Sunteți pe pagina 1din 9

Chemical Engineering and Processing 70 (2013) 7785

Contents lists available at SciVerse ScienceDirect

Chemical Engineering and Processing:


Process Intensication
journal homepage: www.elsevier.com/locate/cep

Characteristics of integrated micro packed bed reactor-heat


exchanger congurations in the direct synthesis of dimethyl ether
Fatemeh Hayer a , Hamidreza Bakhtiary-Davijany a , Rune Myrstad b , Anders Holmen a ,
Peter Pfeifer c , Hilde J. Venvik a,
a

Department of Chemical Engineering, Norwegian University of Science and Technology (NTNU), N-7491 Trondheim, Norway
SINTEF Materials and Chemistry, N-7465 Trondheim, Norway
c
Karlsruhe Institute of Technology (KIT), Institute for Micro Process Engineering (IMVT), Hermann-von-Helmholtz-Platz 1, 76344
Eggenstein-Leopoldshafen, Germany
b

a r t i c l e

i n f o

Article history:
Received 11 October 2012
Received in revised form 16 January 2013
Accepted 27 March 2013
Available online 16 May 2013
Keywords:
Dimethyl ether
Microstructured reactor
Mass transfer
Heat transfer
Residence time distribution
Pressure drop
Direct DME synthesis

a b s t r a c t
The performance of three integrated micro packed bed reactor-heat exchangers (IMPBRHEs) for direct
DME synthesis over physical mixtures of CuOZnOAl2 O3 and -Al2 O3 catalysts was experimentally
investigated. Systematic variations in reactor and slit dimensions and conguration were analyzed in
terms of thermal behaviour, mass transfer, pressure drop and residence time distribution (RTD). The
pressure drop was always small (<0.12 bar) relative to the total pressure (50 bar), and linear dependence
with GHSV conrms the predicted laminar ow for Re = 0.12. A narrow RTD was estimated by the dispersion analysis. Careful temperature measurements conrmed that the reaction temperature is mainly
controlled by the oil heat exchange to give a practically uniform temperature prole for set inlet oil temperatures of 220320 C. The micro packed beds were found free of the internal as well as external mass
transfer limitations, as showed by no signicant change in the CO conversion and DME yield for different
catalyst particle sizes, no effect of varying the linear gas velocity, and no effect of manipulating reactant diffusion coefcient. Packed bed microstructured reactors hence provide an isobaric and isothermal
environment free from transport limitations for the direct DME synthesis, in the kinetic regime as well
as at equilibrium conversion.
2013 Elsevier B.V. All rights reserved.

1. Introduction
The International Energy Agency predicts the worlds energy
consumption to increase by 49% from 2007 to 2035 [1]. Over this
period, the demand for petroleum fuels in the transportation sector is expected to grow by 1.3% annually. Global warming, local
emissions, energy security concerns and increasing cost of unconventional oil resources promote the development of processes
for production of liquid fuels from natural gas, biomass or coal.
Dimethyl ether (DME) is a volatile colorless gas, but non-carcinogen
and non-toxic and already in use as aerosol propellant [2]. The
high cetane number (5560) and low combustion exhaust emissions turn DME into a promising diesel substitute [35]. DME may
also be applied as liqueed petroleum gas (LPG) for heating and
home cooking [6,7], and can be reformed to produce hydrogen for
fuel cells or combustion [814].
DME is conventionally produced from methanol in a two step
process. Recently, direct DME synthesis from synthesis gas in a

Corresponding author. Tel.: +47 73592831; fax: +47 73594080.


E-mail address: hilde.venvik@chemeng.ntnu.no (H.J. Venvik).
0255-2701/$ see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cep.2013.03.021

single reactor has received attention from academia as well as


industrial companies [1517]. Combining the methanol synthesis
and methanol dehydration reactions shifts the methanol synthesis equilibrium towards higher CO conversion relative to methanol
synthesis alone. Alternatively, the process can be modied to operate for example at lower pressure. Moreover, the cost of utility
units between the methanol synthesis and methanol dehydration
steps is avoided. Synthesis gas conversion to form DME involves
the following reactions:
CO + 2H2 CH3 OH H0298 = 90.4kJ/mol

(1)

CO2 + 3H2 CH3 OH + H2 O H0298 = 49.4 kJ/mol

(2)

CO + H2 O CO2 + H2

H0298 = 41.0 kJ/mol

2CH3 OH CH3 OCH3 + H2 O H0298 = 23.0 kJ/mol

(3)
(4)

This process requires a dual function catalyst to form methanol


from syngas (1)(3) and convert methanol to DME (4). Since
methanol is not the thermodynamically favoured product from
H2 /CO/CO2 mixtures, a highly selective catalyst is required. Cu in

