Sei sulla pagina 1di 7

Catalysis Today 216 (2013) 135141

Contents lists available at SciVerse ScienceDirect


Catalysis Today
j our nal homepage: www. el sevi er . com/ l ocat e/ cat t od
Mass transport and kinetics in structured steel foam reactor with
Cu-ZSM-5 catalyst for SCR of NO
x
with ammonia
Joanna Ocho nska-Kryca
a,b
, Marzena Iwaniszyn
b
, Marcin Pi atek
b
,
Przemysaw J. Jodowski
a
, Jacob Thomas
a
, Andrzej Koodziej
b,c
, Joanna ojewska
a,
a
Faculty of Chemistry, Jagiellonian University, Ingardena 3, Krakw, 30-060, Poland
b
Institute of Chemical Engineering, Polish Academy of Sciences, Batycka 5, Gliwice, 44-100, Poland
c
Faculty of Civil Engineering, Opole University of Technology, Katowicka 48, Opole, 45-061, Poland
a r t i c l e i n f o
Article history:
Received 16 March 2013
Received in revised form25 May 2013
Accepted 25 May 2013
Available online 20 July 2013
Keywords:
Metallic foam
Cu-ZSM-5
NH
3
-SCR
Heat transfer
Mass transfer
a b s t r a c t
Continued research on improving heat and mass transfer properties of structured catalysts is essential
for the development of an effective catalytic system. In this paper, the transfer parameters (mass and
heat) and pressure drop across reactors prepared with metal foams of different materials (FeCrAl and
Al) and PPI (pores per inch) are presented. The mass transfer parameter was determined by correlation
using the ChiltonColburn analogy, and the heat transfer and pressure drop parameters were determined
experimentally.
Additionally the kinetic properties of a short channel reactor structure based on metallic foams and
template-less synthesized ZSM-5 were determined for the selective catalytic reduction (SCR) of NO with
ammonia and compared to commonly used commercial solutions i.e. packed bed and honeycomb
monolith reactor structures.
2013 Elsevier B.V. All rights reserved.
1. Introduction
Since their discovery, metallic foams materials with cellular
structure of high porosities 8590% have been gaining research
interest as a new prospective material with applications in many
elds. A broad review on metal foams is presented by Banhart [1]
where foam applications are divided into three sections: struc-
tural industrial application (automotive, building industry etc.),
functional applications (heat exchangers, ltration etc.) and dec-
oration/arts. Metallic foams seemto be very promising as supports
for heterogeneous catalyst. It has already been suggested that they
could be used in oxidation of organic substances like alcohols [2],
hydrocarbons [3,4], ammonia [5], methane [6] or in CO
2
reforming
process [7]. Thanks to their low pressure drop, low density, high
geometric surface areas, high convective heat transfer and radial
mixing, pre-shapedmetallic foams may replace packedbeds of par-
ticles or even monoliths in environmental catalysis. Application of
metal foam based reactors, operating at high space velocity with
lowpressure drop will shorten the needed reactor length, simplify
loading procedure and lower the overall installation costs.
The structural morphology of foamand its specic surface area
are greatly responsible for their performance. With increasing

Corresponding author. Tel.: +48 126632087.


E-mail address: lojewska@chemia.uj.edu.pl (J. ojewska).
precision in foam production techniques, the possibility to design
optimized characteristics of a porous isotropic structure, taking
intoaccount the dual properties structural andfunctional arises.
New materials, like metallic foams, shaped into short channel
structures with enhanced heat transfer properties, might shorten
the reactor length by up to 50 times [8,9], but the reactors still
require an appropriate active catalytic material. In this paper, a
catalyst system for the selective catalytic reduction (SCR) of NO
x
based on copper-exchanged zeolite was explored. In the literature
Cu-exchanged zeolites have been demonstrated to be remarkably
active bothinthe reductionof nitrogenoxides withNH
3
(NH
3
-SCR)
[10] and in the direct decomposition of NO [11,12]. Considering
the preparation of a structured catalyst based on metallic foam
with washcoated or in situ synthesized catalyst, FeCrAl foams are
advantageous as upon calcination, a layer of -alumina forms on
the metal foamsurface thus providing a good bonding area for the
coating [13].
Until recently, research has focused on metal foam properties
such as inertia coefcient, thermal conductivity and permeability
[1416], but to reliably design a structured catalyst reactor and
simulate its performance, data on the heat [1720] and mass trans-
fer coefcients [21,22] and pressure drop [23,24] are required. The
results reported in the literature for the above mentioned param-
eters are highly inconsistent. This is not surprising as the foams
studied by the various authors differed substantially depending
on the suppliers, foam materials, various PPI values, as well as
0920-5861/$ see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cattod.2013.05.018
136 J. Ocho nska-Kryca et al. / Catalysis Today 216 (2013) 135141
Nomenclature
A
0
pre-exponential coefcient in Arrhenius equation
(ms
1
)
C
A
reagent concentration (mol m
3
)
D
A,eff
effective diffusivity in porous media (m
2
s
1
)
D
pore
pore diameter in foam(m)
D
p
particle (sphere) diameter (m)
D
h
hydraulic diameter (m)
E
a
activation energy (kJ mol
1
)
f Fanning friction factor
f
p
catalyst particle external surface area (m
2
)
j Colburn factor
k kinetic rate constants of reaction (ms
1
)
k
c
mass transfer coefcient (ms
1
)
L length of reactor (m)
l length of diffusion path in catalytic phase (m)
Nu Nusselt number, Nu=D
foam

