Вы находитесь на странице: 1из 10

Applied Catalysis A, General 570 (2019) 120–129

Contents lists available at ScienceDirect

Applied Catalysis A, General


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

Pd-Cu catalysts supported on anion exchange resin for the simultaneous T


catalytic reduction of nitrate ions and reductive dehalogenation of
organochlorinated pollutants from water
Corina Bradua,d, , Constantin Căpăţa, Florica Papab, Ligia Frunzac, Elena-Alina Olarua, Grégorio Crinid,
Nadia Morin-Crinid, Élise Euvrardd, Ioan Balintb, Irina Zgurac, Cornel Munteanub
a
University of Bucharest, PROTMED Research Centre, Splaiul Independenţei 91-95, sect. 5, 050107 Bucharest, Romania
b
Institute of Physical Chemistry of the Romanian Academy, 060021 Bucharest, Romania
c
National Institute of Materials Physics, 077125 Măgurele, Romania
d
Université de Bourgogne Franche-Comté, UMR 6249 Chrono-environnement usc INRA, UFR Sciences et Techniques, 25030 Besançon cedex, France

ARTICLE INFO ABSTRACT

Keywords: The present work proposes the simultaneous removal of these classes of pollutants by a catalytic hydrotreatment processes.
Pd-Cu catalyst For this purpose, bimetallic Pd-Cu catalysts (with mass ratio Pd:Cu of 4:1) supported on macroporous strong base anion resin
Anion exchange resin were prepared by different methods. The catalysts were characterized (by XRD, SEM-EDX, XPS, AAS and H2
Nitrate reduction chemisorption) and tested in a continuous flow system. The selected catalyst preparation protocol consists in a two-step
Hydrodechlorination
method, which implies the deposition of palladium by ion exchange and the subsequent deposition of copper by controlled
Water treatment
reaction on the surface of the pre-reduced palladium. The ef-fectiveness of the catalyst in the simultaneous reduction of nitrate
and hydrodechlorination of 4-chlorophenol was demonstrated. By adjusting the initial pH and the flow rate of the aqueous
solution, nearly complete hy-drodechlorination of 4-chlorophenol can occur together with selective nitrate reduction at a
conversion of 95% and a selectivity to N2 of 92% (this value contains the contribution of all gaseous products, including the
eventually formed NOx). The bimetallic catalyst was found to remains relatively stable after 100 h of test time.

1. Introduction Organochlorine compounds are often characterized by high toxicity and


persistence, being designated as priority pollutants [4]. They are readily
Nitrate ions and organochlorine compounds such as pesticides and their bioaccumulated in aquatic and terrestrial organisms, being notorious for their
degradation by-products are pollutants of concern found in nat-ural waters ability to biomagnify in the food chain.
originating mainly from agricultural areas [1]. Therefore, significant research efforts are devoted to the develop-ment of
The nitrogen cycle is altered by anthropogenic activities, particu-larly by practical methods which are able to effectively remove these pollutants from
the intensive agricultural practices, which lead to a gradual increase of nitrate natural waters, especially from those used as water supplies.
ions concentration in natural water bodies. High concentration of nitrates in
surface water negatively affects aquatic ecosystems through its toxic effect on A series of physico-chemical and biological methods were devel-oped for
aquatic fauna or by water eu-trophication. Excess nitrates in groundwater the elimination of nitrate ions from water. Among these, cat-alytic reduction
(frequently used as a drinking water source) can lead to enhanced exposure of of nitrate is an attractive technique that would be suitable for water treatment
human po-pulation being responsible for specific diseases such as [5,6]. This method was reported for the first time by Vorlop and Take [7] and
methemoglo-binemia and several types of cancer [2,3]. presents several advantages. One of its major benefits is that no waste is
generated during the treatment process as in the case of ion exchange, reverse
Contamination of natural waters by pesticide residues, especially osmosis or biological processes. Nevertheless, a crucial aspect remains the
organochlorine compounds, also causes considerable concern. Despite their catalyst selectivity issue because the presence of nitrite and ammonium ions
usefulness in agriculture, pesticides may produce a wide range of toxic side by-products is highly undesirable in drinking water.
effects representing a potential hazard to the environment.

Corresponding author at: University of Bucharest, PROTMED Research Centre, Splaiul Independenţei 91-95, sect. 5, 050107 Bucharest, Romania. E-mail address:
corina.bradu@g.unibuc.ro (C. Bradu).

https://doi.org/10.1016/j.apcata.2018.11.002
Received 5 August 2018; Received in revised form 29 October 2018; Accepted 2 November 2018
Available online 03 November 2018
0926-860X/ © 2018 Elsevier B.V. All rights reserved.
C. Bradu et al. Applied Catalysis A, General 570 (2019) 120–129

