1, [PS] = 1 mM, initial pH = 6.5); (c)

recyclability tests of CC-80 ([AO7] = 20 mg L−1, [CC-80] = 0.2 g L−1, [PS] = 1 mM, initial

pH = 6.5).

Fig. 8. The degradation of different contaminants in CC-80/PS system ([contaminant] = 20

mg L−1, [CC-80] = 0.2 g L−1, [PS] = 1 mM, initial pH = 6.5).

Fig. 9. (a) Effect of radical quenching on AO7 degradation with CC-80 ([AO7] = 20 mg L−1,

[CC-80] = 0.2 g L−1, [PS] = 1 mM, initial pH = 6.5); (b) EPR spectra obtained by spin

trapping with DMPO in the presence of PS and CC-80 ([CC-80] = 0.2 g L−1, [PS] = 1 mM,

[DMPO] = 10 mM, initial pH = 6.5).

26
Fig. 10. Correlations between meso-/macropore volume, particle diameter and reaction rate

of catalytic oxidation of AO7.

Fig. 11. Proposed mechanism of PS activation on carbon cryogels.

27
a b

10 μm 200 nm

c d

200 nm 200 nm

e f

200 nm 200 nm

Fig. 1

28
a b

100 nm 10 nm

c d

100 nm 5 nm

Fig. 2

29
220 CC-0
CC-20

N2 adsorbed (cm g STP)


200 CC-40
CC-60

-1
CC-80
180

3
160

140

120
0.0 0.2 0.4 0.6 0.8 1.0
Relative pressure (P/P0)

Fig. 3

30
a D G

CC-80 ID/IG = 0.81

Intensity (a.u.)
CC-60 ID/IG = 0.88
CC-40 ID/IG = 0.84

CC-20 ID/IG = 0.85

CC-0 ID/IG = 0.87

1000 1250 1500 1750 2000


-1
Raman shift (cm )

b C1s
Intensity (a.u.)

O1s N1s
CC-80
CC-60
CC-40
CC-20
CC-0

1000 800 600 400 200 0


Binding energy (ev)

c Pyridinic N
Pyrrolic N
Graphitic N
Intensity (a.u.)

396 398 400 402 404 406


Binding energy (ev)

Fig. 4

31
20 CA-0
CA-20
10 CA-40
CA-60

Zeta potential (mV)


0 CA-80

-10

-20

-30

-40

2 4 6 8 10 12
pH

Fig. 5

32
a 1.0
0.8

0.6

C/C0
CC-0
0.4
CC-20
CC-40
0.2 CC-60
CC-80

0.0
0 10 20 30 40 50 60
Time (min)

b 1.0
0.8

0.6
C/C0

Only PS
0.4 CC-0
CC-20
CC-40
0.2
CC-60
CC-80
0.0
0 10 20 30 40 50 60
Time (min)

Fig. 6

33
a1.0 pH = 3
b 1.0
pH = 5
0.8 pH = 7 0.8
pH = 9
pH = 11 Fe3O4
0.6 0.6 Co3O4

C/C0
C/C0

MnO2
0.4 0.4 CuO
0
Fe
0.2 CC-80
0.2

0.0 0.0
0 10 20 30 40 50 60 0 20 40 60 80 100 120
Time (min) Time (min)

c 1.0 1st run


2nd run
0.8 1st regeneration
2nd regeneration
3rd regeneration
0.6 4th regeneration
C/C0

5th regeneration
0.4

0.2

0.0
0 20 40 60 80 100 120
Time (min)

Fig. 7

34
1.0
RhB
Phenol
0.8 Tetracycline

0.6
C/C0
0.4

0.2

0.0
0 20 40 60 80 100 120
Time (min)

Fig. 8

35
a1.0 No ethanol
Ethanol / PS = 500
0.8 Ethanol / PS = 1000

0.6
C/C0
0.4

0.2

0.0
0 10 20 30 40 50 60
Time (min)

b 10 min
Intensity (a.u.)

5 min

1 min

PS only

316 318 320 322 324


Magnetic field (mT)

Fig. 9

36
140 0.14
Partical diameter

Meso-/Macropore volume(cm g )
-1
130 Meso-/Macropore volume

3
0.12

Particle diameter (mn)


120

110 0.10
100
0.08
90

80 0.06
70 2
R1=0.9763 2
R2=0.9652
0.04
60

0.00 0.01 0.02 0.03 0.04 0.05


-1
Rate constant (min )

Fig. 10

37
Fig. 11

38
Table 1

Nitrogen adsorption data for all samples.

SBETa Vtotb Vmicroc APDd


Samples
[m2 g-1] [cm3 g-1] [cm3 g-1] [nm]

CS 546 0.20 0.20 5160

CC-0 559 0.23 0.19 132

CC-20 612 0.27 0.20 107

CC-40 584 0.29 0.19 88

CC-60 603 0.32 0.20 75

CC-80 573 0.33 0.20 63

a
surface area calculated by the Brunauer-Emmett-Teller (BET) method at P/P0 = 0.03−0.12.

b
total pore volume at P/P0 = 0.99. cmicropore volume calculated by t-plot analysis. daverage

particle diameter calculated in Nano Measurer software from SEM images.

39
Table 2

The surface elemental composition and the fitting results for the N1s spectra of carbon

cryogel.

XPS (at%) N1s (at%)


Samples
C O N Pyridinic N Pyrrolic N Graphitic N

CC-0 94.17 4.04 1.79 0.51 0.90 0.38

CC-20 94.64 3.63 1.73 0.51 0.78 0.44

CC-40 95.49 2.92 1.59 0.46 0.73 0.40

CC-60 95.32 3.12 1.56 0.46 0.68 0.42

CC-80 94.99 3.55 1.66 0.43 0.78 0.45

40
Carbon cryogels were an efficient metal-free catalyst for activation of persulfate.

Organic pollutants were oxidized by a nonradical reaction.

Polymerization temperature of PANI had a crucial impact on structural properties.

The pore structure and particle size played important roles on catalytic oxidation.

41

You might also like