1
P pressure drop across the reactor
Pr Prandt number, Pr =c
p

1
r kinetic rate (mol s
1
m
2
)
R gas constant (J K
1
mol
1
)
Re Reynolds number, Re =w
0
D
foam

1
Sc Schmidt number, Sc =
1
D
1
A
Sh Sherwood number, Sh = k
c
D
foam
D
1
A
St Stanton number, St
H
= w
1
0

1
c
1
p
=
Nu Re
1
Pr
1
, St
M
= k
c
w
1
0
= Sh Re
1
Sc
1
S
v
external surface area (m
1
)
T temperature (K)
w
0
supercial velocity (ms
1
)
V
air
volumetric owrate (nm
3
h
1
)
V
p
particle volume (m
3
)
X conversion
Greek letters
void volume

e
catalyst effectiveness factor
density (kgm
3
)
catalyst specic surface area (m
2
kg
1
)
Thiele modulus
Indexes
H heat
M mass
S at surface
L at reactor outlet
their experimental approaches and differently dened Reynolds
numbers. There is evidently no agreement on the characteristic
dimension in the criterial numbers, and moreover, no agreement
can be reached on the ow mechanism, and consequently heat,
mass and momentumtransfer description.
Our work focuses on the investigation of the heat transfer
coefcients at the gassolid interface of metallic foams. This is
accomplished by thermoresistive heating the metal foam with an
electrical current, andthencareful analysis of experimental param-
eters. The purpose of this workis tobuilda better correlationfor the
gassolid heat transfer coefcients. The Chilton and Colburn heat
and mass transfer analogy was used to determine the mass transfer
representation. By assuming a simple cubic structure of the metal
foam, where the foampore diameter was consideredas a character-
istic dimension, the pressure drop across a foamreactor is explored
corresponding to different foampore densities and foammaterials
Fig. 1. Image of FeCrAl foam, 20PPI, scale bar 0.5mm.
used. The catalytic studies of NH
3
-SCR on Cu-ZSM-5 enabled us to
determine the kinetic parameters (A
0
, E
a
) which were further used
to select the most effective structural support.
It should be noted that reliable identication of the ow
mechanism, thus proper modelling of the transport and friction
phenomena, requires a large experimental database for compar-
ison. The experiments have to be performed using the same
methodology and denitions of the criterial numbers. More-
over, the morphology of the foams studied has to be examined
using carefully selected techniques and an appropriate geometri-
cal model due to signicant differences between approaches. This
work is the subject of our ongoing research, and the preliminary
results are presented below.
2. Experimental
2.1. Metallic foams
Metal foams of two different composition, FeCrAl (Selee Cor-
poration, 71% Fe, 22% Cr, 6% Al and traces of Mn and Y) and Al
(ERGMaterials andAerospace Corporation) were usedinthis study.
Those foams differed in cell size 20 and 40PPI for FeCrAl and
10, 20 and 40PPI for Al. The geometric characterization of the
foams focused on determining the pore size D
pore
by microscopic
investigation. Images of the different foam samples were taken
with a DSLR camera (Nikon D300s) with a SIGMA 70mm1:2.8 DG
MACRO lens. Clearly focused pores (Fig. 1) were analyzed using
the Leica Application Software (Version 2.) from Leica Microsys-
tems Switzerland Ltd. Over 100 pores were statistically examined
in order to ensure a representative value. A Gaussian curve was
used to t the pore size distribution plot.
The open volume fraction was calculated fromEq. (1)
= 1

foam

s
(1)
The foam density,
foam
, was estimated by dividing the weight
of the foamslice by total measuredvolume while the struts density,

