Vous êtes sur la page 1sur 9

Chemosphere 137 (2015) 78–86

Contents lists available at ScienceDirect

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

Dechlorination of polychlorinated biphenyls by iron and its oxides


Yifei Sun a, Xiaoyuan Liu a, Masashi Kainuma b, Wei Wang c, Masaki Takaoka b,⇑, Nobuo Takeda b
a
School of Chemistry and Environment, Beihang University, Beijing 100191, China
b
Department of Environmental Engineering, Kyoto University, Kyoto 615-8540, Japan
c
School of Environment, Tsinghua University, Beijing 100084, China

h i g h l i g h t s

 N2 was the optimum atmosphere for PCB decomposition using iron or iron compounds.
 No matter which PCB congener was tested, Fe3O4 showed the highest activity.
 More active sites for PCBs decomposition were provided with low O/Fe ratio of Fe3O4.
 Reactive characteristics of Cl atoms on benzene rings followed para > meta > ortho.

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

Article history: The decomposition efficiency of polychlorinated biphenyls (PCBs) was determined using elemental iron
Received 28 September 2014 (Fe) and three iron (hydr)oxides, i.e., a-Fe2O3, Fe3O4, and a-FeOOH, as catalysts. The experiments were
Received in revised form 11 February 2015 performed using four distinct PCB congeners (PCB-209, PCB-153, and the coplanar PCB-167 and
Accepted 15 March 2015
PCB-77) at temperatures ranging from 180 °C to 380 °C and under an inert, oxidizing or reducing
Available online 23 May 2015
atmosphere composed of N2, N2 + O2, or N2 + H2. From these three options N2 showed to provide the best
reaction atmosphere. Among the iron compounds tested, Fe3O4 showed the highest activity for decom-
Keywords:
posing PCBs. The decomposition efficiencies of PCB-209, PCB-167, PCB-153, and PCB-77 by Fe3O4 in an
Iron catalysts
O/Fe ratio
N2 atmosphere at 230 °C were 88.5%, 82.5%, 69.9%, and 66.4%, respectively. Other inorganic chlorine
Coplanar PCBs (Cl) products which were measured by the amount of inorganic Cl ions represented 82.5% and 76.1% of
Hydrogen donor the reaction products, showing that ring cleavage of PCBs was the main elimination process. Moreover,
Reaction pathway the dechlorination did not require a particular hydrogen donor. We used X-ray photoelectron spec-
X-ray absorption near edge structure troscopy to analyze the elemental distribution at the catalyst’s surface. The O/Fe ratio influenced upon
(XANES) the decomposition efficiency of PCBs: the lower this ratio, the higher the decomposition efficiency.
X-ray photoelectron spectroscopy X-ray absorption near edge structure spectra showed that a-Fe2O3 effectively worked as a catalyst, while
Fe3O4 and a-FeOOH were consumed as reactants, as their final state is different from their initial state.
Finally, a decomposition pathway was postulated in which the Cl atoms in ortho-positions were more
difficult to eliminate than those in the para- or meta-positions.
Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction In 1968, the Yusho event in Japan (Yoshimura, 2003), and in


1979 the Yucheng event in Taiwan (Rogan et al., 1988) greatly
In the United States, polychlorinated biphenyls (PCBs) were first amplified the concerns over the safety of using PCBs. Because of
produced commercially in 1929 and used as a heat-transfer med- their persistence and propensity to bio-accumulate in fatty tissues,
ium or insulation oil in electrical transformers and capacitors, the manufacturing of PCBs ceased in Japan in 1974 and in the USA
due to their excellent stability and thermal properties (Wentz, in 1977, with the passing of the Law Concerning the Examination
1995; Qi et al., 2014). Over a 50-year period, approximately 635 and Regulation of Manufacture, Etc. of Chemical Substances and the
million tons of PCBs were produced worldwide (Buckley, 1982). Toxic Substances Control Act, respectively. However, a large fraction
of the PCBs that were produced are still present in the environment
(Bert and Esteban, 2008; Lu et al., 2012). From 1930 through to
⇑ Corresponding author at: Kyoto University C-1-3-461, Nishikyo-Ku, Kyoto 615-
1970, the cumulative losses of PCBs to the environment were esti-
8540, Japan.
E-mail address: takaoka.masaki.4w@kyoto-u.ac.jp (M. Takaoka).
mated at 354,000 tons (Ackerman et al., 1983).

http://dx.doi.org/10.1016/j.chemosphere.2015.03.076
0045-6535/Ó 2015 Elsevier Ltd. All rights reserved.
Y. Sun et al. / Chemosphere 137 (2015) 78–86 79

