You are on page 1of 87

Department of Chemical and Biomolecular Engineering

Modelling and simulation of water gas shift reaction in a packed-bed membrane reactor
system
Wu Chengliang (A0086696H)
Final Report
In partial fulfilment of the requirements for the Bachelor of Engineering (Chemical) Degree.

Foreword
So, there was an Asian guy, you see, hunkered down at a table, in a cafe in Israel. Hes the
only one whos doing any hardcore studying there, with the obligatory, presence-justifying
cappuccino. Hes processing these equations which no layman really understands, with lots of
triangles, dots, and d-something over d-something. Hes reading up on these things called
Fluid Mechanics, Reaction Engineering, Particle Technology and Numerical Simulation. He
seems unsure of whats going to happen, and appears to be contemplating as to whether hed
bitten off more than he can chew.
Fast forward to December, this guys more or less got his simulations in place. They take
approximately 5 minutes to solve each, the rainbow plots look beautiful, and hes got a whole
new depth of understanding in an alien CFD software.
Its been a long, long journey, but a very meaningful FYP. Something that integrates 4-5
courses worth of content is by no means an easy endeavour, compounded by the youre-onyour-own nature of a computational FYP. Ive had a couple of hair-tearing and sleepless
nights, but I guess the patchwork quilt of data, tips and guidance from 1001 sources from
Singapore to Romania, came together eventually.
Thanks aside, during this intensive period of studying and simulating, I have bumped into
many sources of literature where the content appeared excessively-complex, and was just
plain impossible for an undergraduate, much less a layman, to digest. Intentional or not,
nobody, Professor, Graduate, Undergraduate, should have to endure incomprehensible
content. Therefore, every attempt has been made in this report to explain concepts that are
relatable to undergraduate chemical engineering content, so that anybody who reads this
thesis doesnt have to bang his head against the wall so much.
Chengliang, WU
2

Abstract
Packed-bed membrane reactors (PBMRs) have shown promise in improving conversions of
equilibrium reactions, through their ability to remove the product as it is formed, forcing the
forward reaction to be favoured even further.
The research group has expressed interest in understanding the various phenomena, as well as
sensitivity of hydrogen production/CO conversion to various parameters, in one such PBMR
for the water-gas shift (WGS) reaction. This PBMR utilizes a Pd-Ag alloy hollow fiber
membrane separation module and Ni-Cu catalyst particles, and operates under experimental
conditions stipulated by said group.
To this end, a 2D-axisymmetric mathematical model has been developed and simulated in a
Finite Element Method (FEM) solver, COMSOL Multiphysics (Ver 4.4), under said materials
and conditions. The model is capable of describing concentration, velocity, density, viscosity,
and pressure profiles inside the reactor in both radial and axial coordinates, and its behaviour
has been validated, and determined to be consistent with phenomena expected of a PBMR.
In addition, the model has also performed a sensitivity study, inspecting the effect of 8
different parameters on the reactor performance. These parameters include reaction
temperature, flow rate, steam-carbon ratio, and product presence in the feed. The individual
studies produce and compare H2 production/CO conversion/CO equilibrium conversion in
response to parameter variations, and culminate in recommendations for operating said
reactor.
The objectives, observations, and conclusions have been placed in table form towards the end
of this paper, as a concise reference for the readers convenience. In addition, a guide is
appended at the end (Appendix II) to facilitate replication of this model for further study,
along with advice to achieve numerical convergence more quickly.

Contents
Foreword .................................................................................................................................... 2
Abstract ...................................................................................................................................... 3
1.

2.

3.

Background ......................................................................................................................... 6
1.1.

Water-Gas Shift Reaction and Membrane Reactors ................................................... 6

1.2.

Model Set-up ............................................................................................................... 7

1.3.

Objectives of Project ................................................................................................... 8

Theoretical Background/Literature Review ....................................................................... 9


2.1.

Fundamentals .............................................................................................................. 9

2.2.

Literature Review ...................................................................................................... 10

Modelling Process ............................................................................................................ 12


3.1.

Major Assumptions ................................................................................................... 12

3.2.

Key points ................................................................................................................. 13

3.3.

Modelling Approach (Shell Side) ............................................................................. 14

3.3.1.

Momentum Transport Darcys Law................................................................ 14

3.3.2.

Mass Transport and Reaction Maxwell-Stefan Diffusion .............................. 16

3.3.3.

Mass Transport and Reaction Heterogeneous Rate Law ................................ 17

3.3.4.

Energy Transport Pseudo-homogeneous Assumption .................................... 19

3.4.

3.4.1.

Momentum Transport Navier-Stokes Equations ............................................ 20

3.4.2.

Mass Transport Ficks Law of Diffusion ........................................................ 21

3.4.3.

Energy Transport Homogeneous Fluid ........................................................... 22

3.5.

Boundary Conditions................................................................................................. 22

3.5.1.

Momentum Transport Boundary Conditions ..................................................... 23

3.5.2.

Mass Transport Boundary Conditions ............................................................... 24

3.5.3.

Sieverts Law Boundary Condition ................................................................... 25

3.5.4.

Heat Transport Boundary Conditions ................................................................ 26

3.6.
4.

Modelling Approach (Tube Side) ............................................................................. 20

Glossary of Symbols ................................................................................................. 29

Simulation of WGS Reaction First Study and Validation ............................................. 31


4.1.

Meshing and Solver Configuration ........................................................................... 31

4.2.

Model Study and Validation - Conditions................................................................. 32

4.2.1.

Momentum Transport Study and Validation ..................................................... 33

4.2.2.

Mass Transport Study and Validation................................................................ 37

4.2.3.
4.3.
5.

Heat Transport Study ......................................................................................... 44

Counter-Current versus Co-Current Configuration .................................................. 45

Simulation of WGS Reaction Detailed Sensitivity Studies ........................................... 48


5.1.

Experimental Conditions Set 1 (Effect of Temperature) ....................................... 48

5.1.1.

Results Set 1 (Effect of Temperature Concentration Profiles) ................... 50

5.1.2.

Results Set 1 (Effect of Temperature CO Conversion Profiles) ................. 51

5.2.

Experimental Conditions Set 2 (Effect of Flow Rates) .......................................... 53

5.2.1.

Results Set 2 (Effect of Flow Rates Hydrogen Concentrations).................. 53

5.2.2.

Results Set 2 (Effect of Flow Rates CO Conversion Profiles) .................... 55

5.3.

Experimental Conditions Set 3 (Effect of Steam-Carbon Ratio) ........................... 56

5.3.1.

Results Set 3 (Effect of SC Ratios Hydrogen Concentrations).................... 57

5.3.2.

Results Set 3 (Effect of SC Ratios CO Conversion Profiles) ...................... 59

5.4.

Experimental Conditions Set 4 (Effect of Shell Pressure) ..................................... 61

5.4.1.

Results Set 4 (Effect of Shell Pressure Hydrogen Concentrations) ............. 62

5.4.2.

Results Set 4 (Effect of Shell Pressure Conversion Profiles) ...................... 63

5.5.

Experimental Conditions Set 5 (Effect of Sweep Gas Rate) .................................. 64

5.5.1.
5.6.

Experimental Conditions Set 6 (Effect of Inlet H2 Presence) ................................ 67

5.6.1.
5.7.

Results Set 6 (Effect of Hydrogen Presence Conversion Profiles) .............. 68

Experimental Conditions Set 7 (Effect of Inlet CO2 Presence) .............................. 69

5.7.1.
5.8.

Results Set 5 (Effect of Sweep Rate Conversion Profiles) .......................... 65

Results Set 7 (Effect of Inlet CO2 Presence Conversion Profiles) ............... 70

Experimental Conditions Set 8 (Effect of Permeability) ....................................... 72

5.8.1.

Results Set 8 (Effect of Permeability Conversion Profiles)......................... 73

6.

Conclusions ...................................................................................................................... 75

7.

Future Work ...................................................................................................................... 77

8.

Acknowledgements .......................................................................................................... 78

9.

References ........................................................................................................................ 79

Appendix I Derivation of Rate Constant for Rate Law ........................................................ 81


Appendix II - Step-by-Step Modelling Guide in COMSOL 4.4 Update 1 .............................. 81

1. Background
1.1.

Water-Gas Shift Reaction and Membrane Reactors

The water-gas shift reaction is a widely-utilized industrial reaction used in the production of
hydrogen gas, and can be found in examples such as shift converters downstream of steammethane reforming units. It is a mildly-exothermic reversible reaction.
CO(g) + H2O(g)

CO2(g) + H2(g)

H = -41.8kJ/mol

It has been of interest to conduct this reaction in a membrane reactor, which is an integrated
reaction-separation device that removes H2 as it is being produced, allowing equilibrium
limits to be bypassed, and achieving greater production as a result. A pictorial representation
can be seen below:

Fig 1.1 Dynamics of a packed-bed membrane reactor (PBMR)


Multiple types of membrane reactors are available (Doraiswamy, 2014), but in the context of
this simulation, the packed bed membrane reactor, (PBMR), also known as the inert
membrane reactor (IMR) model, will be considered. This model consists of a packed bed

reactor in the annular region, containing an inert, selectively-permeable membrane tube as the
tubular region. Its schematic will be visualized and discussed in the next section.

1.2.

Model Set-up

Fig 1.2 The experimental set-up for the reaction.


The experimental set-up stated by the research group consists of a hollow fiber membrane
module inserted into a packed bed of fine catalyst particles (5Ni/5Cu supported on CeO2),
with bed porosity of 0.4, synthesized by Saw et. Al (2014). The feed, whose temperature and
flow rate varies between 400-600C and 50-100mL, is inserted into the annular region of the
PBMR, where the catalytic reaction occurs at 2 barg, and product H2 gas subsequently
diffuses into the gaseous sweep region, maintained at 1 atm, with a helium sweep gas set at
0.5m/s by the candidate. The default mode of operation is co-current.

1.3.

Objectives of Project

Through this process, the candidate hopes to achieve the following:


S/N
1

Objective
Produce and simulate a sufficiently-rigorous
model for the reactor.
Verify the operational advantage that a
membrane reactor offers, over an equivalent
reactor without membrane activity.

Achieved via
Modelling with reference to
appropriate literature sources.
2
Using a validation model,
comparing the outlet CO2
concentrations between 2 reactors;
1 whose membrane is enabled, 1
whose membrane is disabled.
3
Study and validate the general phenomena
Simulating a validation model,
associated with the experimental conditions,
based on median experimental
including velocity, density, and pressure profiles. conditions, and performing a study
of said parameters.
4
Study of effect of operating temperature on H2
Temperature Sweep
production and conversion.
5
Study of effect of residence time (flow rate) on
Flow Rate Sweep
H2 production and conversion.
6
Study of effect of steam-carbon ratio in feed on
Steam-Carbon Ratio Sweep
H2 production and conversion.
7
Study of effect of reaction (shell) pressure on H2 Pressure Sweep
production and conversion.
8
Study of effect of Helium gas sweep rate on CO Helium Gas Sweep
conversion.
9
Study of effect of inlet H2 on CO conversion.
H2 inlet concentration Sweep
10
Study of effect of inlet CO2 on CO conversion
CO2 inlet concentration Sweep
11
Study of effect of various membrane
Membrane permeability Sweep
permeabilities on CO conversion
Table 1.3.1 Various Objectives to be achieved in the Project
Four terminologies in the above table warrant further elaboration; (1) Sufficiently-rigorous
refers to modelling with minimal simplifying assumptions, to assume only when clearly
validated/not a major consideration in literature. A case in point is to use the momentum
conservation equation to create velocity profiles instead of assuming a constant velocity
profile, which is a common practice in undergraduate Reactor Engineering coursework. (2)
Sweep refers to performing the same simulation, under similar conditions, with variations to
a single variable to study its impact. CFD terminology refers to this as a Parametric Sweep,
henceforth the Sweep term. (3) H2 Production refers to H2 produced in terms of
concentration, in the tube side, considering that the objective of the membrane reactor is to
produce high purity, usable H2 (which is basically the H2 in the tube side). (4) Study refers
8

to identifying trends, as well as the sensitivity, between a certain parameter and the result (in
terms of H2 production or CO conversion or both, where appropriate), and providing a
recommendation at the end of the evaluation.

2. Theoretical Background/Literature Review


2.1.

Fundamentals

In typical undergraduate coursework, the end goal of a Reactor Engineering coursework is to


calculate conversion upon a certain reactor volume or catalyst weight. This approach
normally assumes 1D plug-flow behaviour, meaning that there is a constant, flat velocity
profile that has no r-component. This practice allows one to synthesize a mole balance
equation around a differential section of the plug flow, packed-bed reactor, and integrate the
constant velocity value into said equation. According to Doraiswamy (2014), such an
approach is valid for an isothermal situation where no axial or radial gradients exist
(temperature/velocity gradients in this case). However, this method of thinking falls apart in
this project, because PBMRs are characterized by marked compositions and temperature
radial profiles (Falco, Marrielli, & Iaquaniello, 2011), and therefore require a 2D approach at
the minimum. This claim is further justified by the non-isothermal nature of the water-gas
shift reaction, and fluid property alterations, such as density and viscosity, along the length of
the reactor due to reaction and departure of H2 species.
Therefore, 2D fundamental equations of momentum, mass, and energy transport have to be
solved in order to evaluate concentration profiles of individual species. The momentum
equations calculate pressure and velocity profiles in the reactor, which will be coupled to
mass and energy transport equations, to solve for concentration and temperature profiles.
Simultaneously, the mass and energy transport equations will register changes in fluid
properties (density/viscosity/temperature/heat capacities), in turn affecting the momentum

equations. Therefore, the coupling of these 3 aspects, in both the shell and tube side, will be
elaborated upon, and simulated in the next chapter, to enable the reader to fathom their
interplay.

