Sunteți pe pagina 1din 53

Catalysis Reviews

ISSN: 0161-4940 (Print) 1520-5703 (Online) Journal homepage: http://www.tandfonline.com/loi/lctr20

Monolithic Catalysts for Nonautomobile


Applications
Said Irandoust & Bengt Andersson
To cite this article: Said Irandoust & Bengt Andersson (1988) Monolithic
Catalysts for Nonautomobile Applications, Catalysis Reviews, 30:3, 341-392, DOI:
10.1080/01614948808080809
To link to this article: http://dx.doi.org/10.1080/01614948808080809

Published online: 03 Jan 2007.

Submit your article to this journal

Article views: 127

View related articles

Citing articles: 107 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=lctr20
Download by: [Bibliotheek TU Delft]

Date: 12 September 2016, At: 10:04

CATAL REV.-SCI.

E N G . , 3 0 ( 3 ) . 341-392 (1988)

Monolithic Catalysts for Nonautomobile


Applications
SAID IRANDOUST AND
Department of Chemical
Chalmers University of
S - 4 1 2 96. Gothenburg.

BENGT ANDERSSON*
Reaction Engineering
Technology
Sweden

..................
MODELING . . . . . . . . . . . . . . . . . . . . .
A . Single-Phase Flow . . . . . . . . . . . . . . .
B . Cross Flow . . . . . . . . . . . . . . . . . .
C . Two-Phase Flow . . . . . . . . . . . . . . . .
111. COMMERCIAL MONOLITHS . . . . . . . . . . . . .
A . Material . . . . . . . . . . . . . . . . . . . .
B . Geometric Shape of Monolith Channels . . . . .
C . Selection of Monolith Substrate . . . . . . . .
IV . APPLICATIONS . . . . . . . . . . . . . . . . . .
A . Methanation . . . . . . . . . . . . . . . . . .
B . Hydrogenation and Dehydrogenation . . . . . .
C . Hydrotreatment . . . . . . . . . . . . . . . .
D . Biochemical Reactions . . . . . . . . . . . . .
E . Electrochemical Reactions . . . . . . . . . . .
F . Oxidation . . . . . . . . . . . . . . . . . . .
G . Catalytic Combustion . . . . . . . . . . . . .
.
I1 .
I

INTRODUCTION

*To whom correspondence should be addressed


34 1

Copyright 0 1988 by Marcel Dekker. Inc .

342
343
343
355
358
365
365
372
372
374
374
375
377
378
379
380
381

IRANDOUST AND ANDERSSON

342

......
..............
CONCLUSION. . . . . . . . . . . . . . . . . . . .
SYMBOLS. . . . . . . . . . . . . . . . . . . . . .
ACKNOWLEDGMENT. . . . . . . . . . . . . . . . .
REFERENCES. . . . . . . . . . . . . . . . . . . .
Stationary-Source Emission Control
Other Applications.

H.
I.

V.

I.

383
385
385
386
387
388

INTRODUCTION

The use of monoliths in automotive emission control systems


has dominated the applications of monoliths. However, other aplications have also become of interest due to unique features of
monoliths
Several excellent reviews on emission control [ l - 3 1 and catalytic oxidations [ 4 ] have been published during the last few
years. Deluca and Campbell [51 have reviewed monoliths used
as catalyst supports. This paper reviews flow characterization,
heat and mass transfer, mathematical modeling, properties of
monolithic supports, and nonautomobile applications of monoliths.
The main features of monolithic catalyst support are:

0
0
0
0
0

Low pressure drop


Large external surface
Uniform flow
Low axial dispersion
Low radial heat flow (adiabatic)

The flow in monoliths is usually laminar in straight channels,


which gives very low pressure drop. The pressure drop is two
to three orders of magnitude lower than for a bed of spherical
particles with a particle diameter of the same order of magnitude
as the channel width of the monolith [ 6 1 .
Mass and heat transfer in monoliths is in general somewhat
lower than in conventional fixed-bed reactors. There is no radical
bulk transport of fluid between the parallel channels in a monolith. Consequently there is no radial heat transfer owing to fluid
flow. The radial heat transfer occurs only by conduction through
the laminar flow and the solid walls. Ceramic monoliths with very
low thermal conductivities are usually adiabatic. Metallic monoliths have high heat conduction in the solid materid and the
radial heat transfer is of the same magnitude as in packed beds.
Table 1 shows a comparison between a monolithic catalyst reactor
and a fixed-bed reactor.

MONOLITHIC CATALYSTS

343

TABLE 1
Comparison Between a Monolithic Catalyst Reactor
and a Fixed-Bed Reactor
Fixed bed
(d,=1.5 mm)

Ceramic Monolith
(d = 1.5 mm)
2.8

10-3

1.8

10-4

Pressure drop, gas, bar /m


(Air, 1 . 1 m/s at 293 K)

0.23

Pressure drop, liquid, barlm


(Water, 1 . 3 . 1 0 - 3 m / s
at 293 K)

5.1

Solid fraction, %

60

30

Surface-to-volume ratio, rn-

2400

1900

Diffusion length in the


catalyst, m m

0.75

(0.15

11.

10-

MODELING

Monoliths a r e manufactured for many different applications


and in very different shapes. However, for modeling, three
major flow patterns of monoliths describe most applications :
single-phase flow, cross flow, and two-phase flow. In singleand two-phase flow the fluids flow in parallel channels, while in
cross-flow monoliths one of the reactants or a cooling fluid flows
in channels perpendicular to the main flow.

A.

Single-Phase Flow

Most work on monoliths has been performed with singlephase flow. The flow is usually laminar with a developing velocity profile. Only at very high velocity in gas turbines is the
flow turbulent. The typical flow regimes for different reactions
are given in Table 2 .
1. Pressure Drop

It is chiefly the low pressure drop that has given the monoliths their applications. In one of the first articles on ceramic
honeycomb structures, Johnson et al. 101 measured the pressure

IRANDOUST AND ANDERSSON

344

TABLE 2

Conditions for Different Reactions


Pressure, Temperature,
bar
K

Application

Velocity,
m Is

Re

Ref.

Car exhaust

1- 3

900- 1200

14

200

171

Gas turbine

3-25

400-2000

23-34

2000- 3000

[ 71

Methanation

60

670

21

100

181

Hydrogenation

500-600

10-3-0.05

0.1-10

91

drop. Figure 1 shows their results. The measured pressure


drop is in close agreement with those calculated from the Poiseuille equation
AP =

3 2 LV
~
-

d2
(All symbols used are defined in the list at the end of the text.)
Other studies by Pratt and Cairns [ 111 and Howitt 1121 confirm
these results. Metallic monoliths exhibit lower pressure drops
than ceramic types with similar cell densities, because of larger
open area resulting from low wall thickness 1111.
In Figure 1 is also plotted the pressure drop, calculated
from the Ergun equation [ 131, of a corresponding packed bed
with the same fluid-solid contact area as the monolith. The pressure drop is about two orders of magnitude higher in the packed
bed than in the monolith. Moreover, the pressure drop in packed
beds is often increased further due to deposition of small particles from the feed o r from crushing of the catalyst particles.
2.

Entrance Effects

The entrance effect is usually small. The velocity profile is


fully developed very close to the inlet. Skelland [141 has given
a criterion for the entrance length in which the velocity profile
becomes fully developed :

MONOLITHIC CATALYSTS

345

CALCULATED CURVES BY POlSEUlLLE'S

--,

LAW:

---

FLOW,SCFM/IN'
F I G . 1. Variation of pressure drop with gas flow per unit
of cross section (D: channel diameter). Reprinted with permission
from Ref. 10. Copyright 1961, American Institute of Chemical
Engineers.

-Z =

0.05 Re

(2)

This length is small compared to the total length of the monolith.


In most calculations the authors have neglected the entrance
effect and considered the flow laminar throughout the monolith.
The length required to obtain a fully developed concentration
profile can be estimated from
-Z =

0.05 Re Sc

(3)

IRANDOUST AND ANDERSSON

346

and the corresponding thermal entry length is obtained by replacing Sc with Pr:
Z

0.05 Re P r

(4)

For most monoliths with a channel diameter d < 3 mm and laminar


gas flow, both concentration and thermal profiles are fully developed in all but a negligible part of the reactor. In liquids,
however, the effect of the developing concentration profile cannot be neglected due to the low radial diffusion and corresponding high Sc number (-1000). This is usually no problem
in actual calculations. The distributed parameter models take
this into account by the boundary conditions, and the lumped
parameter models have to use local mass transfer coefficients
which vary with distance from the inlet.
Boersma et al. [15] studied the difference between developed
and developing profiles. They used a honeycomb containing 169
tubes, each with a diameter of 2 . 4 mm and a length of 10 cm.
The first 5 cm was impregnated and highly active while the last
5 cm was nonimpregnated. When the feed gas enters the impregnated part, first the catalyst will be exposed to a developing
profile. By turning the monolith 180 the velocity profile in the
catalytic part will be fully developed. The difference between
the experiments are s m a l l but significant. Figure 2 shows these
differences. The conversion is, as expected , somewhat higher
with a developing profile.
Wendland 1161 has studied the effect of entrance regions on
conversion efficiency for oxidizing monolithic catalytic converters.
He has shown that segmenting the monolith normal to the flow
direction enhances conversion efficiency. This increase is smaller
for large monoliths.
3.

Axial Dispersion

Packed beds are often used when high conversion is required, due to their low axial dispersion. In monoliths the axial
dispersion is much more dependent on channel diameter and flow
rate. In laminar flow the axial dispersion can be calculated from
the following equation given by Taylor [ 171 :
Pe

a =

v d

D=

192

D
v
d,

ea

which can be rearranged into

Re Sc

>

(5)

MONOLITHIC CATALYSTS

347

: theoretical profile, fully developed flow


: theoretical profile, developing flow
: theoretical profile, undeveloped flow
0

experimental, developing flow,' Da = 1.28, Sc

experimental, fully developed flow, Da = 1.28, Sc = 1.09

1.09

1 : Da = 1.28, Sc = 1.09

2,3,4 : Da = 2.7, Sc = 1.09

L / ( P e r R)

FIG. 2. Entrance effect in highly active monolith. Oxidation


of ethylene (inlet concentration C , = 300-350 ppm) at 400 K .
Reprinted with permission from Ref. 15. Copyright 1978, American
Chemical Society.