Prior to 1980, the whole amounts of PCBs produced in China (Soekawa Chemical Co. Ltd, Tokyo, Japan), Fe3O4 (1 l) of 99% purity
were 10,000 tons (MEPPRC, 1991; Chen et al., 2008). Due to the (Soekawa), and a-FeOOH powder of 99% purity (Soekawa).
poor storage methods applied to PCB-contaminated equipment in Four congeners of PCBs (PCB-209, PCB-167, PCB-153, and
China, a great deal of soil became contaminated: approximately PCB-77) were selected to investigate the influence of the ortho-po-
50,000 tons with high concentrations (>500 ppm) and approxi- sition on dechlorination; using PCB-209, the entire dechlorination
mately 500,000 tons with low concentrations (50–500 ppm) of pathway could be established. The other three types of PCBs are
PCBs (MEPPRC, 1991; Chen et al., 2008). In 2004, China joined major components of Kanechlor, a commercial Japanese PCB
the Stockholm Convention on Persistent Organic Pollutants, which mixture.
requires all PCB-contaminated waste to be eliminated before 2028.
In developed countries, including Europe, the United States, and 2.2. Apparatus
Japan, a great deal of progress has been achieved in disposing of
PCBs by incineration or other methods, such as chemical destruc- The test decomposition system was designed and operated on
tion using palladium/carbon (Pd/C) as a catalyst (Kume et al., the basis of a pulse reactor. The length of the silica reactor tube
2008). Since the 1980s, zero-valent metals have been used to used was 700 mm, and the inner diameter was 20 mm. A total of
reduce and dechlorinate organic halogen compounds, and the tech- 0.5 g of Fe or Fe compounds was placed in the middle of the fur-
nique has widely been used in the field of environmental engineer- nace, and then 1 mL PCB solution (10 lg/mL hexane) was injected
ing (Fang and Al-Abed, 2008). Zero-valent metals, mainly Fe, Sn, at the inlet of the reactor, using a JP-S-W.6 microfeeder (Furue
and Zn (Boronina et al., 1995; Su et al., 2014), allow the treatment Science Co. Ltd, Japan) and decomposition was continued for
of halogenated hydrocarbons. Zero-iron (Fe) can convert PCBs into 30 min under an N2, N2 + O2, or N2 + H2 flow at a rate of
biphenyl in 10 min at 400 °C (Chuang et al., 1995), and nano-sized 50 mL/min. The electric furnace temperature was controlled at
Fe powder can be used as a catalyst to destroy 1,10 ,2,20 -tetrachlor 230 °C, 330 °C, or 380 °C for the different reaction atmospheres,
oethane (Chuang et al., 1995). Zero-valent Fe acts as an electron respectively. After the decomposition time had elapsed, the
donor and thus becomes oxidized to Fe2+, while C2Cl4 gains an elec- reaction atmosphere was still maintained and the decomposition
tron and is dechlorinated (Li et al., 2006). The decomposition effi- equipment was allowed to cool down to room temperature.
ciency of 1,2,3,4-Tetrachlorodibenzodioxin (TCDD) attains 95% The exhaust gas was sent to impingers filled with 100 mL
with Fe–Pd as a catalyst (Wang et al., 2010). Catalytic hydrodechlo- toluene to collect the decomposition products. Following
rination with noble-metal catalysts has been widely used for the PCB-decomposition, the catalyst was extracted using Soxhlet
dechlorination of a variety of chlorinated organic compounds extraction to quantify the residual products remaining in the cata-
under relatively mild conditions (Ukisu et al., 1996; Urbano and lyst zone. Finally, the PCB homologs and biphenyl were quantified
Marinas, 2001; Alonso et al., 2002; Ukisu and Miyadera, 2003). in the toluene solution and in the extract, using gas chromatogra-
However, noble metals are very expensive, and it is essential to phy and/or mass spectrometry.
find cheaper substances to eliminate PCBs. Activated carbon sup- The analysis was performed using an HP6890 series gas chro-
ported iron showed high PCBs decomposition efficiencies with matograph (Hewlett Packard, Palo Alto, CA), which was connected
inert atmosphere (Sun et al., 2006, 2012). However, as the iron to an HP 5973 mass-selective detector (electron impact, 70 eV)
was easily oxidized, it was not sure what compounds of iron played operated under the selected ion-monitoring (SIM) mode, using
a role to destroy PCBs. the molecular ion of each compound at 1.5–2 scan/s. The capillary
In the present study, we used iron and its (hydro)oxides column was an HP-5MS (Hewlett Packard, Palo Alto, CA), of 60 m in
(a-Fe2O3, Fe3O4, and a-FeOOH) as catalysts to treat several PCB length and an inside diameter of 0.250 mm, with a film thickness
congeners (PCB-209, PCB-167, PCB-153, and PCB-77) under an of 0.25 lm. The carrier gas was helium at a constant flow-rate of
N2, N2 + O2, or N2 + H2 atmosphere. A decomposition pathway 1 mL/min.
was postulated using PCB-209, and we investigated the influence XPS (ESCA-3200, Shimadzu, Japan) using Mg Ka (1253.6 eV)
of the position of chlorine (Cl) atoms on the ease of dechlorination. radiation and a retarding potential analyzer, was used to character-
In addition, the dechlorination mechanism of PCBs was identified ize the different catalysts. The sample chamber vacuum pressure
using coplanar PCBs, i.e. PCB-77, PCB-167 and PCB-153, which was 105–106 Pa. The catalysts were analyzed before and after
are non-ortho, mono-ortho and di-ortho coplanar PCBs, respec- reaction under atmospheres of N2 and N2 + O2 at 330 °C. X-ray
tively. Coplanar PCBs such as polychlorinated dibenzodioxins absorption fine structure spectroscopy (XAFS) measurements were
(PCDDs) and polychlorinated dibenzofurans (PCDFs) are highly carried out at a synchrotron radiation facility in Japan, used a
toxic (Sun et al., 2006), compared to non-coplanar PCBs, and the beamline BL01B1 in SPring8 (Aritani et al., 2002). The spectra were
toxic effects of coplanar PCBs occur at relatively smaller concentra- collected in the fluorescence mode using a 19-element Ge
tions than those of non-coplanar PCBs (Sun et al., 2012). The solid-state detector for the Fe or Fe compound disks, and in the
molecular structures and C–Cl bonds were different when the transmission the mode using an ionization chamber for the model
coplanar PCB species were different, which is important to decide AC. K edge XANES spectra of a-Fe2O3, Fe3O4, a-FeOOH before and
the Cl position easy to dechlorinate. In order to select the most use- after decomposition were compared.
ful iron or iron compounds, X-ray photoelectron spectroscopy
(XPS) and X-ray absorption near edge structure (XANES) analyses
3. Results and discussion
were used to analyze the structural changes in the iron or iron
compounds of the catalysts used during the reaction (Tanaka
The decomposition efficiency and the biphenyl yield were cal-
et al., 1988).
culated as follows:

