Вы находитесь на странице: 1из 8

Separation and Purification Technology 132 (2014) 346–353

Contents lists available at ScienceDirect

Separation and Purification Technology


journal homepage: www.elsevier.com/locate/seppur

Removal of 2,4-dichlorophenol from contaminated soil


by a heterogeneous ZVI/EDTA/Air Fenton-like system
Haiyan Zhou a,b, Qian Sun a,b, Xun Wang a, Linling Wang a,⇑, Jing Chen a, Jingdong Zhang b, Xiaohua Lu a,b,⇑
a
Environmental Science Research Institute, Huazhong University of Science and Technology, Wuhan 430074, PR China
b
School of Chemistry and Chemical Engineering, Huazhong University of Science and Technology, Wuhan 430074, PR China

a r t i c l e i n f o a b s t r a c t

Article history: The removal of 2,4-dichlorophenol (2,4-DCP) from contaminated soil by an amino carboxylic acid-
Received 8 December 2013 enhanced zero-valent iron/Air (ZVI/Air) Fenton-like system has been performed in this study. The effects
Received in revised form 13 May 2014 of ethylenediaminetetraacetic acid (EDTA) and ethylenediamine-N,N0 -disuccinic acid (EDDS) on the cat-
Accepted 15 May 2014
alytic degradation of 2,4-DCP were primarily studied in the presence of ZVI catalyst and O2 oxidant. The
Available online 2 June 2014
result indicated that the degradation of 2,4-DCP in ZVI/Air system was hindered by EDDS but enhanced
by EDTA. Complete destruction of 2,4-DCP in contaminated soil was achieved in the heterogeneous ZVI/
Keywords:
EDTA/Air (ZEA) Fenton-like system under ambient air and room temperature. ZVI and EDTA were the pre-
Soil remediation
ZVI/EDTA/Air
dominant factors influencing the degradation of 2,4-DCP. The degradation of 2,4-DCP was ascribed to het-
EDDS erogeneous catalytic reaction over the ZVI surface in situ modified by EDTA via the formation of
2,4-Dichlorinphnol monodentate inner-sphere complexes. In this system, two main reactive oxygen species of O 
2 /HO2
Superoxide anion radicals and Fe(IV) were generated, in which O 
2 /HO2 was predominant. The novel ZEA oxidation system provides
a promising eco-friendly alternative for the remediation of contaminated soils.
Ó 2014 Elsevier B.V. All rights reserved.

1. Introduction in soil is advanced oxidation techniques (AOTs) with Fenton’s


reagent (Eqs. (1)–(3)) [2,3]. The Fenton reaction has been proved
The contamination of soil is becoming a global problem to be efficient in remediation of organic compounds contaminated
of increasing concerns. Remediation of soil contaminated by soils [4–6]. However, the pH restriction (pH < 4) hinders the practi-
toxic and carcinogenic chemical compounds known as persistent cal applications of Fenton reaction to natural soil pH which is
organic pollutants (POPs) has been demonstrated as one of the approximately neutral or slightly alkaline [7,8].
most difficult challenges. Amongst the group of POPs, 2,4-dichloro-
phenol (2,4-DCP) has been classified as the priority pollutant by US Fe2þ þ H2 O2 ! Fe3þ þ  OH þ OH ð1Þ
EPA and European Commission. 2,4-DCP is wildly utilized in many
agricultural and industrial manufactures such as herbicide, fungi- Fe2þ þ  OH ! Fe3þ þ OH ð2Þ
cide, and wood preservatives. More, chlorophenols (CPs) may
emerge as degradation products from a number of pesticides [1].
H2 O2 þ  OH ! HO2 þ H2 O ð3Þ
Due to their high toxicity and poor biodegradability, contamination
of soil by CPs poses deleterious threats to water resources, public
health and environmental security. Fe0ðsÞ þ H2 O2 þ 2Hþ ! Fe2þ þ 2H2 O ð4Þ
In order to eliminate the risk of CPs contaminated soil, numerous
To overcome this problem, iron-bearing solids [9–11] have been
methods have been developed. One of the techniques used to
employed to form heterogeneous Fenton-like reaction with hydro-
remove toxic and biorefractory contaminants of high concentration
gen peroxide (Eq. (4)). Alternatively, chelating agents can be uti-
lized to enhance the efficiency of heterogeneous Fenton’s
reaction under natural soil pH conditions [12–14]. Nevertheless,
⇑ Corresponding authors. Address: Environmental Science Research Institute,
when H2O2 encounters iron and manganese oxyhydroxides, or soil
Huazhong University of Science and Technology, Wuhan 430074, PR China (X. Lu).
Tel./fax: +86 27 87792159. organic matters commonly existing in surface soils, it quickly
E-mail addresses: wanglinling@mail.hust.edu.cn (L. Wang), hust-esri2009@ decomposes. Therefore, a large amount of H2O2 is required, which
hotmail.com (X. Lu). results in the problems of low efficiency and high cost.

http://dx.doi.org/10.1016/j.seppur.2014.05.037
1383-5866/Ó 2014 Elsevier B.V. All rights reserved.
H. Zhou et al. / Separation and Purification Technology 132 (2014) 346–353 347