2.2.

Literature Review

Modelling of PBMRs in the literature, specifically with a shell packed-bed and a tube sweep,
are generally few and far in between, mainly because there are many different configurations
of membrane reactors. This issue is exacerbated when the reaction in question is the watergas shift. As of this date, no literature source has been found which utilizes similar CFD
software (COMSOL Multiphysics), for the project PBMR.
With the lack of a direct reference to build upon, a fresh modelling approach is necessary.
Therefore, the scope of the review has expanded to include mainly membrane reactors of
different configurations/reactions, simulated in COMSOL, which contain useful modelling
information that will be deployed in the project PBMR. In this section, 3 major references
which helped in determining the modelling equations/practices in this project will be
discussed and their relevancies elaborated upon briefly.
Iyoha (2008) Modelling and simulation of high temperature water gas shift reaction.
Iyoha (2008), in his PhD thesis, modelled the water-gas shift reaction in the case of a nonpacked bed reactor, where the reaction was in the tube side, and the permeation was in the
shell side. His model, which considered only the reaction side, assumed laminar flow,
Maxwell-Stefan diffusion, and was isothermal. His results managed to achieve a good
agreement with his experimental data.
Iyohas context starkly varied from this version in terms of operating configuration (reaction
was on the tube side instead), packed bed presence (his had none), and isothermality (his
context is, this project is not). Nonetheless, his Maxwell-Stefan diffusivity data, which
10

reflects the interaction between all 4 components of the water-gas shift reaction, can be used
as a good approximation of mass diffusion in the shell side of the project PBMR, given the
good agreement with experimental data.
Carcadea, Varlam, & Stefanescu, (2012) Heat transfer modelling of steam methane
reforming
Carcadea et.al (2012) utilized a model which was much closer to the project PBMR, using a
shell side packed bed (albeit with a serpentine shape, which is not practised here) and a tube
side sweep as well. However, the major difference is that the reaction in their model is the
steam-methane reforming reaction, and not the water-gas shift.
Nonetheless, there were useful observations made; the work deployed Darcys Law in the
packed bed, along with Maxwell-Stefan diffusion, and the heat equation. In particular,
Caracadea et.als assumption of Darcys Law as the momentum transport equation in the
packed bed is verified by a separate article (Manundawee, Assabumrungrat, & Wiyaratn,
2011), which meant that Darcys Law is a credible momentum transport equation in the
packed bed. In addition, the heat equation is also a good lead to pursue for this project.
COMSOL Inc., (2008) Fixed Bed Reactor for Catalytic Hydrocarbon Oxidation
This demonstration model provided by COMSOL Inc, a simulation software company,
illustrated reacting flow through a packed bed with heat transfer effects (no membrane). The
noteworthy point is that a pseudo-homogeneous assumption was applied for heat transfer,
meaning that the catalyst pellets and fluid in the packed bed were assumed to be a single
phase. This simplified the modelling, but required an effective thermal conductivity for this
single phase. The pseudo-homogeneous assumption will later be deployed in the modelling of
the project PBMR.

11

3. Modelling Process
Drawing upon the theoretical background acquired earlier, the modelling equations will be
applied. A summary of the modelling approach is enclosed here.

Fig 3: Modelling Summary

3.1.

Major Assumptions

These following assumptions will be deployed in the entire model. Subsequent assumptions
made will be specific to the context of that particular phenomenon to facilitate structure in
explanation.
Ideal Gas Law:
Owing to the high temperature (>500C), and relatively low pressure (<10 bar), the fluid is
assumed to be an ideal gas.
No electromagnetic phenomena:
According to Borman & Chuzhinov (1971), transport coefficients such as heat conductivity
vary with magnetic and electric fields. Owing to time constraints, and to simplify the

12

discussion, it will be assumed that these phenomena do not manifest to a significant extent.
Moreover, these effects were also not captured in the 3 references.
Gravity effects not considered:
To simplify the modelling process, it is assumed that the impact of gravity is negligible.
Similarly, these effects were not captured in the references.
Properties do not vary in the angular direction:
This allows the model to be simplified to a 2D, axisymmetric model, drastically-reducing
computational time. This approach has been validated in the literature, including (Carcadea et
al., 2012; Iyoha, 2007).
Modelling is steady-state:
All time-dependent derivatives are automatically eliminated from the governing equations, as
there are no transient properties necessitating inspection.
Membrane is reflected as having negligible thickness in the model:
This is to facilitate ease of visual comparison between shell and tube sides. Also note that the
effect of thickness on conversion/hydrogen production has been absorbed into the membrane
permeability term, which will be discussed downstream of this report as a sensitivity study.

3.2.

Key points

Before the modelling approach is discussed in-depth, the reader is requested to consider that:
1. All velocities listed in the governing equations are velocity fields, and are vector
quantities, unless otherwise stated (for instance, through the z suffix, to indicate the
z-direction velocity). Therefore, the word velocity refers to the vector quantity.

13

2. The gradient and divergence operators operate in a cylindrical geometry. However,


because of the axi-symmetric nature, angular derivatives (theta) are eliminated.

3.3.

Modelling Approach (Shell Side)

3.3.1. Momentum Transport Darcys Law

As the shell side of the reactor consists of a packed bed with spherical catalyst pellets, the
governing equations describing flow and pressure drop must be compatible with this
phenomenon. Before determining the appropriate transport phenomena, some understanding
of flow properties has to be applied.
Laminar Flow
The flow is assumed to be laminar, owing to the low velocity of the fluid. This property can,
and will be verified using the simulation results downstream.
Incompressible Flow
As the reactor is operating at a moderate pressure (2 barg maximum), and that Mach number
(that is, the fluid velocity compared to the speed of sound) is easily less than 0.3,
incompressible flow is assumed. This is not to imply that fluid density is an absolute constant
throughout the reactor; incompressible flow merely decouples pressure from density. Owing
to H2 diffusion, the velocity profile is expected to be impacted. This will be reflected in terms
of changes to gas dynamic viscosity and density, which are molar averaged values of their
individual components, as per the following equations:

Where ,

and

refer to dynamic viscosity, molecular weight, and density respectively.


14

Darcys Law
Darcys Law is a simplified version of the Navier-Stokes equations. Darcys law assumes
laminar flow, an incompressible fluid, and considers bed porosity and permeability. This law
generally applies to low velocity fluids (i.e. Rep < 10), and has been utilized by similar IMR
models,

including

Carcadea,

Varlam,

&

Stefanescu,

(2012)

and

Manundawee,

Assabumrungrat, & Wiyaratn, (2011). The governing equations are as follows:


Continuity

Discharge rate per unit area (m/s), or Darcy Velocity, or Superficial Velocity

Where

for a packed bed should be determined by the work of Carman-Kozeny (Reed,

2008), which states that:

Where K = 5 for packed beds, and the bed porosity, , is 0.4, according to the conditions of
the experiment for which this model is tied to.
It is important to note that Darcys Velocity is not the true velocity at which the fluid moves
(Honrath, 1995). This can be explained with a simple analogy; suppose fluid is leaving a
nozzle at velocity U. Thereafter, an object blocks part of the nozzle. As a result, the true
velocity of the fluid, u, is faster due to the smaller flow area, even though the discharge rate
per unit area (i.e. Darcys Velocity) remains unchanged.

15

Therefore, the true velocity, u, is Darcys Velocity divided by the porosity of the bed, as
expressed by:

As COMSOLs Darcys Law interface produces Darcy velocity fields, it is important to


correct the Darcy Velocity to the true velocity when integrating this physics with the
subsequent sections (heat and momentum balance).
3.3.2.

Mass Transport and Reaction Maxwell-Stefan Diffusion

As the reaction system is a concentrated one, where reactants are of comparable orders of
magnitude, and there is no single solvent, rules for a binary system (i.e. Ficks Law) cannot
be applied. The alternative is the Maxwell-Stefan diffusion, which describes mass transport in
multi-component systems. In such a system, for instance, a ternary system, where species A,
B, C are concerned, the interactions between A-B, A-C, and B-C have to be considered in the
evaluation of the molar fluxes. In a similar vein, the interactions between the 4 species in the
water-gas shift system have to be considered, which the Maxwell-Stefan equations are
capable of illustrating. Considering these interactions gives rise to the individual MaxwellStefan component mass transport equations, which are listed as follows, for each of the 4
water-gas shift components.
Component Mass Balance

Mass Flux Vector, relative to mass-averaged velocity, ji

16

Mass Flux Vector, relative to fixed axis,

The diffusional driving force in the relative mass flux vector,

Most prominently,

, can be represented as:

reflects the Maxwell-Stefan diffusivity between 2 distinct species, the

ith and kth. Its values for a water-gas shift system, is described by (Iyoha, 2007) as 2e-5 for all
species. As can be seen, the inter-species interactions have been considered.
One noteworthy point is the thermal diffusion coefficient,

, which is not to be confused

with thermal diffusivity. Thermal diffusion is the coupled effect between gradients of
concentration and temperature (Leahy-Dios, Zhuo, & Firoozabadi, 2008). This phenomenon
is also referred to as the Soret effect. Values for these coefficients have to be determined
experimentally, and none have been found for the water-gas shift reaction as of the time this
report is made. Therefore, they are assumed to be zero. While this inadvertently leads to some
loss in accuracy, this assumption is partially-validated by the mildly-exothermic nature of the
water-gas shift reaction; temperature gradients are expected to manifest, but should not
interfere with the diffusion phenomena too significantly. The validity of this assumption will
be evaluated in the Results section.
3.3.3.

Mass Transport and Reaction Heterogeneous Rate Law

The rate law in this context derives from the work of (Saw et al., 2014), which has been
converted to concentration basis instead of partial pressure basis, via the ideal gas law. This
measure is necessary to allow COMSOL, which produces concentration profiles throughout
the entire reactor, to calculate the rate of reaction. The original rate law is reflected here for
its conciseness, which allows further elaboration in the next section.
17

The original form is as below:


)]
There are a few noteworthy points in this rate law. Firstly, the value of the pre-exponential
factor

, was not published along with the rate law for unknown reasons. However, an

estimated value was back-calculated using the initial conditions from an experiment (see
Table 3 of Saws work), and was deduced to be ~5493 (units are not covered here to avoid
confusion). This may not be the exact value, but is in the same order of magnitude
nonetheless, which is critical for accuracy. The derivation is covered in Appendix I.
A second point which warrants elaboration is the effectiveness factor

, which is a factor

accounting for the mass transfer resistance encountered when the reactants diffuse from the
bulk fluid to catalyst surface. The effectiveness factor is limited to this context because the
catalyst particle has little pore, and therefore the assumption follows that mass transfer
resistance to reaction occurs solely through external mass transfer. Subsequently, the value of
0.9 has been agreed upon with one of the authors of the rate law to be a valid representation
of the effectiveness factor.
Thirdly, the

term represents the reversibility factor, which is common in catalytic rate laws

where equilibrium is concerned. It is a function of concentrations of individual species, in this


case,

Where:

The variables involved in the reversibility factor will also be treated as such; they will be
input into COMSOL as variables which have to be evaluated at every single position along
the reactor.
18

3.3.4. Energy Transport Pseudo-homogeneous Assumption

To evaluate heat transfer, the heat equation is deployed. The heat equation is essentially the
differential form of the First Law of Thermodynamics combined with Fouriers Law of heat
conduction. This equation can be solved separately from the Navier-Stokes equations because
of the incompressible assumption (Incompressibility means that an additional equation of
state relating density to pressure does not need to be solved, since they are decoupled). In
addition, there is an application of the pseudo-homogeneous assumption. This assumption,
commonly-used in modelling approaches where 2D concentration profiles within the reactor
are required (Gallucci, 2011), assumes that the catalyst pellets and fluid phase form a single
coherent phase. This assumption has been utilized in a packed-bed reactor model (COMSOL
Inc., 2008).

In this equation, the density, fluid heat capacity, true velocity, and enthalpy of reaction have
all been factored in. Details of the implementation will be reflected in Appendix II.
Of particular interest is the effective thermal conductivity,

term. This is an empirical

value which describes the combined thermal conductivity of the packed bed and its fluid.
Extensive literature studies have been performed for heat transfer in packed beds,
corresponding to different shapes, packing patterns, and reactor geometries. Owing to time
constraints, it is not possible to sieve out the most accurate one, and therefore a classic model
by Bruggemann (1935), as proposed by Madhusudana, (2014), is used, which is less
convoluted to use compared to most correlations. The model states the following
relationships:

19

Where

= Ratio of effective to fluid thermal conductivities, and

= Ratio of solid to fluid

thermal conductivities.
The following values were used to arrive at the value of .
Species
Ni-Cu catalyst

Thermal Conductivity
40 W/m.K

Fluid (mass-averaged
thermal conductivity, assume
5:1 steam-carbon ratio)

0.01664 W/m.K

Pseudo-homogeneous phase

35.46 W/m.K

Source
(Ho, Ackerman, Wu, Oh, &
Havill, 1978)
National Institute of
Standards and Technology,
US Dept of Commerce
(1984)
Calculated by Bruggemann
relation

Table 3.3.4: Values used in calculating Lambda.


As the value of effective thermal conductivity lies between fluid and catalyst particle
conductivities, the Bruggemann relation has proven to yield a logical value, which will be
used in the simulation.

3.4.

