You are on page 1of 11

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/222348507

Methanation of carbon dioxide by hydrogen


reduction using the Sabatier process in
microchannel reactors

Article in Chemical Engineering Science February 2007


DOI: 10.1016/j.ces.2006.11.020

CITATIONS READS

95 1,524

4 authors, including:

Kriston P. Brooks Jianli hu


Battelle Memorial Institute Princess Alexandra Hospital (Queensland He
46 PUBLICATIONS 541 CITATIONS 54 PUBLICATIONS 1,207 CITATIONS

SEE PROFILE SEE PROFILE

Huayang Zhu
Colorado School of Mines
90 PUBLICATIONS 2,457 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

SAE World Congress View project

All content following this page was uploaded by Kriston P. Brooks on 18 March 2015.

The user has requested enhancement of the downloaded file.


Chemical Engineering Science 62 (2007) 1161 1170
www.elsevier.com/locate/ces

Methanation of carbon dioxide by hydrogen reduction


using the Sabatier process in microchannel reactors
Kriston P. Brooks a , Jianli Hu a , Huayang Zhu b, , Robert J. Kee b
a Pacic Northwest National Laboratory, Richland, WA 99352, USA
b Engineering Division, Colorado School of Mines, Golden, CO 80401, USA

Received 18 July 2006; received in revised form 3 November 2006; accepted 6 November 2006
Available online 15 November 2006

Abstract
This paper describes the development of a microchannel-based Sabatier reactor for applications such as propellant production on Mars or
space habitat air revitalization. Microchannel designs offer advantages for a compact reactor with excellent thermal control. This paper discusses
the development of a Ru.TiO2 -based catalyst using powdered form and its application and testing in a microchannel reactor. The resultant
catalyst and microchannel reactor demonstrates good conversion, selectivity, and longevity in a compact device. A chemically reacting ow
model is used to assist experimental interpretation and to suggest microchannel design approaches. A kinetic rate expression for the global
Sabatier reaction is developed and validated using computational models to interpret packed-bed experiments with catalysts in powder form.
The resulting global reaction is then incorporated into a reactive plug-ow model that represents a microchannel reactor.
2006 Elsevier Ltd. All rights reserved.

Keywords: Sabatier process; CO2 reduction; Ru.TiO2 catalyst; Microchannel reactor

1. Introduction low process temperature (around 350 C) without generating


hot spots or quenching the reaction.
As originally reported by Sabatier in 1902, it is well known The Sabatier process provides a means of producing propel-
that CO2 can be reduced by H2 over a catalyst as lants from CO2 , which is the dominant species in the Martian
atmosphere, and hydrogen transported from Earth. By utiliz-
CO2 + 4H2 CH4 + 2H2 O. (1) ing the indigenous resources on Mars such as the atmospheric
CO2 , the initial mass of a spacecraft launched from low Earth
The global Sabatier reaction, which is reversible and exothermic orbit can be reduced by 2045% (Sanders et al., 2001), with
(H = 167 kJ/mol), proceeds catalytically at relatively low even greater leverage if Martian water can be used. The CH4
temperatures on a catalyst such as ruthenium. Fig. 1 illustrates generated from the Sabatier process is fuel for the return jour-
the equilibrium composition for an initial four-to-one molar ney, and H2 O can be electrolyzed to provide H2 and O2 . The
mixture of H2 and CO2 (in 4.47% Ar) at 1 atm and temperatures O2 is the oxidant for the rocket fuel and the H2 is recycled. The
ranging from 200 to 600 C (Fig. 1 also shows experimental Sabatier process can also be used to reduce CO2 generated in a
data, which are discussed in a later section). It is evident that space habitat to reclaim the oxygen for recycle. In this case, the
lower operating temperatures favor high conversion to CH4 CH4 produced would be a byproduct that could be discarded
and H2 O. Sabatier processes operate at temperatures around or used as fuel, and the oxygen is fed back into the habitat.
400 C. A practical challenge in reactor design is to remove heat The Sabatier reaction has broader relevance than the
produced by the exothermic reaction and maintain a relatively space-based applications discussed above. There are certainly
environmental reasons to reduce CO2 . As a greenhouse gas, re-
search is underway to collect, sequester, and recycle CO2 . The
Corresponding author. Tel.: +1 303 273 3890; fax: +1 303 273 3602. Sabatier process could potentially play a role, although suf-
E-mail address: hzhu@mines.edu (H. Zhu). cient hydrogen would be needed to be available and a suitable
0009-2509/$ - see front matter 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2006.11.020
1162 K.P. Brooks et al. / Chemical Engineering Science 62 (2007) 1161 1170

small packed-bed reactor. Promising catalysts are then loaded


onto an engineered support and tested in a single channel reac-
tor. Using the results of the single channel tests, a multichannel,
microchannel reactor is developed and tested.
A reactive porous-media model is developed and applied to
the packed-bed experiments to develop a kinetic rate expression
for the global reaction. The global reaction is then incorporated
into a reactive channel-ow model that describes the Sabatier
process in both an individual channel of a microchannel reactor
and the multichannel reactor. This model is then used to ex-
plore reactor congurations and operating parameters, seeking
a high-performance reactor design.