It has been reported that the reaction of zero-valent iron (ZVI) 2.2. Apparatus
with dioxygen can generate hydrogen peroxide in situ (Eq. (5)) [15].
Experiments were performed in a 500 mL Pyrex Jacket Reactor
Fe0ðsÞ þ O2 þ 2Hþ ! Fe2þ þ H2 O2 ð5Þ equipped with a two-stage tail gas absorber. A magnetic stirrer
with a temperature detector was applied to mix the slurry in reac-
However, the ZVI/Air system also encounters the problem of
tion evenly and keep the slurry temperature constant. Purified air
iron precipitation under neutral or alkali conditions. Thus, chelat-
or nitrogen was supplied through the glass diffuser regulated by a
ing acids are used to enhance the efficiency of the ZVI/Air system.
rotermeter.
It has been revealed that the complexes of EDTA with Fe(II) is able
to break down the O–O bond of molecular oxygen and generate
H2O2 in situ [16,17]. ZVI/EDTA/Air (ZEA) system has been investi- 2.3. Experimental procedures
gated and applied to oxidize various hazardous substances [18,19].
This novel heterogeneous Fenton-like reaction system is A total of 20 g 2,4-DCP contaminated soil, certain iron powder
arousing great interests of public and researchers from scientific and 400 mL EDTA solution of certain concentration were added
institutes; however, almost all the relevant works were carried into the jacket reactor. The reactor was placed on magnetic stirring
out in solution rather than in soil or slurry. Our group has studied apparatus to mix the slurry uniformly, regulated the temperature
the degradation of DDTs in the soil using the ZVI/EDTA/Air system at certain degree and then aerated to enable the reaction. At prede-
[20]. However, the mechanism for the degradation was barely termined time, 5 mL sample was taken out. At least triplicate runs
explored. It has been a controversial issue concerning reactive were carried out.
oxygen species being hydroxyl radical (OH) or high-valent iron The taken-out slurry was separated by centrifugation at
species (Fe(IV)) in the heterogeneous reaction of ZVI with oxygen 4000 rpm and filtered through 0.45 lm filtration membrane. The
[21,22], while superoxide anion radicals (O 
2 /HO2) was rarely supernatant pH was measured by pH meter and EDTA and 2,4-
mentioned, which was also observed in some heterogeneous DCP were determined by HPLC. The residual 2,4-DCP in solid phase
Fenton-like systems [23]. Pang et al. [24] have proposed that was extracted by 1 mol/L NaOH solution with the assistance of
peroxyl radical should be taken into account to reasonably explain ultrasonication (20 kHz) for 30 min. The recovery of 2,4-DCP in
the discrepancy between their researches. It has been demon- Kaolin was 92.7%. The extraction mixture was then centrifuged,
strated that O
2 radicals are important active oxidative species in and the supernatant was regulated to pH 3–4 by using 1 mol/L
the degradation of a wide range of organic contaminants, including sulfuric acid for analysis of 2,4-DCP. Finally 2,4-DCP in the slurry
chlorinated solvents, pesticides, and dioxins [25,26]. Besides, it can was quantified.
also reduce 2-CB and Fe(III) to biphenyl and Fe(II) respectively
[27,28]. To provide new insight into the ZVI/EDTA/Air reaction,
2.4. Analytical methods
such mechanism is necessary to be explored.
As a structural isomer of EDTA, (S,S)-N,N0 -ethylenediamine dis-
Quantitative analyses of 2,4-DCP, EDTA and EDDS were per-
uccinic acid (EDDS) has been successfully applied to the remedia-
formed with high performance liquid chromatography (Agilent
tion of heavy metals contaminated soils instead of EDTA [29–31].
1200, HPLC) equipped with an Agilent Eclipse XDB-C18 column
Recently, Fe(III)–EDDS in H2O2-based homogeneous Fenton-like
(5 lm, 4.6 mm  150 mm) and a VWD detector at a flow rate of
system has been employed to degrade bisphenol A in solution,
1.0 mL/min. For 2,4-DCP, mobile phase was a mixture of methanol
which offers a new treatment option under higher pH conditions
and 1% acetic acid aqueous solution (V:V, 65:35) and UV absor-
[32,33].
bance wavelength was 280 nm. For EDTA and EDDS determination,
Herein, The comparison of EDTA and EDDS on the removal of
FeCl3 solution and glacial acetic acid was added to sample to con-
2,4-DCP from contaminated soil in the ZVI/Air Fenton-like system
vert EDTA and EDDS to Fe(III) complex. The determination of EDTA
is studied firstly to explore the feasibility of EDDS substitution
was conducted with mobile phase of 0.03 mol/L acetic acid and
for EDTA. Based on the comparative result, the effects of variables
sodium acetate buffer solution (pH = 4) under best UV absorbance
such as EDTA, ZVI, soil pH and air aeration in ZVI/EDTA/Air system
wavelength at 258 nm. The determination of EDDS was conducted
were investigated, and the reaction and degradation mechanisms
with mobile phase of a mixture of methanol and acetate buffer
were clarified as well.
(2 mmol/L tetrabutylammonium hydrogen sulfate and 15 mmol/L
sodium acetate) solution (V:V, 20:80) with UV absorbance wave-
2. Materials and methods length of 236 nm.
Released Cl and low molecular organic acids were detected by
2.1. Materials a Dionex ICS-1500 ion chromatography system equipped with CD
25 conductivity detector, IonPac AS 23 analytical column, IonPac
2,4-Dichlorophenol (2,4-DCP, 99.9%), 2-monochlorophenol AG 23 guard column, and anion ASRS electrolytic suppressor. Inter-
(99.9%), 4-monochlorophenol (99.9%), phenol (99.9%), catechol mediate products were determined by HPLC-MSD. HPLC was Agi-
(99.9%) were purchased from Sigma–Aldrich, USA. Iron powder lent 1100 series and mass spectrometry was carried out using a
(325 mesh), H2Na2EDTA (99%) and HNa3EDDS (35%) were obtained XCT ion trap with electrospray ionization source. Attenuated total
from Sinopharm Chemical Reagent Co. Ltd. Methanol (HPLC grade) reflectance-Fourier transform infrared (ATR-FTIR) analysis was car-
was ordered from Merk, USA. N2 (99.5%) was delivered from ried out with a VERTEX 70 Micro Fourier Transform Infrared/
Wuhan oxygen cylinder test factory. De-ionized water was used Raman spectroscope equipped with a high sensitivity DLATGS
to prepare all solutions. All other reagents used were above analyt- detector and a horizontal ATR-FTIR attachment (Bruker, Germany).
ical grade. The concentration of total dissolved iron ions was detected by
Kaolin clay (chemical purity, Tianjin Kermel Chemical Reagent AA-300 atomic absorbing spectrometer (Perkin Elmer Company,
Development Center, China) was used as simulated soil. 2,4-DCP USA) and dissolved ferrous ion concentration was by 1,10-phe-
contaminated kaolin clay was prepared by adding 2,4-DCP ethanol nantroline spectrometric method. Dissolve oxygen (DO) was mon-
solution into kaolin clay, mixed evenly, vibrated vigorously for 24 h itored by a dissolved oxygen meter (Shanghai Precision & Scientific
in 25 °C thermostat water bath, dried in air to evaporate the Instrument Co., Ltd., JPB-607A). A pH meter (Sartorius universal
methanol thoroughly and grinded to sieve through 100 mesh. type, PB-10) was utilized to track pH variation.
348 H. Zhou et al. / Separation and Purification Technology 132 (2014) 346–353

3. Results and discussions Table 1


kobs values in different systems.

