Sie sind auf Seite 1von 11

Catalysis Today xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Catalysis Today
journal homepage: www.elsevier.com/locate/cattod

Activity of an iron Colombian natural zeolite as potential geo-catalyst for


NH3-SCR of NOx
John-Freddy Gelvesa, Ludovic Dorkisb, Marco-A. Márquezc, Andrés-Camilo Álvarezd,

Lina-María Gonzálezd, , Aída-Luz Villad
a
Grupo de investigación competitividad y sostenibilidad para el desarrollo, Facultad de Ingeniería, Universidad Libre seccional Cúcuta, Cúcuta, Colombia
b
Grupo de investigación en Catálisis y Nanomateriales, Facultad de Minas, Departamento de Materiales y Minerales, Universidad Nacional de Colombia sede Medellín,
Medellín, Colombia
c
Grupo de Mineralogía Aplicada y Bioprocesos, Facultad de Minas, Departamento de Materiales y Minerales, Universidad Nacional de Colombia sede Medellín, Medellín,
Colombia
d
Environmental Catalysis Research Group, Engineering Faculty, Chemical Engineering Department, Universidad de Antioquia UdeA, Calle 70 No. 52-21, Medellín,
Colombia

A R T I C L E I N F O A B S T R A C T

Keywords: The Selective Catalytic Reduction of NOx by ammonia (NH3–SCR of NOx) was studied over an iron Colombian
Geo-catalysts natural zeolite (Nat-zeo) based geo-catalysts after several treatments (H-zeo-Fe3+ and H-zeo-Fe2+) under lean,
NH3-SCR dry and wet conditions. Nat-zeo from the Combia geological formation in Colombia (South America) was treated
Nitrogen oxides with NH4NO3, calcined in air for obtaining H-zeo-Fe3+ and then reduced with hydrogen to obtain the H-zeo-
Selective catalytic reduction
Fe2+ catalyst. Catalysts were characterized by XRD, FTIR, Mössbauer spectroscopy, SEM/EDX, XRF BET,
Iron catalysts
H2–TPR, NH3–TPD, and NOx–TPD. H-zeo-Fe3+ showed better catalytic performance than Nat-zeo and H-zeo-
Fe2+ in the NOx conversion. The presence of Fe3+ ions and the acidity in the catalyst explains partially the high
activity towards NOx reduction. Water inhibits the NO and improves the NO2 adsorption on the catalyst H-zeo-
Fe3+ surface according with the NOx-TPD analysis. After 30 h on-stream, under wet conditions, the H-zeo-Fe3+
catalyst showed a decreased in the NOx conversion. The decrease of activity could be related with the loss of
catalyst surface acidity detected by NH3-TPD analysis after reaction and with the contraction of the zeolite
channels due to the metaheulandite phase formation that was confirmed by XRD, BET and FTIR analysis.

1. Introduction compared with commercial V2O5 catalysts [8,9], and operates at a


wider range of temperatures [8–10]. The iron-based catalysts have at-
Motor vehicles are the main emission source of nitrogen oxides tracted attention owing to high NH3-SCR activity, low cost and lack of
(NOx) into the atmosphere, which reached dangerous levels and pro- toxicity relative to vanadium [10]. Iron oxides catalysts have been used
jected emission values above 1.6 Mt/year only in European light duty in the NH3-SCR, but they show low-temperature activity (between 150
diesel vehicles [1]. NOx are known to cause the formation of acid rain, and 300 °C) due to the low surface area and acidity [9,10]. On the other
photochemical smog, and ground-level ozone [2,3]. Ground-level ozone hand, iron-based zeolites have high activity at temperatures above
is a very potent greenhouse gas and it has a direct warming effect on 300 °C and in comparison with copper-based zeolites, they produce
climate [3,4]. Several researches are addressed in decreased NOx ve- fewer amounts of side products such as N2O and have better hydro-
hicular emissions varying the engine type, operating conditions, and thermal stability [11].
fuel used [5]. Notwithstanding, these efforts for reducing NOx emis- Volcanic rocks containing natural zeolites have been mined world-
sions do not solve completely the problem, and even increase them wide for more than 1000 years for use as cements and building stone.
[5,6]. In the treatment of NOx from exhausts, the three major tech- Since the 1950’s, natural zeolites have found a variety of applications in
nologies are lean NOx traps (LNT), ammonia (or urea) Selective Cata- adsorption, catalysis, building industry, agriculture, soil remediation,
lytic Reduction (NH3-SCR) and Hydrocarbon Selective Catalytic Re- and energy [12,13]. The use of natural zeolites for environmental ap-
duction (HC-SCR) [7]. The leading technology is the NH3-SCR using plications is gaining new research interests mainly due to their prop-
zeolites with Fe and Cu ions, which have better hydrothermal stability erties, and significant worldwide occurrence. A conversion of NOx


Corresponding author.
E-mail address: lina.gonzalez@udea.edu.co (L.-M. González).

https://doi.org/10.1016/j.cattod.2018.01.025
Received 26 September 2017; Received in revised form 11 January 2018; Accepted 21 January 2018
0920-5861/ © 2018 Elsevier B.V. All rights reserved.

Please cite this article as: Gelves, J.-F., Catalysis Today (2018), https://doi.org/10.1016/j.cattod.2018.01.025
J.-F. Gelves et al. Catalysis Today xxx (xxxx) xxx–xxx

