You are on page 1of 15

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/271140988

Model-based Experimental Screening for DOC


Parameter Estimation
ARTICLE in COMPUTERS & CHEMICAL ENGINEERING JANUARY 2015
Impact Factor: 2.78 DOI: 10.1016/j.compchemeng.2015.01.004

READS

91

5 AUTHORS, INCLUDING:
Bjrn Lundberg

Jonas Sjblom

Chalmers University of Technology

Chalmers University of Technology

3 PUBLICATIONS 5 CITATIONS

19 PUBLICATIONS 319 CITATIONS

SEE PROFILE

SEE PROFILE

Derek Creaser
Chalmers University of Technology
60 PUBLICATIONS 1,071 CITATIONS
SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

Available from: Jonas Sjblom


Retrieved on: 18 January 2016

Computers and Chemical Engineering 74 (2015) 144157

Contents lists available at ScienceDirect

Computers and Chemical Engineering


journal homepage: www.elsevier.com/locate/compchemeng

Model-based experimental screening for DOC parameter estimation


Bjrn Lundberg a, , Jonas Sjblom b , sa Johansson c , Bjrn Westerberg d , Derek Creaser a
a

Department of Chemical and Biological Engineering, Chemical Reaction Engineering, Chalmers University of Technology, Gteborg SE-412 96, Sweden
Department of Applied Mechanics, Division of Combustion, Chalmers University of Technology, Gteborg SE-412 96, Sweden
Johnson Matthey AB, Vstra Frlunda SE-421 31, Sweden
d
Scania AB, Sdertlje SE-151 87, Sweden
b
c

a r t i c l e

a b s t r a c t

i n f o

Article history:
Received 4 August 2014
Received in revised form 4 December 2014
Accepted 7 January 2015
Available online 19 January 2015
Keywords:
Parameter estimation
D-optimal design
Diesel Oxidation Catalyst
Multivariate Data Analysis
Engine rig experiments

In the current study a parameter estimation method based on data screening by sensitivity analysis
is presented. The method applied Multivariate Data Analysis (MVDA) on a large transient data set to
select different subsets on which parameters estimation was performed. The subset was continuously
updated as the parameter values developed using Principal Component Analysis (PCA) and D-optimal
onion design. The measurement data was taken from a Diesel Oxidation Catalyst (DOC) connected to a
full scale engine rig and both kinetic and mass transport parameters were estimated. The methodology
was compared to a conventional parameter estimation method and it was concluded that the proposed
method achieved a 32% lower residual sum of squares but also that it displayed less tendencies to converge
to a local minima. The computational time was however signicantly longer for the evaluated method.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
Common kinetic mechanisms involve a large number of parameters. These mechanisms and the complex interaction between
different reactions often result in a high correlation between
the kinetic parameters (Bernaerts et al., 2000; Franceschini and
Macchietto, 2008b). This severely complicates the parameter
estimation process and increases the importance of Design of
Experiments (DoE) and the parameter estimation algorithm. DoE
in reaction kinetics has a long tradition reaching back to the pioneering work by Box and Lucas (1959), Box and Hunter (1965),
Draper and Hunter (1966, 1967a,b), Box (1968), Hill et al. (1968),
and Hunter et al. (1969). The focus has however been generally
more directed toward steady-state systems rather than transient
ones (Franceschini and Macchietto, 2008a) even though the need
and benets of transient experiments have been demonstrated
(Goodwin, 1977; Berger et al., 2008). The system studied in the
current paper is a reactor model of a full scale Diesel Oxidation
Catalyst (DOC) connected to a Heavy Duty Diesel engine rig. The
system is transient and the model parameters highly correlated;
DoE is therefore of highest importance.
A wide range of catalyst models with varying level of detail can
be found in literature. Generally, the kinetics are either described

Corresponding author. Tel.: +46 317722942.


E-mail address: bjorn.lundberg@chalmers.se (B. Lundberg).
http://dx.doi.org/10.1016/j.compchemeng.2015.01.004
0098-1354/ 2015 Elsevier Ltd. All rights reserved.

by a microkinetic model (Dumesic et al., 1993; Olsson et al., 2002;


Crocoll et al., 2005; Salomons et al., 2006), where every reaction is
described by elementary steps, or by more empirical global kinetic
models (Voltz et al., 1973; Ansell et al., 1996; Lafossas et al., 2011;
Watling et al., 2012). However, the catalyst model is not only
dened by the kinetics but the inuence of heat and mass transport must also be taken into account. The general trend for monolith
converters is still to describe heat and mass transfer with simple
one-dimensional models (Gthenke et al., 2007), especially for the
purpose of parameter estimation, but different levels of complexity
epnek et al., 2012)
up to three-dimensional CFD-based models (St
exist. The preference for simple, and thereby fast, catalyst models in parameter estimation tends to make the kinetic parameters
case specic, resulting in kinetic models only giving reliable predictions for a catalyst similar to the one used in the experiments from
which the model was derived (Wang et al., 2008). This means that
efcient methods to tune model parameters from existing models
to measurement data from new catalysts are highly desirable.
To tune the catalyst model to measurement data a method of
parameter optimization is needed. The methods can broadly be
divided into either global or local optimization algorithms where
the latter historically have been far more favored in automotive
catalysis parameter estimation. The local optimization methods
are generally based on an evaluation of the parameter sensitivity to identify the direction of steepest reduction of the residual.
The advantage of these algorithms is that they are fast and accurate but the drawback is that they have a tendency to converge to a

