Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Applied Surface Science 326 (2015) 66–72

Contents lists available at ScienceDirect

Applied Surface Science


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

Catalytic decomposition of H2 O2 over Fe-based catalysts for


simultaneous removal of NOX and SO2
Xianming Huang a,b , Jie Ding a,b , Qin Zhong a,b,∗
a
School of Chemical Engineering, Nanjing University of Science and Technology, Nanjing 210094, Jiangsu, PR China
b
Nanjing AIREP Environmental Protection Technology Co., Ltd, Nanjing 210091, Jiangsu, PR China

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

Article history: Simultaneous flue gas desulfurization and denitrification were achieved with • OH radicals from the
Received 26 August 2014 decomposition of H2 O2 over hematite (Fe) as well as hematite supported on alumina (Fe–Al) and anatase
Received in revised form (Fe–Ti). Under all conditions, SO2 achieved 100% removal, whereas NOX removal varies with the cata-
12 November 2014
lysts. The supporting of Fe over aluminum enhances the catalytic removal of NOX , whereas that of anatase
Accepted 18 November 2014
presents negative effect. The NOX removal is determined by the decomposition rate of H2 O2 into • OH radi-
Available online 24 November 2014
cals over OH bonded with Fe (Fe–OH). The supporting of Fe over alumina enhances the content of Fe–OH
and the points of zero charge (PZC) values, which are beneficial for the production of • OH radicals. The
Keywords:
H2 O2 decomposition
supporting of Fe over anatase results in the formation of FeOTi, which cannot decompose H2 O2 into • OH
Points of zero charges radicals. Furthermore, H2 O2 tends more to be reacted with TiOH to produce O2 over Fe–Ti. Finally, the
Fe-based catalyst enhancement mechanism of H2 O2 decomposition over Fe-based catalysts is speculated. It has a contri-
Denitrification bution to the correct choice for supports and active ingredients of the catalyst in the future industrial
Desulfurization applications.
Flue gas purification © 2014 Elsevier B.V. All rights reserved.

1. Introduction prepared a series of Cr based catalysts for selective catalytic oxida-


tion of nitric oxide (NO), which was simultaneously removed with
Nitrogen oxides (NOX ) and sulfur dioxide (SO2 ) emitted from SO2 in the subsequent absorption process. Khan and Adewuyi [16]
coal-fired power plants have caused considerable environmental investigated the oxidant oxidation–absorption of NO and SO2 by
and health problems [1–3]. Flue gas treatment for NOX and SO2 aqueous solution of sodium persulfate in a bubble column reactor.
has attracted a growing number of attentions [4–6]. But most of Wang et al. [17] utilized oxidation–absorption with the injection
them paid attention on calcium-based wet flue gas desulfuriza- of ozone and a glass made alkaline washing tower. However, these
tion (WFGD-Ca) and ammonia-based selective catalytic reduction above methods always suffer from high operation costs or unsatis-
(NH3 -SCR) [7–9]. Although the combination of WFGD-Ca and NH3 - fied removal efficiency, which limits its wide utilizations. Therefore,
SCR process has achieved industrial application on a large scale it is urgently to develop a low-cost and high-efficiency technology.
for flue gas treatment in coal-fired power-plants, the large and In previous studies, researchers have utilized H2 O2 for desul-
complex systems and the high capital and operating costs limit furization and denitrification through Fenton pathway. Liu et al.
its utilization in the developing world [10–12]. Many technolo- [18,19] photocatalyzed H2 O2 into • OH radicals under UV light,
gies have been developed to simultaneously remove NOX and SO2 which was to remove NOX and SO2 simultaneously. Guo et al. [20]
due to great advantages in low capital and operating costs and and Zhao et al. [21] used iron ions and H2 O2 mixed solutions for
simultaneous removal of multiple pollutants [12–14]. Among these absorption-removal for NOX and SO2 and high removal efficiency
simultaneously removal processes, oxidation–absorption process was obtained. Theses process all used absorption method, which
has prospects in industrial applications due to the simple and has many disadvantages such as the unregeneration of the absorp-
low-cost alteration from the conventional process. Cai et al. [15] tion liquids, high consummation of H2 O2 and difficult separation
of catalysts and solutions. In order to overcome these defects, a
novel process for oxidation of NOX and SO2 with • OH radicals from
the decomposition of H2 O2 over Fe-based catalysts, followed by
∗ Corresponding author at: School of Chemical Engineering, Nanjing University of
the absorption with ammonia-based washing tower has been pro-
Science and Technology, Nanjing 210094, Jiangsu, PR China. Tel.: +86 25 84315517;
fax: +86 25 84315517. posed by our previous works [22,23]. In this process, much smaller
E-mail address: zq304@mail.njust.edu.cn (Q. Zhong). amount of H2 O2 were injected into fixed-bed reactor and mixed