78

F. Hayer et al. / Chemical Engineering and Processing 70 (2013) 7785

combination with ZnO on a high surface area support is preferred


commercially due to cost and selectivity, but methanol is known
also to form over other metals, such as Pd [18]. Methanol dehydration occurs over solid acid catalysts, of which -Al2 O3 and zeolite
HZSM-5 are the most studied and applied, but silicoaluminophosphate (SAPO), zeolite HY and ferrierite are also active [1921]. The
overall reaction is highly exothermic, and control of the temperature is therefore imperative to the reactor design. Two different
reactor types are used in commercial, direct DME synthesis technology; xed bed [10] and slurry phase [11].
The increase in DME demands along with the possibility to
use renewable (biomass, agricultural by-products and wastes from
pulp and paper mills) and fossil (natural gas and coal) feedstock,
encourages intensication of the DME production process. Because
the existing technology generally prots from economy of scale,
utilization of biomass as well as remote/associated natural gas calls
for efcient processes in the small to medium scale. The assessment of microstructured reactors could be key to intensication of
the direct DME synthesis and the development of technology for
commercial production of DME fuel.
Miniaturization of chemical reactors is known to add advantages such as enhanced mass and heat transfer due to high surface
to volume ratio and short characteristic dimensions. Microchannel
reactors with feature sizes in the micrometer range are thus suited
for both highly exothermic (FischerTropsch, methanol and DME
synthesis) or endothermic (methane reforming) reactions [2224].
In addition, the small volumes and quantities of chemicals in the
microreactors make it possible to control the process even under
failed operation (run-away, leak, etc.). Hence, inherent safety characteristics exist that lead to application in high-risk processes (e.g.
cyanides, peroxides, acids) [25] or high risk environments (oating
vessels, agricultural areas). Ajmera et al. [26] used a silicon micro
packed bed for phosgene synthesis to demonstrate the application
of a microchannel reactor in a hazardous process. Microchannel
reactors have signicantly forwarded chemical syntheses that were
previously limited to batch processes and associated with complicated reactor scaling, by conversion to continuous operation and
elimination of reactor scaling phenomena associated with transfer
from the laboratory to the plant [27]. Introduction of this technology to bulk chemicals and fuels production is more challenging,
since the investments are large and the overall process efciency
often improves with scale. Microreactors, however, may in the
future signicantly reduce process footprint as well as the time
and cost involved in taking a process all the way to commercial
operation [22,28,29].
Different approaches exist on how to introduce catalyst to micro
channels/slits/volumes. In the case of gassolid heterogeneous
reactions packed bed microreactors and catalytic wall microreactors may be envisaged, as well as congurations with features that
combine the two, such as inserts made from coated material. The
latter has been termed catalytic bed microreactors [30]. Among
them, micropacked bed reactors offer the easiest way to load and
unload the catalyst as well as the option of direct utilization of existing, optimized catalysts. Their possible disadvantages may include
high pressure drop and maldistribution between reaction channels.
In the present work, a detailed analysis of different microstructured packed bed reactors with integrated heat exchange channels
in the direct DME synthesis is made in order to study the effect
of varying reaction slit cross section and the number of reaction
slits as well as differences in the internal geometry of the slit (pillar
structure).

Table 1
Reaction slit geometry and dimensions of the three different micro packed bed
reactors (A, B and C).

Number of reaction slits


Slit cross section [mm2 ]
Slit length [mm]
Number of pillars in one slit
Surface to volume ratio

Reactor A

Reactor B

Reactor C

1
8.8 1.5
6
26
1731

8
8.8 0.8
6
26
2897

8
8 0.8
6
192
4963

Process Engineering (IMVT) at the Karlsruhe Institute of Technology (KIT) in Germany. These reactors are made of stainless steel
and can operate at pressures up to 100 bar and temperatures up to
350 C [3136]. The design and fabrication method, i.e. etching of
channels into rectangular, stainless steel foils, allows varying the
number of channels in a reactor module (Table 1). The microreactors consist of rectangular reaction slits sandwiched between
cross ow channels for heat transfer by oil circulation (Fig. 1a),
and the resulting, diffusion bonded stack is sealed to a housing
with inlets and outlets for each channel direction. The oil channels
with 0.25 mm 0.5 mm cross section and 0.25 mm ns between
the channels were made from 0.5 mm thick stainless steel foils.

2. Materials and methods


Three different integrated micro packed bed reactor-heat
exchangers (IMPBRHE) were fabricated by the Institute of Micro

Fig. 1. Illustration of (a) the stacking of reaction and heat exchange oil slits in the
three different IMPBRHEs, (b) internal pillar structure inside the reaction slits of
reactors A and B, and (c) pillar structure inside the slits of reactor C.

F. Hayer et al. / Chemical Engineering and Processing 70 (2013) 7785


0,12

0,10

Microreactor B
Microreactor A
Microreactor C

0,08

P, [bar]

The reaction slits were also etched from 1 mm thick foils to 0.4 mm
depth while including hexagonally arranged cylindrical pillars with
0.8 mm diameter and 0.8 mm distance from center to center to
increase the reaction slit surface. Face to face bonding of two such
foils gives the high aspect ratio reaction slit cross sections depicted
in Table 1. The differences between the three micro packed bed
reactors is found in the number of reaction slits, reaction slit cross
section and the number of internal pillars in the slits (Table 1).
Fig. 1b and c schematically illustrates the pillar structure inside the
reaction slits. Whereas reactor C contains pillars throughout its 8
slits, reactor A and B have three rows of pillars at the inlet and outlet only of the same length as the slit height, 1.5 or 0.8 mm. Note
also that the pillar volume was always subtracted from the total
slit volume for GHSV (Nml/gcat/min) calculations to obtain a true
reaction volume.
The heat exchange oil (Thermal H350 Heat Transfer Oil, Marlotherm SH) from a high temperature thermostat (Julabo HT30-M1)
with 18 l/min maximum ow rate, ows through the cooling channels and removes the heat of reaction at a temperature gradient
between inlet and outlet of less than 1 K. The set point is adjusted
to approach a desired temperature inside the reaction slit. Good
external insulation of the reactor module minimizes the heat losses
and stabilizes reactor housing (skin) temperatures. Three holes
of 0.5 mm diameter along the centerline of the reaction zone on
top (and also three holes mirrored on the bottom) of the outer
surface of the reactors was used for the skin temperature measurements, providing close proximity to the cooling channels (1 mm)
and reaction slits (2 mm). These holes were also applied to obtain
the temperature prole along the reactor during the experiment.
In addition, there is a missing row of pillars in the middle reaction slit of reactor C to allow insertion of a thermocouple and
temperature monitoring inside the catalytic bed. Two additional
thermocouples were placed in the inlet and outlet of the oil channels.
The reaction slits were packed by gently shaking a 2:1 mass
ratio physical mixture of a commercial methanol synthesis catalyst (CuOZnOAl2 O3 ) and a commercial -Al2 O3 as methanol
dehydration catalyst. A sieved fraction in the range 5080 m was
applied unless otherwise specied. The catalyst particles were kept
in place by two frits (porous metallic plates) at the inlet and outlet of the slits. No practical difference was experienced during
catalyst loading between the three reactor congurations studied.
The catalysts were reduced with 3% H2 in N2 at 250 C and
atmospheric pressure. Premixed synthesis gas with a composition
of H2 :CO:CO2 :N2 :CH4 = 56:28:5:5:6 (mol%) was preheated close to
the oil circulation temperature and fed from the top of the reactor. The syngas was initially passed through a purifying trap to
eliminate possible carbonyls, and the molar ow rate was controlled by a mass ow controller (Bronkhorst). The pressure was
regulated by a backpressure controller (Bronkhorst). The reaction
experiments were conducted at temperatures from 220 to 320 C
and pressures 5070 bar. The pressure difference between inlet
and outlet of reactors could be measured by a differential pressure
transmitter (APLISENS, APR-2000) with an error of 0.0001 bar.
Maximum allowed operating temperature of the sensor is 120 C.
The pressure drop investigations were therefore carried out at low
temperature in a ow of N2 .
The feed and product gases were analyzed by an online
Agilent 7890 gas chromatograph (GC) equipped with a thermal conductivity detector (TCD) and a ame ionization detector
(FID). To avoid any condensation of products, the outlet gas
lines were heated (180 C) until entering the GC. N2 (TCD) and
CH4 (FID) were used as internal standards for the analysis.
The accuracy of the carbon mass balance was always within
5%