Great efforts are made in order to obtain catalyst with high activity and performed by Centi and Perathoner [33]. They reported a conversion of 90%
selectivity towards nitrogen formation. The catalysts developed for this for chlorinated compounds (no other information provided) and suggested the
purpose are mainly supported bimetallic systems based on noble metals such possibility of simultaneous removal of nitrate and halo-genated compounds.
as Pd, Pt, Rh and Ru, promoted especially with Cu, Sn and In [8–11]. Pd-Cu Several studies investigated the influence of ni-trate and other inorganic
is, by far, the most studied metallic pair for this ap-plication and offers anions on the aqueous phase reduction of organochlorine compounds with
promising prospects [12]. Generally, Pd-Cu cata-lysts were found to have zero-valent iron or with hydrogen in the presence of Pd supported catalysts
higher activity and/or selectivity to nitrogen than Pd-Sn or Pd-In catalysts [34–38]. Different results re-garding the influence of nitrate on the
[13,14]. However, several studies claim that Pd-Sn catalysts perform better hydrodehalogenation of organic compounds on Pd/Al2O3 catalysts were
than the Pd-Cu ones [15,16]. reported. It was found that, the presence of nitrate at moderate to high
The nature of the support is also important, playing a crucial role in the concentrations (0.37–1.29 mM) had no effect on the hydrodechlorination rate
activity and selectivity of the catalytic system. Various carriers have been of trichloroethylene (0.031 mM) [37]. Other authors concluded that the
proposed, γ-Al2O3, SiO2, TiO2, hydrotalcite and carbonaceous materials hydrodehalogena-tion of 1,2-dibromo-3-chloropropane (0.018 mM) was
being among the most common ones [17–21]. Ion exchange resins have also strongly affected by the presence of 0.45 mM nitrate, which lowered the
been considered promising support candidates. There is a relatively small conversion of halogenated compounds with about 50% [38]. The absence of
number of studies dedicated to the investigation of this type of carrier. any ef-fect of NO3− on trichloroethylene dehalogenation was explained by
the lack of activity of monometallic palladium catalyst on nitrate reduction
The use of ion exchange resins as catalyst carriers for the selective (no nitrate conversion). In contrast, in the study of hydrodehalogena-tion of
reduction of nitrate was proposed by Gašparovičova et al. [22–24]. Both 1,2-dibromo-3-chloropropane, 90% of NO3− was removed.
cation and anion exchange resins were studied. Although the presence of the
acid sites of the cation exchange resin contributes to increase the selectivity
This work assesses the effectiveness of catalytic hydrotreatment process
towards N2 of the Pd-Cu supported catalysts (e.g. from 20%, for a resin free for simultaneous removal of nitrate and 4-chlorophenol in aqueous solution.
of acid sites, to 60%, for a resin with acid sites), Pd-Cu catalysts deposited on For this purpose, bimetallic Pd-Cu catalysts supported on macroporous strong
anion exchange resin showed higher specific activity (e.g. 94% selectivity to
base anion resin were prepared by different methods, characterized and tested
nitrogen at 61% nitrate conversion) [22,23]. The superior performances of the
using a continuous flow system. The reasons for choosing anion exchange
catalysts sup-ported on anion exchange resin were attributed to the higher
mobility of the anions compared to cation exchange resin support. It was also resin as support are: (i) to ensure a high dispersion and controlled distribution
highlighted that the selectivity to nitrogen is higher on anion exchange resin of the active components; (ii) to enable the NO3− internal diffusion.
carriers than on γ-Al2O3 [24]. Comparable results were obtained by Neyertz
et al. (2010) for the reduction of nitrate ions on Pd-Cu/anion exchange resin,
at a more elevated concentration, 150 ppm NO3−-N (expressed as nitrogen
2. Experimental
from nitrate) instead of the usually employed concentration of 22.6 ppm
NO3−- N (100 ppm NO3−). More recently, N2 selectivity of 100% for
2.1. Bimetallic Pd-Cu catalyst synthesis
complete nitrate conversion was claimed for bimetallic Pd-Cu catalyst
supported on anion exchange resin [25]. This notable performance was
In this work, a series of bimetallic Pd-Cu catalysts (with mass ratio Pd:Cu
achieved in a batch reactor with NO3− pre-viously retained on catalyst surface
of 4:1) supported on an anion exchange resin was prepared. The synthesis of
by ionic exchange and then main-tained for 6 to 24 h under 6 bar H2/CO2 bimetal-resin catalysts was carried out via two main routes:
pressure (hydrogen partial pressure of 3 bar) at moderate temperature (298 K (i) introduction of both metals by ion exchange with the functional groups of
and 333 K). The selective conversion of NO3− to N2 is explained by the fast the resin; (ii) introduction of palladium by ion exchange and subsequent
retention of the HO−, resulted in NO3− hydrogenation process, on the ion deposition of copper by controlled surface reaction. The carrier was a
exchange sites of the resin vacated by NO3− consumption during reaction commercial anion exchange resin, A-520E Purolite, having poly-(styrene di-
vinyl benzene) matrix (S-DVB) with quaternary am-monium functional
pro-gress. A 100% selectivity to nitrogen has never been reported pre-viously.
The only drawbacks of this method could be the demanding time or groups (-CH2-N(CH3)3+) (ion exchange capacity of 2.8 molc kg−1; ionic
temperature and pressure requirements in the regeneration step. form: Cl−) and a macroporous texture (average pore size 50 nm), kindly
provided by Purolite Romania.
For the ion exchange route, the metal precursors were tetra-
chloropalladate(II) and tetrachlorocuprate(II) complexes (Reaction 1). The
To the best of our knowledge, the catalytic liquid-phase reduction of precursor solutions were obtained by dissolution of PdCl2 (p.a. – Fluka) and
nitrate on bimetallic catalysts supported on ion exchange resin was so far
CuCl2·2H2O (p.a. – AppliChem) in a HCl solution (0.1–8.0 mM). The copper
investigated in semi-batch or batch reactors and no studies have been deposition was carried out simultaneously or following palladium deposition.
published on fixed-bed reactors operating in continuous flow. The obtained catalysts are denoted as PdCu-1 and PdCu-2 respectively (Table
With respect to organochlorine compounds in water, several methods like
1). For the simultaneous de-position method, the dried A520E resin in the Cl−
adsorption or advanced oxidation processes have been proposed for their
ionic form (for 24 h, at 50 °C) was contacted with a solution containing both
removal [26,27]. An alternative approach might be their catalytic metal pre-cursors and maintained under continuous mixing (GFL Test Tube
hydrodechlorination which leads to less toxic organic compounds, thereby Rotator 3025, at 15 rpm) for 0.5–24 h. The reduction of metals previously re-
reducing the ecotoxicity of the treated water. Noble metals, preponderantly tained on the resin by ion exchange was achieved using a solution of sodium
Pd, on various supports have been stu-died for the liquid phase formate (1.5 M) (contact time of 0.5 h, at 50 °C). Finally, the obtained
hydrodechlorination of organochlorine com-pounds from water such as catalyst was washed with demineralized water until the complete elimination
organochlorine pesticides, chlorophenols and chloroacetic acids [11,28–30]. of residual Cl− and HCOO− ions (verified through ion chromatography). In
Complete dehalogenation of mono-and poly-chlorinated organic compounds the case of successive metal deposi-tion by ion exchange, an intermediate
in aqueous solution (e.g. chloroacetic acids and chlorophenols at 0.02 - 0.2 reduction step with HCOONa was realised between the palladium and the
mM) has been achieved in the presence of palladium-based catalysts using copper retention stages (Table 1).
hydrogen under ambient temperature and pressure [29,31,32].

The studies reported in the literature dedicated to treatment of poly-


contaminated waters are scarce. Preliminary catalytic reduction tests with tri- 2(S-DVB)-N(CH3)3+Cl− + [MCl4]2− → [(S-DVB)-N
and tetra-chloroethylene added to a nitrate solution were (CH3)3+]2[MCl4]2− + 2Cl− (1)

121
C. Bradu et al. Applied Catalysis A, General 570 (2019) 120–129

Table 1
− − −1
Synthetized Pd-Cu/anionic resin catalysts and their performance in the reduction of NO 3 (Preliminary tests: NO3 initial concentration = 300 mg·L ; solution flowrate = 1.0
−1 −1
mL·min ; H2 flow rate = 25 mL·min , pHi = 6.0).
Catalyst, notation Metal deposition Metal amount, %** NO3− convertion,
%
First stage* Second stage* Pd Cu

Pd-0 Pd – ion exchange – 2.01 – 13.9


PdCu-1 Pd & Cu – ion exchange – 1.97 0.45 14.7
PdCu-2 Pd – ion exchange Cu – ion exchange 1.98 0.48 32.5
PdCu-3 Pd – ion exchange Cu – controlled surface reaction 2.01 0.49 48.8

* followed by reduction with HCOONa.


