You are on page 1of 24

processes

Article
Computational Fluid Dynamics Modeling of the
Catalytic Partial Oxidation of Methane in
Microchannel Reactors for Synthesis Gas Production
Junjie Chen * ID
, Wenya Song ID
and Deguang Xu ID

Department of Energy and Power Engineering, School of Mechanical and Power Engineering,
Henan Polytechnic University, Jiaozuo 454000, China; wentpj@163.com (W.S.); gaotpj@163.com (D.X.)
* Correspondence: comcjj@163.com or cjj@hpu.edu.cn; Tel.: +86-15138057627

Received: 27 May 2018; Accepted: 29 June 2018; Published: 30 June 2018 

Abstract: This paper addresses the issues related to the favorable operating conditions for the
small-scale production of synthesis gas from the catalytic partial oxidation of methane over rhodium.
Numerical simulations were performed by means of computational fluid dynamics to explore the
key factors influencing the yield of synthesis gas. The effect of mixture composition, pressure,
preheating temperature, and reactor dimension was evaluated to identify conditions that favor a
high yield of synthesis gas. The relative importance of heterogeneous and homogenous reaction
pathways in determining the distribution of reaction products was investigated. The results indicated
that there is competition between the partial and total oxidation reactions occurring in the system,
which is responsible for the distribution of reaction products. The contribution of heterogeneous and
homogeneous reaction pathways depends upon process conditions. The temperature and pressure
play an important role in determining the fuel conversion and the synthesis gas yield. Undesired
homogeneous reactions are favored in large reactors, and at high temperatures and pressures,
whereas desired heterogeneous reactions are favored in small reactors, and at low temperatures and
pressures. At atmospheric pressure, the selectivity to synthesis gas is higher than 98% at preheating
temperatures above 900 K when oxygen is used as the oxidant. At pressures below 1.0 MPa, alteration
of the dimension in the range of 0.3 and 1.5 mm does not result in significant difference in reactor
performance, if made at constant inlet flow velocities. Air shows great promise as the oxidant,
especially at industrially relevant pressure 3.0 MPa, thereby effectively inhibiting the initiation of
undesired homogeneous reactions.

Keywords: catalytic microreactors; synthesis gas production; partial oxidation; microchannel reactors;
reactor design; reaction pathway; hydrogen production; computational fluid dynamics

1. Introduction
There has been an increasing interest in the production of synthesis gas through the catalytic
partial oxidation of methane [1,2], due to its potential applications in many fields such as fuel cells [3,4]
and gas turbines [5,6]. Currently, the primary techniques used in industry to produce synthesis gas
from methane are steam reforming [7,8], autothermal reforming [9], and partial oxidation [7]. Steam
reforming of methane remains the main commercial process for the production of synthesis gas [7,8].
It is important to develop new reaction routes for the production of synthesis gas from methane.
A promising reaction route that has received much attention recently is the catalytic partial oxidation
of methane in short contact time reactors in high temperature environment [10–12], where a high
yield of synthesis gas (higher than 90%) can be achieved [13]. In comparison with other synthesis gas
production routes, this technology shows great promise because higher selectivity and better efficiency
can be achieved [14].

Processes 2018, 6, 83; doi:10.3390/pr6070083 www.mdpi.com/journal/processes


Processes 2018, 6, 83 2 of 24

Significant progress has been made recently in the understanding of the mechanism of this
reaction [15,16]. The reaction proceeds at a lower temperature than the gas-phase partial oxidation
route [17–20], thus offering many advantages such as a reduction of undesired by-products and
the ability to control the process temperature [21–24]. The catalysts used for this reaction usually
contain group VIII transition metals, such as rhodium [25,26], ruthenium [27,28], platinum [29,30],
palladium [31,32], nickel [33,34], iridium [35,36], and cobalt [37,38]. The reaction can proceed in short
contact time reactors under high-temperature conditions [39–41]. This technology provides a novel
route for the production of synthesis gas from methane [1,2], since the yield of the desired products
can be greatly improved under controlled conditions [42,43].
It is important to understand the mechanism of the catalytic partial oxidation reaction to
improve the yield of synthesis gas [44,45]. However, the pathways for the reaction are still debated,
and the prospect for industrial applications is not yet clear. In the case of noble metal catalysts,
both heterogeneous and homogeneous reaction pathways may be significant [46,47]. In addition,
homogeneous reactions decrease the yield of synthesis gas, thus seriously hindering reactor performance.
It is therefore necessary to understand the competition phenomenon between the two reaction pathways
during a catalytic partial oxidation process. Unfortunately, the relative contribution of the two reaction
pathways to the formation of synthesis gas has not yet been addressed, and thus further research is
needed to clarify the mechanism responsible for improving the yield of synthesis gas in a catalytic
partial oxidation system.
Microreactor technology is expected to offer many advantages for process development [48,49],
especially for fast, exothermic reactions such as catalytic partial oxidation. Precise temperature control
is possible, thus significantly reducing the undesirable side reactions occurring during a catalytic
partial oxidation process [48,49]. Furthermore, higher yields can be achieved for these processes under
well-controlled conditions, by taking advantage of enhanced transport in small dimensions [50,51].
For partial oxidation micro-chemical systems, one of the engineering design challenge is to balance
the gains made in heat and mass transfer by going to smaller dimensions against the increases in
pressure drop [52,53]. However, microfabrication methods have the potential to realize reactor designs
that combine excellent thermal uniformity, enhanced transport rates, and low-pressure drop for a
catalytic partial oxidation process [54,55]. The intrinsic kinetics of these partial oxidation processes are
typically very fast, and thus the realization of this process technology will require continued advances
in the development of microreactors [15,16]. The small dimensions associated with these reactors can
effectively inhibit the gas-phase reaction occurring during a catalytic partial oxidation process [56].
It is therefore important to determine the favorable operating conditions under which the yield of
desired products can be maximized.
In addition to providing an explanation of experimental data, numerical simulations can serve as
an efficiency design tool for the development of a catalytic partial oxidation system [56,57]. Iterative,
costly experimental design processes can be avoided. Furthermore, numerical simulations are necessary
to better understand the operating characteristics of a micro-structured device required to implement
a catalytic partial oxidation process, and to evaluate the disadvantages and benefits associated with
an innovative design of the process [56,57]. A number of commercial software tools are available, but,
unfortunately, none of them is universally applicable. To accurately predict the operating characteristics
during the process of a catalytic partial oxidation reaction occurring in microreactors and accurately
reflect experimental observations, detailed mathematical modeling is often necessary [58]. Detailed
computational fluid dynamics modeling can be used to evaluate design changes, such as geometric
parameters, Reynolds numbers, and reaction temperature, during a catalytic partial oxidation process [57].
The main focus of this paper is on determining the favorable operating conditions for the
small-scale production of synthesis gas from the catalytic partial oxidation of methane. High transport
rates are possible in microchannel reactors, thus allowing the catalytic partial oxidation reaction to be
carried out under more favorable conditions. Computational fluid dynamics simulations serve as a
means to understand the role of heterogeneous and homogenous reaction pathways in determining
Processes 2018, 6, x FOR PEER REVIEW 3 of 24

Processes 2018, 6, 83 3 of 24
means to understand the role of heterogeneous and homogenous reaction pathways in determining
the distribution of reaction products. The effect of reactor dimension, pressure, mixture composition,
the distribution
and of reaction products.
preheating temperature The effectto
was investigated of better
reactorunderstand
dimension,the
pressure, mixture
operating composition,
characteristics of
and preheating temperature was investigated to better understand the operating characteristics
the partial oxidation reactor. The favorable conditions for the production of synthesis gas were of
the partial oxidation
determined. The major reactor. Theisfavorable
objective conditions
to understand the for the production
relative importanceofofsynthesis
differentgas were
reaction
determined.inThe
pathways major objective
determining is to understand
the distribution the relative
of reaction importance
products. of different
Special reaction
emphasis pathways
is placed on
in determining the distribution of reaction products. Special emphasis is placed on identifying
identifying favorable operating conditions for the production of synthesis gas in high temperature
favorable operating conditions for the production of synthesis gas in high temperature environments.
environments.

2. Model
2. Model Development
Development

2.1. Reaction System


2.1. Reaction System
The reaction system used in the present investigation is the catalytic partial oxidation of methane
The reaction system used in the present investigation is the catalytic partial oxidation of methane
taking place in a microchannel reactor. For microchannel reactors, the width of each of the channels is
taking place in a microchannel reactor. For microchannel reactors, the width of each of the channels
about one order of magnitude larger than its height [59,60], and thus the reactor used in this paper is
is about one order of magnitude larger than its height [59,60], and thus the reactor used in this paper
modeled as a two-dimensional system. A premixed methane and either air or oxygen mixture is fed
is modeled as a two-dimensional system. A premixed methane and either air or oxygen mixture is
to the reactor, after the reactant stream is preheated to a desired temperature. The reactor contains
fed to the reactor, after the reactant stream is preheated to a desired temperature. The reactor contains
multiple parallel channels having sub-millimeter dimensions, thus offering advantages from enhanced
multiple parallel channels having sub-millimeter dimensions, thus offering advantages from
heat and mass transfer [61]. The reactor modeled in this paper is shown schematically in Figure 1.
enhanced heat and mass transfer [61]. The reactor modeled in this paper is shown schematically in
Unless otherwise specified, the reactor consists of two infinitely wide parallel plates of length 8.0 mm,
Figure 1. Unless otherwise specified, the reactor consists of two infinitely wide parallel plates of
separated by a gap distance 0.8 mm between the two plates.
length 8.0 mm, separated by a gap distance 0.8 mm between the two plates.

Figure 1. Two-dimensional schematic diagram of the microchannel reactor geometry used in


Figure 1. Two-dimensional schematic diagram of the microchannel reactor geometry used in
computational fluid dynamics.
computational fluid dynamics.

The physical properties the walls are the same as those of stainless steel. The rhodium catalyzed
The
partial physicalof
oxidation properties
methane the walls are the
is considered in same as those
the present of stainless
work, steel.
since the The rhodium
catalyst has been catalyzed
reported
partial oxidation of methane is considered in the present work, since the catalyst has
to give a high synthesis gas yield with good long-term stability [25,26]. Additionally, a “base been reported
case”,
to give a high synthesis gas yield with good long-term stability [25,26]. Additionally,
where typical operating conditions and design parameters are considered for the catalytic partial a “base case”,
where typical operating conditions and design parameters are considered for the catalytic
oxidation process, is given in Table 1. In this context, the effect of various operating conditions and partial
oxidation
design process, can
parameters is given in Table
be easily 1. In this
evaluated. context, the
Numerical effect of are
simulations various operating
carried out, andconditions and
the difference
Processes 2018, 6, 83 4 of 24

design parameters can be easily evaluated. Numerical simulations are carried out, and the difference
between a methane–oxygen system and a methane–air system is also investigated. Note that most of
the catalyst properties listed in Table 1 are taken from the works related to the reaction mechanism
used; the reaction mechanism used in this paper will be described in detail in Section 2.3. Reaction
mechanisms. The wash coat thickness is specified based on the reaction system considered.

Table 1. Nominal values of the operating conditions and design parameters used for the base case.