s
was determined using a He pycnometric method. At this stage
of our studies we have used the external surface area S
v
supplied
by the manufacturer.
Geometrical parameters (D
pore
, and S
v
) for the tested metal
foamsamples are given in Table 1.
2.2. Catalyst
The hydrothermal synthesis of ZSM-5 zeolite (Si/Al =15) with-
out an organic template was described in our previous work [25],
where the catalyst was synthesized on a metal plate. XRD analysis
J. Ocho nska-Kryca et al. / Catalysis Today 216 (2013) 135141 137
Table 1
Geometric characteristics of the samples and correlations for mass transfer and Fanning friction factor used in modelling.
Material type Cell density [%] Dpore [m] Sv [m
2
m
3
] Mass transfer correlations Fanning friction factor
FeCrAl foam 20PPI 0.94 1.110
3
1400 h=0.55Re
0.5864
Sc
1/3
f =11.12/Re +0.16
FeCrAl foam 30PPI 0.96 0.910
3
2100 h=0.41Re
0.6147
Sc
1/3
f =6.66/Re +0.11
Al foam 10PPI 0.92 1.210
3
1800 h=1.13Re
0.467
Sc
1/3
f =10.5/Re +0.13
Al foam 20PPI 0.92 0.810
3
2153 h=0.68Re
0.355
Sc
1/3
f =3.66/Re +0.08
Al foam 40PPI 0.94 0.710
3
4300 h=0.37Re
0.398
Sc
1/3
f =3.08/Re +0.08
Packed bed 0.4 0.310
3
12,000
Shp = 2 +1.1Sc
1/3
Re
0.6
p
[38]
Ergun equation [39]
Packed bed 0.4 210
3
1570
Cordierite monolith 400cpsi 0.85 1.1610
3 a
2709 Sh = 3.608(1 +0.095/L
M
)
0.45
[40] f Re = 14.23(1 +0.045/L

)
0.5
[40]
L
*M
=L/D
h
Re Sc; L* =L/D
h
Re.
a
Hydraulic diameter.
of powder sample conrmed the MFI crystallographic form while
SEM images conrmed a crystal size of 8m (Fig. 2). The synthe-
sized MFI sample was ion-exchanged with copper ions, according
to the procedure found in [26]. For the kinetic tests the commer-
cial ZSM-5 (Si/Al =15, Zeolyst) with crystal size of 1m was also
examined.
2.3. Catalytic tests
The catalytic activity of Cu-ZSM-5 sample was examined in
selective catalytic reduction of NO with ammonia as the test
reaction. The tests were performed in a microreactor (Hiden Ana-
lytical CATLAB-PCS) with quadrupole mass spectrometer (QMS)
positioned downstreamof the reactor to analyze the efuent gases
including NH
3
, NO, NO
2
, N
2
O, N
2
and H
2
O. 25mg sample of the
unsupported catalyst as a powder (crystals pressed and sieved
into fractions of average particle diameter =300m) was placed
insidea quartz tubereactor mountedina tubefurnacecoupledwith
a temperature controller. Prior to the kinetic experiments, the cat-
alyst was conditioned in 5% O
2
in He at 500

C. Once the maximum


temperature was reached the sample was cooled to the required
experimental temperature. In all experiments He was used as an
inert (balance) gas and the total owrate of 75ml min
1
was kept
constant. Catalytic reactions were carried out in the temperature
range 40550

C under atmospheric pressure with 2000ppm NO,


2000ppm NH
3
(stoichiometric ratio), 1800ppm H
2
O and 14vol.%
O
2
in He.
In order to eliminate internal mass-transport (Fig. 3) within the
pores of 5.45.6

A diameter in the ZSM-5 the performance of two
Cu-exchangedzeolites of different crystal sizewas compared(8m
synthesizedZSM-5 and1m commercial ZSM-5). Bothzeolites,
Fig. 2. SEMimages of synthesized ZSM-5.
Fig. 3. NO conversion over Cu-ZSM-5 powder catalyst with particle size 1mand
8 m.
Experimental conditions: 2000ppmNO, 2000ppmNH
3
, 1800ppmH
2
Oand 14vol.%
O
2
in He, total owrate 75ml min
1
, 25mg of a catalyst.
in catalytic tests, were sieved into fraction of an average particle
diameter =300m:
r = kC
A
(2)
where subscript A denotes NO included in the reaction mixture.
Calculation of the reaction rate constant was based on Arrhenius
equation (3) where the apparent activation energy was obtained
Fig. 4. Reaction rate in Arrhenius coordinates (k, ms
1
) for powder Cu-ZSM-5 cat-
alyst.
138 J. Ocho nska-Kryca et al. / Catalysis Today 216 (2013) 135141
Fig. 5. Schematics of the apparatus for heat transport and pressure drop measure-
ments.
fromArrhenius plot (Fig. 4) at a lowconversion (<10% in all exper-
iments) in the temperature range 100200