above 80% in NH3-SCR was obtained over natural chabazite with iron, 2.2. Catalyst characterization
however low stability of the material was observed under hydrothermal
conditions [14]. The presence of an iron-rich natural zeolite (around XRD patterns were performed at 5° ≤ 2θ ≤ 55° in a Rigaku Miniflex
12 wt% of Fe2O3) with heulandite (HEU) framework mixed with small II diffractometer, with graphite monochromator using Cu Kα radiation
amounts of the zeolites chabazite (CHA), phillipsite (PHI) and morde- at 30 KV and 15 mA, the scan step was 0.02°, and the counting time per
nite (MOR) was recently reported in the andesitic basalts from the step was 2 s. The software used for qualitative and semi-quantitative
Combia geological formation in Colombia (South America) [15]. This analysis (Rietveld refinement) was PANalytical Xpert Highscore Plus.
natural zeolite showed activity for the simultaneous NOx and o-di- The American Mineralogist Crystal Structure Database was used for
chlorobenzene removal (conversions of: NO around 5% and o-DCB identifying the crystalline phases in the material. The FTIR spectra were
around 25% at 550 °C) using methane as reducing agent under wet and recorded in a Shimadzu 8400S spectrophotometer. Spectra were col-
lean conditions [16]. lected at room temperature, in KBr pellets ((1 mg sample) (100 mg KBr
The present study aim is to report the characterization and potential pressed at 10 t)−1), using 48 scans, at 4 cm−1 of resolution in trans-
of an iron Colombian natural zeolite [15] as geo-catalyst for NH3-SCR mittance mode, in the range of 4000 and 400 cm−1. The software used
of NOx under lean and hydrothermal conditions. The natural zeolite for the analysis was IR Solution v.1.4. Local iron ion configurations in
was tested in the reaction without any treatment and after the proton the Nat-zeo were determined by Mössbauer spectroscopy. The
incorporation for obtaining an acid material. Because Fe3+ and Fe2+ Mössbauer spectra were collected in transmission geometry at room
species may have a different effect on NH3-SCR of NO reaction [18–20], temperature, using a conventional Wissel™ Mössbauer spectrometer,
this effect was also evaluated on the natural zeolite used as catalyst. The equipped with a 57Co (Rh) source working at constant acceleration. The
fresh catalysts were characterized by X-ray powder diffraction (XRD), Mössbauer 57Fe spectrum was calibrated by α-Fe as standard. The
Fourier transform infrared spectroscopy (FTIR), Mössbauer spectro- spectrum was adjusted and analyzed using the Recoil program in order
scopy, scanning electron microscopy with energy dispersive X-Ray to obtain the hyperfine parameters (hyperfine field, quadrupole and
spectroscopy (SEM/EDX), X-ray fluorescence spectroscopy (XRF), BET isomeric change). External morphology of the natural zeolite was
surface area, temperature-programmed reduction under hydrogen (H2- evaluated with JEOL scanning electron microscopy, model JSM-
TPR), and temperature-programmed desorption of ammonia (NH3-TPD) 5910LV. The chemical microanalysis was done using SEM coupled with
and nitrogen oxides (NOx-TPD). The catalyst with the best performance EDX from INCA Oxford instruments (accelerating voltage of 15 kV,
in the NH3-SCR of NO reaction was tested during 30 h on-stream for counting time of 80 s and working distance of 10 mm). The chemical
evaluation of its durability; furthermore, the material was characterized composition of the natural zeolite was performed in an Axios X-ray
after reaction by XRD, FTIR, BET, and NH3-TPD. fluorescence spectrometer from PANalytical, using the software
WROXI. For the XRF analysis, the sample was compacted into a pellet at
20 t pressure. The surface area, pore volume and pore size distribution
2. Experimental
were determined by BET analysis in an ASAP 2020 V4.00 (V4.00 J)
equipment from Micromeritics (sample degassed at 523 K for 12 h, and
2.1. Catalyst preparation
N2 adsorption at 77 K). H2–TPR and NH3–TPD experiments were carried
out in an AutoChem II 2920 instrument (Micromeritics) equipped with
2.1.1. Extraction of natural zeolite
a thermal conductivity detector (TCD). H2–TPR of fresh samples were
Natural zeolite used in this work as Geo-catalyst was collected from
obtained by flowing 50 mL min−1 of a mixture 10% H2/Ar and heating
the riverbed of the creek “La Sucia”, located north of the La Pintada
from 50 to 1050 °C using a heating rate of 10 °C min−1. Before H2–TPR
municipality in the state of Antioquia-Colombia (South America) that is
analysis, the samples were treated under 50 mL min−1 of flowing argon
part of the volcanic member of the Combia geological formation [15].
at 350 °C for 2 h. NH3–TPD was used to determine the total acidity of
The separation process of the zeolite from the rock was performed with
samples. Before analysis the samples were flushed with flowing He
dense liquids method [21] using LST heavy liquid from Central Che-
(50 mL min−1) for 2 h at 350 °C, then the sample was cooled to 50 °C.
mical Consulting − Australia (density of 2.25 g cm−3 operating at room
The saturation adsorption of ammonia was performed at 50 °C by
temperature) [22]. Initially the rock was manually grinded in an agate
flowing 50 mL min−1 of a gas mixture 5% NH3/He for 60 min. In order
mortar, sieved in a Tyler 20 mesh (< 840 μm), and settled in the LST
to remove any physical adsorbed ammonia, the sample was flushed
heavy liquid. The material rich in the zeolites (with a density less than
with 50 mL min−1 of He for 2 h. Desorption of ammonia was obtained
2.25 g cm−3) floated and was separated by filtration. Subsequently, it
by rising the temperature at 10 °C min−1 up to 700 °C. For NOx-TPD
was cleaned in an ultrasonic bath with distilled water (Elma LC20/H)
analysis two types of experiments were performed: NO-TPD and NO2/
for 5 min in order to remove residues of LST bonded to the surface of
N2O-TPD in the presence or absence of water vapor. Prior to NOx-TPD,
the zeolites. Finally the solid was dried at 110 °C in an oven, milled and
50 mg of the sample was flushed with 50 mL min−1 of an inert gas at
sieved to a diameter < 250 μm (120 Tyler mesh). This material was
350 °C for 2 h. In the NO-TPD, the sample was saturated with 500 ppm
coded as Nat-zeo.
of NO at 50 °C for 2 h in the presence of water vapor (0–5% vol). The
sample was flushed for 4 h and finally the temperature was raised to
2.1.2. Synthesis of H-zeo-Fe3+ and H-zeo-Fe2+ materials 550 °C at 5 °C min−1. A similar procedure was followed in the NO2/
In order to facilitate the ammonium ion exchange in the Nat-zeo N2O-TPD, the sample was saturated with a mixture of 350 ppm NO2 and
material, the solid was treated with a 0.5 M solution of NaCl (Merck 20 ppm N2O at 50 °C for 2 h in the presence of water vapor (0–5% vol).
99.5%) at 90 °C and 400 rpm magnetic agitation speed, with a ratio of The gas concentration was monitored with an Antaris IGS, Thermo
(1 g of Nat-zeo) (10 mL of solution)−1. The process was performed for Scientific FTIR equipment, equipped with a MCT detector, KBr
5 days, renewing the solution every 24 h [17]; then the sample was beamsplitter, 2 m pathlength gas cell, 200 mL gas cell volume and a
washed with distilled water until negative presence of chlorine. Finally, ZnSe window. Data were collected with a spatial resolution of 0.5 cm−1
the solid was filtrated and dried at 110 °C for 24 h. The sodium rich with 32 scans for each sampling. Gas cell was operated at atmospheric
material obtained was then exchanged with ammonium nitrate pressure and the temperature was fixed at 150 °C in order to avoid
(NH4NO3) (Merck 99%) using a similar procedure as described above. condensation of the reactants inside the gas cell. A quantification model
The solid material was calcined in a muffle furnace at 360 °C during 6 h. was developed by taking several spectrums of known gas mixtures and
The final material was coded H-zeo-Fe3+. H-zeo-Fe2+ catalyst was adjusting the absorbance and concentrations data by means of partial
obtained by reduction of H-zeo-Fe3+ with 50 mL min−1 of 10% H2/Ar least square quantitative analysis, the software TQ Analyst 9.4.45
flow with a heating rate of 10 °C min−1 and at 600 °C for 2 h. (Thermo Fisher Scientific Inc.) was used for this purpose.

2
J.-F. Gelves et al. Catalysis Today xxx (xxxx) xxx–xxx

2.3. Catalytic reactions labradorite), and celadonite with 5 wt%. Similar composition in natural
zeolites used as catalysts has been previously reported [23–25]. The
Catalytic reactions were carried out in a quartz tube fixed–bed re- XRD pattern of the acid catalyst H-zeo-Fe3+ shows much similarity with
actor (0.015 m i.d. × 0.4 m length) containing a frit to hold the powder the diffraction pattern of the starting Nat-zeo material. However, there
catalyst samples. The reactor was operated at atmospheric pressure are two important changes in its pattern. The first one is associated with
(0.84 atm) in steady state plug–flow mode. Reactor temperature was the decrease in the intensity of the reflection planes of the heulandite
varied from 150 to 550 °C. The reactant gases were fed using electronic phase, Fig. 1. The second aspect is related with the presence the me-
mass flow controllers (Sierra Instruments and Brooks 5850E) and mixed taheulandite phase. The formation of this phase is associated with the
in line before entering the reactor. Ammonia was entered 100 mm transformation of the heulandite phase at temperatures above 300 °C
above the bed to assure reaction just in the catalyst, and to minimize [26], which in our case occur during the elimination of the NH4+ ions
homogenous reactions. The total flow was 100 mL min−1 at 360 °C. This dehydrated heulandite exhibits a decrease in the unit
(GSHV = 30000 h−1). The gas composition of the reactant mixture was cell parameters and in the size of the zeolitic channels [26].
445 ppm NOx (400 ppm NO, 45 ppm NO2), 400 ppm NH3, 8% O2, 0–5% In the case of the reduced catalyst H-zeo-Fe2+, the diffraction pat-
H2O, and argon balance. Water vapor was introduced into the reaction tern (Fig. 1) shows that the intensity of the reflection plane (020) of
system by means of a saturator at 60 °C, using Ar as carrier gas. Gases metaheulandite (2θ around of 11°) increased in comparison with the
were analyzed by an Antaris IGS, Thermo Scientific FTIR, as described intensity in H-zeo-Fe3+ material. The increment of this phase can be
above. NO and NH3 conversions were calculated according to Eqs. (1) associated with the disappearance of the low stability phase of the
and (2), respectively. heulandite structure during the reduction process at 600 °C; which is
evidenced by the notorious decrease in the intensity of the peak at 10°.
[NO]in − [NO]out ⎞
NO Conversion (%) = 100 ⎛ ⎜ ⎟