Parameter Variable Value


Geometry
Channel length l 8.0 mm
Channel height d 0.8 mm
Solid wall
Thickness δ 0.8 mm
Thermal conductivity λs 16 W/(m·K) (300 K)
Gas phase
Inlet methane-to-oxygen molar ratio φ 2.0
Inlet pressure pin 0.1 MPa
Inlet temperature Tin 300 K
Inlet velocity uin 0.8 m/s
Catalyst
Washcoat thickness δcatalyst 0.08 mm
Mean pore diameter dpore 20 nm
Porosity εp 0.5
Tortuosity factor τp 3
Catalyst/geometric surface area Fcat /geo 8
Density of rhodium surface sites Γ 2.72 × 10−9 mol/cm2
Other conditions
Ambient temperature Tamb 300 K
Surface emissivity ε 0.8
External heat loss coefficient ho 20 W/(m2 ·K)

2.2. Mathematical Model


Since the Reynolds number is less than 380, which implies laminar flow. Therefore, it is possible to
fully characterize heat and mass transfer in the system. Detailed reaction mechanisms are necessary to
accurately predict the partial oxidation process. A two-dimensional numerical model is developed by
using the commercial software ANSYS Fluent® Release 16.0 (ANSYS Inc., Canonsburg, PA, USA) [62].
Detailed reaction mechanisms are handled with related external procedures; please refer to Section 2.3.
Reaction mechanisms for more details. The steady-state two-dimensional conservation equations are
solved in the gas phase:
∂(ρu) ∂(ρv)
+ =0 (1)
∂x ∂y
     
∂(ρuu) ∂(ρvu) ∂p ∂ ∂u 2 ∂u ∂v ∂ ∂u ∂v
+ + − 2µ − µ + − µ + =0 (2)
∂x ∂y ∂x ∂x ∂x 3 ∂x ∂y ∂y ∂y ∂x
     
∂(ρuv) ∂(ρvv) ∂p ∂ ∂v ∂u ∂ ∂v 2 ∂u ∂v
+ + − µ + − 2µ − µ + =0 (3)
∂x ∂y ∂y ∂x ∂x ∂y ∂y ∂y 3 ∂x ∂y
Kg Kg
! !
∂(ρuh) ∂(ρvh) ∂ ∂T ∂ ∂T
∂x k∑ ∂y k∑
+ + ρ Yk hk Vk,x − λ g + ρ Yk hk Vk,y − λ g =0 (4)
∂x ∂y =1
∂x =1
∂y
Processes 2018, 6, 83 5 of 24

∂(ρuYk ) ∂(ρvYk ) ∂ ∂   .
+ + (ρYk Vk,x ) + ρYk Vk,y − ω k Wk = 0, k = 1, . . . , K g . (5)
∂x ∂y ∂x ∂y
The diffusion velocity vector is given as follows [63]:
" !# " #
→ Yk W DkT W
V k = − Dk,m ∇ ln + ∇(ln T ) (6)
Wk ρYk W

The ideal gas equation of state is given by

ρRT
p= (7)
W
The caloric equation of state is given by
Z T
hk = hok ( To ) + c p,k dT. (8)
To

The coverage equation of surface species can be expressed as


.
sm
ϑm = 0, m = K g + 1, . . . , K g + Ks . (9)
Γ
The steady-state energy equation in the walls is given by
   
∂ ∂T ∂ ∂T
λs + λs =0 (10)
∂x ∂x ∂y ∂y

The gaseous species equation at each of the fluid-washcoat interfaces is specified by the boundary
condition taken the form
. 
 
ρYk Vk,y + ηFcat/geo Wk sk inter f ace = 0, k = 1, . . . , K g . (11)
inter f ace

The catalyst/geometric surface area, Fcat / geo , is defined as follows [56]:

A0 catalyst
Fcat/geo = (12)
A0 geometric

The effect of diffusional limitation in the catalyst wash coat may be significant [64], and thus
included in the model .
si,e f f tanh(Φ)
η= . = (13)
si Φ
. !0.5
si γ
Φ = δcatalyst (14)
Di,e f f Ci,inter f ace

Fcat/geo
γ= . (15)
δcatalyst
The effective diffusivity can be written as
 
1 τp 1 1
= + (16)
Di,e f f εp Di,molecular Di,Knudsen
Processes 2018, 6, 83 6 of 24

The Knudsen diffusivity is defined as


s
d pore 8RT
Di,Knudsen = . (17)
3 πWi

The energy equation at each of the fluid wash coat interfaces is specified by the boundary condition
taken the form
    Kg
. ∂T ∂T .


qrad − λ g + λs + sk hk Wk inter f ace
= 0. (18)
∂y inter f ace− ∂y inter f ace+ k =1

For the total heat loss to the surroundings, the equation can be written as
 
4 4
q = ho ( Tw,o − Tamb ) + εFs−∞ σ Tw,o − Tamb (19)

The external heat loss coefficient, ho , is assumed to be 20 W/(m2 ·K) [65].

2.3. Reaction Mechanisms


Computational fluid dynamics modeling of the catalytic partial oxidation process is
complex [15,16], especially when the role of reaction pathway needs to be determined [66]. Therefore,
detailed reaction mechanisms are included in the model. The possible reactions involved in the
catalytic partial oxidation process are listed as follows:
Partial oxidation

1 Θ
CH4 + O2 → CO + 2H2 , ∆r Hm (298.15 K) = −35.7 kJ · mol−1 (20)
2
Θ
CH4 + O2 → CO + H2 + H2 O, ∆r Hm (298.15 K) = −278 kJ · mol−1 (21)

Total oxidation

Θ
CH4 + 2O2 → CO2 + 2H2 O, ∆r Hm (298.15 K) = −802.3 kJ · mol−1 (22)

Competition between the two reaction routes is responsible for the distribution of reaction
products. However, the highly exothermic total oxidation reaction serves as a heat source to ensure
self-sustained operation of the system. Therefore, this reaction route can improve the selectivity
towards the desired products. However, the products produced by this route decrease the selectivity.
Overall, the final yield of synthesis gas is determined by this competitive effect.
Steam reforming

Θ
CH4 + H2 O → CO + 3H2 , ∆r Hm (298.15 K) = +206.2 kJ · mol−1 (23)

Water–gas shift reaction

Θ
CO + H2 O → CO2 + H2 , ∆r Hm (298.15 K) = −41.2 kJ · mol−1 , (24)

Dry reforming

Θ
CH4 + CO2 → 2CO + 2H2 , ∆r Hm (298.15 K) = 246.9 kJ · mol−1 (25)

The reaction mechanism has attracted increasing attention recently [15,16,67–69]. The reaction may
proceed through a combination of direct partial oxidation and steam reforming [39–41]. Both partial
and total oxidation products can be formed [40], and carbon dioxide has little or no role during the
Processes 2018, 6, 83 7 of 24

process [39,41]. More importantly, both water-gas shift and carbon dioxide reforming do not contribute
to the formation of synthesis gas [39,41].
The detailed heterogeneous reaction mechanism developed by Schwiedernoch et al. [70], as given
in Table 2, is included in the model. The mechanism consists of 11 surface-adsorbed species and 6
gaseous species involved in 38-step elementary reactions. Since each of the reactive intermediates
and elementary reaction steps involved in the catalytic partial oxidation process is included in the
reaction mechanism, the global reactions such as steam and carbon dioxide reforming are automatically
accounted for [70]. Note that the symbol * used in Table 2 denotes an adsorbed species or an empty site.

Table 2. Heterogeneous reaction mechanism used for the partial oxidation of methane over rhodium.

A (cm, mol,
Reactions s Ea (kJ/mol)
s)
Adsorption
H2 + * + * => H * + H * 1.0 × 10−2
O2 + * + * => O * + O * 1.0 × 10−2
CH4 + * => CH4 * 8.0 × 10−3
H2 O + * => H2 O * 1.0 × 10−1
CO2 + * => CO2 * 1.0 × 10−5
CO + * => CO * 5.0 × 10−1
Desorption
H * + H * => * + * + H2 3.0 × 1021 77.8
O * + O * => * + * + O2 1.3 × 1022 355.2–280ΘO*
H2 O * => H2 O + * 3.0 × 1013 45.0
CO * => CO + * 3.5 × 1013 133.4–15ΘCO*
CO2 * => CO2 + * 1.0 × 1013 21.7
CH4 * => CH4 + * 1.0 × 1013 25.1
Surface reactions
H * + O * => OH * + * 5.0 × 1022 83.7
OH * + * => H * + O * 3.0 × 1020 37.7
H * + OH * => H2 O * + * 3.0 × 1020 33.5
H2 O * + * => H * + OH * 5.0 × 1022 104.7
OH * + OH * => H2 O * + O * 3.0 × 1021 100.8
H2 O * + O * => OH * + OH * 3.0 × 1021 171.8
C * + O * => CO * + * 3.0 × 1022 97.9
CO * + * => C * + O * 2.5 × 1021 169.0
CO * + O * => CO2 * + * 1.4 × 1020 121.6
CO2 * + * => CO * + O * 3.0 × 1021 115.3
CH4 * + * => CH3 * + H * 3.7 × 1021 61.0
CH3 * + H * => CH4 * + * 3.7 × 1021 51.0
CH3 * + * => CH2 * + H * 3.7 × 1024 103.0
CH2 * + H * => CH3 * + * 3.7 × 1021 44.0
CH2 * + * => CH * + H * 3.7 × 1024 100.0
CH * + H * => CH2 * + * 3.7 × 1021 68.0
CH * + * => C * + H * 3.7 × 1021 21.0
C * + H * => CH * + * 3.7 × 1021 172.8
CH4 * + O * => CH3 * + OH * 1.7 × 1024 80.3
CH3 * + OH * => CH4 * + O * 3.7 × 1021 24.3
CH3 * + O * => CH2 * + OH * 3.7 × 1024 120.3
CH2 * + OH * => CH3 * + O * 3.7 × 1021 15.1
CH2 * + O * => CH * + OH * 3.7 × 1024 158.4
CH * + OH * => CH2 * + O * 3.7 × 1021 36.8
CH * + O * => C * + OH * 3.7 × 1021 30.1
C * + OH * => CH * + O * 3.7 × 1021 145.5
Processes 2018, 6, x FOR PEER REVIEW 8 of 24

Processes 2018, 6, 83 8 of 24
C * + OH * => CH * + O * 3.7 × 1021 145.5

Leedsmethane
The Leeds methaneoxidation
oxidation mechanism
mechanism [71,72]
[71,72] is included
is included in thein the model
model to describe
to describe the
the reaction
reaction
taking taking
place place
in the gasinphase.
the gas phase.
The The mechanism
reaction reaction mechanism is also accounted
is also accounted forradical
for the free the free radical
coupling
coupling such
reactions reactions such as coupling
as oxidative oxidativeof
coupling
methane. of The
methane. The selectivity
selectivity to C2-hydrocarbons
to C2 -hydrocarbons may be as mayhigh be
as
as high
20% as 20%
for the for the
reaction reactionunder
proceeded proceeded
certainunder certainthrough
conditions, conditions, through reaction
a gas-phase a gas-phase
routereaction
[73,74].
route [73,74].
These These C2-hydrocarbons
C2 -hydrocarbons areduring
are undesirable undesirable during the
the catalytic catalytic
partial partial
oxidation oxidation
process, process,
as they can
as theythe
cause can cause the
problem problem
related related
to the to theofformation
formation coke. of coke.
CHEMKIN transport
The CHEMKIN transport database
database [63]
[63] isis used
used inin the
the model.
model. The homogeneous and
heterogeneous reaction rates are handled through the CHEMKIN [75] and Surface-CHEMKIN [76]
respectively.
interfaces, respectively.

2.4.
2.4. Computation
Computation Scheme
Scheme
An orthogonal staggered
An orthogonal staggered gridgrid isisused
usedfor forthe
thebase
basecase,
case,consisting
consisting of of
200200 axial
axial nodes
nodes by by
80
80 transverse nodes. Typical fluid node spacing near the catalyst wash coat
transverse nodes. Typical fluid node spacing near the catalyst wash coat is 40 μm in the axial direction is 40 µm in the axial
direction
and 5 μmand 5 µm
in the in the transverse
transverse direction. direction. For thereactor
For the largest largest dimension,
reactor dimension, a grid consisting
a grid consisting of
of 20,000
20,000
nodes in nodes
totalinis total is utilized.
utilized. Adequate Adequate grid resolution
grid resolution is verified
is verified by doubling
by doubling the number
the number of grid
of grid points.
points. Figure 2 shows the profiles of the hydroxyl radical concentration along
Figure 2 shows the profiles of the hydroxyl radical concentration along the centerline between the the centerline between
the
twotwo parallel
parallel plates
plates for for
somesome of the
of the grids
grids usedusedforfor
thethe methane–oxygen
methane–oxygen system.
system. The The inlet
inlet pressure
pressure is
is
3.03.0 MPa.
MPa. TheTherestrest of the
of the parameters
parameters usedused
herehere are listed
are listed in Table
in Table 1. As1.the
Asgrid
thedensity
grid density increases,
increases, there
there is a convergence
is a convergence of theof the solution.
solution. The numerical
The numerical modelmodel
with thewith the coarsest
coarsest grid, consisting
grid, consisting of 4000ofnodes
4000
nodes
in total,incannot
total, cannot accurately
accurately capture capture the hydroxyl
the hydroxyl radicalradical concentration
concentration withinwithin the channel
the channel and itsand its
peak
peak location, and thus fails to accurately predict the onset of gas-phase combustion
location, and thus fails to accurately predict the onset of gas-phase combustion within the system. In within the system.
In contrast,
contrast, thethe solution
solution obtained
obtained byby thethe numerical
numerical model
model with
with a grid
a grid consisting
consisting of of
tenstens of thousands
of thousands of
of nodes
nodes is reasonably
is reasonably accuratefor
accurate forthe
thebase
basecase
casegiven
givenininTable
Table1.1.There
Thereisislittle
littleor
or no
no advantage
advantage for the
largest grid density,
density, upup toto 32,000
32,000nodes
nodesin intotal.
total.