79

0,06

0,04

0,02

0,00
0

5000

10000

15000

20000

25000

-1

GHSV,[hr ]
Fig. 2. Measured pressure drop as function of GHSV [h1 ] over the micro packed
bed reactors at 50 bar and 20 C under pure N2 ow.

The CO conversion (XCO ), and DME yield (YDME ) were dened as


follows:
XCO =

nCO,in nCO,out
100%
nCO,in

YDME =

2nDME
100%
nCO,in + nCO2 ,in

(5)

(6)

3. Results and discussion


3.1. Pressure drop
Between the three micro packed bed reactors, the main difference resides in the ow through catalyst lled reaction slits. The
inuence of hydrodynamics is studied in terms of pressure drop and
ow distribution. Decrease of particle size and hydraulic diameter
may increase the resistance to gas ow; i.e. while packed beds with
characteristic dimensions in the m range provide better contact
between the gas and the catalyst surface, they have been claimed
to suffer from high pressure drop [37].
Fig. 2 shows the measured pressure drop versus GHSV (ow
of N2 ) for the three micro packed bed reactors with the reaction
slit arrangements as depicted in Table 1 and Fig. 1. The pressure
drop shows linear dependence with GHSV as expected for laminar
ow, and is always small (<0.12 bar) relative to the total pressure
(50 bar). The measured pressure drop is highest in reactor B, which
has 0.8 mm high slits and pillars at the inlet and outlet only (reactor
B). The single slit reactor (A), which has somewhat larger slit height
(1.5 mm) but otherwise similar slit geometry, shows considerably
(50%) lower pressure drop in repeated measurements. This is not
easily explained since the numbering up of nearly identically structured reaction slits is not expected to increase pressure drop versus
GHSV signicantly. The dimensions and the aspect ratio of the slit
cross section, i.e. the effective hydraulic diameter, could play a role
to ow restrictions. On the other hand, the packing density should
be facilitated by increased rather than decreased dimensions. The
pillar structure throughout the channels (reactor C) instead of at
inlet and outlet only (reactor B) reduces the pressure drop considerably. The only difference between reactors B and C is the number of
pillars, since they both have 8 channels of similar dimension. Since
the bed volume of the pillar structure is accounted for, it appears
that it causes less dense packing and smaller ow restrictions.

80

F. Hayer et al. / Chemical Engineering and Processing 70 (2013) 7785

100

0,25
DME catalyst (50-80 m)
Inert (50-80 m)
DME catalyst (125-200 m)
Ergun equation (50-80 m)

Dp=50-80 m
Dp=80-125 m
Dp=125-200 m
Equilibrium

80
CO conversion, [%]

P, [bar]

0,20

0,15

0,10

60

40

0,05

20

0,00
0

10000

20000

30000

40000

50000

60000

0
200

70000

220

GHSV, [hr-1]

(7)

where P denotes the pressure gradient over a bed of length L


for a gas of density g and supercial velocity us . dp is the catalyst
diameter, for which an average value was considered for particles in
range of 5080 m. The total energy loss f is considered as the sum
of the viscous energy loss expressed by the BlakeKronzey equation
[39] for laminar ow, and the Burke-Plummer equation [40] for the
kinetic energy loss in highly turbulent ow, respectively:
150 (1 )2
(1 )
+ 1.75
Re
3
3

1+

dh /dp 2
dh /dp

320

(8)

3.2. Mass transfer


The transport of reactants from the gas phase to the catalyst surface and products from the surface to the bulk of the gas
may affect reactor performance. External and intraparticle transport limitations must be assessed in different ways. A limitation
by internal rather than external mass transfer is likely for largescale xed beds, whereas particle dimensions in the micron range
are expected to eliminate internal mass transfer limitations [43].
Possible internal limitations were nevertheless checked by applying particle sizes of 5080 m, 80125 m and 125200 m in
reactor A. Figs. 4 and 5 present the CO conversion against temperature and GHSV, respectively. There are no signicant differences in
observed activity for the three particle sizes except a slight decrease
in CO conversion for the largest particle size in Fig. 5, possibly
within experimental error. We thus conclude that the system is
free of internal mass limitations, at least for particles smaller than
125 m.
In a packed bed, the solid catalyst particles remain in a xed
position and the uid ows through void spaces between the particles. In the bulk, reactants and products are mixed and transported
at rates that depend on the nature of the ow. Conventionally,

100

2 

80

(9)

The calculated pressure drop for 5080 m particles lies


between the corresponding sieved fractions of the hybrid catalyst
and SiC. This shows the importance of particle morphology to the
ow characteristics of micro packed beds [42]. A uniformly packed
bed of particles of similar size distribution should provide a uniform
pressure drop throughout the catalyst bed. However, in addition
to deviations expected in the size distribution between different
materials from the same sieve fraction, the particles have different
shape and agglomeration properties that may signicantly affect
the void fraction of the catalyst bed and consequently the pressure
drop. Particle agglomeration and bed densication was experienced to a larger extent by using HZSM-5 as dehydration catalyst
in the hybrid mixture instead of -Al2 O3 .