3.1. Comparison of 2,4-DCP degradation in systems with different System kobs(2,4-DCP) (h1) kobs(EDTA) (h1) kobs(EDDS) (h1)
chelating ligands ZVI/Air 0.69 – –
ZVI/EDTA/Air 1.68 1.02 –
Fig. 1a demonstrates the degradation of 2,4-DCP in ZVI/Air, ZVI/EDDS/Air 0.28 – 0.88
ZVI/EDTA/Air and ZVI/EDDS/Air systems. In ZVI/Air system, 72%
2,4-DCP could be removed after 3-h reaction, in which 13%
2,4-DCP was removed by volatization. ZVI was quickly oxidized Moreover, considering that the different effects of EDTA and
to ferrous (Fe(II)) and ferric (Fe(III)) irons by O2, which gave rise EDDS might also arise from their different coordination modes
to the production of reactive oxygen species (ROS), capable of with mineral surface, the modified surfaces of ZVI by EDTA and
decomposing both organic and inorganic contaminants [18,19,34]. EDDS were studied with ATR-FTIR spectroscopy. The ATR-FTIR
Therefore, 2,4-DCP was partly removed in ZVI/Air system. spectra of ZVI before and after aeration for 1 h were recorded
It is worth noting that ZVI corrosion causes the solution pH to (Fig. 2a and b). The peaks at 1613 and 1390 cm1 in ZVI/EDTA/
rise to above 8 in ZVI/Air system. While in ZVI/EDTA/Air and ZVI/ Air system and 1583 and 1393 cm1 in ZVI/EDDS/Air system were
EDDS/Air systems, the pH values were observed to be 5–7 and assigned to characteristic asymmetric stretching vibration (vas) and
5.5–7.5, respectively, without regulating pH. It is believed that symmetric stretching vibration (vs) spectral bands of carboxyl
the chelating agents and their degradation products of low molec- group (–COO). The shifted peaks revealed the direct chemical
ular weight organic acids play a predominant role in the self-buf- bonding between carboxylate groups and ZVI surface.
fering ability [19]. Interestingly, the addition of chelating agents It has been reported that the coordination modes of carboxylate
not only provides the acidic pH buffering ability, but also alters and metal could be distinguished by the difference of wavenumber
the degradation rate of 2,4-DCP. When EDTA was introduced into between the carboxylate stretching bands (Dm = vas(–COO) 
the ZVI/Air system, the degradation of 2,4-DCP was significantly vs(–COO)) [39]. When Dm is larger than 200 cm1, the coordination
promoted. Within 3 h, 98% 2,4-DCP was removal (only including between carboxylate and metal is monodentate binding. In the ZVI/
4% volatile 2,4-DCP). In this system, EDTA could interact with EDTA/Air system, the Dm value of 223 cm1 indicates that EDTA is
metal surface to weaken and break the metal oxygen bonds in surface coordinated with ZVI by monodentate binding, either a
the lattice of the surface oxide layer [35], accelerating the reaction monodentate inner-sphere directly coordinated to surface metal
of iron to Fe(II) and Fe(III). The Fe(II) and Fe(III) were then complex ions or outer-sphere complexes involving ester linkages through
with EDTA to form stable metal-chelate complexes, evidently acti- one carboxylate group of EDTA [23]. Moreover, vas(COO) in the
vating oxygen to generate H2O2 in situ [19,36]. In contrast, an spectra of ZVI in ZVI/EDTA/Air system shifts up to higher wavenum-
inhibitory effect of EDDS ligand on the oxidation reaction was ber in comparison with that of pure EDTA, implying the formation
observed in ZVI/EDDS/Air system, which is consistent with previ- of a direct bonding between O atom of COO in EDTA and the Fe
ous reports in traditional Fenton reaction [37,38]. atom on ZVI surface, which would shorten the C@O bond and
Further analysis indicated that in ZVI /Air, ZVI/EDTA/Air and increase the vas(COO) [40]. Hence, it is concluded that the
ZVI/EDDS/Air systems, degradation of 2,4-DCP followed the adsorption of EDTA onto the surface of ZVI is dominated by the for-
pseudo-first-order kinetic (R2 > 0.95). The pseudo-first-order rate mation of monodentate inner-sphere complexes. The interaction of
constant (kobs) was obtained from the slope of the straight line EDTA on ZVI surface probably enhances the catalytic activity by
for the first-order plot (Table 1). It can be seen that the kobs values competition of affinity of O2 to ZVI, dissolution of iron from solid
for 2,4-DCP were 1.68, 0.69 and 0.28 h1 in ZVI/EDTA/Air, ZVI/Air catalysts and enhancement on the redox properties of iron [41].
and ZVI/EDDS/Air systems, respectively. It is known that in the Consequently, it plays an important role in intensifying the produc-
presence of chelating acids, 2,4-DCP and the two chelating acids tion of ROS.
are simultaneously degraded, and competitive degradation rela- However, in ZVI/EDDS/Air system, the Dm value was 190 cm1,
tionship forms between target pollutant and chelating agents. In less than 200 cm1, indicating a different coordination of EDDS
ZVI/EDTA/Air system, kobs for 2,4-DCP was larger than that for on ZVI surface. Wang et al. [41] have mentioned that on the surface
EDTA (1.02 h1), implying that 2,4-DCP is preferentially oxidized. of FeBiO3, EDTA formed a cave suitable for two H2O2 molecules,
While in ZVI/EDDS/Air system, kobs for 2,4-DCP is less than that and thus the local concentration of H2O2 was higher than that on
for EDDS (0.88 h1), meaning that EDDS is more readily oxidized. the surface of bare or oxalic acid modified FeBiO3. Therefore, it
As the result, 2,4-DCP degradation was restrained. accelerated the generation of ROS. The enhancement of EDTA in