C.
k = A
0
exp

E
a
RT

(3)
2.4. Heat transfer and pressure drop measurements
The heat transfer coefcients were evaluated between elec-
trically heated metallic foam and owing gas. The experimental
results were further transformed by the use of the heat and mass
transport analogy into the mass transfer representation. The details
concerning this operation, with rigorous description of the analo-
gies applied, can be found in [30].
Both heat transfer and pressure drop experiments were carried
out in the same test rig (Fig. 5) used by [9]. Foam samples of size
453011mmwere cut fromthe original metal foam. The foam
sample was mounted into the customproduced cartridge thereby
forcing the gas ow to pass directly through the foam samples.
Then the cartridge was inserted into a well isolated tube reac-
tor perpendicularly to the direction of the gas stream (air). The
air volumetric ow rate was measured with a calibrated rotame-
ter and was varied over wide range (V
air
=0.3100nm
3
h
1
). The
physical properties of the air were calculated as a function of the
average temperature of the air in the reactor. The heat transfer
experiments were run on each foam sample within a Reynolds
number range of Re =4800 (Re refers to the cell diameter D
pore
and supercial empty tube velocity w
0
). The metal foams were
heated with electric current (up to 150A) owing directly through
them, to attain an appropriate temperature gradient between the
foamand owing gas (515K). In this way the boundary condition
of constant heat ux was secured. The temperatures at the foam
surface were measured by eight type K thermocouples (4 at the
inlet and 4 at the outlet side of the foam sample) that were xed
by composite glue characterized by good heat conduction and no
electric conduction. The temperatures of the owing air were mea-
suredusing six thermocouples (three placedinthe inlet sectionand
three in the outlet section of the test reactor). All thermocouples
were connectedtoa computer data acquisitionsystemthat allowed
continuous overviewof the experimental results.
Heat transfer coefcients were calculated fromthe mean loga-
rithmic temperature difference between the foamsurface and the
owing air streamas well as fromthe measured electrical heating
power (product of the current and the voltage drop at the foam).
The heat transfer data obtainedfor all testedfoams were ttedwith
a correlation in the following form:
Nu = C Re
A
Pr
1/3
(4)
Fig. 6. Mass transfer representation vs. Re for FeCrAl foams (20 and 30PPI) and Al
foams (10, 20 and 40PPI). Lines correlations for Sh according to Table 1.
This representation was used to turn into the mass transfer rep-
resentation applying the ChiltonColburn analogy [8,31]:
j
H
= St
H
Pr
2/3
=
Nu
Re Pr
1/3
=
f
2
= j
M
= St
M
Sc
2/3
=
Sh
Re Sc
1/3
(5)
According to the above, the mass transfer equation has the same
constants as in Eq. (4) but Nu is replaced by Sh and Pr by Sc [8,31].
All derived mass transfer equations are collected in Table 1 and the
relations between experimental and correlated Sh number versus
Re are presented in Fig. 6. The experimental data obtained for the
20PPI FeCrAl foam were compared to a correlations derived by
Groppi et al. [21] andInceraGarridoet al. [33] (Fig. 7). Thecharacter-
istic dimension in Re, Nu and Sh numbers is the foamcell diameter
D
pore
; the values measured for the foams studied are shown in
Table 1.
The pressure drop caused by air owing across the foam sam-
ple was measured by a Recknagel micromanometer lled with 96%
ethanol (POCH). Before each reading, in order to avoid hystere-
sis indication, the micromanometer capillary was carefully wetted
with alcohol. The ow resistance of foams was expressed as Fan-
ning friction factor f, derived from the DarcyWeisbach equation:
P
L
= 2f
w
2
0

2
D
pore
(6)
0
2
4
6
8
10
12
14
0 50 100 150 200
S
h
Re
Groppi
Incerra
Fig. 7. Experimental points for FeCrAl foam 20PPI (this work) vs. correlations of
Groppi et al. [21] and Incera Garrido et al. [33].
J. Ocho nska-Kryca et al. / Catalysis Today 216 (2013) 135141 139
Fig. 8. Fanning friction factor vs. Re for FeCrAl foams (20 and 30PPI) and Al foams
(10, 20 and 40PPI). Lines correlations for f according to Table 1.
Fanning friction factors with correlated values are plotted as a
function of Reynolds number in Fig. 8 and the appropriate correla-
tions for each foamare gathered in Table 1.
3. Reactor modelling
The performances of the characterized foams were compared
to the monolith (400cpsi) and packed bed of 300m and 2mm
beads. While the catalyst efciency factor of extremely small cat-
alyst grains (D
pore
=300m) applied for the kinetics tests at the
highest temperature was estimated to 0.87 (as the Thiele modulus
was 0.67), additional particle bed with particles of 2mm in diam-
eter was compared. The experimental kinetic data derived for the
300m catalyst grains have therefore been re-calculated for the
case of 2mmbeds.
The mass transfer in porous medium coupled with chemical
reaction is usually described using the catalyst effectiveness factor

e
being function of a Thiele modulus as follows:

e
=
1
C
AS

l
0
C
AI
(z)
l
dx =
tgh()