B. Lundberg et al. / Computers and Chemical Engineering 74 (2015) 144157

local minimum close to the initial guess and thereby failing to converge to the global minima (Blockley, 2010). The global methods
generally evaluate a larger parameter space which means that they
are more likely to nd the global optima. The main drawback with
the global methods is very high computational cost which means
that they often are not realistic to use for complex systems with
many parameters. To be successful the global algorithms also are
more dependent on robust models since a wide parameter span will
be needed to be evaluated. Examples from literature where global
optimization algorithms have been applied in catalysis modeling
include Glielmo and Santini (1999) who used the Genetic Algorithm
(an evolutionary algorithm) and Ramanathan and Sharma (2011)
who showed that an improved t could be achieved if a global algorithm was followed by a local algorithm to increase the accuracy.
A method of reducing the tendencies of the local methods to converge to a local minimum close to the initial guess may therefore
be the more desirable way of improving parameter optimization
for the current application.
In our previous study (Lundberg et al., 2014), an experimental
plan that included several different DOC catalyst congurations,
was applied to a full scale engine rig system and used for parameter tuning. The best model t was achieved with a model that
included internal transport resistance with the effective diffusivities of the species being tuned in addition to the kinetic parameters.
In the current study, the data and catalyst model formulation will be
re-used but with a new approach to parameter tuning will be presented and evaluated. The method is based on a study by Sjoblom
and Creaser (2008) who developed methods for model-based DoE
including transient data. The methodology presented by Sjblom
and Creaser includes Multivariate Data Analysis (MVDA) to reduce
the dimensionality of the largely correlated Jacobian and subsequent selection of a D-optimal sub-set of parameters to estimate
based on the parameter sensitivity of simulation data. The relatively
simple global kinetic model in the current study makes the number
of parameters to estimate less of a concern. Instead the method was
revised to select a set of time points from the large transient data set
with high sensitivity for parameter changes and at the same time
retaining a good representation of the entire data set. To achieve
these properties the D-optimal onion design, introduced by Olsson
et al. (2004), was selected as an algorithm for data point selection. The algorithms combine the D-optimal design, which by itself
tends to only select the most extreme points (Pinto et al., 1990;
Zullo, 1991), with a space lling approach making the selection
more evenly distributed over the entire data set. Parameter estimation was performed based only in the selected time points with
the aim of making use of improved statistical properties to improve
the overall t to measurement data. Since the parameter sensitivity, and thereby the data selection, will depend on the parameter
values, the parameter estimation was performed in an iterative process where the data point selection was updated as new parameter
values were found. This change of data points used for parameter
tuning may also help to avoid local optima. As an evaluation, the
efciency and resulting t to measurement data achieved by the
presented method will be compared to the results of a traditional
method of parameter estimation.

2. Experimental
The most important characteristics of the reactor model and
adjustable parameters are summarized in the current section and
for completeness the dening equations are given in Appendix. A
thorough description of the experimental set-up, measurements
and engine operation design is given in the preceding study
(Lundberg et al., 2014).

145

2.1. Reactor model


A full scale catalyst, connected to an engine with varying inlet
properties displays a highly dynamic behavior. This means that the
catalyst outlet conditions will not only be inuenced by the current
inlet conditions but also those at previous time points. To describe
this behavior a transient catalytic reactor model is needed. In this
work a uniform radial ow and concentration distribution over the
catalyst cross section was assumed which makes it sufcient to
model only one channel. The single channel model, closely based
on the model presented by Ericson et al. (2008), was discretized as
tanks in series where the catalyst washcoat was discretized both
radially and axially while the gas phase was only discretized axially. This 1D/2D (gas phase/washcoat) structure was chosen since
it was considered a good compromise between accuracy and computational speed (Mladenov et al., 2010). A lm theory model was
used to model the external heat and mass transport between gas
bulk and washcoat surface. Axial diffusion and radial temperature
gradients in the washcoat were neglected.
The global kinetics of the DOC has been described in literature
with several different levels of detail. The simplest kinetic models
only describe the oxidation of CO, HC and NO (Voltz et al., 1973; Oh
and Cavendish, 1982; Wang et al., 2008) (where HC is represented
as one molecular species) but more detailed models with a more
complex description of HC (Stamatelos et al., 1999; Lafossas et al.,
2011) and other additional reactions (Ansell et al., 1996; Salomons
et al., 2006; Pandya et al., 2009) have also been developed. The
kinetic model used in this study is of LangmuirHinshelwood type
and was originally suggested in the classical work by Voltz et al.
(1973) and later modied by Oh and Cavendish (1982). The model,
which has been widely and frequently used in DOC modeling over
the years, only includes three reactions of which one is equilibriumlimited:
CO + 0.5O2 CO2
C3 H6 + 4.5O2 3CO2 + 3H2 O
NO + 0.5O2  2NO2
The reaction rates were calculated according to Eqs. (1)(5).
r1 =
r2 =
r3 =
K =

k1 yCO yO2

(1)

G(yi , Ts )
k2 yC3 H6 yO2
G(yi , Ts )
k3 yNO yO2
G(yi , Ts )

(2)


1

K
Kp


(3)

yNO2

(4)

1/2

yNO yO

2
G(yi , Ts ) = Ts (1 + K4 yCO + K5 yC3 H6 )2 (1 + K6 yCO
yC2

3 H6

0.7
)(1 + K7 yNO
)

(5)
where Kj is the reaction rate coefcient for the inhibition terms
in the denominator G and Kp is the equilibrium constant for NO
oxidation. At thermodynamic equilibrium, Kp will be equal to K and
reaction rate r3 will be equal to zero. Both the rate coefcients kj and
the inhibition terms Kj were described by Arrhenius expressions:
kj = Aj eEA,j /RTs