** by AAS.

where M = Pd or Cu Quanta 3D FEG model, operating at 20 kV, equipped with an energy


The second preparation route consisted in a controlled surface re-action in dispersive X-ray (EDS) spectrometer, Apollo X. The analyses were per-
which copper ions were reduced by the hydrogen adsorbed on the pre-reduced formed in high vacuum mode, with Everhart–Thornley secondary electron
Pd surface. (Reaction 2) [39]. Thus, after the retention of palladium by ion (SE) detector. The preparation of the samples was minimal and consisted in
exchange and its reduction with sodium formate (stages described in the immobilizing the beads on a double-sided carbon tape, without coating.
precedent protocol), a Cu(NO3)2 solution ([Cu (NO3)2] = 3.0 g·L−1, pH < 3)
was contacted with the monometallic precursor for 0.5 h. at 50 °C. The
obtained catalyst, denoted as PdCu-3 (Table 1), was washed with
demineralized water until the complete elimination of residual NO3−, Cl− and 2.3. Experimental setup and analytical methods
HCOO− ions (verified through
ion chromatography). The catalytic reduction tests were carried out at room temperature
(21.0 ± 1.0 °C) in a continuous flow system with a fixed bed reactor (trickle
2Pd-H + Cu2+ → Pd2Cu + 2H+ (2) bed reactor) containing 0.4 g catalyst (1cm3 bulk expanded volume). The
The theoretical concentration of Pd and Cu in the bimetallic cata-lysts was aqueous solution circulates at the flow rate of 0.25 to 1.0 mL·min−1,
2% and 0.5%, respectively. Monometallic catalyst with 2% Pd (denominated concurrently with the gas flow (25 mL·min−1). The in-itial concentrations of
nitrate and 4-chlorophenol (4-CP) were 100 mg·L−1 and 50 mg· L−1
Pd-0) was also prepared for comparison.

respectively. For the preliminary tests, a concentration of 300 mg L−1 NO3−


2.2. Catalyst characterization was used. The pH correction (when necessary) was achieved using an
adequate NaNO3/HNO3 ratio (total NO3− concentration was maintained at
The exposed Pd surface area on resin was estimated by H2 chemi-sorption 100 mg ·L−1). Prior steady-state establishment, the elimination of nitrate ions
at 25 °C using a ChemBet-3000 Quantachrome Instrument equipped with a
from solution takes place both by ion exchange and catalytic reduction
thermal conductivity detector (TCD). Before the H2 chemisorption pathways as suggested in Fig. 1.
measurements, the catalysts were reduced with H2 at 70 °C for 1 h then The same experimental setup was employed to evaluate the reten-tion of
cooled to room temperature under Ar flow to remove the adsorbed H2 on the 4-CP onto catalyst (by adsorption or by ion-exchange). The re-tention tests
catalyst surface. were performed in Ar flow (25 mL·min−1) for different pH values.
Atomic absorption spectrometry (AAS) was used to determine the metal
loading on catalysts. A Thermo Elemental, Solaar M5 atomic ab-sorption
spectrometer with flame atomizer was used in absorption mode (measuring The removal of nitrate, the formation of nitrite and ammonia, and the
release of the chloride anion by the anion exchange resin were quantified
time = 4 s; air : acetylene flow = 1.2 L·min−1; bandwidth = 0.5 nm;
through ion chromatography. The analyses were performed on a Dionex ICS-
background - D2 Quadline lamp). For the ana-lysis, the samples were digested 900 apparatus equipped with a conductivity detector and anion or cation
in a microwave digester (Ethos Sel, Milestone) using a mixture of hydrogen suppressors. An AMMS 300 column was used for the anionic species. The
peroxide and HNO3 : HCl (3:1). mobile phase was a solution of mixed sodium carbonate and sodium acid
The crystalline structure of the active phase constituents was ana-lysed
carbonate flowing with 1 mL·min−1. For ammonia, a CMMS 300 column and
with D8 Advance Bruker-AXS diffractometer using Cu Kα radiation (k = 1.54
a solution of methanesulfonic acid as mobile phase were used. The
Å). The diffraction patterns were recorded in the 2Ɵ = 30 −135° domain. The
selectivities to nitrite, ammonium and molecular nitrogen were calculated by
average crystallite size was calculated from peak broadening by using a
the equations below:
Lorentzian function to fit the peak shapes.

X-ray photoelectron spectroscopy (XPS) measurements were carried out,


using a SPECS photoelectron spectrometer with a PHOIBOS 150 analyser, to
determine the chemical state of the active components in the synthesized
catalysts. The X-ray source, XR-50, operates on an Al anode (hν = 1486.7
eV) at 300 W, 24 mA and 12.5 kV. The acquisition was made with a pass
energy of 10 eV for individual spectra and 50 eV for the extended ones. The
samples were deposited on a silicon sub-strate, and a flood gun system was
used for neutralization of the sample charging. The C 1s line of the support
was used as standard for internal calibration. The data were fitted with Voigt
functions using SCF 2.3 software.

The distribution of the metals on the resin surface was evaluated by


Scanning Electron Microscopy coupled with Energy-dispersive X-ray Fig. 1. Schematic representation of nitrate removal over Pd-Cu/anionic resin catalysts.
spectroscopy (SEM-EDX), using a high-resolution microscope, FEI

122
C. Bradu et al. Applied Catalysis A, General 570 (2019) 120–129

Fig. 2. Dependence of copper retention onto A520E resin on HCl concentration in precursor solution ([Cu] = 1.3 g·L −1).

S −=
n −
NO 2 t
NO2 x100 (at λ = 275 nm) and Zorbax SB-C18 column. The mobile phase was a
n − −n −
NO3 0 NO3 t (3) mixture of phosphate buffer solution (20 mM)/acetonitrile, with a ratio
n of 50/50, flowing with 1.2 mL/min.
+
NH4 t Chromatographic analyses were performed in duplicate. The con-
SNH+ = x100
4
n −
NO3 0
−n −
NO3 t (4) versions and selecvities were calculated using mean values of NO3−,
NO2− and NH4+ concentration in 5 aliquots collected in 2.5 h of time
S N = 100 − (SNO − + SNH+) (5) interval. An exception was made for the endurance test where the va-
2 2 4
lues were calculated for periods of 5 h (average calculated from 10
where nNO3−0 is the initial amount of nitrate (mmol) andnNO3−t , nNO 2−t and consecutive aliquots). Typical investigated steady-state period was 6 h
nNH4+t are the amounts of nitrate, nitrite and ammonium at the time t and 86 h for the endurance test. Also, three selected experiments were
(mmol). The selectivity to nitrogen was calculated by equation (5). The performed twice in order to verify the reproducibility. According with
SN2 value contains the contribution of all gaseous products, which our calculation, the relative uncertainty lies between 0.6% and 1.3% for
eventually are formed beside N2 (i.e. N2O or NO). This is considered −
NO3 and 4-CP conversions and between 0.9%–3.4% for the selectiv-
reasonable assumption since the NOx intermediates generated in the ities to nitrogen species.
reduction of nitrate (Reactions 3–9) on Pd-based catalysts (Pd, Pd-Cu, For the hydrodechlorination of 4-CP, the carbon mass balance of the
Pd-In) have been often reported to have negligible small concentration reaction was closed to > 95%. Total N mass balance cannot be calcu-
in the gas phase [25,40–45]. For bimetallic Cu-Pd catalysts with atomic lated since gaseous nitrogen species were not assessed.
ratio Pd/Cu = 0.25, the selectivity to N2O in the hydrogenation of ni- The Pd and Cu leaching, during the experiments was evaluated
trate or nitrite was significantly increased as reported by Nakamura through inductively coupled plasma mass spectrometry (ICP-MS)
et al. [46]. However, in the presence of bimetallic Pd-Cu catalyst with measurements, using a Thermo Fisher Scientific X Series 2 system II with
atomic ratio Pd/Cu ≥ 1, it is probable that N2O produced is rapidly an ESI-SC2 DX fast bundle autosampler. The analysis reliability was
reduced to nitrogen as suggested in [45] which could explain the low or assessed with certified reference materials (drinking water ERM-
undetectable concentration of N2O. Regarding the nitric oxide, its CA011b, European Reference Materials) and was within 10% of the
presence was evidenced exclusively in the adsorbed form [42,44]. certified values.
Nevertheless, the presence in our outlet gas of other nitrogen species
beside N2 cannot be ruled out completely. So, the calculated SN2 could
3. Results and discussion
be more precisely defined as selectivity to nitrogen gaseous species