(7)
where
= l

p
k
D
A,eff
(8)
For any particle shape, the characteristic dimension l is equal to
the ratio of the particle volume V
p
to its external surface area f
p
.
This, for the geometry approaching sphere, equals to D
p
/6, where
D
p
is the particle (sphere) diameter. The Thiele modulus (and the
effectiveness factor
e
) for the kinetics tests catalyst (300m) and
the modelled packed bed (2mm) are denoted
0
(
e0
) and
1
(
e1
),
respectively. As thegeometrical anddiffusional catalyst parameters
(under the square root sign) are the same, the ratio of the Thiele
moduli is the same as that of the diameters of the grains:

0
=
D
p1
D
p0
=
2
0.3
= 6.67 (9)
AccordingtoEq. (7) theratioof theefciencies (
e1
/
e0
) strongly
depends on the Thiele moduli value.
For the case under study, i.e. D
p1
=2mmand D
p0
=300m, the
ratio (
e1
/
e0
) amounts to 0.31. For this factor (of 0.31) the kinetic
rate constant k has to be multiplied when modelling the packed
bed of 2mmgrains.
For the higher temperatures used in the model the effectiveness
factor and Thiele modulus were recalculated. In those calculations
Fig. 9. Relation between the length of different reactor llings related to the tem-
perature of the DeNOx reaction with assumed supercial velocity w
0
=1ms
1
and
90% NO conversion.
the thickness of a zeolite layer on a foam support was estimated
(based on SEMimages of cross sections of deposited samples) to be
10mand the effective diffusivity, D
A,eff
, of ammoniaair mixture
inzeolitelayer was estimatedas 0.0073m
2
s
1
(assessedtoapprox-
imately 1/10 of the molecular diffusivity determined fromGilliland
equation [32]). The internal diffusion was calculated as a function
of catalytic layer internal porosity and the tortuosity factor.
The comparison of all samples (Table 1) were based on the eval-
uation criteria described in [34]. According to the plug-owmodel
developed for the structures [31] the reactor length for the given
yield was calculated using the following formula:
L =
w
0
S
v

1
k
c
+
1
k

ln

1
1 X

(10)
Pressure drop for each structure was evaluated from Eq. (6)
where L was taken fromEq. (10).
The supercial velocity was set to 1ms
1
and the nal con-
version of NO to 90%. The temperature range 100300

C was
modelled. The comparison of reactor llings is presented as the
length of a reactor (or pressure drop) versus the process tempera-
ture of the DeNO
x
reaction (Figs. 9 and 10).
Fig. 10. Relationbetweenthepressuredropof different reactor llings relatedtothe
temperature of the DeNOx reaction with assumed supercial velocity w
0
=1ms
1
and 90% NO conversion.
140 J. Ocho nska-Kryca et al. / Catalysis Today 216 (2013) 135141
The mass transfer and pressure drop characteristics used for
monolith and packed bed modelling quoted fromthe literature can
be found in Table 1.
4. Results and discussion
The catalytic investigation of a reaction rate over CuZSM-5 in
NH
3
-SCRwas carriedout assuming isothermal conditions along the
reactor and rst order reaction rate with respect to NO [2729]. A
test for internal diffusion was carried out comparing the perfor-
mance of Cu-ZSM-5 samples which differed in particle size (Fig. 3).
Both samples of 1mand 8mgrain sizes showed the same trend
in NOconversion to N
2
upon temperature increase, which excludes
the mass limitation within the grain. Additionally, in order to eval-
uate the inuence of pore diffusion we estimated the WeiszPrater
criterion [35] for NO and NH
3
molecules. As for both molecules
the mean free path is larger than pores in the zeolites used (about
5.5

A) the pore diffusion will be dominated by Knudsen diffusion.
Thus, the WeiszPrater number for both molecules was estimated
to 0.05, the Thiele modulus derived amounts to 0.67 and related
efciency factor was 0.87. The results indicated that the internal
diffusion within a catalyst particle weakly affects kinetic results.
The integral method of analysis was used to evaluate the
apparent rate constant, then the activation energy was deter-
mined according to the Arrhenius equation (Eq. (3)). The kinetics
data derived were: A
0
=3.1310
7
ms
1
and E
a
=89kJ mol

1. The
received data are higher than those found in the literature (around
3550kJ mol
1
) [27]. The reason for the observed discrepanicies
can be the differences in zeolite samples used (ZSM-5 with higher
Si/Al ratio) and different ow condition (no H
2
O in the main
stream). In the Arrhenius plot (Fig. 4) (logarithm of reaction rate
constant versus 1/T) it can clearly be noticed that the transi-
tion between kinetic and diffusional regimes occurs above 200

C
where the reaction is limited by external mass transfer (from
bulk gas phase to the catalytic grain surface). However, no such
behaviour was observed for lower temperature range, i.e. below
200

