Sie sind auf Seite 1von 10

Chemical Engineering Journal 344 (2018) 114–123

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Pd and Pt nanoparticles supported on the mesoporous silica molecular sieve T


SBA-15 with enhanced activity and stability in catalytic bromate reduction

Zhiqiang Zhanga, Yan Luob, Yuan Guoa, Wenxin Shia, , Wei Wanga, Bing Zhanga, Ruijun Zhanga,

Xian Baoa, Shuyan Wuc, Fuyi Cuid,
a
State Key Laboratory of Urban Water Resource and Environment, School of Municipal and Environmental Engineering, Harbin Institute of Technology, Harbin 150090,
People’s Republic of China
b
Wuhan Planning and Design Institute, Wuhan 430014, People’s Republic of China
c
China Nerin Engineering Co. Ltd Hainan Branch, Haikou 570125, People’s Republic of China
d
Urban Construction and Environmental Engineering, Chongqing University, Chongqing 400044, People’s Republic of China

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

• Pd and Pt supported on SBA-15 were


prepared and used in catalytic bro-
mate reduction.
• The Pt/SBA-15 catalysts exhibited
higher activities than the Pd/SBA-15
catalysts.
• Bromate reduction on the Pt surface
rather than the diffusion process con-
trolled the reaction rate.
• The 4% Pt/SBA-15 catalyst was reused
5 times without significant losses in
activity and Pt loading.

A R T I C LE I N FO A B S T R A C T

Keywords: Bromate is a highly carcinogenic disinfection by-product in drinking water. Catalytic reduction is an emerging
Bromate reduction technology that can effectively remove bromate from water without forming additional brines and sludges. In
Pt/SBA-15 catalyst this study, Pd and Pt nanoparticles supported on SBA-15 were prepared for catalytic bromate reduction. The
Activity activity investigation results showed that the Pt/SBA-15 catalysts exhibited higher activities than the Pd/SBA-15
Stability
catalysts and the commercial 5% Pd/C catalyst, and the highest catalytic activity of 24.72 mg·min−1·gcat−1 was
obtained using 4% Pt/SBA-15 as the catalyst. This activity is approximately 2.33- and 6.50-fold higher than the
activities of the 4% Pd/SBA-15 and 5% Pd/C catalysts, respectively. Analysis of the intrinsic mechanism revealed
that the activation energy of bromate reduction over the 4% Pt/SBA-15 catalyst was 47.31 kJ·mol−1, and the
reaction followed the Langmuir-Hinshelwood mechanism, suggesting that the reduction of adsorbed bromate on
the Pt surface rather than the diffusion process controlled the overall reaction rate. Catalytic bromate reduction
was strongly dependent on the reaction pH, and the catalytic activity increased with decreasing pH. Bromate
reduction was suppressed in the presence of coexisting anions, including PO43−, SO42−, and Cl−, and the in-
hibition effects followed the order of PO43− > SO42− > Cl−. Surprisingly, bromate removal was obviously
enhanced when a low concentration (< 1.5 mM) of nitrate coexisted in the reaction solution. In addition, the


Corresponding authors.
E-mail addresses: swx@hit.edu.cn (W. Shi), cuifuyi@hit.edu.cn (F. Cui).

https://doi.org/10.1016/j.cej.2018.03.056
Received 23 November 2017; Received in revised form 8 February 2018; Accepted 12 March 2018
Available online 13 March 2018
1385-8947/ © 2018 Elsevier B.V. All rights reserved.
Z. Zhang et al. Chemical Engineering Journal 344 (2018) 114–123

losses in catalytic activity and Pt after five repeated tests were only 7.8% and 2.6%, respectively, revealing the
high stability of the 4% Pt/SBA-15 catalyst in the catalytic reduction of bromate.

1. Introduction and structured porous materials (e.g., Al2O3, SiO2, AC, TiO2, and
MWCNT) have been used as the catalyst support, and investigation
Bromate is a representative disinfection byproduct generally formed results indicated that the activities of the monometallic catalysts were
from bromide oxidation during the ozonation or chlorination of water largely dependent on the support due to the different interactions be-
[1–4]. As a typical carcinogenic substance, bromate in drinking water tween the support and the metallic catalyst [28]. It has been demon-
poses a significant threat to human health [5]. Therefore, the World strated that hexagonally ordered mesoporous silicates have been con-
Health Organization (WHO) has stipulated that the maximum con- sidered to be excellent catalyst supports owing to their large BET
taminant level (MCL) of bromate in drinking water should not exceed surface areas, uniform ordered porosities and high chemical stabilities
0.01 mg/L [6]. To meet this standard, many methods have been ex- [33]. Among the mesoporous silica materials, SBA-15 is widely used as
tensively developed for removing bromate from drinking water. a catalyst support, especially for catalysts that are applied in aqueous
Although the classical physicochemical methods such as reverse catalytic reactions, profiting from its adjustable pore distribution (range
osmosis [7], ion exchange [8], electrodialysis [9], and adsorption [10] from 5 to 30 nm), thick pore walls (between 3.1 and 6.4 nm), and ex-
can be used for the removal of bromate with high efficiency, the col- cellent hydrothermal stability [34]. Many studies have shown that SBA-
lected bromate cannot be degraded by those processes, and the con- 15-based catalysts generally display a high catalytic activity and re-
centrated bromate brine produced by these methods requires further cyclability. For example, Bhuyan et al. [35] reported that a Pd@SBA-15
treatment, which would increase the processing cost and might cause a nanocomposite catalyst showed a high activity and selectivity in the
secondary environmental pollution [11,12]. Another process frequently hydrogenation of phenylacetylene, and its activity was well-maintained
used for bromate removal is biodegradation [13,14], which reduces even after four reaction cycles.
bromate to harmless bromide using glucose, pyruvate or starch as an Considering the outstanding properties of the mesoporous mole-
electron donor in a biological reactor under anaerobic conditions. Al- cular sieve SBA-15, we aimed to synthesize a highly active and stable
though the bromate can be thoroughly degraded, the additional of an metal/SBA-15 catalyst for the catalytic reduction of bromate in
organic carbon source, the low bromate degradation rate, and the risk drinking water. To achieve this aim, Pd and Pt nanoparticles supported
of microorganism limit the practical use of this technology. Alter- on SBA-15 materials were synthesized in this study, and the physico-
natively, a chemical reduction approach, such as zero-valent iron (Fe0) chemical properties of these catalysts were characterized by Fourier
reduction [15–17], can efficiently reduce bromate in water under mild transform infrared (FTIR) spectroscopy, specific surface area analysis
conditions with an appreciable treatment rate. However, Fe2+ would be (BET), transmission electron microscopy (TEM), powder X-ray diffrac-
released into the water during the reduction process, and additional tion (p-XRD), and zeta potential (ξ) measurements. Comparison of the
post-treatment processes are generally necessary for the removal of activity of these catalysts in liquid-phase catalytic bromate reduction
Fe2+ due to the strict regulation of Fe2+ concentration in drinking was performed and discussed. Furthermore, the mechanism of the
water. catalytic reaction and the effects of the pH and the presence of co-ex-
Recently, liquid-phase catalytic hydrogenation has emerged as one isting ions on bromate reduction were also comprehensively studied.
of the most efficient processes for removing a variety of priority organic Finally, the stability of the optimal catalyst was evaluated using cycling
and inorganic contaminants from water, including halogenated com- experiments.
pounds (e.g., trichloroethylene [18], perchloroethylene [19], and dia-
trizoate [20]), N-nitrosamines (e.g., N-nitrosodimethylamine (NDMA)
2. Experimental
[21]), oxyanions (e.g., nitrate [22], nitrite [23], chlorate [24], per-
chlorate [25], bromate [11], and hexavalent chromium (CrVI) [26]).
2.1. Chemicals
Compared with the conventional water treatment technologies dis-
cussed above, catalytic reduction is considered to be a promising water
All chemicals involved in this work were of analytical grade and
treatment technology because it usually uses easily available hydrogen
used as received without further purification. Sodium bromate
as a clean reductant and can simultaneously eliminate multiple con-
(NaBrO3), sodium chloride (NaCl), monosodium phosphate (NaH2PO4),
taminants from water without generating brines or secondary waste
sodium hydrogen phosphate (Na2HPO4), sodium phosphate (Na3PO4),
streams. Therefore, the development of catalytic bromate reduction
sodium nitrate (NaNO3), sodium sulfate (Na2SO4), aqueous NH3·H2O,
over a noble metal catalyst has been extensively investigated in recent
acetone (CH3OCH3), and hydrochloric acid (HCl) were purchased from
decades. In principle, bromate is reduced by hydrogen atom into bro-
Sinopharm Chemical Reagent Co., Ltd. The 5% Pd/C catalyst, po-
mide, a harmless species in water, over a monometallic active site (e.g.,
tassium hexachloroplatinate (K2PtCl6), palladium chloride (PdCl2),
Pd, Pt, and Ru) at room temperature and atmospheric pressure. Among
tetraethoxysilane (TEOS), and pluronics (P123, PEO20PPO70PEO20)
various noble metals, Pt and Pd have been widely considered to be the
were purchased from Sigma-Aldrich. Hydrogen (99.99%) was produced
most effective catalyst for reducing bromate to bromide, and this pro-
by a high-purity hydrogen generator (SGH-500, Beijing Country
cess can be briefly described as the following formula [11,27–29]:
Spectrum Technology Co., Ltd.). Ultrapure water (18.25 MΩ·cm) was
Pd,Pt used for all experiments.
BrO3− + 3H2 ⎯⎯⎯⎯⎯→ Br − + 3H2 O (1)