Modelling Approach (Tube Side)

The tube side is also known as the sweep region. Hydrogen diffuses through the membrane
from the shell to this region, where a flowing fluid, also known as a sweep gas, carries it
away for usage. In this context, helium gas will be used, for its inerting characteristic.
3.4.1. Momentum Transport Navier-Stokes Equations

This region assumes the incompressible Navier-Stokes equations, owing to the generally low
velocity of the sweep gas. These equations read:
Continuity

Momentum

20

Note that the density, velocity, viscosity and pressure values are appended with a 2 suffix to
indicate that they are in the tube side. The volume force term, F, which reflects gravity, is
ignored for simplicity purposes, as per the assumption in the earlier chapter.
3.4.2. Mass Transport Ficks Law of Diffusion

The tube side consists of helium gas and a much smaller quantity of hydrogen. Therefore, this
is a binary system, for which Ficks Law of diffusion is valid (Ficks Law applies to dilute
systems or binary systems). However, it is of interest to consider only the hydrogen
concentration profile in the reactor. Therefore, only one mass transport equation for hydrogen
is necessary. Note that H2 is referred to as component i for standardization purposes.

It is important to also note that the right hand side of the equation is zero. This is because
there is no reaction term in the sweep/tube region, and therefore the equation is purely a
transport-based version, with diffusion and convection transport modes reflected in the first
and second source terms of the equation.
The diffusion coefficient,

, in the tube side, reflects the diffusion of H2 through the sweep

gas. Its mass diffusivity can be evaluated by the correlation of Hirschfelder et.al (1949), as
proposed by (Welty, Wicks, Wilson, & Rorrer, 2008), where:

At 293K and 1 atm, Welty et al., (2008) has identified


Hirschfelder correlation, at 773K and 1 atm,

as 1.64cm2/s. Using the

becomes 7.02cm2/s. Note that this

relation assumes weak temperature dependency of the collision integral, a term which existed
in the original correlation (this claim was also verified by the earlier authors), allowing the
relation to be simplified to the above. In addition, there is an assumption that the temperature
21

does not vary significantly during the course of the reaction, so as to allow a fixed value of
diffusivity to manifest for simplification purposes. This assumption will later be proven valid.
3.4.3. Energy Transport Homogeneous Fluid

The energy balance in the tube side is more straightforward as the fluid is almost 100%
helium. Therefore, the same heat equation as seen in the shell side will be deployed, this time,
with the thermal conductivity of helium factored in. Also note that there is no heat source
term Q, due to the absence of reaction.

3.5.

Boundary Conditions

The boundary conditions for each of the 6 phenomena above will be described. As the model
is 2D, 3 sets of 4 boundary conditions each are needed. 3 sets reflect momentum, mass and
energy conservation equations, each set having 2 r and z-direction boundary conditions (4).
As stated in the initial section, the modelling is 2D-axisymmetric, meaning that COMSOL
will only solve a segment of the model, and perform a solid of revolution to get a 3D model,
as is visualized below. Therefore, the boundary condition images subsequently do not imply
that the modelling in COMSOL is done as per the images; these images are purely for
informative purposes.

Fig 3.5: 2D Axisymmetric method of solution


22

3.5.1. Momentum Transport Boundary Conditions

Fig.3.5.1: Momentum physics and boundary conditions


The boundary conditions for the momentum balances depend upon the experimental
conditions. Defining r = 0 as the centreline of the tube side, and z = 0 as the reactor base,
Shell Side (Darcys Law)
Inlet Velocity

Outlet Pressure

No Flow

No Flow

No-Slip

No-Slip

Tube Side (Navier-Stokes Equations)


Inlet Velocity

Outlet Pressure

23

3.5.2. Mass Transport Boundary Conditions

Fig 3.5.2: Mass transport physics and boundary conditions


The boundary conditions for the mass transport balances depend upon the experimental
conditions as well, and particularly affect the initial conditions. The reader is reminded that in
the shell side, due to the presence of 4 species, each of whom affects the rate law
calculations, all 4 mass transport equation sets have to be solved, whereas in the tube side,
there is only H2 which needs to be considered/evaluated, therefore only 1 set.
Shell Side (Maxwell-Stefan Diffusion)
Inlet Mass Fractions

Convective Dominance

Sieverts Law (H2), No


Flux (others)

No Flux

(H2 only)
(CO, H2O, CO2)
24

Tube Side (Ficks Law)


Inlet Concentration

Convective Dominance
.

Sieverts Law (H2)

Sieverts Law (H2)

3.5.3. Sieverts Law Boundary Condition

The seventh noteworthy phenomenon (aside from heat, mass, and momentum balances for
each side) is the Sieverts Law boundary condition. Sieverts Law basically describes the flux
of a species across a membrane, and is a function of the partial pressures of hydrogen on each
side of the IMR. In the context of hydrogen, the relationship is:

where

is an empirical value known as the base-case membrane permeability, and

membrane thickness. The actual membrane permeability is

is the

scaled via the Arrhenius

relation, as shown from the exponent term.


Owing to the lack of a permeability value for the membrane in question, an arbitrary value of
1.5e-5 (mol/m2Pa0.5s), from Iyoha (2007) will be used for the entire permeation term (this
factors in the membrane thickness as well) . A study of the impact of this parameter on CO
conversion will be discussed in Chapter 5.

25

3.5.4. Heat Transport Boundary Conditions

Fig.3.5.4: Heat transport physics and boundary conditions


The heat transport boundary conditions are slightly more unique; the wall surrounding the
shell is assumed to be of constant temperature, and therefore serves as a heat sink for the
exothermic water-gas shift reaction. The tube side is also another sink as there is no reaction.
Convective heat transfer coefficients have to be determined for 3 cases, wall to shell fluid,
shell fluid to membrane, and membrane to tube fluid as a result, which will be performed
immediately.
Shell Side (Pseudo-homogeneous Model)
Inlet Temperature

Convective Dominance

Heat Flux (to tube)

Heat Flux (to wall)

26

Tube Side (Homogeneous Model)


Inlet Temperature

Convective Dominance

Heat Flux (to shell)

Heat Flux (to shell)

Convective Heat Transfer Correlations


Wall to Shell fluid, and shell fluid to membrane wall
The shell region is an annulus, which is the space between 2 concentric cylinders. To
evaluate the heat transfer coefficient, Incropera & Dewitt, (2011) have cited the need to first
evaluate the ratio of the inner to outer diameter of the tube and shell regions, which is:

Based on this, the Nusselt numbers for fully-developed flow in a circular annulus with one
surface insulated, and the other at constant temperature can be evaluated according to the
text. The values are:

0.25
0.375
0.5

7.37
6.55
5.74

4.23
4.33
4.43

While there is some loss in accuracy due to the condition of an insulated surface, the loss in
accuracy is mitigated once again by the low temperature gain of the WGS reaction (to be
proven in the validation model), which means that heat transfer from shell to fluid will be of a
very low quantity, allowing the insulated surface condition to hold well.

27

Based on the above 2 relations, h1 and h2 are evaluated as 309684 W/m2K and 76770.9
W/m2K. It is important to note that these high values arise out of considering the shell phase
as pseudo-homogeneous, and therefore the effects of the high thermal conductivity of the bimetallic catalyst and the small diameter are prominently featured in these coefficients.
Membrane wall to tube fluid
In this context, fluid can be visualized as flowing in a closed conduit. To evaluate the heat
transfer coefficient, the pipe flow Reynolds Number has to be evaluated.

Fluid density, viscosity, and velocity are assumed to be identical to that of heliums, since
hydrogen is of a much lower concentration compared to helium. Based on helium property
data below, sourced from the National Institute of Standards and Technology, US Dept of
Commerce (1984),
Density
Velocity
Viscosity
Thermal Conductivity
Membrane tube diameter

0.524
0.5
1.81*10-5
0.1513
0.0015

Kg/m3
m/s
Pa.s
W/(m.K)
m

The Reynolds number is calculated to be 22. This indicates that flow is laminar (Red <
2300), verifying that the physics employed in the tube side is feasible.
The next step is to evaluate whether the fluid flow is fully-developed or not. This can be
evaluated via the entrance length relation, which states that:

28

The entrance length,

, is determined to be 0.002m. As this value is smaller than the length

of the entire tube, this means that the fluid flow is fully-developed. This result is also
verifiable in the simulation results.
For laminar, fully-developed flow in a closed conduit, with a constant heat flux, the Nusselt
number (Nu) is 4.364, as described by (Thirumaleshwar, 2006). Since the Nusselt number is
the ratio of convective to conductive heat transfer, which is:

The convective heat transfer coefficient for membrane wall to tube fluid, h, is determined as
439.78 W/(m2K).

3.6.
Symbol

Glossary of Symbols
First Appears in
All phenomena
All phenomena
Darcys Law
Darcys Law
All phenomena
Darcys Law
Darcys Law
Darcys Law
Maxwell-Stefan
Maxwell-Stefan
Maxwell-Stefan
Maxwell-Stefan
Maxwell-Stefan
Maxwell-Stefan
Maxwell-Stefan
Maxwell-Stefan
Maxwell-Stefan
Rate Law

Meaning
Gas Viscosity (Pa.s)
Gas Density (kg/m3)
Discharge Rate per unit area or Darcy Velocity(m/s) Field,
not to be confused with fluid velocity field (u)
Permeability (m2/s)
Absolute Pressure (Pa)
Bed Porosity (unitless) Value of 0.4
Empirical constant for determining bed permeability
(unitless)
Catalyst particle diameter (m)
Mass reaction rate of species i (kg/m3.s)
True fluid velocity field (m/s)
Mass fraction of species i (unitless)
Relative Mass Flux Vector, relative to mass-averaged
velocity (kg/m2s)
Diffusional driving force of species k (m-1)
Maxwell-Stefan diffusion coefficient (m2/s)
Mass Flux of component i, relative to a fixed axis (kg/m2.s),
not to be confused with molar flux (which is not bolded)
Mole fraction of component i (unitless)
Thermal Diffusion Coefficient for component i (zero in this
case, due to mild temperature changes)
Effectiveness Factor to account for mass transfer resistance
to catalyst surface (unitless)
29

Rate Law
Rate Law
Rate Law
Rate Law
Rate Law
Rate Law
Rate Law
Rate Law
All phenomena
All phenomena
Rate Law
Rate Law
Energy Balance
Energy Balance
Energy Balance
Tube Side - Navier
Stokes
Tube Side - Navier
Stokes
Tube Side Ficks
Law
Tube Side Ficks
Law
Boundary
Conditions
Sieverts Law
Sieverts Law
Sieverts Law
Sieverts Law
Sieverts Law
Sieverts Law
All boundary
conditions
Conversion

Catalyst surface area per unit volume (m2/m3); for a catalyst


particle of diameter 187.5 , translates to 32000 m2/m3
Reaction rate for the water gas shift reaction (mol/m3.s)
CO partial pressure (atm)
H2O partial pressure (atm)
H2 partial pressure (atm)
CO2 partial pressure (atm)
Pre-exponential factor (rendered unitless)
Activation energy for the catalyzed water gas shift reaction
(J/mol)
Molar Gas Constant (J/mol K)
Shell/Reaction side temperature (K)
Reversibility Factor (unitless)
Equilibrium constant (unitless)
Constant pressure heat capacity (J/kg.K)
Effective thermal conductivity (W/m.K)
Heat source term from reactions (W/m3)
Identity matrix
Volume Force (to account for effects of gravity) (N/m3)
Mass diffusivity of species i in k (m2/s)
Molar flux of component i, relative to a fixed axis
(mol/m2.s) Note that the mass flux above is bolded.
Length of reactor (0.05m)
Molar flux of hydrogen (mol/m2.s)
Base-case membrane permability (mol/(atm0.5.m.s))
(Units of pressure can change type depending on
experimenter)
Membrane thickness (m)
Membrane Activation Energy (J/mol)
Hydrogen partial pressure in shell side (atm) (assuming that
membrane permeability is based on atm units)
Hydrogen partial pressure in tube side (assumed to be 0)
Normal vector (unitless)
Molecular weight of an arbitrary species i (g/mol)

30

4. Simulation of WGS Reaction First Study and Validation


4.1.

Meshing and Solver Configuration

The presence of the membrane necessitates a higher resolution at the membrane boundary to
ensure accuracy of results. Therefore, a triangular mesh, with added refinement at aforesaid
boundary, is chosen, with a distribution of 300X (length) by 50X (width) on both shell and
tube sides, as seen. This was determined by trial-and-error to produce accurate results; further
increment resulted in a <1% change in results.

Fig 4.1 : Meshing of the PBMR: Notice the higher resolution at the membrane boundary.
The solver of choice is the PARDISO solver, as demonstrated by COMSOL Inc., (2008) in its
solution of a pseudo-homogeneous fixed bed reactor for hydrocarbon oxidation. PARDISO is
one of 4 finite element method (FEM) solvers available in the COMSOL interface.

31

4.2.

Model Study and Validation - Conditions

With the model completed, it is imperative to simulate and verify that the results are logical
and that the assumptions made earlier can be fulfilled. In addition, the validation can also
serve as a study to inform the reader of what phenomena to expect. Thus, an initial
experimental condition has been proposed, which takes the mid-range of temperature, and the
highest flow rate set by the research group. The rationale is explained in the Remarks column
for each table.
Note that the addressing of the second main objective, which is to verify that the membrane
reactor confers an operational advantage over a conventional annular packed-bed reactor, is
parked under Mass Transport Study and Validation for structure in report organization. The
reader may wish to skip to that section directly.
Shell Side
Boundary Parameter
Outlet Absolute Pressure

Inlet Mass Fractions

Value
3 bar

CO
H2O
H2
CO2

Inlet and wall


temperatures

500C

Inlet Feed Rate

100mL/min

0.1666
0.833
1e-6
1e-6

Remarks
Experimental condition is at
2 barg, therefore absolute
pressure should be ~3 bar.
5-1 steam-carbon ratio is
preserved, with a small
quantity of product gases
added for convergence
purposes.
Wall and inlet temperatures
are kept the same to preserve
temperature constancy.
The reactor is operated at the
highest rate set by the
research group to verify that
laminar flow persists.