0.009
(16000 nodes are used in
Hydroxyl radical mass fraction

this paper for the base case)

4000 nodes
0.006
8000 nodes
16000 nodes
32000 nodes

0.003

0.000
0 2 4 6 8
Axial distance (mm)

Figure 2.
Figure 2. Profiles
Profiles of
of the
the hydroxyl
hydroxyl radical
radical concentration
concentration along
along the
the centerline
centerline between
between the
the two
two parallel
parallel
plates for some of the grids used for the methane–oxygen system. The inlet pressure is 3.0 MPa. The The
plates for some of the grids used for the methane–oxygen system. The inlet pressure is 3.0 MPa. rest
rest
of ofparameters
the the parametersusedused
here here are listed
are listed in Table
in Table 1. 1.

The physical properties of the mixture depend on the local conditions of component and
The physical properties of the mixture depend on the local conditions of component and
temperature. The physical properties the walls, such as the thermal conductivity and specific heat
temperature. The physical properties the walls, such as the thermal conductivity and specific heat
capacity, depend on the local temperature. The conservation equations are discretized by using a
capacity, depend on the local temperature. The conservation equations are discretized by using a
finite-volume method. The momentum, energy, and species equations are discretized by using a two-
finite-volume method. The momentum, energy, and species equations are discretized by using a
order upwind approximation. The pressure-velocity coupling is discretized using the “SIMPLE”
two-order upwind approximation. The pressure-velocity coupling is discretized using the “SIMPLE”
method. The convergence criterion is 10−6 by examining the values of the residuals for all of the
method. The convergence criterion is 10−6 by examining the values of the residuals for all of the
conservation equations. Convergence of the solution is usually difficult due to the inherent stiffness
conservation equations. Convergence of the solution is usually difficult due to the inherent stiffness of
Processes 2018, 6, 83 9 of 24
Processes 2018, 6, x FOR PEER REVIEW 9 of 24

the detailed
of the detailed reaction
reaction mechanism
mechanism used.
used.Figure
Figure3 3shows
showsthe
theresiduals
residualsfor
forthe
theconservation
conservation equations
at the end of each solver iteration. The The residual
residual plot
plot indicates
indicates that
that after
after approximately
approximately 800
800 iterations,
iterations,
the convergence
Processes 2018, 6, criterion
x FOR PEERis satisfied.
satisfied.
REVIEW 9 of 24

of the detailed reaction mechanism used. Figure 3 shows the residuals for the conservation equations
at the end of each solver iteration. The residual plot indicates that after approximately 800 iterations,
the convergence criterion is satisfied.

Figure 3.
Figure 3. Residuals
Residuals for
for the
the conservation
conservation equations
equations at
at the
the end
end of
of each
each solver
solver iteration.
iteration. The
The mesh
mesh used
used
here consists of 16,000 nodes in total. The parameters used in the numerical simulations
here consists of 16,000 nodes in total. The parameters used in the numerical simulations conducted conducted
here Figure
here are the
are the3.same
same as those
those
Residuals
as adopted
for the in Figure
Figure
conservation
adopted in 2.
equations
2. at the end of each solver iteration. The mesh used
here consists of 16,000 nodes in total. The parameters used in the numerical simulations conducted
Modelhere are the same as those adopted in Figure 2.
2.5. Model Validation
2.5. Validation
In order
2.5.
In order
Modelto to verify the
Validation
verify the model,
model, the the experimental
experimental results results reported
reported in in the
the literature
literature [77]
[77] are
are utilized.
utilized.
The reactor
The reactor used
used is
In order
is made
to made
verify of
of a
theamodel,
quartz
quartzthe tube, tube, and the
and the inside
experimental
inside
results diameter diameter
reportedof inthe
of
the reactor
the
literature
reactor
is [77]
18 mm, isas18described
are utilized.
mm, as
described
in theThe in
literaturethe literature
reactor [77].
used The [77].
inlet of
is made The inlet
temperature temperature
a quartz tube, of a and
methane of a methane
and oxygen
the inside diameter and
mixtureoxygen
of the is mixture

20 C,isand
reactor is 20 °C,
as and
the pressure
18 mm,
themaintained
is pressure
describedisin atmaintained
the MPa atat[77].
0.12literature 0.12
the MPa
The
outlet. at temperature
inlet the outlet.
Nitrogen Nitrogen
of a methane
dilution is 30%, dilution
and the
and is inlet
oxygen 30%, and rate
mixture
flow the inlet
is 20is°C, flow
and rate
maintained
is maintained
the pressure
3 atis5000 cm
maintained3/min. A grid consisting of 36,000 nodes in total is used here. Numerical
at 0.12 MPa at the outlet. Nitrogen dilution
at 5000 cm /min. A grid consisting of 36,000 nodes in total is used here. Numerical simulations areis 30%, and the inlet flow rate
simulations
performed forare
is maintained theperformed
experimental forcase
at 5000 cm3/min. the experimental
Atested
grid consisting
under theof case tested
36,000
operating nodesunder
in total
conditions theand
isoperating
used here.
design conditions
Numericalgiven
parameters and
designsimulations
parameters are performed for the experimental case tested under the operating conditions and
in the literature [77].given in the literature
The results obtained for [77].the
The results obtained
selectivity for theconversion
and the outlet selectivityforand thethe outlet
mixture
design parameters
conversion for the given inwith
mixture the literature
different [77]. The results obtained
compositions is for theto
compared selectivity
the and the outlet
experimental data in
with different compositions is compared to the experimental data in Figure 4. The maximum difference
conversion for the mixture with different compositions is compared to the experimental data in
Figure 4. the
between Thenumerical
maximum results difference andbetween
the the numerical
experimental data results
is about and5.6%.
the experimental
Therefore, thedata is about
numerical
Figure 4. The maximum difference between the numerical results and the experimental data is about
5.6%. Therefore, the numerical results
results are in consistent with the experimental data. are in consistent with the experimental data.
5.6%. Therefore, the numerical results are in consistent with the experimental data.
100100 100
Lines:numerical
Lines: numerical results 100
results
Outlet methane conversion (%)

Symbols:experimental
experimental data
Outlet methane conversion (%)

Symbols: data
95 95 95 95
Selectivity (%)
Selectivity (%)

90 90 90 90

85 Hydrogen selectivity 85
85 Hydrogen selectivity 85
Carbon monoxide selectivity
Carbon monoxide selectivity
Outlet methane conversion
80 Outlet methane conversion 80
1.5 1.6 1.7 1.8 1.9 2.0 2.1
80 80
1.5 1.6 1.7 1.8
Inlet methane-to-oxygen 1.9 ratio2.0
molar 2.1
Inlet methane-to-oxygen molar ratio
Figure 4. Comparison between the numerical results and the experimental data obtained for a
Figure 4. Comparison between the numerical results and the experimental data obtained for a methane
methane
Figure and oxygen between
4. Comparison mixture with various compositions. The experimental data aredata
taken from the for a
and oxygen mixture with variousthe numerical
compositions. results and the experimental
The experimental data are taken obtained
from the previous
previous work of Bodke et al. [77]. A grid consisting of 36,000 nodes in total is used here.
methane
work and oxygen
of Bodke mixture
et al. [77]. with
A grid variousof
consisting compositions.
36,000 nodesThe experimental
in total data are taken from the
is used here.
previous work of Bodke et al. [77]. A grid consisting of 36,000 nodes in total is used here.
Processes 2018, 6, 83 10 of 24

3. Results and Discussion


Processes 2018, 6, x FOR PEER REVIEW 10 of 24
In the following sections, the reactor performance in terms of conversion, selectivity,
3. Results and
and temperature Discussion
under various operating conditions is discussed in detail, and the relative importance
of different
In reaction pathways
the following in determining
sections, the distribution
the reactor performance of reaction
in terms products
of conversion, is investigated.
selectivity, and
Additionally, comparisons
temperature areoperating
under various made in terms of reactor
conditions performance
is discussed between
in detail, and the results
the relative obtained
importance of for
different reaction
a methane–oxygen pathways
system in determining
and for a methane–airthe system.
distribution of reaction products is investigated.
Additionally, comparisons are made in terms of reactor performance between the results obtained
3.1. Base
for aCase
methane–oxygen system and for a methane–air system.

For the base


3.1. Base Case case given in Table 1, a methane-to-oxygen molar ratio of 2.0, i.e., a stoichiometric
mixture for the production of synthesis gas, is used. This ratio is ideal for the downstream
For the base case given in Table 1, a methane-to-oxygen molar ratio of 2.0, i.e., a stoichiometric
processing such as in the synthesis of methanol and in the production of Fischer-Tropsch products.
mixture for the production of synthesis gas, is used. This ratio is ideal for the downstream processing
The temperature is set as 300 K at the inlet for the base case, thus effectively avoiding gas-phase
such as in the synthesis of methanol and in the production of Fischer-Tropsch products. The
combustion; please
temperature is refer
set asto300theKSection
at the 4. Further
inlet Discussion
for the base case,forthus
more details. avoiding gas-phase
effectively
Figure 5 shows contour plots of the methane and carbon
combustion; please refer to the Section 4. Further Discussion for more details. monoxide concentrations and
temperature within the fluid in the methane–oxygen system.
Figure 5 shows contour plots of the methane and carbon monoxide Despite the small dimension involved,
concentrations and
the temperature
temperatureand species
within gradients
the fluid change significantly
in the methane–oxygen in the
system. vicinity
Despite the of reaction
small region.
dimension This can be
involved,
the temperature
attributed and species
to the difference gradients
in the change
time scale significantly
between in the vicinity of
the heterogeneous reactionand
reaction region. This can
the heat transfer
in thebe attributeddirection.
transverse to the difference in the time
The significant scale in
change between the heterogeneous
both temperature reaction
and species and themay
gradients heat need
to usetransfer in the transverse
a two-dimensional direction. Thefluid
computational significant change
dynamics in both
model, astemperature andof
axial diffusion species gradients
species and energy
may need to use a two-dimensional computational fluid dynamics model, as axial diffusion of species
is not negligible. Furthermore, mass-transfer limitations, typical characteristics of a catalytic partial
and energy is not negligible. Furthermore, mass-transfer limitations, typical characteristics of a
oxidation reaction, are observed here, despite the small dimension involved. Accordingly, the effect of
catalytic partial oxidation reaction, are observed here, despite the small dimension involved.
transport phenomena
Accordingly, willof
the effect betransport
discussed in the following
phenomena sections.in the following sections.
will be discussed

Figure 5. Contour plots of the methane and carbon monoxide concentrations and temperature within
Figure 5. Contour plots of the methane and carbon monoxide concentrations and temperature within
the fluid in the methane–oxygen system. The operating conditions and design parameters used are
the fluid in the methane–oxygen system. The operating conditions and design parameters used are
listed in Table 1.
listed in Table 1.
Processes 2018,
Processes 6, x83FOR PEER REVIEW
2018, 6, 11 of
11 of 24
24

3.2. Effect of Preheating for Oxygen Feed


3.2. Effect of Preheating for Oxygen Feed
Figure 6 shows the influence of preheating temperature on the performance of the methane–
Figure 6 shows the influence of preheating temperature on the performance of the methane–oxygen
oxygen system. As the pressure increases, there is a sharp drop in the selectivity to synthesis gas
system. As the pressure increases, there is a sharp drop in the selectivity to synthesis gas (Figure 6a,b)
(Figure 6a,b) and conversion (Figure 6c), but a sharp rise in wall temperature (Figure 6d). This sharp
and conversion (Figure 6c), but a sharp rise in wall temperature (Figure 6d). This sharp drop implies
drop implies the initiation of the total oxidation reaction occurring in the gas phase. After the
the initiation of the total oxidation reaction occurring in the gas phase. After the initiation of gas-phase
initiation of gas-phase combustion, the contribution of heterogeneous reactions is still considerable,
combustion, the contribution of heterogeneous reactions is still considerable, as indicated by the
as indicated by the selectivity to synthesis gas at high pressures (Figure 6a,b), but there is lack of
selectivity to synthesis gas at high pressures (Figure 6a,b), but there is lack of oxygen for the catalytic
oxygen for the catalytic partial oxidation reaction.
partial oxidation reaction.