Similar to most heterogeneous catalytic reactions, the practical 2.2. Synthesis of catalysts
application of this technology is mainly dependent on the catalyst ac-
tivity and stability [30]. Corresponding investigations have indicated 2.2.1. Synthesis of SBA-15
that the properties of the catalyst support play an important role in the The SBA-15 mesoporous silica material was synthesized according
catalyst behaviour. This relationship is because the support can strongly to the previously reported procedure [36]. Tetraethoxysilane (TEOS,
affect the nanoparticle size, dispersion, and electronic structure of the 98%, Aldrich) and the pluronic P123 (PEO20-PPO70-PEO20) were em-
supported metal and the chemical and mechanical stability of the cat- ployed as the source of silica and the structure-directing agent, re-
alyst [31,32]. In the case of bromate hydrogenation, various powder spectively. Typically, 1.65 g of P123 was dissolved in 50 ml of a 2 M HCl

115
Z. Zhang et al. Chemical Engineering Journal 344 (2018) 114–123

solution, and after stirring at 38 °C for 2 h, 3.7 ml of TEOS was added where ni is the number of Pd particles with a diameter of di and the total
n
dropwise into the solution, which was continuously stirred at 38 °C for number of particles, N = ∑i , is larger than 200.
another 24 h. After this, the resulting white gel was subsequently The zeta potential (ξ) of the support and catalysts was measured on
transferred to a 100 ml Teflon-lined autoclave and heated at 100 °C for a zeta potential analyser (Nano-Z, Malvern, UK). Briefly, 10 mg of
24 h under static conditions. The final product was filtered, washed sample was dispersed in 100 ml of a 0.01 M KBr solution with different
thoroughly with ultrapure water and ethanol, and dried in air at 100 °C pH values obtained by adding 0.05 M HCl or 0.05 M NaOH, and the
for 12 h. Finally, the structure-directing agent was removed by calci- suspension was equilibrated for 24 h prior to measurement.
nation in air at 550 °C for 3 h.
2.4. Bromate catalytic reduction
2.2.2. Synthesis of Pd/SBA-15 and Pt/SBA-15
Uniform Pd and Pt catalysts with different loadings were synthe-
Bromate catalytic reduction tests were carried out in a 100 ml
sized by the adsorption method reported by Li et al. [37]. Typically, a
homemade semi-batch reactor under atmospheric pressure. The reac-
certain amount of K2PtCl6 or PdCl2 precursor solution was transferred
tion temperature and stirring speed were maintained at the desired
to a 100 ml beaker and diluted to ca. 10 ml with ultrapure water. Then,
values using a water bath and magnetic stirrer. Typically, 90 ml of ul-
the pH value of the solution was adjusted to 11–12 by the dropwise
trapure water was added into the reactor under stirring at ca. 500 rpm
addition of aqueous NH3·H2O to form [Pd(NH3)4]2+ or [Pt(NH3)4]2+,
and bubbled with a flow of H2 (50 ml·min−1, Fig. S1(b)) for 30 min to
which adsorbed to the surface silanol groups (negatively charged at this
remove dissolved oxygen from the water (Fig. S2). Afterwards, 30 mg of
pH, Fig. 6) through electrostatic interactions. Afterwards, 0.5 g of the
the catalyst (Fig. S1(a)) pre-dispersed in 5 ml of ultrapure water was
as-synthesized SBA-15 was slowly added into the precursor solution,
introduced into the reactor with a syringe. Subsequently, 5 ml of a
and the mixture was sonicated until it became homogeneous. After
concentrated bromate solution was immediately added into the reactor
stirring for another 24 h, the synthesized solid was separated carefully
to obtain the desired initial bromate concentration and start the reac-
by filtration and washed with ultrapure water and acetone, dried at
tion. Samples were withdrawn at specific reaction times, and the cat-
65 °C under vacuum for 12 h, and activated at 300 °C for 3 h under a
alyst particles were removed by filtering through a 0.22 μm membrane.
10% H2/Ar flow of 100 ml/min.
The adsorption experiment was carried out under identical conditions
except that H2 was replaced with N2. The bromate concentration of the
2.3. Catalyst characterization
filtrate was determined using a ICS2100 ion chromatograph (Dionex,
USA). After the reaction was finished, the catalyst dispersed in the re-
The actual metal loadings of the different catalysts were measured
action solution was collected by centrifugation, washed with water and
using ICP-AES (5300DV, Perkin Elmer, USA). Infrared spectra of the
acetone and dried overnight at 65 °C under vacuum for further reusa-
catalysts were collected on a Nicolet 380 FTIR spectrometer (Thermo
bility experiments.
Scientific, USA) from 4000 to 400 cm−1 using KBr pellets. Low-angle
The catalytic activity and the initial activity were respectively es-
and wide-angle powder XRD patterns were obtained from an Empyrean
timated at a bromate conversion of 80% and a reaction time of 5 min.
XRD (Panalytical, Netherlands) and a D8 Advance XRD (Bruker-Axs,
Germany), using Cu Kα radiation (λ = 0.1541 nm) in a scanning range
of 0.5–5° and 10–90° (2θ), respectively, with a 0.02° step size. N2 ad- 3. Results and discussion
sorption-desorption isotherms were measured on a QuadraSorb SI
analyser (Quantachrome, USA) at a bath temperature of −195.85 °C. 3.1. Catalyst characterization
The corresponding pore size distribution curves were obtained from the
adsorption branches of the isotherms using the BJH method. The 3.1.1. Texture, structure and composition
morphologies of the support and catalysts were observed by SEM The actual Pd and Pt loadings in all the catalysts were analysed
(Supra 55 Sapphire, Zeiss, Germany) and TEM (JEM 2100, JEOL, using ICP-AES, and the results revealed that the Pt and Pd species could
Japan). Prior to the TEM measurements, approximately 5 mg of sample be effectively deposited on the SBA-15 support by the adsorption
was dispersed in 10 ml of ethanol and sonicated for 20 min, and then a method and that the loading could be easily adjusted by altering the
drop of the suspension was taken out and deposited on a Cu grid cov- precursor solution concentration (Table 1). The normalized FTIR
ered with an ultrathin film of carbon. The grid was then dried in a spectra of fresh SBA-15 and the 4% Pt/SBA-15 and 4% Pd/SBA-15
vacuum oven overnight at room temperature. The Pd nanoparticle size catalysts are shown in Fig. 1, and they exhibit three clear absorption
distributions of catalysts were statistically analysed via the TEM peaks at 1076, 809, and 463 cm−1, which can be ascribed to the
images, and the average Pd particle sizes (dmean) were calculated using asymmetric or symmetric vibrations of the Si-O-Si bonds in the in-
Eq. (2) below [11]: organic framework [36]. The characteristic peak of adsorbed H2O ob-
served at 1631 cm−1 [38] in both the 4% Pt/SBA-15 and 4% Pd/SBA-15
∑ ni di3
dmean = catalysts decreased compared to that in SBA-15, indicating that H2O
∑ ni di2 (2) adsorption was suppressed following metal loading.