2. Catalyst preparation and performance

The purpose of catalyst synthesis is to develop active, stable,


and supported catalysts that operate at temperatures between
300 and 400 C. Both Ru and Rh are found to have high turn-
over frequency for methanation reactions. The supports include
TiO2 (Rutile), mixed-phase TiO2 (Rutile : Anatase = 60 : 40),
and ZrO2 (Engelhard Corp.), as well as Al2 O3 and Al2 O3 MgO
Fig. 1. Measured exhaust composition as a function of bed temperature spinel (SASOL). Metal precursors include a Rh-nitrite solution
in a small reactor with a bed of Ru.TiO2 catalysts in powder form. The (10 wt% Rh) (Engelhard Corp.) and a Ru-nitrate solution (Colo-
dashed lines show the equilibrium composition and the solid lines are model nial). A catalyst-promoter precursor Ce(NO3 )2 is obtained from
predictions. The initial mixture is 4 moles H2 and 1 mole CO2 , carried in Aldrich Chemicals.
4.47% Ar.
Catalyst synthesis is carried out by the incipient-wetness
technique. Solutions containing Ru nitrate or Rh nitrate are pre-
pared and impregnated on different catalyst supports to obtain a
end use for CH4 would be needed. Furthermore, Sabatier CO2 range of metal compositions. Typical Rh or Ru loadings varies
reduction could potentially be incorporated into any number of from 16% (by weight). After introducing metals on catalyst
industrial processes. In a hydrogen economy, assuming a source supports, catalyst samples are dried at 110 C overnight. All the
of low-cost renewable hydrogen, the Sabatier reaction could be catalyst samples are subjected to nal calcination at 450 C for
important in production of methane, alcohols, and long-chain 3 h using a ramp rate of 2 C/min.
hydrocarbons from waste CO2 . Catalyst performance is analyzed by measuring CO2 conver-
The Sabatier process is often implemented in a packed-bed sion and CH4 selectivity, using catalysts in powder form. By
or porous-foam reactor conguration. This paper considers im- comparing CO2 conversion and CH4 selectivity of 3% Ru cata-
plementation in a microchannel reactor. Microchannel reactors lyst supported on all the support materials at 1 atm, 350 C and
can deliver high heat- and mass-transfer between the reactive- H2 /CO2 = 4 : 1, the mixed-phase TiO2 is found to be the best
gas ow and the channel walls, thus enabling precise temper- support for the Ru catalyst. Moreover, the 3% Ru supported on
ature control in a compact reactor (Cao et al., 2004, 2005; mixed-phase TiO2 is found to be stable during lifetime testing.
Aartun et al., 2004; Drost et al., 1997; Holladay et al., 2002; The catalytic stability was determined by monitoring CO2 con-
Palo et al., 2005; TeGrotenhuis et al., 2003; Brooks et al., 2005). version as a function of time on stream (TOS). During a total
By controlling the heat transfer along the length of the reactor, of 77 h operation, including 16 shut-down and re-start cycles,
catalyst deactivation due to excess temperature can be avoided. loss of CO2 conversion was less than 2% (Hu et al., 2007).
The microchannel Sabatier reactor discussed in this paper is In a separate study, maximum CO2 conversion is found to oc-
designed to deliver approximately 16 g/h of methane (Brooks cur at a metal loading of about 3% (Hu et al., 2007). Generally,
et al., 2006). lower metal loadings can result in a higher percentage of metal
The Sabatier process has been widely investigated using al- dispersion and smaller metal particle size, which is translated
ternative catalyst metals (i.e. Ni, Ru, Rh, and Co) on various ox- to higher turn over frequency (TOF). However, at lower metal
ide supports (i.e. TiO2 , SiO2 , MgO, and Al2 O3 ) (Lunde, 1974; loading, the site density required to achieve a certain level of
Lunde and Kester, 1973, 1974; Henderson and Worley, 1985; CO2 conversion may be lower. In contrast, at higher metal load-
Mori et al., 1996; Nakayama et al., 1997; Ohya et al., 1997; ing, the percentage of metal dispersion may be lower; however,
Li et al., 1998; Prairie et al., 1991). The work reported here the site density could be higher. The typical volcano-shaped
uses a Ru.TiO2 system, which is one of the most active and curve is expected. At higher metal loading, due to the difculty
stable combinations (Prairie et al., 1991). A range of catalysts in dispersing the metal, average metal particle size becomes
and supports are prepared in powder form and evaluated in a larger, therefore ultimately affecting hydrogenation activity.
K.P. Brooks et al. / Chemical Engineering Science 62 (2007) 1161 1170 1163

Fig. 1 shows results of packed-bed experiments. As synthe-


sized, the 3% Ru.TiO2 catalyst powders are in the size range
of 70100 mesh. A small sample of catalyst powders (0.42 cm
diameter and 0.87 cm long) is packed loosely within a quartz
tube. The packed tube is placed in a temperature-controlled fur-
nace for testing. Reactive gases ow through the bed and the
exhaust composition is measured by gas chromatography. For
the data shown in Fig. 1 the inlet composition is 74.96% H2 ,
20.57% CO2 and 4.47% Ar. The ow rate is 90 sccm, which Fig. 3. Illustration of a single channel with an interior catalyst-loaded porous
corresponds to a space velocity of 45000 h1 , and the operat- metal felt.
ing pressure is 1 atm. The concentrations of CH4 , H2 , CO and
CO2 are measured directly, but the H2 O mole fraction is cal-
culated based on element balances. As illustrated in Fig. 1, the
3% Ru.TiO2 catalyst achieves near-equilibrium conversion and
selectivity to CH4 for temperatures above 350 C.