100 100
300 K 700 K
Carbon monoxide selectivity (%)

Oxygen feed
500 K 900 K

Hydrogen selectivity (%)


90 90
No
p

300 K
re
he

500 K
at

80 80
in

700 K
g

900 K

70 T 70
pre
h eati
ng : 90
0K Oxygen feed
60 60
0.1 0.5 1.0 1.5 2.0 2.5 3.0 0.1 0.5 1.0 1.5 2.0 2.5 3.0
Pressure (MPa) Pressure (MPa)
(a) (b)

100 1800
Oxygen feed
Maximum wall temperature (K)
Outlet methane conversion (%)

300 K
500 K
90 700 K 1600 300 K
900 K 500 K
700 K
80 1400 900 K

70 1200
Oxygen feed

60 1000
0.1 0.5 1.0 1.5 2.0 2.5 3.0 0.1 0.5 1.0 1.5 2.0 2.5 3.0
Pressure (MPa) Pressure (MPa)
(c) (d)
Figure 6. Influence of preheating temperature on the selectivity, conversion, and maximum wall
Figure 6. Influence of preheating temperature on the selectivity, conversion, and maximum wall
temperature in the methane–oxygen system. Hereafter, all other parameters are kept at their base case
temperature in the methane–oxygen system. Hereafter, all other parameters are kept at their base
values shown in Table 1. The sharp drop in selectivity and conversion indicates the initiation of gas-
case values shown in Table 1. The sharp drop in selectivity and conversion indicates the initiation
phase combustion. (a) Selectivity to carbon monoxide; (b) selectivity to hydrogen; (c) outlet
of gas-phase combustion. (a) Selectivity to carbon monoxide; (b) selectivity to hydrogen; (c) outlet
conversion;
conversion; (d)
(d) maximum
maximum wall
wall temperature.
temperature.

At atmospheric pressure, the selectivity to synthesis gas (Figure 6a,b) and the conversion (Figure
At atmospheric pressure, the selectivity to synthesis gas (Figure 6a,b) and the conversion
6c) increase with increasing preheating temperature. Very high selectivity to synthesis gas (>98%) is
(Figure 6c) increase with increasing preheating temperature. Very high selectivity to synthesis
possible at atmospheric pressure when the preheating temperature is above 900 K. At atmospheric
gas (>98%) is possible at atmospheric pressure when the preheating temperature is above 900 K.
pressure, the main product is synthesis gas, and the catalytic partial oxidation reaction is favored at
At atmospheric pressure, the main product is synthesis gas, and the catalytic partial oxidation
high temperatures. On the other hand, gas-phase combustion is favored at high temperatures, at
reaction is favored at high temperatures. On the other hand, gas-phase combustion is favored at
which the initiation of the combustion reaction is possible at lower pressures. For example, as the
high temperatures, at which the initiation of the combustion reaction is possible at lower pressures.
inlet temperature increases from 300 to 900 K, the initiation pressure decreases from about 2.5 to 0.7
For example, as the inlet temperature increases from 300 to 900 K, the initiation pressure decreases
MPa. At high pressures, the catalytic partial oxidation reaction is favored at low temperatures. The
from about 2.5 to 0.7 MPa. At high pressures, the catalytic partial oxidation reaction is favored at low
situation is the reverse of the results obtained at atmospheric pressure. In all of the cases examined
here, the selectivity to synthesis gas (Figure 6a,b) and the conversion (Figure 6c) decrease with
Processes 2018, 6, 83 12 of 24

Processes 2018, 6, x FOR PEER REVIEW 12 of 24


temperatures. The situation is the reverse of the results obtained at atmospheric pressure. In all of
the cases examined here, the selectivity to synthesis gas (Figure 6a,b) and the conversion (Figure 6c)
increasing pressure. After the initiation of gas-phase combustion, however, the pressure has little or
decrease with increasing pressure. After the initiation of gas-phase combustion, however, the pressure
no effect on the conversion. Furthermore, the loss in conversion at between atmospheric pressure and
has little or no effect on the conversion. Furthermore, the loss in conversion at between atmospheric
the highest pressure is almost the same.
pressure and the highest pressure is almost the same.
On the other hand, as the pressure increases, there is a transition of primary reaction pathway
On the other hand, as the pressure increases, there is a transition of primary reaction pathway
from catalytic partial oxidation to gas-phase combustion, as depicted by the inflection point in the
from catalytic partial oxidation to gas-phase combustion, as depicted by the inflection point in the
conversion profile (Figure 6c). After the reactants have been ignited in the gas phase, there is a
conversion profile (Figure 6c). After the reactants have been ignited in the gas phase, there is a
significant drop in conversion. This is due to lack of oxygen for the gas-phase combustion reaction,
significant drop in conversion. This is due to lack of oxygen for the gas-phase combustion reaction,
despite the fact that a stoichiometric methane–air mixture is used for the production of synthesis gas
despite the fact that a stoichiometric methane–air mixture is used for the production of synthesis gas
from the catalytic partial oxidation reaction. Please refer to the stoichiometric coefficients described
from the catalytic partial oxidation reaction. Please refer to the stoichiometric coefficients described in
in the chemical Equations (20)–(22) for more details.
the chemical Equations (20)–(22) for more details.
3.3. Effect of Preheating for Air Feed
3.3. Effect of Preheating for Air Feed
One of the major challenges during the catalytic partial oxidation process is a reduction of the
One of the major challenges during the catalytic partial oxidation process is a reduction of the cost
cost of pure oxygen separation [15,16], since the production of pure oxygen is highly expensive. The
of pure oxygen separation [15,16], since the production of pure oxygen is highly expensive. The reaction
reaction system in the presence of a catalyst can be operated at much milder conditions as compared
system in the presence of a catalyst can be operated at much milder conditions as compared to the
to the gas-phase partial oxidation system, which can effectively inhibit the formation of nitrogen
gas-phase partial oxidation system, which can effectively inhibit the formation of nitrogen oxides in
oxides in the gas phase, making it possible to use air instead of pure oxygen [15,16]. This shows great
the gas phase, making it possible to use air instead of pure oxygen [15,16]. This shows great promise
promise for the catalytic partial oxidation process.
for the catalytic partial oxidation process.
Figure 7 shows the effect of preheating temperature on the performance of the methane–air
Figure 7 shows the effect of preheating temperature on the performance of the methane–air
system. The pressure has a small effect on the selectivity to synthesis gas, but with a tendency to shift
system. The pressure has a small effect on the selectivity to synthesis gas, but with a tendency to shift
towards a higher yield of synthesis gas as the pressure decreases, as shown in Figure 7a,b
towards a higher yield of synthesis gas as the pressure decreases, as shown in Figure 7a,b Additionally,
Additionally, high preheating temperatures tend to improve the yield of synthesis gas. Therefore, the
high preheating temperatures tend to improve the yield of synthesis gas. Therefore, the production of
production of synthesis gas is favored at low pressures and high temperatures. In this context, the
synthesis gas is favored at low pressures and high temperatures. In this context, the outlet concentration
outlet concentration of carbon dioxide is very low (Figure 7a), but the small amount of carbon dioxide
of carbon dioxide is very low (Figure 7a), but the small amount of carbon dioxide must be removed to
must be removed to meet the downstream processing requirements [15,16]. The amount of the total
meet the downstream processing requirements [15,16]. The amount of the total oxidation products
oxidation products and C2-hydrocarbons increases with increasing pressure, leading to a decrease in
and C2 -hydrocarbons increases with increasing pressure, leading to a decrease in both conversion and
both conversion and the selectivity to synthesis gas. On the other hand, the conversion is favored at
the selectivity to synthesis gas. On the other hand, the conversion is favored at high temperatures and
high temperatures and at low pressures, as shown in Figure 7c. In contrast, the maximum wall
at low pressures, as shown in Figure 7c. In contrast, the maximum wall temperature increases with
temperature increases with increasing pressure, especially at high preheating temperatures (Figure
increasing pressure, especially at high preheating temperatures (Figure 7d) where the total oxidation
7d) where the total oxidation reaction is favored. In comparison with the results obtained for the
reaction is favored. In comparison with the results obtained for the methane–oxygen system (Figure 6),
methane–oxygen system (Figure 6), the initiation of gas-phase combustion is impossible for the
the initiation of gas-phase combustion is impossible for the methane–air system, and the contribution
methane–air system, and the contribution of homogeneous reactions is small.
of homogeneous reactions is small.

100 50 100 60
300 K 700 K
Carbon monoxide selectivity (%)

500 K 900 K
Carbon dioxide selectivity (%)

90 50
Hydrogen selectivity (%)

90 40
Water selectivity (%)

80 40
80 300 K 30
500 K Air feed 70 Air feed 30
700 K
70 900 K 20
60 20

60 10
50 10
300 K 700 K
500 K 900 K
50 0 40 0
0.1 0.5 1.0 1.5 2.0 2.5 3.0 0.1 0.5 1.0 1.5 2.0 2.5 3.0
Pressure (MPa) Pressure (MPa)
(a) (b)

Figure 7. Cont.
Processes 2018, 6, x FOR PEER REVIEW 13 of 24
Processes 2018, 6,
Processes 2018, 6, 83
x FOR PEER REVIEW 13
13 of
of 24
24

100 1400
100 300 K 1400

(K)(K)
(%)
300 K
500
(%)

temperature
90 500 K
700
conversion

1300

temperature
90 700 K
900
conversion

1300
900 K
80
80 1200
methane

wall
1200
methane

wall
70

Maximum
300 K
70

Maximum
1100 300 K
500
Outlet

Air feed
1100 500 K
Outlet

Air feed 700


60 Air feed 700 K
900
60 Air feed 900 K
1000
0.1 0.5 1.0 1.5 2.0 2.5 3.0 10000.1 0.5 1.0 1.5 2.0 2.5 3.0
0.1 0.5 1.0 Pressure
1.5 (MPa)2.0 2.5 3.0 0.1 0.5 1.0 Pressure
1.5 (MPa)2.0 2.5 3.0
Pressure (MPa) Pressure (MPa)
(c) (d)
(c) (d)
Figure 7. Effect of preheating temperature on the selectivity, conversion, and maximum wall
Figure 7. Effect of preheating temperature on the selectivity, conversion, and maximum wall
temperature in the methane–air system. (a) Selectivity to carbon monoxide and carbon dioxide; (b)
Figure 7. Effect
temperature in of
thepreheating temperature
methane–air system. on(a)the selectivity,
Selectivity toconversion, and maximum
carbon monoxide wall temperature
and carbon dioxide; (b)
selectivity to hydrogen and water; (c) outlet conversion; (d) maximum wall temperature.
in the methane–air system. (a) Selectivity to carbon monoxide and carbon dioxide;
selectivity to hydrogen and water; (c) outlet conversion; (d) maximum wall temperature. (b) selectivity to
hydrogen and water; (c) outlet conversion; (d) maximum wall temperature.
3.4. Effect of Reactor Dimension for Oxygen Feed
3.4. Effect of Reactor Dimension for Oxygen Feed
The reactor
3.4. Effect of Reactordimension
Dimensioncan for significantly
Oxygen Feed affect the effect of mass transfer [78]. For the system
The reactor dimension can significantly affect the effect of mass transfer [78]. For the system
examined here, it is unclear whether there is an optimal dimension in which the yield of synthesis
The reactor
examined here, itdimension
is unclear can significantly
whether there is affect the effect
an optimal of mass
dimension intransfer
which the[78]. Forofthe
yield system
synthesis
gas can be maximized. To provide a way to reduce the contribution of homogeneous reactions, the
examined
gas can behere, it is unclear
maximized. whetherathere
To provide way istoan optimal
reduce thedimension
contributionin which the yield of reactions,
of homogeneous synthesis gas
the
effect of reactor dimension is investigated.
can beof
effect maximized. To provide
reactor dimension a way to reduce the contribution of homogeneous reactions, the effect
is investigated.
Figure 8 presents the results obtained for the selectivity, conversion, and maximum wall
of reactor dimension
Figure 8 presents is investigated.
the results obtained for the selectivity, conversion, and maximum wall
temperature at different dimensions of the methane–oxygen system. The selectivity to synthesis gas
Figure 8atpresents
temperature differentthe results obtained
dimensions for the selectivity,
of the methane–oxygen conversion,
system. and maximum
The selectivity to synthesiswall
gas
is favored in smaller reactors, as shown in Figure 8a,b. The dimension has little effect on the reactor
temperature at different
is favored in smaller dimensions
reactors, as shown of in
theFigure
methane–oxygen system. has
8a,b. The dimension Thelittle
selectivity
effect ontothe
synthesis
reactor
performance at low pressures, where the system is operated in a surface kinetically-controlled regime
gas is favoredat
performance inlow
smaller reactors,
pressures, as shown
where in Figure
the system 8a,b. The
is operated indimension has little effect on the regime
a surface kinetically-controlled reactor
and the yield of synthesis gas is excellent. At high pressures, however, both mass-transfer limitations
performance at low pressures, where the system is operated in a surface kinetically-controlled
and the yield of synthesis gas is excellent. At high pressures, however, both mass-transfer limitations regime
and homogeneous reactions become significant. For the smallest reactor studied, the dominant
and the yield of synthesis
homogeneous gas is
reactions excellent.
become At high pressures,
significant. however,reactor
For the smallest both mass-transfer
studied, the limitations
dominant
chemistry is heterogeneous at all of the pressures examined. At the highest pressure examined, the
and homogeneous
chemistry reactionsatbecome
is heterogeneous all of thesignificant. For the smallest
pressures examined. reactor pressure
At the highest studied, examined,
the dominant the
selectivity to synthesis gas is quite high in the smallest reactor, but drops sharply in the largest
chemistry
selectivity isto heterogeneous
synthesis gas isatquiteall ofhigh
the pressures examined.
in the smallest Atbut
reactor, thedrops
highest pressure
sharply examined,
in the largest
reactor. Furthermore, the yield of synthesis gas is favored in smaller reactors, in which the dominant
the selectivity
reactor. to synthesis
Furthermore, gas is
the yield of quite highgas
synthesis in the smallest
is favored inreactor,
smaller but dropsinsharply
reactors, in the
which the largest
dominant
chemistry is heterogeneous, as discussed above. Therefore, the design shows great promise for the
reactor.
chemistry Furthermore, the yield
is heterogeneous, as of synthesisabove.
discussed gas is Therefore,
favored in smaller reactors,
the design showsingreat
which the dominant
promise for the
production of synthesis gas, but only at low pressures. To achieve a high yield of synthesis gas at
chemistry
productionisof heterogeneous,
synthesis gas, as butdiscussed above.
only at low Therefore,
pressures. the design
To achieve shows
a high great
yield promise for
of synthesis gasthe
at
high pressures, the reactor dimension must be carefully designed to reduce the mass-transfer
production of synthesis
high pressures, gas, but
the reactor only at low
dimension pressures.
must To achieve
be carefully a high to
designed yield of synthesis
reduce gas at high
the mass-transfer
limitations.
pressures,
limitations.the reactor dimension must be carefully designed to reduce the mass-transfer limitations.