Decomposition efficiency ð%Þ ¼ ð1  c=c0 Þ  100% ð1Þ


2. Experimental section
Biphenyl yieldð%Þ ¼ cb =c0  100% ð2Þ
2.1. Materials and reagents
where c0 is the original amount of PCBs injected into the system,
Four kinds of catalysts were used: Fe powder of 177 lm and c and cb are the total amounts of PCBs and biphenyl remaining
(Nakalai Tesque Inc., Kyoto, Japan), a-Fe2O3 (1 lm) of 99.9% purity on the catalyst and absorbed in toluene, respectively. All
80 Y. Sun et al. / Chemosphere 137 (2015) 78–86

experiments were conducted twice under the same experimental with 87.1% efficiency; however, Fe and a-Fe2O3 had much lower
conditions with a RSD less than 2%. efficiencies of 41% and 13.1%, respectively. As shown in
Fig. 1(b), a maximum efficiency of 82.5% of PCB-167 was reached
with Fe3O4, followed by a-FeOOH at 65.6%, a-Fe2O3 at 42.9%, and
3.1. Effects of iron and its (hydr)oxide on PCB decomposition
Fe at 14.8% efficiency. For PCB-153 (shown in Fig. 1c), a maximum
efficiency of 69.9% was reached with Fe3O4, followed by a-FeOOH
At 180 °C, the decomposition efficiency was fairly low. PCB-209
at 58.1%, a-Fe2O3 at 26.4%, and Fe at 6.7% efficiency. For PCB-77
was not decomposed by either Fe or a-Fe2O3 and only 45.1% and
(shown in Fig. 1d), a maximum efficiency of 66.4% was reached
10.4% decomposition efficiencies were achieved with Fe3O4 and
with Fe3O4, followed by a-FeOOH at 51.9%, Fe at 47.7%, and
a-FeOOH at 180 °C. However, at 380 °C, the decomposition effi-
ciencies attained with all four catalysts exceeded 90%, for example,
a-Fe2O3 at 44.7% efficiency. The same tendency is shown: a maxi-
mum decomposition efficiency for the 4 kinds of PCB congeners is
being 92.1%, 92.5%, and 98.7% for PCB-167 with Fe, a-Fe2O3, and
reached with Fe3O4, and then followed by a-FeOOH, a-Fe2O3, and
Fe3O4, respectively. Fig. 1 shows the PCB decomposition efficien-
Fe in that order. The dechlorination of PCBs appeared to be more
cies. Panel (a) shows that for PCB-209, for which a maximum effi-
effective for compounds with more Cl atoms. The average biphenyl
ciency of 88.5% was reached with Fe3O4, followed by a-FeOOH
yields for levels of PCB-209 of 1.58% and PCB-77 of 0.4% also
provide evidence for this inference. The activities of the different
iron forms for PCB decomposition were most active in the follow-
ing order: Fe3O4 > a-FeOOH > a-Fe2O3 > Fe. That is, Fe3O4 had the
highest activity for decomposing each of the four PCB congeners
under the experimental conditions tested. Its decomposition
efficiency was highest in the following order: PCB-209 >
PCB-167 > PCB-153 > PCB-77.
This pattern follows their respective PCB orbital bond energies
(4.1–5.5 eV) and zero-point energies (0.38–1.65 eV). The bond dis-
sociation energy varies inversely with the number of Cl atoms pre-
sent in chlorinated aromatic compounds (Yamauchi et al., 2006).
This effect may be related to the p–p conjugation of the C–Cl bond
that reduces molecular reactivity. As on the phenyl ring, the C–H
bond is substituted by a C–Cl bond, which results in a conjugation
between the p electron orbit and the p electron orbit of the Cl
atom. The electron density of the Cl atom is dispersed to keep
the molecule more stable. Therefore, molecules with more Cl will
show greater p–p conjugations, or lower molecular activities, and
consequently express reduced stability (Zhang et al., 2008).
The two hexachlorobiphenyls PCB-167 and PCB-153 differ in
their substituent positions: PCB-167 has a meta-position and
PCB-153 has an ortho-position. No matter which catalyst was used,
the decomposition of PCB-167 was always higher than that of
PCB-153. During decomposition, a Cl atom at an ortho-position is
more difficult to remove than a meta-substituted Cl atom. In gen-
eral, PCBs dechlorinate in a predictable manner with chlorine
removal from each position following the general trend para P me-
ta » ortho chlorines (Lowry and Johnson, 2004). According to the
molecular orbital theory of chemical reactivity, the lowest unoccu-
pied molecular orbital (LUMO) energy property expresses the
quantum activation state; therefore, it serves as a generalized
descriptor of soft electrophilicity, or the tendency to accept elec-
trons from a donor species (Campos et al., 2001). The correlation
suggests that the LUMO energy of the test chemical is involved
in either a charge transfer interaction with a cellular receptor or
a process of accepting electrons.
The LUMO energies of PCB-167 and PCB-153 are 0.0658 and
0.0527 eV, respectively (Zhou et al., 2005). The carbon atoms
attached to the electron-withdrawing Cl atoms may act as elec-
trophiles and react with nucleophilic species, forming a covalent
bond by replacing Cl atoms (Zhang et al., 2008). The lower the
LUMO energy, the easier it is for the PCB molecule to receive elec-
trons and the easier it is for dechlorination to occur; therefore,
PCB-167 displayed a higher dechlorination efficiency than
PCB-153.