(6)

The start values for estimation of kinetic parameters were taken


from the start values in Wang et al. (2008) where results from
several studies (Chen et al., 1988; Koltsakis et al., 1997; Dubien

146

B. Lundberg et al. / Computers and Chemical Engineering 74 (2015) 144157

et al., 1998; Koltsakis and Stamatelos, 1999; Kandylas and Koltsakis,


2002; Tsinoglou and Koltsakis, 2003; Guojiang and Song, 2005)
were compiled. The initial values for kinetic parameter estimation
used in this study are shown in Section 3.
The kinetic parameters in Eqs. (1)(5) are highly correlated and
since the initial parameter values were taken from different studies, the t of the model to experimental data was expected to be
poor before any parameter tuning was performed. However, the
parameters were successfully used as a starting point for parameter
tuning of a DOC model against engine rig data in Wang et al. (2008)
which was also the intended application in the present study.
2.2. Adjustable parameters
To tune the model to the measurement data a number of parameters in scaled and centered form (Sjoblom and Creaser, 2008) were
adjusted. These parameters can be divided into kinetic parameters,
mass transfer parameters and heat transfer parameters. In the current section only a brief introduction will be given to the adjustable
parameters, a more detailed description can, however, be found in
Section A.2 in Appendix.
Three types of kinetic parameters are adjusted; pre-exponential
factors, activation energies and activity scaling factors. The activity
scaling factor (actq in Eq. (A2)) is simply a scale factor for all reaction rates on a certain site on a certain catalyst (parameters were
tuned to several different catalysts simultaneously) with the purpose of accounting for different properties of catalyst samples, such
as active noble-metal dispersion.
Previously (Lundberg et al., 2014), the effective diffusivities
were evaluated as a method of tuning the transport resistance
with the aim of reducing the correlation between mass transport
and kinetic parameters. The species were divided into two groups,
where the rst group contained O2 , NO, NO2 , and CO and the second
group contained HC. In the rst group all species are well dened
with similar diffusivities (Massman, 1998) and could be expected
to have similar mass transport properties in the washcoat with presumably the same bias from their true values. To reduce the number
of parameters to tune the same scale factor, fDscl ,S in Eq. (A21), was
used for all species in this group. The second group contained HC
which was represented as C3 H6 but in reality it is a wide range of
hydrocarbons with different mass transport properties. The scale
factor used for HC was denoted fDscl ,L .
The heat transfer parameters including thermal mass (sheat ),
environmental temperature (T ), and a lumped heat transfer coefcient (tot ) were earlier tuned to a satisfactory t in (Lundberg
et al., 2014) for the system at hand and will therefore not be retuned
in the current study.
2.3. Experimental set-up
A 13 L heavy duty diesel engine was used as the exhaust source
and Swedish MK1 diesel, a commercial low-sulfur (approximately
5 ppm S) diesel, was used as fuel. The engine was equipped with
a dynamometer control system enabling independent control of
load and speed. A change in operating point for the engine could
be quickly implemented, although the dynamics of the engine rig
(as well as the monolith) itself were much slower due to the large
thermal mass.
To further widen the experimental range, four different catalyst congurations with different noble metal loading, lengths, and
washcoat thicknesses were used. All catalysts, shown in Table 1,
were model catalysts provided by Johnson Matthey.
All congurations have a 30.5 cm diameter (12 inch),
62 channels/cm2 (400 cpsi) cell density and a wall thickness
of 0.152 mm. The substrate material used was cordierite. The

Fig. 1. Temperature engine map and selected operating points.

catalysts were thermally conditioned for one hour at 600 C prior


to use.
2.4. Engine operation design
As mentioned in Section 1 the available exhaust composition,
ow and temperature is limited by the operating points of the
engine. It also takes several minutes for the catalyst inlet conditions to stabilize when switching between operating points, which
means that experimental time will be important when deciding the
number of different operating points when the full transient behavior is of interest. Since the number of catalyst congurations was
large (Table 1) and some replicates also were necessary only 8 different operating points were selected. The operating points were
selected to make as large steps as possible in the different variables including concentrations of NO, NO2 , HC, CO and O2 as well as
temperature and ow rate with the purpose of making the experimental space as large (and orthogonal) as possible. Fig. 1 shows
how the eight operating points span over the engine map for the
case of temperature.
Every engine operating point was run for 15 min before switching to another, thereby generating both transient data as well as
data for operation close to steady-state. Further details about the
engine operation design were reported earlier (Lundberg et al.,
2014).
3. Parameter estimation method
With the engine operating points selected above, both transient
and stationary data are available over a wide range of temperatures, ow rates and concentrations which will aid the parameter
tuning. There is however a risk that some parts of the data set will
dominate the parameter tuning simply because there still is a large
number of data points at similar conditions. This similarity in the
data may result in the parameter tuning being less efcient since
some parameters may have a very low sensitivity for a large part
of the data points. Data reduction by Multivariate Data Analysis
(MVDA) is a means by which to make the inuence from data points
at different conditions more equal. In the current study a method
of data reduction is evaluated that can be divided into three steps.
In the rst step the sensitivity for all parameters, time points and
residuals is calculated by nite differences. The resulting matrix is
commonly denoted the Jacobian matrix. In the second step a principal component analysis (PCA) is performed on the Jacobian matrix
(J) resulting in a linear model consisting of a loadings matrix (P) and
a scores matrix (T) according to Eq. (7).
J T P