NO3− + H2 → NO2− + H2O (6) 3.1. Catalyst preparation


2 NO2− + H2 → 2 NO + 2 HO− (7) It was found that immobilization of metals by ion exchange onto the
2 NO + H2 → N2O + H2O (8) resin depends on acidity of precursor solutions. The effect of the HCl
concentration on the metal retention from precursor solution was sig-
2 NO + 5 H2 → 2 NH4+ + 2 HO− (9) nificant in the case of copper. As an example, for Cu2+ solutions of
N2O + H2 → N2 + H2O (10) 1.3 g·L−1, the HCl concentration was varied from 0.5 to 8.0 M. The
maximum copper retention was obtained for HCl 6.0 M (see Fig. 2).
2 NO2− + 3 H2 → N2 + 2 H2O + 2 HO− (11) This outcome might be explained by the interplay of two phenomena:
− + − (i) dependence of [CuCl4]2− complex formation on the concentration of
NO2 + 3 H2 → NH4 + 2 HO (12) chloride ions; (ii) competition between [CuCl4]2− and Cl− for the ex-
The concentrations of 4-chlorophenol and phenol (identified as the change sites of resin. Thus, low concentrations of HCl may be in-
only hydrodechlorination reaction product) were measured using a sufficient for the formation and stabilization of the [CuCl4]2− complex
Varian ProStar liquid chromatograph equipped with an UV-VIS detector [47], while an excessive Cl− concentration would limit the retention of

123
C. Bradu et al. Applied Catalysis A, General 570 (2019) 120–129

Fig. 3. Micro-photos of Pd deposition on A520E spherical beads: (a) [HCl] = 0.3 M, contact time of 0.5 h; (b) [HCl] = 0.5 M, contact time of 1 h; (c) [HCl] = 2.0 M, contact time of 24
h. (complete retention of palladium on the ion exchange resin).

[CuCl4]2− on the ion exchange sites due to the competition between the two to nitrogen was obtained [12]. However, Gašparovičová et al. [22] obtained a
anions. more active catalyst by simultaneous deposition of Pd and Cu onto cation
The amount of Pd deposited onto resin was only slightly affected by the exchange resin, as compared to a successive deposition method.
HCl concentration (data not shown). However, the variation the HCl
concentration from 0.1 to 6.0 M and/or the contact time between pre-cursor
solution and anionic resin (0.5–24 h), induced changes in Pd radial NO3− + 2Cu ⇆ Cu2O + NO2− (13)
distribution. Thus, peripheral to uniform distribution of palla-dium could be Cu2O + 2Pd-H ⇆ 2 CuPd + H2O (14)
obtained, as illustrated in Fig. 3. We have chosen a broad shell distribution of
metal (see Fig. 3b) because it is a convenient compromise between an optimal Based on the results obtained in terms of nitrate removal degree, PdCu-3
use of the support and minimization of mass-transport phenomena, which is was selected for further investigations.
an important factor in NH4+ for-mation mechanism [19].
3.3. Catalyst characterization
It is worth mentioning that copper retention from the Cu(NO3)2 solution
It is expected that selected deposition protocol ensures a good dis-persion
on the palladium monometallic catalyst was complete by controlled surface
reaction. It can be safely assumed that the Cu de-position occurred on pre- of the metal onto support. To verify this assumption, catalysts with higher
reduced Pd surface (see catalyst character-isation). metal contents were prepared (palladium up to 7%, keeping similar Pd-Cu
ratio). The palladium crystallite dimension cal-culated from the XRD
diffractograms was around 7 nm for the bime-tallic PdCu-3 catalyst (2% Pd).
Increasing palladium content to 4% and 7%, the crystallite average size
3.2. Preliminary tests remains practically unchanged (Table 2). This reveals a suitable dispersion of
palladium deposited through the ion exchange method.
The catalysts prepared as afore mentioned, were subjected to pre-liminary
tests for nitrate reduction. In the beginning, the removal of nitrate ions takes It should be noted that the XRD diffraction pattern of Pd-Cu cata-lysts
place both by ion exchange and catalytic reduction pathways (Fig. 1). After a showed only the main characteristic peaks of palladium metal with face-
period of 390–520 minutes, depending on the catalyst, the steady-state was centered-cubic structure. The characteristic diffraction lines of copper could
attained, and the elimination of nitrate ions can be assumed to take place only not be observed (Fig. 4). The low crystallinity of the cooper phase is a
by catalytic reduction process. The nitrate removal degree calculated at pertinent explanation for the absence of XRD dif-fraction lines [51].
steady-state for PdCu-1, PdCu-2 and PdCu-3 bimetallic catalysts as well as
for Pd-0 monometallic reference, is present in Table 1. The PdCu-3 diffractogram does not provide any information re-garding a
possible Pd-Cu interaction, but the calculated lattice para-meter (Table 2)
It is generally accepted that copper acts as a promoter in the first reaction allowed us to assume that the formation of a Pd-Cu alloy is possible. Thus, an
step consisting in nitrate conversion to nitrite [21,48,49]. As expected, all average composition of the nanocrystals calculated from XRD data according
bimetallic Pd-Cu catalysts lead to higher nitrate conver-sion than the to Vegard's law:
monometallic Pd catalyst (Table 1). Comparing the bi-metallic catalysts, it
can be observed that the second metal deposition procedure strongly affect aalloy = x·aPd +(1 − x)aCu
catalyst performance. Higher nitrate removal was obtained for the catalyst where a represents the lattice constant (aPd = 3.8907 Å, aCu = 3.6149 Å) and
prepared by deposition of copper by controlled surface reaction on the pre- x is the fraction of the element was found to be
reduced palladium, PdCu-3. This might be the result of a better contact Pd Cu .
98.8 1.2
between the two metals achieved in the case of PdCu-3, as compared to
The chemisorption of H2 was used to asses Pd qualitatively and
catalyst in which cooper was introduced by ion exchange, especially to PdCu-
quantitatively on monometallic and bimetallic catalysts (H2 is adsorbed
1 prepared by si-multaneous deposition of cooper and palladium. selectively only by the exposed Pd sites). The stoichiometry of the