- v2d2

EE

Dea -

In turbulent flow Taylor 1181 gives


pea

-- -0 . 2 8

10'

Jf

<

Re

<

lo4

(7)

in which the friction factor can be calculated from


f = 0.046 Re-

(8)

In packed beds the axial dispersion is given by [19]:


v d
Pe

=-

= 2,

'ea

Re

<

500

IRANDOUST AND ANDERSSON

348

and
v d

-2

Dea - 2

A condition for Eqs. (9) and (10) is that dt/dp > l o . In a typical case of air at room temperature with v = 0 . 1 m / s and dt =
m , the dispersion coefficient will be 2
m2/s
dp =
for the monolith and 5
l o w 5m2/s for the packed bed. For gasphase flow the axial dispersion is much less in the monolith than
in the packed bed. In liquid phase, however, the diffusivity
is four orders of magnitude smaller, and the axial dispersion
will be significant in the monolith; while in packed beds where
the dispersion is independent of the diffusivity, the flow pattern
is close to ideal plug flow.
Dispersion in liquid flow can be decreased substantially by
introducing a gas or a nonmiscible liquid into the flow. In twophase slug flow the residence time distribution is close to ideal
plug flow. The flow patterns of two-phase flow are discussed
further in a later section.
4.

Mass and Heat Transfer

Mass and heat transfer has been studied theoretically by


Shah and London [ 61 and Young and Finlayson 1201. Young and
Finlayson solved the steady-state heat transfer equations for
constant wall temperature and linear wall temperature for different geometries. Table 3 shows their calculated asymptotic
Nusselt numbers for different geometries.
The local Nusselt number near the inlet decreases with distance from the inlet. For the Graetz problem in a circular tube,
Sellars et al. [21] showed that the Nusselt number in the entry
length follows the formula

(:

Nu = B - Re Pr

y3

Young and Finlayson [201 have calculated the constant B for


different geometries. Their result is shown in Table 3.
Experimental studies of the mass and heat transfer have been
performed by several authors [15, 2 2 , 231. Hawthorn 1241 has
given the following correlations :

MONOLITHIC CATALYSTS

349
TABLE 3

Calculated Asymptotic Nusslet Numbers for Different Geometries


Young and Finlayson [201

Constant B
in Eq. ( 11)

Geometry

Nusselt ,
constant
wall temp.

Circular

Shah and London [ 6 ]


Nusselt , Nusselt
constant constant
wall temp. heat flux
3.657

4.364
3.608

Square :
b/a = 1

1.16

2.976

2.967

Rectangle :
b/a = 0 . 5
b/a = 0 . 2 5

1.23
1.38

3.392
4.441

3.391
4.439

Equilateral triangle

2.491

2.470

3.111

Trapezoid :
b = l
c = 0.5
b = 0.75 c = 0.25

2.744
2.120

2.617

Sinusoidal

1 + 0.095 Re Pr
1 + 0.095 Re Sc

Here A is a constant depending on the shape of the tube cross


section: A = 3.66 for circular tubes and A = 2.35 for triangular
tubes.
Votruba et al. [231 soaked the monolith in different liquids
and measured the weight loss at different gas flow rates. They
fitted their mass and heat transfer results to the equations
Sh = 0.705
Nu = 0.571

(
(

r3
q3

Re -

Re -

ScoeS6

IRANDOUST AND ANDERSSON

350

<

Re

<

480

0.57

<

Sc

<

3.3

Pr = 0.74
They used air in all their experiments, and no heat transfer dependence on the Prantl number could be calculated.
Both Boersma et al. [251 and Heck et al. [ 2 6 ] found that
their data correlated slightly better with Hawthorn's model.
Figure 3 shows the two models and some experimental data. The
main difference between the models is at low Reynolds number,
where Votruba's model predicts a continuous decrease in Sherwood number with decreasing Reynolds number while Hawthorn's
model reaches an asymptotic value.
The observed mass transfer coefficients are within a factor
of 2 of the theoretical calculations presented by Young and
Finlayson [ 201. The reason for this deviation of experiments
from theory is not yet established. Possible causes are experimental errors, errors in the values of diffusivities and Viscosities used in the calculations, or that the theoretical models are
too simple.
In Figure 3 the Sherwood number is also plotted for a packed
bed with the same surface area per volume of reactor a s the
monolith. The Sherwood number for the packed bed is the same
at low flow rates, but at high flow values it becomes more than
one order of magnitude higher. However, the lower Sherwood
number can be compensated by using a monolith with a higher cell
density. Decreasing the channel diameter by a factor of 3 increases the mass transfer coefficient kLa by a factor of 9. The
lower limit of channel diameter is determined by manufacturing
methods and is at present about 1 m m .
An enhanced mass transfer in liquid phase can be obtained
by introducing plugs of a nonmiscible liquid or a gas into the
flow. This disturbs the laminar flow and increases the radial mass
transfer. This effect is more thoroughly discussed later in connection w i t h two-phase flow.
5.

Mathematical Modeling

Heat and mass transfer with homogeneous and heterogeneous


reactions has been studied by many authors. Table 4 gives a list
of some of the theoretical calculations that have been published.

MONOLITHIC CATALYSTS

351

Sh
30 -

25-

Packed bed E3]


Hawthorn [24]
Votruba e t al.p3]
+++ Data from Boersma et aI. @

--_

20 -

___.-.--.+*+$
_ _

o--0

.-. -.

._ .-.

-. -. -.-.

- ----------------

-. -.-.

_ _ _ _ _ _
+^ _ _ _ _ _ _ _

200

400

600

800

1000

1200

1400

1600

1800
Re ;Rep

FIG. 3 . Mass transfer in active wall reactor and packed bed


with laminar flow. Combustion of ethene.

a. Distributed-Parameter Models
All published distributed-parameter models have approximated
the monolith with an adiabatic cylinder. The mass and heat transfer in a circular tube with fully developed incompressible laminar
flow can be written

a
at
Ia
and

=D. i -(ra
1 r ar

+ Di-

az2
IIa

IIIa

IVa

x)

a if

(16)

IRANDOUST AND ANDERSSON

352

TABLE 4

Mathematical Modeling of Heat and Mass Transfer


Given in the Literature

Ref.

Model

Condition

Heat
conduction
in
solid

Lumped

Steady

No

No

No

Distributed

Steady

Yes

No

No

Distributed

Steady

No

No

No

Lumped
Distributed

Transient
Steady

No
No

No
No

No
No

Lumped

Steady

No

No

No

Distributed

Transient

Yes

No

Yes

Lumped

Transient

Yes

No

No

Lumped

Steady

Yes

No

No

Lumped

Steady

Yes

No

Yes

Lumped

Steady

Yes

No

No

Distributed

Steady

Yes

No

No

Lumped

Steady

Yes

No

No

Distributed

Steady

Yes

Yes

No

I : lumped
11: distr.
111: distr.

Transient
Transient
Transient

Yes
Yes
Yes

No
No

No
No

Yes

N0

2)

Internal
mass
transfer
in solid

Radiation
-

aTf +
at

a 2Tf
+ 2v(l

hf-

az2
IIb

Ib

5)

IIIb

aaTf
=
z

Xf

a r (r
IVb

Only symmetrical solutions are usually considered, and the


boundary conditions at r = 0 are written

acif

-=-

ar

aTf
ar

= o

(17)

MONOLITHIC CATALYSTS

353

The mass and heat transfer in the porous wall is described by

3
at

7
a2cis= D

+D

ei

1a ( r 7
a is )
+
e. r ar
1

Va

VIa

Vb

VIIa

VIb

VIIb

vijrj

(19)

VIIIa

VIIIb

There is no diffusion or heat transfer to the neighboring channel :


aCis -

ar

a TS = O
ar

a t r = R + S

At the boundary between t h e fluid and solid ( r = R ) , the balances are written

a is
-De
i

xs

ar

a TS
ar

--

a if

D. 1 ar

a Tf

ar

(23)

All parts in these equations are not equally important. Young


and Finlayson 1201 have solved the equations for a catalytic converter. They found that accumulation ( i a ) and axial diffusion
(IIa) are small compared to convective transport (IIIa) and radial
diffusion (IVa). Also, heat accumulation (Ib) and axial heat conduction (IIb) can be neglected compared to convection (IIIb) and
radial heat transport ( I V b )
In the solid phase the concentration accumulation (Va) and
axial dispersion (VIa) are small. It is only in very fast reactions that the mass transfer resistance in the very thin washcoat
will affect the reaction rate. Zygourakis and A r i s 1361 have calculated the effectiveness factor in the washcoat for oxidation of
hydrocarbons and carbon monoxide. They found effectiveness
factors between 0.05 and 1 . 3 . In the heat balance the radial
conduction (VIIb) is the only part which usually can be neglected.

IRANDOUST AND ANDERSSON

354

b. Lumped-Parameter Model
Most calculations have been done with lumped-parameter
models. In these models only the average concentration and
temperature in the phases are considered. The lumped-parameter models are sufficient in many applications where t h e gradients in the fluid and the solid are small.

-a '-if

at

-a T- f
at

-., a a'ifz

"

- - kfi(Cif

- Cis)

a Tf
-v az

with the boundary conditions


kfi (Cif

- Cis) =

wijrj

~,'ps

aTS

a 2TS

at

az

- + A-

r. ( - A H . )

+
j

(27)

Young and Finlayson [ 371 compared a distributed-parameter


model with a lumped-parameter model. They found that the
local Sherwood and Nusselt numbers have a wide variation.
Their calculated local Sherwood and Nusselt numbers for oxidation of carbon monoxide in a catalytic converter are plotted in
Fig. 4. A s expected from Eq. (11) there is an inlet effect on
the Nusselt number. Before the light-off, the best average
Nusselt number is for constant flux, while after light-off the
best Nusselt number is for constant wall temperature. In the
ignition area the Nusselt number is several times higher. The
axial transport of heat is important. Model 3 in Fig. 4 does not
allow for axial heat transfer in the monolith. With increasing
axial heat conduction in the solid wall, the predictions by a
lumped-parameter model will be less accurate.
Radial heat transfer in m e t a l monoliths has been studied by
Flytzani-Stephanopoulos et al. [ 381. In Fig. 5 the radial heat
transfer in a metal monolith is compared to a fixed bed. The
radial heat transfers in metal monoliths and fixed beds are of the
same magnitude.

MONOLITHIC CATALYSTS

7.0-

355

II
11

B.

Cross Flow

In cross-flow monoliths, the two fluids are separated by a


permeable or impermeable monolithic wall. In this construction
the two fluids can be different reactants diffusing from opposite
sides into the catalytic layer; or one of the fluids can be a cooling media, thus obtaining very efficient cooling (Fig. 6).
The flow characteristics, pressure drop, and external mass
transfer are described by the same models as in the parallelchannel monoliths.
1.

Heat Transfer

Heat transfer in cross-flow monoliths with catalytic walls is


very efficient since the coolant is in close contact w i t h the solid
wall on which the reaction occurs. In metallic monoliths there
is only about 50 p m of metal between the reactant zone and the
coolant.