100 60 100 100


100 60 100 100
(%)

(%)
(%)

90 50
(%)

(%)

80 80
selectivity

90 50
selectivity

(%)
(%)

80 80
selectivity

selectivity

(%)
selectivity

80 40
selectivity
selectivity

80 Plate separation distance 40 60 60


selectivity

Plate separation distance 60 0.3 mm 60


monoxide

70 0.3 mm 30
dioxide

0.3 mm
monoxide

70 0.3 mm
0.8 30 0.8
dioxide

Hydrogen

40 0.8 mm 40
0.8 mm 1.5
Water

1.5
Hydrogen

60 20 40 40
1.5 mm
Water

1.5 mm
Carbon

60 20 Oxygen feed
Carbon

Oxygen feed
Carbon

20 Oxygen feed 20
Carbon

50 Oxygen feed 10
20 20
50 10
40 0 0 0
400.1 0.5 1.0 1.5 2.0 2.5 3.00 00.1 0.5 1.0 1.5 2.0 2.5 3.00
0.1 0.5 1.0 Pressure
1.5 (MPa)2.0 2.5 3.0 0.1 0.5 1.0 Pressure
1.5 (MPa)2.0 2.5 3.0
Pressure (MPa) Pressure (MPa)
(a) (b)
(a) (b)

Figure 8. Cont.
Processes2018,
Processes 2018,6,6,x83
FOR PEER REVIEW 1414ofof2424

20 85 1800
Total selectivity to C2 products (%)

Maximum wall temperature (K)


Outlet methane conversion (%)
80
15 1600
Oxygen feed
0.3 mm 75
0.8 mm 0.3 mm 0.3 mm
10 1.5 mm 1400
0.8 mm 0.8 mm
70 1.5 mm 1.5 mm
Oxygen feed
5 1200
65

0 60 1000
0.1 0.5 1.0 1.5 2.0 2.5 3.0 0.1 0.5 1.0 1.5 2.0 2.5 3.0
Pressure (MPa) Pressure (MPa)
(c) (d)
Figure 8. Effect of reactor dimension on the selectivity, conversion, and maximum wall temperature
inFigure 8. Effect of reactor
the methane–oxygen dimension
system. on the selectivity,
(a) Selectivity to carbonconversion, andcarbon
monoxide and maximum wall (b)
dioxide; temperature
selectivityin
the methane–oxygen system. (a) Selectivity to carbon monoxide and carbon dioxide;
to hydrogen and water; (c) total selectivity to C2 products; (d) outlet conversion and maximum (b) selectivity
wall
to hydrogen
temperature. and water; (c) total selectivity to C 2 products; (d) outlet conversion and maximum
wall temperature.
Figure 8a,b also shows the selectivity to the total oxidation products. For the largest dimension
examined Figure 8a,bthe
here, alsoamount
shows the of selectivity to the totalproducts
the total oxidation oxidationisproducts. For the
considerable, largest dimension
especially at high
examinedas
pressures, here, the amount
shown in Figureof8a,b.
the total oxidation
In this context,products
there is an is increases
considerable, especially atdue
in temperature high
topressures,
the heat
as shown
released byinthe
Figure
total8a,b. In this reaction
oxidation context, there
Figure is an
8d,increases in temperature
thus increasing due toofthe
the amount heat released
undesired by-
products
by the totalsuch as C2-hydrocarbons
oxidation reaction Figure (Figure 8c). increasing
8d, thus Therefore,the theamount
contribution of homogeneous
of undesired by-productsreactions
such as
atChigh pressures is(Figure
2 -hydrocarbons considerable for the largest
8c). Therefore, dimension of
the contribution examined.
homogeneousAfter the initiation
reactions of gas-phase
at high pressures
combustion,
is considerable therefor
is the
a sharp drop
largest in conversion
dimension for the moderate
examined. to large reactors
After the initiation studied,
of gas-phase as shown
combustion,
inthere
Figureis a8d. Thedrop
sharp inflection point offor
in conversion thethe
pressure
moderate decreases
to largewith increasing
reactors studied,theas reactor
shown indimension.
Figure 8d.
The
Themaximum wall of
inflection point temperature
the pressureincreases
decreaseswith withincreasing
increasingpressure.
the reactorThis is mainly
dimension. Thedue to the
maximum
increased amount increases
wall temperature of the reactants. A sharp pressure.
with increasing rise in temperature
This is mainly represents initiation amount
due to the increased of gas-phase
of the
combustion,
reactants. A as shown
sharp in temperature
rise in Figure 8d. Finally,
represents thethe
maximum
initiation wall temperature
of gas-phase levels off
combustion, at highin
as shown
pressures.
Figure 8d.Overall,
Finally, the maximum
yield of synthesis gas is favored
wall temperature levelsatoff
low at pressures for theOverall,
high pressures. methane–oxygen
the yield of
system.
synthesis gas is favored at low pressures for the methane–oxygen system.

3.5.Effect
3.5. EffectofofReactor
ReactorDimension
Dimensionfor
forAir
AirFeed
Feed
Asdiscussed
As discussed above,
above, thethedimension
dimensioncan significantly
can significantlyaffect the performance
affect the performanceof theofmethane–oxygen
the methane–
system operated at high pressures. To gain a better understanding of the design
oxygen system operated at high pressures. To gain a better understanding of the design described described in this paper,
in
the paper,
this effect ofthe
reactor
effectdimension
of reactorisdimension
investigatedisfor the methane–air
investigated for thesystem operatedsystem
methane–air at various pressures.
operated at
Figure
various pressures.9 presents the results obtained for the selectivity, conversion, and maximum wall
temperature
Figure 9 at different
presents thereactor
resultsdimensions.
obtained forThe thereactor dimension
selectivity, has little
conversion, andor maximum
no effect on the
wall
selectivity toat synthesis
temperature different gas (Figure
reactor 9a,b) under
dimensions. The the conditions
reactor dimension studied here. orThe
has little no production
effect on theof
C -hydrocarbons
selectivity
2 is favored in lager reactors (Figure 9c). On the other hand,
to synthesis gas (Figure 9a,b) under the conditions studied here. The production the reactor dimension
of C2-
has little effect on the conversion (left vertical axis, Figure 9d), and the maximum
hydrocarbons is favored in lager reactors (Figure 9c). On the other hand, the reactor dimension has wall temperature
(right
little vertical
effect on theaxis, Figure 9d).
conversion (leftAs expected,
vertical axis, the dominant
Figure 9d), andchemistry
the maximum is heterogeneous
wall temperature in all(right
of the
vertical axis, Figure 9d). As expected, the dominant chemistry is heterogeneous in all of the casesin
cases under the conditions studied here, and the initiation of gas-phase combustion is impossible
the methane–air
under the conditionssystem examined.
studied here, and the initiation of gas-phase combustion is impossible in the
methane–air system examined.
Processes 2018,
Processes 6, x83FOR PEER REVIEW
2018, 6, 15 of
15 of 24
24

86 15 88 28
0.3 mm
Carbon monoxide selectivity (%)

Carbon dioxide selectivity (%)


0.8 mm

Hydrogen selectivity (%)


1.5 mm

Water selectivity (%)


84 24
84 13
0.3 mm
80 0.8 mm 20
1.5 mm
82 11
76 16
Air feed
Air feed

80 9 72 12
0.1 0.5 1.0 1.5 2.0 2.5 3.0 0.1 0.5 1.0 1.5 2.0 2.5 3.0
Pressure (MPa) Pressure (MPa)
(a) (b)

0.9 70 1200
Total selectivity to C2 products (%)

Maximum wall temperature (K)


Outlet methane conversion (%)
67 1150
0.6 0.3 mm
0.8 mm 0.3 mm
1.5 mm 64 0.8 mm 1100
1.5 mm
0.3
Air feed 61 1050
Air feed

0.0 58 1000
0.1 0.5 1.0 1.5 2.0 2.5 3.0 0.1 0.5 1.0 1.5 2.0 2.5 3.0
Pressure (MPa) Pressure (MPa)
(c) (d)
Figure 9. Effect of dimension on the performance of the methane–air system operated at various
Figure 9. Effect of dimension on the performance of the methane–air system operated at various
pressures. (a) Selectivity to carbon monoxide and carbon dioxide; (b) selectivity to hydrogen and
pressures. (a) Selectivity to carbon monoxide and carbon dioxide; (b) selectivity to hydrogen and water;
water; (c) total selectivity to C2 products; (d) outlet conversion and maximum wall temperature.
(c) total selectivity to C2 products; (d) outlet conversion and maximum wall temperature.

3.6. Effect of Nitrogen Diluent


3.6. Effect of Nitrogen Diluent
For the design examined, the onset of gas-phase combustion can be inhibited by utilizing a
For the design examined, the onset of gas-phase combustion can be inhibited by utilizing a
methane–air system (Figures 7 and 9), while a methane–oxygen system (Figures 6 and 8) will allow
methane–air system (Figures 7 and 9), while a methane–oxygen system (Figures 6 and 8) will allow
the gas mixture to be ignited at a certain pressure, as discussed above. It is therefore important to
the gas mixture to be ignited at a certain pressure, as discussed above. It is therefore important to
determine the critical dilution to reduce the contribution of the reaction occurring in the gas phase.
determine the critical dilution to reduce the contribution of the reaction occurring in the gas phase.
Figure 10 shows the influence of nitrogen diluent on the performance of the reactor operated at
Figure 10 shows the influence of nitrogen diluent on the performance of the reactor operated at
preheating temperature 700 K. The contribution of homogeneous reactions is small, but considerable
preheating temperature 700 K. The contribution of homogeneous reactions is small, but considerable
under certain conditions. At high pressures, the methane–air system has the advantage of reducing
under certain conditions. At high pressures, the methane–air system has the advantage of reducing
undesired by-products and improving the selectivity to synthesis gas, as shown in Figure 10a,b. At
undesired by-products and improving the selectivity to synthesis gas, as shown in Figure 10a,b. At the
the highest pressure examined, there appears to be an optimal dilution ratio, of about 38% nitrogen
highest pressure examined, there appears to be an optimal dilution ratio, of about 38% nitrogen in the
in the mixture, which exhibits the maximum yield of synthesis gas. The methane–air system suffers
mixture, which exhibits the maximum yield of synthesis gas. The methane–air system suffers a slight
a slight drop in conversion (Figure 10d), but offers an economical solution to the production of
drop in conversion (Figure 10d), but offers an economical solution to the production of synthesis gas
synthesis gas (Figure 10a,b). At moderate pressure 1.5 MPa, only 8% nitrogen diluent is needed to
(Figure 10a,b). At moderate pressure 1.5 MPa, only 8% nitrogen diluent is needed to avoid the initiation
avoid the initiation of the combustion reaction occurring in the gas phase. At atmospheric pressure,
of the combustion reaction occurring in the gas phase. At atmospheric pressure, the contribution of
the contribution of homogeneous reactions is rather small, and thus nitrogen dilution has little or no
homogeneous reactions is rather small, and thus nitrogen dilution has little or no effect on the yield
effect on the yield of synthesis gas. At moderate to high pressures, there exists a sharp rise in the yield
of synthesis gas. At moderate to high pressures, there exists a sharp rise in the yield of synthesis
of synthesis gas, as shown in Figure 10. The inflection point of the nitrogen diluent increases with
gas, as shown in Figure 10. The inflection point of the nitrogen diluent increases with increasing
increasing pressure. For the methane–air system, the yield of synthesis gas also increases with
pressure. For the methane–air system, the yield of synthesis gas also increases with increasing pressure.
increasing pressure. For the system operated at moderate to high pressures with a low dilution, the
For the system operated at moderate to high pressures with a low dilution, the initiation of gas-phase
initiation of gas-phase combustion is possible, and both the amount of C 2-hydrocarbons and the
maximum wall temperature increases with increasing pressure, as shown in Figure 10c,d.
Processes 2018, 6, 83 16 of 24

combustion is possible, and both the amount of C2 -hydrocarbons and the maximum wall temperature
Processes 2018, 6, x FOR PEER REVIEW 16 of 24
increases with increasing pressure, as shown in Figure 10c,d.