http://dx.doi.org/10.1016/j.apsusc.2014.11.088
0169-4332/© 2014 Elsevier B.V. All rights reserved.
X. Huang et al. / Applied Surface Science 326 (2015) 66–72 67

with transition metal oxides largely declining the capital invest-


ment and operating cost, which had a broad application prospects.
In the present study, the catalyst was found to play a critical effect
on the production of • OH radicals. Therefore, it is significant to
study a highly effective catalyst and concerning theories. In recent
years, iron oxide catalysts have been paid much attention for H2 O2
decomposition due to relatively low cost, nontoxicity, and envi-
ronmentally friendly character [24,25]. Most of papers focused on
the effect of processing conditions and surface groups in solutions
[26–30], but few of them have reported the critical factors for H2 O2
decomposition into • OH radicals and the corresponding mecha-
nisms in the system like this novel. Furthermore, higher active
catalysts should be developed for the wide application of this pro-
cess.
Herein, the preparation of Fe as well as Fe–Al and Fe–Ti were Fig. 1. XRD patterns of Fe, Fe–Al and Fe–Ti (vertical lines indicate PDF#33-0664).
carried out, with the aim of evaluating their performance as H2 O2
decomposition catalysts for the simultaneously removal of NOX and
2.3. Data process
SO2 . For that purpose, the effect of the critical factors such as the
electron cloud density, valence of metal oxide, acid sites and PZC
NOX and SO2 removal efficiencies at different conditions were
values on catalytic desulfurization and denitrification was investi-
obtained after 1 h at steady state. The removal efficiency was
gated, and the possible catalytic mechanism over Fe-based catalysts
obtained by using the equation:
was preliminarily speculated.
Cin − Cout
= × 100% (1)
Cin
2. Experimental  is the removal efficiency (%), Cin and Cout are the inlet and
outlet of NOX or SO2 content, respectively.
2.1. Experimental apparatus
2.4. Characterization
The experimental system for simultaneous desulfurization and
denitrification in the duct is reported by our previous paper [22]. 2.4.1. Decomposition rate of H2 O2
The duct is a quartz tube (I.D. 20 mm, length 500 mm). The sim- The decomposition rate of H2 O2 was determined by mass titra-
ulated flue gas was prepared by mixing N2 and small amount of tion method with potassium permanganate (KMnO4 ) [32]. The
concentrated NO gas and SO2 . • OH radicals from the decomposition consumed KMnO4 was used to represent the decomposition rate
of H2 O2 over Fe-based catalyst were injected into the center of the of H2 O2 in the solution.
duct, which oxidized NOX or SO2 to HNO3 or H2 SO4 , respectively.
Finally, these oxidation products and remaining contaminants
2.4.2. Characterization of catalyst
were absorbed in the glass made ammonia-based washing tower.
The power X-ray diffraction (XRD) was carried out with Beijing
The absorbent is 0.05 wt% ammonia solutions and 1 L is used in
Purkinjie general instrument XD-3 X-ray diffraction (CuK␣, voltage
one test. The H2 O2 flow is 0.4 mL min−1 . The purged gas flow is
35 kV, electrical current 20 mA, 2 from 5◦ to 80◦ ). Specific surface
40 mL min−1 , and the total gas flow is 240 mL min−1 . NOX content is
area was measured using N2 adsorption at −196◦ C and determined
550 ppm, and SO2 content is 2000 ppm. The catalyst in the fix-bed is
by BET method using a V-Sorbet 2800S automated gas sorption
0.6 g.
system. X-ray photoelectron spectroscopy (XPS) was carried out
on an RBD upgraded PHI-5000C ESCA system (Perkin Elmer) with
Mg K␣ radiation (h = 1253.6 eV), calibrated by the C 1s peak at
2.2. Preparation of catalysts 284.6 eV with an accuracy of 0.1 eV. The PZC values of catalysts
were measured with a mass titration method [33,34].
The Anatase was prepared by precipitation method with Ammonia (NH3 ) temperature-programmed desorption (NH3 -
Ti(SO4 )2 (Sinopharm Chemical Reagent Co., Ltd, AR) and 25 wt% TPD) was performed in a flow of He (70 cm3 min−1 ) over 100 mg of
ammonia (Sinopharm Chemical Reagent Co., Ltd., 25 wt%) [31]. catalyst using a heating rate of 10 ◦ C min−1 . The NH3 adsorption was
The 25 wt% ammonia was gradually added into Ti(SO4 )2 solu- conducted with the flow rate of 70 mL min−1 at room temperature.
tions under stirring agitation until the pH reaches 9. The anatase The surface acidity of catalysts was investigated by means of
was obtained after filtration, drying and calcination at 500 ◦ C. The a Fourier transform infrared spectroscopic (FTIR) study of NH3
hematite was prepared by Fe(NO3 )3 ·6H2 O (Sinopharm Chemical adsorption. NH3 adsorbed FTIR spectra (NH3 -FTIR) were recorded
Reagent Co., Ltd, AR) through calcining at 500 ◦ C [22]. The alu- in an IS10 FTIR spectrometer. The samples were absorbed by NH3
mina (␥-phase alumina) was purchased from Shanghai Yuejiang for 5 h. Then, the spectra (200 scans, resolution: 1 cm−1 , range
Co., Ltd. of acquisition: 650–4000 cm−1 ) were recorded after evacuation
The hematite supported on anatase and alumina was impreg- (10−5 Torr or 0.0013 Pa) for 30 min.
nated by incipient wetness with aqueous solution of 25% (mol/mol)
ferric nitrate [23]. The anatase or alumina was mixed in the dis- 3. Results and discussions
tilled water, and suitable amount of Fe(NO3 )3 ·6H2 O was gradually
added into the solutions. The mixtures were vigorous stirred at 3.1. Structure characterizations
80 ◦ C until drying. The hematite supported on anatase and alu-
mina were obtained after drying at 120 ◦ C for 24 h and calcined at XRD patterns of Fe, Fe–Al and Fe–Ti have been shown in Fig. 1.
500 ◦ C for 3 h. The hematite supported on the anatase, alumina and Within the detection limits of this technique, hematite (PDF#33-
hematite catalyst is denoted as Fe–Ti, Fe–Al and Fe, respectively. 0664) has been detected in all samples [35,36]. The pattern peaks
68 X. Huang et al. / Applied Surface Science 326 (2015) 66–72