(7)

B. Lundberg et al. / Computers and Chemical Engineering 74 (2015) 144157

147

Table 1
Catalyst conguration used for parameter estimation (ad). Catalyst conguration c consists of two catalysts in series. The Pt-loading is shown in both mass% of washcoat
and g/ft3 total catalyst volume.
Conguration

Pt-loading [mass% washcoat]

Pt-loading [g/ft3 monolith]

a
b
c
d

0.30
0.59
0.30
0.59

15
30
15
15

The scores matrix gives a simplied description of the parameter sensitivity of all data points given in the Jacobian. The scores
matrix has fewer columns than the Jacobian but the same number of rows, see Appendix for details. In the third step a D-optimal
onion design is performed on the resulting scores matrix giving a
selection of data points that will have a more equal inuence from
the different parameters. The reduced number of columns in the
scores matrix compared to the Jacobian will reduce the computational time of nding a D-optimal design which would be very high
if the Jacobian would be used directly. The method is described in
detail in Appendix and is schematically summarized in Fig. 2.

Length [cm]

Washcoat thickness [mm]

10.2
10.2
2 10.2
10.2

0.110
0.110
0.110
0.055

account for differences in the average concentration the residuals


were also weighted against the inverse of the average outlet mole
fractions for the respective component. In the current study the
data points used for parameter estimation will change and a residual calculation that is more adaptive to a change data is needed.
To achieve a more balanced inuence from the different components, the residual was calculated as the difference in conversion
instead of the difference in mole fraction, as demonstrated in Eq.
(8).

yi,in 
yi,in yi,out
yi,out

yi,in
yi,in

yi,out yi,out

3.1. Residual calculation

res =

The denition of the residuals and their weighting are of utmost


importance to obtain a balanced description of the parameter
sensitivity at every time point of the data set and thereby also
for the effectiveness of the parameter search algorithm itself. In
the previous study and parameter estimation on the presented
data (Lundberg et al., 2014), the difference in outlet mole fraction
between measured and simulated data was used as the residual. To

yi,out is the
where wi is a weighting factor (further described below), 
simulated outlet mole fraction, yi,out is the measured outlet mole
fraction and yi,in is the measured inlet mole fraction. To compensate
for the fact that the NO conversion is limited by thermodynamic
constraints the residual for NO was instead calculated relative to the

w =

yi,in

wi (8)

Original parameters

Parameter sensivity
calculaons
(at all me points)

Jacobian matrix

PCA

Residual sum of squares


calculaon
(at all me points)

Scores matrix

D-opmal onion design


Selected me points

Parameter tuning
(at selected me points)

No

Improved parameter
values?

Yes

Fig. 2. Summary of the method used for parameter estimation. Improved parameter values means that parameters were found that reduced the residual sum of squares at
the selected time points which will result in a right hand loop.

148

B. Lundberg et al. / Computers and Chemical Engineering 74 (2015) 144157

equilibrium conversion possible at the current condition as shown


in Eq. (9).

res =

(yNO,in yNO,out,lim ) (
(yNO,in yNO,out,lim ) (yNO,out yNO,out,lim )
yNO,out yNO,out,lim )

yNO,in yNO,out,lim
yNO,in yNO,out,lim

where yout,lim is the lowest molar fraction of NO possible via NO


oxidation at the inlet conditions due to thermodynamic constraints.
Since the model did not account for NOx reduction, and since the
NOx reduction was below 3% for more than 99% of the data set, the
t for NOx is considered to be described well enough by only the NO
residual. In addition to NO, the residuals of both HC and CO were
used for parameter estimation.
It is not desirable to t parameters to data points with close to
100% conversion, since the kinetic parameter sensitivity for these
points will be very low. Therefore it was necessary to set the residuals to zero for certain components at such points in the data set. For
the current data set the following limits for outlet mole fractions,
based on measurement accuracy, were selected to generate a zero
residual for the three different components:
yHC,out < 2 ppm
yCO,out < 5 ppm
yNO,out yNO,out,lim < 5 ppm
When starting simulations with the catalyst model the only
known conditions were those of the inlet and outlet ow. This
means that starting properties such as concentrations and temperature inside the catalyst needed to be estimated which resulted
in the rst seconds of every simulation being unreliable. To avoid
parameters being tuned against unreliable simulation results the
residuals from the rst 120 s of every experimental data set was
set to zero.
The weighting factor wi is calculated according to Eq. (10):
1
wi =
ni

(10)

where ni is the number of data points used for residual calculation for species i after points with close to 100% conversion and
points from the rst 120 s of the experimental data sets have been
removed.
3.2. Parameter tuning
After data selection using D-optimal onion design a parameter estimation with the gradient search method is undertaken for
only the selected time points. The gradient search method is very
efcient for linear systems but can also be applied for nonlinear
systems such as catalyst models. For a nonlinear system the residual function is rst linearized for all parameters and then a step
in the parameter space is made in the direction of the steepest
descent (lowest residual). This process is usually repeated until
the change in residual is below a certain tolerance. The gradient
search method of choice in this work is the trust-region-reective
method (Coleman and Li, 1996) which is the standard method in
Matlab, the software used for parameter estimation in this project,
for over-determined nonlinear least square problems (function
call lsqnonlin). In theory the proposed methodology with PCA
and D-optimal design should make a gradient search more efcient compared to if all data points were used, since the correct
direction for parameter adjustment should be better dened by a
well-balanced subset of more sensitive data points.
Since the properties of the sensitivity matrix (Jacobian in all data
points) are dependent on the actual parameter values, the best
selection of time points for parameter tuning will change as the
parameters are tuned. To make the parameter tuning efcient it