C. Assuming sufcient mass transfer rate below 200

C, the
derived Arrhenius equation should describe well even higher tem-
perature range. The derived kinetic parameters were applied to
the model that describes the performance of structured catalysts
during DeNO
x
reaction.
Measurements of gassolid heat transfer coefcients were per-
formedinatest rigshowninFig. 5at different owrates andheating
powers. Experimental data were used to determine the correlation
for Nu number, where the percent mean error for the regression
was 4%. The ChiltonColburn analogy was used to derive the mass
transfer representation for each metal structure (Table 1).
Fig. 6 shows the dependence of the Sherwood number on the
Reynolds number andthe correlatedSherwoodnumber. The higher
gas velocity (or Re number) the higher heat transfer coefcient (Sh
number). Sh decreases distinctly as foam density (thus foam spe-
cic surface area) increases (or D
pore
decreases). It is too early yet
to interpret the results reliably. However, the Sherwood number
increases strongly (approximately linearly) with the D
pore
for any
assumed value of Re. The mass transfer coefcient k
c
derived from
the Sh depends on the pore diameter similarly to the Sherwood
number (linearly as well). The mass transfer coefcient is referred
to the unit surface area of the foam neglecting the surface shape,
thus the pore size as well. Consequently, as k
c
increases with the
pore diameter D
pore
, the transfer mechanism has to change with
D
pore
. Perhaps, larger pore diameter enhances the mixing within
the foamstructure. For larger pores (or lower pore density PPI) the
turbulences are slightly more probable.
The distinct difference between aluminium (Al) and FeCrAl
foams results from different morphology of the metallic struts
connecting the pores. The Al struts are more rounded and have a
smooth surface while that of FeCrAl are sharp and rough thereby
causing higher heat/mass transfer intensity. These relations are
partially, although not perfectly, reected by the ow resistance
behaviour (Fig. 8). The ow resistance through the foam packing
was determined in the same experimental set-up. Results shows
that the ow resistance (pressure drop) increases with density
(data not shown) contrary to the Fanning friction factor, f, which
decreases with the PPI increase (Fig. 8). Similar to the relationship
for the Sherwood number, the Fanning friction factor for the FeCrAl
foams is distinctly higher than that of the Al foam displaying the
same PPI value. This phenomenon may be explained, just like for
the mass transfer, by the sharp edges and rough surface of the
FeCrAl foam.
The Fanning friction factor is seriously inuenced by the geo-
metric properties of the foams including general geometry shape
and parameters, roughness of the surface and shape of the struts.
These parameters should be determined with great care by the
researchers themselves rather than relying on the data supplied by
the foamproducers (the manufacturer supplied data might signif-
icantly differ from the actual value because producers commonly
present the characteristic parameters of the initial polyurethane
foamwhich serves as a template in metal foamproduction). How-
ever, the proper determination of mean diameter of the megapores
as well as mean strut diameter is not an easy task because of highly
randomand irregular foamstructure. While different models were
used in the literature to calculate the specic surface area (i.e.
dodecahedron [36], tetrakaidecahedra [37] and cubic cel [17]) we
believe that the most reliable model should be based on a statisti-
cal analysis of very precise determination of geometric parameters
(e.g. by tomography). This will be the origin of our next studies.
The experimental mass transfer data for FeCrAl 20PPI were
compared with the correlation found in literature (Fig. 7). Groppi
et al. [21] determined the correlation for both metallic and ceramic
foams withporedensity5-15PPI for theReupto200andInceraGar-
ridoet al. [33] studiedthe correlationfor ceramic foams (1045PPI)
within the range of Re from 7 to 1100 and the inuence of topol-
ogy on mass transfer parameters. A comparison shown in Fig. 7
reveals that for Re from 50 to 200, the data t to the calculated
values based on literature correlations within 20% error. Discrep-
ancy for Re lower than 50 can be due to different experimental
procedure employed to determine the mass transfer coefcient
we have used the ChiltonColburn analogy based on heat trans-
fer experiments while the cited authors used chemical reaction on
foamwith washcoated catalyst.
Once the heat and mass transport properties and pressure drop
of foams had been determined, it became interesting to compare
metal foams with other structures that are commonly used as a
catalyst supports, e.g. honeycomb monoliths and packed beds of
pellets. Since many applications in environmental catalysis require
lightweight and compact reactors we decided to use the size of the
reactor as an important factor to consider in selection of a cata-
lyst geometrical structure. The reaction rate constants (dependent
on reaction temperature) developed from detailed kinetic stud-
ies of catalytic selective reduction of NO were used to predict the
reactor sizerequiredfor (assumed) nal NOconversionof 90%. Clas-
sical literature expressions (Table 1) were implemented for packed
beds of spheres (mean particle size 2mm and 300m) and honey-
comb monoliths with square cells (400cpsi), whereas correlations
derived in this work were used for the metal foams. As shown in
Fig. 9 for gas velocity 1ms
1
anda nal NOconversionof 90%, more
compact reactors are required using an Al foam of 40PPI than a
honeycomb monolith. At temperature 250