IRANDOUST AND ANDERSSON

356
860
G = 10 Ib/hr Np-l
S.V. = 17.000 hr

805

PACKED BED
1/8" x 1/8" alumina pellets

--- METAL MONOLITH

Fecralloy@with 400 holedin

750

Y
I-

*, ,a

695

3
l-

a
w

640

I-

0
UI

m
-I

5
a

585

---

530

475

-/

420

0.0

CENTER

0.33

0.67

REDUCED RADIAL DISTANCE, x/R

1 .o

WALL

FIG. 5. Experimental radial temperature profiles in the


metal monolith and packed bed ( H : axial monolith length, x :
radial distance, R : monolith bed radius). Reprinted with permission from Ref. 38. Copyright 1986, Pergamon Press.

MONOLITHIC CATALYSTS

357

FIG. 6. Principle of the cross-flow catalyst. Reprinted with


permission from Ref. 39. Copyright 1982, Pergamon Press.
The heat and mass transfer into the reactant fluid is described by the equations in Sec. I I . A . 5 . The boundary condition for the mass transfer in E q . ( 2 1 ) is changed to
-a is
-

ar

- 0

The heat transfer into the cross-flow fluid is described by


conduction from the washcoat through the monolith wall

-a T
-s at

aTm
ar

a t r = R + S

and into the cooling fluid

The flow of the cooling fluid is perpendicular to the reactant


flow, which makes Eqs. (16) to ( 3 0 ) difficult to solve. The
temperature will have a radial and axial variation.

358

IRANDOUST AND ANDERSSON

Degnan and Wei [ 4 0 , 411 calculated the heat transfer between


the flows with the assumption that the different monolith flows
were isothermal. They found that their model described the experimental data well. The largest deviation occurred with thinwall monoliths where the heat transfer between the fluids was
high and the heat transfer in the walls parallel to the fluids was
low.
2.

Mass Transfer and Reaction

In a series of articles, de Vos et al. [39, 42, 431 have described a monolith reactor in which the reactants were separated
by a catalytic wall. The diffusion and reaction are described by
Eq. (19) but the diffusion is from opposite sides, which gives
the concentration profiles shown in Fig. 7. These concentration
profiles will give very different selectively conditions compared
to catalysts where the reactants diffuse from the same side [43].
The thick porous walls (>0.1 mm) give severe mass transfer
resistance for fast reactions. In order to improve the mass
transfer, Hatziantoniou et al. [44] developed a catalytic wall for
gas-liquid reactions in which most of the wall is not wetted by
the liquid. Only a narrow reaction zone, in which the gas and
liquid have to be in contact, is wetted. Since the gas-phase
diffusivity is about four orders of magnitude higher than the
liquid-phase diffusivity , the effectiveness factor is increased
substantially (Fig. 8).
In this improved catalytic wall the mass transfer resistance
in the liquid channel becomes the rate-determining step. Static
mixers in the liquid channel were installed to increase the mass
transfer.
C.

Two-Phase Flow

Two-phase flow in monoliths gives a very beneficial flow


pattern. The presence of slugs in the flow decreases the axial
dispersion and increases the radial mixing in the channels.
Monolithic reactors with two-phase flow are characterized by
very efficient gas-solid and liquid-solid mass transfer.
Several flow patterns are possible in gas-liquid two-phase
flow. Figure 9 shows four common patterns. The transition
between different flow regimes has been studied by many authors
[ 45-48]

In bubble flow the bubbles are much smaller than the channel
diameter. The bubbles disturb the laminar flow and increase the

-0
1

-0
1

FIG. 7. Dimensionless concentration profiles of hydrogen and


reactant A in the pore in a cross-flow monolith at various stoichiometric ratios (5 < 4 < 80). (CH: concentration of hydrogen in
liquid, CHO: solubility of hydrogen in liquid, 4: Thiele modul u s ) . Reprinted with permission from Ref. 42. Copyright 1982,
Pergamon Press.

IRANDOUST AND ANDERSSON

360
C" x 10

FIG. 8. Concentration profiles for hydrogen ( H ) and nitrobenzoic acid ( A ) in hydrogenation with a thin and active plate.
( b : bulk, z: dimensionless length coordinate in the plate).
Reprinted with permission from Ref. 44. Copyright 1986, American Chemical Society.

radial dispersion and mass transfer. The axial dispersion will


also decrease due to the radial mixing.
Slug flow gives the best mass transfer properties. In this
flow the liquid slugs are separated by large bubbles with a
length which is larger than the channel diameter. The fluid
recirculates in the liquid slug, which increases the radial mass
transfer and decreases the axial dispersion. Between the gas
bubble and the wall there is only a very thin liquid f i l m . Mass
transfer from the gas bubble to the wall is hindered only by this
thin f i l m at the wall. Axial mass transfer between two successive
slugs can only occur through this thin film, which gives a very
low axial dispersion.
In annular flow the fluids are separated into a gas phase
along the tube in the core and a wavy liquid film at the walls.
There is no radial mixing in either flow and the liquid film is
much thicker than in slug flow. The annular flow gives the
monolith reactor very poor mass transfer and residence time
properties.

MONOLITHIC CATALYSTS

361

.. .. ... ...
..
. ...
.. ..
....
., ... ..

. ... .' .. ..
....
.....
.. .......
....

a.

0
Bubble
flow

FIG. 9.

Slug
flow

Taylor
flow

:a.

.. .. .. .. .
. . .. .. ..
....

......

.. . .. .. .
.. .. . .
:. . .
. *. . .'
*

Annular
flow

Flow patterns in vertical flow [ 4 5 ] .

Both bubble flow and slug flow are desirable in monolith reactors. Slug flow gives better mass transfer and residence time
but requires a higher gas hold-up.
The transition between the flow regimes is difficult to predict.
Most experiments are done with channel diameters larger than
1 cm.

Taitel et al. [ 4 5 ] give a qualitative discussion of the phenomena involved in the transition between flow patterns, which can
give some guidelines about where to expect the transition. The
bubble-slug transition requires an agglomeration of bubbles.
When the gaslliquid flow ratio is increased, the bubbles come
closer to each other and the coalescence rate increases. A theoretical upper limit of the gas hold-up is 0.52, which is the maximum allowable packing of spherical gas bubbles. However, the
gas bubbles move and collide randomly, and published data [ 4 5 1
indicate that at hold-up above 30% the bubble flow will irreversibly transform into slug flow. For void fractions less than 0.20,
coalescence is rarely observed. Only at high velocities are the
forces large enough to break the large bubbles into smaller ones.
Transition to annular flow occurs at high gas velocities.
Turner et al. [ 4 9 ] suggested that annular flow cannot exist un-

IRANDOUST AND ANDERSSON

362

less the gas velocity in the gas core is sufficient to lift the entrained droplets. When the gas rate is insufficient, the droplets
fall back, accumulate, and form a bridge, and slug flow takes
place.
Figure 10 shows the annular and slug transition regimes for
the system air-water in a 2-mm capillary tube. There is a wide
area in which the flow pattern is dependent on how the gas is
introduced into the fluid.
1.

Pressure Drop

The pressure drop of two-phase flow in monoliths is the sum


of frictional pressure drop, the static head, and orifice effects:
L

200

Transition

Transition

1 50

.
5
VI

>
k

Slug Flow

100

>

7
4
/%

50

Depends on
Method of
Liquid introduction

7
/

&

&
0
0

LIQUID VELOCITY cm/s

FIG. 10. Transition regime in a single capillary. Reprinted


with permission from Ref. 50. Copyright 1977, American Chemical Society.

MONOLITHIC CATALYSTS

363

Satterfield and &el [50] measured the pressure drop in monoliths and separated it into its components. They concluded that
only at high cell density (greater than 250 cells/in. 2 , did the
orifices have any effects. The frictional pressure drop and
static head accounted for more than 70% of the total pressure
drop in all monoliths tested.
At low velocity and narrow channels, the frictional pressure
drop is viscosity-dominant and can be calculated from the weighted sum of the liquid- and gas-phase pressure drops [ 131 :

Figure 11 shows the pressure drop measured by Satterfield


and Ozel [501.

0 4 mm beads Vp = 2.80 cmls


0 4 mm beads Vp = 5.22 cm/s
0 6 mm beads Vp = 2.80 cm/s
6 mm beads Va = 5.22 cm/s

Monoliths

Wp = 2.63 cmls

Vp = 5.22 cmls

~~

30

60

90

120

SUPERFICIAL GAS VELOCITY

150

180

cm/s

FIG. 11. Comparison of pressure drop in monoliths (200cell, 15.2-cm blocks) with that in packed beds. Reprinted with
permission from Ref. 50. Copyright 1977, American Chemical
Society.

IRANDOUST AND ANDERSSON

364
2.

Film Thickness

The film thickness of the continuous phase was studied at


an early date by Fairbrother and Stubbs [ 511, in 1935. Later
Taylor [52] studied this kind of flow, which is also called Taylor
flow or Taylor bubbles. These workers related the f i l m thickness
to the tube diameter ( d t ) , the viscosity ( u ) , the surface tension ( o) and the velocity of the fluid ( v ) .
-=a(...)"
6d

(33)

Bretherton [ 531 determined the parameters a and b to be 0 . 6 7


and 213 while Fairbrother and Stubbs [ 5 1 ] found 1 . 0 and 0 . 5
for ( p v ) 1 0 < 0 . 1 . Irandoust and Andersson [ 541 have recently
measured the film thickness spectrometrically
They summarized
their results in the empirical formula

Ti6

= 0.18

1-

exp ( - 3 . 1

(~y-~~)],

lJV < 2
U

(34)

to their data.
3.

Mass Transfer

Mass transfer between the continuous phase and the solid


wall has been studied by Horvath et al. 1551 and Hatziantoniou
and Andersson [ 561. The mass transfer in slug flow with recirculation within the slug was 3-10 times faster than with continuous liquid flow. Hatziantoniou and Andersson [ 561 summarized their results in the empirical formula

(7)

Sh = 3 . 5 1 Re Sc

o-44 B-OSo9

(35)

Irandoust et al. 1571 have recently studied gas-liquid mass


transfer in capillaries in a segmented gas-liquid flow. It was
shown that both the rear end of the gas plug and its cylindrical
part have the same importance for the gas-liquid mass transfer.
The influence of the liquid film on the gas-liquid mass transfer
was also considered.

MONOLITHIC CATALYSTS
4.

365

Axial Dispersion

Both Bretherton [ 531 and Pedersen and Horvath [ 581 used


residence-time distribution experiments to study the film thickness and the axial dispersion in the gas-liquid flow. The axial
dispersion between succeeding slugs occurs through the transport in the liquid film between the slugs. Usually the film is
very thin and the axial dispersion is very low. In most cases
the two-phase slug flow can be considered a s ideal plug flow.
In the bubble-flow regme the axial dispersion will be in the
intermediate range between continuous flow and ideal plug flow,
depending on bubble size.
111.