⎝ [NO]in ⎠ (1)
3.1.2. FTIR
[NH3]in − [NH3]out ⎞ Infrared analysis corroborates the presence of the heulandite
NH3 Conversion (%) = 100 ⎛ ⎜ ⎟
structure in the catalysts, Fig. 2. The “fingerprint” region of the heu-
⎝ [NH3]in ⎠ (2)
landite phase in the FTIR spectra of the three catalysts present the
Durability test was carried out for 30 h of continues operation at the characteristic vibrations of this zeolite, such as vibration of the pseu-
temperature where the highest conversion was obtained with the cat- dolattice rings at 604 cm−1, the asymmetric stretching vibration of the
alyst with the best performance. TO group at 1198 cm−1, and the stretching band of the TO4 group
(between 1020 and 1060 cm−1) [27,28]. However, a widening in the
3. Results and discussion band at 604 cm−1 and in the TO4 group bands is evidenced after the
heat treatment in the Nat-zeo (calcination of ammonium ions and re-
3.1. Catalysts characterization duction treatment). This widening is associated with a decreasing in the
crystallinity of the heulandite phase [27,28]. FTIR analysis also shows
3.1.1. XRD the symmetrical stretching of the OH groups at 3620 cm−1 and
X-ray diffraction (XRD) patterns of Nat-zeo, H-zeo-Fe3+, and H-zeo- 3440 cm−1 [27,28]. The presence of calcite or dolomite in the catalysts
Fe2+ catalysts are shown in Fig. 1. According with XRD analysis is is evidenced by the band at 1450 cm−1, which is associated with the
possible to establish the presence of three crystalline phases in the Nat- asymmetric tension of the CO3 group [27]. Calcite concentration in the
zeo material. Heulandite/clinoptilolite and plagioclase feldspar (anor- catalysts must be low since it was not identified by XRD.
tita and labradorite) are found as major phases, while celadonite ap- In the case of the H-zeo-Fe3+ catalyst, the band at 1405 cm−1
pears as minority phase. By means of Rietveld refinement it can be corresponds to the vibration of some NH4+ ions that remain as ex-
established that the heulandite (heulandite + clinoptilolite) phase re- change cations in the zeolite [29]. The presence of celadonite and
presents about 70 wt% of the composition of Nat-zeo material, followed plagioclase is not very noticeable by FTIR, possibly due to the over-
by calcium plagioclase 25 wt% (equal amount of anortite and lapping of the characteristic vibrations bands of the heulandite zeolite.

Fig. 1. XRD patterns of Nat-zeo, H-zeo-Fe2+, H-zeo-Fe3+ and H-zeo-


Fe3+, U (U: used, reaction conditions: 400 ppm NO, 45 ppm NO2,
400 ppm NH3, 8% O2, 5% H2O, and balance Ar, GHSV = 30000 h−1,
400 °C, 30 h on stream). O: orthoferrosilite, H: heulandite/clinoptilite,
C: celadonite, M: metaheulandite, P: plagioclase feldspar.

3
J.-F. Gelves et al. Catalysis Today xxx (xxxx) xxx–xxx

Table 1
Hyperfine parameters obtained from Mössbauer spectrum of Nat-zeo.

Site IS (mm s−1) ΔQ (mm s−1) W (mm s−1) Concentration (%)*

1 (doublet 1) 0.21 0.33 0.14 79


2 (doublet 2) 0.38 0.21 0.10 21

IS: isomer shift, ΔQ: quadruple splitting, W: width.


* Based on the peak area percentage calculated by the integration of the area under the
curve.

(see Fig. 3) allows to corroborate that the iron is found as Fe3+


[30,31,33]. According with the isomeric shift (IS) and quadrupole
splitting (ΔQ) we can establish that the iron in “site 1” is in a tetra-
hedral coordination, while in “site 2” is in octahedral coordination
[30,34]. An approximate concentration of the sites 1 and 2 is 79% and
21%, respectively (Table 1).
The hyperfine parameters obtained for Fe3+ ions at site 2 are quite
Fig. 2. FTIR spectroscopy analysis of Nat-zeo, H-zeo-Fe2+, and H-zeo-Fe3+ catalysts similar to those reported for the celadonite phase at octahedral sites.
before and after reaction (reaction conditions: 400 ppm NO, 45 ppm NO2, 400 ppm NH3, Fe3+ in this environment has been associated with the IR vibration
8% O2, 5% H2O, and balance Ar, GHSV = 30000 h−1, 400 °C, 30 h on stream). band of (Fe3+-Mg)-OH at 3550 cm−1 observed in the FTIR analysis of
the Nat-zeo, Fig. 2 [35–37]. Respecting to the site 1, this iron could be
However, the vibration band of the Mg-Fe3+ species of the celadonite at present in the structure (tetrahedral) of the zeolite and feldspar type
3550 cm−1 was evidenced in the spectra of natural zeolite (Nat-zeo) plagioclase [35].
and in the catalyst H-zeo-Fe3+ (Fig. 2) [27,28]. In the reduced material
H-zeo-Fe2+ this band is no longer evident, possibly due to the trans- 3.1.4. SEM/EDX
formation of the celadonite to a new ortoferrosilite phase (according to Fig. 4 shows the morphology of the Nat-zeo material, which allows
the result of XRD, see Fig. 1). the recognition of heulandite zeolite crystals with laminar habit and
predominance of the plane (020). This observation is in agreement with
3.1.3. Mössbauer spectroscopy the high intensity of peak at 2θ = 10° found in the XRD analysis (see
Mössbauer spectroscopy was used to elucidate the chemical en- Fig. 1). The predominant laminar habit in the material could be related
vironment and the coordination of the iron ions present in natural with the presence of the potassium, Table 2, which acts as compensa-
zeolite extracted (Nat-zeo). This technique allows checking the pre- tion cation in the structure [38]. The presence of celadonite film is
sence of the iron in the zeolite or in the minority phases such as cela- observed like a coating the surface of the zeolite crystals, Fig. 4b.
donite (Fig. 3 and Table 1). The chemical microanalysis displayed in Table 2 confirms the pre-
According with Fig. 3, the Mössbuaer spectrum of Nat-zeo catalyst sence of heulandite phase due to the predominance of calcium and
shows an asymmetric doublet with lower intensity in the left and high potassium which are their characteristic compensation cations [38].
dispersion in the baseline, which suggest a low iron content in the However, the presence of calcium as compensating cation reduces the
sample. Based on these experimental results, it is possible to establish a thermal stability of the heulandite structure [26,39]. In the case of the
model considering two sites for the iron in the natural zeolite. The celadonite film, the composition found (Table 2) is typical for this
values of the isomeric shift (IS), Table 1, lead us to conclude that the material [15]. The content of Fe3+ present in the celadonite could
iron in the Nat-zeo material is present as Fe3+ in agreement with lit- generate a synergistic effect with the acidic sites suitable for its use in
erature reports [30–32]. The data adjustment allows to identify the the selective catalytic reduction of NOx [18,23,24].
existence of quadrupole splitting in the material spectrum (Fig. 3). This
fact implies that the iron in this material is not located in an isometric 3.1.5. XRF
crystal structure. The distance between the peaks of the sites 1 and 2 The chemical analysis by X-ray fluorescence, expressed as oxides, is
presented in Table 3. The main elements present in the natural material
are silicon and aluminum, in a Si/Al ratio of 3.76. This fact permits to
identify the phases of the zeolite group as a silica heulandite with
medium thermal stability [40].
The presence of calcium is associated with the existence of felds-
pathic and zeolitic phases according to the XRD analysis. Magnesium is
directly related to the presence of celadonite and in a lesser extent to
heulandite, as evidenced in EDX results. In the case of potassium, it can
be associated to the clinoptilolite [41], celadonite [42], and in small
amounts to heulandite [43] phases. The amount of iron (4.51 wt% as
Fe2O3 or 3.15 wt% as Fe) present in the natural material (Nat-zeo)
could be sufficient for being an active site in the NH3-SCR of NOx
[18,23,24]. Traces of Ba and Sr in the Nat-zeo are associated to the
heulandite phase [44]. Ti, Sr, Mn and other elements are also present;
these elements may come from residual phases of basalt after the
concentration process, which cannot be identified by FTIR and XRD, or
be present as compensation cations in some zeolite crystals.

Fig. 3. Experimental (EXP) and fitted (FIT) room temperature Mössbauer spectrum of 3.1.6. BET analysis
Nat-zeo catalyst. Spectrum was fitted using Recoil program.
The adsorption/desorption isotherm for the natural material (Nat-

4
J.-F. Gelves et al. Catalysis Today xxx (xxxx) xxx–xxx

Fig. 4. SEM analysis of Nat-zeo catalyst.

Table 2
EDX-Microanalysis of heulandite and celadonite phases in Nat-zeo.

Heulandite Celadonite

Element % Atomic Element % Atomic

O 54.59 O 52.84
Si 33.77 Si 26.06
Al 7.47 Al 5.45
Ca 3.08 Ca 0.84
K 0.95 K 2.16
Mg 0.14 Mg 4.86
Fe 0.00 Fe 7.77

Table 3
Chemical analysis of the Nat-zeo material by XRF.