CO conversion, [%]

= 0.38 + .073

300

Fig. 4. CO conversion as function of reaction temperature in reactor A for different


catalyst particle size range at 50 bar and GHSV = 7500 Nml/gcat/h.

The porosity, , was calculated using the correlation developed


by Haughey and Beveridge [41] for a packed bed of spheres and
reaction slit hydraulic diameter dh :

280

Temperature, [ C]

Fig. 3 displays the pressure drop in the single slit reactor (A)
for different catalyst particle size, as well as with the catalyst
replaced by inert (SiC) material. The pressure drop is diminished
by increasing the particle size from 5080 m to 125200 m. It is
also considerably smaller for 5080 m SiC particles relative to the
mixture of CuOZnOAl2 O3 and -Al2 O3 particles in the same size
range. Fig. 3 also shows the pressure drop as calculated by the Ergun
equation, which has been veried for 101 < Re < 103 and therefore
can be applied for both laminar and turbulent ow regimes [38]:

f =

260
o

Fig. 3. Calculated and measured pressure drop as function of GHSV [h1 ] over reactor A for varying catalyst particle size range and bed material at 50 bar and 20 C
under pure N2 ow.

g u2s
P
=f
L
dp

240

Dp=50-80 m
Dp=80-125 m
Dp=125-200 m

60

40

20

0
0

10000

20000

30000

40000

50000

GHSV, [Nml/gcat/h]
Fig. 5. CO conversion as function of GHSV in reactor A for different catalyst particle
size range at 50 bar and 250 C reaction temperature.

F. Hayer et al. / Chemical Engineering and Processing 70 (2013) 7785

80

YDME , [%]

multi-component gas mixtures. A simple relation that may be


applied is called Blances law [48]:

100

Microreactor A
Microreactor B
Microreactor C
Equilibrium

60

60

40

40

20

20

0
0

10000

20000

30000

40000

80

CO conversion, [%]

100

0
50000

GHSV, [Nml/min/gcat]
Fig. 6. Comparison of (a) CO conversion (b) yield of DME as function of GHSV in the
three micro packed bed reactors at 50 bar and 250 C reaction temperature.

perfect mixing is assumed in the bulk, and the concept of a socalled stagnant lm has been assigned to the region close to the
outer catalyst particle surface, through which the mass transfer
is proportional to the molecular diffusion coefcient. The mass
transfer in micro packed beds may be described by similar fundamentals as in conventional xed bed reactors, but the ow through
the microchannels is generally laminar and it has been questioned
whether the stagnant lm model is applicable [44].
Different methods exist to study possible effects of external
mass transfer on packed bed performance, and the method should
not affect the reaction kinetics. Given that the stagnant lm concept can be applied, the linear gas velocity is claimed to affect the
thickness of the lm and hence the transfer between the bulk of
the gas and the catalyst surface [45]. By varying the reactor diameter, the linear velocity can be changed while keeping the residence
time and other parameters constant.
Fig. 6 shows the CO conversion and DME yield against GHSV
at 250 C for the three different reactors. The reactor performance
remains the same given the different linear gas velocities for the
different reaction slit cross sections (6.4, 7.04 and 13.2 mm2 ). Under
the assumption that mass transfer is affected by the linear velocity,
the reaction slits may be claimed free of external mass transfer
limitations. As expected, the CO conversion decreases as the contact
time decreases, with the DME formation in the consecutive reaction
even more affected by GHSV.
Another possibility for assessing external mass transfer in
packed beds is the variation of the molecular diffusion coefcient.
The molecular diffusion coefcient (Dij ) of reactants at high pressure can be calculated by the Fuller equation (10) [46], which is
a modication of ChapmanEnskog theory [47]. The temperature
power dependence of 1.75 in the Fuller equation represents a reasonable agreement with measured values for most binary systems,
whereas ChapmanEnskog suffers from larger deviations between
theory and experiment with respect to the temperature power
dependence.
0.001 T 1.75
Dij =
P


1/3
i


+

1
Mi

1
Mj

1/2


1/3 2

81

(10)

j

where T is temperature (K), P is pressure (atm), M is molecular mass (g) and  is diffusion volume (atomic volume). There
exist a few approaches to the calculation of molecular diffusion in

Dim =

n

xj

j=1

1

Dij

(11)