COMMERCIAL MONOLITHS
A.

Material

Monolithic support material is mainly used for emission control. Most of the development of monolithic material has been
directed toward high-temperature applications required in that
area. However, there is a growing interest from the chemical
industry to find applications for monoliths a s catalysts in chemical processes. These applications often require a high-surfacearea support.
Today monoliths are made of many different materials. Table
5 gives a list of monolithic materials that are available or under
development.
1.

Low- Surface- Area Monoliths


Ceramic monoliths for automotive emission control require

0
Thermal shock and thermal stress resistance, involving
a very low coefficeint of thermal expansion
0
Refractoriness
0
Porosity and pore size distribution for catalyst coating
application, that is, 30-40% open porosity and 5-15 p m medium
pore size
0
Axial crushing strength > 20 MN/m2

Cordierite meets these demands for standard automotive applications ; but in some applications, for example, heavy-duty
trucks and gas turbines, the low melting point (1735 K) limits
its application. However, refractory oxide materials have high

366

IRANDOUST AND ANDERSSON


TABLE 5
Materials of Commercial Monolithic Carriers

Name

Composition

a- and y-Alumina

A1 2

C or dierite

2Mg0 -2A1 20 -5SiO

Cordierite - mullite

2Mg0 -2A1 20 -5SiO2- 2A1 20 -2SiO

Magnesium aluminate- spinel

MgO-MgO *A12

Mullite

3A1 2O 4 i O

Mullite - aluminum titanate

3A1 2O -2SiO2-Al 20 .Ti0

Silica

SiO

Silicon carbide

Sic

Silicon nitride

Si 3N ,,

Spinel

MgO .A1 2

Titania

Ti0

Zeolites

A1 2O3-

Zirconia

ZrO

Zirconia- spinel

ZrO 2-Mg0 *A120

Metallic

Fe- C r -Al- Y t

03

03

03

SiO

thermal expansion and consequently low thermal shock resistance.


The thermal characteristics of some commercial monoliths are
given in Table 6.
The shock resistance can be increased by developing a
microcracked microstructure. Microcracks result from microstresses that arise after heat treatment. Anisotropic material or
phases with different thermal expansion can develop these stresses.
By adding aluminum titanate or zirconia to mullite, microcracks
are produced which increase the shock resistance but decrease
the mechanical strength [591.

MONOLITHIC CATALYSTS
2.

367

High- Surface-Area Monoliths

Application of monoliths in the chemical industry requires


monoliths with a high surface area, while the thermal stability
is less important. Table 7 gives a list of high-surface-area
monolithic materials that have been developed or are under
development for industrial applications. Most of the common
support material can now be manufactured in the shape of
monoliths. However, large high-surface-area monoliths are more
difficult to extrude than cordierite monoliths. Lachman [60, 611
has studied the effect of composition and heat treatment on surface area, porosity, and strength for some high-surface-area
monoliths.
3.

Metallic Monoliths

Metallic monoliths are made to withstand oxidating atmosphere


at high temperatures. Austenitic steels and nickel-chromium
alloys cannot meet these conditions. Aluminum containing ferritic steels produce a protective layer of aluminum oxide which
makes the alloy operative up to 1500 K 1111.
Metallic monoliths are usually made of the alloys Kanthal and
Fecralloy (Table 6 ) . Fecralloy contains iron, chromium, aluminum, and a small percentage of ytterium. Very thin foil sheets
(typically 50 p m ) can be made from these alloys. These thin
foils make it possible to obtain monoliths with a very high cell
density (Table 8 ) .
4.

Manufacturing

Ceramic monoliths have been manufactured by different techniques IS]. The two most frequently used methods are corrugation and extrusion.
In the corrugation method, a slurry of inorganic oxides and
salts with particles 1-5 p m in diameter is mixed w i t h organic
binders and plasticizers and impregnated onto a fabric or paper.
The sheet is often enforced with glass fibers or other inorganic
fibers. These sheets are then stacked or rolled together, with
every other sheet corrugated. The corrugated sheets are put
parallel to form conventional monoliths, or criss-crossed to form
cross-flow monoliths. The monolith is then fired at high temperature. Monoliths of almost any shape or size can be made by
the corrugation method. Heat exchangers with a diameter of
more than 5 m have been manufactured by this method.

++
++

+
+

Mullite

MulliteC

Cordierite

+++
++

+++

Cor dierite

Cordierite

++
++
++
++

++
++
++

Strength

+++

Shack
resistance

Cordieriteb

Alumina

Alumina

Alumina

Substrate
composition

1950

DuPont
Amer. Lava

Torvex
Thermacomb MD- 3

1430

Gen. Refr.

Ver sagrid

1400
1350

W. R . Grace

corning
Poramic

Celcor 9475

Amer. Lava

1200

1200

Thermacomb 795

1200

Amer. Lava

Thermacomb 614

1540

Coors

Dupont

Manufacturer

Torvex

Trade
name

1500

Upper temp.
limit, C

Thermal characteristics

Thermal Characteristics of Some Commercial Monoliths

TABLE 6

++
++
++
++
++

++*
+++

++
++

+
+

++
+

++

++
++

+++

+++

Zirconia- spinel

Zircon - mullite

Silicon carbide

Metallic

Metallic

Reprinted from Ref. 8 , with permission.


a+-fair, ++-good, +++-excellent.
bMagnesium aluminum silicate.
CAluminurn silicate.

Silicon nitride

Zirconia

Mullite

Mullite

Fecralloy

Kanthal
~~

SpectramicRX 384

SpectramicR X 3 8 7

Thermacomb 784

Celcor

Versagrid

Copyright 1982, Gulf Publishing Company.

1250

1400

1540

1650

1480

1700

2200

1700

1650

Johnson Matth.

K Y Metal

Norton

Norton

Amer. Lava

Corning

Corning

Cwrs

Gen. Refr.

IRANDOUST AND ANDERSSON

370

TABLE 7
High-Surface-Area Monolithic Material
Heat treat.
temp.,

Surface
area,
m /g

Porosity,
%

Median
pore
size,

Composition

y - Alumina

800

200

63

35

1300

80

59

140

1300

155

800

155

47

1300

60

60

800

50

63

1300

30

46

Transition alumina
93%Alumina + 7 %

silica
Silica

Bimodal

Magnesium
Aluminate spinel
Titania
Titania- silica

4 10

310

Metal monoliths are manufactured in a similar way. Very


thin metal f i l m s are rolled together, with every other sheet corru gate d.
Most monoliths are made by extrusion. In this process the
starting material is a very find powder to which plasticizers and
binders are added to form a plastic composition. This plastic
composition is then extruded through a special die to give a
monolithic structure. The monolith is carefully dried and then
fired at high temperature. Extruded monoliths can be made up
to 25 cm in diameter, but larger monoliths can then be made by
putting square-shaped monoliths together.
The manufacturing procedure has a great effect on the physical properties of the monoliths. The arrangements and sizes
of the crystalline and glass phases and the pore structure a s
well as the chemical composition determine the thermal expansion,
thermal conductivity, strength, melting point, and surface area.

MONOLITHIC CATALYSTS

37 1

TABLE 8
Comparison of Ceramic and Metal Substrate Parameters I: 111
Ceramic cell
density, cell /in.

Cell area,

Metal cell
density, cell/in

200

300

400

400

500

600

2.30

1.43

1.21

1.50

1.10

0.97

1890

2205

2790

3230

3580

3940

70

60

76

89

86

83

0.28

0.30

0.15

0.05

0.05

0.05

m2.1O6

Surface to volume
ratio, m-
Open area, %
Wall thickness,
m -103

5.

Washcoatin g

Monoliths are usually made with a surface area of only 0 . 1 - 1 . 0


m2/g. A large surface area is obtained by coating the monolith
surface with 5- 20 wt % of high-surface-area ( 50- 200 m 2 / g ) oxides.
This coating gives a surface area of 2 . 5 - 4 0 m2/g based on total
support weight. The desired properties of the washcoat are
high uniformity, high surface area, good adherence to the monolith, and good thermal stability.
There are three major ways of producing a high-surface-area
washcoat [ 51. One method is to dip the monolith into a slurry
of a fine powder of a high-surface-area oxide, resulting in a
washcoat after a heat treatment. A second method is to impregnate the monolith with a salt solution of the desired metal ion
and then heat the system to decompose the salt and form an
oxide. This impregnation can be repeated to obtain a washcoat
of desired thickness. The last method is to treat the monolith
with a metal salt and then precipitate the metal with a precipitating agent 1621. The produced solid, usually a hydroxide, is
subsequently heated to give an oxide.
The properties of the washcoat are dependent on the choice
of preparation method and the properties of the monolith. A
monolith w i t h large pores ( 5 - 1 5 pm) and open porosity in the
range 30-40% i s required to give good adherence. The thermal
stability of the washcoat is often improved by adding s m a l l
amounts of different oxides.

IRANDOUST AND ANDERSSON

372

B. Geometric Shape of Monolith Channels


The monoliths are manufactured in a wide range of sizes and
shapes. The cell shapes frequently produced are square, round,
and triangular.
Hegedus 1631 investigated the effect of the geometric shape
of the monolith channels on the performance of oxidative catalytic converters. He showed that significant saving in the weight
and size of the monolith can be obtained by proper selection of
channel shape, channel dimension, and feed flux. For a given
conversion, elongated rectangles gave the shortest monoliths.
Decreasing the hydraulic radius of a given geometric structure
could reduce the monolith weight without changing the pressure
drop essentially. Neither mechanical nor thermal stresses were
considered.
The effect of cell geometry on thermal shock resistance of
monoliths was studied by Gulati 1641. He showed that the cell
geometry has an important role in determining the monolith
characteristics. Triangular cell configuration requires a lower
thermal expansion coefficeint than square cell configuration, to
obtain equivalent shock resistance for a given monolith size and
surface area. A high cell density requires a lower thermal expansion coefficeint in order to keep the same thermal shock
resistance. The author concluded that the selection of cell geometry depends on various factors such as pressure drop, heat
and mass transfer, and conversion efficiency.
Howitt [ 121 has recently described various geometries for
monolithic catalyst supports. The effect of geometry on some
important properties such as pressure drop, thermal shock
resistance, mechanical strength, and catalytic activity was
studied. It was found that the cell density and web thickness
have a direct relationship to mechanical strength and light-off
time of the monolith.
C.

Selection of Monolith Substrate

Monoliths can be manufactured as either metallic or ceramic.