Oxide Weight%

SiO2 63.31
Al2O3 14.90
CaO 5.48
Fe2O3 4.51
K2O 1.98
MgO 1.30
BaO 0.18
SrO 0.11
TiO2 0.07
MnO 0.06
SO3 0.04
CuO 0.03
As2O3 0.03
SnO2 0.02
V2O5 0.02
SeO2 0.010
Loss of ignition 7.96

zeo) is shown in Fig. 5a and the textural properties obtained from BET
analysis of the Nat-zeo and the catalyst H-zeo-Fe3+ before and after Fig. 5. N2 adsorption/desorption isotherms of a) Nat-zeo catalyst and b) H-zeo-Fe3+
reaction are presented in Table 4. The Nat-zeo material shows a typical catalyst before (solid line) and after reaction (dotted line with symbol) (reaction condi-
adsorption/desorption isotherm type IV associated to mesoporous ma- tions: 400 ppm NO, 45 ppm NO2, 400 ppm NH3, 8% O2, 5% H2O, and balance Ar,
terials, with an hysteresis loop H2 that indicates pores with narrow GHSV = 30000 h−1, 400 °C, 30 h on stream).
mouths (ink-bottle pores), uniform channels and connectivity [40,45].
According with this BET analysis is also possible to establish that in Nat- regions where the cleavage planes are given (see Fig. 4a) [46,47].
zeo material there is a mixture of microporous (monolayer-multilayer An important change is observed in the Nat-zeo after the acid
adsorption in p/p0 lower than 0.6) and mesoporous (capillary con- treatment and calcination to obtain H-zeo-Fe3+, Fig. 5b. BET analysis of
densation in mesoporous with p/p0 above 0.9). The mesoporosity in the H-zeo-Fe3+ displays a typical shape of a zeolitic material with an iso-
material can be associated with the presence of the minor phases of therm type I and the hysteresis loop of the type H4 [45,46]. In addition,
phyllosilicates such as celadonite mica. Similarly, heulandite type it is observed a lower average pore size of H-zeo-Fe3+ in comparison
zeolite with lamellar morphology can generate mesoporosity in the with Nat-zeo material (Table 4).

5
J.-F. Gelves et al. Catalysis Today xxx (xxxx) xxx–xxx

Table 4 around 150 °C and 510 °C which are associated to weak and strong acid
Textural parameters of Nat-zeo and H-zeo-Fe3+ catalysts before and after reaction. sites, respectively [53]. After the modification of the natural zeolite, by
ion exchange, is evident the increment in the total acid concentration in
Material SBETa db (Å) Amicroc(m2 g−1) Amesod(m2 g−1) Vpe(cm3 g−1)
(m2 g−1) the H-zeo-Fe3+ catalyst (Table 5). These changes are directly associated
with the type of compensation cation present in the zeolite structure,
Nat-zeof 30.61 193.8 3.41 20.51 0.029 since it modifies the electronegativity, the ability to form hydroxyl
H-zeo- 28.28 124.5 2.58 25.70 0.088
groups and the diffusion of the NH3 into the channels of the zeolite
Fe3+
fresh [54–56].
H-zeo- 26.94 136.7 5.59 21.35 0.092
Fe3+
3.2. Activity measurements
usedg

a
Specific surface area calculated by the Brunauer–Emmett–Teller (BET) method. The conversions of NO and NH3 as a function of temperature over
b
Average pore size, derived from the BJH adsorption (4 V/A). the natural zeolite and the modified materials (H-zeo-Fe3+ and H-zeo-
c
Micropore area calculated by t-plot method. Fe2+) are compared in Figs. 8 and 9, respectively.
d
Mesopore area calculated by t-plot method. The Nat-zeo material showed a NO conversion around of 20% in the
e
Total pore volume estimated from the N2 adsorbed amount at p/p0 = 0.99.
f temperature range between 300 °C and 450 °C. This activity can be
Analysis performed in a Tristar 3000 from Micromeritics.
g
After 30 h in reaction. associated with the presence of isolated Fe species on the zeolite fra-
mework [51], probably in tetrahedral positions, according with Moss-
In comparison with many of the synthetic zeolites [48], the surface bauer spectroscopy (Fig. 3) [57]. In addition, iron species have high
area of natural materials can be considered low (Table 4); however, oxygen storage capacity and mobility based on their redox Fe2+/Fe3+
these values are typically reported for heulandite and celadonite ma- equilibrium [58].
terials [49,50]. The mesoporosity of the material is reflected by the In our case, with H-zeo-Fe3+ and H-zeo-Fe2+ catalysts (Fig. 8a), the
average pore size of the material (193.8 Å) [49]. After the treatment for maximum NO conversion was attained at around 400 °C and then it
obtaining the proton form in catalyst H-zeo-Fe3+, both the pore size decreased because of NH3 oxidation (Fig. 9). Fe3+ and Fe2+ are active
and the surface area decrease. This fact can be due to the changes in the species in the NOx reduction [59,60]. Here the catalyst H-zeo-Fe3+
crystallinity observed by XRD and FTIR (Figs. 1 and 2, respectively). showed the highest NO conversion. Fe3+ species improve the reactivity
of the catalyst [11,29,59]. This fact can be associated with the high
concentration of NO2 observed with this catalyst (Fig. 8b), that is the
3.1.7. H2–TPR
determining step in the SCR process over zeolites [60,61]. NOx can be
The H2–TPR profile of Nat-zeo, is presented in Fig. 6. In the natural
converted according with the reactions 3 and 4, but in the presence of
zeolite, Nat-zeo, the reduction profile showed the presence of two peaks
NO2 the reaction is more efficient due to the existence of a fast SCR
at 523 °C and 623 °C, associated with the reduction of cations or oxo-
reaction [9,59,61]. NO2 can be provided in the inlet gas or produced by
cations Fe3+ to Fe2+ ions in different sites of the zeolite [18]. In the
the oxidation of NO [59].
case of Fe2+ sites detected in the Mössbauer analysis (Fig. 3), the TPR
analysis showed the peaks at 698 °C and 794 °C may associated with the 4NO + 4NH3 + O2 → 4N2 + 6H2 O (3)
reduction of Fe2+ to Fe0 [18,51,52]. The peaks at 954 °C and 1021 °C
may be associated with some residual iron species located in the more NO + 2NH3 + NO2 → 2N2 + 3H2 O (4)
stable structure orthoferrosilite (observed in the XRD analysis) [18]. Oxidized Fe 3+
species interact with NO producing NO2; however,
this reaction is controlled by the NO2 desorption [59] and in our case its
3.1.8. NH3–TPD presence in the gas-phase suggests a slow reduction rate. Regeneration
NH3–TPD profiles of Nat-zeo and H-zeo-Fe3+ fresh samples are of the Fe2+ sites takes place by the presence of the oxygen in the stream
shown in Fig. 7, and ammonia consumptions for each material are [58,59,62].
presented in Table 5. The acid–strength distribution is determined with Acidity of the materials also plays an important role in the NH3-SCR
the ammonia desorbed peaks above 100 °C. Two peaks are observed at of NOx reaction; Wang and coworkers [9] reported than NH4+ ions

Fig. 6. H2–TPR of Nat-zeo catalyst.

6
J.-F. Gelves et al. Catalysis Today xxx (xxxx) xxx–xxx

Fig. 7. NH3–TPD profiles of Nat-zeo and H-zeo-Fe3+catalysts.

located at the Brönsted acid sites are unstable and react with the NO2 remained almost constant (Fig. 10a).
species generated at the surface to form intermediate species, which The NOx-TPD results over the catalyst H-zeo-Fe3+, Fig. 11, corro-
further react with the NO to produce N2. Also, the key intermediates in borate the facts described above. The NOx-TPD profiles of the catalysts
the reaction are NH and NH2 species, which can be generated when show bands between 100 °C and 550 °C. This indicates the presence of
NH3 is adsorbed on the Lewis acid sites [63]. weakly adsorbed NOx and the formation of different nitrogenated
On the other hand, according with Fig. 8b, there is formation of N2O species on the catalyst surface [67–70]. In the NO-TPD (Fig. 11a), the
during the NO reduction over these catalysts and it is associated with peak at 103 °C could be attributed to physisorbed NOx; the peaks at 150
the presence of NO2 and its reaction with ammonia, Eq. (5) [61]. It is and 190 °C were due to the decomposition of monodentate nitrate
noticeable that over the Nat-zeo catalyst the formation of N2O is the species or iron nitrosyl species [71,72]; and the broad peaks above
highest (Fig. 8b). 250 °C were due to the decomposition of nitrate species (bridging and
bidentate) with higher thermal stability [73,74]. The presence of water
2NO2 + 2NH3 → N2 O + N2 + 3H2 O (5)
vapor significantly inhibited the NO adsorption as we expected
3+
NH3 conversion with the catalysts H-zeo-Fe and H-zeo-Fe2+ (Fig. 11a) [73]. On the other hand, NO2 and N2O-TPD profiles
(Fig. 9) was almost 100% at temperatures above 400 °C. The conversion (Fig. 11b) show the importance of water in their adsorption on the H-
observed at low temperatures over the Nat-zeo and H-zeo-Fe2+ cata- zeo-Fe3+ catalyst surface. NO2 presented two desorption broad peaks
lysts could be related with the NH3 adsorption at the beginning of the between 50 and 300 °C and 330–550 °C. The amount of NO2 desorbed is
reaction (temperatures above 150 °C) than starts to desorb at about higher when the experiment was made in the presence of water, which
200 °C. At temperatures above 400 °C, oxidation of NH3 to NO is evi- corroborates the formation of HNO2 and HNO3 on the active Fe sites
dent in iron zeolites although ammonia preferentially participates in [65]. N2O desorption is observed at temperatures above 300 °C [74],
NO reduction [64]. and the improvement in the adsorption capacity of the catalyst H-zeo-
Effect of water on the performance of the catalyst H-zeo-Fe3+ in the Fe3+ in the presence of water can be related with the formation of Fe
NO and NH3 conversion is shown in Fig. 10a. Water affected NO con- (OH)2 species in the catalyst, that can absorb N2O [75].
version in great extent and only good activity was observed between
400 and 450 °C. The effect of water at temperatures below 350 °C is 3.3. Catalyst stability
because competitive adsorption between H2O and NH3, and NH3 ad-
sorption is a key step of the NO reduction [65,66]. A durability test of the catalyst H-zeo-Fe3+ at 400 °C under wet
In the presence of water the formation of NO2 is improved and the conditions is shown in Fig. 12. After the first 4 h of reaction the NO
N2O is avoided (Fig. 10b). NO2 can react with water to form a mixture conversion decreased from 98% to 80% while the NH3 conversion in-
of nitrous acid (HNO2) and nitric acid (HNO3) which likely remain creased from 83% to 97%. With 30 h on stream the NO conversion
adsorbed on the active Fe sites required for the NO oxidation reaction decreased until 65% and the NH3 conversion until 93%. However, the
[65]. Regarding to NH3 conversion, the presence of water decreases loss of conversion in time seems to be frequent in catalysts made from
NH3 conversion until 500 °C, after this temperature the conversion naturally occurring materials and in some zeolites [38,76,77]. Two