3.2. Effect of atmosphere on PCB decomposition

Fig. 1. Decomposition efficiency and biphenyl yield of (a) PCB-209, (b) PCB-167, (c)
To study the effect of atmospheric composition on the decom-
PCB-153, and (d) PCB-77, using Fe, a-Fe2O3, Fe3O4 and a-FeOOH in an N2 position of PCBs, three different inert, oxidizing or reducing gases
atmosphere at 230 °C. were used as a carrier: N2, N2 + O2, and N2 + H2. For these
Y. Sun et al. / Chemosphere 137 (2015) 78–86 81

experiments, Fe3O4 and a-FeOOH, the two most active catalysts, efficiency was as low as 4% with Fe3O4 or a-FeOOH under an
were used at a temperature of 230 °C. Fig. 2(a) shows the decom- N2 + O2 atmosphere. Therefore, N2 was an optimum reaction atmo-
position efficiencies of PCB-209 with Fe3O4 and a-FeOOH using sphere not only in terms of decomposition efficiency, but also in
these three different gases, and the resultant biphenyl yields are terms of biphenyl yield. The Biphenyl yield was 1.6% in an N2
also shown. When Fe3O4 was used as a catalyst, the PCB-209 atmosphere, higher than under either an N2 + O2 or N2 + H2 atmo-
decomposition efficiencies were 88.5% and 80.3% in a flow of N2 sphere. These results were similar when a-FeOOH was used. The
and N2 + H2, respectively. However, the PCB-209 decomposition N2 + O2 atmosphere inhibited PCB decomposition, either by

Fig. 2. The decomposition efficiency of PCB-209, ratio of products, and decomposition products using Fe3O4 and a-FeOOH at 230 °C under different atmospheres. (a)
Decomposition efficiency (DEP) and biphenyl yield (BY) of PCB-209. (b) Ratio of decomposition products including biphenyls (BY), CBzs, PCBs except PCB-209, PCB-209 and
others. (c and d) The decomposition product distribution of PCB-209 with Fe3O4 or a-FeOOH under N2. (e and f) The decomposition product distribution of PCB-209 with
Fe3O4 or a-FeOOH under N2 + O2. (g and h) The decomposition product distribution of PCB-209 with Fe3O4 or a-FeOOH under N2 + H2.
82 Y. Sun et al. / Chemosphere 137 (2015) 78–86

competing for electrons with PCBs or by disrupting metal catalysis. were used to characterize the chemical forms before and after
Fig. 3 shows a TG-DTA diagram of Fe3O4 and a-FeOOH under an decomposition by each catalyst at a decomposition temperature
N2 + O2 atmosphere, while heating from room temperature to of 330 °C. The main elements of Fe and O were analyzed quantita-
500 °C. The TG curve of Fe3O4 showed an exothermic reaction peak tively based on XPS analysis.
at 230 °C; meanwhile, the weight of Fe3O4 increased. The reaction Fig. 4(a) shows the O/Fe and C/Fe ratios of different catalysts
in Eq. (3) took place: the product’s weight can increase by 3.5%, before and after reaction in an N2 or N2 + O2 atmosphere. The ratios
which is in accordance with the TG curve of Fe3O4 shown in Fig. 3. were 7.85 for Fe, 7.22 for a-Fe2O3, 4.34 for Fe3O4, and 6.67 for
a-FeOOH. The elemental Fe used in this study showed a high ratio
2Fe3 O4 þ 1=2O2 ! 3a  Fe2 O3 ð3Þ
because it is easily oxidized in the air during the XPS measurement
Using a-FeOOH at 250 °C, conversion takes place according to process. However, a-Fe2O3 is a stable form of iron oxide, so it dis-
reaction (4): in theory the weight of reactants may be reduced played an oxygen-saturated state before and after decomposition,
by 11%, which is in accordance with the TG curve of a-FeOOH even under an N2 or N2 + O2 atmosphere. The decomposition of
shown in Fig. 3. PCBs followed a two-phase gas–solid reaction, taking place on
the catalyst surface; as discussed above, Fe3O4 was the most active
a  FeOOH () a  Fe2 O3 þ H2 O ð4Þ
catalyst for PCB decomposition. Therefore, the lower the O/Fe ratio,
In the presence of H2, the decomposition efficiency came close the more active was the catalyst. For lower ratios, there appeared
to that in a pure N2 atmosphere; however, biphenyl yields were to be more oxygen-unsaturated sites, which were the active sites
only 0.8% with Fe3O4 and 0.6% with a-FeOOH. Under N2, the biphe- and these were able to promote the dechlorination of PCBs.
nyl yields were up to 1.6% with Fe3O4 and 2.7% with a-FeOOH. The The detection depth of XPS increased over time. Accordingly,
reaction atmosphere also affected the decomposition product dis- the O/Fe ratios of the four catalysts steeply declined, especially that
tribution. Fig. 2(b) shows the product ratio after PCB-209 break- for elemental Fe, which decreased from 7.85 to 0.83. Apparently,
down under the three types of atmosphere. The N2 atmosphere the surface of the elemental Fe was oxidized from the start. The
showed the highest ratio of chlorobenzene formation with both tests revealed that N2 was the optimum atmosphere for PCB
kinds of catalysts. The percent of chlorobenzene was 3.9% and decomposition, while N2 + O2 inhibited the reaction. This was
6.0% using Fe3O4 and a-FeOOH, respectively. Other inorganic Cl undoubtedly associated with the high O/Fe ratio. The introduction
products which were measured by the amount of inorganic Cl ions of O2 reduced the activity of the catalyst. Oxygen atoms occupied
represented 82.5% and 76.1% of the reaction products, when using active sites at the surface, in competition with PCBs. This indirectly
Fe3O4 and a-FeOOH as catalysts under an N2 atmosphere. These explains the suppression of decomposition in an N2 + O2
results show that ring cleavage of PCBs was the main process. atmosphere.
Moreover, dechlorination did not require a hydrogen donor. Fig. 4(b–d) shows the XANES spectra of a-Fe2O3, Fe3O4, and
Fig. 2(c–h) shows the product distribution of PCB-209 using a-FeOOH before and after decomposition under an N2 atmosphere
Fe3O4 or a-FeOOH under each atmosphere at 230 °C. A total of at a decomposition temperature of 330 °C. Because Fe was very
20.01 nmol PCB-209 was injected into the decomposition system. easily oxidized during the sample pretreatment process, it was
Under N2, the level of biphenyl formation was 0.33 nmol with not measured by XAFS. The pattern-fitting calculation was
Fe3O4 and 0.54 nmol with a-FeOOH; under N2 + H2, there was performed using REX2000 (RIGAKU, Japan). The best-fit patterns
0.16 nmol and 0.11 nmol biphenyl with Fe3O4 and a-FeOOH, obtained by minimizing residual (R-factor) are shown with the
respectively. In contrast, the sum of di- to nona-CBs was 0.08 nmol observed XANES spectra. As shown in Fig. 4(b), the XANES spectra
under N2 and 0.40 nmol under N2 + H2 with Fe3O4. Therefore, of a-Fe2O3 did not change before or after decomposition, in accor-
among the three different atmospheres used in this study, N2 dance with the results of XPS. According to the pattern-fitting
was the most appropriate for PCB dechlorination and biphenyl results of Fe3O4 shown in Fig. 4(c), the ratio of Fe3O4 was 82%,
yield. while the ratio of a-Fe2O3 was 18% after decomposition. Reaction
Eq. (4) is reversible, and was calculated using the
3.3. Effects of catalyst characteristics on decomposition efficiency Materials-oriented Little Thermodynamic (MALT2) database
(Japan Society of Calorimetry and Thermal Analysis). The reversible
Among the three catalysts, Fe3O4 was the most active, followed reaction of Eq. (3) occurred to produce a-Fe2O3. This implies that
by a-FeOOH, a-Fe2O3, and elemental Fe. XPS and XANES analyses the XANES spectra of a-Fe2O3 did not change after decomposition
(Fig. 4b). For a-FeOOH, according to the pattern-fitting calculation,
98% of the structure was a-Fe2O3 after decomposition (shown in
100 10 Fig. 4d). There was only 2% of a-FeOOH that did not react. This con-
250rC firms the results of TG-DTA and reaction Eq. (4) at 250 °C, where
230rrC the reaction with a-FeOOH generated a-Fe2O3. Hence, a-Fe2O3
95 5
worked as a catalyst, while Fe3O4 and a-FeOOH were consumed
as reactants as final state is different from the initial state.
Weight (%)