3. Microchannel reactor development

A major challenge in designing a Sabatier process is to


control temperature. It is evident from Fig. 1 that selectivity to
desired products (H2 O and CH4 ) is favored at low temperature.
However, because reaction kinetics are slow at low temperature,
it is difcult to initiate the reaction at low temperature. Once
the reaction is initiated at temperatures around 350 C, remov-
ing the heat from the exothermic reaction can be difcult. If the
heat is not removed effectively, the process increases temper-
ature, reduces the effectiveness, and shifts the selectivity un-
favorably. Microchannel reactors are used in the present work
because they offer the possibility of efcient and exible ther-
mal management. There are some potential benets associated
with initiating reaction at relatively high temperature, then de-
creasing the temperature along the length of the microchannels.
Fig. 2 illustrates a microchannel reactor with active side-
wall cooling. In this approach, the reactive gases ow in the
rectangular channels, which contains the catalyst. A counter
ow of cooling oil ows in the outer oval-shaped channels. Fig. 4. Engineering drawing of the Sabatier microchannel reactor assembly.
The reactor body is metal, providing good thermal distribution.
Depending on the ow rates and inlet temperatures, the oil ow
is used to control wall-temperature proles along the length of the reactive microchannels. In the data reported here, the oil
has been used to hold wall temperatures nearly uniform at 300,
350, and 400 C, and the oil ow rates have also been varied
to establish temperature variations along the channel length of
around 50 C.
It is typical to apply the catalyst directly to the channel wall
as a porous washcoat. However, Fig. 3 illustrates an alternative
approach used here. A thin catalyst-loaded porous-metal felt is
placed within each channel. The felt is a FeCrAlY intermetallic
alloy that is coated with the Ru.TiO2 catalyst. The felt is not
bonded directly to the channel walls.
Fig. 4 shows a CAD rendering of the microchannel reactor
that was designed and used in the present effort. There are
two parallel columns of 15 microchannels each (illustrated in
Fig. 2). The reactor is fabricated from 316 stainless steel, with
a net volume of 40.8 cm3 and weighing 90 g. In addition to
Fig. 2. Illustration of a section of microchannel reactor. In this approach the the microchannels and oil passages, two additional features are
reacting gases ow in the central rectangular channels and oil ows in the included in this reactor. The rst is a small foam-based catalytic
outer oval channels. reactor at the gas inlet. The second is a mixing volume that
1164 K.P. Brooks et al. / Chemical Engineering Science 62 (2007) 1161 1170

permits gas mixing between the microchannels midway along catalyst surfaces, but homogeneous chemistry is neglected. The
the channel length. model used here is a subset of a more-comprehensive model
As the gas enters the reactor, the exothermicity of the Sabatier that considers thermal as well as mass balances (Zhu et al.,
reaction usually causes a large temperature spike at the leading 2006). However, because the powder sample is small and held
edge of the catalyst, resulting in catalyst sintering. To prevent within a temperature-controlled furnace, the ow within the
this damage to the microchannels, the gases ow through al- bed is approximated to be isothermal at the measured furnace
ternating layers of catalyst and heat exchange upstream of the temperature.
microchannels. The catalyst is coated on three 0.08-cm-thick The species and overall mass-continuity equations are written
layers of high-thermal-conductivity copper foam that is com- in conservative form as
pressed between two heat-exchanger panels that are 0.21 cm j(Yk )
thick. This feature is illustrated just downstream of the reac- + jk = As sk Wk , (2)
jt
tant inlet in Fig. 4. Reaction gases and cooling oil ow through Kg Kg
alternating cross-ow channels to remove the excess reaction j()  
+ jk = As sk Wk . (3)
heat. Because of the close contact between the catalyst and the jt
k=1 k=1
heat-exchanger panels, heat generated by the reaction is trans-
ferred both by convection through the gases and conduction In these equations  is the gas-phase density, Yk are mass frac-
through the foam to the heat exchanger panels. The partially tions, Wk are molecular weights. The bed porosity is  and
reacted gases then enter the microchannel section. As is the specic surface area of the active catalysts (i.e., ac-
The heat-transfer oil (Julabo HT 350) used to cool the reactor tive surface area per unit volume of the porous media). The
has a maximum operating temperature of 350 C. The reactor molar production rates of gas-phase species by heterogeneous
is designed so that 46 W can be removed when the oil is ow- reaction are represented as sk . These production rates are eval-
ing at 20 mL/min. By pumping cooler oil into the outlet end of uated the using local gas concentrations within the pore space.
the reactor at low ow rates, the high temperatures from the In other words, there is no mass-transfer resistance around in-
exothermic reaction create a temperature gradient from the inlet dividual particles. At the small pore scale the gas is presumed
to the outlet of the reactor. The temperature gradient can pro- to be perfectly mixed and thus uniform. Of course, the compo-
mote fast Sabatier kinetics near the channel inlet, where tem- sition varies along the length of the bed. The effective catalyst
peratures are high, and higher conversion to CH4 at the channel area and the effects of any mass-transfer resistance, around a
exit, where the temperatures are lower. At high oil ow rates, particle or within a particle, are represented via the effective
the oils heat capacity overwhelms the reaction heat and the specic area As . The pressure p is determined from the equa-
reactor operates nearly isothermally. tion of state as
As the oil leaves the reactor it is cooled in a recuperative heat Kg