Fig. 1. Performance of different comparison systems: (a) 2,4-DCP decomposition; (b) EDTA decomposition. Initial conditions: iron 12.5 g/L; EDTA(EDDS) 0.4 mmol/L.
H. Zhou et al. / Separation and Purification Technology 132 (2014) 346–353 349

Fig. 2. ATR-FTIR spectra of (a) pure EDTA, ZVI in ZVI/EDTA/Air system and pure ZVI; (b) pure EDDS, ZVI in ZVI/EDDS/Air system and pure ZVI.

this system could be explained reasonably by the similar state- increased from 0.69 to 3.72 h1. While the EDTA concentration
ment. However, due to its different coordination from EDTA, EDDS was further improved to be higher than 1.2 mmol/L, kobs decreased
adsorbed on ZVI surface might fail to form a cave or form a cave of slightly and then approached a constant value. Meanwhile, the
a different size, which reduces the surface Fe sites available for the degradation of EDTA occured but it was not so sensitive to EDTA
interactions with O2, and leads to lower local concentration of ROS concentration as that of 2,4-DCP. With 1.2 mmol/L EDTA, 96%
than that on the bare ZVI surface. Therefore, 2,4-DCP degradation 2,4-DCP and 59% EDTA were degraded in 45 min; and prolonging
was decelerated by ZVI/EDDS/Air system. the reaction time to 6 h, 96% EDTA was removed. In fact, EDTA
remaining in post-reaction solution could be separated for further
3.2. Effects of reaction conditions on2,4-DCP degradation in ZVI/EDTA/ reuse in ZEA system. Although EDTA enhances the oxidation abil-
Air system ity, the degradations of both 2,4-DCP and EDTA are hindered by
excessive EDTA. Excessive EDTA not only impedes the formation
3.2.1. EDTA and ZVI of H2O2–FeII/IIIEDTA and O2–FeII/IIIEDTA adducts necessary for
Both EDTA and ZVI serve as vital agents. Fig. 3a illustrates the H2O2 production [18,42], but also consumes ROS as an organic
effect of initial EDTA concentration on 2,4-DCP degradation. When compound competitively.
the concentration of EDTA was increased from 0 to 1.2 mmol/L, the Fig. 3b shows the effect of ZVI on the degradation of 2,4-DCP
degradation rate constant kobs for 2,4-DCP was significantly and EDTA. Without ZVI, 2,4-DCP was only volatilized by 27%. As

Fig. 3. Effect of EDTA and ZVI: (a) kobs(2,4-DCP) and kobs(EDTA) as concentration of EDTA changed (initial ZVI 12.5 g/L); (b) kobs(2,4-DCP) and kobs(EDTA) as ZVI increased; (c)
concentrations of total dissolved iron and ferrous ion (left) and molar ratio of EDTA to dissolved iron (right); (d) removal efficiency of 2,4-DCP in different systems (Fe2+ and
Fe3+ 1 mmol/L). Initial conditions: iron 7.5 g/L; EDTA 1.2 mmol/L.
350 H. Zhou et al. / Separation and Purification Technology 132 (2014) 346–353

ZVI dosage increased from 1 to 2.5 g, kobs for 2,4-DCP increased reaction and no more than 20% 2,4-DCP was removed partly via
from 2.15 to 4.01 h1 and kobs for EDTA increased from 0.41 to reductive dehalogenation of ZVI [44–47]. Due to the stable struc-
0.56 h1. Although more ZVI in the system accelerates the decom- ture of CPs, the chloride on the aromatic ring is not readily to be
position of contaminants, simply adding ZVI to the system cannot substituted by hydrogen atom generated in the anaerobic system
intensify the degradation. With increasing the amount of ZVI from [48]. When air was pumped into the system at a rate of 1 L/min,
2.5 to 5 g, both kobs(2,4-DCP) and kobs(EDTA) were observed to be kobs(2,4-DCP) rose to 3.95 h1 rapidly. Nevertheless, kobs(2,4-DCP)
slightly decreased from 4.01 to 3.72 h1 and 0.56 to 0.47 h1, decreased to 2.75 h1 as the aeration rate was improved to 2 L/min.
respectively. Generally, more ZVI means more FeII/IIIEDTA complex At the same time, kobs(EDTA) increased to 0.66 h1 at the aeration rate
and better regeneration of FeIIEDTA from FeIIIEDTA, which acceler- of 1 L/min, and then stabilized.
ates the production of H2O2. Nevertheless, excessive ZVI would During the reaction, dissolved oxygen (DO) variation was mon-
also consume the generated H2O2 (Eq. (4)), which appears to itored. As shown in Fig. 4b, without aeration, DO in solution was
overwhelm the benefit of ZVI in H2O2 production (Eq. (5)). There- zero. With increasing the aeration rate to 0.5 L/min, DO was almost
fore, balance of the opposite parts of ZVI should be maintained. zero in the first 2 h and increased rapidly afterword, suggesting that
Since individually increasing ZVI or EDTA addition amount the deficiency of DO restricted the degradation reaction within the
restrains the degradation, the coaction of EDTA and ZVI should initial 2 h. When aeration rate was raised to 1 L/min, sufficient DO
be considered. It has been reported that [EDTA]:[FeII/III] molar ratio was supplied. However, too high aeration might accelerate the for-
above 1:1 inhibited the formation of [FeII(EDTA)(O2)]2+ and mation of hydrated ferric oxide film over the surface of ZVI, imped-
sequential production of H2O2, and consequently decelerated the ing the complexing of EDTA with iron and the generation of ROS.
degradation rate [43]. Fig. 3c shows the results of [EDTA]:[FeII/III]
molar ratio increased from below 1:1 to above 1:1, as ZVI addition 3.2.3. Soil pH
was altered from 2.5 to 5 g with 1.2 mmol/L EDTA. Fig. 5a illustrates the effect of initial pH on the remediation of
Chelating acids could strongly intensify the iron dissolution from 2,4-DCP contaminated soil. At initial pH 3.05, 4.91, 7.03 and
iron-bearing catalysts and promote the Fe3+/Fe2+ recycling. The dis- 10.62, kobs(2,4-DCP) values were 3.30, 3.95, 3.13, 2.70 h1, respec-
solved iron could further propagate homogeneous Fenton reaction tively. In addition, at pH 4.91, which was the original pH of the sys-
along with generated H2O2. To determine the contribution of tem, more than 96% 2,4-DCP was degraded within 45 min, faster
homogeneous Fenton reaction, total dissolved iron and dissociative than the others, which might be contributed by EDTA adsorption
ferrous iron (Fe2+) were detected. As shown in Fig. 3c, the concentra- on ZVI. As presented in Fig. 5b, EDTA adsorption increased as pH
tion of total dissolved iron increased at the initial 10 min but then increased from 3.05 to 4.91, and then decreased as pH continued
gradually decreased. However, no dissociative ferrous iron was to increase to 10.62. Unlike 2,4-DCP, kobs(EDTA) decreased slightly
detected during the whole reaction period. The observed decrease as pH increased, because the acidic condition was generally favor-
of total dissolved iron concentration might be primarily caused by able for EDTA degradation [49].
the degradation of EDTA and generation of passive film on ZVI Regardless of slight differences in degradation rate constants
surface during the reaction. This indicates that the degradation is under all studied pH conditions, 2,4-DCP were all removed by more
not ascribed to homogeneous Fenton reaction. On the other hand, than 98% in 1.5 h. It is inferred that pH in ZEA system is not so
Fe2+ and Fe3+ were employed to substitute ZVI. It was found that important as that in traditional Fenton or Fenton-like system, but
the 3-h removal efficiency of 2,4-DCP was less than 20% (including influence the degradation of contaminants to certain extent by
volatile parts) in both Fe2+/EDTA/Air and Fe3+/EDTA/Air system exerting the effects on surface interaction between EDTA and
(Fig. 3d). Accordingly, we deduce that degradation of organic com- ZVI. That means the contaminants could be degraded effectively
pounds in ZEA system is not attributed to the generated homoge- equally at acid, neutral and even alkaline conditions, which shows
neous Fenton reaction, but mainly ascribed to heterogeneous momentous significance for practical applications.
catalytic reaction over the ZVI surface in situ modified by EDTA.