100 40 100 40
Carbon monoxide selectivity (%)

Carbon dioxide selectivity (%)


0.1 MPa

Hydrogen selectivity (%)


1.5 MPa

Water selectivity (%)


90 30 90 30
3.0 MPa

0.1 MPa
80 1.5 MPa 20 80 20
3.0 MPa

0.1 MPa
70 10 70 1.5 MPa 10
3.0 MPa

60 0 60 0
0 10 20 30 40 50 55.6 0 10 20 30 40 50 55.6
Oxygen feed Nitrogen mole fraction (%) Air feed Oxygen feed Nitrogen mole fraction (%) Air feed
(a) (b)

20 100 1800
0.1 MPa
Total selectivity to C2 products (%)

Maximum wall temperature (K)


Outlet methane conversion (%)

1.5 MPa
90 3.0 MPa
15 1600

80
0.1 MPa
10 1.5 MPa 1400
3.0 MPa 70

5 1200
60

0 50 1000
0 10 20 30 40 50 55.6 0 10 20 30 40 50 55.6
Oxygen feed Nitrogen mole fraction (%) Air feed Oxygen feed Nitrogen mole fraction (%) Air feed
(c) (d)
Figure
Figure 10.
10. Effect
Effectof
of nitrogen
nitrogen diluent
diluentonon the
the selectivity,
selectivity, conversion,
conversion, and
and maximum
maximum wall wall temperature
temperature
at
at different pressures when the reactor is operated at preheating temperature 700 K. (a)
different pressures when the reactor is operated at preheating temperature 700 K. (a) Selectivity
Selectivitytoto
carbon monoxide and carbon dioxide; (b) selectivity to hydrogen and water; (c) total selectivity
carbon monoxide and carbon dioxide; (b) selectivity to hydrogen and water; (c) total selectivity to C22 to C
products;
products; (d)
(d) outlet
outlet conversion
conversion and
and maximum
maximum wall wall temperature.
temperature.

3.7. Air Feed Versus Oxygen Feed


3.7. Air Feed Versus Oxygen Feed
Comparisons are made between the results obtained for the two systems operated at
Comparisons are made between the results obtained for the two systems operated at atmospheric
atmospheric pressure. Figure 11 shows the influence of preheating temperature on the performance
pressure. Figure 11 shows the influence of preheating temperature on the performance of the two
of the two systems operated at atmospheric pressure. Both the selectivity to synthesis gas (Figure
systems operated at atmospheric pressure. Both the selectivity to synthesis gas (Figure 11a,b) and the
11a,b) and the outlet conversion (Figure 11c) are higher for the methane–oxygen system than for the
outlet conversion (Figure 11c) are higher for the methane–oxygen system than for the methane–air
methane–air system, especially at lower temperatures. For both systems, the production of synthesis
system, especially at lower temperatures. For both systems, the production of synthesis gas is favored at
gas is favored at high preheating temperatures, but the methane–air system can benefit more from
high preheating temperatures, but the methane–air system can benefit more from preheating, as shown
preheating, as shown in Figure 11. Therefore, the effect of preheating is more pronounced for the
in Figure 11. Therefore, the effect of preheating is more pronounced for the methane–air system.
methane–air system. For each of the two systems, the maximum wall temperature increases with
For each of the two systems, the maximum wall temperature increases with increasing preheating
increasing preheating temperature under the conditions examined here (Figure 11c). As expected, the
temperature under the conditions examined here (Figure 11c). As expected, the maximum temperature
maximum temperature within the walls is higher for the methane–oxygen system than for the
within the walls is higher for the methane–oxygen system than for the methane–air system.
methane–air system.
Processes 2018, 6,
Processes 2018, 6, 83
x FOR PEER REVIEW 17
17 of
of 24
24

100 20 100 20
Carbon monoxide selectivity (%)

feed

Carbon dioxide selectivity (%)


Oxygen en fee
d
Oxyg

Hydrogen selectivity (%)

Water selectivity (%)


95 15 95 15

d d
fee fee
90 Air 10 90 Air 10

Air Air
fee fee
d d
85 5 85 5
Oxyg
Oxygen en fee
feed d

80 0 80 0
300 500 700 900 300 500 700 900
Preheating temperature (K) Preheating temperature (K)
(a) (b)

100 1400

Maximum wall temperature (K)


Outlet methane conversion (%)

d
en fee
90 Oxyg 1300

feed
80 Air 1200

ed
en fe
Oxyg
70 1100
eed
Air f

60 1000
300 500 700 900
Preheating temperature (K)
(c)
Figure 11. Comparisons
Figure 11. Comparisons between
between the
the methane–oxygen
methane–oxygen system
system and the methane–air
and the methane–air system
system operated
operated
at atmospheric pressure. (a) Selectivity to carbon monoxide and carbon dioxide; (b) selectivity
at atmospheric pressure. (a) Selectivity to carbon monoxide and carbon dioxide; (b) selectivity to to
hydrogen and water;
hydrogen and water; (c)
(c) outlet
outlet conversion
conversion and
and maximum
maximum wall
wall temperature.
temperature.

4. Further Discussion
4. Further Discussion
The results presented above indicate that the design shows great promise for the production of
The results presented above indicate that the design shows great promise for the production of
synthesis gas from methane. When the dominant chemistry is heterogeneous, the effect of pressure
synthesis gas from methane. When the dominant chemistry is heterogeneous, the effect of pressure is
is small. In contrast, when the gas-phase chemistry is considerable, the pressure can significantly
small. In contrast, when the gas-phase chemistry is considerable, the pressure can significantly affect
affect the reactor performance such as the yield of synthesis gas. In this context, the initiation of gas-
the reactor performance such as the yield of synthesis gas. In this context, the initiation of gas-phase
phase combustion is possible, the temperature is high, the amount of undesired by-products is
combustion is possible, the temperature is high, the amount of undesired by-products is relatively
relatively large, and the out conversion is low, as shown in Figure 8. For the base case, a stoichiometric
large, and the out conversion is low, as shown in Figure 8. For the base case, a stoichiometric mixture
mixture for the production of synthesis gas is utilized. In this context, homogeneous reactions can
for the production of synthesis gas is utilized. In this context, homogeneous reactions can lead to a
lead to a lower outlet conversion, because they tend to consume more amount of oxygen than
lower outlet conversion, because they tend to consume more amount of oxygen than heterogeneous
heterogeneous reactions, which leads to lack of oxygen.
reactions, which leads to lack of oxygen.
At the highest pressure examined, it may be difficult to optimize operating conditions of the
At the highest pressure examined, it may be difficult to optimize operating conditions of
methane–oxygen system. This is because high preheating temperatures increase the outlet
the methane–oxygen system. This is because high preheating temperatures increase the outlet
conversion, but decrease the selectivity to synthesis gas, as shown in Figure 6. Furthermore, the
conversion, but decrease the selectivity to synthesis gas, as shown in Figure 6. Furthermore, the
selectivity to synthesis gas is not high enough, at all of the preheating temperatures examined for the
selectivity to synthesis gas is not high enough, at all of the preheating temperatures examined for the
methane–oxygen system. The selectivity to synthesis gas decreases with increasing the preheating
methane–oxygen system. The selectivity to synthesis gas decreases with increasing the preheating
temperature, as shown in Figure 6. However, smaller reactors show great promise, since they can
temperature, as shown in Figure 6. However, smaller reactors show great promise, since they can
delay the initiation of gas-phase combustion, as shown in Figure 8.
delay the initiation of gas-phase combustion, as shown in Figure 8.
At high pressures, the methane–air system has the advantage of reducing the contribution of
At high pressures, the methane–air system has the advantage of reducing the contribution of
homogeneous reactions, as shown in Figures 6 and 7. At low pressures, the dimension has little or no
homogeneous reactions, as shown in Figures 6 and 7. At low pressures, the dimension has little or
effect on the performance of each of the two systems, as shown in Figures 8 and 9. At all of the
no effect on the performance of each of the two systems, as shown in Figures 8 and 9. At all of the
pressures examined here, the dimension has little effect on the performance of the methane–air
system, and the contribution of homogeneous reactions is negligible, as illustrated in Figure 9. For
Processes 2018, 6, 83 18 of 24

pressures examined here, the dimension has little effect on the performance of the methane–air system,
and the contribution of homogeneous reactions is negligible, as illustrated in Figure 9. For the mixture
of methane and oxygen diluted with a large amount of nitrogen, the collision between the reactant
molecules becomes less frequent, which can greatly reduce the contribution of homogeneous reactions,
as shown in Figure 10. However, the contribution may still be considerable at high pressures, as shown
in Figure 10, because mass-transfer limitations are significant under the conditions examined here.
At the highest pressure examined here, the methane–air system can offer a good yield of
synthesis gas, especially at high preheating temperatures, as shown in Figure 7. For each of the
two systems, the yield of synthesis gas decreases with increasing pressure, as shown in Figures 6
and 7. At high pressures, the production of synthesis gas is favored at low preheating temperatures
for the methane–oxygen system (Figure 6), but at high preheating temperatures for the methane–air
system (Figure 7). At atmospheric pressure, the use of oxygen and the use of preheated air have
the similar effect on the reactor performance, due to the negligible contribution of homogeneous
reactions (Figure 11). At high pressures, however, the effect played by the above two feeding methods
is quite different from each other, because in this context the contribution of homogeneous reactions is
considerable. For each of the two systems, the outlet conversion decreases with increasing pressure
due to the mass-transfer limitations and the possible occurrence of gas-phase combustion, especially
in larger dimensions. The mass-transfer effect is more pronounced at high pressures, as shown in
Figures 7 and 8. As the dimension of the methane–oxygen system increases, the outlet conversion
drops sharply (Figure 8), because in this context the mass-transfer effect may be important [79,80] and
the initiation of gas-phase combustion is possible.

5. Conclusions
Catalytic partial oxidation of methane in microchannel reactors in high temperature environments
was studied numerically. This investigation provided knowledge on how reaction conditions affect the
operating characteristics and the distribution of reaction products in the reactor. Comparisons were
made with published results, suggesting that the model developed in this paper is reliable and helpful
to the design of the system. Furthermore, the catalytic partial oxidation process under extremely short
contact time conditions can be accurately described by the model via the combination of detailed
heterogeneous and homogenous reaction mechanisms.
The results have shown that the relative role of heterogeneous and homogeneous reaction
pathways depends strongly upon the operating conditions. The distribution of reaction products
depends upon a number of factors, such as the reactor dimension, pressure, temperature, and feed
composition. The operating temperature can significantly affect the yield of synthesis gas. Undesired
homogeneous reactions are favored in large reactors, and at high temperatures and pressures. It is
more economical to utilize air instead of oxygen as the oxidant. Such an arrangement is particularly
beneficial, since the onset of homogeneous reactions can thereby be inhibited or avoided. It is necessary
to use the reactant mixture diluted with a sufficient amount of nitrogen to avoid homogenous reactions
at high pressures. When air is used as the oxidant, preheating is needed, and the production of
synthesis gas is practically favored at high pressures. The use of air as the oxidant is preferred at
industrially relevant pressure 3 MPa, at which the contribution of undesired homogeneous reactions is
usually small.
Further research is needed on the principles underlying the catalytic partial oxidation process.
Catalyst deactivation may be an important risk factor, which is not addressed in this paper.
This deactivation can significantly decrease the yield of synthesis gas, thus reducing the performance
of the system. There are potential solutions to this issue, such as process control to maintain the
temperature below a certain damaging threshold, non-uniform catalyst distribution, and thicker
catalyst layers. Furthermore, the success of microchannel reactors is highly dependent on the robust
catalysts suitable for the operating conditions of these small-scale chemical systems. On the other
hand, high temperatures obtained within the system may destroy the catalyst wash coat employed
Processes 2018, 6, 83 19 of 24

and impose severe constraints on the materials used. Lower reactor temperatures are essential for the
stability of the catalyst and materials used. The problem related to the materials stability limit is also
not addressed in this paper. A temperature threshold should be well defined in the practical design,
which will be the subject of future work.