wNO =

yNO,out yNO,out
wNO
yNO,in yNO,out,lim

(9)

was therefore necessary to select new time points for parameter


tuning as new parameter values developed. In the current work, the
parameter tuning was limited to only a few steps in the parameter
space which decreased the risk of obtaining a too large inuence
from the selected time points on the overall t. To investigate the
inuence of the number of steps on the nal t, two parameter
tuning methods were evaluated. The rst method (referred to as 1
step) selected new time points after every step in parameter space
while the second method (referred to as 5 steps) took ve steps
in parameter space before new time points were selected. The 1
step and 5 steps methods will together be referred to as the MVDA
methods.
The motivation for the 1 step approach was that the points
selected with D-optimal onion design were selected by evaluating
the sensitivity for all data points (Jacobian matrix) for a dened
set of parameter values. When the parameter values are changed,
as during a gradient search, the sensitivity also changes and the
point selection should be updated to ensure that the parameter
estimation is performed based on the statistically best data points.
By limiting the number of steps to one, new points will be selected
every time a new set of parameter values is found. The drawback
with the method is that the sensitivity for all data points needed to
be frequently calculated. This is computationally expensive since all
time points are used for these calculations and more computational
time was spent selecting new data points than performing actual
parameter estimation with the selected data points.
In the second approach (5 steps case), the number of steps
taken with the gradient search method (with the points selected
with D-optimal onion design) was instead limited to 5. Since the
steps with the gradient search method are still limited, the selected
time points may still be relevant, despite the changes in parameter
values. As a result more computational time is be spent on parameter tuning and less on selecting data points compared to the 1 step
method.
As a reference, a conventional parameter tuning with a gradient search method was also performed based on the entire data set
using the same data points as for the sensitivity analysis described
in Section A.3 in Appendix. The parameter tuning in the preceding study was performed with this method and data points but
the changed formulation of the residual calculation between the
two studies makes a new tuning necessary for comparison with
the MVDA-cases.
3.3. Stop criteria
Two stop criteria were applied in the study. The rst stop criterion was applied to the reduction of the residual sum of squares for
the full data set.
1. The parameter estimation was stopped if a new set of parameters
that reduced the residual sum of squares for the full data set by
more than 1 percent could not be found after six consecutive
steps in parameter space (six right hand loops in Fig. 2).
The second criterion was applied to the parameter estimation
performed with the data points selected with D-optimal design. If
no reduction of the residual sum of squares was achieved (referred

B. Lundberg et al. / Computers and Chemical Engineering 74 (2015) 144157

to as improved parameter values in Fig. 2) with the gradient search


method in the data points selected with D-optimal design, a new Doptimal design will be performed. Even though the scores matrix is
unchanged (since the parameters were unchanged) the D-optimal
design is likely to select different data points since the selection of
a design matrix with D-optimal design uses a random seed in the
selection process. A new parameter estimation with the gradient
search method was then performed on the new data set. This process was repeated until parameter values were found that reduce
the residual sum of squares of the points selected with D-optimal
design or until the second stop criteria was fullled.
2. The parameter estimation was stopped if no improved parameter values were found after more than six selections of data
points for the same parameter values (corresponding to more
than six left hand loops in Fig. 2)
If none of the criteria were fullled a new sensitivity calculation
on the whole data set (Jacobian matrix calculation) was performed
and the cycle returned to Parameter Sensitivity Calculations in
Fig. 2.
4. Results and discussion
4.1. Data point selection with D-optimal onion design
The data point selection with D-optimal onion design was
updated as the parameters were estimated according to the method
described previously in Section 3. Fig. 3 shows one example of a
selection from which it is clear that certain areas were frequently
selected from the onion layers with high sensitivity, whereas other
areas were mainly selected from the onion layers with low sensitivity.
The points selected were similarly for the different catalyst congurations even though points from catalyst conguration c (data
point 28,81843,240) were less frequently selected for the depicted
case. The points selected were taken from both transient and stationary sections of the data. As expected the data points selected
in layers with high sensitivity were mainly taken from transient
areas while data points selected in layers with low sensitivity were
mainly taken from more stationary areas.
4.2. Evolution of residuals
In Fig. 4 the evolution of the residual sum of squares of the
gas phase components (NO, HC, and CO) for the whole data set is
shown. The gure also shows the nal residual sum of squares for
the reference method as a benchmark for the two MVDA methods.
The residual sum of squares for the whole data set is only calculated after a completed parameter estimation in the time points
selected by D-optimal onion design. Neither of the methods (1 step
or 5 steps) fullled either of the two stop criteria in Section 3.3 and
were instead stopped after about 3 104 core hours of simulation.
For the reference case, the rst stop criterion was fullled
after about 5 103 core hours of simulation with a residual sum
of squares of 3580. The reference parameter tuning method was
then approaching parameter values where the catalyst model
was unable to nd a stable solution and thereby converging to a
local minimum. After 5 103 core hours of simulation the 5 steps
method had a residual sum of squares of 3256 but the 1 step method
on the other hand had a residual sum of squares of 8919, it is therefore not possible to conclude that the MVDA-methods enabled a
faster residual reduction. Instead the advantage with the methods
appears to be a reduced risk of converging to a local minimum.
Even after the comprehensive parameter tuning already performed,

149

Table 2
Residual sum square of each component and complete data set together with the
summation of residual sum square (rightmost column) for the different methods.