j=
/ i
where the diffusion coefcient depends on the gas composition,
and this represents an opportunity to manipulate mass transfer
rates without affecting the kinetics of the reaction, by keeping
temperature, reactant contact time and partial pressures constant.
Walter et al. [45] changed the inert gas for a rst order reaction and
found no signicant change in conversion at low reaction rate. At
higher reaction rate, they observed a difference in conversion that
correlated with predicted differences in diffusion coefcients, and
concluded that the reactant turnover was controlled by the mass
transfer.
We adopted this method to analyze the external mass transfer
by adding helium (He) to the syngas mixture. The partial pressure
and contact time of reactants was kept constant by increasing the
pressure from 50 to 63.5. In this manner, the diffusion coefcient of
the main reactant CO was decreased by 20%. Fig. 7 shows that the CO
conversion and yield of DME are not affected by changes in molecular diffusion for any of the three micro packed bed reaction slit
congurations. This approach is complementary to the assessment
by changing the linear velocity, irrespective of the applicability of
the stagnant lm concept. The results support the conclusion that
the external mass transfer is not limiting the reactor performance
for the DME synthesis process under the conditions studied.
3.3. Thermal behavior
The direct DME synthesis is highly exothermic. The adiabatic
temperature rise for a syngas of composition as applied here,
250 C feed temperature and 50 bar pressure is calculated to 148 C.
Hence, temperature control is crucial to maintain a favorable
equilibrium conversion and good catalyst stability. As mentioned,
microchannel reactors provide high surface to volume ratios and
excellent integration between reaction and heat removal channels. Several studies have shown that most microchannel reactors,
including micro packed beds, exhibit uniform wall temperature
[31,32,35,49].
The effects of varying the oil temperature in the range
220320 C, as well as gas ow rate on the measured temperature
difference between the heat exchange oil and inside the middle
reaction slit in reactor C are shown in Figs. 8 and 9, respectively.
Fig. 8 shows that the reaction slit temperature is mainly controlled
by the oil heat exchange. The measurements reveal a maximum
difference of 1 C under the reaction conditions applied. Moreover,
the difference within inlet and outlet oil temperatures was always
within 1 C. The maximum deviation is systematically found at low
GHSV and corresponding high conversion (Fig. 9). Fig. 10 presents
measured reactor skin temperatures of the three reactors obtained
at equivalent reaction conditions in close proximity to the reaction slits as explained in the experimental section. The deviation of
the measured skin temperature to the set inlet oil temperature is
always less than 3 C. Hence, the bed temperature prole is practically isothermal, and the controlling factor is the temperature of the
oil. The reaction slit wall temperature may therefore be assumed
constant. We have previously simulated the effect of varying slit
height under the assumption that the heat transfer is sufcient
to maintain a constant temperature at the two walls adjacent to
the heat exchange channels [35]. It was found that in the range
0.83 mm, although the bed temperature gradient varies consistently with the slit dimension, in remains small (<4 C). Although

82

F. Hayer et al. / Chemical Engineering and Processing 70 (2013) 7785


100

320

80

80

300

60

60

40

40

20

20

100

YDME, [%]

CO conversion, [%]

Reaction slit temperature, [ C]

Reactor A

syngas at 50 bar
Diluted syngas with He at 63.5 bar

280

260

240

220

100
0

100
0

Reactor B

200

80

200

80

220

240

260

280

300

320

60

60

40

40

YDME, [%]

251,3
20

20

251,2

Syngas at 50 bar
Diluted syngas with He at 63.5 bar
0
100

0
100

80

80

60

60

40

40

20

20

YDME, [%]

Reactor C

CO conversion, [%]

Fig. 8. Reaction slit temperature inside reactor C versus inlet oil temperature at
50 bar and GHSV = 7500 Nml/gcat/h.

Reaction slit temperature, [ C]

CO conversion, [%]

Inlet oil temperature, [oC]

251,1

251,0

250,9

250,8

250,7
0

10000

20000

Syngas at 50 bar
Diluted syngas with He at 63.5 bar
0
240

250

260

270

280

40000

50000

GHSV, [Nml/gcat/h]

0
230

30000

290

Fig. 9. Reaction slit temperature inside reactor C versus GHSV at 50 bar and
Toil = 250 C.

Temperature, [oC]
Fig. 7. CO conversion and yield of DME versus reaction temperature in the three
micropacked bed reactors at GHSV = 7500 Nml/gcat/h for undiluted (50 bar) and He
diluted (63.5 bar) syngas. Reactant partial pressure and contact time was maintained
constant.

280
270

Microreactor A
Microreactor B
Microreactor C

Skin temeprature, [ C]

Fig. 10 shows temperatures obtained in close proximity to the channels rather than inside, it supports this conclusion since the three
reactors display practically identical temperature proles.
Fig. 11 compares the performance of the reactors in terms of CO
conversion (a) and yield of DME (b) versus temperature at GHSV. In
general, increased reaction temperature increases the syngas conversion and yield of DME in the kinetic regime [34]. A maximum CO
conversion corresponding to near equilibrium for the DME synthesis reaction could be achieved in all three reactors. Naturally, the
experimental parameters determine at which temperature equilibrium conversion is reached, but here it particularly reects the
somewhat poor kinetics over the -Al2 O3 dehydration catalyst at
temperatures below 280 C. The results also demonstrate, provided
that both catalyst functions can be maintained stable and the bed
temperature can be controlled, how the direct synthesis can be
intensied by somewhat elevated reaction temperatures since the
methanol equilibrium is completely alleviated.

290

260
250
240
230
220
210
0

Thermocouple number from top


Fig. 10. Reactor skin temperature for the three micro packed bed reactors at
Toil = 220, 250 and 280 C, 50 bar and GHSV = 7500 Nml/gcat/h.

F. Hayer et al. / Chemical Engineering and Processing 70 (2013) 7785

83

100

(a)

CO conversion, [%]

80

60

Microreactor A
Microreactor B
Microreactor C
xCO,eq (DME)

40

xCO,eq (MeOH)

20

0
200

220

240

260

280

300

320

340

Temperature, [ C]
60

(b)
50

Fig. 12. Axial and radial Peclet number for packed beds (IUPAC).
Adopted from Wilhelm [51].
Microreactor A
Microreactor B
Microreactor C

YDME, [%]

40

30

20

10

0
200

220

240

260

280

300

320

340

Temperature, [ C]
Fig. 11. Comparison of (a) CO conversion (b) yield of DME as function of
reaction temperature in the three micro packed bed reactors at 50 bar and
GHSV = 7500 Nml/gcat/h.