Table 1
The textural and elemental characterization of the catalysts and fresh SBA-15 support.

Samples Textural characterization Loading Particle size

2 −1 3 −1 a
SBET (m ·g ) Vpore (cm ·g ) Dpore (nm) Theoretical (wt%) Actual (wt%) dmean (nm)

SBA-15 831 1.27 7.55 – – –


2% Pd/SBA-15 592 0.95 6.56 2.00 1.93 4.54
3% Pd/SBA-15 563 0.91 6.55 3.00 2.95 4.56
4% Pd/SBA-15 469 0.80 6.56 4.00 3.89 4.40
2% Pt/SBA-15 605 0.97 7.33 2.00 1.86 4.67
3% Pt/SBA-15 558 0.93 6.63 3.00 2.93 4.51
4% Pt/SBA-15 479 0.84 6.34 4.00 3.91 4.28
5% Pd/C 591 0.58 1.69 5.00 4.75 4.24

116
Z. Zhang et al. Chemical Engineering Journal 344 (2018) 114–123

various metal loadings are shown in Fig. S4 and Fig. 4(d) to (i). Ob-
H2 O Si-O-Si viously, both the Pd and the Pt nanoparticles exhibit a spherical or
ellipsoidal morphology and deposit well on SBA-15 without aggrega-
Transmittance (a.u.)

tion. The nanoparticle size distributions of these catalysts were de-


termined (inserted images), and the average nanoparticle sizes (dmean)
Si-O-Si were calculated to be 4.54, 4.56, 4.40, 4.67, 4.51, and 4.28 nm for 2%
Pd/SBA-15, 3% Pd/SBA-15, 4% Pd/SBA-15, 2% Pt/SBA-15, 3% Pt/SBA-
15, and 4% Pt/SBA-15, respectively (Table 1), indicating that the pre-
sence of strong interactions between the SBA-15 support material and
the metal species prevented aggregation. In addition, the nearly com-
mensurate dmean values of the catalysts reflect the presence of similar
metal dispersions due to the metrological relationship between the
4% Pt/SBA-15 metal dispersion and metal particle size [42].
Si-OH
4% Pd/SBA-15
Si-O-Si
SBA-15
3.2. Bromate reduction over Pd/SBA-15, Pt/SBA-15, and Pd/C
2000 1750 1500 1250 1000 750 500
Supported Pd and Pt nano-catalysts are widely used in catalytic
Wavenumber (cm-1)
hydrogenation due to their high H2 activation capacities at low pressure
Fig. 1. FTIR spectra of the fresh SBA-15 support, 4% Pd/SBA-15, and 4% Pt/SBA-15. and/or temperature [43]. The catalytic performance of the Pd and Pt
catalysts in catalytic bromate reduction was evaluated in water solu-
The typical N2 adsorption-desorption isotherms and pore diameters tions. As shown in Fig. S5, bromate could not be reduced by H2 without
of pure SBA-15 and the 4% Pt/SBA-15 and 4% Pd/SBA-15 catalysts are a catalyst. Nevertheless, without metal loading, ca. 20% of the bromate
shown in Fig. 2(a) and (b), respectively. The BET specific surface area was removed within 120 min due to adsorption to SBA-15 (Fig. 5(a) and
(SBET), pore volume (Vpore), and pore diameter of SBA-15 are (b)). Unlike the unmodified SBA-15, 80% bromate removal was
831 m2·g−1, 1.27 cm·g−1 and 7.55 nm, respectively, while those of 4% achieved over 2% Pd/SBA-15, 3% Pd/SBA-15, 4% Pd/SBA-15, 2% Pt/
Pt/SBA-15 and 4% Pd/SBA-15 dropped to 479 m2·g−1, 0.84 cm·g−1 and
6.34 nm and 469 m2·g−1, 0.80 cm·g−1, and 6.56 nm, respectively a)
(Table 1). In addition, the SBET, Dpore, and Vpore values of all the catalysts
exhibited a moderate decrease following an increase in the Pt or Pd
N2 uptake (cm3·g-1 at STP)

250
loading, indicating that the mesoporous pores of SBA-15 were partially
occupied or blocked with metal nanoparticles.
Fig. 3(a) and (b) shows the low-angle and wide-angle p-XRD spectra
of SBA-15 and the representative 4% Pt/SBA-15 and 4% Pd/SBA-15 4%Pt/SBA-15
catalysts. SBA-15 and the two typical catalysts exhibited three clear
diffraction peaks at 2θ values of 0.9°, 1.6° and 1.8°, which are respec-
tively assigned to the (1 0 0), (1 1 0) and (2 0 0) planes, revealing the 4%Pd/SBA-15
presence of a typical two-dimensional hexagonal ordered mesoporous
structure [39], and further suggesting that SBA-15 maintained its basic
structure after metal loading. In addition, the intensity of these peaks
dropped, indicating that the number of ordered mesopores decreased SBA-15
following an increase in the metal loading and revealing occupation of
the pores by the Pt or Pd nanoparticles, which is consistent with the
texture analysis results (Table 1). In Fig. 3(b), the wide-angle p-XRD 0.0 0.2 0.4 0.6 0.8 1.0
pattern of 4% Pt/SBA-15 compared to that of SBA-15presents four P/P0
characteristic Pt diffraction peaks at 2θ values of 39.8°, 46.1°, 67.8°,
and 81.5°, which are assigned to the (1 1 1), (2 0 0), (2 2 0), and (3 1 1) b)
interplanar spacings of Pt0, respectively [40], indicating that Pt nano- 4
crystals with multiple facets existed in the Pt/SBA-15 catalyst. How-
dV/d(logD) (cm3·g-1)

ever, only one characteristic Pd diffraction peak at a 2θ value of 40.2°, 4%Pt/SBA-15


which corresponded to the (1 1 1) interplanar spacing of face-centred
cubic (fcc) Pd0 [41], was observed in the 4% Pd/SBA-15p-XRD pattern.
In addition, an increase in the metal loading led to increases in the
intensities of the diffraction peaks, reflecting a higher content of Pt or
Pd nanoparticles supported on the SBA-15pores and/or surface with the 4%Pd/SBA-15
increase in the metal loading (Fig. S3), which is highly consistent with
the ICP-AES results (Table 1).