90 0
DTA (rC)

3.4. Decomposition product distribution and decomposition pathway


315rrC analysis
85 -5
Fe3O4 TG It was also necessary to detect the different dechlorination
α-FeOOH TG products after the reaction, to acquire a better understanding of
80 -10
Fe3O4 DTA
the PCB decomposition process. Fig. 5 presents the decomposition
α-FeOOH DTA
efficiency and product distribution of PCB-167, PCB-153, and
75 -15 PCB-77 at 230 °C, 330 °C, and 380 °C under an N2 atmosphere with
0 100 200 300 400 500
Fe, a-Fe2O3 and Fe3O4. Note the larger extent of dechlorination giv-
Temperature (rC)
ing rise to biphenyl as the decomposition temperature raised. More
Fig. 3. A TG-DTA diagram of Fe3O4 and a-FeOOH under an N2 + O2 atmosphere from biphenyl was produced under an N2 atmosphere than in the other
room temperature to 500 °C. atmospheres, using the same types of catalyst. Furthermore, as
Y. Sun et al. / Chemosphere 137 (2015) 78–86 83

Fig. 4. O/Fe ratio by XPS analysis and Fe K edge XANES spectra. (a) O/Fe ratio before and after decomposition under different conditions, (b) Fe K edge XANES spectra of a-
Fe2O3, (c) Fe K edge XANES spectra of Fe3O4, and (d) Fe K edge XANES spectra of a-FeOOH.