3.3. Mechanism for 2,4-DCP degradation by ZVI/EDTA/Air system


3.2.2. Aeration
As O2 is indispensable for the oxidation of ZVI to produce reac- As mentioned above, the argument concerning reactive oxygen
tive oxygen species, aeration is considered to be a critical factor in species has never been stopped in the heterogeneous reaction of
the ZEA system. As shown in Fig. 4a, with nitrogen pumped into ZVI with oxygen. In order to assess the importance of possible reac-
the reactor instead of air to create an anoxic system, no obvious tive oxygen species being hydroxyl radical (OH), high-valent iron
appearance of iron oxides was observed throughout the whole species (Fe(IV)) or superoxide anion radicals (O2/HO2), several

Fig. 4. Effect of air aeration: (a) trends of kobs(2,4-DCP) and kobs(EDTA) as air aeration changed; (b) concentration of DO. Initial conditions: ZVI 7.5 g/L; EDTA 1.2 mmol/L.
H. Zhou et al. / Separation and Purification Technology 132 (2014) 346–353 351

Fig. 5. Effect of pH: (a) trends of kobs(2,4-DCP) and kobs(EDTA) as pH changed; (b) EDTA adsorption rate on ZVI surface. Initial conditions: ZVI 7.5 g/L; EDTA 1.2 mmol/L.

Fig. 6. Linear plot of ln(C/C0) versus time with different probe compounds added (a) for 2,4-DCP; (b) EDTA. Initial conditions: ZVI 7.5 g/L; EDTA 1.2 mmol/L and probe
compounds 2 mmol/L.

compounds were employed based on their different reaction rates


with radical species. Benzoquinone was reported as scavenger for
O2/HO2, while methanol was chosen as a probe for OH and
Fe(IV). Besides, tert-Butyl alcohol (TBA) was chosen specially to
distinguish OH, for it can only be oxidized by OH and other stron-
ger oxidants. Therefore, the addition of these probes would change
the degradation rates of 2,4-DCP and EDTA.
As presented in Fig. 6a, compared with 2,4-DCP degradation
rate of 3.94 h1 in blank system (no scavengers added), addition
of benzoquinone restrained the degradation rate most severely to
2.23 h1, followed by methanol (kobs = 2.64 h1); and TBA had little
impact on the degradation rate (kobs = 3.59 h1), which were in
agreement with their impacts on EDTA degradation (Fig. 6b). The
inhibitory influence in the order of benzoquinone > metha-
nol > TBA suggests that in addition to the participation of OH,
O 
2 /HO2 radical and Fe(IV) were mainly responsible for oxidizing
organic pollutants, of which O 
2 /HO2 radical was the dominant
Fig. 7. Released Cl during reaction. Initial conditions: Kaolin clay 20 g, average 2,4-
reactive oxygen species.
DCP concentration 731 mg/kg, ZVI 7.5 g/L, EDTA 1.2 mmol/L.
On the basis of the information obtained above, a possible reac-
tion mechanism of O2 activation by ZVI in the presence of EDTA
was proposed. Firstly, EDTA is bonded on ZVI surface via monoden- proved that the reaction of O 
2 /HO2 production is superior to the

tate binding, facilitating dissolution and redox recycle of iron. reaction of Fe(IV) production, which might be accounted for by
Then, Fe0(s) is oxidized to „Fe2+ or ferrous ion by O2 to produce the reactants supply to some extent. O2 is adequate during the
H2O2 (Eq. (5)). Herein, „Fe2+ stands for Fe(II) sites on the catalyst reaction by successive aeration, while H2O2 is generated midway
surface, produced from Fe0(s) corrosion, „Fe3+ reduction by Fe0, in a small amount and meanwhile consumed rapidly by other
H2O2 and O 
2 /HO2 (Eqs. (6)–(8)). The further reaction between
reactions (Eqs. (1), (2), (7), (11)–(14)).
„Fe2+ species and H2O2 is responsible for the formation of Fe(IV)
BFe3þ þ Fe0ðsÞ ! BFe2þ ð6Þ
(Eq. (9)), where Fe(IV) represents all forms of this species.
Meanwhile, O2 can compete with H2O2 to react with „Fe2+ species,
leading to the production of O  BFe3þ þ H2 O2 ! BFe2þ þ OH ð7Þ
2 /HO2 (Eqs. (10) and (12)). It is
352 H. Zhou et al. / Separation and Purification Technology 132 (2014) 346–353