SBA-15
3.1.2. Morphologicales analysis
In Fig. 4(a), the SEM image shows that SBA-15 had a typical rod-like
morphology with a uniform size of ca. 0.8 μm × 1.5 μm. The TEM 0 10 20 30 40 50
images of SBA-15 shown in Fig. 4(b) and (c) exhibit a highly uniform Pore diameter (nm)
ordered 2D mesoporous structure with a pore diameter of ca. 7.4 nm,
which is highly consistent with the textural characterization results. Fig. 2. (a) N2 adsorption-desorption isotherms of the fresh SBA-15 support, 4% Pd/SBA-
15, and 4% Pt/SBA-15. (b) Corresponding pore-size distribution curves.
The TEM images of 5% Pd/C and Pd/SBA-15 and Pt/SBA-15 with

117
Z. Zhang et al. Chemical Engineering Journal 344 (2018) 114–123

a) partly depends on the electrical properties of the catalyst surface [44].


However, the zeta potentials of the 4% Pd/SBA-15 and 4% Pt/SBA-15
20000 catalyst were roughly the same (approximate −18 mV) at the reaction
pH of 6.5 ± 0.25 (Fig. 6), proving the zeta potential is not the main
4%Pt/SBA-15 reason for the activity differences between the Pd and Pt-based cata-
Intensity (a.u.)

lysts. In addition, it has been reported that Pd and Pt metals have


identical adsorption heats and affinities towards H2 [45], thus, the
higher catalytic activity of Pt/SBA-15 probably results from the pre-
4%Pd/SBA-15 sence of more facets in the Pt nanoparticles compared to in the Pd
nanoparticles according to the p-XRD results above (Fig. 3(b)). Since
the 4% Pt/SBA-15 catalyst exhibited a higher catalytic performance
than the 4% Pd/SBA-15 catalyst, the following discussion mainly fo-
SBA-15 cuses on bromate reduction over 4% Pt/SBA-15.
(100) Activated carbon has been widely employed as a catalyst support in
(110) (220) many heterogeneous catalytic processes [32,46–48]. Thus, catalytic
bromate reduction by a commercial 5% Pd/C catalyst was also eval-
0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 uated under identical reaction conditions for comparative purposes in
this study. The reaction rate constant and catalytic activity for this
2 Theta (°)
commercial catalyst were 0.0133 min−1 and 4.2 mg·min−1·gcat−1, re-
b) spectively, revealing that the reduction of bromate over Pd/C was
significantly inferior to that by the Pd/SBA-15 and Pt/SBA-15 catalysts.
(111)
Considering that the significant pore size differences between the SBA-
(200)
(220) 15 and Pd/C, the excellent catalytic activities of the SBA-15-based
(311) catalysts resulted from the two-dimensional straight pores of SBA-15,
Intensity (a.u.)

which are beneficial for improving the mass transfer process; thus, SBA-
(111)
4%Pt/SBA-15 15 was a more suitable catalyst support material for the catalytic re-
● duction of bromate.
(200)

(220)
● (311) 3.3. Intrinsic mechanism analysis

4%Pd/SBA-15
As a heterogeneous catalytic reaction, the catalytic bromate re-
500 duction rate is mainly controlled by three successive steps: diffusion of
bromate from the bulk into the catalyst pores and solid-liquid inter-
SBA-15
faces, adsorption of bromate onto active sites, and reduction of the
10 20 30 40 50 60 70 80 90 adsorbed bromate by H atoms. To better understand the intrinsic me-
chanism of catalytic reduction of bromate over 4% Pt/SBA-15, the ef-
2 Theta (°) fects of the temperature and initial bromate concentration on the bro-
Fig. 3. (a) Low-angle and (b) wide-angle p-XRD patterns of the fresh SBA-15 support, 4% mate removal efficiency were evaluated. As shown in Fig. 7(a) and (b),
Pd/SBA-15, and 4% Pt/SBA-15. an increase in the reaction temperature from 15 to 30 °C led to an in-
crease in the catalytic activity from 12.21 to 29.03 mg·min−1·gcat−1.
The corresponding activation energy (Ea) was also estimated (Fig. 7(c))
SBA-15, 3% Pt/SBA-15, and 4% Pt/SBA-15 within ca. 118, 45, 31, 69,
from the thermodynamic analysis to be 47.31 kJ·mol−1 (r2 = 0.96),
22, and 13 min, respectively, indicating that the catalytic bromate re-
implying that bromate diffusion from the bulk to the active sites is not
duction efficiency significantly improved over the Pd/SBA-15 and Pt/
the main limiting step in the reduction process [49]. This result can
SBA-15 catalysts. The reaction rate constants for 4% Pd/SBA-15 and 4%
probably be explained by the presence of highly uniform ordered 2D
Pt/SBA-15 were estimated to be 0.0551 and 0.1802 min−1, respec-
mesoporous channels, which promote diffusion of the reaction substrate
tively, and 4% Pt/SBA-15 exhibited a 3.27-fold enhancement relative to
[50] and further eliminate the effect of the mass transfer process.
4% Pd/SBA-15 (Fig. 5(c)). The catalytic activities of the supported Pd
The catalytic bromate reduction performance of 4% Pt/SBA-15
and Pt catalysts were calculated at a bromate removal of 80%, and the
under various initial bromate concentrations as a function of reaction
results shown in Fig. 5(d) indicate that the activities of the Pt/SBA-15
time is presented in Fig. 8. This figure shows that bromate could be
catalysts were higher than those of the Pd/SBA-15 catalysts for the
completely degraded within ca. 30 min, indicating that the reduction
same loadings. For example, the activity of 4% Pt/SBA-15 was
rate improved with the increase in the initial substrate concentration.
24.72 mg·min−1·gcat−1, which is approximately 2.33- and 6.50-fold
The corresponding catalytic activities were calculated and are pre-
higher than those of the 4% Pd/SBA-15 and 5% Pd/C catalysts, re-
sented in the inset. Obviously, increasing the initial bromate con-
spectively.
centration from 10 to 100 mg·L−1 resulted in a rapid increase in the
Specifically, the consumption of bromate was accompanied by the
catalytic activity from 1.57 to 24.72 mg·min−1·gcat−1. According to this
production of bromide, and the total concentrations of bromate and
result, H2 adsorption on the noble metal surface is much stronger than
bromide were very close to the initial bromate concentration (Fig.
that of bromate [11], and the H atom coverage of the Pt surface can be
S6(a)), suggesting that bromate degradation occurred through catalytic
considered to be constant due to the continuous aeration with a con-
reduction rather than an adsorption process. In addition, none of the
stant H2 flow rate during the reaction; thus, the catalytic reduction
intermediate oxyanions (e.g., BrO2− and BrO−) could be observed in
process can be approximatively described as a simple decomposition
the IC curves (Fig. S6(b)), indicating that the reduction of such inter-
reaction based on the Langmuir-Hinshelwood mechanism, as shown
mediates was much faster than that of bromate and that bromide was
below [51,52]:
the only product.
Generally, the zeta potential of a catalyst strongly influences the dC k1 k2 θC
R0 = − =
catalytic activity because the adsorption capacity for the substrate dt k1 C + k−1 + k2 (3)