exchanger. The cooling is done because the recirculation pump p = RT Yk /Wk , (4)
and oil ow meter cannot operate at temperatures above 150 C. k=1
The oil is then reheated with cartridge heaters before re-entering
where T is the temperature, and R is the universal gas constant.
the microchannel reactor. A recuperative heat exchanger is also
The species mass uxes jk through the open pore structure
used in the oil line to recover heat generated by the reaction
are determined from the dusty-gas model (DGM) (Mason and
and help cool the exiting oil.
Malinauskas, 1983; Zhu et al., 2005, 2006). The DGM is writ-
ten as an implicit relationship among the molar concentrations,
4. Model development
molar uxes, concentrations gradients, and the pressure gradi-
ent as
Lunde and coworkers have previously reported experimental
 [X ]Jk [Xk ]J Jk [Xk ] Bg
and modeling work on the Sabatier process (Lunde and Kester, + e = [Xk ] e p.
1973, 1974; Lunde, 1974). Lunde and Kester (1973) developed [XT ]Dk
e Dk,Kn Dk,Kn 
=k
an empirical rate expression to describe the Sabatier kinetics (5)
for a Ru.Al2 O3 catalyst. The present effort follows this ap-
proach, but rets the rate expression to represent measurements In this relationship Jk is the molar ux of gas-phase species
for Ru.TiO2 catalysts. Lunde (1974) also developed a reac- k, [Xk ] are the molar concentrations, [XT ] = p/RT is the to-
tive porous-media model to represent a packed-bed reactor. The tal molar concentration, and Bg is the permeability. The mass
models developed in the present effort are described below. uxes are related simply to the molar uxes as jk = Wk Jk .
The mixture viscosity is given as  and Dke and D e
k,Kn are the
4.1. Porous-media model effective binary and Knudsen diffusion coefcients.
The effective diffusion coefcients are written as

A one-dimensional, porous-media, chemically reacting ow  4 rp  8RT
model is based on mass conservation within the packed bed. Dk = Dk , Dk,Kn =
e e
. (6)
 3  Wk
A dusty-gas formulation represents gas-phase species transport,
considering ordinary and Knudsen diffusion as well as pressure- The binary diffusion coefcients Dk and the mixture viscosi-
driven Darcy ow. Chemical kinetics is considered on the ties are determined in the ordinary way from kinetic theory
K.P. Brooks et al. / Chemical Engineering Science 62 (2007) 1161 1170 1165

Table 1 ow models is well established (Kee et al., 2003; Raja et al.,


Parameters for porous media ow within the catalyst powder bed 2000). The governing equations are summarized here as
Parameters Value Units
Kg
j ju Pc 
Bed length (L) 0.87 cm + = sk Wk , (8)
Porosity () 0.75 jt jx Ac
k=1
Tortuosity () 3  
m
jYk jYk u jjk Pc
Pore radius (rp ) 60 + + =  k + sk Wk , (9)
Particle diameter (dp ) 180 m jt jx jx Ac
Specic catalyst area (As ) 1500 cm1 ju ju2 jp Ph
+ = + w , (10)
jt jx jx Ac

(Kee et al., 2003). Knudsen diffusion represents mass transport and


assisted by gaswall collisions. The Knudsen diffusion coef- je jqg
cients depend on the porous-media geometry, including poros- + = qgw . (11)
jt jx
ity, average pore radius rp , and tortuosity . The permeability
can be determined from the KozenyCarman relationship as The independent variables are time t and the channel position
x. Dependent variables are mass density , mass fractions Yk ,
3 dp2 mean velocity u, temperature Tg , and pressure p. The gas-phase
Bg = , (7)
72(1 )2 specic internal energy is e. The channel geometry is char-
acterized by cross-sectional ow area Ac , the hydrodynamic
where dp is the particle diameter. Further details of the DGM perimeter Ph , and the catalytically active perimeter Pc . The
and its numerical implementation can be found in Zhu et al. molar production rates by homogeneous and heterogeneous re-
(2005). actions are represented as 
k and sk , respectively. In the work
The steady-state governing equations form a boundary-value reported here, however, gas-phase chemistry is negligible. Gas-
problem. The independent variable is the spatial coordinate x. phase species diffusion ux along the channel length is mod-
The dependent variables are the mass density , mass fractions eled using a mixture-averaged formulation as
Yk , and pressure p. Boundary conditions are needed at the inlet
and exit of the bed. At the inlet the species mass-ow rates are Wk jXk
jk =  Dkm . (12)
specied and the outlet boundary is modeled by specifying the W jx
exit pressure.
The model is implemented to solve either the transient or The mixture-averaged diffusion coefcients are
the steady problem. In either case, a one-dimensional nite- 1 Yk
volume spatial discretization is used. The transient problem is Dkm = K , (13)
g
solved using a method-of-lines algorithm, with the time march- j =k Xj /Dj k
ing accomplished using the LIMEX software, which is designed
to solve systems of differential-algebraic equations (Deuhard where Xk are the mole fractions, W is the mixture mean molec-
et al., 1987). The steady-state problem is solved either as the ular weight, and Dj k are the binary diffusion coefcients (Kee
long-time solution of the transient problem, or directly as a et al., 2003).
boundary-value problem using a hybrid-Newton method (Grcar The wall shear stress w can be represented in terms of a
et al., 1986; Kee et al., 2003). friction factor f = 2w /u2 . For a rectangular channel
Table 1lists the physical parameters that are used for model- Re f 13.74 + 10.38 exp(3.4/), (14)
ing the catalyst-powder experiments.
where =H /W is the channel aspect ratio. The local Reynolds
4.2. Microchannel model number is dened in terms of the mean velocity u and the
hydraulic diameter (Dh = 2H W /(H + W )) as Re = uD h /,
Chemically reacting ow within the microchannels is mod- where  is the mixture dynamic viscosity.
eled by solving mass-, momentum-, and energy-conservation The axial heat ux qg , considering both heat conduction and
equations. The channel-wall temperature can be specied, ei- species transport, is written as
ther as uniform or as a prole. Consistent with one-dimensional
Kg
plug-ow approximations, any variation across the channel jTg 
width is neglected, but species composition, gas-phase density, qg = + (Yk u + jk )hk , (15)
jx
velocity, pressure, and temperature vary along the length of the k=1