Each type has its advantages and limitations.
Thermal characteristics of monoliths are shown in Table 6.
The upper temperature limit ranges from 1520 K to 1670 K for
metallic and from 1470 K to 2470 K for ceramic monoliths.
Difference in thermal conductivity is of particular importance.
Metallic monoliths have about two orders of magnitude higher
thermal conductivity than ceramic monoliths. Ceramic monoliths

MONOLITHIC CATALYSTS

373

are almost adiabatic, while radial heat transfer in metallic


monoliths is of the same magnitude as in packed beds.
In Table 8 some important substrate parameters are compared
between ceramic and metal substrates. The higher surface-tovolume ratio of metallic monoliths provides higher reactivity per
unit volume for fast mass-transfer-limited reactions. This means
that a smaller metallic monolith could have the same activity a s
a larger ceramic monolith. In addition, the low thermal mass of
metallic monoliths makes them respond faster to temperature
changes. The higher open frontal area of the metallic monoliths
gives a lower pressure drop.
Both ceramic and metallic monoliths can be washcoated. The
metallic monolith is easier to regenerate. The washcoat can be
stripped without degrading the metal substrate considerably.
Flyt zani- Stephanopoulos et al. [ 381 have recently studied
radial heat transfer in metal honeycomb monoliths under nonadiabatic conditions. Heat transfer in the metal monolith was
compared to a packed bed of ceramic pellets. The main drawback of using packed beds of ceramic pellets is poor heat conduction and the resulting large radial and axial temperature
gradients within the bed. These kinds of temperature gradients are of vital importance for endothermic reactions where
external heating must be provided. For packed beds, the
radially flowing gas answers for radial heat transfer. The
authors conclude that the metal monoliths may be used to enhance
energy efficiency of existing reactors or heat exchangers by replacing a certain length of pellets at the inlet part of the bed
where the energy demand is greatest. They also emphasized
that the combination of the best heat-transfer features of both
the monolith and the packed bed of pellets may result in an
optimized design for catalyst systems.
According to Pratt and Cairns 1111. who studied the use of
metal catalysts on metallic substrates, the use of metal substrates is favorable for applications involving high-flow , hightemperature environments.
Applications involving very high temperatures and lor corrosive environments should employ ceramic monoliths, while when
mechanical and thermal shock are of greater interest, metallic
monoliths will be superior to ceramic ones.

IRANDOUST AND ANDERSSON

374
1V.

APPLICATIONS

The main features of the monolithic catalyst that determine its


applications are high surface-to-volume ratio, low pressure drop,
and uniform flow distribution. These characteristics have given
the monolith applications in areas with high reaction rate and
high volumetric flow rate. The monolithic construction provides
an attrition-free operation, reducing plugging and channeling.
New materials and geometries of the monoliths in the last few
years have broadened the applications. Metallic and cross-flow
monoliths have been tested when heat transfer is critical. Further high-surface monoliths made of y-alumina have been used
for slow reactions where high catalyst loading to reactor volume
is needed.
An analysis of monolithic catalyst reactor has been given by
Senkan et al. [ 331.
A.

Methanation

T h e methanation of carbon oxides is a rapid, exothermic reaction which makes the temperture control of the reactor difficult.
In such a case the use of metallic monoliths is very suitable,
especially when high reactant gas flow is desired.
Tucci and Thomson [65] have compared metal monolith and
pellet methanation catalysts based on equivalent metal loading.
The catalyst used was Ru in both cases. Pelleted catalyst showed
high initial activity which decreased slightly and finally regained
its original high activity. Monolith catalyst held a constant,
lower activity. Selectivity for converting CO to CHI,was about
97% for ruthenized monolith and about 83%for pelletized catalyst.
Jarvi et al. [661 have investigated the catalytic activity and
selectivity of nickel supported on washcoated ceramic monoliths
and pellets. The reactions were performed over a wide range of
temperture, pressure, and space velocity. It was shown that
washcoated monoliths provide less pore diffusional resistance
than pellet catalysts. This makes monolithic catalysts ideal for
intrinsic kinetic studies. The experimental results also showed
that, due to smaller pore diffusional resistance and consequently
high mass transfer rates, the monolithic catalyst used was 20100% more active than commercial catalyst pellets. The monolithic
catalysts were more active and selective under any given pressure and temperature conditions. Thus, the authors concluded
that monolithic nickel is the ideal catalyst for application in highthroughput recycle methanators.

MONOLITHIC CATALYSTS

375

De Bruijn et al. [671 have also investigated the use of monolith reactors for the methanation of carbon dioxide in hydrogen
at atmospheric pressure. The reaction temperature was between
473 K and 523 K . It was concluded that monolithic reactor systems have better temperature control and are very suitable for
large quantities of gas, which makes them appropriate within
SNG production.
Another study on this subject has been carried out by
Sughrue and Bartholomew [ 681 , who investigated the kinetics of
carbon monoxide methanation on nickel monolithic catalysts. It
was emphasized that the use of the coated monolithic catalysts
made it possible to determine kinetic data at higher temperatures
than was possible for a supported metal.
B . Hydrogenation and Dehydrogenation
1.

Liquid-Phase Reactions

Liquid-phase hydrogenations are common processes in both


large- and small-scale industry. The hydrogenations are usually
performed in the presence of a solid catalyst. The liquid-phase
hydrogenation involves mass transfer limitations in the gas-liquid
interphase, in the liquid film surrounding the catalyst particles,
and finally in the porous catalyst. Two conventional reactors
used in gas-liquid-solid processes are slurry reactors and tricklebed reactors. The main features of the monolithic reactor system are a combination of the advantages of the slurry and the
fixed-bed reactors, avoiding their disadvantages such as great
pressure drop, mass transfer limitation, filtration of the catalyst,
and mechanical stirring.
Liquid-phase hydrogenation of nitrobenzoic acid was performed
in a honeycomb monolith reactor system by Hatziantoniou and
Andersson [69]. The catalyst used was palladium on a ceramic
monolith honeycomb structure. The BET surface of the monolith
used was 1 4 . 7 m2/g, with a porosity of 59%. The catalyst contained 2 . 5 8 Pd. The total specific mass transfer area for gasliquid and gas-solid was 1200 m-. The reactions were performed
at 1 - 4 bar and 300-350 K . The flow through the monolith was
segmented gas-liquid flow in order to enhance mass transfer conditions. The authors also compared the monolithic catalyst reactor with a trickle bed for this hydrogenation reaction. The
effectiveness factor was shown to be much higher for the monolith than for a trickle-bed system under corresponding reaction
conditions. The advantages of the monolithic catalyst were short
diffusion length in the solid catalyst (0.15 mm in monolith used,

376

IRANDOUST AND ANDERSSON

compared to 2.5 mm for a typical trickle bed), good contact


between gas-solid and liquid-solid , uniform flow distribution, low
axial dispersion, and low pressure drop. All these properties
make the monolithic catalyst suitable for fast liquid-phase hydrogenations.
Hatziantoniou et al. [701 have recently studied mass transfer
and selectivity aspects of a honeycomb monolith catalyst reactor
compared to a slurry reactor, in liquid-phase hydrogenation.
The reactant mixture used was nitrobenzene and m-nitrotoluene.
The monolith had a BET-surface of 44 m2/g with a bed porosity
of 65%. The total BET-surface was increased to 80 m2/g by
covering the monolith with a washcoat. The Pd content of the
catalyst was 5.3% and the total volumetric mass transfer area
The reaction conditions were 5-10
of the catalyst was 1322 m-.
bar and 343-373 K . The gas-liquid flow in the monolith channels
was a segmented two-phase flow. It was found that the dominating transport step was the mass transfer of hydrogen from the
gas plugs to the channel wall of the monolith. The selectivity
of methyl aniline formation was higher for monolithic catalyst
compared to slurry catalyst. The authors explained this by the
influence of the mass transfer resistance.
The honeycomb monolith reactor is also used for hydrogenation
of anthraquinone in hydrogen peroxide production. The liquidphase hydrogenation of anthraquinone in a honeycomb monolith
res :tor is already an industria1 process 1711. The long-time
stability and the mass transfer characteristics of this reactor
system are reported to be excellent.
De Vos et al. 1391 have investigated liquid-phase hydrogenation of aqueous nitrobenzoic acid in a cross-flow monolith catalyst
reactor. The reactants, liquid and gas, were flowing in separate
channels perpendicular to each other. The porosity of the carrier was 65%. The BET-surface area was 44 m2/g, and the pore
volume was 0.66 cm3/g. The metal content of the catalyst was
7.1% Pd. The reaction conditions were 1-10 bar and 300-370 K .
The effectiveness factor was compared with a trickle-bed system
under corresonding conditions. The effectiveness factor was
found to be much higher for a cross-flow reactor system compared to a trickle-bed reactor. The main advantages of using
a cross-flow system are increasing the availability of the active
catalyst surface and reducing liquid film resistance.
The selectivity aspects of using a cross-flow monolith catalyst
reactor in liquid-phase hydrogenation of p-xylene, m-xylene, and
1.4-di-tert-butylbenzene in cyclohexane was studied by de V o s
et al. [43]. The results were compared to a slurry reactor

MONOLITHIC CATALYSTS

377

under corresponding conditions. The carrier was of the same


type as used by de Vos et al. 1391, with metal contents of 0 . 5 %
and 0 . 8 % Rh, respectively. The reactions were performed at
1 . 5 - 10 bar and 293-333 K . The particular concentration profile
in the cross-flow system was suggested to be the major factor
affecting the different product distribution obtained in reactions.
2.

Gas-Phase Reactions

Parmaliana et al. [ 9 ] have studied the catalytic activity of a


ceramic monolith-supported platinum catalyst for hydrogenation
of benzene and dehydrogenation of cyclohexane. The reactions
were performed at 523-623 K with contact times ranging from
10 to 1500 g h mol-l. Their result showed a good reproducibility and thermal stability for the monolith used. Long catalyst
lifetime without any isomerization products was observed. Optimal
conditions for conversion of reactants have been discussed.
Parmaliana et al. [72] have also investigated benzene hydrogenation on a ceramic monolith. The reaction conditions were in
the temperature range of 373-553 K with a contact time ranging
from 40 to 1750 g h mol-l. The kinetic behavior was studied
for different reactant compositions. The catalyst used contained
0 . 6 3 wt%nickel. It was concluded that the hydrogenation of
benzene to cyclohexane on nickel honeycomb gives a maximum
yield at 483 K at low P H , / P B ~ratio. The honeycomb catalyst
used showed constant reaction order with respect to the reactants
in a larger range of temperature than other conventional nickel
catalysts.
The investigation of benzene hydrogenation w a s extended to
Pt on honeycomb monoliths at both high 1731 and low I741 temperatures. The monolith with a BET-surface area of 2 8 . 3 2 m2/g
contained 0 . 2 4 % Pt. The reactions were studied at temperatures
ranging from 523 K to 623 K for high-temperature study and
373 K to 413 K for low-temperture study, respectively. The initial H , / B z ratio was 5 to 10. Their result showed that a singlesite adsorption reaction scheme of the Eley-Rideal type is adequate for both low and high temperature levels. The authors
gave notice of further investigations regarding the gas-solid interaction in monolith catalysts.
C.