118
Z. Zhang et al. Chemical Engineering Journal 344 (2018) 114–123

a) b) c)

2 μm 200 nm 50 nm

d)40 e)40 f)40

Frequency (%)
30 30 30

Frequency (%)
Frequency (%)

20 20 20

10 10 10

0 0 0
3 4 5 6 50 nm 3 4 5 6 50 nm 3 4 5 6 50 nm
Diameter (nm) Diameter (nm) Diameter (nm)

g)40 h)40 i) 40
30 30 30
Frequency (%)

Frequency (%)
Frequency (%)

20 20 20

10 10 10

0 0 0 50 nm
3 4 5 6 50 nm 3 4 5 6 50 nm 3 4 5 6
Diameter (nm) Diameter (nm) Diameter (nm)
Fig. 4. Morphologies of the SBA-15 support and catalysts. (a) SEM image of SBA-15; (b) and (c) TEM images of SBA-15; (d), (e) and (f) TEM images of 2%, 3% and 4% Pd/SBA-15,
respectively; (g), (h) and (i) TEM images of 2%, 3% and 4% Pt/SBA-15, respectively. The insets in panels (d)-(i) are the corresponding particle size distributions of the Pd or Pt NPs.

where R0 is the bromate reduction rate (reaction activity), C is the results (Fig. 5(c)). This result indicates that the limiting step is the re-
bromate concentration, t is the reaction time, k1 and k−1 are the bro- duction of bromate on the Pt surface rather than the diffusion process.
mate adsorption and desorption rate constant, respectively, k2 is the The linear fitting curves of the initial bromate concentration C
reaction rate constant, and θ is the total number of sites (occupied or versus reaction rate (activity) R0 is given in the insert image of Fig. 8.
unoccupied). The linear correlation index r2 of R0 with C is 0.962, implying that the
Based on the analysis above, the main limiting step in the reduction catalytic bromate reduction over 4% Pt/SBA-15 catalyst is a first order
process is largely dependent on bromate hydrogenation on the Pt sur- reaction in component bromate under low concentration condition,
face, namely: k2 ≤ k1C, k−1. Thus, which is consistent with the above pseudo-first reaction kinetic fitting
results (Fig. 5(c)). This result indicates that the limiting step is the re-
k1 k2 θC Kk2 θC duction reaction of bromate on the Pt surface rather than the diffusion
R0 ≈ =
k1 C + k−1 KC + 1 (4) process.
where K = k1/k2 is the adsorption equilibrium constant.
In this study, the initial bromate concentration was below 100 mg/L
3.4. Effect of reaction pH on catalytic bromate reduction
(< 7.8 × 10−4 mol/L), thus the above equation can be simplified as
follows:
The reaction pH is one of the most important parameters affecting
R 0 ≈ Kk2 θC = K ′C (5) the catalytic activity in heterogeneous catalytic reactions in water.
Catalytic bromate hydrogenation was performed over 4% Pt/SBA-15 at
The linear fitting curves of the initial bromate concentration, C, various pH values obtained by adjustment with phosphate buffer. It
versus reaction rate (activity), R0, is given in the inset. The linear should be noted that to avoid the effects of phosphate ions, the phos-
correlation index, r2, of R0 with C is 0.962, implying that catalytic phate concentration was constant in all the parallel tests. As presented
bromate reduction over 4% Pt/SBA-15 catalyst is a first-order reaction in Fig. 9(a), an increase in the pH from 3 to 11 significantly decreased
in terms of bromate under low concentration conditions, which is the bromate removal rate, and the corresponding initial catalytic ac-
consistent with the above pseudo-first-order reaction kinetic fitting tivity decreased from 62.31 to 1.59 mg·min−1·gcat−1 (Fig. 9(b)), which

119
Z. Zhang et al. Chemical Engineering Journal 344 (2018) 114–123

a) 100 SBA-15 5%Pd/C 2%Pd/SBA-15 b) 100 SBA-15 5%Pd/C 2%Pt/SBA-15


3%Pd/SBA-15 4%Pd/SBA-15 3%Pt/SBA-15 4%Pt/SBA-15

80 80

Bromate (mg·L-1)
Bromate (mg·L-1)

60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Reaction time (min) Reaction time (min)
c) 5% Pd/C
d) 30
0 4% Pd/SBA-15 Pd/C
4% Pt/SBA-15 25

Activity (mg·min-1·gcat-1)
Pd/SBA-15
-1
r12=0.989 Pt/SBA-15
-2 20
ln(Ct/C0)

-3 r22=0.971 15
-4
r32=0.963 10
-5
-6 5
-7
0
0 20 40 60 80 100 120 5% 2% 3% 4%
Reaction time (min) Metal loading
Fig. 5. (a) and (b) Catalytic bromate reduction over various catalysts; (c) corresponding plots of ln(Ct/C0) versus reaction time for bromate reduction over 4% Pd/SBA-15, 4% Pt/SBA-15
and 5% Pd/C; (d) corresponding catalytic activities of the various catalysts. Reaction conditions: bromate: 100 mg·L−1, catalyst: 30 mg/L, pH: 6.5 ± 0.25, and temperature: 25 °C.

10 enhanced the repulsive forces between bromate and the catalyst, which
weakened the adsorption of bromate ions onto the catalyst active sites
0 and subsequently decreased the catalytic bromate reduction rate. Fur-
thermore, BrO3− can be protonated to HBrO3 at a high H+ con-
Zeta potential (mV)

centration (low pH) [53], which will probably change the electronic
-10 structure of the bromate and further improve the bromate reducibility
and vice versa. In Fig. 9(a), the bromate reduction rate in the absence of
-20 buffer is obviously higher than that at the condition of pH = 5; this
difference is probably because of the inhibition effect of phosphate ions,
-30 which will be discussed below.