temperature increased, decomposition increased according to the decomposition products, they were still associated with the kind
amount of biphenyl. The phenomenon of biphenyl increasing with of atmosphere. Both the product distribution and product yield
rising temperature was same as in Zhang’s reports (Zhang and Hua, varied with the atmosphere applied.
2000). There are two possible pathways for formation of biphenyl: The decomposition products of Deca-CB were as follows:
one is cleavage of C–Cl bond followed by H addition, another path-
way is combination of two phenyl radicals.  Nona-CBs: PCB-207 > PCB208 >> PCB-206.
The profiles present a maximum of biphenyl among the prod-  Octa-CBs: PCB-197 P PCB-200 P PCB-199 > PCB-204.
ucts, such as PCB-167, products of biphenyl accounting for 59.5%,  Hepta-CBs: PCB-176 > PCB-184 > PCB-179.
77.2% and 94.1% with Fe, a-Fe2O3 and Fe3O4 at 380 °C. A total of
27.7 mmol PCB-167 was used at the start. After the reaction, the Through analysis of the products, the pathway of PCB-209
residual PCB-167 contents were 0.5 mmol using Fe3O4, 2.1 mmol dechlorination was determined (Fig. 6). During the decomposition
with a-Fe2O3, and 2.3 mmol with Fe. The DiPenta-CBs were of PCB-209, the Cl atoms located at the para- and meta-positions
0.1 mmol using Fe3O4, 0.1 mmol with a-Fe2O3, and 0.7 mmol with are broken off preferentially and form the two main nona-CB prod-
Fe, while biphenyl contents were 10.9 mmol with Fe3O4, 7.9 mmol ucts (PCB-208 and PCB-207). Next, these atoms are removed, form-
with a-Fe2O3, and 4.4 mmol with Fe. In short, the higher the ing other octa- and hepta-CB products. Throughout the conversion,
decomposition efficiency, the higher the biphenyl yield, and the the Cl atoms at ortho-positions remain difficult to remove. This was
lower the amount of chlorinated PCBs. Decomposition efficiency in accordance with Choi et al.’s opinion, which was not regard for
and product yields are closely related to the forms of the iron cat- substrate effects on independent or competitive conditions (Choi
alyst; however, the distribution of PCB congeners was independent et al., 2009). The dechlorination pathways of PCB-167 and
from the forms of iron. No matter which kind of catalyst was used, PCB-153 followed the same rules. The four tetra-CB products
the product sequence always proceeded as follows: (PCB-66, PCB-67, PCB-70, and PCB-74) were all detected after
PCB-167 decomposition. In the same way, PCB-153 was dechlori-
biphenyl > Hexa-CBs > Penta-CBs > Tetra-CBs > Tri-CBs > Di-CBs: nated producing two kinds of tri-CBs: PCB-17 and PCB-18.

When PCB-153 and PCB-77 were decomposed, the results fol-


lowed the same pattern. Biphenyl created from PCB-153 decompo- 3.5. Kinetic study of decomposition
sition was 3.0, 6.3 and 8.9 mmol, from PCB-77 decomposition was
9.3, 9.8, 13.6 mmol with Fe, a-Fe2O3 and Fe3O4 at 380 °C, account- The kinetics of the decomposition reaction is believed to con-
ing for 43.9%, 80.5% and 95.2% from PCB-153, 80.7%, 80.4% and form to the first-order reaction model (Björk and Gilek, 1997). It
97.6% from PCB-77 with Fe, a-Fe2O3 and Fe3O4 at 380 °C. can be calculated quantitatively by using the first-order reaction
Although the catalyst did not affect the distribution of model, because the surface absorption turns out to be the
84 Y. Sun et al. / Chemosphere 137 (2015) 78–86

Fig. 5. The decomposition efficiency and product distribution of PCB-167, PCB-153, and PCB-77 at 230 °C, 330 °C, and 380 °C under an N2 atmosphere with three different
catalysts. (a) The decomposition efficiency (DEP) and biphenyl yield (BY) of PCB-167, PCB-153, and PCB-77 at 230 °C, 330 °C, and 380 °C under an N2 atmosphere with three
different catalysts. (b–d) Product distribution of PCB-167, PCB-153, and PCB-77 at 380 °C under an N2 atmosphere with three different catalysts. Note the large extent of
dechlorination giving rise to biphenyl.

rate-controlling step whose absorption rate is faster than the rate respectively. Taking the catalytic decomposition of PCB-167 as an
of decomposition. We can obtain the first-order kinetics by the example, the apparent activation energy of the PCB-167 decompo-
characteristics of the first-order reactions, as Eq. (5): sition are 66.32 kJ/mol with Fe, 44.90 kJ/mol with a-Fe2O3 and
22.18 kJ/mol with Fe3O4 which decrease 66.6% compared with Fe
lnð1=1  xÞ ¼ kt ð5Þ at 330 °C and 41.83 kJ/mol with Fe, 27.87 kJ/mol with a-Fe2O3
and 16.62 kJ/mol with Fe3O4 which decrease 60.3% compared with
where x represents decomposition efficiency of reactants, %; k rep-
Fe at 380 °C. Nevertheless, the effects of different catalysts on the
resents the apparent rate constant, min1; t represents the reaction
apparent activation energy for PCB-77 at 330 °C and 380 °C has
time, maintaining 30 min in this study.
equally matched levels of 27.0–30.5 kJ/mol. This discovery meets
The apparent rate constant, k can be described by the Arrhenius
exactly the catalytic effects of the three materials above on PCB
equation (Eq. (6)):
decomposition as the following order: Fe < a-Fe2O3 < Fe3O4.
k ¼ AeEa=RT ð6Þ Catalysts promote reactions by decreasing their activation energy
(Murena and Schioppa, 2000). Obviously PCB decomposition cat-
where A represents the pre-exponential factor; Ea represents appar- alyzed by Fe3O4 with the minimum apparent activation energy
ent activation energy, kJ/mol; R is the gas constant, 8.315 J/(K mol); has the maximum decomposition efficiency. Moreover, the appar-
T represents the reaction temperatures, K. ent activation energy of the PCB-153 decomposition are
From Eqs. (5) and (6) (Lin et al., 2012), the apparent activation 69.22 kJ/mol with Fe, 61.46 kJ/mol with a-Fe2O3 and 27.78 kJ/mol
energy of the catalytic decomposition of PCB-167, PCB-153 and with Fe3O4 at 330 °C which are relatively higher than the apparent
PCB-77 by using Fe, a-Fe2O3 and Fe3O4 are shown in Fig. 7. The activation energy of 65.02 kJ/mol with Fe, 41.63 kJ/mol with
apparent activation energy of the decomposition for both a-Fe2O3 and 23.74 kJ/mol with Fe3O4 at 380 °C. Higher tempera-
PCB-167 and PCB-153 catalyzed by Fe, a-Fe2O3 and Fe3O4 has the ture will promote the decomposition of PCB-167 and PCB-153 gen-
similar trend of Fe > a-Fe2O3 > Fe3O4 at 330 °C and 380 °C, erally since they have lower apparent activation energy at 380 °C,
Y. Sun et al. / Chemosphere 137 (2015) 78–86 85