channel. Streamwise transport, including convective and diffu- where hk are the specic enthalpies of the gas-phase species.
sive gas-phase species transport, is included. This formulation The convective heat transfer between the channel ow and
enables upstream species diffusion, which can be important in the channel wall can be represented as
low-ow situations. Heterogeneous catalytic chemistry is con-
sidered at the channel walls. Deriving such one-dimensional qgw = hconv (Tw Tg )Ph /Ac , (16)
1166 K.P. Brooks et al. / Chemical Engineering Science 62 (2007) 1161 1170

where Tw is the channel-wall temperature. The heat-transfer totic approach to equilibrium. The rate of progress for a global
coefcient hconv is based on the heat-transfer correlations in Sabatier reaction is written as
form of a Nusselt number, N u = hconv Dh / . A Nusselt-number
correlation for developing ow in a rectangular channel is given q = kfn [CO2 ]n [H2 ]4n krn [CH2 ]n [H2 O]2n , (18)
as (Groppi et al., 1995) where n is the empirical factor. Forward and reverse rate con-
 0.5368   stants are represented as kf and kr , respectively. The rate con-
1000 42.49
Nu = 2.977 + 6.854 exp . (17) stants and the concentrations are all raised to the power n,
Gz Gz which preserves microscopic reversibility through the equilib-
rium constant. The forward rate constant is represented in Ar-
The Graetz number is dened as Gz = Re P r Dh /x.
rhenius form,
In circumstances where the catalyst is bonded directly to the
channel walls, a mass-transfer coefcient may be used to repre- kf = A exp(Ea /RT ), (19)
sent mass-transfer resistance between the mean composition in
the channel and the catalyst surface. In this case a Sherwood- where A and Ea are the pre-exponential factor and activation
number correlation is typically used. In the situation here, how- energy, respectively. The reverse rate constant is determined
ever, the catalyst is loaded onto a relatively open felt structure from microscopic reversibility as kr =Keq /kf , where Keq is the
that is not bonded to the channel walls (Fig. 3). For this con- equilibrium constant. Because all reactants are in the gas-phase,
guration, it may be assumed that the heterogeneous reaction the equilibrium constant is evaluated easily from the gas-phase
rates can be evaluated using the mean gas composition along thermodynamics. The production rate of species k is determined
the channel. Further, it may be assumed that the felt and the from the rate-of-progress and the reaction stoichiometry as
gas are at the same temperature, which may differ from the
sk =
k q, (20)
wall temperature. Consequently, heterogeneous reaction rates
are evaluated at the gas temperature, not the wall temperature. where
k is the stoichiometric coefcient of species k in the
The catalytically active perimeter Pc is used to represent the global reaction (i.e., Eq. (1)).
active catalyst area per unit length of the channel. This parame- The model is used to determine reaction-rate parameters,
ter is adjusted empirically to represent a particular felt structure seeking a best t to the measured exhaust species. As illus-
and catalyst loading. In the modeling results that follow, the trated in Fig. 1, an excellent t to the powder data is achieved
catalytic perimeter is taken to be Pc = 65 cm (approximately using the parameter n = 0.30 and an activation energy of Ea =
45 times the hydraulic perimeter Ph ). This represents a specic 69.06 kJ/mol. The pre-exponential factor is A = 7.75 106
area of the catalyst within the felt to be As 1150 cm1 . in appropriate cm-mol-s units. It should be noted that in this
Eqs. (8)(11) form a boundary-value problem, whose solu- model the product of specic catalyst area As and the Arrhenius
tion requires boundary conditions. At the channel inlet, the tem- pre-exponential factor A appear as a single parameter. Without
perature, velocity, and species composition are specied. At the an independent measurement of the catalyst specic area, the
channel exit the pressure is specied. Also, because species dif- pre-exponential factor cannot be determined uniquely.
fusion is included, an exit boundary condition is needed for the Fig. 1 shows that at temperatures of around 400 C and above
species continuity equations (Eq. (9)). In this case, the mass- the data and the models are very close to chemical equilibrium.
fractions gradients are assumed to vanish. At temperatures below 400 C the reaction is slow, leading to
The channel-ow problem is mathematically similar to the signicant departures from equilibrium.
porous-bed problem described above. That is, the steady-state It is interesting to note that the early Sabatier experiments
problem is a one-dimensional boundary-value problem that can by Lunde (Lunde and Kester, 1973, 1974; Lunde, 1974) us-
be solved computationally. Here, the channel ow problem is ing Rualumina catalysts found parameters very similar to
solved using a method-of-lines approach to solve the associ- those we found here. The Lunde work found n = 0.225 and
ated transient problem to a steady solution. The time-marching Ea = 74.46 kJ/mol. Other work by Ohya et al. (1997), using
algorithm is implemented with the LIMEX software (Deuhard similar experiments and Rualumina catalysts, found correlat-
et al., 1987). ing parameters of n = 0.85, and Ea = 69.06 kJ/mol.