Table 1
Amount of KMnO4 consumed by different catalysts.

Catalysts Amount of consumed KMnO4 (mL)

Run 1 Run 2 Run 3

Fe–Al 2.56 2.49 2.43


Fe 3.21 3.15 3.12
Fe–Ti 3.76 3.72 3.67

were conducted as shown in Fig. S4. It can be obtained that the


NOX removal apparently decreases when the isopropanol, a kind
of typical radical annihilation agent, is injected into the system,
indicating that the NOX removal is proportional to the • OH concen-
trations. The decomposition rate of H2 O2 over catalysts has also
Fig. 2. NOX and SO2 removal efficiencies with Fe, Fe–Al and Fe–Ti. NOX content is
550 ppm, SO2 content is 2000 ppm, H2 O2 flow is 0.4 mL min−1 , the flow of purged been measured by aqueous KMnO4 solution titration method. The
gas is 40 mL min−1 , the total gas flow is 240 mL min−1 . results are shown in Table 1. The more KMnO4 consumed means the
lower decomposition rate of H2 O2 . The decomposition rate of H2 O2
of Fe–Al is apparently weakened, illustrating that the alumina pre- decreases as the following order: Fe–Al > Fe > Fe–Ti, conceding with
vents the crystallization of hematite. Some previous papers have the results of NOX removal and • OH radical concentration. Combin-
reported that the hematite–alumina compositions were formed on ing with the above facts, it can be deduced that the NOX removal
the surface, which was the reason for the weakening of pattern is mainly determined by the H2 O2 decomposition over catalysts
peaks. Compared to Fe–Al, only very weak characteristic peaks of into • OH radicals. The surface properties such as pore structures,
hematite were observed for Fe–Ti, maybe resulted by the better acid sites, PZC values and surface hydroxyl groups ( OH) play the
dispersion of hematite on the surface of TiO2 . As shown in Fig. important role in the decomposition rate of H2 O2 [42,43]. How-
S1, TiO2 is mainly composed of anatase. Compared to anatase, the ever, results of SEM and BET characterizations have indicated that
peak intensity of Fe–Ti is dramatically reduced, indicating that the the surface structures of Fe-based catalysts have little effect on the
hematite also has an inhibition on the crystallization of anatase. H2 O2 decomposition.
These two facts indicate the possible formation of Fe O Ti linkage
bonds. 3.3. Effect of surface acid sites
SEM images of Fe–Al and Fe–Ti are shown in Fig. S2. The catalysts
were irregular in shape, and larger aggregates were observed on NH3 -TPD experiments were performed over Fe-based catalysts
the surface of alumina. Combining with the results of XRD, it can to study the surface acid sites. Fig. S4 shows that the NH3 -TPD pro-
be inferred that the hematite tends to crystallize on the surface of files of the Fe-based catalysts. There is a strong desorption peak
alumina. at around 450 ◦ C in the NH3 -TPD profiles of Fe and Fe–Ti. Fe–Al
The specific surface areas of Fe, Fe–Al and Fe–Ti are measured by shows weaker desorption peak at approximately 350 ◦ C. It indi-
BET method as shown in Table S1. It can be obtained from Table S1 cates that the supporting of alumina apparently reduces the acidity
that the surface areas of these three catalysts are small and remain of the catalyst. The nature of acid sites (Lewis and Brønsted) on
almost constant, illustrating that the surface areas are not the main the surface of Fe-based catalysts has been studied by NH3 -FTIR.
reason affecting the catalytic activities. The spectra arising from NH3 adsorption at room temperature fol-
Fig. S3 shows the FTIR spectra of TiO2 , Al2 O3 , Fe, Fe–Al and Fe–Ti. lowed by evacuation at the same temperature are shown in Fig. 3.
It is worth noting that two apparent absorption peaks are observed The band at about 1600 cm−1 has been assigned to the vibration
at around 500 cm−1 for Fe, corresponding to the Fe O Fe lattice modes of Lewis site coordinated NH3 [44–46]. Similarly, the band at
movements in crystals [37]. However, the Fe O Fe movements about 1500 cm−1 corresponds to a NH4 + bonded to a Brønsted site
for Fe–Ti is much weaker maybe due to the formation of Fe O Ti [44,45,47]. Comparing with the Fe, the Lewis acid sites increases
linkage bonds [38]. The apparent widening of the peak ranging from apparently for Fe–Ti, but remains almost constant for Fe–Al. It is
450 to 1250 cm−1 further indicates the formation of Fe O Ti link- clear that a main effect for Fe–Ti is to create more Lewis acid sites
age bonds because of the overlapped vibrations of Fe O Ti and on the catalyst surface.
Fe O Fe [39]. Furthermore, a weak peak at 1037 cm−1 is found,
maybe attributing the bending vibrations of MOH (M is Fe, Ti or Al)
[40,41].