Hydrotreatment

Hydrodesulfurization and hydrodenitrogenation are slow reactions and not suitable for standard monolith reactors. The stand-

IRANDOUST AND ANDERSSON

378

ard washcoated cordierite monolith has less than 5% porous support in the washcoat layer per volume reactor, while the usual
hydrotreatment in a trickle-bed reactor has more than 10 t i m e s
as much porous support per volume reactor.
Soni and Crynes [75] have performed these reactions on pure
y-alumina monoliths. These monoliths had 92 m2/g surface area
and were impregnated with 3.37% COO and 7.25% MOO3 . The
reactions were performed at 100 bar and 644 K. The activity
of the catalysts for hydrodesulfurization and hydrodenitrification
was about the same a s a commercial hydrotreatment catalyst
The catalyst was tested both on
(Nalcomo 747 from Katalco)
a high-boiling stock and on a low-boiling coal-derived liquid.
The authors conclude that the mass transfer , liquid distribution ,
and catalyst wetting are better in the monolith than in a trickle
bed. The monolith also minimizes bed plugging when fine solid
particles are present.

D.

Biochemical Reactions

The application of monolithic geometry is biochemical reactions, such as immobilization and precision separations , has
recently attracted attention from many authors. The honeycomb
support is in many applications superior to conventional supports. Advantages gained by the former are high surface-areato-volume ratio, low pressure drop , and even flow of liquid
medium as well as adequate oxygen diffusion. For systems w i t h
viscous liquid flow and high oxygen demand, this novel reactor
has proved to be very efficient.
In general packed-bed columns, fluidized-bed reactors, and
monolith reactors can be used as bioreactors. For fermentation
systems with gas evolution, the use of honeycomb monolith a s
a bioreactor is the most appropriate choice. The packed-bed
columns and fluidized-bed reactors are not recommended due to
channeling and plugging of liquid flow in the former and lower
conversion due to backmixing in the latter.
Ariga et al. [76] have recently studied the immobilization of
microorganisms on honeycomb monolith supports for biochemical
reactions with and without gas evolution. The effects of mass
transfer resistance and axial dispersion on the conversion of
substrates were studied. They found that the monolithic reactor had characteristics close to plug flow, that i s , minimum reactor volume for a given conversion of substrate, even for systems
w i t h gas evolution. The residence time distribution of the monolithic reactor was comparable to four completely mixed tanks in

MONOLITHIC CATALYSTS

379

series. This can be compared to a fluidized-bed reactor which


was comparable to 1.6 tanks in series. The effect of gas evolution on liquid film resistance in the monolithic reactor was also
studied. It was shown that at low superficial gas velocity the
gas bubbles may adhere to the surface, which decreases the effective surface area for the reaction.
The ceramic honeycomb structures are important in the largescale culture of animal cells. Lydersen et al. 1771 have recently
studied the use of ceramic matrix for this purpose. The authors
concluded that the high surface area per unit volume and the
special configuration of ceramic matrix provide an even distribution and growth of a wide variety of cells to densities equal to
or greater than obtained with other methods. This method
proved especially useful for scale-up studies when a scale-up
from 0.9 m 2 to 18.5 m 2 of surface area showed no losses in
efficiency of surface utilization. Other more conventional methods such as roller bottles, microcarriers, hollow fibers, or semipermeable capsules have shown several limitations. The limitations are due to problems in scale up, sensitivity of some cells
to the agitation, and the practical difficulty in separation. With
the ceramic matrix there is no need for either agitation or
separation of cells from the spent medium. It was also shown
that cells grown on the ceramic are readily harvested.
The monolithic reactor system has also been used for continuous saccharification of sol. starch by use of an immobilized glucoamylase monolith [ 781
Cross-flow monolith structures are also used as bioreactors
1791. The reaction mixture and a coolant flow in separate
channels prependicular to each other. This reactor controls
fermention heat more efficiently than a conventional bioreactor

Electrochemical Reactions

A novel solid-state electrocatalytic reactor is the cross-flow


monolith, which is used in fuel cells. Solid-electrolyte fuel cells
convert a significant portion of the Gibbs free energy change of
exothermic reactions into electricity. The reactor consists of
two sets of channels perpendicular to each other. One set of
these channels is coated with a catalyst such as Pt o r Ag and
serves for the oxidant flow where the reduction of O 2 to 0'takes place. These oxygen ions migrate through the solid electrolyte wall to the next set of channels, which serve for the fuel
flow. The w a l l s of these channels are coated with an appropriate
metal or conductive metal oxide catalyst, which catalyzes the

IRANDOUST AND ANDERSSON

380

anodic oxidation of the fuel. Thus the solid in this field of


application is used not only as a catalyst support, but also as
an electrolyte across which oxygen ion transport takes place.
Different kind of fuels can be used. Appropriate choice of
catalyst electrode materials depends on the nature of the fuel
used and the oxidation product. Pt is commonly used for ammonia conversion to NO, or for H , and CO combustion. and Ag
for ethylene epoxidation.
Vayenas et al. [801 have recently developed an analysis for
cross-flow monolithic designs in this area. They concluded that
the cross-flow geometry provides high 0-transfer area per unit
reactor volume. This leads to higher volume power densities
than in conventional designs, by a factor of 4 to 5 .
F.

Oxidation

Strongly exothermic catalytic reactions involving large quantities of gas such as oxidation reactions can take advantage of
certain properties of the monolithic catalyst such as low pressure
drop. The cross-flow monolith reactor can ideally be used for
oxidation reactions providing facile temperature control of the
catalytic surface. Furthermore, simple flow pattern and structure of the monolithic reactor make the scale-up process of this
reactor more reliable. Finally, better mechanical properties of
catalysts are of great value for this kind of reactions. One
limitation for monolithic reactors is observed for operations in
the laminar flow regime where mass transfer to the tube wall
limits the conversion.
The oxidation of ammonia to nitrogen was studied by Bernauer
et al. [ 2 8 1 , who found a monolithic reactor suitable for kinetic
investigation. The kinetics of this reaction were obtained by
use of data from a tubular wall reactor. They derived dimensionless equations which permit evaluation of temperature and
concentration profiles in the catalyst.
Boersma et al. [25] have compared the behavior of the empty
and packed tubular wall reactor both theoretically and experimentally using the air oxidation of ethene catalyzed by platinum.
Their experiments were applied to three different hydrodynamic
conditions: : laminar flow, turbulent flow, and the transition
region between.
Zabar and Sheintuch [811 have studied a monolithic reactor
for oxidation of SO, over platinum catalyst. The application of
cross-flow design was compared to two monolith blocks w i t h interstage cooling. They optimized the operating conditions for

MONOLITHIC CATALYSTS

381

this reaction in order to minimize expenses of catalyst and


energy required to achieve a certain degree of conversion. Industrially this reaction is carried out in a series of several
fixed-bed reactors with interstage cooling. The temperature
control in the cross-flow design of monolith is sufficient due to
large heat transfer area and close contact between the catalyst
and the coolant stream. It was shown that the cross-flow design
was superior to the design of two monolith blocks with interstage cooling. N o comparison between conventional fixed-bed
reactors and monolithic reactors was given.
G.

Catalytic Combustion

Catalytic combustion is a flameless combustion of various fuels


promoted by a catalyst. Catalytic combustion is used for production of heat from waste gases, flameless heaters, combustion
of organic fumes to avoid atmospheric pollutior., and direct power
production by, for example, gas turbine engines. Generally,
catalytic combustion is used to produce pollution-free energy.
The catalytic combustion can be initiated over a wide range of
fuel-air ratios and at lower temperatures than noncatalytic combustion. Combustion at lower temperature prevents the formation of pollutants such as nitric oxide.
The use of wood stoves as a source of heat has again become
common. Wood heat compared to costs for natural gas, oil, and
electricity is an attractive alternative. However, most wood
stoves have low efficiency because 114 to 112 of the usable
wood energy will be unburned. This unburned fraction in wood
smoke pollutes the air. The condensed smoke in the chimney
can be ignited, causing a fire danger. In such a case a catalytic
combustor can be fitted into the smoke passage, resulting in a
reduction of the temperature at which the smoke will burn.
Monoliths with very wide channels giving a low pressure drop
are ideal for this application,
Catalytic converters can also be used for gas turbines, where
they convert carbon monoxide and hydrocarbons in turbine exhaust to harmless carbon dioxide and water. The main advantage
of a catalytic gas turbine system is a lower level of pollution due
to lower combustion temperatures. A catalytic gas turbine consists essentially of three sections. The first section is a preheat burner where the combustion process starts. The next section is the catalyst unit. The gas stream enters the catalyst
unit having sufficient temperature to retain the combustion reaction. The catalysts are platinum-group metals supported by a

382

IRANDOUST AND ANDERSSON

stainless steel honeyconbed monolith. The advantages of using


metallic monoliths are high surface area, low pressure drop,
high mechanical resistance, high thermal shock resistance , and
even combustion due to a good heat conduction through the
monolith. Spent catalyst can be recovered and replaced on the
same support.
The third section is a thermal region where unreacted fuel
reacts by gas-phase free-radical chain processes. For maximum
efficiency of such a system, the pressure drop through the
catalyst section must be minimal.
Ablow and Wise 1821 have developed a theoretical model of
catalytic combustion which provides calculation of temperature
and fuel product distribution. The catalytic monolith effectiveness is generally mass transfer limited. This means that it is
desirable to increase the degree of conversion for reaction by
further homogeneous gas-phase reactions which are promoted by
the heat and intermediate species generated in the catalytic reaction. According to Ablow and Wise 1821 these gas-phase
reactions increase the fuel consumption but do not change the
temperature and species distributions.
Bensalem et al. [27] have recently developed a mathematical
model in order to study the behavior of a homogeneous-heterogeneous combustion in catalytic monoliths , applied to carbon
monoxide oxidation. The behavior of the catalytic monolith was
studied over a wide temperature range. The catalytic reaction
which occurs before ignition is generally kinetic limited. The
catalytic light-off is characterized by a sharp increase of the wall
temperature. In this region, the catalytic reaction is generally
mass transfer limited. The heat generated by the catalytic reaction ignites the homogeneous reaction, which improves the reactor efficiency. The authors have reported that for high inlet
temperatures, the contribution of the homogeneous reaction increases exponentially across the reactor and the catalyst is
needed only in the first part of the reactor. Conditions for
which both homogeneous and heterogeneous reactions are important were also investigated. The effect of inlet conditions on the
conversion was discussed. It was shown that for an intermediate
range of inlet temperatures from about 650 K to 1100 K , both
the homogeneous and heterogeneous reactions are important. The
reaction is mass transfer limited only in the inlet zone of the
reactor in this temperature range. The importance of the thermal
combustion rate increases along the tube. The authors suggest
a two-chamber reactor performance at high temperatures. The
first section is a small-diameter catalytic monolith concerned with