Table 5
Ammonia consumption and total acidity of Nat-zeo and H-zeo-Fe3+ catalysts before and after reaction.

Material Peak 1 Peak 2 Peak 3 Total acidity (μmol g−1


cat )

T (°C) Acidity (μmol g−1


cat ) T (°C) Acidity (μmol g−1
cat ) T (°C) Acidity (μmol g−1
cat )

Nat-zeo 160 229.9 510 427.3 – – 657.2


H-zeo-Fe3+ 183 920.6 526 1609.9 665 46.17 2576.7
H-zeo-Fe3+ used 170 665.6 500 155.7 – – 821.3

7
J.-F. Gelves et al. Catalysis Today xxx (xxxx) xxx–xxx

Fig. 10. Effect of water with the H-zeo-Fe3+ catalyst on a) NO (fully symbols) and NH3
(open symbols) conversions and b) NO2 (fully symbols) and N2O (open symbols) con-
Fig. 8. Catalytic activity of Nat-zeo, H-zeo-Fe2+, and H-zeo-Fe3+ catalysts on a) NO centration. Reaction conditions: 400 ppm NO, 45 ppm NO2, 400 ppm NH3, 8% O2, 5%
conversion and b) NO2 (fully symbols) and N2O (open symbols) concentration. Reaction H2O and balance Ar, GHSV = 30000 h−1.
conditions: 400 ppm NO, 45 ppm NO2, 400 ppm NH3, 8% O2, and balance Ar,
GHSV = 30000 h−1. spectroscopy (Fig. 2), X-ray diffraction (Fig. 1), textural properties
(Table 4) and TPD-NH3 analysis (Table 5) were performed to establish
structural variations in the used H-zeo-Fe3+ catalyst. The XRD pattern
of the H-zeo-Fe3+ catalyst before and after 30 h reaction in presence of
water at 400 °C showed some structural variations in the catalyst, Fig. 1.
Two aspects are evident in the material, the first of them is an incre-
ment in the metaheulandite phase (2θ = 11°), which is derived from the
low thermal stable heulandite (as described in Section 3.1.1). The
second aspect is the decrease of the interplanar distance of the plane
(020) of the heulandite from 9.07 Å to 9.00 Å in the used catalyst
(Fig. 1). These two aspects suggest a decrease in the cross-sectional area
of the zeolite channels, affecting also the diffusion of the reactants to-
wards the active centers of the catalyst.
The FTIR of the used catalyst did not show new bands, Fig. 2. The
absence of the bands at 1387 cm−1 and 1242 cm−1 discards the for-
mation of solid deposit of nitrate (NO3−) and nitrite (NO2−) ions in the
used catalyst [24,57,78]. The band at 1405 cm−1 associated with the
vibration of the ammonium ions in the H-zeo-Fe3+ catalyst remains
after the durability test [78]. The presence of these ions in the fresh
catalyst is associated with the cation exchange process and the low
Fig. 9. NH3 conversion with Nat-zeo, H-zeo-Fe2+, and H-zeo-Fe3+ catalysts. Reaction calcination temperature used (360 °C, 6 h), which is not sufficient for
conditions: 400 ppm NO, 45 ppm NO2, 400 ppm NH3, 8% O2, and balance Ar, their total decomposition towards NH3. In the case of the material after
GHSV = 30000 h−1. reaction, a small increase in the intensity of the vibration of the NH4+
band is evidenced and it could be associated with the adsorption of NH3
causes could be associated to the blockage of active sites by the water on the Brönsted sites of the zeolite. Fig. 2 also shows changes in the
that affects the NO oxidation reaction [65], or the deactivation of the vibration band associated to the TO4 group (T: Si or Al) of the heu-
NO/NH3 adsorption sites and the degradation of the structure by water landite structure in the used catalyst, observed between 1020 cm−1 and
effects and reaction time. 1070 cm−1 [28]. Mozgawa et al. [28] correlated the displacement of
In order to corroborate the previous approaches, infrared this band with changes in the Si/Al ratio of the zeolite. It was

8
J.-F. Gelves et al. Catalysis Today xxx (xxxx) xxx–xxx

Fig. 11. NOx desorption profiles from the H-zeo-Fe3+ catalyst after: (A)
NO adsorption and (B) NO2/N2O adsorption at 50 °C under wet and dry
conditions.

established for the heulandite structure with high silicon content that
there is a shift of the band in the direction of the near infrared, whereas
an increase in the content of aluminum will lead to a displacement
towards the far infrared [28]. The TO4 band in the used catalyst shifted
from 1040 cm−1 to 1067 cm−1, compared with the material before
reaction, which could be associated to dealumination of the catalyst.
The dealumination process has been associated to the presence of water
vapor and the high reaction temperature (500 °C). This fact may lead to
the reduction of Brönsted acid sites, as well as the structural/textural
degradation of the zeolite, which may affect catalyst activity
[24,79,80]. However, the adsorption of water and ammonia during the
reaction can also cause this shifted since the OH or NH groups near to
the Si-O group generally lowers the stretching frequency due to the
hydrogen bonding [81]. In addition, the degradation of the tetrahedral
structure of the zeolite can be observed with the small widening of the
TO4 band after reaction (Fig. 2), indicating a decrease in crystallinity
[28]. This change in crystallinity was also observed by XRD (Fig. 1), but
Fig. 12. Durability test for H-zeo-Fe3+ catalyst. Reaction conditions: 400 ppm NO, due to the formation of metaheulandite.
45 ppm NO2, 400 ppm NH3, 8% O2, 5% H2O, and balance Ar, GHSV = 30000 h−1, 500 °C, The changes in the textural properties (Table 4) and in the acidity
30 h on stream.
(Table 5) of the catalyst H-zeo-Fe3+ after the reaction were evaluated.