PCB-209

PCB-208
PCB-207

PCB-204
PCB-199

PCB-200
PCB-197

PCB-179 PCB-184
PCB-176

Fig. 6. The dechlorination pathway from Deca-CB to Hepta-CBs.

the atmosphere was changed from N2 to N2 + O2, the decomposi-


tion efficiency of PCB-209 using Fe3O4 decreased from 88.5% to
0% as the oxygen atoms occupied the active sites and inhibited
the catalyst’s activity. Therefore, N2 was the optimum atmosphere
for PCB decomposition. If the catalyst was active or the atmosphere
was beneficial for the decomposition reaction, the PCBs not only
dechlorinated efficiently, but the resulting dechlorination products
were decomposed at the same time, and the biphenyl yield
increased. The distribution of decomposition products was
independent from the catalyst, but it was closely associated with
the atmosphere used. During the dechlorination of PCBs, the Cl
atoms located at the para- and meta-positions were eliminated
preferentially, whereas those at the ortho-positions were difficult
to remove. The apparent activation energy of the first-order
Fig. 7. The apparent activation energy of the catalytic decomposition of PCB-167, decomposition reaction followed the order of
PCB-153 and PCB-77 by using the catalysts of Fe, a-Fe2O3, Fe3O4 at 330 °C and Fe3O4 < a-FeOOH < a-Fe2O3 < Fe.
380 °C, respectively.

Acknowledgements
whereas, it has little influence on the decomposition of PCB-77.
Because with temperature increasing, reaction with a greater The synchrotron radiation experiments were performed at
apparent activation energy will have a more obvious change than SPring-8 (2005B0439). The authors are grateful to the National
that with a smaller one. Natural Science Foundation of China (Project No. 21277010,
2141101075), the National Commonweal Project of the Ministry
of Environmental Protection (No. 201209005) for providing finan-
4. Conclusions cial assistances.

Iron or iron compounds can be used as catalysts for decompos- References


ing PCBs in different gas flows (under an N2 or N2 + H2 atmo-
sphere), because the dechlorination process does not involve a
hydrogen donor. No matter which PCB congener was tested, Ackerman, D. G., Scinto, L., Bakshi, P., Delumyea, R., Johnson, R., Richard, G., Takata,
A., Sworzyn, E., 1983. Destruction and Disposal of PCBs by Thermal and Non-
Fe3O4 among the iron compounds showed the highest activity Thermal Methods. Park Ridge: New Jersey.
and the highest decomposition efficiency for decomposing PCBs. Alonso, F., Beletskaya, I.P., Yus, M., 2002. Metal-mediated reductive
The decomposition efficiencies of PCB-209, PCB-167, PCB-153, hydrodehalogenation of organic halides. Chem. Rev. 102, 4009–4091.
Aritani, H., Nishimura, S., Tamai, M., Yamamoto, T., Tanaka, T., Nakahira, A., 2002.
and PCB-77 using Fe3O4 in an N2 atmosphere were 88.5%, 82.5%,
Local structure of framework iron in Fe-substituted Al-mordenites by Fe K edge
69.9%, and 66.4%, respectively. At a relatively low O/Fe ratio of XAFS. Chem. Mater. 14, 562–567.
4.34 at the surface of Fe3O4, more active sites for the decomposi- Bert, B.V., Esteban, A., 2008. Long-term worldwide QA/QC of dioxins and dioxin-like
tion of PCBs were provided. The surface of elemental Fe was highly PCBs in environmental samples. Anal. Chem. 80, 3956–3964.
Björk, M., Gilek, M., 1997. Bioaccumulation kinetics of PCB 31, 49 and 153 in the
oxidized, and some adsorbed oxygen also appeared at the surface, blue mussel, Mytilus edulis L. as a function of algal food concentration. Aquat.
limiting the decomposition activity of PCBs by elemental Fe. When Toxicol. 38, 101–123.
86 Y. Sun et al. / Chemosphere 137 (2015) 78–86