MONOLITHIC CATALYSTS

383

the mass-transfer-controlled surface reaction. The second


section is a noncatalytic section w i t h a larger diameter where
the bulk gas temperature is high enough to support homogeneous
combustion of the fuel. This reactor was reported to minimize
the pressure drop and shows a good efficiency w i t h lean fuelair mixtures.
Madgavkar et al. [ 8 3 , 841 have investigated the catalytic
combustion of low heat value (LHV) gases. For many operations in which the hydrocarbon content of LHV gases varies with
time, resulting in undesirable temperature variation of the exhaust gas, the authors showed that a stable catalytic combustion
can be achieved using oxygen-deficient atmosphere.
Enga and Thompson [85] have demonstrated the use of catalytic cornbustion in gas turbine technology. The activity of the
catalyst was tested by the light-off temperature. A decrease in
activity with each start-up test was observed. The authors
recommend the use of a metal support rather than a ceramic one,
due to the fact that no ceramic support could withstand more
than 20 start-ups.
The use of monolithic structures in catalytic combustion has
been reviewed by Tucci [8]. The author emphasizes the careful
choice of various parameters such as gas inlet temperature, residence time, adiabatic combustion temperature, fuel /air ratio,
combustor substrate configuration, and catalytic material. The
monolithic systems were recommended to achieve maximum catalytic combustion efficiency with high flexibility.
Other studies in this field have been carried out by Trimm
[ ? I , Harrison and Ernst 291, and Trevino and Sen 1861.
H.

Stationary-Source Emission Control

The emission of volatile organic compounds and carbon monoxide from fossil fuel combustion, gas turbine exhaust, petroleum,
and petrochemical processes results in grave environmental destruction.
These emissions are significant sources of energy and power.
Thus, combustion of flue gases results not only in pollution control, but also in enhancing energy recovery. The control of
environmental pollution has received much attention for well over
a decade.
The existing emisison control systems are generally based on
0
0
0

Incineration
Selective catalytic reduction (SRC)
Non-selective catalytic reduction (NSCR)

384

IRANDOUST AND ANDERSSON

Fume incineration is a thermal or catalytic oxidation of the


organic compounds present in the process exhaust. The products of oxidation are carbon dioxide and water. The thermal
oxidation requires an exhaust temperature of 923-1273 K and
fume retention of 0.3-1 8 . The catalytic oxidation, on the other
hand, requires a preheating temperature of 523-773 K with essentially no fume retention.
SCR is a technology to reduce NOx content of exhaust gases
by use of a reducing agent such as ammonia. NOx is reduced
selectively into N and water through a reaction between gasphase NOx and adsorbed NH,. The control of NOx emissions
by SCR technology does not affect the CO and HC emissions.
NSCR proceeds via a redox mechanism on the surface of a
catalyst. The catalyst surface is reduced by, for example, CO
and then oxidized by adsorbed NO,.
The catalyst surface can
also be oxidized by adsorbed O 2 when oxygen is present. In
such a case no NO reduction occurs,
x
In all these applications, the monolithic catalyst configuration
offers certain advantages. Since a catalytic reaction is a surface
phenomenon, the higher surface area per unit volume of monoliths reduces the catalyst volume. Furthermore, the low pressure drop in monoliths is important in emission control applications where the reactions are usually carried out at high temperatures and very high space velocities.
Kenson [ 8 7 , 881 has summarized the application and costeffectiveness of catalytic incineration compared to thermal incineration. The author showed that the catalytic incinerations are
the most cost-effective alternative in many cases.
Williams et al. [89] have recently investigated use of ceramic
honeycombs as supports for various catalysts for removal of
nitrogen oxides. They examined various catalysts such as V 205 ,
C r ,O 3 , Fe ,O 3 , COO, CuO, NiO, MOO 3, and WO on alumina and
titania honeycombs. The method used was NO reduction by SCR
with NH as reducing agent. Various reactant gas compositions
were used: N O - N H 3 , NO-NH 3 - 0 2 , NO-NH s-O ,-H ,O, and NON H 3-02-H
2 0 - S 0 2 . The authors concluded that for all catalysts
studied, O 2 enhances the reduction of NO with N H , and, further,
that the reduction fo NO to N , predominates over the oxidation
of N H 3 for most catalysts. Catalysts on Ti02 and MgAlOt, showed
the highest activity for all reaction conditions.
De Biasi [ g o ] , in an investigation of SCR catalysts, has recommended the use of metal monoliths as "an excellent choice" in
SCR applications. It was shown that stoichiometric amounts of
ammonia to NOx are necessary to provide optimal conditions for

MONOLITHIC CATALYSTS

385

the SCR method. Further, the removal of No increases with increasing temperature up to about 720 K , above which the r e moval rate of NO is decreased with increasing temperature. This
is probably due to oxidation of NH, by oxygen that occurs at
high temperatures.
Bartholomew 1911 has studied reduction of nitric oxides by
use of monolithic supported palladium-nickel and palladiumruthenium alloys. The effect of reaction parameters such as
temperature, space velocity, and C O / O p ratio on the reduction
of NO are reported.
Becker and Zygourakis [ 921 have recently considered the
design parameters for a catalytic monolithic reactor. The authors
developed a mathematical model which clarifies the effect of various design parameters on reactor light -off characteristics. The
parameters con sidered were nonuniform velocity distribution in
the honeycomb channels, thermal mass of the reactor, geometric
surface area of monolith, and stoichiometry of the exhaust gases.
It was found that reducing the thermal mass of the reactor, and
increasing exhaust gas temperature and the geometric surface
area, improved the light -off performance of the reactor. Moreover, the authors emphasized that a radially nonuniform velocity
profile may drastically increase the reactor light-off temperature.
The most important effect of heat losses proved to be a decrease
of steady-state conversion.
To sum up, the special performance of the catalyst system
depends on the type of contaminants in exhaust gases, desired
level of pollutant removal, exhaust temperature, and flow rates.
I.

Other Applications

The monoliths have also been successful in other applications


such as ozone abatement in jet aircraft passenger cabins, filtering molten metals, ultrafiltration, heat exchanging, and gasliquid contacting in absorption columns.
Lachman and McNally [93] and Lachman [so] have recently
summarized some industrial applciations of ceramic honeycombs.
V.

CONCLUSION

The development of new materials for monoliths has opened


many new fields of application for monoliths. Monolithic structures have been tested successfully within chemical processing
and refining industries as chemical catalysts.

386

IRANDOUST AND ANDERSSON

The main advantages of a monolithic substrate are due to


its large external surface and its low pressure drop. For fast
reactions when the process is mass transfer limited, the use of
monoliths has enhanced total effectiveness of the catalyst.
There are few published applications of monoliths for moderately fast reactions and slow reactions, where the reaction rate
per volume reactor is limited by the amount of the active material that could be introduced into the washcoat. We believe that
the introduction of new high-surface monolithic material will increase the number of applications in this area.
SYMBOLS
A
a
B
b
C

3
d

Da
De
DP
dt
f

g
(-AH)
h

kLa
L

L1
Nu

P
Pe
Pr
R

r
r
Re
sc
Sh

constant in E q s . ( 1 2 ) and (13)


parameter in E q . ( 3 3 )
constant in E q . (11)
parameter in E q . ( 3 3 )
molar concentration, mol/m
heat capacity at constant pressure, J / k g K
molecular diffusivity , m * / s
channel diameter, m
Dahmkiiler number
effective molecular diffusivity , m / s
particle diameter, m
tube diameter, m
friction factor
acceleration due to gravity, m / s 2
heat of reaction, Jlmol
thermal conductivity, W/m s
mass transfer coefficient, m / s
volumetric mass transfer coefficient, s - l
reactor length, m
length of liquid plug, m
Nusselt number
pressure, Pa
Peclet number
Prantl number
radius, m
radial coordinate, m
reaction rate, moI/m3 s
Reynolds number
Schmidt number
Sherwood number

MONOLITHIC CATALYSTS

t
T
V

387

time, s
temperature, K
linear fluid velocity, m / s
axial coordinate, m

Greek Letters
B
Y
6

5
E

x
!J
V

P
U

dimensionless length = Ll/dt


dimensionless length = L/dt
liquid film thickness, m
washcoat thickness, m
void fraction
thermal diffusivity = (h/pCD), m 2 / s
viscosity, kg/m s
stoichiometric coefficient
density, kglm
surface tension, k g / s

Subscripts
a

axial
benzene
cooling fluid
effective
fluid phase
frictional
gas phase
hydrogen
component number
reaction number
liquid phase
monolith
orifice
particle
solid phase
tube

wall

Bz
C

e
f
fr

H2
i
j
1

m
Or

P
S

ACKNOWLEDGMENT
We acknowledge useful discussions concerning this work with
D r . I . M. Lachman, Corning Glass Works.

IRANDOUST AND ANDERSON

388

REFERENCES

K . C Taylor, in Catalysis and Automotive Pollution


~- Control ( A . Crucq and A . Frennet, e d s . ) , Elsevier, A m s t e r dam, 1987, p. 97.
[ 2 ] K . Koberstein and G. Wannemacher, in Catalysis and
Automotive Pollution Control ( A . Crucq and A . Frennet,
e d s . ) , Elsevier, Amsterdam, 1987, p . 155.
t 31 J. S. Howitt, in Catalysis and Automotive Pollution Control
( A . Crucq and A . Frennet, e d s . ) , Elsevier, Amsterdam,
1987, p. 301.
R
. Prasad, L. A. Kennedy, and E. Ruckenstein, Catl.
[ 41
Rev.-Sci.
Eng., 26, 1 ( 1 9 8 4 ) .
[ 51 J. P. Deluca and L. E . Campbell, in Advanced Materials
in Catalysis ( J . J. Burton and R . L. Garten, e d s . ) ,
Academic Press, New York, 1977, p . 293.
[ 61 R . K . Shah and A . L . London, Tech. Rep. No. 7 5 , Dept.
Mech. Eng., Stanford University, Stanford, C A , 1971.
D . L. Trirnm, Appl. Catal., 7 , 249 ( 1 9 8 3 ) .
E . R . Tucci, Hydrocarbon Processing, 6 1 , 159 (March
1982).
[ 91 A . Parmaliana, C . Crisafulli, R . Maggiore, J. C. J. Bart,
and N . Giordano, React. Kinet. Catl. Lett., 18, 295
( 1981).
[ 101 L. L. Johnson, W . C . Johnson, and D. L . O'Brien, Chem.
Eng. Prog. Syrnp. Ser., 35, 55 ( 1 9 6 1 ) .
A . S . Pratt and J. A. Cairns, Platinum Metals Rev., 21,
74 ( 1 9 7 7 ) .
[ 121 J. S . Howitt, Paper 800082 Presented at SAE Automotive
Engineering Congress, Detroit, February 1980.
R . B . Bird, W . E . Stewart, and E. N . Lightfoot, Transport Phenomena, Wiley, New York, 1960.
A . H . P. Skelland, Diffusional Mass Transfer, Wiley, N e w
York, 1974.
[ 151 M . A. M. Boersma, W . H . M. Tielen, and H . S . Van D e r
Baan, ACS Syrnp. Ser., 6 5 , 72 ( 1 9 7 8 ) .
D. W . Wendland, Trans. ZSME, 102, 194 ( 1 9 8 0 ) .
G . I. Taylor, Proc. R . Soc., A L K 186 ( 1 9 5 3 ) .
G. 1 . Taylor, Proc. R . SOC., A223, 446 ( 1 9 5 4 ) .
G . F. Frornent and K . B . Bischoff, Chemical Reactor
Analysis and Design, Wiley, New York, 1979.
2 2 , 331
[ 201 L. C . Young and B . A . Finlayson, AIChE J . , ( 1976).
[ 11

MONOLITHIC CATALYSTS

1211

389

J. R . Sellars, M. Tribus, and J. S. Klein, Trans. ASME,

78, 441 (1956).