Tube Side
Boundary Parameter
Outlet Absolute Pressure

Inlet Concentration

Value
1 atm

H2

1e-6 mol/m3

Remarks
Experimental condition:
Tube side to be maintained at
atmospheric pressure.
Small quantity of H2 added
for tube side convergence
32

Tube Inlet Temperature

500C

Inlet Feed Velocity

0.5m/s

Tube inlet temperature to be


kept the same as shell and
wall temperatures to preserve
temperature consistency.
Arbitrary value specified.

4.2.1. Momentum Transport Study and Validation

The earlier assumptions of laminar flow in packed bed, tube side, and fully-developed flow
are critical in determining the validity of Darcys Law as well as the heat/mass transfer
coefficients. Therefore, the flow properties will be inspected in detail, and the assumptions
verified where appropriate.
Firstly, pertaining to the validation model, the results for the values of viscosity, density, and
velocity can be visualized as follows:

Fig 4.2.1.1: Shell-side effective (mixture) viscosity under the validation conditions
Based on the results above, the viscosity is not expected to vary, or at least significantly, for
subsequent models.

33

Fig 4.2.1.2: Shell-side effective (mixture) density under the validation conditions
As can be seen, the diffusion of H2 across the membrane has produced a slight decrease in
density in the shell side, albeit somewat insignificant (~2%) in the context of this experiment.
Therefore, the reader may choose, at his discretion, to consider the system a constant-density
one, unless future experimental contexts dictate a significantly-increased rate of diffusion
across the membrane which will cause greater velocity variations.

34

Fig 4.2.1.3: Shell-side Darcy/Superficial velocity under the validation conditions. The left
legend reflects profiles in the tube; the right reflects profiles in the shell.
Due to a density decrease, the velocity has to increase from the shell inlet to the outlet, from a
simple mass balance as evidenced below. (In this context, the reader is requested to consider
the diffusion of H2 out of the shell as a decreasing-density reaction instead, to avoid
confusion.)

Therefore, the superficial velocity increase from ~0.15 to ~0.16 m/s is expected.
Validation of Darcys Law
To verify Darcys Law, the flow has to prove itself as laminar. Therefore taking the highest
attained values of density and viscosity (the legend indicated on the right hand sides of the

35

results in the earlier sections), and verifying using the particle Reynolds number for a packed
bed, as per expressed:

As evidenced, the particle Reynolds number at the highest throughput is easily less than 10,
which is the maximum for laminar flow. Therefore, the assumption of laminar flow in the
shell side is valid.
Validation of Pressure Drop

Fig 4.2.1.4: Shell-side pressure drop (centreline) under the validation conditions

36

The pressure drop in the shell centreline is around 5800 Pa as indicated by the simulation.
This can be verified using Erguns equation, an empirical relation that calculates pressure
drop in packed beds.

Substituting all parameters in, including the median value of density, Erguns equation
predicts a pressure drop of 4816 Pa. As Erguns equation only serves as a rough estimate in
this context, considering the 2D nature of system compared to the 1D nature of Erguns
equation, as well as variations in fluid properties, the pressure drop as calculated by Darcys
Law can be considered agreeable with this empirical relation, having a deviation of ~ 17%.
Conclusively-speaking, based on the laminar flow verification and pressure drop agreements,
the Darcys Law physics is a good fit for this model.
Validation of Entrance Length
Special mention is needed for the velocity profile in the tube side; observe that the velocity
profile transits from flat to parabolic at around 0.002m mark, which is exactly the entrance
length that was calculated earlier in the evaluation of closed conduit heat transfer coefficient.
Therefore, the entrance length calculated is validated in the simulation result.
4.2.2. Mass Transport Study and Validation

Studying the mass transport outcomes (concentration/conversion profiles) is more relevant in


the 8-parameter sensitivity study. Therefore, this section will be dedicated to verification
purposes.

37

Validation of Membrane Reactor Performance


The critical factor in evaluating the performance of a membrane reactor lies in its ability to
increase the overall conversion, compared to a regular annular packed-bed reactor; hence it is
listed as a key objective. This model has successfully-demonstrated the operational
advantage. Recall that the water-gas shift is of the form:
CO(g) + H2O(g)

CO2(g) + H2(g)

And that a membrane reactor is hypothesized to shift the reaction position to the right, as
opposed to a packed bed reactor. Therefore, if 2 models were to run under the same
conditions, one with the membrane enabled, one with the membrane disabled, the CO2
concentration at the exit of the reactor should be higher in the membrane reactor, since the
equilibrium position is shifted more towards the right. H2 concentration is not chosen because
the basis of comparison becomes more complex, given that H2 flux has to be evaluated on
both shell and tube sides.
Note that the system does not necessarily have to reach equilibrium to prove the operational
advantage; as long as membrane action exists, Le Chateliers principle will manifest in the
shell side and there will be greater production of CO2, which is the ultimate indicator of an
operational advantage.

38

Fig 4.2.2.1: Centerline concentration profile of CO2 (membrane enabled vs disabled)


With the membrane disabled (i.e. a normal annular packed bed reactor), the exit
concentration of CO2 is 0.203 mol/m3. However, once the membrane is enabled, the
centreline tube concentration at the reactor exit increases to ~0.205 mol/m3. In simpler terms,
once the membrane is enabled, the overall equilibrium position shifts to the right even more,
as evidenced from the increase in product CO2 concentration at the exit. The difference is
small owing to the small size of the reactor and the low permeability value used from the
literature, but nonetheless proves that the model in this context has captured the operational
advantage.

39

Validation of H2 diffusion phenomenon


Evaluating any form of results requires the flow profile of H2 to adhere to a logical format,
that is, it has to be produced in the shell side, and subsequently diffuse into the tube side of
the reactor. To verify that this has occured, a H2 concentration profile was evaluated and
plotted along the centre position of shell and tube sides, against the length of the reactor (zcoordinate).

Fig 4.2.2.3: Center-line Concentration profile of H2 compared to CO2, verifying flux.


As shown in the picture, H2 was produced in the shell side, as noted from its initial
concentration, and subsequently diffused into the tube side, as seen from the grey line, which
is always lower than the red. The diffusion effect is also substantiated by the lower

40

concentration of H2 relative to that of CO2 in the shell side. Therefore, this observation
validates the models ability to capture the flux effect of H2 from the shell to the tube side.
Validation of H2 radial concentration profile
The radial concentration profiles are also evaluated midway through the length of the reactor
(z=0.025m). It is observed that at the shell side, the concentration dipped towards the
membrane boundary. At the tube side, concentration was high at the boundary, and dipped
towards the bulk fluid. These observations are valid according to the two-phase film theory in
mass transfer, which suggests a decrease in hydrogen concentration from shell bulk fluid
(centre-line), to the membrane interface, and from said interface to the tube bulk fluid
(centre-line).

Fig 4.2.2.4: Shell Side: Concentration profile of H2 dips from bulk fluid towards membrane
boundary (r-coordinate 0.00075m)

41

Fig 4.2.2.5: Tube Side: Concentration profile of H2 dips towards from boundary (right-hand
side of the plot) towards bulk fluid (r-coordinate 0m)

Of particular interest is the concentration profile in the tube side; from the centre-line of the
tube axis, to the edge, the difference is miniscule (0.0001 mol/m3). This means that the
centreline concentration is a good gauge of the average concentration of H2 gas in the tube
side, and can be used as an indicator of H2 production in said membrane reactor.
Validation of Equilibrium effect/Study on maximum theoretical yield
Lastly, it is necessary to validate that equilibrium effects can manifest in the reactor, such that
equilibrium conversions can be captured for sensitivity studies downstream. Equilibrium in
this context means 2 things; (1) equilibrium in Sieverts Law (net flux from shell to tube side
reaches zero), simultaneously meaning (2) equilibrium in the water-gas shift reaction.
Therefore, these two criteria have to manifest in the reactor to ensure model robustness.
42

To capture the equilibrium effect, the residence time for reaction was increased by 600X
(flow rate 10ml, reactor length multiplied by 60 to 3.00m), and the simulation operated once
more. This multiplier is an arbitrarily-large value meant for verification purposes only.

Fig 4.2.2.6: Line concentration profile for H2 in shell and tube sides, with a heavily-extended
residence time to capture equilibrium effects.
The first observation is that the CO2 concentration has stagnated along with H2
concentrations, indicating/validating that the reaction rate is indeed pegged to the membrane
performance; as soon as flux to the tube side ceased, equilibrium position ceased to alter,
leading to a halt in CO2 production. This indicates that the model is sufficiently-viable to
study equilibrium conversions.
The second observation is that equilibrium in Sieverts Law has been achieved at around the
2.56m mark, which also indicates a maximum theoretical yield in the neighbourhood of
43

~2.00mol/m3 under the validation conditions. The sensitivity studies in the later sections will
indicate that the results achieved in the experimental context set by the research group are far
from this maximum theoretical yield, and sets a precedent for the reader to anticipate results
which are far from equilibrium conversion.
4.2.3. Heat Transport Study

Based on the results demonstrated below, the temperature profile in the reactor is virtually
isothermal. This is due to the mildly-exothermic nature of the WGS reaction, the excellent
thermal conductivity of the shell pseudo-homogeneous phase, and the fact that the wall and
tube fluid temperatures were kept identical to the inlet temperatures; causing the temperature
rise to be around 0.002K (not reflected in the diagram below due to program plotting
limitations).

Fig 4.2.3.1: Surface Plot temperature profile in tube and shell sides. The red spectrum of
773K reflects 773.002K, but is not reflected in COMSOLs plot.
44

Validation of temperature phenomena of WGS


The temperature profile at the shell side reflects hotspots, that is, a circular region of
relatively-higher temperature, which is common in exothermic reactions for which cooling is
applied. (COMSOL Inc., 2008). Therefore, the exothermic/cooling processes have been
successfully captured in the model.
Validation of Thermal Diffusion Coefficient and Hirschfelder Mass Transfer coefficient
Owing to the virtually-isothermal nature of this reaction, the claims of the thermal diffusion
coefficient being zero (see Section 3.3.2), as well as the Hirschfelder correlation assumption
(see Section 3.4.2), are validated.

4.3.

Counter-Current versus Co-Current Configuration

It is of interest to inspect whether there is any implication arising from co/counter-current


configurations (counter meaning sweep and feed originating from opposite directions), to
verify that the original co-current configuration is the recommended course of action.
Therefore, the same validation model will be once again applied. The figure of initial
reference will be Fig 4.2.2.3.

45

Fig 4.2.2.3 (or 4.3.1): Center-line Concentration profile of H2 compared to CO2, co-current.
The counter-current configuration was achieved by simply swapping the boundary conditions
in the tube side for momentum, mass, and energy balances, from one side to another. The
result is as follows:

46

Fig 4.3.2: Center-line Concentration profile of H2 compared to CO2, counter-current.


Based on the two plots, it can be seen from the exit concentration of CO2 that there was
virtually no alteration in equilibrium position by switching to a counter-current flow in the
context of this experiment, in the sense that the exit concentrations of CO2 were almost
identical. However, note that this is strictly-limited to the experimental context.
Secondly, the counter-current configuration was observed to produce a faster initial rise in
tube H2 concentration, as opposed to the co-current one. For instance, at the 0.004m mark in
the co-current configuration, the tube H2 concentation was 0.005mol/m3, whereas at the
0.046m mark in the counter-current configuration, the tube H2 concentration was
0.001mol/m3. Such a phenomenon is expected, mainly because of the stronger initial
concentration gradient offered by a counter-current mode of operation.

47

However, as far as overall performance is concerned, the co-current configuration outperformed the counter-current configuration; the co-current configuration registered an outlet
H2 concentration of 0.56mol/m3, whereas the counter-current configuration registered
0.50mol/m3, and produced a declining concentration nearing the end of the reactor. Such a
phenomenon is expected in the counter-current configuration; since the exit of the tube (high
H2 concentration) corresponds to the entrance of the shell (low H2 concentration), H2 is
expected to diffuse from tube to shell at this region instead, therefore negating some of the H2
gain made. To mitigate this backflow, it is recommended to operate the shell at a higher
pressure (>2 barg), such that the difference between tube and shell H2 partial pressures is
reduced, cutting H2 flux back to the shell side. Having said that, the initial experimental
condition of co-current configuration remains preferable, and this mode of operation will be
practised for the detailed sensitivity studies in the next Chapter.

5. Simulation of WGS Reaction Detailed Sensitivity Studies


5.1.

Experimental Conditions Set 1 (Effect of Temperature)

With the modelling complete and validated, the sensitivity studies will be performed,
pertaining to Objectives 4 through 11. The first study is on the effect of inlet shell
temperature on hydrogen production and CO conversion. The conditions are similar to those
seen in the validation model, and are as follows:
Shell Side
Boundary Parameter
Outlet Absolute Pressure

Inlet Mass Fractions

Value
3 bar

CO
H2O
H2
CO2

0.1666
0.833
1e-6
1e-6

Remarks
Experimental condition is at
2 barg, therefore absolute
pressure should be ~3 bar.
5-1 steam-carbon ratio is
preserved, with a small
quantity of product gases
added for convergence
purposes.
48

Inlet and wall


temperatures*

673, 723, 773, 823, 873K

Inlet Feed Rate


Tube Side

75 mL/min

Boundary Parameter
Outlet Absolute Pressure

Value
1 atm

Inlet Concentration

H2

1e-6 mol/m3

Tube Inlet
Temperature*

673, 723, 773, 823, 873K

Inlet Feed Velocity

0.5m/s

Wall and inlet temperatures


are kept the same to preserve
temperature constancy.

Remarks
Experimental condition:
Tube side to be maintained at
atmospheric pressure.
Small quantity of H2 added
for tube side convergence
Tube inlet temperature to be
kept the same as shell and
wall temperatures to preserve
temperature consistency.
Arbitrary value specified.