Author Contributions: Conceptualization, J.C.; Methodology, J.C.; Software, W.S.; Validation, W.S.; Formal
Analysis, J.C. and D.X.; Investigation, J.C. and W.S.; Resources, D.X.; Supervision, J.C.; Project Administration,
J.C. and D.X.; J.C. conceived and designed the analyses and simulations; W.S. implemented the C code for the
ANSYS Fluent program; D.X. contributed to concepts and notebooks; J.C. and D.X. provided chemical expertise
and insight; J.C. wrote the paper.
Funding: This research was funded by the National Natural Science Foundation of China (No. 51506048) and the
Fundamental Research Funds for the Universities of Henan Province (No. NSFRF140119).
Acknowledgments: J.C. is grateful for financial support from the National Natural Science Foundation of China
and the Henan Polytechnic University. The authors would also like to acknowledge the DETCHEM website for
their computing resources.
Conflicts of Interest: The authors declare no conflict of interest.

Nomenclature
A pre-exponential factor
A0 surface area
C concentration
c specific heat capacity
D diffusivity
DT thermal diffusivity
Deff effective diffusivity
Dm mixture-averaged diffusivity
d channel height
dpore mean pore diameter
Ea activation energy
Fcat/geo catalyst/geometric surface area
F view factor
∆r Hm Θ standard molar enthalpy of reaction
h specific enthalpy
ho external heat loss coefficient
Kg , Ks number of gaseous species and number of surface species
l length
m total number of gaseous and surface species
p pressure
q heat flux
R ideal gas constant
.
s rate of appearance of a heterogeneous product
s sticking coefficient
T, To absolute temperature and reference temperature
u, v streamwise and transverse velocity components

V, V diffusion velocity and diffusion velocity vector

W, W relative molecular mass and relative molecular mass of the mixture
x streamwise coordinate
Y mass fraction
y transverse coordinate
Processes 2018, 6, 83 20 of 24

Greek variables
γ surface area per unit catalyst volume
ε emissivity
δ thickness
εp catalyst porosity
λ thermal conductivity
η effectiveness factor
µ dynamic viscosity
ρ density
σ Stefan-Boltzmann constant
ϑ site occupancy
τp catalyst tortuosity factor
φ inlet molar ratio
.
ω rate of appearance of a homogeneous product
Γ site density
Θ surface coverage
Φ Thiele modulus
Subscripts
amb ambient
eff effective
g gas
i, k, m species index, gaseous species index, and surface species index
in inlet
o outer
rad radiation
s solid
w wall
x, y streamwise and transverse components

References
1. Guo, S.; Wang, J.; Ding, C.; Duan, Q.; Ma, Q.; Zhang, K.; Liu, P. Confining Ni nanoparticles in honeycomb-like
silica for coking and sintering resistant partial oxidation of methane. Int. J. Hydrogen Energy 2018, 43,
6603–6613. [CrossRef]
2. Grundner, S.; Luo, W.; Sanchez-Sanchez, M.; Lercher, J.A. Synthesis of single-site copper catalysts for methane
partial oxidation. Chem. Commun. 2016, 52, 2553–2556. [CrossRef] [PubMed]
3. Kee, R.J.; Zhu, H.; Sukeshini, A.M.; Jackson, G.S. Solid oxide fuel cells: Operating principles, current
challenges, and the role of syngas. Combust. Sci. Technol. 2008, 180, 1207–1244. [CrossRef]
4. Baldinelli, A.; Barelli, L.; Bidini, G. Performance characterization and modelling of syngas-fed SOFCs (solid
oxide fuel cells) varying fuel composition. Energy 2015, 90, 2070–2084. [CrossRef]
5. Pramanik, S.; Ravikrishna, R.V. Numerical study of rich catalytic combustion of syngas. Int. J.
Hydrogen Energy 2017, 42, 16514–16528. [CrossRef]
6. Zheng, X.; Mantzaras, J.; Bombach, R. Homogeneous combustion of fuel-lean syngas mixtures over platinum
at elevated pressures and preheats. Combust. Flame 2013, 160, 155–169. [CrossRef]
7. Sengodan, S.; Lan, R.; Humphreys, J.; Du, D.; Xu, W.; Wang, H.; Tao, S. Advances in reforming and partial
oxidation of hydrocarbons for hydrogen production and fuel cell applications. Renew. Sustain. Energy Rev.
2018, 82, 761–780. [CrossRef]
8. Tran, A.; Pont, M.; Crose, M.; Christofides, P.D. Real-time furnace balancing of steam methane reforming
furnaces. Chem. Eng. Res. Des. 2018, 134, 238–256. [CrossRef]
9. Lu, N.; Gallucci, F.; Melchiori, T.; Xie, D.; Van Sint Annaland, M. Modeling of autothermal reforming of
methane in a fluidized bed reactor with perovskite membranes. Chem. Eng. Process 2018, 124, 308–318.
[CrossRef]
10. Goralski, C.T.; O’Connor, R.P.; Schmidt, L.D. Modeling homogeneous and heterogeneous chemistry in the
production of syngas from methane. Chem. Eng. Sci. 2000, 55, 1357–1370. [CrossRef]
Processes 2018, 6, 83 21 of 24

11. Shelepova, E.; Vedyagin, A.; Sadykov, V.; Mezentseva, N.; Fedorova, Y.; Smorygo, O.; Klenov, O.; Mishakov, I.
Theoretical and experimental study of methane partial oxidation to syngas in catalytic membrane reactor
with asymmetric oxygen-permeable membrane. Catal. Today 2016, 268, 103–110. [CrossRef]
12. Zhang, Q.; Liu, Y.; Chen, T.; Yu, X.; Wang, J.; Wang, T. Simulations of methane partial oxidation by CFD
coupled with detailed chemistry at industrial operating conditions. Chem. Eng. Sci. 2016, 142, 126–136.
[CrossRef]
13. Chakrabarti, R.; Kruger, J.S.; Hermann, R.J.; Blass, S.D.; Schmidt, L.D. Spatial profiles in partial oxidation
of methane and dimethyl ether in an autothermal reactor over rhodium catalysts. Appl. Catal. A 2014, 483,
97–102. [CrossRef]
14. Chen, W.-H.; Lin, S.-C. Characterization of catalytic partial oxidation of methane with carbon dioxide
utilization and excess enthalpy recovery. Appl. Energy 2016, 162, 1141–1152. [CrossRef]
15. York, A.P.E.; Xiao, T.C.; Green, M.L.H. Brief overview of the partial oxidation of methane to synthesis gas.
Top. Catal. 2003, 22, 345–358. [CrossRef]
16. Christian Enger, B.; Lødeng, R.; Holmen, A. A review of catalytic partial oxidation of methane to synthesis
gas with emphasis on reaction mechanisms over transition metal catalysts. Appl. Catal. A 2008, 346, 1–27.
[CrossRef]
17. Guo, D.; Wang, G.-C. Partial oxidation of methane on anatase and rutile defective TiO2 supported Rh4
cluster: A density functional theory study. J. Phys. Chem. C 2017, 121, 26308–26320. [CrossRef]
18. Kraus, P.; Lindstedt, R.P. Microkinetic mechanisms for partial oxidation of methane over platinum and
rhodium. J. Phys. Chem. C 2017, 121, 9442–9453. [CrossRef]
19. Urasaki, K.; Kado, S.; Kiryu, A.; Imagawa, K.-i.; Tomishige, K.; Horn, R.; Korup, O.; Suehiro, Y. Synthesis gas
production by catalytic partial oxidation of natural gas using ceramic foam catalyst. Catal. Today 2018, 299,
219–228. [CrossRef]
20. Gil-Calvo, M.; Jiménez-González, C.; de Rivas, B.; Gutiérrez-Ortiz, J.I.; López-Fonseca, R. Effect of Ni/Al
molar ratio on the performance of substoichiometric NiAl2 O4 spinel-based catalysts for partial oxidation of
methane. Appl. Catal. B 2017, 209, 128–138. [CrossRef]
21. Kumar Singha, R.; Shukla, A.; Yadav, A.; Sain, S.; Pendem, C.; Kumar Konathala, L.N.S.; Bal, R. Synthesis
effects on activity and stability of Pt-CeO2 catalysts for partial oxidation of methane. Mol. Catal. 2017, 432,
131–143. [CrossRef]
22. Osman, A.I.; Meudal, J.; Laffir, F.; Thompson, J.; Rooney, D. Enhanced catalytic activity of Ni on η-Al2 O3
and ZSM-5 on addition of ceria zirconia for the partial oxidation of methane. Appl. Catal. B 2017, 212, 68–79.
[CrossRef]
23. Singha, R.K.; Ghosh, S.; Acharyya, S.S.; Yadav, A.; Shukla, A.; Sasaki, T.; Venezia, A.M.; Pendem, C.; Bal, R.
Partial oxidation of methane to synthesis gas over Pt nanoparticles supported on nanocrystalline CeO2
catalyst. Catal. Sci. Technol. 2016, 6, 4601–4615. [CrossRef]
24. Luo, Z.; Kriz, D.A.; Miao, R.; Kuo, C.-H.; Zhong, W.; Guild, C.; He, J.; Willis, B.; Dang, Y.; Suib, S.L.; et al.
TiO2 Supported gold-palladium catalyst for effective syngas production from methane partial oxidation.
Appl. Catal. A 2018, 554, 54–63. [CrossRef]
25. Boukha, Z.; Gil-Calvo, M.; de Rivas, B.; González-Velasco, J.R.; Gutiérrez-Ortiz, J.I.; López-Fonseca, R.
Behaviour of Rh supported on hydroxyapatite catalysts in partial oxidation and steam reforming of methane:
On the role of the speciation of the Rh particles. Appl. Catal. A 2018, 556, 191–203. [CrossRef]
26. Scarabello, A.; Dalle Nogare, D.; Canu, P.; Lanza, R. Partial oxidation of methane on Rh/ZrO2 and
Rh/Ce-ZrO2 on monoliths: Catalyst restructuring at reaction conditions. Appl. Catal. B 2015, 174–175,
308–322. [CrossRef]
27. Figen, H.E.; Baykara, S.Z. Effect of ruthenium addition on molybdenum catalysts for syngas production via
catalytic partial oxidation of methane in a monolithic reactor. Int. J. Hydrogen Energy 2018, 43, 1129–1138.
[CrossRef]
28. Zhu, Y.; Barat, R. Partial oxidation of methane over a ruthenium phthalocyanine catalyst. Chem. Eng. Sci.
2014, 116, 71–76. [CrossRef]
29. Wang, F.; Li, W.-Z.; Lin, J.-D.; Chen, Z.-Q.; Wang, Y. Crucial support effect on the durability of Pt/MgAl2 O4
for partial oxidation of methane to syngas. Appl. Catal. B 2018, 231, 292–298. [CrossRef]
Processes 2018, 6, 83 22 of 24