A slight difference in conversion between the three reactors


can be observed in Fig. 11, particularly at high temperature, and
a similar behavior is found for the yield of DME. Since we are
convinced that all reactors operate nearly isothermally, under the
same experimental conditions (temperature, pressure, concentration, residence time and catalyst), and in absence of mass and heat
transfer limitations, identical results are expected for the three
reaction slit arrangements. An important issue to the conversion
as well as the formation of DME in consecutive reactions is the
homogeneity of the catalytic bed. The methanol synthesis and the
methanol dehydration catalyst were carefully mixed to the desired
mass ratio and loaded into the micro packed bed reactor. But the
mechanical agitation upon loading could have affected the distribution between the two phases as well as the density of the packing,
and this becomes important as both the methanol synthesis and
methanol dehydration approach local equilibria. In future application for the direct DME synthesis, using a truly bifunctional catalyst
with both catalytic functions incorporated in the same material
could possibly eliminate effects of bed inhomogeneity.
3.4. Residence time distribution
The uid ow in most conventional packed beds is laminar
and a parabolic prole is assumed. However, there are different
approaches to describe nonideal ow in packed beds. Dispersion

models are widely used to represent reactor behavior over the


range of PFRs to CSTRs. The axial diffusion increases the dispersion,
while the radial dispersion diminishes the effect of the parabolic
ow prole. The dispersion coefcient (De ) is introduced through a
dimensionless group, the Peclet number (Pe), as a ratio of convective to dispersive ow. The Peclet number is normally calculated
via correlations as a function of the Reynolds (Re) and Schmidt (Sc)
numbers, which depend on ow regime. Wilhelm [51] plotted the
Peclet number for packed beds as a function of the Reynolds number (Fig. 12), and discussed the dispersion in terms of different ow
regions.
Since the three micro packed beds in this paper operate at laminar ow and 0.006 Re 1.0, we concentrate our discussion to this
range. It is noted that for Re < 100, viscous ow is approached and
dispersion becomes mainly the results of molecular diffusion. We
therefore have estimated the axial dispersion (De,a ) in our micro
packed bed reactors with the correlation proposed by Wen and Fan
(12) [52] and the radial dispersion (De,r ) with correlation proposed
by de Ligny (13) [53]:
0.5
1
0.3
=
+
Pee,a
Re Sc
1 + (3.8/(Re Sc))

(12)

0.7
0.12
1
+
=
Pee,r
Re Sc
1 + (78/(Re Sc))

(13)

where Pee,a = uint dp /De,a , Pee,r = uint dp /De,r , Re = g us dp /g and


Sc = g /g Dm , and the axial and radial Peclet numbers, Pee,a and
Pee,r , are calculated for a gas of viscosity g , density g , molecular
diffusion of mixture Dm and interstitial velocity uint .
The estimated Pee,a and Pee,r are linear functions of Re in the
ow regime relevant to our micro packed bed reactor, as shown in
Fig. 13. Therefore, the comparison between Figs. 12 and 13 indicates that the molecular diffusivity plays an important role in axial
and radial dispersion or RTD. The RTD described by the dispersion model is characterized by the Bodenstein number (Bo = tax /),
which is the ratio between the axial dispersion time (tax = L2 /De,a )
and the mean residence time ( = Lt /us ). For Bo 0 complete back
mixing exists and for Bo no dispersion occurs [54]. The axial
dispersion is normally neglected for Bo 100.
Fig. 14 illustrates the calculated Bo = us L/De,a for the three micro
packed beds against Re. The Bo number is the same function of
Re for the three reactors over the complete ow range, meaning

84

F. Hayer et al. / Chemical Engineering and Processing 70 (2013) 7785

1,4
1,2
Pee,a

1,0

Pee,r

Pee

0,8
0,6
0,4
0,2
0,0
0,0

0,2

0,4

0,6

0,8

1,0

Re
Fig. 13. Calculated axial and radial Peclet number versus Reynolds number for the
three micro packed bed reactors at 50 bar and 250 C.

1200

1000

Microreactor A
Microreactor B
Microreactor C

Acknowledgments
The nancial support from the Research Council of Norway
through the GASSMAKS research program (182531/S10) is thankfully acknowledged. HJV also acknowledges the nancing by
Statoil ASA through the Gas Technology Centre (NTNU-SINTEF)
(www.ntnu.no/gass/).

800
Bo

adding He to the synthesis gas at constant reactant partial pressure


and increased total pressure. A 20% decrease of molecular diffusivity does not inuence the CO conversion. Neither were changes
in the reactant conversion observed upon changing the linear gas
velocity. We therefore conclude that the micro packed bed is not
limited by external mass transfer under the conditions studied.
The three micro packed bed reactors operate at laminar ow
and 0.006 Re 1. The estimated radial and axial Peclet numbers
were linear functions of Re, indicating that the molecular diffusivity plays an important role in the dispersion model. The dispersion
characteristics were investigated by assessing the Bodenstein number versus Re to show that the axial dispersion is insignicant
and the radial dispersion is high. Although we could not study
the RTD experimentally, the analysis is indicative of a narrow RTD.
The pressure drop measured over our micro packed bed reaction
slits at 50 bar were linear with GHSV and always less than 0.12 bar.
However, the results demonstrate that the pillar structure and the
catalyst morphology, and possibly the reaction slit cross section
aspect ratio, affects the packing density and consequently the pressure drop. The analysis of the three different micro packed bed
reactors shows that the direct DME synthesis could potentially be
intensied by the characteristic features of microstructured reactors.

600

400

References
200

0
0,0

0,2

0,4

0,6

0,8

1,0

Re
Fig. 14. Bodenstein number versus Reynolds number for the three micro packed
bed reactor at 50 bar and 250 C.

that the axial dispersion is insignicant in the three micro packed


beds. This result can be understood by comparing the radial diffusion time (tD = dh2 /Dm ) to the mean residence time. The former is
in the order of 10, while the mean residence time is in the order
of 1001000 units. Moreover, for tD <  < tax , diffusion in the radial
direction attens the parabolic velocity prole. Altogether, this
analysis is indicative of a narrow RTD.
4. Conclusions
Three integrated micro packed bed reactor-heat exchangers
(IMPBRHEs) were applied in the direct DME synthesis over a physical mixture containing CuOZnOAl2 O3 and -Al2 O3 in mass ratio
2:1. Careful temperature measurements inside the slits, along the
reactor housing (skin) and at oil inlet and outlet, conrmed that
the reaction slit temperature is mainly controlled by the oil heat
exchange to give a practically uniform temperature prole for set
inlet oil temperatures in the range 220320 C.
The CO conversion and DME yield obtained for three different
catalyst particle size fractions as function of temperature and GHSV
conrm that the micro packed beds are free of internal mass limitations, at least for particles smaller than 125 m and temperatures
lower 280 C. The molecular diffusivity of CO was decreased by