Start
1 step
5 steps
Ref

NO

HC

CO

Sum

18,986
515
558
1238

9736
1161
1193
1173

16,954
730
742
1169

45,675
2405
2494
3580

the residuals in Fig. 4 still show a declining trend even when both
methods have reached residual sum of squares below 2500.
Although the rate of decline appears to be decreasing, no clear
optima appear to have been reached.
4.3. Comparison to measurement data
The parameter values that gave the lowest residual sum of
squares, shown in Fig. 4, shall be regarded as the nal tuning results
of the methods. A compilation of the residual sum of sqares for the
nal results and the starting paramter values are shown for every
gas phase component in Table 2.
From these results it appears that HC was the most difcult component to t to measurement data for the MVDA methods which
may have several different reasons. Firstly HC and CO concentrations are highly correlated in the exhaust of a HDD engine which
means that it can be difcult to generate experimental data, let
alone subsets of data, where the parameters that inuence each
component can be distinguished. Since both components have high
concentrations at similar time points, the larger residual for CO at
the starting parameters may have inuenced the parameter estimation to initially focus more on reducing the CO residual at the
cost of decreased sensitivity for HC concentration. HC also differs from NO and CO since it is modeled as an average of a wide
range of hydrocarbons. The physical properties are more difcult
to model and the HC composition at the catalyst inlet will even
change between different engine operating points. For the reference case, the HC residual was at about the same level as for
the MVDA methods but both NO and CO residuals are noticeably higher. The results in Table 2 also indicate that both MVDA
methods achivieved a very similar nal t, with HC being the
largest residual, but that the reference method differs because the
total residual is more equally distibuted over the three components.
Two examples of the nal results of the tuning of the kinetic and
mass transport parameters for the different parameter estimation
methods are shown in Figs. 5 and 6.
Fig. 5 shows a very similar t for the two MVDA methods for all
components. The reference method shows a similar t for CO and
HC but for NO the transient was poorly captured. The NO production
before the change in engine operating point is likely a result of
NO2 reduction by CO or HC. These reactions are not included in
the kinetic model and can therefore not be captured by any of the
models.
Fig. 6 also shows a very similar t for the two MVDA methods, with the most noticeable difference compared to the reference
method, again mainly being the better t for NO. However the
NO concentrations after the transient did not agree as well with
measurement data as for the case in Fig. 5. The t of HC and CO
is somewhat better for the MVDA methods, but there were still
notable differences compared to measurement data.
The above gures show only two of 16 transients used for
parameter estimation (4 catalyst congurations and 4 operating
point changes for every conguration). The transients were, as previously mentioned, selected to give a wide temperature, ow and
concentration window. The combinations of torque and speed that

150

B. Lundberg et al. / Computers and Chemical Engineering 74 (2015) 144157

Fig. 3. Measured outlet mole fraction of NO for an example of a selected data set colored by relative sensitivity. The order of the engine operating points is 1, 7, 2, 8, 3, 4, 5, 6
which were run sequentially for catalyst congurations a, b, c and d (separated by blue vertical lines). The relative sensitivity is an indication of how sensitive the data point
is to changes in parameter value relative to the rest of the data set. A value of 0.9 for example, means that the data point has a higher sensitivity than 90% of the data set. (For
interpretation of the references to color in this gure legend, the reader is referred to the web version of this article.)

Fig. 4. Residual sum of squares for every set of parameters for the two evaluated methods. The dashed line is the residual sum of squares for the nal set of parameters from
the reference method. (a) 1 step method (b) 5 steps method (c) 1 step method, magnied (d) 5 steps method, magnied.

Table 3
Residual sum of squares for the experiment displayed in Figs. 5 and 6 for every component.
NO

HC

CO

Figure

1 step
5 steps
Ref

40.9
42.3
85.5

40.4
38.7
43.9

94.9
79.4
68.2

5
5
5

1 step
5 steps
Ref

66.2
76.6
212.8

112.6
111.9
112.2

80.2
101.9
157.2

6
6
6

B. Lundberg et al. / Computers and Chemical Engineering 74 (2015) 144157

b)

a)
1600

1200
[ppm]

1000

60
NO in
NO out measured
NO out sim, 1 Step
NO out sim, 5 Steps
NO out sim, Ref

C H in
3 6
50

[ppm]

1400

151

800
600

C3H6 out measured

40

C3H6 out sim, 1 Step

30

C3H6 out sim, Ref

C H out sim, 5 Steps


3 6

20

400
10

200
0
0

200

400

600

0
0

800 1000 1200 1400 1600 1800


Time [s]

200

400

600

c)

d)

250

500
CO in
CO out measured
CO out sim, 1 Step
CO out sim, 5 Steps
CO out sim, Ref

200
150

450
400
350
[C]

[ppm]

800 1000 1200 1400 1600 1800


Time [s]

100

300
250
Tin

200

50

150
0
0

200

400

600

100
0

800 1000 1200 1400 1600 1800


Time [s]

out

measured

Tout simulated
200

400

600

800 1000 1200 1400 1600 1800


Time [s]

Fig. 5. Measured and simulated outlet concentrations of NO (a), C3 H6 (b), CO (c) and temperature (d) for a change in operating point from 1 to 7 for catalyst conguration c.