5. Results and discussion 5.2. Microchannel reactor

5.1. Global Sabatier reaction mechanism Table 2 summarizes 20 different experimental conditions for
microchannel-reactor operation. The operating conditions con-
There is prior literature in tting global reaction mechanisms sider different gas temperatures, wall temperatures, and ow
for the Sabatier processes. We follow the approach developed rates. In all cases the reactor is operating at atmospheric pres-
by Lunde and Kester (1973, 1974) and Lunde (1974), who sure. The temperatures, CO2 conversion, and CH4 outlet mass
use an empirical factor n to modify a kinetic rate expression. ow rates are measured directly. The inlet velocity is com-
This factor is adjusted to match the experimental observations puted from the measured volumetric ow rate and the channel
in a packed-bed reactor. Although empirical, the approach is cross-sectional area (excluding the catalyst). The inlet velocity
consistent with microscopic reversibility, thus assuring asymp- is evaluated at the reactor inlet temperature. The residence time
K.P. Brooks et al. / Chemical Engineering Science 62 (2007) 1161 1170 1167

Table 2
Experimental ow conditions

Case H2 : CO2 Tg,in Tg,out To,in To,out Uin CO2 conversion CH4 out T
(molar) ( C) ( C) ( C) ( C) (cm/s) (%) 104 (g/s) (ms)

A 4 251 254 251 253 14.62 53.7 0.36 489.8


B 4 285 255 251 284 15.56 75.1 0.50 460.1
C 4 311 280 281 308 16.30 89.4 0.60 439.5
D 4 357 301 302 349 35.23 89.5 1.20 203.3
E 4 304 303 299 306 32.29 89.4 1.07 221.8
F 4 300 300 295 301 32.04 80.6 1.08 223.4
G 4 357 302 300 348 46.90 88.4 1.58 152.7
H 6 338 300 300 329 68.35 99.8 1.91 104.8
I 6 301 276 270 296 64.20 90.7 1.74 111.5
J 6 278 284 274 279 61.59 64.0 1.23 116.3
K 4 398 327 323 388 74.99 88.4 2.37 95.48
L 4 353 346 343 354 69.97 88.2 2.37 102.3
M 4 353 305 301 345 69.95 85.4 2.29 102.4
N 4 303 308 294 308 64.37 64.0 1.72 111.2
O 4 298 302 290 299 63.79 64.8 1.74 112.2
P 4 402 338 306 379 150.9 82.7 4.44 47.46
Q 4 366 347 350 362 142.8 84.1 4.51 50.16
R 4 352 305 296 343 139.7 79.4 4.26 51.24
S 4 311 301 293 314 130.5 68.4 3.67 54.87
T 4 302 308 289 304 128.5 55.9 3.00 55.70

As illustrated in Fig. 4 and discussed in Section 3, the reactant


gases ow through a short catalytic heat-exchanger section prior
to entering the microchannels. The intent of this feature is to
limit an initial temperature spike owing to exothermic reaction.
Although not measured directly, the catalyst loading and coolant
ow is designed to convert approximately 20% of the reactants
and limit the temperature to about 350 C. To simulate this
behavior, the channel model is modied to increase the nominal
heat-transfer coefcient by a factor of eight in the initial 0.7 cm
region of the channel.
Fig. 6 shows model predictions for velocity, temperature, and
Fig. 5. Predicted methane mass ow rates at the exit of the reactor compared mole-fraction proles along the channel length for situations
to measured methane ow rates as functions of reactor residence time. The where the channel-wall temperature varies from approximately
temperatures refer to the inlet temperatures of the gas and the oil. The letter 350 C at the gas inlet to 300 C at the gas exit. In all three
label associated with each data point refers to the cases summarized in cases, the gas enters at approximately 350 C, but each case has
Table 2.
a different inlet velocity. Following the initial heat-exchanger
section (shaded region in Fig. 6), the gas-phase temperature
is estimated as T = L/Uin , where L is the channel length and increases noticeably as the ow enters the microchannels. This
Uin is the channel inlet velocity. In all cases the channel length increase is due to the exothermic reaction and resistance to
is L = 7 cm. transferring heat to the channel walls. When the heat-exchanger
Fig. 5 shows measured and modeled methane outlet mass- section is not used, the initial temperature in the channels can
ow rates as functions of residence time. The channel-wall increase to around 650 C, potentially causing catalyst damage
temperature is assumed to vary linearly between the measured in the initial section of the microchannels.
inlet and the outlet oil temperature. However, the local gas As the gases ow through the microchannels, conversion con-
temperature is computed from the energy equation in the model. tinues and the temperatures approach the channel-wall temper-
The data points in Fig. 5 are labeled by letter corresponding to atures. As the gas-phase temperature decreases toward the wall
the experiments in Table 2. The data symbols are also grouped temperature, the velocity also decreases because the density in-
according to the inlet gas and oil temperatures. For example, creases. The higher-velocity cases (e.g., Case R) result in lower
the data marked by circles represent cases with nearly uniform CO2 conversion and less CH4 and H2 O in the exhaust than the
wall temperatures of Tw = 300 C and inlet gas temperatures of lower velocity cases. This is the anticipated (and measured) re-
Tg = 300 C. Four corresponding sets of modeling predictions sult, owing to shorter residence time in the high-velocity cases.
are shown as solid lines. It is evident that the model predicts Solution proles in all the simulations share some common
the measured methane-production rates very well. attributes. Near the channel inlet, the gas temperatures increase
1168 K.P. Brooks et al. / Chemical Engineering Science 62 (2007) 1161 1170

Fig. 7. Predicted single-channel performance with three different channel


temperature proles.