Fig. 8. Proposed reaction pathway for the degradation of 2,4-DCP and EDTA.

BFe3þ þ O
2 =HO2 ! BFe
 2þ
þ O2 ðþHþ Þ ð8Þ the aromatic ring ruptured and low molecule weight (LMW)
organic acids such as acetic acid and formic acid are produced. As
BFe2þ þ H2 O2 ! FeðIVÞ ðe:g: FeO2þ Þ þ H2 O ð9Þ an organic compound, EDTA is simultaneously degraded and the
main intermediates are proved to be iminodiacetic acid (IMDA),
oxalic acid, acetic acid and formic acid by IC analysis (Fig. 8), in
BFe2þ þ O2 ! BFe3þ þ O 
2 =HO2 ð10Þ
accordance with previous work [18].

BFe2þ þ O þ
2 =HO2 þ H ! BFe
 3þ
þ H 2 O2 ð11Þ
4. Conclusions
3þ 2þ þ
BFe þ H2 O2 ! BFe þH þ HO2 ð12Þ
We found that the oxidation ability of ZVI/Air system was
2þ þ 3þ greatly enhanced by the addition of EDTA but inhibited by EDDS,
BFe þ H2 O2 þ H ! Fe 
þ OH þ H2 O ð13Þ which was probably resulted from their different degradabilities
for 2,4-DCP and coordination modes with ZVI surface. Accordingly,

OH þ H2 O2 ! HO2 þ H2 O ð14Þ the remediation of 2,4-DCP contaminated soil by the novel hetero-
In response to the generated ROS, both 2,4-DCP and EDTA was geneous ZVI/EDTA/Air Fenton-like system was investigated sys-
oxidized by 96% and 63% within 45 min, respectively. As chloride tematically. The 2,4-DCP degradation was dependent upon the
substituents are responsible for the toxicity of aromatic com- amount of added EDTA and ZVI, supplied O2 and pH. With the sup-
pounds, the dechlorination degree was also measured to estimate plements of 7.5 g/L ZVI, 0.4 mmol/L EDTA and 1 L/min aeration,
the detoxification degree. As shown in Fig. 7, Cl gradually released approximate 800 mg/kg 2,4-DCP in 20 g contaminated soil was
from 2,4-DCP during the reaction, and the dechlorination rate of effectively degraded by 96% within 45 min under ambient atmo-
2,4-DCP almost reached 100%, suggesting that the chlorine on the sphere without pH regulation or H2O2 addition, which is of great
aromatic ring was released and no final chlorinate byproducts significance for practical remediation. Moreover, EDTA was simul-
were generated. The slight decrease of Cl after 45 min might be taneously degraded, exerting little safety threaten towards the
due to the adsorption of iron and iron oxides on the surface. environment. The degradation was mainly ascribed to heteroge-
According to HPLC and HPLC–MS analyses, the detected neous catalytic reaction over the ZVI surface in situ modified by
aromatic intermediates were found to be mono-chlorophenol, cat- EDTA. During the heterogeneous Fenton-like reaction, two main
echol, 2-chlorohydroquinone and hydroxylated product of 2,4-DCP. reactive oxygen species namely O 
2 /HO2 and Fe(IV) were gener-

Considering the experimental results as well as the information ated, in which O 2 /HO 
2 was predominant. The investigation
reported in the literatures [50–52], a possible degradation pathway provides a new insight into the mechanism of ZEA system remedi-
for 2,4-DCP was proposed in Fig. 8. In this process, some part of ating organic compounds contaminated soil. The ZEA reaction pro-
2,4-DCP is dechlorinated to mono-chlorophenols (m/z = 126.5 in vides a promising eco-friendly alternative for the remediation of
the chromatograph under negative ionization) by ZVI and O 2 . Fang
contaminated soils.
et al. [28] have found that the degradation of 2-CB by O
2 is a reduc-
tive dechlorination process. The mono-chlorophenols are further Acknowledgements
oxidized to corresponding benzenediol (m/z = 108.5). Some other
part of 2,4-DCP is oxidized by ROS, leading to the formation of cer- The work was supported by National Natural Science Funds of
tain intermediate products, e.g. hydroxylation product of 2,4-DCP China (No. 21077038) and National High Technology Research
(m/z = 177), which has also been observed in several researches and Development (863) Program of China (No. 2012AA06A304)
[51,52]. They were further transformed to 2-chloro-benzenediol and Fundamental Research Funds for the Central Universities
and 4-chloro-benzenediol (m/z = 144.5). As the reaction proceeds, (No. 2013QN065). We appreciated the help from Analytical and
H. Zhou et al. / Separation and Purification Technology 132 (2014) 346–353 353