30. Singha, R.K.; Shukla, A.; Yadav, A.; Sasaki, T.; Sandupatla, A.; Deo, G.; Bal, R. Pt-CeO2 nanoporous spheres -
an excellent catalyst for partial oxidation of methane: Effect of the bimodal pore structure. Catal. Sci. Technol.
2017, 7, 4720–4735. [CrossRef]
31. Li, B.; Li, H.; Weng, W.-Z.; Zhang, Q.; Huang, C.-J.; Wan, H.-L. Synthesis gas production from partial
oxidation of methane over highly dispersed Pd/SiO2 catalyst. Fuel 2013, 103, 1032–1038. [CrossRef]
32. Yashnik, S.A.; Chesalov, Y.A.; Ishchenko, A.V.; Kaichev, V.V.; Ismagilov, Z.R. Effect of Pt addition on sulfur
dioxide and water vapor tolerance of Pd-Mn-hexaaluminate catalysts for high-temperature oxidation of
methane. Appl. Catal. B 2017, 204, 89–106. [CrossRef]
33. Singha, R.K.; Shukla, A.; Yadav, A.; Sivakumar Konathala, L.N.; Bal, R. Effect of metal-support interaction on
activity and stability of Ni-CeO2 catalyst for partial oxidation of methane. Appl. Catal. B 2017, 202, 473–488.
[CrossRef]
34. Kaddeche, D.; Djaidja, A.; Barama, A. Partial oxidation of methane on co-precipitated Ni-Mg/Al catalysts
modified with copper or iron. Int. J. Hydrogen Energy 2017, 42, 15002–15009. [CrossRef]
35. Nakagawa, K.; Ikenaga, N.; Suzuki, T.; Kobayashi, T.; Haruta, M. Partial oxidation of methane to synthesis
gas over supported iridium catalysts. Appl. Catal. A 1998, 169, 281–290. [CrossRef]
36. Nakagawa, K.; Ikenaga, N.; Teng, Y.; Kobayashi, T.; Suzuki, T. Partial oxidation of methane to synthesis gas
over iridium-nickel bimetallic catalysts. Appl. Catal. A 1999, 180, 183–193. [CrossRef]
37. De SantanaSantos, M.; Neto, R.C.R.; Noronha, F.B.; Bargiela, P.; da Graça Carneiro da Rocha, M.; Resini, C.;
Carbó-Argibay, E.; Frétya, R.; Brandão, S.T. Perovskite as catalyst precursors in the partial oxidation of
methane: The effect of cobalt, nickel and pretreatment. Catal. Today 2018, 299, 229–241. [CrossRef]
38. López-Ortiz, A.; González-Vargas, P.E.; Meléndez-Zaragoza, M.J.; Collins-Martínez, V. Thermodynamic
analysis and process simulation of syngas production from methane using CoWO4 as oxygen carrier. Int. J.
Hydrogen Energy 2017, 42, 30223–30236. [CrossRef]
39. Horn, R.; Williams, K.A.; Degenstein, N.J.; Schmidt, L.D. Syngas by catalytic partial oxidation of methane
on rhodium: Mechanistic conclusions from spatially resolved measurements and numerical simulations.
J. Catal. 2006, 242, 92–102. [CrossRef]
40. Nogare, D.D.; Degenstein, N.J.; Horn, R.; Canu, P.; Schmidt, L.D. Modeling spatially resolved data of methane
catalytic partial oxidation on Rh foam catalyst at different inlet compositions and flowrates. J. Catal. 2011,
277, 134–148. [CrossRef]
41. Horn, R.; Williams, K.A.; Degenstein, N.J.; Bitsch-Larsen, A.; Dalle Nogare, D.; Tupy, S.A.; Schmidt, L.D.
Methane catalytic partial oxidation on autothermal Rh and Pt foam catalysts: Oxidation and reforming
zones, transport effects, and approach to thermodynamic equilibrium. J. Catal. 2007, 249, 380–393. [CrossRef]
42. Zhan, Z.; Lin, Y.; Pillai, M.; Kim, I.; Barnett, S.A. High-rate electrochemical partial oxidation of methane in
solid oxide fuel cells. J. Power Sources 2006, 161, 460–465. [CrossRef]
43. Lee, D.; Myung, J.; Tan, J.; Hyun, S.-H.; Irvine, J.T.S.; Kim, J.; Moon, J. Direct methane solid oxide fuel cells
based on catalytic partial oxidation enabling complete coking tolerance of Ni-based anodes. J. Power Sources
2017, 345, 30–40. [CrossRef]
44. Wang, B.; Albarracín-Suazo, S.; Pagán-Torres, Y.; Nikolla, E. Advances in methane conversion processes.
Catal. Today 2017, 285, 147–158. [CrossRef]
45. Taifan, W.; Baltrusaitis, J. CH4 conversion to value added products: Potential, limitations and extensions of a
single step heterogeneous catalysis. Appl. Catal. B 2016, 198, 525–547. [CrossRef]
46. Védrine, J.C.; Fechete, I. Heterogeneous partial oxidation catalysis on metal oxides. C. R. Chim. 2016, 19,
1203–1225. [CrossRef]
47. Arutyunov, V.S.; Strekova, L.N. The interplay of catalytic and gas-phase stages at oxidative conversion of
methane: A review. J. Mol. Catal. A Chem. 2017, 426, 326–342. [CrossRef]
48. Tanimu, A.; Jaenicke, S.; Alhooshani, K. Heterogeneous catalysis in continuous flow microreactors: A review
of methods and applications. Chem. Eng. J. 2017, 327, 792–821. [CrossRef]
49. Yao, X.; Zhang, Y.; Du, L.; Liu, J.; Yao, J. Review of the applications of microreactors. Renew. Sustain.
Energy Rev. 2015, 47, 519–539. [CrossRef]
50. Geyer, K.; Codée, J.D.C.; Seeberger, P.H. Microreactors as tools for synthetic chemists—The chemists’
round-bottomed flask of the 21st century? Chem. Eur. J. 2006, 12, 8434–8442. [CrossRef] [PubMed]
51. Kashid, M.N.; Kiwi-Minsker, L. Microstructured reactors for multiphase reactions: State of the art. Ind. Eng.
Chem. Res. 2009, 48, 6465–6485. [CrossRef]
Processes 2018, 6, 83 23 of 24

52. Kiwi-Minsker, L.; Renken, A. Microstructured reactors for catalytic reactions. Catal. Today 2005, 110, 2–14.
[CrossRef]
53. Kolb, G.; Hessel, V. Micro-structured reactors for gas phase reactions. Chem. Eng. Sci. 2004, 98, 1–38.
[CrossRef]
54. Jensen, K.F. Microreaction engineering—Is small better? Chem. Eng. Sci. 2001, 56, 293–303. [CrossRef]
55. Jähnisch, K.; Hessel, V.; Löwe, H.; Baerns, M. Chemistry in microstructured reactors. Angew. Chem. Int. Ed.
2004, 43, 406–446. [CrossRef] [PubMed]
56. Deutschmann, O. Modeling of the interactions between catalytic surfaces and gas-phase. Catal. Lett. 2015,
145, 272–289. [CrossRef]
57. Bawornruttanaboonya, K.; Devahastin, S.; Mujumdar, A.S.; Laosiripojana, N. A computational fluid dynamic
evaluation of a new microreactor design for catalytic partial oxidation of methane. Int. J. Heat Mass Transf.
2017, 115, 174–185. [CrossRef]
58. Navalho, J.E.P.; Pereira, J.M.C.; Pereira, J.C.F. Multiscale modeling of methane catalytic partial oxidation:
From the mesopore to the full-scale reactor operation. AIChE J. 2018, 64, 578–594. [CrossRef]
59. Tonkovich, A.; Kuhlmann, D.; Rogers, A.; McDaniel, J.; Fitzgerald, S.; Arora, R.; Yuschak, T. Microchannel
technology scale-up to commercial capacity. Chem. Eng. Res. Des. 2005, 83, 634–639. [CrossRef]
60. Tonkovich, A.Y.; Perry, S.; Wang, Y.; Qiu, D.; LaPlante, T.; Rogers, W.A. Microchannel process technology for
compact methane steam reforming. Chem. Eng. Sci. 2004, 59, 4819–4824. [CrossRef]
61. Suryawanshi, P.L.; Gumfekar, S.P.; Bhanvase, B.A.; Sonawane, S.H.; Pimplapure, M.S. A review on
microreactors: Reactor fabrication, design, and cutting-edge applications. Chem. Eng. Sci. 2018. [CrossRef]
62. ANSYS Fluent User’s Guide; Release 16.0; ANSYS Inc.: Canonsburg, PA, USA, 2014.
63. Kee, R.J.; Dixon-lewis, G.; Warnatz, J.; Coltrin, M.E.; Miller, J.A.; Moffat, H.K. A Fortran Computer Code
Package for the Evaluation of Gas-Phase, Multicomponent Transport Properties; Report No. SAND86-8246B; Sandia
National Laboratories: Livermore, CA, USA, 1998.
64. Von Rickenbach, J.; Lucci, F.; Narayanan, C.; Dimopoulos Eggenschwiler, P.; Poulikakos, D. Effect of washcoat
diffusion resistance in foam based catalytic reactors. Chem. Eng. J. 2015, 276, 388–397. [CrossRef]
65. Bergman, T.L.; Lavine, A.S.; Incropera, F.P.; DeWitt, D.P. Fundamentals of Heat and Mass Transfer, 8th ed.; John
Wiley & Sons, Inc.: Hoboken, NJ, USA, 2017; ISBN 978-1-119-32042-5.
66. Glarborg, P.; Miller, J.A.; Ruscic, B.; Klippenstein, S.J. Modeling nitrogen chemistry in combustion.
Prog. Energy Combust. Sci. 2018, 67, 31–68. [CrossRef]
67. Messaoudi, H.; Thomas, S.; Djaidja, A.; Slyemi, S.; Chebout, R.; Barama, S.; Barama, A.; Benaliouche, F.
Hydrogen production over partial oxidation of methane using NiMgAl spinel catalysts: A kinetic approach.
C. R. Chim. 2017, 20, 738–746. [CrossRef]
68. Pruksawan, S.; Kitiyanan, B.; Ziff, R.M. Partial oxidation of methane on a nickel catalyst: Kinetic Monte-Carlo
simulation study. Chem. Eng. Sci. 2016, 147, 128–136. [CrossRef]
69. Neagoe, C.; Boffito, D.C.; Ma, Z.; Trevisanut, C.; Patience, G.S. Pt on Fecralloy catalyses methane partial
oxidation to syngas at high pressure. Catal. Today 2016, 270, 43–50. [CrossRef]
70. Schwiedernoch, R.; Tischer, S.; Correa, C.; Deutschmann, O. Experimental and numerical study on the
transient behavior of partial oxidation of methane in a catalytic monolith. Chem. Eng. Sci. 2003, 58, 633–642.
[CrossRef]
71. Hughes, K.J.; Turányi, T.; Clague, A.R.; Pilling, M.J. Development and testing of a comprehensive chemical
mechanism for the oxidation of methane. Int. J. Chem. Kinet. 2001, 33, 513–538. [CrossRef]
72. Turanyi, T.; Zalotai, L.; Dobe, S.; Berces, T. Effect of the uncertainty of kinetic and thermodynamic data on
methane flame simulation results. Phys. Chem. Chem. Phys. 2002, 4, 2568–2578. [CrossRef]
73. Aseem, A.; Harold, M.P. C2 yield enhancement during oxidative coupling of methane in a nonpermselective
porous membrane reactor. Chem. Eng. Sci. 2018, 175, 199–207. [CrossRef]
74. Gambo, Y.; Jalil, A.A.; Triwahyono, S.; Abdulrasheed, A.A. Recent advances and future prospect in catalysts
for oxidative coupling of methane to ethylene: A review. J. Ind. Eng. Chem. 2018, 59, 218–229. [CrossRef]
75. Kee, R.J.; Rupley, F.M.; Meeks, E.; Miller, J.A. CHEMKIN-III: A Fortran Chemical Kinetics Package for the Analysis
of Gasphase Chemical and Plasma Kinetics; Report No. SAND96-8216; Sandia National Laboratories: Livermore,
CA, USA, 1996. [CrossRef]
Processes 2018, 6, 83 24 of 24

76. Coltrin, M.E.; Kee, R.J.; Rupley, F.M.; Meeks, E. SURFACE CHEMKIN-III: A Fortran Package for Analyzing
Heterogeneous Chemical Kinetics at a Solid-Surface-Gas-Phase Interface; Report No. SAND96-8217; Sandia
National Laboratories: Livermore, CA, USA, 1996. [CrossRef]
77. Bodke, A.S.; Bharadwaj, S.S.; Schmidt, L.D. The effect of ceramic supports on partial oxidation of
hydrocarbons over noble metal coated monoliths. J. Catal. 1998, 179, 138–149. [CrossRef]
78. Hunt, G.; Karimi, N.; Torabi, M. Two-dimensional analytical investigation of coupled heat and mass transfer
and entropy generation in a porous, catalytic microreactor. Int. J. Heat Mass Transf. 2018, 119, 372–391.
[CrossRef]
79. Venvik, H.J.; Yang, J. Catalysis in microstructured reactors: Short review on small-scale syngas production
and further conversion into methanol, DME and Fischer-Tropsch products. Catal. Today 2017, 285, 135–146.
[CrossRef]
80. Faridkhou, A.; Tourvieille, J.-N.; Larachi, F. Reactions, hydrodynamics and mass transfer in micro-packed
beds—Overview and new mass transfer data. Chem. Eng. Process 2016, 110, 80–96. [CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like