SBA-15
-40 4%Pd/SBA-15 3.5. Effect of coexisting anions on the catalytic bromate reduction
4%Pt/SBA-15
-50 Pd/C Based on the fact that catalytic bromate reduction occurs on the Pt
surface, competitive adsorption on the Pt active sites by coexisting
-60 anions may inhibit the bromate reduction rate. Therefore, the effects of
2 3 4 5 6 7 8 9 10 11 the ubiquitous anions present in water sources, including PO43−,
SO42−, NO3−, and Cl−, on catalytic bromate reduction by 4% Pt/SBA-
pH 15 were investigated. As shown in Fig. 10, at a reaction time of 10 min,
the bromate removal rate was 23.2%, 31.1%, and 62.4% in the presence
Fig. 6. Zeta potentials of SBA-15, 4% Pd/SBA-15, 4% Pt/SBA-15 and 5% Pd/C. (The
shaded region represents the reaction pH).
of PO43−, SO42−, and Cl−, respectively, which were significantly lower
than that of 77.6% in the absence of coexisting anions. This result in-
dicates that these anions markedly suppressed bromate reduction, and
indicated that the catalytic activity of aqueous bromate reduction was the inhibition effects followed the order of PO43− > SO42− > Cl−.
inhibited at a high pH. This result is mainly because the redox potential The inhibition effects can be mainly explained by the occupation of the
of bromate/bromide decreases with increasing pH, which indicates that active sites by the coexisting anions, which decreased the number of
the reducibility of bromate is poor at high pH values, and the catalytic active sites available for bromate reduction. The order of inhibition
bromate reduction rate automatically decreases [27]. In addition, from likely results from the fact that anions with high ionic charges generally
the zeta potential analysis results, the negative charge of the catalyst have high adsorption affinities on a metal surface [11]. These results
increased with increasing pH (Fig. 6). Therefore, an increase in the pH indicate that satisfactory catalytic reduction of bromate cannot be

120
Z. Zhang et al. Chemical Engineering Journal 344 (2018) 114–123

Temperature ( )
a) b) 35 c) 30 25 20 15
100
15 20

Activity (mg·min-1·gcat-1)
30 -5.6
25 30
80
Bromate (mg·L-1)

-5.8
25
-6.0
60 20 r2 = 0.96

ln (k)
-6.2
15
40 -6.4
10 Ea = 47.31 kJ/mol
-6.6
20
5 -6.8

0 0 -7.0
0 10 20 30 40 50 60 15 20 25 30 3.30 3.35 3.40 3.45 3.50
Reaction time (min) Temperature ( ) 1000/T (K-1)

Fig. 7. (a) Effects of temperature on the bromate reduction over the 4% Pt/SBA-15 catalyst; (b) corresponding catalytic activity; (c) estimation of the bromate reduction activation energy
(Ea).

100 mg·L-1 80 mg·L-1 60 mg·L-1


100 None PO43- SO42-
1.0
40 mg·L-1 20 mg·L-1 10 mg·L-1 Cl- NO3-
80

Bromate (mg·L-1)
0.8 25
Activity (mg·min-1·gcat-1)

20
60
Ct/C0

0.6 15
r2=0.962
10

0.4 5
40

0
0.2 0 20 40 60 80 100 20
Initial concentration (mg·L-1)

0.0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60
Reaction time (min) Reaction time (min)
Fig. 8. Catalytic bromate reduction over 4% Pt/SBA-15 under various initial bromate Fig. 10. Effects of coexisting anions on catalytic bromate reduction (coexisting anion
concentrations, and catalytic activity of 4% Pt/SBA-15 as a function of the initial bromate concentration: 1 mmol/L).
concentration (insert image).

catalytic activity. Bromate removal was obviously enhanced at a low


achieved in water containing high concentrations of sulfate and/or nitrate concentration (< 1.5 mM) in the reaction solution, while it was
phosphate. moderately inhibited at a high nitrate concentration (> 1.5 mM). It is
In addition, it is noticed in Fig. 10 that the bromate reduction rate well known that nitrate can also be easily reduced to N2 or ammonia in
was obviously enhanced when 1.0 mM nitrate coexisted in the reaction catalytic hydrogenation processes [22,41]. However, no nitrate reduc-
solution. To better understand how the nitrate concentration affects the tion was observed in this study. Actually, nitrate can only be reduced by
catalytic activity, catalytic bromate reduction in the presence of various a bimetallic catalyst such as Pd-Cu, Pt-Cu, and Pd-Sn because nitrate
nitrate concentrations was carried out. The results showed a volcano- cannot adsorb on a monometallic site (e.g., Pt, Pd) without a promotor
type relationship between the coexisting nitrate concentration and the metal [54–56]. Therefore, the nitrate ions could not adsorb on the Pt

a) b) 70
100
Initial activity (mg·min-1·gcat-1)

60
80
Bromate (mg·L -1)

50

60 40
No buffer pH=3
pH=5 pH=7
30
40 pH=9 pH=11

20
20
10

0 0
0 10 20 30 40 50 60 3 5 7 9 11
Reaction time (min) pH
Fig. 9. (a) Effect of reaction pH on catalytic bromate reduction, and (b) corresponding initial catalytic activity at different pH values.

121
Z. Zhang et al. Chemical Engineering Journal 344 (2018) 114–123

a) b)
Run 1 Run 2 Run 3 Run 4 Run 5
100
(111)
Bromate (mg·L-1)

Intensity (a.u.)
80 (200)
(220) (311)
60
used 4%Pt/SBA-15
40

20 500 fresh 4%Pt/SBA-15

0
0 10 20 0 10 20 0 10 20 0 10 20 0 10 20 30 10 20 30 40 50 60 70 80 90
Reaction time (min) 2 Theta (°)
Fig. 11. (a) Reusability of the 4% Pt/SBA-15 catalyst in bromate reduction. (b) Wide-angle p-XRD patterns of the fresh and used 4% Pt/SBA-15 catalyst.

catalyst surface and decrease the number of Pt active sites, which re- Science Foundation of China (Grant No.51778172), and the National
sulted in no significant inhibition of bromate reduction. Natural Science Foundation of China (Grant No.51573034).