3.2. Effect of supports on removal efficiency

The NOX and SO2 removal as a function of the simulated flue


gas temperature are shown in Fig. 2. It can be seen that SO2
achieved 100% removal under all conditions as indicated by our
previous works [22,23], whereas the NOX removal increases with
the temperature rising. It is most probably caused by the increase
of molecular activity at higher temperature, thus promoting these
oxidation reactions. As shown in Fig. 2, the Fe achieves 71% NOX
removal. The Fe–Al obtains a slightly higher NOX removal, about
74%, whereas the NOX removal of Fe–Ti is a little lower, about 68%.
The variation of NOX removal may be directly related to the concen-
tration of • OH radicals. To confirm such relationship between NOX
removal and • OH radicals, the isopropanol injecting experiments Fig. 3. FTIR spectra of NH3 adsorbed Fe, Fe–Al and Fe–Ti.
X. Huang et al. / Applied Surface Science 326 (2015) 66–72 69

3.4. Effect of surface properties and transfer electron to TiOH as shown in the following equations
[54,55]:
3.4.1. XPS
Ti(III)OH + H2 O2 → (H2 O2 )S
Surface information on catalysts was analyzed by XPS character-
ization. XPS spectra over the spectral regions of Fe2p and O1s were
evaluated (as shown in Fig. 4). For the Fe and Fe–Al, the binding (H2 O2 )S → H2 O + Ti(IV) + HO2 •
energies centered at about 709.7 and 710.4 eV can be assigned to The loss of surface OH on surface of TiO2 after reaction
Fe2+ and Fe3+ in FeOFe structure (FeOFe), respectively [48,49]. The decreases the density of electron cloud. However, no apparent vari-
binding energy centered about 712 eV is attributed to Fe3+ in FeOH ation can be observed in Al2 O3 .
[50,51]. The binding energies of Fe species for Fe–Ti are splitted
into four peaks at 708.9, 710.1, 711.4 and 712.7 eV, which is cor- 3.4.2. FTIR and XRD
responding to Fe2+ , Fe3+ in FeOTi, Fe3+ in FeOFe and Fe3+ in FeOH, Fig. S7 shows the FTIR spectra of fresh and reacted Fe, Fe–Al
respectively. Compared to Fe, the Fe2+ and Fe3+ in FeOFe for Fe–Ti and Fe–Ti. The number and location of the peaks have no changes,
shift to lower binding energies, but Fe3+ in FeOTi and in FeOH shift indicating the stability of the Fe-based catalysts. It is noted that the
to higher binding energies, indicating that the Fe atoms for Fe2+ and peak of MOH (M is Fe, Ti or Al) bending vibrations is apparently
Fe3+ in FeOTi show higher density of electron cloud, whereas Fe3+ intensified in Fe and Fe–Al, indicating that the MOH is generated
in FeOTi and in FeOH show lower density of electron cloud around during the H2 O2 treatment, which is consistent with the results of
Fe atoms. The lower density of electron cloud should benefit the XPS characterization. XRD patterns of fresh and reacted Fe, Fe–Al
electron transfer from the H2 O2 to Fe3+ , thus favoring the genera- and Fe–Ti are shown in Fig. S8. Only the peak intensities are varied,
tion of • OH radicals. But the results of this paper indicate that the whereas no changes for peak number and locations are observed,
electron cloud around Fe atoms has less effect on the generation of further confirms the stability of these catalysts in reaction process.
• OH radicals.