[1] International Energy Outlook 2010 Highlights, US Energy Information Administration (EIA), May 25, 2010 http://www.eia.doe.gov/oiaf/ieo/highlights.html
[2] C.W. Simons, J. ONeill, J.A. Gribens, Aerosol propellant for personal products,
US Patent 4,041,148 (1977).
[3] F. Maroteaux, G. Descombes, F. Sauton, J. Jullien, Investigation on exhaust emissions of a common rail high-speed direct injection diesel engine running with
dimethyl ether, International Journal of Engine Research 2 (2001) 199207.
[4] C. Arcoumanis, C. Bae, R. Crookes, E. Kinoshita, The potential of di-methyl ether
(DME) as an alternative fuel for compression-ignition engines: a review, Fuel
87 (2008) 10141030.
[5] Study of Dimethyl Ether (DME) as an Alternative Fuel for Diesel Engine Applications, Advanced Engine Technology Ltd. for CANMET Energy Technology Centre,
Natural Resources Canada and Transportation Development Centre, Transport
Canada, 2001.
[6] IDA Fact Sheet DME/LPG Blends, International DME Association, 2010
http://www.aboutdme.org/aboutdme/les/ccLibraryFiles/Filename/
000000001519/IDA Fact Sheet 1 LPG DME Blends.pdf
[7] E.D. Larson, H. Yang, Dimethyl ether (DME) from coal as a household cooking
fuel in China, Energy for Sustainable Development 8 (2004) 115126.
[8] K. Xu, S.J. Lao, H.Y. Qin, B.H. Liu, Z.P. Li, A study of the direct dimethyl ether fuel
cell using alkaline anolyte, Journal of Power Sources 195 (2010) 56065609.
[9] A. Serov, C. Kwak, Progress in development of direct dimethyl ether fuel cells,
Applied Catalysis B: Environmental 91 (2009) 110.
[10] J.-H. Yoo, H.-G. Choi, C.-H. Chung, S.M. Cho, Fuel cells using dimethyl ether,
Journal of Power Sources 163 (2006) 103106.
[11] M.C. Lee, S.B. Seo, J.H. Chung, Y.J. Joo, D.H. Ahn, Industrial gas turbine combustion performance test of DME to use as an alternative fuel for power generation,
Fuel 88 (2009) 657662.
[12] E.G.R. Jones, H. Holm-Larsen, D. Romani, R.A. Sills, DME for power generation
fuel: supplying Indias Southern region, in: PETROTECH, New Delhi, India, 2001.
[13] C. Ledesma, U.S. Ozkan, J. Llorca, Hydrogen production by steam reforming of
dimethyl ether over Pd-based catalytic monoliths, Applied Catalysis B: Environmental 101 (2011) 690697.
[14] C. Ledesma, J. Llorca, Hydrogen production by steam reforming of dimethyl
ether over CuZn/CeO2 ZrO2 catalytic monoliths, Chemical Engineering Journal 154 (2009) 281286.
[15] Z. Wang, J. Wang, J. Diao, Y. Jin, The synergy effect of process coupling for
dimethyl ether synthesis in slurry reactors, Chemical Engineering & Technology
24 (2001) 507511.
[16] Y. Ohno, H. Yagi, N. Inoue, K. Okuyama, S. Aoki, Slurry phase DME direct synthesis technology 100 tons/day demonstration plant operation and scale up

F. Hayer et al. / Chemical Engineering and Processing 70 (2013) 7785

[17]

[18]

[19]

[20]

[21]

[22]

[23]

[24]

[25]
[26]
[27]
[28]

[29]

[30]
[31]

[32]

study, in: M.S. Fbio Bellot Noronha, S.-A. Eduardo Falabella (Eds.), Studies in
Surface Science and Catalysis, Elsevier, 2007, pp. 403408.
J.H. Kim, M.J. Park, S.J. Kim, O.S. Joo, K.D. Jung, DME synthesis from synthesis
gas on the admixed catalysts of Cu/ZnO/Al2 O3 and ZSM-5, Applied Catalysis A:
General 264 (2004) 3741.
X.K. Phan, H.D. Bakhtiary, R. Myrstad, J. Thormann, P. Pfeifer, H.J. Venvik,
A. Holmen, Preparation and performance of a catalyst-coated stacked foil
microreactor for the methanol synthesis, Industrial & Engineering Chemistry
Research 49 (2010) 1093410941.
J. Fei, Z. Hou, B. Zhu, H. Lou, X. Zheng, Synthesis of dimethyl ether (DME) on
modied HY zeolite and modied HY zeolite-supported CuMnZn catalysts,
Applied Catalysis A: General 304 (2006) 4954.
K.S. Yoo, J.H. Kim, M.J. Park, S.J. Kim, O.S. Joo, K.D. Jung, Inuence of solid acid
catalyst on DME production directly from synthesis gas over the admixed catalyst of Cu/ZnO/Al2 O3 and various SAPO catalysts, Applied Catalysis A: General
330 (2007) 5762.
P.S. Sai Prasad, J.W. Bae, S.-H. Kang, Y.-J. Lee, K.-W. Jun, Single-step synthesis of
DME from syngas on CuZnOAl2 O3 /zeolite bifunctional catalysts: the superiority of ferrierite over the other zeolites, Fuel Processing Technology 89 (2008)
12811286.
F. Daly, L. Tonkovich, Enabling offshore production of methanol by use of an
isopotential reactor, in: B. Xinhe, X. Yide (Eds.), Studies in Surface Science and
Catalysis, Elsevier, 2004, pp. 415420.
W.W. Simmons, R.D. Litt, T. Mazanec, A.L. Tonkovich, Process for converting
a carbonaceous material to methane, methanol and/or dimethyl ether using
microchannel process technology, US Patent 20,090,259,076 (2009).
J. Wang, L. Tonkovich, T. Mazzanee, F.P. Daly, J. Hu, C. Kibby, X.S. Li, M.D. Briscoe,
N. Gano, Y.H. Chin, FischerTropsch synthesis using microchannel technology
and novel catalyst and microchannel reactor, US Patent 7,084,180 (2006).
K.F. Jensen, Microreaction engineering is small better? Chemical Engineering
Science 56 (2001) 293303.
S.K. Ajmera, M.W. Losey, K.F. Jensen, M.A. Schmidt, Microfabricated packed-bed
reactor for Phosgene synthesis, AIChE Journal 47 (2001) 16391647.
V. Hessel, Novel process windows gate to maximizing process intensication
via ow chemistry, Chemical Engineering & Technology 32 (2009) 16551681.
A.Y. Tonkovich, S. Perry, Y. Wang, D. Qiu, T. LaPlante, W.A. Rogers, Microchannel
process technology for compact methane steam reforming, Chemical Engineering Science 59 (2004) 48194824.
T. Jarosch Kai, Y. Tonkovich Anna Lee, T. Perry Steven, D. Kuhlmann, Y. Wang,
Microchannel reactors for intensifying gas-to-liquid technology, in: Microreactor Technology and Process Intensication, American Chemical Society, 2005,
pp. 258272.
V. Hessel, A. Renken, J.C. Schouten, J. Yoshida, Micro Process Engineering: System, Process and Plant Engineering, WILEY-VCH, 2009.
H. Bakhtiary-Davijany, F. Hayer, X.K. Phan, R. Myrstad, H.J. Venvik, P. Pfeifer,
A. Holmen, Characteristics of an integrated micro packed bed reactor-heat
exchanger for methanol synthesis from syngas, Chemical Engineering Journal
167 (2011) 496503.
H. Bakhtiary-Davijnay, F. Hayer, X.K. Phan, R. Myrstad, H.J. Venvik, P. Pfeifer,
A. Holmen, Performance of a multi-slit packed bed microstructured reactor in
the synthesis of methanol: comparison with a laboratory xed bed reactor,
Chemical Engineering Science 66 (2011) 63506357.