Boronina, T., Klabunde, K.J., Sergeev, G., 1995. Destruction of organohalides in water Rogan, W.J., Gladen, B.C., Hung, K.L., Koong, S.L., Shih, L.Y., Taylor, J.S., Wu, Y., Yang,
using metal particles: carbon tetrachloride/water reactions with magnesium, D., Ragan, N.B., Hsu, C.C., 1988. Congenital poisoning by polychlorinated
tin, and zinc. Environ. Sci. Technol. 29, 1511–1517. biphenyls and their contaminants in Taiwan. Science 241, 334–336.
Buckley, E.H., 1982. Accumulation of airborne polychlorinated biphenyls in foliage. Su, G., Lu, H., Zhang, L., Zhang, A., Huang, L., Liu, S., Li, L., Zheng, M., 2014. Thermal
Science 216, 520–522. degradation of octachloronaphthalene over as-prepared Fe3O4
Campos, M., del Carmen Núñez, M., Rodrı´guez, A., Entrena, A., Hernández-Alcoceba, micro/nanomaterial and its hypothesized mechanism. Environ. Sci. Technol.
R., Fernández, F., Lacal, J.C., Gallo, M.A., Espinosa, A., 2001. LUMO energy of 48, 6899–6908.
model compounds of bispyridinium compounds as an index for the inhibition of Sun, Y., Takaoka, M., Takeda, N., Matsumoto, T., Oshita, K., 2006. Kinetics on the
choline kinase. Eur. J. Med. Chem. 36, 215–225. decomposition of polychlorinated biphenyls with activated carbon-supported
Chen, J., Lu, C., Liu, X., Zhang, J., 2008. PCBs Disposal Initial Strategic Research in iron. Chemosphere 65, 183–189.
China. China Environmental Science Press, Beijing. Sun, Y., Takaoka, M., Takeda, N., Wang, W., Zeng, X., Zhu, T., 2012. Decomposition of
Choi, H., Al-Abed, S.R., Agarwal, S., 2009. Catalytic role of palladium and relative 2,20 ,4,40 ,5,50 -hexachlorobiphenyl with iron supported on an activated carbon
reactivity of substituted chlorines during adsorption and treatment of PCBs on from an ion-exchange resin. Chemosphere 88, 895–902.
reactive activated carbon. Environ. Sci. Technol. 43, 7510–7515. Tanaka, T., Yamashita, H., Tsuchitani, R., Funabiki, T., Yoshida, S., 1988. X-ray
Chuang, F., Larson, R.A., Wessman, M.S., 1995. Zero-valent iron-promoted absorption (EXAFS/XANES) study of supported vanadium oxide catalysts.
dechlorination of polychlorinated biphenyls. Environ. Sci. Technol. 29, 2460– structure of surface vanadium oxide species on silica and c-alumina at a low
2463. level of vanadium loading. J. Chem. Soc., Faraday Trans. 1: Phys. Chem. Condens.
Fang, Y., Al-Abed, S.R., 2008. Dechlorination kinetics of monochlorobiphenyls by Fe/ Ph. 84, 2987–2999.
Pd: effects of solvent, temperature, and PCB concentration. Appl. Catal. B: Ukisu, Y., Miyadera, T., 2003. Hydrogen-transfer hydrodechlorination of
Environ. 78, 371–380. polychlorinated dibenzo-p-dioxins and dibenzofurans catalyzed by supported
Kume, A., Monguchi, Y., Hattori, K., Nagase, H., Sajiki, H., 2008. Pd/C-catalyzed palladium catalysts. Appl. Catal. B: Environ. 40, 141–149.
practical degradation of PCBs at room temperature. Appl. Catal. B: Environ. 81, Ukisu, Y., Iimura, S., Uchida, R., 1996. Catalytic dechlorination of polychlorinated
274–282. biphenyls with carbon-supported noble metal catalysts under mild conditions.
Li, X., Elliott, D.W., Zhang, W., 2006. Zero-valent iron nanoparticles for abatement of Chemosphere 33, 1523–1530.
environmental pollutants: materials and engineering aspects. Crit. Rev. Solid Urbano, F.J., Marinas, J.M., 2001. Hydrogenolysis of organohalogen compounds over
State Mater. Sci. 31, 111–122. palladium supported catalysts. J. Mol. Catal. A: Chem. 173, 329–345.
Lin, S., Su, G., Zheng, M., Ji, D., Jia, M., Liu, Y., 2012. Synthesis of flower-like Co3O4– Wang, Z., Huang, W., Peng, P., Fennell, D.E., 2010. Rapid transformation of 1,2,3,4-
CeO2 composite oxide and its application to catalytic degradation of 1,2,4- TCDD by Pd/Fe catalysts. Chemosphere 78, 147–151.
trichlorobenzene. Appl. Catal. B: Environ. 123, 440–447. Wentz, C.A., 1995. Hazardous Waste Management. McGraw-Hill, New York. p. 72.
Lowry, G.V., Johnson, K.M., 2004. Congener-specific dechlorination of dissolved Yamauchi, T., Kobayashi, S., Itoh, S., Yamasaki, K., Kamei, Y., Kanno, Y., 2006.
PCBs by microscale and nanoscale iron in a water/methanol solution. Environ. Dissociation characteristics of light absorbed dioxin analogues calculated by
Sci. Technol. 38, 5208–5216. density functional theory. Environ. Sci. 19, 507–515.
Lu, D., Wang, D., Ip, H.S., Barley, F., Ramage, R., She, J., 2012. Measurements of Yoshimura, T., 2003. Yusho in Japan. Ind. Health 41, 139–148.
polybrominated diphenyl ethers and polychlorinated biphenyls in a single drop Zhang, G., Hua, I., 2000. Cavitation chemistry of polychlorinated biphenyls:
of blood. J. Chromatogr. B 891, 36–43. decomposition mechanisms and rates. Environ. Sci. Technol. 34, 1529–1534.
MEPPRC, 1991. Control Standard on Polychlorinated Biphenyls for Wastes. China Zhang, H., Zhang, J., Zhu, Y., 2008. In vitro investigations for the QSAR mechanism of
Standards Press, Beijing. lymphocytes apoptosis induced by substituted aromatic toxicants. Fish
Murena, F., Schioppa, E., 2000. Kinetic analysis of catalytic hydrodechlorination Shellfish Immunol. 25, 710–717.
process of polychlorinated biphenyls (PCBs). Appl. Catal. B: Environ. 27, 257– Zhou, W., Zhai, Z., Wang, Z., Wang, L., 2005. Estimation of n-octanol/water partition
267. coefficients (Kow) of all PCB congeners by density functional theory. J. Mol.
Qi, Z., Buekens, A., Liu, J., Chen, T., Lu, S., Li, X., Cen, K., 2014. Some technical issues in Struct.: Theo. 755, 137–145.
managing PCBs. Environ. Sci. Pollut. Res. 21, 6448–6462.

Vous aimerez peut-être aussi