As shown in Fig. 4, for Fe, the O peaks are mainly centered at


3.4.3. PZC values
528.82 eV, as expected for the transition metal oxides. Another
The production of • OH radicals plays an important role in
oxygen species centered at 530.48 eV is assigned to O in hydroxyl
the NOX removal, which is closely related to the surface proper-
group ( OH) [52,53]. For Fe–Al, the O peak of transition metal
ties of catalyst, such as PZC values. PZC values, as an important
oxides shifts to lower binding energies, which may be one of rea-
parameter of the catalyst surface, may affect the generation of
sons for the decline of acid sites especially Lewis acid sites as • OH radicals because titania, alumina and hematite exhibit dif-
indicated by Fig. 3. For Fe–Ti, both O peaks shift to higher bind-
ferent PZC values, which decrease as the following order [56]:
ing energies, indicating lower density of electron cloud around
Al2 O3 > Fe2 O3 > 7 > TiO2 . As shown in Fig. 3, the NOX removal effi-
O atoms, conceding with the results of NH3 -FTIR as shown in
ciency of Fe is close to that of Fe–Al, but Fe–Ti has much lower
Fig. 3.
NOX removal. From XRD characterizations, the characteristic peaks
Table 2 shows that the content of Fe and O species in Fe, Fe–Ti
of hematite in Fe and Fe–Al are clearly observed. However, those
and Fe–Al. It can be obtained that the Fe–Ti has the highest con-
peaks of hematite in Fe–Ti are much weaker. It indicates that the
tent of surface -OH, but lowest content of FeOH. The Fe has the
effect of hematite on surface properties may be weakened, but that
lowest content of surface OH, whereas the Fe–Al has the high-
of supports increases. Results of mass titration show that the PZC
est content of FeOH. It indicates that most of surface OH is on
values of Fe and Fe–Al (>9) are higher than those of Fe–Ti (<5.5). The
the TiO2 surface for Fe–Ti. It has reported that the MOH was the
state of surface OH will be varied with the PZC values. The cata-
active sites in producing the • OH radicals. However, experimen-
lyst surface is in equilibrium by the surface coordination reactions
tal results as shown in Fig. 2 indicate that the NOX removal with
according to the following two equations:
Fe–Ti is the lowest, but that of Fe–Al is the highest. Furthermore,
as shown in Fig. S5, the Fe–Ti presents much higher O2 content. MeOH + H+ → MeOH2 + (pH < PZC)
It demonstrates that H2 O2 tends to be mainly decomposed into
• OH radicals over FeOH, whereas decomposed into O over TiOH.
2 MeOH + OH− → MeO− + H2 O (pH > PZC)
This may be caused by the weak oxidizing properties of Ti4+ or
Ti3+ . When the pH of aqueous solution is near the PZC of the catalyst,
As shown in Fig. S6, comparing with the fresh and reacted Fe–Ti, most of the surface OH are in a neutral state, namely nearly no
the content of surface OH decreases dramatically, whereas the surface charge exists. Otherwise, the catalyst surface becomes pro-
content of FeOH increases. It demonstrates that the TiOH is the tonated or deprotonated when the pH of aqueous solution is below
main H2 O2 reactive sites in Fe–Ti, further confirming that the O2 or above the PZC values [57]. At circumneutral pH values (pH 5.5),
is produced on TiOH sites. The high FeOH content in reacted Fe–Ti it is expected that the surfaces of Fe–Ti will be negatively charged,
reveals the low reactivity of FeOH in Fe–Ti as demonstrated by the whereas Fe and Fe–Al surfaces will be positively charged. The H2 O2
low NOX removal. is a kind of typical electron donor, and more tended to be adsorbed
Compared to the fresh Fe–Al, the content of Fe2+ and Fe3+ in on the positively charged surface. The adsorption of H2 O2 on the
FeOFe for reacted Fe–Al is slightly higher, whereas the Fe3+ in catalyst has been reported to be the rate-determining step in H2 O2
FeOH is slightly lower. Combining with the results of the above decomposition [58].
mentioned, it can deduce that the FeOH is the active sites for the
H2 O2 decomposition into • OH radicals. The content of FeOH is not 3.5. Enhancement mechanism
declined apparently maybe due to the easily conversion from Fe2+
and Fe3+ to FeOH. However, for Fe–Ti, the content of Fe3+ in FeOTi From the experimental results and the theory motioned above,
remains almost constant, but the Fe3+ in FeOFe declines dramati- the enhancement mechanism of H2 O2 decomposition on the
cally, indicating that the FeOH is mainly conversed by Fe3+ in FeOFe hematite or hematite over supports is proposed and illustrated in
not the Fe3+ in FeOTi. As shown in Fig. S6 and Table S2, the Ti3+ of Fig. 5.
fresh sample is lower than that of reacted one, and the binding First, with the injection of H2 O2 solution into the fixed bed. The
energy of fresh sample tends to be lower than those of the reacted H2 O2 solution is mixed with Fe-based catalyst. The surface pro-
one. It can be deduced that the H2 O2 may adsorb on the TiOH tonated or deprotonated hydroxyl groups will be formed. Then the
70 X. Huang et al. / Applied Surface Science 326 (2015) 66–72