* Denotes a parameter whose effects are to be studied

The method of inspection will be through center-line profiles, marked in blue in the picture
below. This method of inspection will be deployed for the rest of the studies as well.

Fig 5.1: Center-line plot analysis. Note that this center-line approach was also deployed for
the previous examples.
49

5.1.1. Results Set 1 (Effect of Temperature Concentration Profiles)

The effect of varying operating temperatures on the yield of hydrogen is reflected as below.

Fig 5.1.1.1: H2 concentration profile in shell and tube sides, study of inlet temperature effect.
Based on the above plot, it can be seen that a temperature increase enhances the forward
reaction based on the increased yield of hydrogen in shell and tube sides. While this may
seem counter-intuitive based on the exothermic nature of the water-gas shift, it has been
verified that at this temperature, the kinetic effect overrides the thermodynamic effect. This
can be checked using the rate law supplied, through evaluating its constituent parameters, to
determine the cumulative effect on the rate of reaction (see column rate constant *(1-B))
T (K)

Rate Constant
KEQ
Beta
1-Beta
Rate Constant * (1-B)
673
3.41
11.85
0.084
0.92
3.12
773
8.86
4.91
0.203
0.80
7.06
873
18.51
2.49
0.401
0.60
11.09

Table 5.1.1.2: Evaluation of kinetic versus thermodynamic effects.


50

In the above table, it can be seen that a temperature increment of 200K raised the kinetic
effect (rate constant), by 6X, whereas the thermodynamic effect, (1-Beta), reduced by 33%,
therefore, it comes as no surprise that the rate of reaction receives a net enhancement from an
increase in temperature. With an increase in rate of reaction, equilibrium position is further
rightward at the end of the reactor, which explains the higher hydrogen production achieved.
Based on the data provided, the tube side hydrogen yield ranged from 0.04-0.09 mol/m3. As
stated earlier in the verification section, these results may be taken as an indicator of the
average hydrogen concentration in the tube side. In addition, this value range can be used as
an order-of-magnitude estimate for the reader in downstream sections.
5.1.2. Results Set 1 (Effect of Temperature CO Conversion Profiles)

The conversion profiles for CO in the shell side is defined by (Iyoha, 2007) as:

As can be seen, calculating the inlet mole fraction of CO, which is

, requires a

conversion from mass to mole fractions, given that the Maxwell-Stefan relations have been
used previously. The mole fraction of any species i,

, can be evaluated via the following

relation:

As an example, for steam-carbon ratio of 5:1, the molecular weight of the mix is 19.66g/mol.
Applying the above relations, the conversion profiles of CO are as follows:

51

Fig 5.1.1.2 : CO conversion profile in shell side, Equilibrium conversions are labelled EC.
Conversion of CO is generally low based on the experimental conditions, which is expected
given the low H2 yields compared to those seen in the validation section. As observed, the
range varies from 0.035 to ~0.09mol/m3. Equilibrium conversions were determined by
increasing the residence time by 1000X (flow rate of 75mL/min reduced to 0.075mL/min),
which created a plateauing conversion profile. The plot is not reflected here to avoid
confusion with the actual results, but the equilibrium conversions are reflected as ECs.
At 873K, equilibrium conversion of CO is observed to be 98%, and is therefore the most
ideal temperature to operate the shift reaction. On the other end of the spectrum, at 673K,
equilibrium conversion was significantly lower, clocking at 68.5%.

52

5.2.

Experimental Conditions Set 2 (Effect of Flow Rates)

The study will now switch to inspect the effect of flow rate, or more appropriately, residence
time, on H2 production. Experimental conditions are as follows:
Shell Side
Boundary Parameter
Outlet Absolute Pressure

Inlet Mass Fractions

Value
3 bar

CO
H2O
H2
CO2

0.1666
0.833
1e-6
1e-6

Inlet and wall


temperatures

500C/773K

Inlet Feed Rate*


Tube Side

50, 60, 70, 80, 90, 100 mL/min

Boundary Parameter
Outlet Absolute Pressure

Value
1 atm

Inlet Concentration

H2

Tube Inlet Temperature

500C/773K

Inlet Feed Velocity

0.5m/s

1e-6 mol/m3

Remarks
Experimental condition is 2
barg, therefore absolute
pressure should be ~3 bar.
5-1 steam-carbon ratio is
preserved, with a small
quantity of product gases
added for convergence
purposes.
Wall and inlet temperatures
are kept the same to preserve
temperature constancy.

Remarks
Experimental condition:
Tube side to be maintained at
atmospheric pressure.
Small quantity of H2 added
for tube side convergence
Tube inlet temperature to be
kept the same as shell and
wall temperatures to preserve
temperature consistency.
Arbitrary value specified.

* Denotes a parameter whose effects are to be studied

5.2.1. Results Set 2 (Effect of Flow Rates Hydrogen Concentrations)

The effect of varying operating inlet flow rates on the production of hydrogen was studied by
taking the centreline positions of shell and tube sides as well.

53

Fig 5.2.1.1 : H2 profile in shell and tube sides (effect of Flow Rate). Tube H2 concentrations
are reflected as dotted lines (not elaborated upon in the legend due to space constraints)
As the flow rate decreases, the residence time increases in the reactor. As the reactor is still
far from equilibrium, this means that there is further forward reaction. Interestingly enough, it
can be seen that increasing the residence time by a fixed quantity produces increasing returns,
as evidenced by the increasing hydrogen concentration. This can be further verified by
inspecting the CO2 concentration in Figure 5.2.1.2, which follows a similar progression of
increasing returns.

54

Fig 5.2.1.2 : Line flow rate CO2 profile in shell side. Note that CO2 remains trapped in the
shell side and therefore there is no tube profile.
Based on the plot in Figure 5.2.1.1, tube side H2 production ranges from 0.055 to
0.075mol/m3, for flow rates of 100mL/min to 50mL/min respectively. This value range is in
the same order of magnitude as seen in the temperature study, and serves as a source of
verification for the expected low hydrogen production in the reactor.
5.2.2. Results Set 2 (Effect of Flow Rates CO Conversion Profiles)

As seen in Section 5.2.1, the same expression for conversion was used, and the results
plotted. As the reaction is conducted only at 773K, the equilibrium conversion is 89.5%, or
0.895 (see Fig. 5.1.1.2).

55

Fig 5.2.1.3 : CO conversion profile in shell side, Equilibrium conversions are labelled EC.
In a similar vein to the earlier section, the conversion is low at the same order of magnitude,
ranging from 0.045 to 0.075. Therefore, a longer residence time is recommended, alongside
operating at 873K, to get the maximum CO conversion possible.

5.3.

Experimental Conditions Set 3 (Effect of Steam-Carbon Ratio)

The steam-carbon (SC) ratio is a common terminology, specifically-referring to the ratio of


H2O-CO at the inlet. In this Set, it is of interest to inspect the effect of varying SC ratios on
reactor performance. Note that the ratios are on a mass fraction basis, given the MaxwellStefan relations. The experimental conditions as per listed below will be deployed.
Shell Side
Boundary Parameter
Outlet Absolute Pressure

Value
3 bar

Remarks
Experimental condition is 2
barg, therefore absolute
56

Inlet Mass Fractions*

CO

H2O

Inlet and wall


temperatures

H2
CO2
500C/773K

Inlet Feed Rate


Tube Side

75mL/min

Boundary Parameter
Outlet Absolute Pressure

Value
1 atm

Inlet Concentration

H2

Tube Inlet Temperature

500C/773K

Inlet Feed Velocity

0.5m/s

0.1666, 0.2,
0.25, 0.333,
0.49999
0.8333,
0.79999,
0.74999,
0.6666,
0.49999
1e-6
1e-6

pressure should be ~3 bar.


The list of values on the
column to the left reflects the
various mass fractions
required, for each steamcarbon ratio.
E.g. CO, 0.1666, paired with
H2O, 0.8333, makes a ~5:1
SC Ratio.
Wall and inlet temperatures
are kept the same to preserve
temperature constancy.

1e-6 mol/m3

Remarks
Experimental condition:
Tube side to be maintained at
atmospheric pressure.
Small quantity of H2 added
for tube side convergence
Tube inlet temperature to be
kept the same as shell and
wall temperatures to preserve
temperature consistency.
Arbitrary value specified.

* Denotes a parameter whose effects are to be studied

5.3.1. Results Set 3 (Effect of SC Ratios Hydrogen Concentrations)

The hydrogen concentrations are as follows:

57

Fig 5.3.1.1 : Hydrogen shell and tube concentrations under varying Steam-CO Ratios.

As steam-carbon ratio tends towards 1:1, the tube and shell sides generally show increasing
H2 concentrations; in particular, at the tube side, a 1:1 SC yields a tube concentration of
0.10mol/m3, whereas a 5:1 SC yielded 0.06mol/m3, a 40% drop. This observation can be
attributed to the 1:1 stoichiometric ratio of the CO:H2O reaction; the initial validation
condition which proposed a 1:5 ratio made CO the limiting reactant, and caused H 2
production to be on the low side, simply due to the lack of CO. The closer the ratio became a
1:1, the more CO feed there was into the system, and the more H2 was produced as a result.
While a 1:1 ratio on paper sounds ideal, it is important to note that a S:C ratio has to be kept
above a certain threshold, otherwise there will be the risk of forming various side products.
As an example, in the steam-methane reforming reaction, where the WGS also occurs,
58

methane cracking can occur, leading to coking of the reactor. Annesini, Piemonte, &
Turchetti, (2002) state that to mitigate such an outcome, reactions practised in an industrial
context generally adhere to a 3:1 ratio for the steam-methane reforming process
(recommended ratios for the WGS could not be found in the literature, but considering its
manifestation in the reforming process, this value was used). Using this value as a
benchmark, it can be observed that the tube side hydrogen concentration is 0.076mol/m3,
which is still a respectable increase from 0.06mol/m3, a 27% increase. Therefore, in
accordance with industrial practices, it is recommended to operate the reactor at a 3:1 SC
ratio, instead of 5:1.
5.3.2.

Results Set 3 (Effect of SC Ratios CO Conversion Profiles)

The CO conversion profiles are reflected as follows in the below plot.

Fig 5.3.2.1 : CO Conversions under varying Steam-CO Ratios.


59

Fig 5.3.2.2 : Equilibrium CO Conversions under varying Steam-CO Ratios.


There are 2 points to discuss based on the above plots. The first is that as the SC ratio tends
towards 1:1, the faster the initial rate of reaction, which is expected by virtue of the 1:1
stoichiometric ratio. This explains the discontinuities observed at the entry point of the
reactor; at a 1:1 ratio, the CO conversion is already at ~0.047. The reader is encouraged to
think of this high starting point as a spike in conversion from 0.
The second is the cross observed at the 1:1 SC ratio; that is, initially, the 1:1 ratio sees the
highest conversion, but it becomes the lowest at the end as well. Further verification can be
seen by increasing the residence time by 10000X to capture equilibrium conversion (See Fig
5.3.2.2); the 1:1 ratio saw a dismal 71% conversion, compared to the 5:1 ratio which saw a
92% conversion. This appears to contradict the earlier discovery in Section 5.3.1, that a 1:1
ratio is the most ideal operating regime. However, there is actually no contradiction, because
(1) The absolute values of CO were different to begin with for each run, therefore, comparing
60

conversions to determine which SC ratio was ideal is inherently flawed and should not be
done, and (2) A high SC ratio (5:1) will naturally produce a higher conversion; since the ratio
of water to CO molecules is much greater, therefore, the chance of a CO molecule reacting
(and therefore CO conversion) is higher.
Conclusively-speaking, as far as operating regime is concerned; the lowest ratio permissible
by industry standards, that is, 3:1, is ultimately the way to go. This yields an 87.5%
conversion under the set experimental conditions, and produces a tube hydrogen
concentration yield that is 27% above the originally-quoted 5:1 ratio.

5.4.

Experimental Conditions Set 4 (Effect of Shell Pressure)

As observed in the validation section, shell pressure plays a role in Sieverts Law, governing
the net flux of H2 into the tube side, and ultimately affecting the yield available for usage.
Therefore, it is also of interest to assess the impact of this quantity on hydrogen production
and conversion as well. The experimental quantities are as below:
Shell Side
Boundary Parameter
Outlet Absolute
Pressure*
Inlet Mass Fractions

Inlet and wall


temperatures

Value
1 bar, 2 bar, 3 bar, 4 bar, 5 bar

Remarks
To be varied

CO
H2O
H2
CO2
500C/773K

5:1 SC Ratio

Inlet Feed Rate


Tube Side

75mL/min

Boundary Parameter
Outlet Absolute Pressure

Value
1 atm

Inlet Concentration

H2

0.1666
0.8333
1e-6
1e-6

Wall and inlet temperatures


are kept the same to preserve
temperature constancy.

1e-6 mol/m3

Remarks
Experimental condition:
Tube side to be maintained at
atmospheric pressure.
Small quantity of H2 added
for tube side convergence
61

Tube Inlet Temperature

500C/773K

Inlet Feed Velocity

0.5m/s

Tube inlet temperature to be


kept the same as shell and
wall temperatures to preserve
temperature consistency.
Arbitrary value specified.