H.-R. Stock and A. Lowe, Chem. Eng. Sci., 38, 1039
(1983).
[ 231 J. Votruba, 0. Mikus, and K . Nguen, V. Hlavacek, and
J. Skrivanek, Chem. Eng. Sci., 30, 201 (1975).
1241 R. D. Hawthorn, Shell DevelpmentComp. Rept. P-2121,
Emeryville , USA.
[ 251 M . A . M . Boersma, J. A , M . Spierts, W. J. G . Van Lith,
and H. S. Van Der Baan, Chem. Eng. J . , 20, 177 (1980).
22,
1261 R. H . Heck, J. Wei, and J. R . Katzer, AIChE J . , 477 (1976).
1271 0. Bensalem and W . R . Ernst, Combust. Sci. Technol.,
29, 1 (1982).
[ 281 5 Bernauer, A . Simecek, and J . Vosolsobe, Collect.
Czech. Chem. Commun., 47, 2087 (1982).
[ 291 B. K . Harrison and W . R r E r n s t , Combust. Sci. Technol.,
19, 31 (1978).
L. Hegedus, AIChE J . , 21, 849 (1975).
[ 301
1311 S. T . Lee, PhD Thesis, University of Minnesota, 1977.
321 S. H. Oh and J. C. Cavendish, Ind. Eng. Chem. Prod.
Res. Dev., 21, 29 (1982).
L. B. Evans and J. B. Howard, Ind. Eng.
[ 331 S. M. Senkan:
Chem. Proc. Des. Dev., 18, 125 (1979).
33, 839
1 341 J. Sinkuk and V. Hlavacek, Chem. Eng. Sci., (1978).
[ 351 J. G . Stevens and E . N . Ziegler, Chem. Eng. Sci., 32,
385 (1977).
38, 733
1361 K . Zygourakis and R . Aris, Chem. Eng. Sci., (1983).
22, 343
[ 371 L. C. Young and B . A . Finlayson, AIChE J . , ( 1976).
1381 M Flyt zani- Stephanopoulos, G E Voecks , and T
Charng, Chem. Eng. Sci., 41, 1203 (1986).
and N . H . Schiitin, Chem.
1391 R. De Vos, V. Hatziantoniou:
Enp. Sci., 37, 1719 (1982).
T. F. D e g n E and J. Wei, AIChE J . , 25, 338 (1979).
T. F. Degnan and J. Wei, AIChE J . , 26, 60 (1980).
R. De Vos and C . E. H a m r m Chem. Eng. Sci., 5,
1711 (1982).
[ 431 R. De Vos, G. Smedler, and N . H. Schiitin, Ind. Eng.
Chem. Proc. Des. Dev., 25, 197 (1986).

1223

. .

IRANDOUST AND ANDERSSON

390

V. Hatziantoniou, B. Andersson, T. Larsson, N . H.


SchgCin, L. Carlsson, S. Schwarz, and K.-B. WidBen,
Ind. Eng. Chem. Proc. Des. Dev., 25, 143 (1986).
26,
1451 Y. Taitel, D. Bornea, and A . E. Dukler, AIChE J . , I441

345 (1980).

t 461
[ 471

D. Barena, 0. Shoham, Y . Taitel, and A. E . Dukler,


Int. J. Multiphase Flow, 6, 217 (1980).
R . C . Fernandes, R . Semiat, and A . E. Dukler, AIChE J . ,

29,
481
[ 491

501

981 (1983).

G. W. Govier, B . A. Radford, and J . S. C . Dunn, Can.


J. Chem. Eng., 58 (August 1957).
R. G . Turner, M . G . Hubbard, and A . E . Dukler, J.
Petroleum Tech., 21, 1475 (1969).
C. N. Satterfield E d F. Ozel, Ind. Eng. Chem. Fundam.,
16, 6 1 (1977).

Fairbrother and A . E. Stubbs, J . Chem. Soc., 1, 527

[ 511

( 1935).

G. I. Taylor, J. Fluid Mech., 10, 1 6 1 (1961).


F. P. Bretherton, J. Fluid Mech., 10, 166 (1961).
S. Irandoust and B . Andersson, Paper submitted for publication.
Ind.
551 C. Horvath, B . A. Solomon, and J.-M. Engasser, Eng. Chem. Fundam., 12, 431 (1973).
561 V . Hatziantoniou and B r A n d e r s s o n , Ind. Eng. Chem.
Fundam., 21, 451 (1982).
[ 571 S. IrandouG, S. Ertle, and B . Andersson, Paper submitted for publication.
[ 581 H . Pedersen and C . Horvath, Ind. Eng. Chem. Fundam.,

1521
[ 531
[ 541

20,

181 (1981).

I . M. Lachman and R . N . McNally, Ceram. Eng. Sci.


Proceed., 2, 337 (1981).
I601 I. M. Lachian, Sprechsaal, 119, 1116 (1986).
[611 I . M. Lachman, Presented a t t h e AIChE Chicago Annual
Meeting, November 1985.
621 R . L. Nelson, J. D. F. Ramsay, J. L. Woodhead, J . A .
Cairns, and J. A. A . Crossley, Thin Solid Films, 81, 329

[ 591

(1981).

[ 631

L. L. Hegedus, Am. Chem. Soc. Div. Pet. Chem. Prepr.,

18,
[ 641

487 (1973).
S. T. Gulati, Paper 750171 Presented at SAE Automotive
Engineering Congress, Detroit, February 1975.

[ 651

E. R

. Tucci and- W .

123 (February 1979).

J. Thomson, Hydrocarbon Processing,

MONOLITHIC CATALYSTS

391

G . A . Jarvi, K . B . Mayo, and C . H . Bartholomew, Chem.


Eng. Commun., 4 , 325 (1980).
E . W . De Bruijn, W. A. De Jong, and C . J. Van Der
Spiegel, ACS Symp. Ser., 65, 63 (1978).
E. L. Sughrue and C. H . Bartholomew, Appl. Catal., 2,
239 (1982).
V. Hatziantoniou and B. Andersson, Ind. Eng. Chem.
Fundam., 23, 82 (1984).
V. HatzianGniou, B . Andersson, and N . H . S c h g n , Ind.
Eng. Chem. Proc. Des. Dev., 25, 964 (1986).
T . Berglin and W. Herrmann, Pat. No. EP 102934 B1,
Sweden (1986).
A. Parmaliana, A. Mezzapica, C . Crisafulli, S. Galvango,
R . Maggiore, and N . Giordano, React. Kinet. Catal. Lett.,
19, 155 (1982).
A, Parmaliana, M. El Sawi, G . Mento, U. Fedele, and
N . Giordano, Appl. Catal., 7 , 2 2 1 (1983).
A . Parmaliana, M. El Sawi, fi. Fedele, G . Giordano, F.
Frusteri, G. Mento, and N . Giordano, Appl. Catal., 12,
49 (1984).
D. S. Soni and B . L. Crynes, ACS Symp. S e r . , 156, 207
( 1981).
0. Ariga, M . Kimura, M. Taya, and T. Kobayashi, J.
Ferment. Technol., 64, 327 (1986).
B . K . Lydersen, G . 2 . Pugh, M. S. Paris, B . P. Sharma,
and L . A. Noll, Bio/Technol., 63 (January 1985).
F. Shiraishi, K . Kawakami, and K . Kusunoki, Kagaku
Kogaku Ronbunshu, 12, 492 (1986).
H. Saito, Pat; No. 85256375 A 2 , Japan (1985).
C . G . Vayenas, P. G . Debenedetti, I. Yentekakis, and
L. L. Hegedus, Ind. Eng. Chem. Fundam., 24, 316
(1985).
E. Zabar and M. Sheintuch, Chem. Eng. Commun., 16,
313 (1982).
C . M. Ablow and H . Wise, Combust. Sci. Technol., 5.
35 (1979).
A . M. Madgavkar, R . F. Vogel, and H . E . Swift, Ind.
Eng. Chem. Prod. Res. Dev., 20, 628 (1981).
A . M. Madgavkar, R . F. VogelFand H . E. Swift, Ind.
Eng. Chem. Prod. Res. Dev., 20, 637 (1981).
B . E. Enga and D. T. Thompson, Platinum Metals Rev.,
23, 134 (1979).
Trevino and M. Sen, Chem. Eng. Sci., 41, 2253
(1986).

392

IRANDOUST AND ANDERSSON

f 871 R . E. Kenson, Chem. Eng. Prog., 81, 57 (1985).


[ 881 R . E. Kenson, Plant /Operations Prog., 2, 182 (1983).
[ 891

[ 901

[911
[ 921

931

J. Williams, I. Lachman, and T. Rosenbusch, Paper


presented at the 192nd National Meeting of the American
Chemical Society, Anaheim, 1986.
B . de Biasi, JOHNSON MATTHEY Cogeneration, 21 (JulyAugust 1984).
C . H . Bartholomew, Ind. Eng. Chem. Prod. Res. Dev.,
14, 29 (1975).
R . Becker and K . Zygourakis, Paper presented at
the Annual Meeting of the AIChE, Miami Beach, 1986.
I . M. Lachman and R . N . McNally, Chem. Eng. Prog.,
81, 29 (1985).
-

S-ar putea să vă placă și