Fig. 4. XPS spectra of Fe, Fe–Ti and Fe–Al over the spectral regions of Fe2p and O1s.

chemisorbed OH characterize the oxide/water interface as shown FeOH in Fe–Al benefits the production of • OH radicals, and the TiOH
in state 2. As shown in Fig. S8, the peak of MOH for Fe, Fe–Al and in Fe–Ti tends to produce O2 . Compared with FeOH, H2 O2 tends
Fe–Ti before and after H2 O treatment has no apparent changes, more to react with TiOH in Fe–Ti, thus decreasing the production of
indicating that the OH may be formed with H2 O2 . • OH radicals. It has been shown in state 3. On the other hand, the PZC

The hematite supported on supports can cause the variation of values increase with the supporting of alumina, but decreases with
catalyst surface characteristics. On the one hand, the high content of the supporting of anatase. Specifically, the PZC values of hematite

Table 2
XPS results of Fe, Fe–Ti and Fe–Al.

Samples Surface atomic concentration

Fe O
2+ 3+ 3+
Fe Fe /FeOFe Fe /FeOTi Fe/FeOH O/FeOFe O/FeOH

Fe 0.30 0.42 0.28 0.73 0.27


Fresh Fe–Al 0.38 0.27 0.35 0.69 0.31
Fresh Fe–Ti 0.26 0.23 0.23 0.28 0.50 0.50
Reacted Fe–Al 0.36 0.26 0.38 0.68 0.32
Reacted Fe–Ti 0.24 0.22 0.13 0.40 0.69 0.31
X. Huang et al. / Applied Surface Science 326 (2015) 66–72 71

Fig. 5. Proposed enhancement mechanisms of H2 O2 decomposition on the surface of Fe-based catalysts.

supported on alumina are higher than pH 5.5 of H2 O2 solutions, supports and active ingredients should make a correct choice to
but which of hematite supporting on anatase are lower than 5.5. obtain a high NOX removal efficiency.
Combining with Eqs. (1) and (2) with the results obtained above, it
can be concluded that the surface OH2 + is the predominant surface
groups formed under the present experimental conditions for Fe–Al 4. Conclusions
and Fe. The surface OH2 + benefits the adsorption of H2 O2 due to
the high nucleophilicity of H2 O2 , which accelerates the generation Hematite supported on alumina benefits the enhancement of
of • OH radicals. It is as shown in step 4–6. FeOH and the increase of PZC values. Both are favor of • OH radical
Second, the hematite tends to form Fe O Ti with anatase. And production, which is conductive to the NOX removal. And the higher
the Fe3+ in Fe O Ti is stable and difficult to be transfered into FeOH, PZC values for Fe–Al enhance the adsorption of H2 O2 on FeOH2 + ,
thus inhibiting the generation of • OH radicals. It has been shown in further increasing the production of • OH radicals. Hematite sup-
step 3. In addition, the reaction between FeOH and H2 O2 in Fe–Ti is ported on anatase prevents Fe3+ in FeOTi from transforming into
much slower than that in Fe–Al as shown in step 3. Therefore, the FeOH, and the H2 O2 tends more to react with TiOH to generate
generation of • OH in Fe–Ti is much slower than that in Fe–Al. O2 . The mechanism for H2 O2 decomposition is preliminarily spec-
ulated in this paper. The decrease of • OH radicals inhibits the NOX
removal for Fe–Ti. This paper contributes to the correct choice for
3.6. Environmental implications supports and active ingredients in the catalyst.