C the foamwith 40PPI


wouldneed30%of the honeycomb monolithreactor. Clearly a reac-
tor based on a packed bed of spheres of 300mmean particle size,
due to high mass transfer, performs much better than any of the
J. Ocho nska-Kryca et al. / Catalysis Today 216 (2013) 135141 141
other structures, however, taking into account the consideration
fromexperimental section (excluding of such a packing due to high
pressure drop in the industrial application) and comparing foams
with a packed bed of 2mm spheres, foams present lower reactor
length throughout whole temperature range studied, i.e. at 250

C,
40PPI foamreactor is 18% shorter.
Comparison of Fig. 9 (reactor length vs. temperature) with
Fig. 10 (pressure drop vs. temperature) seems very interesting.
In Fig. 9, the lower the characteristic dimension (grain or pore
diameter) the shorter the reactor, with only single exception: reac-
tor with Al foam 10PPI is longer than that utilizing FeCrAl foam
30PPI. This most probably results from the better mass transfer
properties of FeCrAl foams. On the contrary, the higher foamden-
sity (or the smaller pore size), the lower ow resistance displayed
by the reactor. Allowing for a certain inconsistency between Al
and FeCrAl foams, increased pore density causes shorter reactor
(thus lower catalyst consumption) and lower pressure drop (thus
lower pumping energy). This phenomenon proves very promising
behaviour of foams as structured catalyst carriers.
Overall, metallic foams, thanks to very high void fractions, oper-
ate better than packed beds of spheres whose main drawback is a
high pressure drop across the bed (Fig. 10). The opposite is the case
for monolithreactors where a fullydevelopedlaminar owexhibits
a small pressure drop, and the limiting factor is mass transfer. In
that case, optimizing the foam density we might end up with a
structure exhibiting better mass transfer properties than honey-
comb monolith and a comparable or even smaller pressure drop.
In fact, lower pressure drops over the whole range of temperatures
are presented by 40PPI Al. Such an optimization should be the case
of future studies as there could be difculties in direct synthesis of
the zeolite on the higher density foams without blocking the pores.
5. Conclusions
In the environmental catalysis usually a compromise must be
reached between mass transfer and pressure drop performances of
a catalyst. The ideal structured catalyst should yield very high con-
versions under diffusioncontrolledconditions to meet the required
emission limits while keeping the pressure drop as lowas possible
to reduce energy costs.
In this study we proved that metallic foams are excellent candi-
dates for that purpose. The insitusynthesis of the zeolite ona metal
substrate would largely improve the heat transfer within the cata-
lyst bed while the cell porosity of the metal foam and a thin layer
of a zeolite coating ensure high diffusion and good mass transfer
characteristics of structure catalyst.
Metal foams exhibit marked advantages over packed beds as
for the equivalent conversions they generates pressure drop over
200 times smaller. Additionally they offer considerable reduction
of reactor size when compared to honeycomb monolith. Another
advantage is that they can be produced in any desired shape and
size so they might easily replace industrial working catalyst.
Acknowledgements
This work was supported by BRIDGE Programme (No. 2010-
1/4) within the Foundation for Polish Science co-nanced by the
EU Structured Funds and partly by the project from National Sci-
ence Centre NN 209438639. The work of Przemysaw Jodowski
was nanced by Polish Science Foundation MPD Programme co-
nanced by the EU European Regional Development Fund.
References
[1] J. Banhart, Progress in Materials Science 46 (2001) 559632.
[2] A.N. Pestryakov, V.V. Lunin, A.N. Devochkin, L.A. Petrov, N.E. Bogdanchikova,
V.P. Petranovskii, Applied Catalysis A 227 (2002) 125130.
[3] M. Huff, L.D. Schmidt, Journal of Physical Chemistry 97 (1993) 1181511822.