Assuming that cooper is oxidized during redox conversion of nitrate to Table 2


nitrite and then subsequently reduced by palladium activated hy-drogen Lattice parameter and crystallite size of palladium in the bimetallic Pd-Cu catalyst with
(Reactions 10 and 11), a close contact Pd-Cu is required to enable the different metal load.
hydrogen spill-over from noble metal to promoter [49,50]. These results are in
Palladium content, % Lattice parameter a, Å Crystallite size,
agreement with those obtained by Neyertz et al. (2010), who found a higher ( ± 0.0005) nm
nitrate conversion for resin supported Pd-Cu catalyst prepared by controlled ( ± 0.5)
surface deposition of copper, as compared to Pd-Cu catalyst obtained by ion
exchange. A two-step method was also considered more favourable for the 2 3.8875 7.4
4 3.8894 7.1
preparation of Pd-Cu/Al2O3 catalysts versus a single step procedure as higher 7 3.8908 7.3
selectivity

124
C. Bradu et al. Applied Catalysis A, General 570 (2019) 120–129

Fig. 4. XRD pattern of bimetallic Pd-Cu catalyst with different metal load.

Fig. 5. SEM-EDX analysis of PdCu-3 catalyst.

adsorption H:Pd atoms was assumed to be 1:1 [52]. The H2 chemi-sorption result of the selected preparation protocol. The copper deposited in the second
also reveals a good dispersion of palladium. The surface area of exposed stage by catalytic reduction follows the same pattern as the first metal,
palladium for the monometallic catalyst (Pd-0) is around 2.37 m2·g−1 but confirming thus the deposition of copper on the palladium sur-face.
decreases to 1.15 m2·g−1 for the bimetallic PdCu-3 catalyst. The explanation
The composition and chemical state of the catalytic active phase of PdCu-
in exposed metal surface area is that Pd sur-face became partially covered
3 was also evaluated by XPS. In the region of Pd3d, well sepa-rated spin-orbit
with copper phase. This result is sup-ported by SEM-EDX investigations. The
bimetallic catalyst of choice (PdCu-3) was analysed by Scanning Electron components ( = 5.26 eV) are present, that were also found in the spectra of
Microscopy and the metals were mapped on the cross section of the catalyst our samples. The XPS spectra are shown in Fig. 6, together with the results of
beads by Energy-dis-persive X-ray spectroscopy (Fig. 5). The palladium their decomposition into components using Voigt functions. The curve fitting
mapping shows that the metal is generally well distributed in a broad shell, shows that Pd3d spectra of the samples consist of two main doublets
with a higher density of particles on the outer part and no visible characteristic for palladium atoms in different oxidation states. The Pd in
agglomeration as a metallic state, with binding

125
C. Bradu et al. Applied Catalysis A, General 570 (2019) 120–129

Fig. 6. Position of the binding energy for deconvoluted XPS peaks of (a) Pd3d5/2 and Pd3d3/2 core level and (b) Cu2p3/2 core level for PdCu-3 catalyst: fresh (up) and after the
endurance test (down).

energy (BE) at 335.0–335.3 eV, was the predominant component (≈ 86%)


[53,54]. The position of the 337.0–337.4 eV band in the Pd3d5/2 region is
indicative of Pd2+. This BE value is intermediate between those of PdO and
of PdCl2 reported in the literature [54–56].
It is probable that palladium reduction is not complete, and a small part of
the metal subsists in the chloride form. The spectrum of, [PdCl4]2− precursor
retained by ion exchange on resin shows in Pd3d5/ 2 region a characteristic
BE band at 337.2 eV (SI, Fig. S1). The surface oxidation of the palladium in
contact with air during drying and storage of the catalyst cannot be fully
excluded. The presence of Pd2+-Pd° mixture might have a beneficial
contribution in the hydrodechlorina-tion process. It was reported that both
electro-deficient and zero-valent species are required for the
hydrodechlorination of the organic chlori-nated compounds [32].

Cu2p3/2 binding energies are located near 932.5 eV (Fig. 7b). Al-though
the multivalent state of Cu complicates the discrimination be-tween distinct
Cu species, the deconvoluted peak located at 932.1 eV is close to metallic Cu°
state but can be also attributed to Cu+ in Cu2O (932 eV). The satellites
characteristic for the 2+ oxidation state cannot be well distinguished (XPS
Fig. 7. The evolution of the nitrate, nitrite, ammonium and chloride ions concentrations
spectra of Cu for binding energy in the range of 920–980 eV are presented in
SI, Fig. S2). Therefore, it can be said that copper is preponderantly found in during the hydrotreatment process (NO3− initial concentration = 100 mg·L−1; solution
flow rate of 1.0 mL·min−1; H2 pressure of 1 atm; pHi 6.1).
metallic state and/or as Cu+ [20]. Unfortunately, the Auger CuLMM spectrum
could not be used to discriminate between different oxidation states of copper
due to its very poor resolution.
processes. After this period, NO3− concentration gradually increased reaching
a plateau at 30 mg L−1 when the ion-exchange sites have been exhausted (as
can be deduced from the drop of Cl− concentration of effluent). The dynamic
3.4. Catalytic tests of the nitrite concentration followed a similar trend which is consistent with
the partial retention of NO2− on the resin sites. Since the NH4+ was not
3.4.1. Nitrate reduction retained on the resin, its equilibrium concentration in the effluent was more
The selected catalyst (PdCu-3) was firstly tested for nitrate reduc-tion rapidly established compared to NO3− and NO2−. The steady-state was
reaction. A first experiment was carried out at nitrate solution flow rate of 1.0 attained, after the complete displacement of Cl− from the ion-exchange
mL·min−1 (without pH correction) and H2 pressure of 1 atm. As can be centres of the resin by NO3−, when the removal of nitrate underwent only by
observed from Fig. 7, the nitrate was completely removed within the first 6 h catalytic re-duction. After the steady-state was reached, the conversion of
of experiment. The complete elimination of NO3− was achieved through nitrate
combined ion-exchange and catalytic reduction

126
C. Bradu et al. Applied Catalysis A, General 570 (2019) 120–129

pressure was not anymore considered as an effective strategy in our case.