result in high HC and CO are few and therefore only two of the
eight engine operating points used in the study generate signicant
HC and CO concentrations at the catalyst outlet. To demonstrate
the performance of the different methods, transients with engine
operating points with high HC and CO have been chosen in both
Figs. 5 and 6. In addition, the high temperature at the start of the
experiments provide an interesting transient change in both CO and
HC conversion for the rst 900 s and the change in engine operating
point provides transients in NO and temperature. To put the t of

the depicted experiments in Figs. 5 and 6 in relation to the t of the


remaining 14 experiments the residual sum of squares for the presented experiments are shown in Table 3. The trends in Table 3 for
two of the experiments agree with the results for all experiments
shown in Table 2. The t for NO is better for the MVDA methods
in both experiments while the t for HC is about the same for all
methods. Even though the t for CO is better for the reference case
in one of the experiments, the t of CO over both experiments in
the Table 3 is still better for the MVDA methods.

b)

a)
1000
800

[ppm]

700
600

NO in
NO out measured
NO out sim, 1 Step
NO out sim, 5 Steps
NO out sim, Ref

C3H6 out measured


C3H6 out sim, 1 Step

50

500
400
300

C H out sim, 5 Steps

40

3 6

C3H6 out sim, Ref

30
20

200

10

100
0
0

C H in
3 6

60

[ppm]

900

70

200

400

600

0
0

800 1000 1200 1400 1600 1800


Time [s]

200

400

600

c)

d)

250

450
CO in
CO out measured
CO out sim, 1 Step
CO out sim, 5 Steps
CO out sim, Ref

150

400
350
[C]

[ppm]

200

100

300
250
Tin

200
50
0
0

800 1000 1200 1400 1600 1800


Time [s]

150
200

400

600

800 1000 1200 1400 1600 1800


Time [s]

100
0

out

measured

Tout simulated
200

400

600

800 1000 1200 1400 1600 1800


Time [s]

Fig. 6. Measured and simulated outlet concentrations of NO (a), C3 H6 (b), CO (c) and temperature (d) for a change in operating point from 2 to 8 for catalyst conguration b.

1.00
12.3
10.9
4.85
7.71
1.00
0.706
0.706
0.793
1.00
1.00
0.740
0.730
0.828
0.701

actd a [%]
actb a [%]

3.98
50.6
3.96
1.35
1.93
2.08 10
3.88 105
2.59 105
5.33 104
8.25 104

1.00
1.00
1.00
1.00
0.766
3.10 104
2.28 104
2.68 104
2.87 104
3.05 104
9.65 104
9.67 104
9.81 104
9.74 104
9.69 104

acta , actc a [%]


EA ,7 [J/mol]
EA,6 [J/mol]

65.5
39.0
0.151
77.7
95.8
4.50 10
3.18 1013
1.77 1014
2.34 1017
2.71 1014
4.00 10
4.70 1017
1.06 1018
1.67 1022
1.25 1019
1.00 10
2.17 1015
2.59 1016
2.78 1020
4.89 1016

3.00 103
4.54 103
2.21 103
3.30 103
1.64 103
7.99 103
5.54 103
1.02 104
8.50 103
8.06 103
7.00 104
7.82 104
8.18 104
7.29 104
7.46 104
Start
1 step
5 steps
Ref
Preceding study

EA,5 [J/mol]
EA,4 [J/mol]
EA,3 [J/mol]

Start (Wang et al., 2008)


1 step
5 steps
Ref
Preceding study (Lundberg et al., 2014)

a
acta is activity scaling factor for catalyst conguration a (0.30 wt% Pt), actb is activity scaling factor for catalyst conguration b (0.59 wt% Pt), and actd is activity scaling factor for catalyst conguration d (0.59 wt% Pt with thin
washcoat).
b
Effective diffusivity scaling for small (S) and large (L) molecules.

[]
1.00
1.49
1.34
1.62
2.11

fDscl ,L
[]
b

fDscl ,S

8.00 10
7.33 104
7.85 104
7.90 104
7.67 104
4.79 10
4.85 105
2.99 106
4.84 1010
4.71 107

EA,2 [J/mol]
EA,1 [J/mol]

A7 []
A6 []
A4 []

A5 []

14

17

20

A2 [mol, K/m2 , s]

A3 [mol, K/m2 , s]
Table 4
Results of kinetic parameter tuning.