9
J.-F. Gelves et al. Catalysis Today xxx (xxxx) xxx–xxx

A slight reduction in the surface area of H-zeo-Fe3+ after durability test [8] D.W. Crandell, H. Zhu, X. Yang, J. Hochmuth, M.-H. Baik, The mechanism of se-
is observed (Table 4 and Fig. 5b). This fact may be associated with the lective catalytic reduction of NOx on Cu-SSZ-13–a computational study, Dalton
Trans. 46 (2017) 369–377, http://dx.doi.org/10.1039/c6dt03894h.
structural changes of the zeolite (according to the XRD and FTIR ana- [9] P. Wang, H. Sun, X. Quan, S. Chen, Enhanced catalytic activity over MIL-100(Fe)
lysis) that make difficult the entrance of the N2 to the channels of the loaded ceria catalysts for the selective catalytic reduction of NOx with NH3 at low
zeolite. Likewise, significant changes in the other textural parameters temperature, J. Hazard. Mater. 301 (2016) 512–521, http://dx.doi.org/10.1016/j.
jhazmat.2015.09.024.
are observed in Table 4, showing an increment in the micropore area, [10] Z.-B. Xiong, J. Liu, F. Zhou, D.-Y. Liu, W. Lu, J. Jin, S.-F. Ding, Selective catalytic
average pore size and total pore volume. These changes could be as- reduction of NOx with NH3 over iron-cerium-tungsten mixed oxide catalyst pre-
sociated with the dealumination process described above in the FTIR pared by different methods, Appl. Surf. Sci. 406 (2017) 218–225, http://dx.doi.org/
10.1016/j.apsusc.2017.02.157.
analysis. Finally, the changes in the acidity of the material presented in [11] B. Feng, Z. Wang, Y. Sun, C. Zhang, S. Tang, X. Li, X. Huang, Size controlled ZSM-5
Table 5 allowed to reinforce the hypothesis that the dealumination by on the structure and performance of Fe catalyst in the selective catalytic reduction
the breaking of the SieOHeAl bonds, the decrease in the BET surface of NOx with NH3, Catal. Commun. 80 (2016) 20–23, http://dx.doi.org/10.1016/j.
catcom.2016.02.020.
area due to the metaheulandite formation, and the loss of crystallinity
[12] S. Wang, Y. Peng, Natural zeolites as effective adsorbents in water and wastewater
contribute to decrease the acid centers of the H-zeo-Fe3+catalyst from treatment, Chem. Eng. J. 156 (2010) 11–24, http://dx.doi.org/10.1016/j.cej.2009.
2576.7 to 821.3 μmol g−1 cat , which diminishes the amount of active sites 10.029.
for NH3 adsorption, affecting the efficiency of the catalyst. [13] P. Misaelides, Application of natural zeolites in environmental remediation: a short
review, Micropor. Mesopor. Mat. 144 (2011) 15–18, http://dx.doi.org/10.1016/j.
micromeso.2011.03.024.
4. Conclusions [14] T. Günter, M. Casapu, D. Doronkin, S. Mangold, V. Trouillet, T. Augenstein, J.-
D. Grunwaldt, Potential and limitations of natural chabazite for selective catalytic
reduction of NOx with NH3, Chem. Ing. Tech. 85 (2013) 632–641, http://dx.doi.
An iron Colombian natural zeolite composed mainly by heulandite org/10.1002/cite.201200182.
and celadonite phases was successfully evaluated as a potential geo- [15] J.F. Gelves, G. Sierra-Gallego, M.A. Marquez, Mineralogical characterization of
catalyst for NH3-SCR of NOx. After the treatments for obtaining the acid zeolites present on basaltic rocks from Combia geological formation, La Pintada
(Colombia), Micropor. Mesopor. Mat. 235 (2016) 9–19, http://dx.doi.org/10.1016/
material and the Fe3+ and Fe2+ species, the NO conversion increased j.micromeso.2016.07.035.
with operating temperature window of 350–500 °C. Maximum conver- [16] J.F. Gelves, G. Sierra-Gallego, L.M. Gonzalez, A.L. Villa, Colombian natural zeolite
sion increased from 21% in the natural zeolite until 76% in the presence as potential geo-catalyst for NOx and o-DCB removal, 25th North American Meeting
(NAM25), Denver-USA June 4-9, 2017.
of Fe2+ species and to 86% with the Fe3+ species. The presence of Fe3+ [17] A. Arcoya, J.A. Gonzalez, G. Llabre, X.L. Seoane, N. Travieso, Role of the coun-
favored the NO conversion, but with a 5% mol of water in the inlet tercations on the molecular sieve properties of a clinoptilolite, Microporous Mater.
stream narrowed the operating window to 400–500 °C obtaining a 7 (1996) 1–13, http://dx.doi.org/10.1016/0927-6513(96)00022-3.
[18] A. Ates, A. Reitzmann, C. Hardacre, H. Yalcin, Abatement of nitrous oxide over
conversion near to 100% at 500 °C. NOx-TPD confirmed that water
natural and iron modified natural zeolites, Appl. Catal. A: Gen. 407 (2011) 67–75,
inhibits the NO adsorption on the catalyst H-zeo-Fe3+ surface, but http://dx.doi.org/10.1016/j.apcata.2011.08.026.
improves that of NO2. At 500 °C under wet conditions after 30 h of [19] Z. Sobalik, J. Novakova, Z. Tvaržková, M. Schwarze, D. Kaucký, B. Bernauer,
reaction the NO conversion decreases to 35% over the catalyst H-zeo- Approach to analysis of redox active sites of working metal-ion zeolites. In-situ
analysis of Fe-zeolite catalysts in NOx reduction by NH3, and NO assisted N2O
Fe3+, which according with the FTIR, XRD and BET analysis can be decomposition, Stud. Surf. Sci. Catal. 158 (2005) 1231–1238, http://dx.doi.org/10.
associated with a dealumination process, the contraction of the zeolite 1016/S0167-2991(05)80469-0.
channels due to the metaheulandite formation, and the acidity and [20] L.L. Ren, T. Zhang, Reduction of NO with methane over Fe/ZSM-5 catalysts, Chin.
Chem. Lett. 21 (2010) 674–677, http://dx.doi.org/10.1016/j.cclet.2010.02.001.
crystallinity losses. [21] H.A.W. Kaminsky, T.H. Etsell, D.G. Ivey, O. Omotoso, Characterization of heavy
minerals in the Athabasca oil sands, Miner. Eng. 21 (2008) 264–271, http://dx.doi.
Acknowledgments org/10.1016/j.mineng.2007.09.011.
[22] The Home of LST Heavy Liquid. Central Chemical Consulting [On line]: http://
www.heavyliquids.com/techdata.html. Consulted: July 29, 2017.
JFG, LD and MAM acknowledge financial support to Universidad [23] N. Ghasemian, C. Falamaki, M. Kalbasi, M. Khosravi, Enhancement of the catalytic
Nacional de Colombia sede Medellín; ACA, LMG and ALV thank to performance of H-clinoptilolite in propane SCR-NOx process through controlled
dealumination, Chem. Eng. J. 252 (2014) 112–119, http://dx.doi.org/10.1016/j.
Universidad de Antioquia (UdeA) and Colciencias for financial support
cej.2014.04.039.
through the project 1115-569-33782, and UdeA for project PURDUE [24] R. Moreno-Tost, J. Santamaria-González, E. Rodriguez-Castellón, A. Jiménez-López,
14-2-05. M.A. Autié, E. González, M. Carreras Glacial, C. de las Pozas, Selective catalytic
reduction of nitric oxide by ammonia over Cu-exchanged Cuban natural zeolites,
Appl. Catal. B: Environ. 50 (2004) 279–288, http://dx.doi.org/10.1016/j.apcatb.
References 2004.01.019.
[25] M. Ostrooumov, P. Cappelletti, R. de’Gennaro, Mineralogical study of zeolite from
[1] L. Ntziachristos, G. Papadimitriou, N. Ligterink, S. Hausberger, Implications of new mexican deposits (Cuitzeo area, michoacan, Mexico), Appl. Clay Sci. 55 (2012)
diesel emissions control failures to emission factors and road transport NOx evo- 27–35, http://dx.doi.org/10.1016/j.clay.2011.09.011.
lution, Atmos. Environ. 141 (2016) 542–551, http://dx.doi.org/10.1016/j. [26] A. Alberti, The structure type of heulandite B (Heat-Collapsed phase), TMPM
atmosenv.2016.07.036. Tschermaks Min. Petr. Mitt. 19 (1973) 173–184.
[2] A. Osman Emiroglu, Investigation of NOx reduction activity of Rh/ZnO nanowires [27] H. Van der Marel, H. Beautelspacher, Atlas of Infrared Spectroscopy of Clay
catalyst, Atmos. Pollut. Res. 8 (2017) 149–153, http://dx.doi.org/10.1016/j.apr. Minerals and Their Admixtures, Elsevier Sci Pub., B.V., Amsterdam, 1978.
2016.08.006. [28] W. Mozgawa, M. Sitarz, M. Rokita, Spectroscopic studies of different aluminosili-
[3] D.K. Wasiuk, Md A.H. Khan, D.E. Shallcross, M.H. Lowenberg, A commercial air- catestructures, J. Mol. Struct. 511-512 (1999) 251–257, http://dx.doi.org/10.
craft fuel burn and emissions inventory for 2005–2011, Atmosphere 7 (2016) 1016/S0022-2860(99)00165-9.
78–92, http://dx.doi.org/10.3390/atmos7060078. [29] M. Iwasaki, K. Yamazaki, K. Banno, H. Shinjoh, Characterization of Fe/ZSM-5
[4] C. Roy, S. Fadnavis, R. Müller, D.C. Ayantika, F. Ploeger, A. Rap, Influence of en- DeNOx catalysts prepared by different methods: relationships between active Fe
hanced Asian NOx emissions on ozone in the upper troposphere and lower strato- sites and NH3-SCR performance, J. Catal. 260 (2008) 205–216, http://dx.doi.org/
sphere in chemistry–climate model simulations, Atmos. Chem. Phys. 17 (2017) 10.1016/j.jcat.2008.10.009.
1297–1311, http://dx.doi.org/10.5194/acp-17-1297-2017. [30] V.I. Goldanskii, R.H. Herber, Chemical Application of the Mössbauer Spectroscopy,
[5] J. Thangaraja, C. Kannan, Effect of exhaust gas recirculation on advanced diesel Academic Press, New York, 1968.
combustion and alternate fuels − a review, Appl. Energy 180 (2016) 169–184, [31] P. Gütlich, R. Link, A. Trautwein, Mossbauer Spectroscopy and Transition Metal
http://dx.doi.org/10.1016/j.apenergy.2016.07.096. Chemistry, Springer Verlag, Berlin, 1978.
[6] G. Karavalakis, K.C. Johnson, M. Hajbabaei, T.D. Durbin, Application of low-level [32] J. Fontcuberta, La espectroscopia Mössbauer, principios y aplicaciones, Rev.
biodiesel blends on heavy-duty (diesel) engines: feedstock implications on NOx and Colomb. Quim. 14 (1985) 99–116.
particulate emissions, Fuel 181 (2016) 259–268, http://dx.doi.org/10.1016/j.fuel. [33] A. Ignaszak, S. Komornicki, P. Pasierb, Mössbauer effect study of the Fe-substituted
2016.05.001. NASICON, Ceram. Int. 35 (2009) 2531–2535, http://dx.doi.org/10.1016/j.
[7] M. Nahavandi, Selective catalytic reduction (SCR) of NO by ammonia over V2O5/ ceramint.2009.02.003.
TiO2 catalyst in a catalytic filter medium and honeycomb reactor: a kinetic mod- [34] R. Roque-Malherbe, C. Diaz-Aguila, E. Reguera-Ruiz, J. Fundora-Lliteras, L. Lopez-
eling study, Braz. J. Chem. Eng. 32 (2015) 875–893, http://dx.doi.org/10.1590/ Colado, M. Hernaindez-Velez, The state of iron in natural zeolites: a Mössbauer
0104-6632.20150324s00003584. study, Zeolites 10 (1990) 685–689, http://dx.doi.org/10.1016/0144-2449(90)
90080-B.