The net ow rate is 218 sccm. The heavy lines are for a uniform
temperature of 400 C.
The equilibrium products of the Sabatier reaction prefer the
desired CH4 and H2 O at low temperatures. However, the ki-
netics are too slow at temperatures below about 350 C to op-
Fig. 6. Predicted temperature, velocity, and species proles for three operating
erate the reactor at a uniform low temperature. The results in
conditions. The letter labels refers to the cases summarized in Table 2. Fig. 7 provide some quantitative insight about potential advan-
The shaded sections on the left side of the plots represent the catalytic tages for such a reactor control.
heat-exchanger region, positioned upstream of the microchannel entrance. The light solid line in Fig. 7 represents a situation where
the inlet temperature is 400 C, and the temperature decreases
linearly to 300 C at the channel exit. This prole leads to a
above the wall temperature. Despite relatively high heat trans- roughly 10% improvement in performance. The light dashed
fer to the channel walls, all the heat resulting from the exother- lines represent a situation where the temperature prole de-
mic chemistry cannot be transferred immediately to the walls. creases linearly to 200 C at the channel exit. In this case
However, as the ow proceeds along the microchannnel the gas the temperature decreases too rapidly, thus prematurely slow-
temperature approaches the wall temperature. ing the kinetics. It turns out for the ow rates here that the
The modeling results indicate that the Sabatier chemistry 200 C ramp leads to an overall reactor performance that is
is rapid near the channel inlet, with diminishing rates in the comparable to the uniform 400 C situation. Other tempera-
downstream sections of the channels. This leads to relatively ture proles may be found that lead to a better performance.
low sensitivities to design parameters like channel length. However, it appears that the potential for improving reactor
performance based on temperature-prole control is in the
5.3. Channel temperature proles range of 10%. This production is consistent with experimental
observation.
A potentially interesting design alternative is to control the
channel-wall temperature prole to achieve optimal perfor- 6. Summary and conclusions
mance. One approach is to initiate the reaction at high tem-
perature ( 400 C) near the reactor inlet, then decrease the A microchannel reactor has been designed and demonstrated
temperature along the channel. The hope is that once initiated to implement the Sabatier process for CO2 reduction by H2 ,
the catalytic processes will maintain near-equilibrium perfor- producing H2 O and CH4 . A Ru.TiO2 catalyst is found to pro-
mance at the lower temperatures, where conversion to CH4 vide good performance and stability. Based on experiments
and H2 O is more favorable than at higher temperatures (see with catalysts in powder form and model-based interpretation,
Fig. 1). The objective is to assist optimization of reactor design a rate expression for the global reaction has been established.
and operating conditions, seeking to maximize conversion and This rate expression, together with a reactive plug-ow model,
selectivity. is then used to assist microchannel-reactor design and experi-
Fig. 7 shows results of a study considering alternative tem- mental interpretation. Active thermal control of a microchannel
perature proles along the length of the channel. All cases in reactor can be achieved by a counter-ow of oil to maintain de-
Fig. 7 use an inlet mixture of 76% H2 , 20% CO2 , and 4% Ar. sired microchannel wall temperatures. The reactive-ow model
K.P. Brooks et al. / Chemical Engineering Science 62 (2007) 1161 1170 1169

is validated by comparison with measured conversion rates and [Xk ] gas-phase species mole concentrations,
selectivity in a microchannel reactor. mol/cm3
[XT ] total gas-phase mole concentrations, mol/cm3
Notation Yk gas-phase species mass fractions
Greek letters
A pre-factor of the Arrhenius form
 channel aspect ratio
Ac channel cross-sectional ow area, cm2
heat conductivity, J/cm K s
As specic surface area of the active catalysts,
 gas-phase viscosity, g/cm s
cm1

k reaction stoichiometry
Bg permeability, cm2
 gas-phase mass density, g/cm3
dp particle diameter, cm
 tortuosity
Dh Hydraulic diameter, cm
e w wall shear stress, dynes/cm2
Dk,Kn effective Knudsen diffusion coefcients, cm2 /s
T residence time, s
Dk binary diffusion coefcients, cm2 /s  porosity
Dke effective binary diffusion coefcients, cm2 /s 
k molar production rate by gas-phase reactions,
Dkm mean mixture diffusion coefcients, cm2 /s mol/cm3 s
Ea activation energy, J/mol
f friction factor
Gz Graetz number
Acknowledgment
H channel height, cm
hk species heat enthalpy, J/g
This work was supported by NASA Johnson Space Center
hconv heat transfer coefcient, W/m2 K under contract NNJ05HB58C.
jk gas-phase species mass ux, g/cm2 s
Jk gas-phase species mole ux, mol/cm2 s
Keq reaction equilibrium constant
Kg number of gas-phase species References
kf forward rate constants
Aartun, I., Venvik, H., Holmen, A., Pfeifer, P., Grke, O., Schubert, K.,
kb backward rate constants 2004. Temperature proles and residence time effects during catalytic
L channel length, cm partial oxidation and oxidative steam reforming of propane in metallic
n empirical factor for the global Sabatier reaction microchannel reactors. Catalysis Today 110, 98107.
Nu Nusselt number Brooks, K., Fischer, C., King, D., Pederson, L., Rawlings, G., Stenkamp,
p pressure, dynes/cm2 V., Tegrotenhuis, W., Wegeng, R., Whyatt, G., 2005. Fuel reformation:
microchannel reactor design. In: Microreactor Technology and Process
Pc catalytically active Channel perimeter, cm
Intensication. ACS Symposium Series, vol. 914. pp. 238257.
Ph hydraulic diameter of the channel, cm Brooks, K., Caldwell, D., Holladay, J., Howard, C., Hu, J., Kee, R., Lilley, B.,
Pr Prandtl number Rassat, S., Romig, K., Schlahta, S., Simon, T., Zhu, H., 2006. Microchannel
q rate of progress for the global Sabatier reaction, in situ propellant production system project. Technical Report PNWD-
mol/cm3 s 3670, Battelle. Pacic Northwest Division, Richland, WA.
qg heat ux, W/cm2 s Cao, C., Xia, G., Holladay, J., Jones, E., Wang, Y., 2004. Kinetic studies
of methanol steam reforming over Pd/ZnO catalyst using a microchannel
qqw convective heat transfer rate between the chan-
reactor. Applied Catalysis A: General 262, 1929.
nel ow and wall, W/cm2 s Cao, C., Wang, Y., Rozmiarek, R., 2005. Heterogeneous reactor model for
R universal gas constant, J/mol K steam reforming of methane in a microchannel reactor with microstructured
Re Reynolds number catalysts. Catalysis Today 110, 9297.
rp pore radius, cm Deuhard, P., Hairer, E., Zugck, J., 1987. One-step and extrapolation methods
sk molar production rate by surface reactions, for differential-algebraic systems. Numerische Mathematik 51, 501516.
mol/cm2 s Drost, M., Call, C., Cuta, J., Wegeng, R., 1997. Microchannel
integrated evaporator/combustor thermal processes. Journal of Microscale
t time, s
Thermophysics Engineering 1, 323333.
T pore gas-phase temperature, K Grcar, J., Kee, R., Smooke, M., Miller, J., 1986. A hybrid Newton/time-
Tg channel gas-phase ow temperature, K integration procedure for the solution of steady, laminar, one-dimensional
Tw channel wall temperature, K premixed ames. Proceedings of the Combustion Institute 21, 17731782.
u mean channel ow velocity, cm/s Groppi, G., Belloli, A., Tronconi, E., Forzatti, P., 1995. A comparison of
Uin channel inlet velocity, cm/s lumped and distributed models of monolith catalytic combustors. Chemical
Engineering Science 50, 27052715.
W channel width, cm
Henderson, M., Worley, S., 1985. An infrared study of the hydrogenation
Wk species molecular weight, g/mol of carbon dioxide on supported rhodium catalysts. Journal of Physical
W mean molecular weight, g/mol Chemistry 89, 14171423.
x spatial coordinate, cm Holladay, J., Jones, E., Phelps, M., Hu, J., 2002. Microfuel processor for use
Xk gas-phase species mole fractions in a miniature power supply. Journal of Power Sources 108, 2127.
1170 K.P. Brooks et al. / Chemical Engineering Science 62 (2007) 1161 1170