3.6. Catalyst reusability Appendix A. Supplementary data

The catalyst reusability is also a very important factor for the Supplementary data associated with this article can be found, in the
practical application of catalytic processes. Therefore, the reusability of online version, at http://dx.doi.org/10.1016/j.cej.2018.03.056.
the 4% Pt/SBA-15 catalyst was measured by repeated experiments
under the same reaction conditions in this study. As shown in References
Fig. 11(a), bromate could be completely reduced within 30 min in all
five cycles of the test. In addition, Fig. S8 and Table S1 show that the [1] C. Liu, G.U. Von, J.P. Croué, Enhanced bromate formation during chlorination of
bromide-containing waters in the presence of CuO: catalytic disproportionation of
catalytic activity loss and Pt leaching were only 7.8% and 2.6%, re- hypobromous acid, Environ. Sci. Technol. 46 (2012) 11054–11061.
spectively, after 5 cycles. These low losses are probably because the Pt [2] R. Song, C. Donohoe, R. Minear, P. Westerhoff, K. Ozekin, G. Amy, Empirical
nanoparticle structures supported on SBA-15 remained stable in the modeling of bromate formation during ozonation of bromide-containing waters,
Water Res. 30 (1996) 1161–1168.
repeated tests. The p-XRD characterization of the fresh and used cata- [3] J. Yang, J. Li, J. Zhu, Z. Dong, F. Luo, Y. Wang, H. Liu, C. Jiang, H. Yuan, A novel
lyst (Fig. 11(b)) shows that the Pt characteristic peaks exhibited no design for an ozone contact reactor and its performance on hydrodynamics, disin-
obvious changes, which confirmed the above conjecture. Thus, based fection, bromate formation and oxidation, Chem. Eng. J. 328 (2017) 207–214.
[4] Y. Liu, Y. Yang, S. Pang, L. Zhang, J. Ma, C. Luo, C. Guan, J. Jiang, Mechanistic
on the high activity and stability, the 4% Pt/SBA-15 catalyst can be insight into suppression of bromate formation by dissolved organic matters in sul-
used as a promising reusable catalyst in the catalytic reduction of fate radical-based advanced oxidation processes, Chem. Eng. J. 333 (2018)
aqueous bromate. 200–205.
[5] Y. Kurokawa, A. Maekawa, M. Takahashi, Y. Hayashi, Toxicity and carcinogenicity
of potassium bromate–a new renal carcinogen, Environ. Health Persp. 87 (1990)
4. Conclusions 309.
[6] P. Zhang, F. Jiang, H. Chen, Enhanced catalytic hydrogenation of aqueous bromate
over Pd/mesoporous carbon nitride, Chem. Eng. J. 234 (2013) 195–202.
In this study, Pd and Pt nanoparticles supported on the mesoporous [7] S. Gyparakis, E. Diamadopoulos, Formation and reverse osmosis removal of bro-
silica molecular sieve SBA-15 were prepared and applied in catalytic mate ions during ozonation of groundwater in coastal areas, Sep. Sci. Technol. 42
bromate reduction. The activity investigation results showed that the (2007) 1465–1476.
[8] J.A. Wiśniewski, Bromate removal in the ion-exchange process, Desalination 261
Pt/SBA-15 catalysts exhibited higher activities than the Pd/SBA-15 and (2010) 197–201.
commercial 5% Pd/C catalysts. Analysis of the intrinsic mechanism [9] J.A. Wiśniewski, M. Kabsch-Korbutowicz, S. Łakomska, Donnan dialysis and elec-
revealed that bromate reduction followed a simple decomposition re- trodialysis as viable options for removing bromates from natural water,
Desalination 281 (2011) 257–262.
action based on the Langmuir-Hinshelwood mechanism, and the reac- [10] H. Ji, W. Wu, F. Li, X. Yu, J. Fu, L. Jia, Enhanced adsorption of bromate from
tion of adsorbed bromate rather than the adsorption process mainly aqueous solutions on ordered mesoporous Mg-Al layered double hydroxides (LDHs),
controlled the reaction rate. Catalytic bromate reduction was enhanced J. Hazard. Mater. 334 (2017) 212.
[11] H. Chen, Z. Xu, H. Wan, J. Zheng, D. Yin, S. Zheng, Aqueous bromate reduction by
at a low reaction pH. The bromate removal rate was suppressed in the catalytic hydrogenation over Pd/Al2O3 catalysts, Appl. Catal. B: Environ. 96 (2010)
presence of the coexisting anions PO43−, SO42−, and Cl−, and the in- 307–313.
hibition effects followed the order of PO43− > SO42− > Cl−. [12] Z. Zhang, J. Cheng, Y. Luo, W. Shi, W. Wang, B. Zhang, R. Zhang, X. Bao, Y. Guo,
F. Cui, Pt nanoparticles supported on amino-functionalized SBA-15 for enhanced
Surprisingly, bromate removal was obviously enhanced by the presence aqueous bromate catalytic reduction, Catal. Commun. 105 (2018) 11–15.
of a low concentration (< 1.5 mM) of nitrate in the reaction solution. [13] C.G.V. Ginkel, A.M.V. Haperen, B.V.D. Togt, Reduction of bromate to bromide
Additionally, only a minor loss in catalytic activity of 7.8% after five coupled to acetate oxidation by anaerobic mixed microbial cultures, Water Res. 39
(2005) 59–64.
repeated tests indicated that the 4% Pt/SBA-15 catalyst is reusable in
[14] R. Butler, A.R. Godley, R. Lake, L. Lytton, E. Cartmell, Reduction of bromate in
the catalytic reduction of aqueous bromate. groundwater with an ex situ suspended growth bioreactor, Water Sci. Technol. 52
(2005) 265–273.
[15] K.Y.A. Lin, C.H. Lin, Enhanced reductive removal of bromate using acid-washed
Acknowledgements zero-valent iron in the presence of oxalic acid, Chem. Eng. J. 144–150 (2017).
[16] C.H. Xu, S. Lin, X.H. Wang, Y.M. Chen, L.J. Zhu, Z.H. Wang, Ordered mesoporous
The authors are grateful for the financial support from the State Key carbon immobilized nano zero-valent iron in bromate removal from aqueous so-
lution, J. Taiwan Inst. Chem. Eng. 45 (2014) 3000–3006.
Laboratory of Urban Water Resources and Environment (Harbin [17] H. Zhang, R. Deng, H. Wang, Z. Kong, D. Dai, Z. Jing, W. Jiang, Y. Hou, Reduction
Institute of Technology) (Grant no. 2016DX11), the National Natural of bromate from water by zero-valent iron immobilized on functional