The alumina has the potential to be more effective as the sup-


ports for oxidative treatment of contaminated flue gas with H2 O2 Acknowledgements
solution comparing with the anatase. While large amounts of H2 O2
is catalytic decomposed into oxidants capable of oxidizing and This work was finally supported by the National Natural
removing NOX and SO2 . The oxidation products were analyzed Science Foundation of China (U1162119), the Assembly Foun-
by ion chromatography and the material balance were made as dation of the Industry and Information Ministry of the People’s
shown in Fig. S10 and Table S3. As shown in Fig. S10, NO3 − , SO3 2− Republic of China 2012 (543), Research and Innovation Plan
and SO4 2− are present in the solutions, which indicates that • OH for Postgraduates of Jiangsu Province (CXZZ13 0215), Research
radicals have sufficient strength to oxidize NOX to HNO3 . It has Fund for Scientific Research Project of Environmental Protection
been specifically explained by our previous work [22]. As shown Department of Jiangsu Province (2013003) and (201212), Industry-
in Table S3. There are no other byproducts except for NO3 − , SO3 2− Academia Cooperation Innovation Fund Projects of Jiangsu
and SO4 2− produced in this process. Therefore, minimal waste pro- Province (BY2012025) and Scientific Research Project of Environ-
duction and the resource of contaminant gas in this process may mental Protection Department of Jiangsu Province (201112).
provide advantages over other approaches. Additional research is
needed to further enhance the efficiency of the catalyst and assess
the scaling up of the treatment systems employing the catalyst. Appendix A. Supplementary data
This study has important implications for the design of the
industrial catalyst. Previous studies points out that the MOH is the Supplementary data associated with this article can be
active sites in producing • OH radicals. And this paper reveals that found, in the online version, at http://dx.doi.org/10.1016/j.apsusc.
not all MOH can promote the generation of • OH radicals, thus, the 2014.11.088.
72 X. Huang et al. / Applied Surface Science 326 (2015) 66–72

References [31] H.Y. Huang, R.T. Yang, Langmuir 17 (2001) 4997–5003.