Hu, J., Brooks, K., Holladay, J., Howe, D., Simon, T., 2007. Catalyst Palo, D., Holladay, J., Dagle, R., Chin, Y.-H., 2005. Integrated methanol fuel
development for microchannel reactors in Mars exploration. Catalysis processors for portable fuel cell systems. In: Microreactor Technology and
Today, to appear. Process Intensication. ACS Symposium Series, vol. 914. pp. 209223.
Kee, R., Coltrin, M., Glarborg, P., 2003. Chemically Reacting Flow: Theory Prairie, M., Renken, A., Higheld, J., Thampi, K., Grtzel, M., 1991. A
and Practice. Wiley, Hoboken, NJ. Fourier transform infrared spectroscopic study of CO2 methanation on
Li, D., Ichikuni, N., Shimazu, S., Uematsu, T., 1998. Catalytic properties supported ruthenium. Journal of Catalysis 129, 130144.
of sprayed Ru/Al2 O3 and promoter effects of alkali metals in CO2 Raja, L., Kee, R., Deutschmann, O., Warnatz, J., Schmidt, L., 2000. A critical
hydrogenation. Applied Catalysis A: General 172, 351358. evaluation of NavierStokes, boundary-layer, and plug-ow models of the
Lunde, P., 1974. Modeling, simulation, and operation of a Sabatier reactor. ow and chemistry in a catalytic-combustion monolith. Catalysis Today
Industrial & Engineering Chemistry Process Design Development 13, 59, 4760.
226233. Sanders, G., Peters, T., Wegeng, R., TeGrotenhuis, W., Rassat, S., Brooks, K.,
Lunde, P., Kester, F., 1973. Rates of methane formation from carbon dioxide Stenkamp., S., 2001. Report on development of micro chemical/thermal
and hydrogen over a ruthenium catalyst. Journal of Catalysis 30, 423429. systems for Mars ISRU-based missions. In: 39th American Institute of
Lunde, P., Kester, F., 1974. Carbon dioxide methanation on a ruthenium Aeronautics and Astronautics Space Sciences Meeting. AIAA 2001-0939,
catalyst. Industrial & Engineering Chemistry Process Design and Reno, Nevada.
Development 13, 2733. TeGrotenhuis, W., King, D., Whyatt, G., Fischer, C., Wegeng, R., Brooks,
Mason, E., Malinauskas, A., 1983. Gas Transport in Porous Media: the Dusty- K., 2003. Microchannel reactors with temperature control. US Patent
Gas Model. Elsevier, New York. Application 20030180216.
Mori, S., Xu, W., Ishidzuki, T., Ogasawara, N., Imai, J., Kobayashi, K., 1996. Zhu, H., Kee, R., Janardhanan, V., Deutschmann, O., Goodwin, D., 2005.
Mechanochemical activation of catalysts for CO2 methanation. Applied Modeling elementary heterogeneous chemistry and electrochemistry in
Catalysis A: General 137, 255268. solid-oxide cells. Journal of the Electrochemical Society 152, A2427
Nakayama, T., Ichikuni, N., Sato, S., Nozaki, F., 1997. Ni/MgO catalyst A2440.
prepared using citric acid for hydrogenation of carbon dioxide. Applied Zhu, H., Kee, R., Engel, J., Wickham, D., 2006. Catalytic partial oxidation
Catalysis A: General 158, 185199. of methane using RhSr- and Ni-substituted hexaaluminates. Proceedings
Ohya, H., Fun, J., Kawamura, H., Itoh, K., Ohashi, H., Aihara, M., Tanisho, of the Combustion Institute, in press.
S., Negishi, Y., 1997. Methanation of carbon dioxide by using membrane
reactor integrated with water vapor permselective membrane and its
analysis. Journal of Membrane Science 131, 237247.

View publication stats

You might also like