122
Z. Zhang et al. Chemical Engineering Journal 344 (2018) 114–123

polypropylene fiber, Chem. Eng. J. 292 (2016) 190–198. [37] C. Li, Q. Zhang, Y. Wang, H. Wan, Preparation, characterization and catalytic ac-
[18] G.V.L. And, M. Reinhard, Pd-catalyzed TCE dechlorination in groundwater: solute tivity of palladium nanoparticles encapsulated in SBA-15, Catal. Lett. 120 (2008)
effects, biological control, and oxidative catalyst regeneration, Environ. Sci. 126–136.
Technol. 34 (2000) 3217–3223. [38] M.V. Lombardo, M. Videla, A. Calvo, F.G. Requejo, G.J. Soler-Illia, Aminopropyl-
[19] C.G. Schreier, M. Reinhard, Catalytic hydrodehalogenation of chlorinated ethylenes modified mesoporous silica SBA-15 as recovery agents of Cu(II)-sulfate solutions:
using palladium and hydrogen for the treatment of contaminated water, adsorption efficiency, functional stability and reusability aspects, J. Hazard. Mater.
Chemosphere 31 (1995) 3475–3487. 223–224 (2012) 53–62.
[20] L.E. Knitt, J.R. Shapley, T.J. Strathmann, Rapid metal-catalyzed hydro- [39] J. Liu, H. Huang, X. Liu, J. Xiao, S. Zhong, X. She, Z. Fu, S.R. Kirk, D. Yin,
dehalogenation of iodinated X-ray contrast media, Environ. Sci. Technol. 42 (2008) Preparation of Fe2O3 doped SBA-15 for vapor phase ortho-position C-alkylation of
577–583. phenol with methanol, Catal. Commun. 92 (2017) 90–94.
[21] D. Shuai, D.C. Mccalman, J.K. Choe, J.R. Shapley, W.F. Schneider, C.J. Werth, [40] M.S. Kumar, D. Chen, J.C. Walmsley, A. Holmen, Dehydrogenation of propane over
Structure sensitivity study of waterborne contaminant hydrogenation using shape- Pt-SBA-15: effect of Pt particle size, Catal. Commun. 9 (2008) 747–750.
and size-controlled Pd nanoparticles, ACS Catal. 3 (2013) 453–463. [41] Z. Zhang, Y. Xu, W. Shi, W. Wang, R. Zhang, X. Bao, B. Zhang, L. Li, F. Cui,
[22] Y. Ding, W. Sun, W. Yang, Q. Li, Formic acid as the in-situ hydrogen source for Electrochemical-catalytic reduction of nitrate over Pd-Cu/γAl2O3 catalyst in
catalytic reduction of nitrate in water by PdAg alloy nanoparticles supported on cathode chamber: enhanced removal efficiency and N2 selectivity, Chem. Eng. J.
amine-functionalized SiO2, Appl. Catal. B: Environ. 203 (2017) 372–380. 290 (2016) 201–208.
[23] S. Seraj, P. Kunal, H. Li, G. Henkelman, S.M. Humphrey, C.J. Werth, PdAu alloy [42] X. Zhao, Y. Jin, F. Zhang, Y. Zhong, W. Zhu, Catalytic hydrogenation of 2,3,5-tri-
nanoparticle catalysts: effective candidates for nitrite reduction in water, ACS Catal. methylbenzoquinone over Pd nanoparticles confined in the cages of MIL-101(Cr),
7 (2017) 3268–3276. Chem. Eng. J. 239 (2014) 33–41.
[24] D. Shuai, B.P. Chaplin, J.R. Shapley, N.P. Menendez, D.C. Mccalman, [43] J.K. Noerskov, T. Bligaard, A. Logadottir, J.R. Kitchin, J.G. Chen, S. Pandelov,
W.F. Schneider, C.J. Werth, Enhancement of oxyanion and diatrizoate reduction U. Stimming, Trends in the exchange current for hydrogen evolution, J.
kinetics using selected azo dyes on Pd-based catalysts, Environ. Sci. Technol. 44 Electrochem. Soc. 152 (2005) 23–26.
(2010) 1773–1779. [44] M.M. Hu, H. Emamipour, D.L. Johnsen, M.J. Rood, L. Song, Z. Zhang, Monitor and
[25] J. Liu, J.K. Choe, Z. Sasnow, C.J. Werth, T.J. Strathmann, Application of a Re-Pd control of an adsorption system using electrical properties of the adsorbent for
bimetallic catalyst for treatment of perchlorate in waste ion-exchange regenerant organic compound abatement, Environ. Sci. Technol. 51 (2017) 7581–7589.
brine, Water Res. 47 (2013) 91–101. [45] I. Toyoshima, G.A. Somorjai, Heats of chemisorption of O2, H2, CO, CO2, and N2 on
[26] L. Wei, R. Gu, J. Lee, Highly efficient reduction of hexavalent chromium on amino- polycrystalline and single crystal transition metal surfaces, Catal. Rev. 19 (1979)
functionalized palladium nanowires, Appl. Catal. B: Environ. 176–177 (2015) 105–159.
325–330. [46] A.E. Palomares, C. Franch, T. Yuranova, L. Kiwi-Minsker, E. García-Bordeje,
[27] W. Sun, Q. Li, S. Gao, J.K. Shang, Highly efficient catalytic reduction of bromate in S. Derrouiche, The use of Pd catalysts on carbon-based structured materials for the
water over a quasi-monodisperse, superparamagnetic Pd/Fe3O4 catalyst, J. Mater. catalytic hydrogenation of bromates in different types of water, Appl. Catal. B:
Chem. A 1 (2013) 9215. Environ. 146 (2014) 186–191.
[28] J. Restivo, O.S.G.P. Soares, J.J.M. Órfão, M.F.R. Pereira, Catalytic reduction of [47] J. Yan, Z. Wang, P. Zhang, L. Zhang, Y. Li, X. Li, L. Zeng, Effects of pretreatment of
bromate over monometallic catalysts on different powder and structured supports, activated carbon on the selective hydrogenation performance of Pt/C catalyst,
Chem. Eng. J. 309 (2017) 197–205. Guangdong Chem. Ind. (2017) 61–62.
[29] Y. Wang, J. Liu, P. Wang, C.J. Werth, T.J. Strathmann, Palladium nanoparticles [48] E. Diaz, A.F. Mohedano, J.A. Casas, L. Calvo, M.A. Gilarranz, J.J. Rodriguez,
encapsulated in core-shell silica: a structured hydrogenation catalyst with enhanced Comparison of activated carbon-supported Pd and Rh catalysts for aqueous-phase
activity for reduction of oxyanion water pollutants, ACS Catal. 4 (2014) 3551–3559. hydrodechlorination, Appl. Catal. B: Environ. 106 (2011) 469–475.
[30] C.M.A.S. Freitas, O.S.G.P. Soares, J.J.M. Órfão, A.M. Fonseca, M.F.R. Pereira, [49] S.W. Weller, Kinetics of heterogeneous catalyzed reactions, Catal. Rev. 34 (1992)
I.C. Neves, Highly efficient reduction of bromate to bromide over mono and bi- 227–280.
metallic ZSM5 catalysts, Green Chem. 17 (2015) 4247–4254. [50] Z. Zhu, P. Wu, G. Liu, X. He, B. Qi, G. Zeng, W. Wang, Y. Sun, F. Cui, Ultrahigh
[31] X.F. Yang, A. Wang, B. Qiao, J. Li, J. Liu, T. Zhang, Single-atom catalysts: a new adsorption capacity of anionic dyes with sharp selectivity through the cationic
frontier in heterogeneous catalysis, Acc. Chem. Res. 46 (2013) 1740–1748. charged hybrid nanofibrous membranes, Chem. Eng. J. 957–966 (2016).
[32] R.G. Rao, R. Blume, T.W. Hansen, E. Fuentes, K. Dreyer, S. Moldovan, O. Ersen, [51] R.S. Swathi, K.L. Sebastian, Molecular mechanism of heterogeneous catalysis,
D.D. Hibbitts, Y.J. Chabal, R. Schlögl, Interfacial charge distributions in carbon- Resonance 13 (2008) 548–560.
supported palladium catalysts, Nat. Commun. 8 (2017) 1–10. [52] R.I. Masel, Principles of Adsorption and Reaction on Solid Surfaces, A Wiley-
[33] Y. Wan, D. Zhao, On the controllable soft-templating approach to mesoporous si- Interscience Publication, 1997.
licates, Chem. Rev. 107 (2007) 2821–2860. [53] R.A. Adigun, M. Mhike, W. Mbiya, S.B. Jonnalagadda, R.H. Simoyi, Oxyhalogen-
[34] A.R. Martins, I.T. Cunha, A.A.S. Oliveira, F.C.C. Moura, Highly ordered spherical sulfur chemistry: kinetics and mechanism of oxidation of chemoprotectant, sodium
SBA-15 catalysts for the removal of contaminants from the oil industry, Chem. Eng. 2-mercaptoethanesulfonate, MESNA, by acidic bromate and aqueous bromine, J.
J. 318 (2017) 189–196. Phys. Chem. A 118 (2014) 2196–2208.
[35] D. Bhuyan, K. Selvaraj, L. Saikia, Pd@SBA-15 nanocomposite catalyst: synthesis and [54] U. Prüsse, K. Vorlop, Supported bimetallic palladium catalysts for water-phase ni-
efficient solvent-free semihydrogenation of phenylacetylene under mild conditions, trate reduction, J. Mol. Catal. A: Chem. 173 (2001) 313–328.
Microporous Mesoporous Mater. 241 (2017) 266–273. [55] N. Barrabés, J. Sá, Catalytic nitrate removal from water, past, present and future
[36] V. Hernández-Morales, R. Nava, Y.J. Acosta-Silva, S.A. Macías-Sánchez, J.J. Pérez- perspectives, Appl. Catal. B: Environ. 104 (2011) 1–5.
Bueno, B. Pawelec, Adsorption of lead (II) on SBA-15 mesoporous molecular sieve [56] J. Martínez, A. Ortiz, I. Ortiz, State-of-the-art and perspectives of the catalytic and
functionalized with -NH2 groups, Microporous Mesoporous Mat. 160 (2012) electrocatalytic reduction of aqueous nitrates, Appl. Catal. B: Environ. 207 (2017)
133–142. 42–59.

123

Das könnte Ihnen auch gefallen