10
J.-F. Gelves et al. Catalysis Today xxx (xxxx) xxx–xxx

[35] J. Steven, A. Khasanov, J. Miller, H. Pollak, Z. Li, Mössbauer Mineral Handbook, spectroscopy and kinetic modeling of NOx storage and NO oxidation on Fe-BEA SCR
Mössbauer Effect Data Center, Asheville-North Carolina, 2005. catalysts, Appl. Catal. B: Environ. 148-149 (2014) 446–465, http://dx.doi.org/10.
[36] L.G. Daynyak, V.A. Drits, Interpretation of Mossbauer spectra of nontronite, 1016/j.apcatb.2013.11.018.
Celadonite and glauconite, Clays Clay Miner. 35 (1987) 363–372. [61] C.P. Cho, Y.D. Pyo, J.Y. Jang, G.C. Kim, Y.J. Shin, NOx reduction and N2O emissions
[37] V.A. Drits, L.G. Dainyak, F. Muller, G. Besson, A. Manceau, Isomorphous cation in a diesel engine exhaust using Fe-zeolite and vanadium based SCR catalysts, Appl.
distribution in Celadonites, glauconites and Fe-illites determined by infrared Therm. Eng. 110 (2017) 18–24, http://dx.doi.org/10.1016/j.applthermaleng.2016.
Mössbauer and EXAFS spectroscopies, Clay Miner. 32 (1997) 153–179. 08.118.
[38] G. Rodríguez-Fuentes, I. Rodríguez-Iznaga, Caracterización del mineral zeolítico [62] G. Delahay, D. Valade, A. Guzmán-Vargas, B. Coq, Selective catalytic reduction of
para el desarrollo de materiales nanoestructurados, Revista Cubana de Física 26 nitric oxide with ammonia on Fe-ZSM-5 catalysts prepared by different methods,
(2009) 55–60. Appl. Catal. B: Environ. 55 (2005) 149–155, http://dx.doi.org/10.1016/j.apcatb.
[39] S. Yang, M. Lach-hab, E. Blaisten-Barojas, X. Li, V.L. Karen, Machine learning study 2004.07.009.
of the heulandite family of zeolites, Micropor. Mesopor. Mat. 130 (2010) 309–313, [63] D. Zhang, R.T. Yang, NH3-SCR of NO over one-pot Cu-SAPO-34 catalyst: perfor-
http://dx.doi.org/10.1016/j.micromeso.2009.11.027. mance enhancement by doping Fe and MnCe and insight into N2O formation, Appl.
[40] K.S.W. Sing, D.H. Everett, R.A.W. Haul, L. Moscou, R.A. Pierotti, J. Rouquerol, Catal. A: Gen. 543 (2017) 247–256, http://dx.doi.org/10.1016/j.apcata.2017.06.
T. Siemieniewska, Reporting physisorption data for gas/solid systems with special 021.
reference to the determination of surface area and porosity (Recommendations [64] P. Boron, L. Chmielarz, J. Gurgul, K. Latka, B. Gil, J.-M. Krafft, S. Dzwigaj, The
1984), Pure Appl. Chem. 57 (1985) 603–619. influence of the preparation procedures on the catalytic activity of Fe-BEA zeolites
[41] D. Bal Akkoca, M. Yιlgιn, M. Ural, H. Akçin, A. Mergen, Hydrothermal and thermal in SCR of NO with ammonia and N2O decomposition, Catal. Today 235 (2014)
treatment of natural clinoptilolite zeolite from bigadiç, Turkey: an experimental 210–225, http://dx.doi.org/10.1016/j.cattod.2014.03.018.
study, Geochem. Int. 51 (2013) 495–504, http://dx.doi.org/10.1134/ [65] P.S. Metkar, N. Salazar, R. Muncrief, V. Balakotaiah, M.P. Harold, Selective cata-
S0016702913040022. lytic reduction of NO with NH3 on iron zeolite monolithic catalysts: steady-state and
[42] K.A. Savko, S.M. Piliugin, N.S. Bazikov, Experimental data for high-temperature transient kinetics, Appl. Catal. B: Environ. 104 (2011) 110–126, http://dx.doi.org/
decomposition of natural Celadonite from banded iron formation, Chin. J. 10.1016/j.apcatb.2011.02.022.
Geochem. 34 (2015) 507–514, http://dx.doi.org/10.1007/s11631-015-0066-2. [66] R.Q. Long, R.T. Yang, Catalytic performance of Fe–ZSM-5 catalysts for selective
[43] F. Esenli, I. Kumbasar, X-Ray diffraction intensity ratios I (111)//(311) of natural catalytic reduction of nitric oxide by ammonia, J. Catal. 188 (1999) 332–339,
heulandites and clinoptilolites, Clays Clay Miner. 46 (1998) 679–686. http://dx.doi.org/10.1006/jcat.1999.2674.
[44] A.O. Larsen, F.S. Nordrum, N. Döbelin, T. Armbruster, O.V. Petersen, M. Erambert, [67] F. Liu, H. He, Y. Ding, C. Zhang, Effect of manganese substitution on the structure
heulandite-Ba, a new zeolite species from Norway, Eur. J. Mineral. 17 (2005) and activity of iron titanate catalyst for the selective catalytic reduction of NO with
143–153, http://dx.doi.org/10.1127/0935-1221/2005/0017-0143. NH3, Appl. Catal. B: Environ. 93 (2009) 194–204, http://dx.doi.org/10.1016/j.
[45] A. Mehmani, M. Prodanović, The application of sorption hysteresis in nano-petro- apcatb.2009.09.029.
physics using multiscale multiphysics network models, Int. J. Coal Geol. 128-129 [68] F. Liu, H. He, Structure-Activity relationship of iron titanate catalysts in the se-
(2014) 96–108, http://dx.doi.org/10.1016/j.coal.2014.03.008. lective catalytic reduction of NOx with NH3, J. Phys. Chem. C 114 (2010)
[46] N. Mansouri, N. Rikhtegar, H.A. Panahi, F. Atabi, B.K. Shahraki, Porosity, 16929–16936, http://dx.doi.org/10.1021/jp912163k.
Characterization and structural properties of natural zeolite-clinoptilolite-as a sor- [69] R. Gao, D. Zhang, X. Liu, L. Shi, P. Maitarad, H. Li, J. Zhang, W. Cao, Enhanced
bent, Environ. Prot. Eng. 39 (2013) 139–152, http://dx.doi.org/10.5277/ catalytic performance of V2O5–WO3/Fe2O3/TiO2 microspheres for selective cata-
EPE130111. lytic reduction of NO by NH3, Catal. Sci. Technol. 3 (2013) 191–199, http://dx.doi.
[47] P. Kowalczyk, M. Sprynskyy, A.P. Terzyk, M. Lebedynets, J. Namieśnik, org/10.1039/c2cy20332d.
B. Buszewski, Porous structure of natural and modified clinoptilolites, J. Colloid [70] A. Serrano-Lotina, A.C. Bueno, C. Goberna-Selma, P. Ávila, M.A. Bañares, NO ad-
Interf. Sci. 297 (2006) 77–85, http://dx.doi.org/10.1016/j.jcis.2005.10.045. sorption and influence of the control of temperature over catalytic test results for