[32] N.V. Klassen, D. Marchington, H.C.E. McGowan, Anal. Chem. 66 (1994)
[1] J.D. Felix, E.M. Elliot, S.L. Shaw, Environ. Sci. Technol. 46 (2012) 3528–3535. 2921–2925.
[2] G.E. Marnellos, M.A. Efthimiadis, I.A. Vasalos, Appl. Catal. B: Environ. 48 (2004) [33] J.S. Noh, J.A. Schwarz, J. Colloid Interface Sci. 130 (1989) 157–164.
1–15. [34] J.S. Noh, J.A. Schwarz, Carbon 28 (1990) 675–682.
[3] A. Venkatesh, P. Jaramillo, W.M. Griffin, H.S. Matthews, Environ. Sci. Technol. [35] D. Wodka, R.P. Socha, E. Bielańska, M. Elźbieciak-Wodka, P. Nowak,
46 (2012) 9838–9845. P. Warszyński, Appl. Surf. Sci. (2014), http://dx.doi.org/10.1016/j.apsusc.
[4] Y.K. Kwon, D.H. Han, Ind. Eng. Chem. Res. 49 (2010) 8147–8156. 2014.08.010 (in press).
[5] G. Hu, Z. Sun, H. Gao, Environ. Sci. Technol. 44 (2010) 6712–6717. [36] P.R. Palacios, A. Bustamante, P. Romero-Gómez, J.C. González, Hyperfine Inter-
[6] G. Chen, J. Gao, J. Gao, Q. Du, X. Fu, Y. Yin, Y. Qin, Ind. Eng. Chem. Res. 49 (2010) act. 203 (2011) 113–118.
12140–12147. [37] J.L. Rendon, C.J. Serna, Clay Miner. 16 (1981) 375–381.
[7] D.W. Fickel, E. D’Addio, J.A. Lauterbach, R.F. Lobo, Appl. Catal. B: Environ. 102 [38] F. Liu, K. Asakura, H. He, Y. Liu, W. Shan, X. Shi, C. Zhang, Catal. Today 164 (2011)
(2011) 441–448. 520–527.
[8] P. Forzatti, I. Nova, E. Tronconi, Ind. Eng. Chem. Res. 49 (2010) 10386–10391. [39] L. Chen, B. He, S. He, T. Wang, C. Su, Y. Jin, Powder Technol. 227 (2012) 3–8.
[9] L. Chen, J. Li, M. Ge, Environ. Sci. Technol. 44 (2010) 9590–9596. [40] S.J. Hug, J. Colloid Interface Sci. 188 (1997) 415–422.
[10] M.T. Makhlouf, B.M. Abu-Zied, T.H. Mansoure, Appl. Surf. Sci. 274 (2013) 45–52. [41] C.X. Gao, Q.F. Liu, D.S. Xue, J. Mater. Sci. Lett. 21 (2002) 1781–1783.
[11] T. Huang, S. Hsu, C. Wu, Environ. Sci. Technol. 46 (2012) 2324–2329. [42] M. Hermanek, R. Zboril, I. Medrik, J. Pechousek, C. Gregor, J. Am. Chem. Soc. 129
[12] H. Deng, H. Yi, X. Tang, H. Liu, X. Zhou, Ind. Eng. Chem. Res. 52 (2013) 6778–6784. (2007) 10929–10936.
[13] S.O. Owusu, Y.C. Adewuyi, Ind. Eng. Chem. Res. 45 (2006) 4475–4485. [43] C. Zhang, L. Wang, F. Wu, N. Deng, Environ. Sci. Pollut. Res. 13 (2006) 156–160.
[14] J.L. de Paiva, G.C. Kachan, Ind. Eng. Chem. Res. 37 (1998) 609–614. [44] N.Y. Topsøe, H. Topsøe, J.A. Dumesic, J. Catal. 151 (1995) 226–240.
[15] W. Cai, Q. Zhong, S. Zhang, J. Zhang, RSC Adv. 3 (2013) 7009–7015. [45] T. Barzetti, E. Selli, D. Moscotti, L. Forni, J. Chem. Soc., Faraday Trans. 92 (1996)
[16] N.E. Khan, Y.G. Adewuyi, Ind. Eng. Chem. Res. 49 (2010) 8749–8760. 1401–1407.
[17] Z. Wang, J. Zhou, Y. Zhu, Z. Wen, J. Liu, K. Cen, Fuel Process. Technol. 88 (2007) [46] N.Y. Topsøe, J. Catal. 128 (1991) 499–511.
817–823. [47] Z. Huang, P. Wu, B. Gong, Y. Lu, N. Zhu, Z. Hu, Appl. Surf. Sci. 286 (2013)
[18] Y. Liu, J. Zhang, C. Sheng, Y. Zhang, L. Zhao, Chem. Eng. J. 162 (2010) 1006–1011. 371–378.
[19] Y. Liu, J. Zhang, J. Pan, A. Tang, Energy Fuels 26 (2012) 5430–5436. [48] C.J.G. Peter, M.A.J. Somers, Appl. Surf. Sci. 100/101 (1996) 36–40.
[20] R. Guo, W. Pan, X. Zhang, J. Ren, Q. Jin, H. Xu, J. Wu, Fuel 90 (2011) 3295–3298. [49] X. Hu, J.C. Yu, J. Gong, Q. Li, G. Li, Adv. Mater. 19 (2007) 2324–2329.
[21] Y. Zhao, X. Wen, T. Guo, J. Zhou, Fuel Process. Technol. 128 (2014) 54–60. [50] S. Yang, Y. Guo, N. Yan, Z. Qu, J. Xie, C. Yang, J. Jia, J. Hazard. Mater. 186 (2011)
[22] J. Ding, Q. Zhong, S. Zhang, F. Song, Y. Bu, Chem. Eng. J. 243 (2014) 176–182. 508–515.
[23] J. Ding, Q. Zhong, S. Zhang, RSC Adv. 4 (2014) 5394–5398. [51] S. Yang, Y. Guo, N. Yan, N. Yan, D. Wu, H. He, Z. Qu, J. Jia, Ind. Eng. Chem. Res. 50
[24] Z.R. Lin, X.H. Ma, L. Zhao, Y.H. Dong, Chemosphere 101 (2014) 10–20. (2011) 9650–9656.
[25] J. Hong, S. Lu, C. Zhang, S. Qi, Y. Wang, Chemosphere 84 (2014) 1542–1547. [52] Y. Zhang, M. Yang, X.M. Dou, H. He, D.S. Wang, Environ. Sci. Technol. 39 (2005)
[26] A. Anfruns, M.A. Montes-Morán, R. Gonzalez-Olmos, M.J. Martin, Chemosphere 7246–7253.
91 (2013) 48–54. [53] T. Herranz, M. Rojas, M. Ojeda, F.J. Perez-Alonso, J.L.G. Fierro, Chem. Mater. 18
[27] D. Xiao, Y. Guo, X. Lou, C. Fang, Z. Wang, J. Liu, Chemosphere 103 (2014) (2006) 2364–2375.
354–358. [54] P. Tenjvall, I. Lundström, L. Sjöqvist, H. Elwing, Biomaterials 10 (1989)
[28] S. Valizadeh, M.H. Rasoulifard, M.S. Seyed Dorraji, Appl. Surf. Sci. (2014), 166–175.
http://dx.doi.org/10.1016/j.apsusc.2014.07.139 (in press). [55] A. Corma, P. Esteve, A. Martinez, S. Valencia, J. Catal. 152 (1995) 18–24.
[29] D. Xu, F. Cheng, Q. Lu, P. Dai, Ind. Eng. Chem. Res. 53 (2014) 2625–2632. [56] M. Kosmulski, J. Colloid Interface Sci. 253 (2002) 253, 77–87.
[30] L. Cheng, M. Wei, L. Huang, F. Pan, D. Xia, X. Li, A. Xu, Ind. Eng. Chem. Res. 53 [57] T. Zhang, J. Ma, J. Mol. Catal. A 279 (2007) 82–89.
(2014) 3478–3485. [58] W.P. Kwan, B.M. Voelker, Environ. Sci. Technol. 37 (2003) 1150–1158.

You might also like