Testing Center, Huazhong University of Science and Technology, [25] O. Furman, D.F. Laine, A. Blumenfeld, A.L. Teel, K. Shimizu, I.F. Cheng, R.J.
Watts, Enhanced reactivity of superoxide in watersolid matrices, Environ.
China.
Sci. Technol. 43 (2009) 1528–1533.
[26] O. Furman, Reactivity of Oxygen Species in Homogeneous and Heterogeneous
References Aqueous Environments, Ph.D. Dissertation, Washington State University,
Washington, 2009.
[1] A. Lallai, G. Mura, Biodegradation of 2-chlorophenol in forest soil: effect of [27] A.L. Rose, T.D. Waite, Reduction of organically complexed ferric iron by
inoculation with aerobic sewage sludge, Environ. Toxicol. Chem. 23 (2004) superoxide in a simulated natural water, Environ. Sci. Technol. 39 (2005)
325–330. 2645–2650.
[2] C.K.J. Yeh, Y.A. Kao, C.P. Cheng, Oxidation of chlorophenols in soil at natural pH [28] G.D. Fang, D.M. Zhou, D.D. Dionysiou, Superoxide mediated production of
by catalyzed hydrogen peroxide: the effect of soil organic matter, hydroxyl radicals by magnetite nanoparticles: demonstration in the
Chemosphere 46 (2002) 67–73. degradation of 2-chlorobiphenyl, J. Hazard. Mater. 250–251 (2013) 68–75.
[3] L.L. Bissey, J.L. Smith, R.J. Watts, Soil organic matter–hydrogen peroxide [29] B. Kos, D. Le tan, Influence of a biodegradable ([S,S]-EDDS) and nondegradable
dynamics in the treatment of contaminated soils and groundwater using (EDTA) chelate and hydrogel modified soil water sorption capacity on Pb
catalyzed H2O2 propagations (modified Fenton’s reagent), Water Res. 40 phytoextraction and leaching, Plant Soil 253 (2003) 403–411.
(2006) 2477–2484. [30] C. Luo, Z. Shen, X. Li, Enhanced phytoextraction of Cu, Pb, Zn and Cd with EDTA
[4] R.J. Watts, Matthew D. Udell, Paul A. Rauch, Solomon W. Leung, Treatment of and EDDS, Chemosphere 59 (2005) 1–11.
pentachlorophenol-contaminated soils using Fenton’s reagent, Hazard. Waste [31] X. Wang, Y. Wang, Q. Mahmood, E. Islam, X. Jin, T. Li, X. Yang, D. Liu, The effect
Hazard. Mater. 7 (1990) 335–345. of EDDS addition on the phytoextraction efficiency from Pb contaminated soil
[5] B.W. Tyre, R.J. Watts, G.C. Miller, Treatment of four biorefractory contaminants by Sedum alfredii Hance, J. Hazard. Mater. 168 (2009) 530–535.
in soils using catalyzed hydrogen peroxide, J. Environ. Qual. 20 (1991) 832– [32] W. Huang, M. Brigante, F. Wu, K. Hanna, G. Mailhot, Development of a new
838. homogenous photo-Fenton process using Fe(III)–EDDS complexes, J.
[6] P. Kakarla, R. Watts, Depth of Fenton-like oxidation in remediation of surface Photochem. Photobiol. A: Chem. 239 (2012) 17–23.
soil, J. Environ. Eng. 123 (1997) 11–17. [33] W. Huang, M. Brigante, F. Wu, C. Mousty, K. Hanna, G. Mailhot, Assessment of
[7] S. Gan Venny, H.K. Ng, Current status and prospects of Fenton oxidation for the the Fe(III)–EDDS complex in Fenton-like processes: from the radical formation
decontamination of persistent organic pollutants (POPs) in soils, Chem. Eng. J. to the degradation of Bisphenol A, Environ. Sci. Technol. 47 (2013) 1952–1959.
213 (2012) 295–317. [34] I.A. Katsoyiannis, T. Ruettimann, S.J. Hug, PH dependence of Fenton reagent
[8] J.J. Pignatello, E. Oliveros, A. MacKay, Advanced oxidation processes for organic generation and As(III) oxidation and removal by corrosion of zero valent iron
contaminant destruction based on the Fenton reaction and related chemistry, in aerated water, Environ. Sci. Technol. 42 (2008) 7424–7430.
Crit. Rev. Environ. Sci. Technol. 36 (2006) 1–84. [35] J. Rubio, E. Matijević, Interactions of metal hydrous oxides with chelating
[9] R. Matta, K. Hanna, T. Kone, S. Chiron, Oxidation of 2,4,6-trinitrotoluene in the agents. I. b-FeOOH–EDTA, J. Colloid Interface Sci. 68 (1979) 408–421.
presence of different iron-bearing minerals at neutral pH, Chem. Eng. J. 144 [36] X. Xue, K. Hanna, C. Despas, F. Wu, N. Deng, Effect of chelating agent on the
(2008) 453–458. oxidation rate of PCP in the magnetite/H2O2 system at neutral pH, J. Mol. Catal.
[10] L. Xu, J. Wang, Magnetic nanoscaled Fe3O4/CeO2 composite as an efficient A: Chem. 311 (2009) 29–35.
Fenton-like heterogeneous catalyst for degradation of 4-chlorophenol, [37] X. Xu, N.R. Thomson, An evaluation of the green chelant EDDS to enhance the
Environ. Sci. Technol. 46 (2012) 10145–10153. stability of hydrogen peroxide in the presence of aquifer solids, Chemosphere
[11] H. Che, S. Bae, W. Lee, Degradation of trichloroethylene by Fenton reaction in 69 (2007) 755–762.
pyrite suspension, J. Hazard. Mater. 185 (2011) 1355–1361. [38] W. Huang, M. Brigante, F. Wu, K. Hanna, G. Mailhot, Effect of ethylenediamine-
[12] L.R. Bennedsen, A. Krischker, T.H. Jørgensen, E.G. Søgaard, Mobilization of N,N0 -disuccinic acid on Fenton and photo-Fenton processes using goethite as
metals during treatment of contaminated soils by modified Fenton’s reagent an iron source. Optimization of parameters for bisphenol A degradation,
using different chelating agents, J. Hazard. Mater. 199–200 (2012) 128–134. Environ. Sci. Pollut. Res. 20 (2013) 39–50.