The negative influence of high pH values on the selectivity toward N2 is


well known [6]. The pH control is classically achieved by addition of strong
acids, such as HCl, by bubbling CO2, or by using formic acid both as
reductant and in situ source of CO2 [16,49,58,59]. Catalysts containing acid
sites, like carboxylic group, are also proposed to pre-vent the excessive
increase of pH during the reaction [22,60]. Ac-cording to the literature data, it
is recommended to maintain the pH in the range of 5.0–6.0, which is true for
batch systems [16,18,57,58]. In a continuous flow system, a gradient of pH is
generated along the cata-lytic bed. Thus, the experiments were started with
more acidic solutions (initial pH of 2.7–3.1). In order to limit the number of
anionic species competing for the sites of ion exchanging resin carrier, the pH
correc-tion was realized by conveniently adjusting the NaNO3/HNO3 ratio
(the total NO3− concentration was kept at 100 mg ·L−1). The conversion
values measured in this case are similar with those obtained for a so-lution
with initial pH of 6.1. Moreover, the selectivity to the molecular nitrogen was
higher, increasing up to 93% (Table 3).
Fig. 8. Flow rate influence on the conversion and selectivity of the nitrate re-duction on
− −1
PdCu-3 (NO3 initial concentration = 100 mg·L ; H2 pressure of 1 atm; pHi 6.1).

3.4.2. Hydrodechlorination of 4-chlorophenol


The 4-chlorophenol (4-CP) was chosen as a model pollutant for
was established at around 70%. As expected, the increase of contact time by hydrodechlorination reaction due to its simple molecular structure and the
lowering the flow rate of the aqueous solution lead to aug-mentation in nitrate
large number of studies reported on the phenolic compounds. In addition, 4-
conversion (Fig. 8). The conversion of NO3− was higher than 93% for a flow CP is frequently present in the anthropic impacted sites being a degradation
rate of 0.50 mL·min−1 and 100% at 0.25 mL·min−1. There was also an product of a series of pesticides such as 2,4-di-chlorophenoxiacetic acid and
enhancement of selectivity towards molecular nitrogen from 49% to 65%, due pentachlorophenol [61–63].
to the further reduction of nitrite intermediate. In contrast, the undesired The S-DVB polymer is a good sorbent for phenols in water and it finds
increase of selectivity to ammonium from 19% to 32% was observed. some applications in this field [64]. Moreover, phenols retention by ion
exchange on the functional groups of the resin could be ex-pected. Indeed,
Literature data reveal that the selectivity of catalytic nitrate re-duction retention tests performed for 4-CP (50 mg ·L−1), at different pH values show
strongly depends on pH and on hydrogen pressure, which correlates with its
that the chlorinated phenol could be retained on the A520E resin both by
solubility in the aqueous solution [39,50]. Lowering the hydrogen pressure is sorption on the S-DVB matrix and by ion exchange mechanism, depending on
typically achieved by diluting H2 either with N2 or CO2 [10,40]. The use of the solution pH. As can be seen in Fig. 9, a more important retention of 4-CP
carbon dioxide contributes also to pH adjustment by acting as a buffer, accompanied by chloride release from the resin ion exchange centres is
avoiding thus the pH increase due to hydroxide ions formation (Reactions obtained at pH of 9.8, as compared to the pH of 5.9. It can be assumed that
4,6,8 and 9). Generally, it was observed that ammonia generation is hindered sorption on the S-DVB matrix is the predominant route of 4-CP retention at
at low hydrogen pres-sure [57]. pH below the pKa of the organic molecule (pKa 4-CP = 9.41), while at pH
above the pKa, retention of 4-CP occurs in an increased proportion through
Thus, to limit the formation of ammonium, the hydrogen partial pressure ion exchange. Anyway, considering that the present experiments are per-
was lowered to 0.50 and 0.25 atm by dilution with Ar (we avoided using N 2 formed in acidic medium (selected condition: 3.1 < pH < 7.7), the retention of
which is a reaction product). The nitrate solution flow rate for these catalytic 4-CP and/or phenol resulted in the hydrodechlorination process (pKa phenol =
tests was 0.5 mL·min−1. In these experimental conditions, a reduction in 9.95) takes place preponderantly by sorption.
ammonium selectivity could be observed, but the gain in nitrogen selectivity
is not significant, since the selectivity to nitrite increases as well (Table 3). The 4-CP hydrodechlorination tests performed with PdCu-3 catalyst
Moreover, the conversion drops with 14–23%, when the hydrogen pressure revealed its high efficiency for this reaction. Complete conversion of 4-CP to
was reduced to 0.50 and 0.25 atm, respectively. Consequently, the reduction phenol (the only reaction product) was obtained for the para-meters selected
of hydrogen partial for the nitrate reduction process (aqueous solution flow rate of 0.5 mL·min−1,
initial pH of 3.1 and H2 pression of 1 atm). The conversion of the 4-CP was
Table 3 evaluated for a time longer than 800 min, after the end of the sorption
Influence of H2 partial pressure and pH on catalyst performances. (For the H2 partial process. The reaction advancement fol-lowed the calculation of carbon mass
pressure tests, the initial pH was 6.1; For pH tests, the H2 pressure was of 1 atm; Nitrate balance.
solution flowrate was 0.5 mL·min−1 for all tests).

Parameters Conversion, % Selectivity, %


3.4.3. Simultaneous removal of nitrate and of 4-chlorophenol
Further the PdCu-3 catalyst was tested for simultaneous reduction of
NO2− NH4+ N2 NO3− and dehalogenation of 4-CP under the previously selected con-
ditions. After the steady-state has been reached, the conversion of 4-CP
pH2, atm and that of NO3− are of 97% and 93% respectively, while the selectivity
1.00 93.5 15.8 20.3 63.9
0.50 79.2 18.1 16.9 65.0 to nitrogen of 90% is slightly smaller than for the simple nitrate re-
0.25 69.7 18.3 12.0 69.7 duction process. This small decrease could be assigned to the compe-
pHi/pHf * tition between nitrite ions and 4-CP for the active sites of monometallic
6.1/10.9 93.5 15.8 20.4 63.8 palladium.
3.1/8.7 94.4 1.1 6.1 92.8
2.7/4.1 87.5 1.1 8.8 90.1
Fine tuning of the experimental parameters, aqueous solution flow
rate and pH, led to the improvement of hydrotreatment process per-
* - pHi – initial pH (of the nitrate solution); pHf – final pH (of the treated formances Thus, for an initial pH of 2.9 and a flow rate of
solution). 0.4 mL·min−1, the nitrate conversion reached 95% with a selectivity to

127
C. Bradu et al. Applied Catalysis A, General 570 (2019) 120–129

−1 −1
Fig. 9. Time dependence of 4-CP retention on PdCu-3: at pHi 5.9 (a); and 9.8 (b) ([4-CP] = 50 mg·L ; solution flow rate of 0.5 mL·min ; Ar pressure of 1 atm).

total). Insignificant leaching of palladium was found, the global loss is


estimated at less than 0.02% of the amount contained by the fresh
catalyst. As expected, the copper leaching was higher, especially in the
first ten hours. The maximal concentration of copper in treated water
was of 5.8 μg·L−1. However, the amount of leached copper did not
exceed 0.3% of the initial amount.
Furthermore, the XPS analysis revealed that there are no significant
differences between palladium spectra of fresh and spent catalyst. An
additional contribution situated at 935.7 eV was detected for the used
catalyst as compared to the fresh sample (Fig. 6). This value is a bit far
from the BE for CuO: The corresponding species might be copper hy-
droxide. The presence of Cu2+ in the used catalyst is in agreement with
the role expected for metallic copper in the nitrate reduction [50].