Denotes a parameter whose effects are to be studied

5.4.1. Results Set 4 (Effect of Shell Pressure Hydrogen Concentrations)

Fig 5.4.1.1: Hydrogen Shell and Tube side concentrations under varying Shell Pressures
(units: mol/m3)
The observation here is that the higher the shell side pressure, the greater the shell and tube
hydrogen production. The tube case can be explained by the higher partial pressure of
hydrogen at the shell side, which resulted in a greater flux, according to Sieverts Law. The
shell increment in hydrogen concentrations, on the other hand, is harder to prove qualitatively
due to the complex rate law involved, but it can be seen from the results that the rate law
generally favors a higher pressure operation, as a higher shell pressure led to more H2
production at the exit of the shell side. It is also good to note that such a measure generally
62

produces diminishing returns, as can be seen from the shell and tube concentration profiles,
whose increments in H2 production decrease per bar increase in pressure. Therefore, it is
recommended to operate the process at a high pressure, but not to the extent that the costs of
maintaining a high pressure outweigh the benefits in hydrogen production. This can be
achieved via simulation of the scaled-up unit, with cost estimation studies factored in.
5.4.2. Results Set 4 (Effect of Shell Pressure Conversion Profiles)

In this context, conversions are generally not a useful indicator of reactor performance,
because the variations in pressure also lead to variations in velocity through the momentum
balance. Given that the conversion expression is a function of velocity as well as CO mole
fraction, the result becomes more convoluted and is not useful. That being said, it is still
useful to evaluate the equilibrium conversion, which is approximated in this case by
increasing the residence time by 10,000X, upon which the conversion profiles plateau-ed.

Fig 5.4.2.1: CO Equilibrium Conversions under different shell pressures


63

Based on the plot as shown above, the lowest pressure has the greatest conversion of 0.93,
whereas the highest has a conversion of 0.89. Considering that the difference in equilibrium
conversion is minute (and that equilibrium is unlikely to be totally reached in a practical
scenario, considering that throughput has to be extremely low), the recommendation of
operating at maximum yet economical pressure, stands as is.

5.5.

Experimental Conditions Set 5 (Effect of Sweep Gas Rate)

Helium sweep gas is used in the tube side, to push, or more appropriately, sweep out the
product H2 formed from the shift reaction. Research by Chein, Chen, & Chung, (2015) has
demonstrated improved reactor performance in terms of increased CO conversion, with
increased sweep flow rates. This is presumed to be due to the faster depletion of H2 in the
tube sides, which induces more hydrogen flux from shell to the tube side, and consequently
further conversion of CO. This claim will be verified with the following experimental
conditions.
Shell Side
Boundary Parameter
Outlet Absolute Pressure
Inlet Mass Fractions

Inlet and wall


temperatures

Value
3 bar
CO
H2O
H2
CO2
500C/773K

Inlet Feed Rate


Tube Side

75mL/min

Boundary Parameter
Outlet Absolute Pressure

Value
1 atm

Inlet Concentration
Tube Inlet Temperature

H2
500C/773K

0.1666
0.8333
1e-6
1e-6

Remarks
Experimental conditions
5:1 SC Ratio

Wall and inlet temperatures


are kept the same to preserve
temperature constancy.

1e-6 mol/m3

Remarks
Experimental condition:
Tube side to be maintained at
atmospheric pressure.
Small quantity of H2 added
for tube side convergence
Tube inlet temperature to be
64

Inlet Feed Velocity*

0.1, 0.3, 0.5, 0.7, 0.9m/s

kept the same as shell and


wall temperatures to preserve
temperature consistency.
To be varied

*Denotes a parameter whose effects are to be studied

5.5.1. Results Set 5 (Effect of Sweep Rate Conversion Profiles)

The conversion profiles will be discussed only as the primary objective is to confirm the
presence of increased CO conversion.

Fig 5.5.1.1: CO Conversions under different sweep rates; sweep at 0.9m/s shows the highest
CO conversion. Graph has been zoomed-in.
As evidenced from Fig 5.5.1.1, the highest sweep rate of 0.9m/s saw the highest conversion
of CO, verifying that the higher the sweep rate, the greater the conversion. The difference in
conversions is small in the context of this reactor due to its relatively-small size.

65

The simulation is run to equilibrium once more, by increasing the residence time by 10,000X
to get a plateauing concentration profile, so as to illustrate the impact of sweep rate on
equilibrium conversions.

Fig 5.5.1.2: Equilibrium CO Conversions under different sweep rates; sweep at 0.9m/s shows
the highest CO conversion. Graph has been zoomed-in.
From the study above, the effect of sweep rate on equilibrium conversion is not excessivelysignificant; a 9X increase in sweep velocity raised the equilibrium conversion from 0.896 to
0.924, with diminishing returns to boot. Taking into account that costs go up with increase in
helium sweep rate, it is recommended to operate at the most cost-effective regime, rather than
the regime with the highest permissible helium sweep rate.

66

5.6.

Experimental Conditions Set 6 (Effect of Inlet H2 Presence)

As stated earlier, the WGS is unlikely to occur on its own in an industrial application. For
instance, in the steam methane reformer (SMR), the primary steam-reforming reaction always
takes place first, resulting in the production of some hydrogen.

As evidenced from the SMR equation, hydrogen presence is guaranteed when the water-gas
shift reaction occurs subsequent to the SMR, and therefore it is of interest to inspect the
impact of existing H2 on CO conversion. Hence, the following experimental condition is
proposed. Inlet hydrogen fractions are kept on the low end, to ensure that CO conversions
remain within the same neighbourhood for easier visual comparison.
Shell Side
Boundary Parameter
Outlet Absolute Pressure
Inlet Mass Fractions

Value
3 bar
CO
H2O

H2 *

Inlet and wall


temperatures

CO2
500C/773K

Inlet Feed Rate


Tube Side

75mL/min

Boundary Parameter
Outlet Absolute Pressure

Value
1 atm

Inlet Concentration
Tube Inlet Temperature

H2
500C/773K

Maintained in a
5:1 ratio, after
subtracting H2
and CO2 mass
fractions.
0.01, 0.015,
0.02, 0.025, 0.03
1e-6

Remarks
Experimental conditions
5:1 SC Ratio, with variations
in H2 inlet fractions. Note
that inlet fractions are kept
low.

Wall and inlet temperatures


are kept the same to preserve
temperature constancy.

1e-6 mol/m3

Remarks
Experimental condition:
Tube side to be maintained at
atmospheric pressure.
Small quantity of H2 added
for tube side convergence
Tube inlet temperature to be
kept the same as shell and
wall temperatures to preserve
67

Inlet Feed Velocity

0.5m/s

temperature consistency.
To be varied

*Denotes a parameter whose effects are to be studied

5.6.1. Results Set 6 (Effect of Hydrogen Presence Conversion Profiles)

In a similar fashion to the above Set, only the conversion profiles will be considered in this
context as qualitative evaluation of H2 impact is more important. The results are displayed as
follows:

Fig 5.6.1.1: CO Conversion Profiles under different inlet mass fractions of hydrogen.
In the diagram above, the starting conversions are above zero owing to the presence of
hydrogen in the feed. Therefore, it is difficult to make a visual comparison. A more effective
method of ascertaining the impact of hydrogen in feed is calculating change in CO
conversion, which is tabulated below.

68

H2 Inlet Fraction
Initial Conversion
Final Conversion
0.03
0.202
0.211
0.025
0.178
0.188
0.02
0.152
0.163
0.015
0.1245
0.136
0.01
0.0945
0.108
Fig 5.6.1.2: Conversion Changes under different H2 fractions

Change
+0.009
+0.01
+0.011
+0.0115
+0.0135

As the inlet fraction of H2 increases (feed becomes less pure) by a fixed size, it can be seen
that the increase in conversion diminishes. In simpler terms, the operational advantage
offered by this PBMR diminishes the more the feed gets contaminated with H2. Therefore,
this PBMR is not viable for applications where the feed is likely to be impure, namely the
SMR application, where the bulk of the hydrogen is yielded from the reforming reaction, as
the diminishing return effect means that the PBMR will not be effective. That being said, it is
definitely possible to counterbalance this effect through measures such as improving the
membrane permeability. Such a measure increases the draw rate of hydrogen into the tube
side, and forces more CO conversion.
Conclusively-speaking, it is recommended to operate the PBMR in a context where hydrogen
will be of a negligible quantity in the feed, otherwise, further changes to the reactor will be
needed if it is to have an appreciable contribution to the WGS.

5.7.

Experimental Conditions Set 7 (Effect of Inlet CO2 Presence)

In a similar fashion to the purposes of Set 6, the implications of CO2 presence in the feed on
conversion will be studied as well. The conditions are as below.
Shell Side
Boundary Parameter
Outlet Absolute Pressure
Inlet Mass Fractions

Value
3 bar
CO
H2O

H2

Maintained in a
5:1 ratio, after
subtracting H2
and CO2 mass
fractions.
1e-6

Remarks
Experimental conditions
5:1 SC Ratio, with variations
in H2 inlet fractions. Note
that inlet fractions are kept
low.

69

CO2*
Inlet and wall
temperatures

500C/773K

Inlet Feed Rate


Tube Side

75mL/min

Boundary Parameter
Outlet Absolute Pressure

Value
1 atm

Inlet Concentration

H2

Tube Inlet Temperature

500C/773K

Inlet Feed Velocity

0.5m/s

0.01, 0.015,
0.02, 0.025, 0.03
Wall and inlet temperatures
are kept the same to preserve
temperature constancy.

1e-6 mol/m3

Remarks
Experimental condition:
Tube side to be maintained at
atmospheric pressure.
Small quantity of H2 added
for tube side convergence
Tube inlet temperature to be
kept the same as shell and
wall temperatures to preserve
temperature consistency.
To be varied

*Denotes a parameter whose effects are to be studied

5.7.1. Results Set 7 (Effect of Inlet CO2 Presence Conversion Profiles)

The results collected from the sweep are as below:

Fig 5.7.1.1: CO Conversion Profiles under different inlet mass fractions of CO2.
70

There are 2 observations to note here. The first is that the CO2 equilibrium quantity appears to
be much smaller than that of hydrogen. This is because at a small inlet quantity (mass
fraction ~0.0175, based on interpolation from Fig 5.7.1.1) and above, the initial conversion
becomes negative, that is, the backward reaction manifests, which is clearly undesirable, even
if there is a net positive conversion in the end. Therefore, the feed should ideally contain a
CO2 inlet mass fraction which is less than the aforesaid value.
The second observation is that it is not recommended to operate this reactor completely if the
inlet CO2 fraction exceeds 0.025, as the net conversion is less than 0.005 (or <1%). Above
this value, e.g. at 0.03 mass fraction, the conversion at the exit is fully-negative; the reactor is
facilitating a reverse water-gas shift reaction. Should this suggestion not be possible, a much
longer residence time should be needed to allow the system to reach closer to equilibrium, as
it can be seen from the below plot that equilibrium conversion, regardless of starting inlet
mass fraction, still hovers at the neighbourhood of 0.9.

Fig
5.7.1.2: Equilibrium CO Conversion Profiles under different inlet mass fractions of CO2.
71

The equilibrium conversions as can be seen, generally do not differ significantly based on the
inlet CO2 mass fractions (0.917 versus 0.907, 1% difference, between the lowest and highest
inlet fractions), however, this is considering that the mass fractions are low in value to begin
with. Differences in equilibrium conversion are expected to become more significant as CO2
inlet fraction increases.

5.8.

Experimental Conditions Set 8 (Effect of Permeability)

The membrane forms the linchpin of the PBMR by giving it an operational advantage over
the conventional annular packed-bed reactor, as seen in the validation section. It is easy to
qualitatively-reason that as membrane permeability increases, CO conversion will increase as
well. However, the sensitivity of conversion increment to permeability change is a topic of
interest. Therefore, the last study will be the effect of varying membrane permeabilities on
CO conversion. The experimental context will be as below, which is identical to the
validation condition.
Shell Side
Boundary Parameter
Outlet Absolute Pressure

Inlet Mass Fractions

Value
3 bar

CO
H2O
H2
CO2

Inlet and wall


temperatures

500C

Inlet Feed Rate

75mL/min

0.1666
0.833
1e-6
1e-6

Remarks
Experimental condition is at
2 barg, therefore absolute
pressure should be ~3 bar.
5-1 steam-carbon ratio is
preserved, with a small
quantity of product gases
added for convergence
purposes.
Wall and inlet temperatures
are kept the same to preserve
temperature constancy.
Median value of flow rate
(range 50-100mL) used. This
reasoning is retrospective.

Tube Side
Boundary
Parameter
Outlet Absolute

Value

Remarks

1 atm

Experimental condition:
72

Pressure
Inlet Concentration

H2

1e-6 mol/m3

Tube Inlet
Temperature

500C

Inlet Feed Velocity


Membrane
Permeability*

0.5m/s
0.375e-5,0.75e-5, 1.5e-5, 3e-5, 6e-5
(mol/m2Pa0.5s)

Tube side to be maintained at


atmospheric pressure.
Small quantity of H2 added
for tube side convergence
Tube inlet temperature to be
kept the same as shell and
wall temperatures to preserve
temperature consistency.
Arbitrary value specified.
Permeabilities proposed are
quarter, half, double,
quadruple that of the original
value, 1.5e-5.

*Denotes a parameter whose effects are to be studied

5.8.1. Results Set 8 (Effect of Permeability Conversion Profiles)

Fig 5.8.1.1: CO Conversion Profiles under different inlet membrane permeabilities.


Based on the results above, increased membrane permeability does improve CO conversion,
to the tune of increasing returns, that is, a 1X to 2X permeability yields more conversion
change, compared to a 0.5X to 1X permeability. However, in terms of absolute values, the

73

difference remains highly insignificant. At 4X the usual membrane permeability, conversion


increased barely from 0.05835 to ~0.0587 (0.6% improvement in conversion). Considering
that these values are in the same order of magnitude as those seen in the literature, and
therefore unlikely to see a drastic increment in permeability in the near future, it is
recommended to harness multiple membrane tubes instead of a single one within a reactor, to
gain a more prominent CO conversion advantage.
That being said, these differences become more significant as equilibrium is being
approached. At 100X the residence time, the difference is in the order of percentages in CO
conversion, indicating that a longer residence time will harness the positive effect of the
membrane tube to a greater extent.