85

[33] H. Bakhtiary-Davijnay, F. Hayer, X.K. Phan, R. Myrstad, H.J. Venvik, P. Pfeifer,


A. Holmen, Analysis of external and internal mass transfer limitations at low
Reynolds numbers in a multi-slit packed bed microstructured reactor under
real methanol synthesis operating condition, Industrial & Engineering Chemistry Research 51 (2012) 1357413579.
[34] F. Hayer, H. Bakhtiary-Davijany, R. Myrstad, A. Holmen, P. Pfeifer, H.J. Venvik,
Synthesis of dimethyl ether from syngas in a microchannel reactor simulation
and experimental study, Chemical Engineering Journal 167 (2011) 610615.
[35] F. Hayer, H. Bakhtiary-Davijany, R. Myrstad, H.J. Venvik, P. Pfeifer, A. Holmen, Modeling and simulation of an integrated micro packed bed reactor-heat
exchanger conguration for direct dimethyl ether synthesis, Topics in Catalysis
54 (2011) 817827.
[36] R. Myrstad, S. Eri, P. Pfeifer, E. Rytter, A. Holmen, FischerTropsch synthesis in
a microstructured reactor, Catalysis Today 147 (2009) S04S301.
[37] W. Liu, Ministructured catalyst bed for gas-liquid-solid multiphase catalytic
reaction, AIChE Journal 48 (2002) 15191532.
[38] S. Ergun, Fluid ow through packed columns, Chemical Engineering Progress
42 (1952) 8994.
[39] P.C. Carman, Fluid ow through granular beds, Transactions of the Institution
of Chemical Engineers 15 (1937) 150166.
[40] S.P. Burke, W.B. Plummer, Gas ow through packed columns 1, Industrial &
Engineering Chemistry 20 (1928) 11961200.
[41] D.P. Haughey, G.S.G. Beveridge, Structural properties of packed beds a review,
Canadian Journal of Chemical Engineering 47 (1969) 130140.
[42] F. Schmidt, Morphological Characteristics, WILEY-VCH Verlag GmbH & Co.
KGaA, 2008.
[43] G.B. Froment, K.B. Bischoff, Chemical Reactor Analysis and Design, 2nd ed.,
Wiley, 1990.
[44] V. Hessel, S. Hardt, H. Lwe, Chemical Micro Process Engineering: Fundamentals, Modelling and Reactions, WILEY-VCH, 2003.
[45] S. Walter, S. Malmberg, B. Schmidt, M.A. Liauw, Mass transfer limitations in
microchannel reactors, Catalysis Today 110 (2005) 1525.
[46] E.N. Fuller, P.D. Schettler, J.C. Giddings, New method for production of binary
gas-phase diffussion coefcients, Industrial & Engineering Chemistry 58 (1966)
1827.
[47] S. Chapman, T.G. Cowling, The Mathematical Theory of Non-Uniform Gases,
Cambridge University Press, Cambridge, UK, 1952.
[48] B.E. Poling, J.M. Prausnitz, J.P. OConnell, Properties of Gases and Liquids, 5th
ed., McGraw-Hill, 2001.
[49] C. Cao, D.R. Palo, A.L.Y. Tonkovich, Y. Wang, Catalyst screening and kinetic
studies using microchannel reactors, Catalysis Today 125 (2007) 2933.
[51] R.H. Wilhelm, Progress towards the a priori design of chemical reactors, Pure
and Applied Chemistry 5 (1962) 403422.
[52] C.Y. Wen, L.T. Fan, Models for Flow Systems and Chemical Reactors, Dekker,
New York, 1975.
[53] C.L. de Ligny, Coupling between diffusion and convection in radial dispersion
of matter by uid ow through packed beds, Chemical Engineering Science 25
(1970) 11771181.
[54] N. Kockmann, Micro Process Engineering: Fundamentals, Devices, Fabrication,
and Applications, WILEY-VCH, Weinheim, 2006.

S-ar putea să vă placă și