4. Conclusion

Fig. 10. Catalytic performance of PdCu-3 for simultaneous removal of nitrate It was found that the performances of the bimetallic Pd-Cu catalyst
and 4-CP ([NO3−] = 100 mg·L−1 [4-CP] = 50 mg·L−1; pHi of 2.9; solution flow supported on S-DVB anion exchange resin in the selective reduction of
rate of 0.4 mL·min−1; H2 pressure of 1 atm). nitrate strongly depend on the second metal deposition procedure. The
selected preparation protocol consisted in a two-step method, implying
Table 4 the deposition of palladium by ion exchange and the subsequent de-
Endurances tests ([NO3−] = 100 mg·L−1 [4-CP] = 50 mg·L−1; pHi of 2.9; so- position of copper by controlled reaction on the surface of the pre-re-
−1
lution flow rate of 0.4 mL·min ; H2 pressure of 1 atm). duced palladium. This protocol allowed a controlled deposition of
Time, h 15-20 95-100 metals on the support assuring: (i) a good dispersion of Pd (confirmed
by XRD and H2 chemisorption); (ii) controlled radial distribution of
− −1
NO3 –N, mg·L 1.11 2.20 metals in the resin bead (confirmed by SEM-EDX) and (iii) suitable
NO2− –N, mg·L−1 0.47 0.64 proximity between Pd and Cu (confirmed by SEM-EDX and H2 chemi-
NH4+ –N, mg·L−1 1.15 1.42
NO3− conversion, % 95.1 90.2 sorption). The catalyst obtained through this method, PdCu-3 shows
N2 selectivity, % 92.5 89.9 high nitrate removal degree and high selectivity in the reduction of
4-CP, mg·L−1 0.69 0.85 NO3− to gaseous nitrogen.
4-CP conversion, % 98.6 98.3 The effectiveness of the PdCu-3 catalyst in the simultaneous re-
duction of nitrate and hydrodechlorination of 4-CP was demonstrated
as well. Nearly complete hydrodechlorination of 4-CP can be achieved
nitrogen of 93% and the conversion of 4-CP was 99% (Fig. 10 and
simultaneously with selective nitrate reduction at a conversion of 95%
Table 4).
and a selectivity to N2 of 92% (calculated disregarding NOx contribu-
tion), by adjusting the initial pH and the flow rate of the aqueous so-
3.4.4. Catalyst stability lution.
Endurance test was performed to assess the catalyst stability. It can The PdCu-3 catalyst was found to be relatively stable for 100 h of
be observed that after 100 h of running, the catalyst performances, in running. The conversion of nitrate decreasing slightly from 95% to
terms of nitrate conversion and selectivity toward nitrogen, are little 90%, while the conversion of 4-CP seems to be unaffected. Moreover, it
affected (Table 4). The conversion diminished only with 5%, and the is noteworthy that the metals leaching was insignificant for Pd and
selectivity to N2 with less than 3%. Regarding the hydrodechlorination, moderate for Cu.
the conversion of 4-CP remains unaffected. Considering the promising results obtained in this work the next
`The amount of metals leached from the catalyst during the re- planned research step will consist of testing our proof concept for real
ductive treatment process was also evaluated. The Pd and Cu con- waters at laboratory scale as well as at pilot scale.
centrations were measured at different time intervals (10 samples in