[13] F. Vicente, J.M. Rosas, A. Santos, A. Romero, Improvement soil remediation by [39] X. Gao, D.W. Metge, C. Ray, R.W. Harvey, J. Chorover, Surface complexation of
using stabilizers and chelating agents in a Fenton-like process, Chem. Eng. J. carboxylate adheres Cryptosporidium parvum öocysts to the hematite–water
172 (2011) 689–697. interface, Environ. Sci. Technol. 43 (2009) 7423–7429.
[14] N. Kang, I. Hua, Enhanced chemical oxidation of aromatic hydrocarbons in soil [40] X.H. Guan, G.H. Chen, C. Shang, ATR-FTIR and XPS study on the structure of
systems, Chemosphere 61 (2005) 909–922. complexes formed upon the adsorption of simple organic acids on aluminum
[15] C.R. Keenan, D.L. Sedlak, Factors affecting the yield of oxidants from the hydroxide, J. Environ. Sci. 19 (2007) 438–443.
reaction of nanoparticulate zero-valent iron and oxygen, Environ. Sci. Technol. [41] N. Wang, L. Zhu, M. Lei, Y. She, M. Cao, H. Tang, Ligand-induced drastic
42 (2008) 1262–1267. enhancement of catalytic activity of nano-BiFeO3 for oxidative degradation of
[16] S. Seibig, R. van Eldik, Kinetics of [FeII(EDTA)] oxidation by molecular oxygen bisphenol A, ACS Catal. 1 (2011) 1193–1202.
revisited. New evidence for a multistep mechanism, Inorg. Chem. 36 (1997) [42] C.R. Keenan, D.L. Sedlak, Ligand-enhanced reactive oxidant generation by
4115–4120. nanoparticulate zero-valent iron and oxygen, Environ. Sci. Technol. 42 (2008)
[17] V. Zang, R. Van Eldik, Kinetics and mechanism of the autoxidation of iron(II) 6936–6941.
induced through chelation by ethylenediaminetetraacetate and related [43] T. Zhou, T.T. Lim, X. Lu, Y. Li, F.S. Wong, Simultaneous degradation of 4CP and
ligands, Inorg. Chem. 29 (1990) 1705–1711. EDTA in a heterogeneous ultrasound/Fenton like system at ambient
[18] C.E. Noradoun, I.F. Cheng, EDTA degradation induced by oxygen activation in a circumstance, Sep. Purif. Technol. 68 (2009) 367–374.
zerovalent iron/air/water system, Environ. Sci. Technol. 39 (2005) 7158–7163. [44] R. Cheng, J.L. Wang, W.X. Zhang, Comparison of reductive dechlorination of p-
[19] C. Noradoun, M.D. Engelmann, M. McLaughlin, R. Hutcheson, K. Breen, A. chlorophenol using Fe0 and nanosized Fe0, J. Hazard. Mater. 144 (2007) 334–
Paszczynski, I.F. Cheng, Destruction of chlorinated phenols by dioxygen 339.
activation under aqueous room temperature and pressure conditions, Ind. [45] Y.H. Kim, E.R. Carraway, Dechlorination of pentachlorophenol by zero valent
Eng. Chem. Res. 42 (2003) 5024–5030. iron and modified zero valent irons, Environ. Sci. Technol. 34 (2000) 2014–
[20] M.H. Cao, L.L. Wang, L. Wang, J. Chen, X.H. Lu, Remediation of DDTs 2017.
contaminated soil in a novel Fenton-like system with zero-valent iron, [46] Y.S. Keum, Q.X. Li, Reductive debromination of polybrominated diphenyl
Chemosphere 90 (2013) 2303–2308. ethers by zerovalent iron, Environ. Sci. Technol. 39 (2005) 2280–2286.
[21] S.Y. Pang, J. Jiang, J. Ma, Oxidation of sulfoxides and arsenic(III) in corrosion of [47] X. Cong, N. Xue, S. Wang, K. Li, F. Li, Reductive dechlorination of organochlorine
nanoscale zero valent iron by oxygen: evidence against ferryl ions (Fe(IV)) as pesticides in soils from an abandoned manufacturing facility by zero-valent
active intermediates in Fenton reaction, Environ. Sci. Technol. 45 (2010) 307– iron, Sci. Total Environ. 408 (2010) 3418–3423.
312. [48] T. Zhou, Y. Li, F.S. Wong, X. Lu, Enhanced degradation of 2,4-dichlorophenol by
[22] C.K. Remucal, C. Lee, D.L. Sedlak, Comment on ‘‘Oxidation of sulfoxides and ultrasound in a new Fenton like system (Fe/EDTA) at ambient circumstance,
arsenic(III) in corrosion of nanoscale zero valent iron by oxygen: evidence Ultrason. Sonochem. 15 (2008) 782–790.
against ferryl ions (Fe(IV)) as active intermediates in Fenton reaction’’, [49] S. Chitra, K. Paramasivan, P.K. Sinha, K.B. Lal, Ultrasonic treatment of liquid
Environ. Sci. Technol. 45 (2011) 3177–3178. waste containing EDTA, J. Clean. Prod. 12 (2004) 429–435.
[23] M.Q. Wang, N. Wang, H.Q. Tang, M.J. Cao, Y.B. She, L.H. Zhu, Surface [50] R. Takeuchi, Y. Suwa, T. Yamagishi, Y. Yonezawa, Anaerobic transformation of
modification of nano-Fe3O4 with EDTA and its use in H2O2 activation for chlorophenols in methanogenic sludge unexposed to chlorophenols,
removing organic pollutants, Catal. Sci. Technol. 2 (2012) 187–194. Chemosphere 41 (2000) 1457–1462.
[24] S.Y. Pang, J. Jiang, J. Ma, Response to comment on ‘‘Oxidation of sulfoxides and [51] M. Czaplicka, Photo-degradation of chlorophenols in the aqueous solution, J.
arsenic(III) in corrosion of nanoscale zero valent iron by oxygen: evidence Hazard. Mater. 134 (2006) 45–59.
against ferryl ions (Fe(IV)) as active intermediates in Fenton reaction’’, [52] L.J. Xu, J.L. Wang, Fenton-like degradation of 2,4-dichlorophenol using Fe3O4
Environ. Sci. Technol. 45 (2011) 3179–3180. magnetic nanoparticles, Appl. Catal. B – Environ. 123 (2012) 117–126.

Вам также может понравиться