Fig 5.8.1.2: CO Conversion Profiles under different inlet membrane permeabilities, 100X
residence time.

74

6. Conclusions
To sum up the report, a packed-bed membrane reactor has been studied and modelled with
reference to the appropriate literature. Appropriate boundary conditions and heat/mass
transfer correlations have been developed and integrated into the said model.
From the validation model proposed, it can be seen that the model holds good promise for
future studies as the momentum, mass, and heat transfer studies generally agree with the
assumptions made, and that the behaviour of the model adheres to that of a membrane
reactor, such as greater conversion and H2 transfer across the membrane.
In addition, the studies performed with regards to the 8 parameters have provided insights
into recommended operating regimes. However, CO conversion/hydrogen production under
these conditions is generally low, as a result of the small residence time afforded for reaction.
One recommendation for future work is for the research group to increase the residence times
in the experimental contexts, especially if the PBMR is to be scaled-up to industrial
production levels.
Returning to the list of 11 objectives seen in Section 1-3, the outcomes and conclusions are
tabulated as follows:
S/N
1

Objective
Produce and simulate a
sufficiently-rigorous model for the
reactor.

Verify the operational advantage


that a membrane reactor offers.

Study and validate the general


phenomena associated with the
experimental conditions,
including velocity, density, and
pressure profiles.

Outcome/Conclusion
2D-axisymmetric model developed, utilizing
fundamental differential equations of momentum,
mass, and energy balances, with appropriate
heat/mass transfer correlations from literature.
A PBMR is verified to produce more CO2
compared to an annular packed-bed reactor under
the same conditions.
Validation model operated, with the following
discoveries:
Momentum:
Laminar flow, pressure drop and entrance length
validated.
Velocity/Density not expected to vary
75

significantly based on set conditions.


Mass:
H2 departure from shell side confirmed.
Radial profiles consistent with two-phase mass
transfer phenomena, tube centre-line H2
concentrations found to be accurate indicator of
average H2 concentrations.
Reaction equilibrium effect captured in model.
Heat:
Reactor operation is discovered to be virtually
isothermal, Thermal diffusion coefficient value
/Hirschfelder relation validated.

Study of effect of temperature on


yield and conversion.

Counter-vs-Co-Current
Counter-current produces stronger initial rise in
H2 concentration, but co-current outperforms
counter-current in terms of H2 production in tube.
Higher inlet temperature favours conversion
owing to kinetic effect >> thermodynamic effect.
Recommended to operate at 873K due to 98%
equilibrium conversion.

Study of effect of residence time


(flow rate) on H2 production and
conversion.

Study of effect of steam-carbon


ratio in feed on H2 production and
conversion.

H2 production in the range of 0.04-0.09mol/m3,


which sets a range of expected values.
Lower flow rate favours conversion due to longer
residence time.
Increasing returns to be had per unit increment in
residence time.
Lower SC ratio favours more CO conversion/H2
production from a stoichiometry standpoint.

Study of effect of reaction (shell)


pressure on H2 production and
conversion.

However, a 1:1 ratio is not feasible due to


possibility of side reactions. Recommended to
operate at a steam:CO ratio of 3:1, an industrial
practice. This creates a 27% increase in H2
production compared to 5:1 ratio.
Hydrogen production in both shell and tube sides
increases with shell-side pressure increase, due to
rate law and Sieverts Law.

Study of effect of Helium gas

However, the returns are diminishing, and


therefore recommended to operate at a costeffective pressure.
CO conversion increases as sweep rate increases,
76

sweep rate on CO conversion.

10

Study of effect of inlet H2 on CO


conversion.

albeit not significant.


CO conversion increment diminishes as sweep
rate increases, therefore, recommended to operate
the sweep at a cost-effective rate, rather than ashigh-as-possible.
Inlet H2 presence diminishes operational
advantage conferred by PBMR, but does not
create negative conversion.

As inlet H2 presence increases, the operational


advantage diminishes. Therefore, this PBMR is
not appropriate for high H2 inlet contexts such as
a steam-methane reformer.
Study of effect of inlet CO2 on CO CO2 presence begins to create negative
conversion
conversion effect if inlet mass fraction >0.0175,
although there is still a positive net conversion.
PBMR will have a detrimental effect instead
(reverse WGS) if CO2 inlet mass fraction >0.025,
unless residence time is increased from the
experimental context to allow system to
equilibrate.

11

Equilibrium conversion generally insensitive to


inlet CO2 mass fractions.
Study of effect of various Increased membrane permeability raises CO
membrane permeabilities on CO conversion to the tune of increasing returns, but
conversion
effect is highly-muted in this context.

Recommended to operate PBMR utilizing


multiple membrane tubes to stack the positive
effects.
Table 6-1: List of objectives and corresponding outcomes.

7. Future Work
Based on the Gantt Chart as provided in the interim report, the project has reached total
completion.

Fig 7: Gantt Chart depicting project progress


77

Nonetheless, it is desirable to verify the work with an experimental set-up, which could not
be achieved by the research group within this period in time. Also, further optimization of the
heat and mass transfer coefficients can be done, as more relevant literature becomes
available.

8. Acknowledgements
The candidate would like to express his acknowledgements to:
1. Assoc Prof Kus Hidajat and Dr Usman Oemar for the FYP opportunity.
2. Dr Eldin Lim for CFD and convergence-related advice.
3. Dr Elena Carcadea for her guidance.
4. Authors of membrane reactor modelling texts.

78

9. References
Annesini, M. C., Piemonte, V., & Turchetti, L. (2002). Carbon Formation in the Steam
Reforming Process: a Thermodynamic Analysis Based on the Elemental Composition.
Borman, V. D., & Chuzhinov, V. A. (1971). The Influence of an Electric Field on the
Thermal Diffusion Coefficient for Gases, 33(5), 89.
Carcadea, E., Varlam, M., & Stefanescu, I. (2012). Heat Transfer Modelling of Steam
Methane Reforming. COMSOL Conference, 2012, Milan, (4).
Chein, R. Y., Chen, Y. C., & Chung, J. N. (2015). Sweep gas flow effect on membrane
reactor performance for hydrogen production from high-temperature water-gas shift
reaction. Journal of Membrane Science, 475, 193203.
doi:10.1016/j.memsci.2014.09.046
COMSOL Inc. (2008). Fixed-Bed Reactor for Catalytic Hydrocarbon Oxidation. Burlington,
MA: COMSOL.Inc.
Doraiswamy, L. K. (2014). Chemical Reaction Engineering, Beyond the Fundamentals. Boca
Raton, FL, USA: CRC Press.
Falco, M. de, Marrielli, L., & Iaquaniello, G. (2011). Membrane Reactors for Hydrogen
Production Processes (1st ed.). Springer. doi:10.1007/978-0-85729-151-6
Gallucci, F. (2011). Modeling of Membrane Reactors for Hydrogen Production and
Purification (Vol. 2, pp. 139). doi:10.1039/9781849733489-00001
Ho, C. Y., Ackerman, M. W., Wu, K. Y., Oh, S. G., & Havill, T. N. (1978). Thermal
Conductivity of Ten Selected Binary Alloy Systems. Journal of Physical Chemistry,
7(3). Retrieved from http://www.nist.gov/data/PDFfiles/jpcrd123.pdf
Honrath, R. E. (1995). Mass Transport Processes. Retrieved from
http://www.cee.mtu.edu/~reh/courses/ce251/251_notes_dir/node4.html
Incropera, F. P., & Dewitt, D. P. (2011). Fundamentals of Heat and Mass Transfer
(Seventh.). Jefferson, MI, USA.
Iyoha, O. (2007). H2 Production in Palladium & Palladium-Copper Membrane Reactors at
1173K in the presence of H2S. University of Pittsburgh.
Leahy-Dios, A., Zhuo, L., & Firoozabadi, A. (2008). New thermal diffusion coefficient
measurements for hydrocarbon binary mixtures: viscosity and composition dependency.
The Journal of Physical Chemistry. B, 112(20), 64427. doi:10.1021/jp711090q
Madhusudana, C. V. (2014). Thermal Contact Conductance, Second Edition (Second.).
Sydney, NSW, Australia: Springer.

79

Manundawee, S., Assabumrungrat, S., & Wiyaratn, W. (2011). Two-dimensional


mathematical modeling of oxidative coupling of methane in a membrane reactor. TiCHE
International Conference, 15.
Reed, N. (2008). Modeling Flow through a Fixed Bed Packed Reactor. Pennsylvania State
University. Retrieved from
http://www.ems.psu.edu/~elsworth/courses/EGEE520/2008Deliverables/reports/EGEE5
20FinalpaperReed.pdf
Saw, E. T., Oemar, U., Tan, X. R., Du, Y., Borgna, a., Hidajat, K., & Kawi, S. (2014).
Bimetallic NiCu catalyst supported on CeO2 for high-temperature watergas shift
reaction: Methane suppression via enhanced CO adsorption. Journal of Catalysis, 314,
3246. doi:10.1016/j.jcat.2014.03.015
Thirumaleshwar, M. (2006). Fundamentals of Heat and Mass Transfer (Second.). Patparganj,
Delhi, India.
Welty, J. R., Wicks, C. E., Wilson, R. E., & Rorrer, G. L. (2008). Fundamentals of
Momentum, Heat, and Mass Transfer (Fifth.). Corvallis, Oregon, USA.

80

Appendix I Derivation of Rate Constant for Rate Law


According to (Saw et al., 2014), a kinetic study performed for the 5Ni/5Cu catalyst produced
a rate of 0.33*10-6 mol/m2s.
Considering that the surface area per unit volume is 32000m2/m3, for a particle diameter of
187.5 microns, the rate equates to 0.01053mol/m3s.
The study performed assumed atmospheric pressure (1atm), at mole fractions of 0.07, 0.22,
0.1, and 0.2 for CO, H2O, CO2, and H2 respectively. This means that the partial pressures
were 0.07, 0.22, 0.1 and 0.2atm respectively. In addition, the temperature at which this study
was performed was at 623K.
Substituting these values into the rate law, along with an activation energy of 41300J/mol, as
per the article, gives rise to the value of A as 5493.

Appendix II - Step-by-Step Modelling Guide in COMSOL 4.4 Update 1


Starting out
1. Start COMSOL Multiphysics
2. Select File >> New >> Model Wizard >> 2D Axi-Symmetric
3. Select and add Transport of Concentrated Species, Darcys Law, and Heat
Transfer in Fluids. Note that the pseudo-homogeneous assumption has eliminated
the need for the Heat Transfer in Porous Media Interface.
4. Click on Study >> Stationary >> Done
Variables and geometry
1. Right-click on the Geometry tab and select Rectangle.
2. Click on the Rectangle tab, and enter the following:
Parameters
Type
Size: Width
Size: Height
Base
Rotation

Quantities
Solid
irad
0.05
Corner, r:0, z:0
0

3. Repeat this step to create Rectangle 2.


4. Click on Rectangle 2, and enter the following:
Parameters
Type
Size: Width
Size: Height
Base
Rotation

Quantities
Solid
orad-irad
0.05
Corner, r:irad, z:0
0

Note that orad and irad are parameters which will be entered in the next step, therefore ignore
the fact that they are highlighted in orange.

81

5. We will now enter the variables and parameters that define this model. Right click the
Global Definitions tab under the top-most node, and select Parameters.
6. Enter the following data values.
Parameter

Quantity

Description

T0

773[K]

Inlet temperature

Ea

41.3[kJ/mol]

Activation energy of WGS

8.31

Molar Gas Constant

DeltaH1new

-41[kJ/mol]

Enthalpy of WGS

Per

1.50E-05

Membrane Permeability (Unitless, Iyoha)

dp

187.5e-6[m]

Particle Diameter

M_H2

2e-3[kg/mol]

MW H2 (Kg/mol)

M_H2O

18e-3[kg/mol]

MW H2O (Kg/mol)

M_CO

28e-3[kg/mol]

MW CO (Kg/mol)

M_CO2

44e-3[kg/mol]

MW CO2 (Kg/mol)

Cp_H2

14.25[kJ/(kg*K)]

SHC of H2

Cp_H2O

2[kJ/(kg*K)]

SHC of H2O

Cp_CO

1[kJ/(kg*K)]

SHC of CO

Cp_CO2

1.075[kJ/(kg*K)]

SHC of CO2

eta_H2

2.31e-5[kg/(m*s)]

Viscosity of H2

eta_H2O

4.44e-5[kg/(m*s)]

Viscosity of H2O

eta_CO

4.44e-5[kg/(m*s)]

Viscosity of CO

eta_CO2

4.47e-5[kg/(m*s)]

Viscosity of CO2

por

0.4

Bed Porosity

orad

0.002[m]

Outer Radius

irad

0.00075[m]

Inner Radius

flowrate

100[ml/min]

Feed Rate

htube

439.78[W/(m^2*K)]

Heat Transfer Coefficient (Membrane wall to tube)

ho

76770.9[W/(m^2*K)]

Annulus outer heat transfer coefficient

hi

309684[W/(m^2*K)]

Annulus inner heat transfer coefficient

k_eff

35.46[W/(m*K)]

Pseudo-Homogeneous Effective Thermal Conductivity

7. Under Component 1, right click Definitions, and select Variables. Enter the
following data.
Variable

Expression

Description

yA

cA/c_tot

Mole Frac CO

yB

cB/c_tot

Mole Frac H2O

yC

cC/c_tot

Mole Frac H2

yD

cD/c_tot

Mole Frac CO2

(cC*cD)/(cA*cB*keq1)
Per*(((8.31^0.5)*(T[K])^0.5[1/K]*(cC[m^3/mol])^0.5)((8.31^0.5)*(T2[K])^0.5[1/K]*(cH2[m^3/mol])^0.5))