[4] A.N. Pestryakov, A.A. Fyodorov, M.S. Gaisinovich, V.P. Shurov, I.V. Fyodor-
ova, T.A. Gubaydulina, Reaction Kinetics and Catalysis Letters 54 (1995)
167172.
[5] L. Cambell, Inventor, Catalyst for the production of nitric acid by oxidation of
ammonia, United States patent US 5217939 (8 June 1993).
[6] O.Yu. Podyacheva, A.A. Ketov, Z.R. Ismagilov, V.A. Ushakov, A. Bos, H.J. Veringa,
Reaction Kinetics and Catalysis Letters 60 (1997) 243250.
[7] J.T. Richardson, M. Garrait, J.K. Hung, Applied Catalysis A 255 (2003) 6982.
[8] A. Koodziej, J. ojewska, Catalysis Today 147 (2009) S120S124.
[9] A. Koodziej, J. ojewskac, J. Ocho nska, T. ojewski, Chemical Engineering and
Processing 50 (2011) 869876.
[10] M.A. Gomez-Garcia, V. Pitchon, A. Kiennemann, Environment International 31
(2005) 445467.
[11] P. Pietrzyk, B. Gil, Z. Sojka, Catalysis Today 126 (2007) 103111.
[12] K. Sun, H. Xia, Z. Feng, R. Santen, E. Hensen, C. Li, Journal of Catalysis 254 (2008)
383396.
[13] J. ojewska, A. Koodziej, P. Dynarowicz- atka, A. Weseucha-Birczy nska, Catal-
ysis Today 101 (2005) 8191.
[14] J.W. Paek, B.H. Kang, S.Y. Kim, J.M. Hyun, International Journal of Thermophysics
21 (2000) 453464.
[15] J.F. Despois, A. Mortensen, Acta Materialia 53 (2005) 13811388.
[16] A. Bhattacharya, V.V. Calmidi, R.L. Mahajan, International Journal of Heat and
Mass Transfer 45 (2002) 10171031.
[17] T.J. Lu, H.A. Stone, M.F. Ashby, Acta Metallurgica 46 (10) (1998) 36193635.
[18] K.I. Salas, A.M. Waas, Journal of Heat Transfer 129 (9) (2007) 1217.
[19] I. Ghosh, Journal of Heat Transfer 130 (3) (2008) 034501.
[20] G. Leonardo, G. Groppi, E. Tronconi, American Chemical Society 44 (2005)
90789085.
[21] G. Groppi, L. Giani, E. Tronconi, Industrial and Engineering Chemistry Research
46 (12) (2007) 39553958.
[22] G. Leonardo, G. Groppi, E. Tronconi, Industrial and Engineering Chemistry
Research 44 (2005) 49935002.
[23] J.G. Fourie, J.P. Du Plessis, Chemical Engineering Science 57 (2002) 27812789.
[24] D. Edouard, M. Lacroix, C.P. Huu, F. Luck, Chemical Engineering Journal 44 (2)
(2008) 299331.
[25] J. Ocho nska, A. Rogulska, P. Jodowski, M. Iwaniszyn, M. Michalik, W. asocha,
A. Koodziej, J. ojewska, Topics in Catalysis 56 (2013) 5661.
[26] J. Ocho nska, D. McClymont, P.J. Jodowski, A. Knapik, B. Gil, W. Makowski, W.
asocha, A. Koodziej, S.T. Kolaczkowki, J. ojewska, Catalysis Today 191 (2012)
611.
[27] T. Komatsu, M. Nunokawa, Il S. Moon, T. Takahara, S. Namba, T. Yashima, Journal
of Catalysis 148 (1994) 427437.
[28] S. Brandenberger, O. Krcher, Oliver, A. Tissler, R. Althoff, Catalysis Reviews
Science and Engineering 50 (2008) 492531.
[29] S. Roy, M.S. Hegde, G. Madras, Applied Energy 86 (2009) 22832297.
[30] M. Iwaniszyn, M. Jaroszy nski, J. Ocho nska, J. ojewska, A. Koodziej, Research
Papers of the Institute of Chemical Engineering Z.15, Polish Academy
of Sciences, 2011, Available in the Internet http://www.iich.gliwice.pl/
les/303/Zeszyt%2015%20(2011).pdf
[31] A. Koodziej, J. ojewska, J. Tyczkowski, P. Jodowski, W. Redzynia, M. Iwaniszyn,
S. Zapotoczny, P. Ku strowski, Chemical Engineering Journal 200202 (2012)
329337.
[32] E.R. Gilliliand, Industrial and Engineering Chemistry 26 (1934) 681685.
[33] G. Incera Garrido, F.C. Patcas, S. Lang, B. Kraushaar-Czarnetzki, Chemical Engi-
neering Science 63 (21) (2008) 52025217.
[34] A. Koodziej, J. ojewska, Topics in Catalysis 4243 (2007) 475.
[35] M.A. Vanice, Kinetics of Catalytic Reactions, Springer Science, New York,
2005.
[36] D.L. Duan, R.L. Zhang, X.J. Ding, S. Li, Materials ScienceandTechnology22(2006)
13641367.
[37] A. Inayat, J. Schwerdtfeger, H. Freund, C. Krner, R.F. Singer, W. Schwieger,
Chemical Engineering Science 66 (2011) 27582763.
[38] N. Wakao, S. Kaguei, Heat and Mass Transfer in Packed Beds, Girdon and Breach
Science Publisher, NewYork, 1982.
[39] S. Ergun, Chemical Engineering Progress 48 (1952) 89.
[40] R.D. Hawthorn, AIChE SymposiumSeries 70 (137) (1974) 428.

Potrebbero piacerti anche