128
C. Bradu et al. Applied Catalysis A, General 570 (2019) 120–129

Acknowledgement 33–51.
[27] M.B. Ninković, R.D. Petrović, M.D. Laušević, J. Serb. Chem. Soc. 75 (2010)
565–573.
The authors thank the Executive Agency for Higher Education, Research,
[28] G.S. Pozan, I. Boz, J. Hazard. Mater. 136 (2006) 917–921.
Development and Innovation Funding of Romania (UEFISCDI) for the [29] J. Zhou, Y. Han, W. Wang, Z. Xu, H. Wan, D. Yin, Appl. Catal. B 134–135 (2013) 222–
financial support under the PNII Project No. 100/ 2012. 230.
[30] N. Jadbabaei, T. Ye, D. Shuai, H. Zhang, Appl. Catal. B 205 (2017) 576–586.
[31] S. Zheng, D. Zhu, U.D. Patel, S. Suresh, J. Colloid Interface Sci. 319 (2008)
462–469.
Appendix A. Supplementary data [32] J.A. Baeza, L. Calvo, M.A. Gilarranz, A.F. Mohedano, J.A. Casas, J.J. Rodriguez, J.
Catal. 293 (2012) 85–93.
[33] G. Centi, S. Perathoner, Appl. Catal. B 41 (2003) 15–29.
Supplementary material related to this article can be found, in the online [34] Y. Liu, T. Phenrat, G.V. Lowry, Environ. Sci. Technol. 41 (2007) 7881–7887.
version, at doi:https://doi.org/10.1016/j.apcata.2018.11.002. [35] K. Ritter, M.S. Odziemkowski, R. Simpgraga, R.W. Gillham, D.E. Irish, J. Contam.
Hydrol. 65 (2003) 121–136.
[36] O. Schlicker, M. Ebert, M. Fruth, M. Weidner, W. Wüst, A. Dahmke, Ground Water 38
References (2000) 403–409.
[37] G.V. Lowry, M. Reinhard, Environ. Sci. Technol. 35 (2001) 696–702.
[1] C. Ravier, L. Prost, M.H. Jeuffroy, A. Wezel, L. Paravano, R. Reau, Land use policy 42 [38] D.P. Siantar, C.G. Schreier, C.S. Chou, M. Reinhard, Water Res. 30 (1996)
(2015) 131–140. 2315–2322.
[2] N.S. Bryan, H. van Grinsven, The Role of Nitrate in Human Health, in: D.L. Sparks (Ed.), [39] F. Epron, F. Gauthard, J. Barbier, Appl. Catal., A 237 (2002) 253–261.
Advances in Agronomy, Vol. 119 Academic Press Elsevier, 2013, pp. [40] J. Wärnå, I. Turunen, T. Salmi, T. Maunulasa, Chem. Eng. Sci. 49 (1994)
154–176. 5763–5773.
[3] D.M. Klurfield, N.S. Bryan, J. Loscalzo (Eds.), Nitrite and Nitrate in Human Health and [41] Y. Yoshinaga, T. Akita, I. Mikami, T. Okuhara, J. Catal. 207 (2002) 37–45.
Disease. Nutrition and Health, Humana Press, Cham, 2017, pp. 311–336. [42] S.D. Ebbesen, B.L. Mojet, L. Lefferts, J. Catal. 256 (2008) 15–23.
[4] List of Priority Substances in the Field of Water Policy- Annex II of the Directive [43] Y. Liou, C. Lin, S. Weng, H.H. O U, S.L. Lo, Environ. Sci. Technol. 43 (2009)
2008/105/EC, (2008) (Accessed 14 November 2017), http://eur-lex.europa.eu/ legal- 2482–2488.
content/EN/TXT/HTML/?uri=CELEX:32008L0105&from=EN. [44] R. Zhang, D. Shuai, K.A. Guy, J.R. Shapley, T.J. Strathmann, C.J. Werth,
[5] T.J. Strathmann, C.J. Werth, J.R. Shapley, A. Street, R. Sustich, J. Duncan, ChemCatChem 5 (2013) 313–321.
N. Savage (Eds.), Nanotechnology Applications for Clean Water, second ed., [45] S. Jung, S. Bae, W. Lee, Environ. Sci. Technol. 48 (2014) 9651–9658.
William Andrew Publishing, Oxford, 2014, pp. 339–349. [46] K. Nakamura, Y. Yoshida, I. Mikami, T. Okuhara, Appl. Catal. B 65 (2006) 31–36.
[6] J. Martínez, A. Ortiz, I. Ortiz, Appl. Catal. B 207 (2017) 42–59. [47] H. Yi, F.F. Xia, Q. Zhou, D. Zeng, J. Phys. Chem. A 115 (2011) 4416–4426.
[7] K.D. Vorlop, T. Tacke, Chem. Ing. Tech. 61 (1989) 836–837. [48] F. Epron, F. Gauthard, C. Pineda, J. Barbier, J. Catal. 198 (2001) 309–318.
[8] J. Hirayama, R. Abe, Y. Kamiya, Appl. Catal. B 144 (2014) 721–729. [49] M.A. Bahri, L. Calvo, M.A. Gilarranz, J.J. Rodriguez, F. Epron, Appl. Catal. B 141
[9] Y. Wang, J. Qu, H. Liu, C. Hu, Catal. Today 126 (2007) 476–482. (2013) 138–139.
[10] I. Witonska, S. Karski, J. Rogowski, N. Krawczyk, J. Mol. Catal. A Chem. 287 (2008) [50] W. Gao, N. Guan, J. Chen, X. Guan, R. Jin, H. Zeng, Z. Liu, F. Zhang, Appl. Catal. B 46
87–94. (2003) 341–351.
[11] B.P. Chaplin, M. Reinhard, W.F. Schneider, C. Schüth, J.R. Shapley, [51] M. Yamauchi, R. Abe, T. Tsukuda, K. Kato, M. Takata, J. Am. Chem. Soc. 133 (2011)
T.J. Strathmann, C.J. Werth, Environ. Sci. Technol. 46 (2012) 3655–3670. 1150–1152.
[12] A. Pintar, J. Batista, J. Levec, T. Kajiuchi, Appl. Catal. B 11 (1996) 81–98. [52] N. Mahata, Surf. Interfaces 4 (2016) 51–54.
[13] A. Pintar, J. Batista, I. Muševič, Appl. Catal. B 52 (2004) 49–60. [53] R. State, F. Papa, T. Tabakova, I. Atkinson, C. Negrila, I. Balint, J. Catal. 346 (2017)
[14] L. Lemaignen, C. Tong, V. Begon, R. Burch, D. Chadwick, Catal. Today 75 (2002) 43– 101–108.
48. [54] M. Brun, A. Berthet, J.C. Bertolini, J. Electron Spectrosc. Relat. Phenom. 104 (1999) 55–
[15] U. Prüsse, M. Kröger, K.D. Vorlop, Chem. Ing. Tech. 69 (1997) 87–90. 60.
[16] U. Prusse, M. Hahnlein, J. Daum, K.D. Vorlop, Catal. Today 55 (2000) 79–90. [55] M.C. Militello, S.J. Simko, Surf. Sci. Spectra 3 (1994) 402–409.
[17] O.S.G.P. Soares, J.J.M. Órfão, M.F.R. Pereira, Catal. Lett. 139 (2010) 97–104. [56] S. Sharma, B.D. Mukri, M.S. Hegde, J. Chem. Soc., Dalton Trans. 40 (2011)
[18] O.S.G.P. Soares, J.J.M. Órfão, M.F.R. Pereira, Desalination 279 (2011) 367–374. 11480–11489.
[19] A.E. Palomares, C. Franch, A.A. Corma, Catal. Today 172 (2011) 90–94. [57] Y.X. Chen, Y. Zhang, G.H. Chen, Water Res. 37 (2003) 2489–2495.
[20] F. Papa, I. Balint, C. Negrila, E.A. Olaru, I. Zgura, C. Bradu, Ind. Eng. Chem. Res. 53 [58] U. Prüsse, K.D. Vorlop, J. Mol. Catal. A Chem. 173 (2001) 313–328.
(2014) 19094–19103. [59] A. Garron, F. Epron, Water Res. 39 (2005) 3073–3081.
[21] N. Barrabés, J. Sá, Appl. Catal. B 104 (2011) 1–5. [60] A. Roveda, A. Benedetti, F. Pinna, G. Strukul, Inorg. Chim. Acta 349 (2003)
[22] D. Gašparovičová, M. Králik, M. Hronec, Collect. Czech. Chem. Commun. 64 (1999) 203–208.
502–514. [61] M.D. Mikesell, S.A. Boyd, J. Environ. Qual. 14 (1985) 337–340.
[23] D. Gašparovičová, M. Králik, M. Hronec, A. Biffis, M. Zecca, B. Corain, J. Mol. Catal. [62] C. Guillard, J. Disdier, C. Monnet, J. Dussaud, S. Malato, J. Blanco, M.I. Maldonado,
A Chem. 244 (2006) 258–266. J.M. Herrmann, Appl. Catal. B 46 (2003) 319–332.
[24] D. Gašparovičová, M. Králik, M. Hronec, Z. Vallusova, H. Vinek, B. Corain, J. Mol. [63] P. Benoit, E. Barriuso, S. Houot, R. Calvet, Eur. J. Soil Sci. 47 (1996) 567–578.
Catal. A Chem. 264 (2007) 93–102. [64] C. Pãcurariu, G. Mihoc, A. Popa, S.G. Munteanu, R. Ianoº, Chem. Eng. J. 222 (2013)
[25] Y. Kim, M.Y. Kim, M. Choi, Chem. Eng. J. 289 (2016) 423–432. 218–227.
[26] A.R. Ribeiro, O.C. Nunes, M.F.R. Pereira, A.M.T. Silva, Environ. Int. 75 (2015)

129

Вам также может понравиться