Approach to Equilibrium

JH2

Sieverts Law
82

keq1

exp(4577.8[K]/T-4.33)

Eqm Constant

rho_mix

M_CO*cA+M_H2O*cB+M_H2*cC+M_CO2*cD

Mixture Density

cp_mix

yA*Cp_CO+yB*Cp_H2O+yC*Cp_H2+yD*Cp_CO2

Effective Heat Capcity

eta_mix

yA*eta_CO+yB*eta_H2+yC*eta_H2+yD*eta_CO2

Effective Viscosity

c_tot

p[(m^3*K)/J][mol/m^3]/(R*T)

Total Concentration

cA

chcs.x_wCO*p[(m^3*K/J)][mol/(m^3)]/(R*T)

Concentration CO

cB

chcs.x_wH2O*p[(m^3*K/J)][mol/(m^3)]/(R*T)

Concentration H2O

cC

chcs.x_wH2*p[(m^3*K/J)][mol/(m^3)]/(R*T)

Concentration H2

cD

chcs.x_wCO2*p[(m^3*K/J)][mol/(m^3)]/(R*T)

Concentration CO2

perm

por^3/(5*(1-por)^2*(6/(dp))^2)

dl.w

dl.u

Darcy Permeation Constant


Velocity z component
(Darcy Law)
Velocity r component
(Darcy Law)

u_feed

Feed Velocity

rcalc

flowrate/(pi*(orad^2-irad^2))
0.9*32000*10^-6*5493*exp(41300[K]/(R*T))*(yA*p[1/Pa]*9.87*10^6)^0.95*(yB*p[1/Pa]*9.87*10^6)^0.16*(yC*p[1/Pa]*9.87*10^-6)^0.19*(yD*p[1/Pa]*9.87*10^-6)^-0.34*(1B)[mol/(m^3*s)]

M_mix

((0.833*18)+(0.1666*28))

Molecular Weight at inlet

xCOin

0.1666*M_mix/28

Inlet CO mole fraction

xH2Oin

0.8333*M_mix/18

Inlet H2O mole fraction

Rate Law

Shell Side, Darcys Law


Click on Darcys Law node, and select the shell domain (Domain 2 by default)
Under Gravity effects, select Off
Under the Dependent Variables node, enter p for Pressure.
Click on the Fluid and Matrix Properties 1.
Under Fluid Properties, set Density and Dynamic viscosity to user-defined, and enter
rho_mix and eta_mix respectively.
6. Under Matrix Properties, select Domain Material, and enter por in the porosity field.
7. Under the Permeability model, select Permeability, and enter perm in the field.
8. Click on the No Flow node, and ensure that the sides (Borders 4 and 7), are
highlighted.
9. Click on the Initial Values node, select the Pressure radio button, and enter 1[atm]
in the field.
10. Right click on Darcys Law node, and select Inlet. Repeat, and select Pressure. These
two entities constitute the inlet and outlet boundary conditions.
11. Under Inlet, enter u_feed. Click on Boundary 5 (or the shell side inlet region)
12. Under Pressure, enter 3[bar]. Click on Boundary 6.
1.
2.
3.
4.
5.

83

Shell Side, Heat Transfer in Fluids


1. Click on Heat Transfer in Fluids node. Ensure that the dependent variable is T, and
that the Shell side is checked.
2. Click on Heat Transfer in Fluids 1, under Model Inputs >> Absolute Pressure, select
Pressure (dl/dlm1). This couples Darcys Law pressure to the heat transfer.
3. Under Velocity field, select User-Defined, and enter u/por and v/por for r and zdirections respectively.
4. Under Heat Conduction, Fluid, select User Defined, and enter k_eff in the field. This
reflects the pseudo-homogeneous thermal conductivity.
5. Under Thermodynamics, Fluid, enter rho_mix and cp_mix in the Density and Heat
capacities sections.
6. Click on the Initial Values node, enter T0.
7. Right click on the Heat Transfer in Fluids main node, and add in Outflow,
Temperature, Heat Source, and 2 Heat Flux instances.
8. Under Outflow, select boundary 6. This reflects convective dominance.
9. Under Temperature, select boundary 5, and enter T0 in the field.
10. Under Heat Source, select the General Source radio button >> User Defined, and
enter (-DeltaH1new)*rcalc in the field. Highlight Domain 2 to ensure that heat of
reaction applies in shell side.
11. Under Heat Flux 1, select Boundary 7 (outer shell side), check the General inward
heat flux radio button, and enter -ho*(T-T0).
12. Under Heat Flux 2, select Boundary 4 (inner shell side), check the General inward
heat flux ratio button, and enter -hi*(T-T2).
Shell Side, Transport of Concentrated Species
1. Click on Transport of Concentrated Species node, select the shell domain by clicking
on the shell geometry.
2. Under Transport Mechanisms >> Diffusion Model, select Maxwell-Stefan, check
Convection box.
3. Under Dependent Variables, enter 5 species, and enter these mass fractions, wCO,
wH2O, wH2, wCO2, and wHe.
4. Under Species >> From Mass Constraint, choose He.
Note: COMSOL requires a 5th species to solve for in the shell side. Therefore, a dummy
species Helium is added, even though it is not relevant in the shell side.
5. Click on Convection and Diffusion sub-node.
6. Velocity Field >> Enter u/por for r-direction, and w/por for z-direction. U and w are
actually Darcys Law components, which will be declared later.
7. Under Temperature, choose (ht). This couples the heat transfer module to the
Maxwell-Stefan module.
8. Under Mixture Density, select Ideal Gas.
9. Under the Molar Mass sections, enter M_CO, M_H2O, M_H2, M_CO2, and 100 for
the respective chemical species. 100 is for He.
10. Under the Maxwell-Stefan diffusivity matrix, enter 2e-5 for all values. This is derived
from Iyoha (2007).
11. Under the Thermal Diffusion component, ensure that all values are 0.
12. Under Initial Values, enter 0.1666, 0.833, 1e-6, 1e-6, for CO, H2O, H2, and CO2
respectively. This gives the 5:1 steam-carbon ratio. Vary if desired.
84

13. Right click on Transport of Concentrated Species, add in the following: Reactions, No
Flux, Mass Fraction, Outflow, and Flux.
14. Click on Reactions, and fill in rcalc*M_CO, -rcalc*M_H2O, rcalc*M_H2,
rcalc*M_CO2 for the respective chemical species. Highlight Domain 2 to apply the
Reaction to the shell region.
15. Click on No Flux 2, select Boundaries 4 and 7.
16. Click on Mass Fraction 1 node, select Boundary 5, and ensure 0.166, 0.833, 1e-6, and
1e-6 respectively.
17. Click on Outflow node, select Boundary 6.
18. Click on Flux 1 node, Boundary 4, check species wH2, and under the Inward Flux
box, fill in -JH2[mol/(m^2*s)]*M_H2. This represents Sieverts Law boundary
condition.
Tube Side, Laminar Flow
1. The focus will now shift to tube side modelling. We will start off by entering the
Navier-Stokes equations. On the top ribbon, click on Physics tab, and select Add
Physics option.
2. Enter Laminar Flow, Transport of Dilute Species, and Heat Transfer in Fluids.
3. Click on Laminar Flow top node, select the tube domain (Domain 1) and under
Compressibility, select Incompressible Flow.
4. Select None for Turbulence Model type.
5. Click on the Fluid Properties sub-node, choose User Defined for both density and
viscosity, and enter 0.524 and 1.81e-5 respectively in the fields.
6. Click on Wall sub-node, select No-Slip, and select boundary 4. This applies No-slip to
the tube wall.
7. Click on Initial Values sub-node, enter 0 for both velocity fields, and 1[atm] for the
pressure field.
8. Right click on the parent Laminar Flow node, and append Inlet and Outlet.
9. Under Inlet 1, choose Boundary 2, select Normal inflow velocity node, and fill in 0.5
in the field.
10. Under Outlet 1, choose Boundary 3, select Pressure boundary condition, enter 1[atm]
in Pressure.
Tube Side, Transport of Dilute Species
1. Click on the Transport Dilute Species node, highlight Domain 1 to apply it.
2. Under Transport Mechanisms, check Convection.
3. Under Dependent Variables, Number of Species, enter 1, and under the
concentrations field, enter cH2.
4. Click on Convection and Diffusion 1 sub-node, and enter u2 and w2 in the velocity
field. (u2 and w2 are set as default by the Laminar Flow node. If this happens to
change, check the Dependent Variables in the Laminar Flow node, and identify what
the r and z component velocities are called)
5. Under Diffusion >> Bulk Material, choose None. Under Diffusion Coefficient, select
User Defined, and enter 7.028e-4.
6. Click on Initial Values sub-node, enter 1e-6.
7. Right click on the parent Transport of Dilute Species node, and insert these 3;
Concentration, Outflow, and Flux.
8. Under Concentration, check Species cH2 box, enter 1e-6 in the field, and highlight
Boundary 2.
85

9. Under Outflow, highlight Boundary 3.


10. Under Flux, check Species cH2 box, and enter this into the edit field.
JH2[mol/(m^2*s)]
Tube Side, Heat Transfer in Fluids
1. Click on the newly-created Heat Transfer in Fluids node, and highlight Domain 1.
2. Under Dependent Variables, check that the declared variable is T2.
3. Click on Heat Transfer in Fluids 1 node. Under Absolute Pressure, choose Pressure
(spf2/fp1) from the drop-down. This couples the Navier-Stokes flow field to the heat
transfer module.
4. Under Velocity Field, enter u2 and w2. Alternatively, one may choose (spf2/fp1); no
difference is expected.
5. Under Heat Conduction in Fluid, choose User Defined and enter 0.1513, choose
Isotropic.
6. Under Thermodynamics, Fluid, select Gas/Liquid from the drop-down, and enter
0.524 for Density, 5190 for Cp.
7. Click on Initial Values node, and enter T0 in the edit field.
8. Add in Outflow, Temperature and Heat Flux sub-nodes.
9. Under Outflow sub-node, highlight Boundary 3.
10. Under Temperature sub-node, highlight Boundary 2, and enter T0 in the edit field.
11. Under Heat Flux, highlight Boundary 4, select General inward heat flux radio
button, and enter htube*(T-T2)
Meshing
1. Right-click on Meshing, select Free Triangular.
2. Right-click Free Triangular, select Distribution twice. This creates 2 distribution subnodes.
3. Under Distribution 1, highlight Boundary 4, enter 300 into the edit field.
4. Under Distribution 2, highlight Boundaries 2 and 5, and enter 50 into the edit field.
Solver
1. Under Study 1, Solver Configurations, expand until Stationary Solver 1 is revealed.
2. Right-click, and choose Direct.
3. Under the newly-created Direct node, choose PARDISO for the solver and Nested
Dissection Multithread for pre-ordering algorithm.
4. Click on Compute in the ribbon under the Home tab.
Solving Tips:
1. COMSOL has a retrospective method of seeking convergence, that is, it automatically
takes a previous solution and uses it as a form of reference for future solutions.
Therefore, if convergence cannot be achieved, with errors such as Undefined value
found in the Equation Residual Vector, or Undefined value found in the stiffness
matrix, attempt to relax the conditions by setting higher Initial Values or Mass
Fractions (e.g. from 1e-6, change to 0.01), and re-compute to converge first. Then
slowly reduce the values. This is because solvers generally do not like working with
very small quantities or absolute zeroes, and tend to work up.

86

Study
1. To perform a Surface Plot, right-click Results, select 2D Plot Group to create a 2D
Plot Group node.
2. Click on this Plot Group node, choose Surface, and enter the parameter/variable of
choice. Axis preferences/legends/titles are available for the user to modify within
these 2 nodes.
3. To perform a Line Plot (such as the center-line plots seen in the Simulation section of
this report), a Cut Line has to be made. Right click Data Sets under Results, and select
Cut Line 2D.
4. Under Data Set, choose Solution 1 (note, if it is a new Study, such as a Parametric
study, the results will be under a different Solution).
5. Under Line Data section, select 2 points, and enter the r and z-coordinates. A line will
be drawn between these 2 points, meaning that COMSOL will read the data along this
line. So if the Shell side, center-cut line is desirable, enter these values into the edit
fields.
Point 1
Point 2

r
(0.002+0.00075)/2
(0.002+0.00075)/2

z
0
0.05

6. Create a 1D plot group. Right click on this node and choose Line Graph. Under Data
Set, choose Cut Line 2D, to call for the solutions from the cut line you have just
created.
7. To observe the CO conversion profiles, in the y-axis Data, key in ((xCOin*u_feed)(chcs.x_wCO*w))/(xCOin*u_feed).
Parametric Study
1. To operate the simulation for different inlet values and pressures, right-click Study 1,
and choose Parametric Sweep.
2. Click on the Parametric Sweep node, choose Specified Combinations from the dropdown node, and enter the variable to be altered, and its set of values. E.g. T0 (inlet
temperature) needs to take on these values 673/723/773/823/873. Separate them by a
space.
3. Re-Compute the simulation.
4. For more complex simulations such as Steam-CO ratio sweeps, add in 1 entry for
steam mass fraction, and list down its parameters. Repeat for CO. Choose Specified
Combinations once again, such that COMSOL will run the parameters for both in a
respective fashion.
Note: COMSOL is unable to perform a sweep for values which are not directly entered into
the simulation. For instance, if the inlet velocity is u_feed, which is a variable defined by the
flow rate divided by the unit area, running a parametric sweep for the flow rate will fail. The
user has to calculate the different values of u_feed by himself, run a sweep for u_feed.

- END OF FINAL REPORT 87

You might also like