[48] X. Du, E. Wu, Porosity of microporous zeolites A, X and ZSM-5 studied by small NO oxidation, Catal. Today 297 (2017) 2–9, http://dx.doi.org/10.1016/j.cattod.
angle X-ray scattering and nitrogen adsorption, J. Phys. Chem. Solids 68 (2007) 2017.06.031.
1692–1699, http://dx.doi.org/10.1016/j.jpcs.2007.04.013. [71] V. Blasin-Aubé, O. Marie, J. Saussey, A. Plesniar, M. Daturi, N. Nguyen, C. Hamon,
[49] G.E. Christidis, D. Moraetis, E. Keheyan, L. Akhalbedashvili, N. Kekelidze, M. Mihaylov, E. Ivanova, K. Hadjiivanov, Iron nitrosyl species in Fe-FER: a com-
R. Gevorkyan, H. Yeritsyan, H. Sargsyan, Chemical and thermal modification of plementary Mössbauer and FTIR spectroscopy study, J. Phys. Chem. C 113 (2009)
natural HEU-type zeolitic materials from Armenia, Georgia and Greece, Appl. Clay 8387–8393, http://dx.doi.org/10.1021/jp900699m.
Sci. 24 (2003) 79–91, http://dx.doi.org/10.1016/S0169-1317(03)00150-9. [72] X. Wang, W. Wu, Z. Chen, R. Wang, Bauxite-supported transition metal oxides:
[50] P.P. Diale, E. Muzenda, J. Zimba, A study of south african natural zeolites properties promising low-temperature and SO2-tolerant catalysts for selective catalytic re-
and applications, Proceedings of the World Congress on Engineering and Computer duction of NOx, Sci. Rep. 5 (2015) 9766–9772, http://dx.doi.org/10.1038/
Science 2011 (WCECS 2011), October 19–21, San Francisco, USA, 2011. srep09766.
[51] M. Shimokawabe, T. Chaki, S. Ozawa, M. Arai, Activity and physicochemical [73] L. Chen, Z. Si, X. Wu, D. Weng, Z. Wu, Effect of water vapor on NH3–NO/NO2 SCR
properties of Fe-zeolite catalysts for SCR of N2O with CH4, React. Kinet. Catal. Lett. performance of fresh and aged MnOx–NbOx–CeO2 catalysts, J. Environ. Sci. 31
86 (2005) 363–370, http://dx.doi.org/10.1007/s11144-005-0333-z. (2015) 240–247, http://dx.doi.org/10.1016/j.jes.2014.07.037.
[52] S.S.R. Putluru, A.D. Jensen, A. Riisager, R. Fehrmann, Alkali resistant Fe-zeolite [74] Z. Say, M. Dogac, E.I. Vovk, Y.E. Kalay, C.H. Kim, W. Li, E. Ozensoy, Palladium
catalysts for SCR of NO with NH3 in flue gases, Top. Catal. 54 (2011) 1286, http:// doped perovskite-based NO oxidation catalysts: the role of Pd and B-sites for NOx
dx.doi.org/10.1007/s11244-011-9750-6. adsorption behavior via in-situ spectroscopy, Appl. Catal. B 154-155 (2014) 51–61,
[53] L.J. Lobree, I.-C. Hwang, J.A. Reimer, A.T. Bell, Investigations of the state of Fe in http://dx.doi.org/10.1016/j.apcatb.2014.01.038.
H–ZSM-5, J. Catal. 186 (1999) 242–253, http://dx.doi.org/10.1006/jcat.1999. [75] A. Heyden, B. Peters, A.T. Bell, F.J. Keil, Comprehensive DFT study of nitrous oxide
2548. decomposition over Fe-ZSM-5, J. Phys. Chem. B 109 (2005) 1857–1873, http://dx.
[54] E.G. Derouane, J.C. Vedrine, R.R. Pinto, P.M. Borges, L. Costa, Lemos M.A.N.D.A, doi.org/10.1021/jp040549a.
F.R. Ribeiro, The acidity of zeolites: concepts, measurements and relation to cata- [76] L. Chmielarz, M. Rutkowska, M. Jablonska, A. Wegrzyn, A. Kowalczyk, P. Boron,
lysis: a review on experimental and theoretical methods for the study of zeolite Z. Piwowarska, A. Matusiewicz, Acid-treated vermiculites as effective catalysts of
acidity, Catal. Rev. 55 (2013) 454–515, http://dx.doi.org/10.1080/01614940. high-temperature N2O decomposition, Appl. Clay Sci. 101 (2014) 237–245, http://
2013.822266. dx.doi.org/10.1016/j.clay.2014.08.006.
[55] M. Boronat, A. Corma, Factors controlling the acidity of zeolites, Catal. Lett. 145 [77] I.M. Saaid, A.R. Mohamed, S. Bhatia, Activity and characterization of bimetallic
(2015) 162–172, http://dx.doi.org/10.1007/s10562-014-1438-7. ZSM-5 for the selective catalytic reduction of NOx, J. Mol. Catal. A: Chem. 189
[56] A. Ates, C. Hardacre, The effect of various treatment conditions on natural zeolites: (2004) 241–250, http://dx.doi.org/10.1016/S1381-1169(02)00330-8.
ion exchange, acidic, thermal and steam treatments, J. Colloid Interf. Sci. 372 [78] A.E.-A.A. Said, M.M.A. El-Wahab, M.N. Goda, Selective synthesis of acetone from
(2012) 130–140, http://dx.doi.org/10.1016/j.jcis.2012.01.017. isopropyl alcohol over active and stable CuONiO nanocomposites at relatively low-
[57] R.G. Herman, Q. Sun, C. Shi, K. Klier, C.B. Wang, H. Hu, M.M. Bhasin, Development temperature, Egypt, J. Basic Appl. Sci. 3 (2016) 357–365, http://dx.doi.org/10.
of active oxide catalysts for the direct oxidation of methane to formaldehyde, Catal. 1016/j.ejbas.2016.08.004.
Today 37 (1997) 1–14, http://dx.doi.org/10.1016/S0920-5861(96)00256-8. [79] P. Budi, E. Curry-Hyde, R.F. Howe, Steam deactivation of transition metal MFI
[58] H. Pan, Y. Guo, H.T. Bi, NOx adsorption and reduction with C3H6 over Fe/zeolite zeolite catalysts for NOx reduction, Stud. Surf. Sci. Catal. 105 (1997) 1549–1556,
catalysts: effect of catalyst support, Chem. Eng. J. 280 (2015) 66–73, http://dx.doi. http://dx.doi.org/10.1016/S0167-2991(97)80798-7.
org/10.1016/j.cej.2015.05.093. [80] P. Wang, H. Zhao, H. Sun, H. Yu, X. Quan, Porous metal–organic framework MIL-
[59] S. Brandenberger, O. Kröcher, A. Tissler, R. Althoff, The state of the art in selective 100 (Fe) as an efficient catalyst for the selective catalytic reduction of NOx with
catalytic reduction of NOx by ammonia using metal-exchanged zeolite catalysts, NH3, RSC Adv. 4 (2014) 48912–48919, http://dx.doi.org/10.1039/C4RA07028C.
Catal. Rev. Sci. Eng. 50 (2008) 492–531, http://dx.doi.org/10.1080/ [81] K. Byrappa, B.V. Suresh Kumar, Characterization of zeolites by infrared spectro-
0161494080248012. scopy, Asian J. Chem. 19 (2007) 4933–4935.
[60] S.A. Skarlis, D. Berthout, A. Nicolle, C. Dujardin, P. Granger, Combined IR

11

Das könnte Ihnen auch gefallen