Table 4 shows the nal parameter values for both approaches


and the reference case together with the starting parameters values
from the study by Wang et al. (2008) and the parameters of the best
t from the preceding study by the authors (Lundberg et al., 2014).
Since catalyst conguration c (see Table 1) consisted of the catalyst
in conguration a in series with another identical catalyst, the same
activity scaling factor was assumed for both congurations a and c.
The correlation between activity scaling factors and preexponential factors in the numerator (A13 ) were linear. An increase
by a certain factor of the activity scaling factor gives the same
increase in reaction rate as an increase in the pre-exponential factor by the same factor. This was true since the kinetic model only
uses one type of reaction site. To make the kinetic parameters more
comparable, the activity scaling factors for every set of parameters
(1 step, 5 steps, Ref and preceding study) were scaled with a number
making the highest activity scaling factor of that set of parameters
equal to 1% at the same time as the pre-exponential factors in the
numerator were divided by the same number. This did not change
the behavior of the kinetic model; but merely shifted the factor
from one parameter to another. The activity scaling factor was in
other words only a handle to tune differences between the catalysts
and should only be interpreted as a relative difference and not an
absolute value for activity scaling.
In general it can be expected that a high noble metal loading
should correspond to a low dispersion and thereby also a low activity scaling factor. It is therefore not unexpected that the lowest
loading, i.e. 0.30 wt% Pt loading (acta and actc ), also has highest
activity scaling for all cases of the current study.
The complexity of the kinetic model and the catalyst model complicates a thorough analysis of every single parameter value but
some general remarks can be made. The original parameters gave
a too high conversion for all components and a very poor t in
general. The tuned parameters have therefore been changed in a
direction where the reaction rates in general are slower. This is not
always obvious when examining the parameter values since a large
pre-exponential factor in the numerator (A13 ) may be compensated by a large pre-exponential factor in the denominator (A47 )
and differences in activation energies (EA,j ) will make different preexponential factors signicant at different temperatures.
As indicated by the residual analysis in the previous, section the
parameter values are very similar for the two MVDA methods, while
the reference case stands out by having large pre-exponentials in
the numerator (A13 ). The reference case however also has a value
for A7 that is about ve orders of magnitude larger than for the
other models. The fact that A7 is in the denominator expression
for NO inhibition means that it will always have an effect since
the NO levels are always above zero due to thermodynamic limitations. The difference between the reference case and the MVDA
methods is thus a stronger dependence on NO concentration for
the reference case more than an overall large difference in reaction rates due to large pre-exponentials in the numerator (A13 ) for
the reference case. In the preceding study it was concluded that
the sign of EA,7 made the NO inhibition term mimic internal transport resistance since an increase in temperature increases the term
and thereby affects all reaction rates negatively. The sign of EA,7 is
still positive for all evaluated methods but for both MVDA methods EA,5 has changed from a negative starting value to a positive
nal parameter value. This means that the MVDA model has two
terms mimicking internal transport resistance, one as a function
of NO concentration and the other a function of HC concentration,
compared to one for the reference case. These values together with
the decreased mass transport resistance due to increased diffusivity scaling parameters mean that the actual transport resistance in
the model is reduced and replaced by a ctitious one connected to

4.4. Analysis of nal parameter values

10.0 104
9.51 104
9.52 104
9.86 104
9.80 104

B. Lundberg et al. / Computers and Chemical Engineering 74 (2015) 144157

A1 [mol, K/m2 , s]

152

B. Lundberg et al. / Computers and Chemical Engineering 74 (2015) 144157

the kinetics. The result of the parameter tuning with MVDA shows
the importance of having a well formulated model and also displays
how an efcient parameter tuning method can tune the parameters
to less realistic values, and still achieve a good t, even if the model
formulation is decient.
In general the parameter values from the preceding study are
closer to the values of the MVDA methods than the reference case.
The parameters from the preceding study also gave an overall better t with a residual sum of squares of 3094 compared to 3580
for the reference case. There are two main differences between the
parameter estimation method of the preceding study and the reference case. Firstly the residuals were calculated differently, see
Section 3.1, and secondly a ten times higher parameter weighting
for kinetic parameters (w in Eqs. (A15) and (A17)) was used in the
current study to make the Jacobian less sensitive to numerical noise.
It appears that this increased parameter weighting has resulted in
estimated parameter values further away from the original ones
with an overall poorer t. This could be an effect of some parameters, such as A1 A3 , having a large inuence on the t at an early
stage of the parameter estimation which makes them change fast
and thereby reach values where the sensitivity for the other parameters are low. In other words the risk of nding a local minimum
may be increased by the increased parameter weighting. One of the
aims of the MVDA method was to make the inuence of parameters
more equal and it appears that this has prevented the parameters
from reaching extreme values for both the 1 step and the 5 steps
method even though the parameter weighting was high.
4.5. Final remarks
By continuously changing data points on which the parameter
estimation is performed the risk of nding a local minimum appears
to have been reduced at the cost of longer computational time.
As mentioned in the introduction, these characteristics are typically associated with global parameter estimation methods such
as simulated annealing. The difference with the method at hand is,
however, that since a gradient search method is applied, the parameters will evolve without evaluating a large number of parameter
sets (as is characteristic of a global search method) which will put
less demand on the model stability. This will also reduce computational cost which already is high for such a complex system of
equations and large data set.
The method of parameter estimation should be useful for complex dynamic models where model stability limits what parameter
values are possible and where the number of data points is large.
The model is not likely to be suitable if the diversity of the data
points is low, for example if the model should be tuned to only a
few stationary data points. With low diversity in the data it is not
likely to achieve a notable improvement in parameter sensitivity
by changing the data point selection. The computational effort is
then better spent on tuning the model to the few available data
points directly. For models where the stability is highly dependent
on starting conditions, such as detailed kinetic models, the large
number of short simulation sections where starting conditions are
needed may make the method less suitable. For example it may be
very difcult to nd stable starting conditions if the rst data point
is in the middle of a sharp transient such as a catalyst ignition point.
5. Conclusions

153

method, referred to as the MVDA method, was applied to full scale


catalytic exhaust gas data to estimate kinetic parameters.
The MVDA method displays a better overall t for all components resulting in a residual sum of squares 32% below what
was achieved with a conventional method used as reference. The
method displayed less tendencies to converge to local minima and
to reach areas where the catalyst model is unstable; however, the
method was more computationally expensive than the reference
method. In general the method should be applicable on other complex dynamic models where model stability limits what parameter
values are possible and where the number of data points is large.
Both evaluated congurations of the MVDA method gave similar
nal t to measurement data. Neither method could be identied
as less computationally expensive than the other.
Acknowledgements
The computations were performed on resources provided by
the Swedish National Infrastructure for Computing (SNIC) at C3SE.
Financial support from Swedish Energy Agency (Grant Nos. 322821 and 32282-2) is gratefully acknowledged.
Appendix.
A.1. Reactor model
Quasi-steady state was considered to prevail for gas phase
species (Eqs. (A1) and (A2)) and gas temperature (Eq. (A6)) since
the characteristic time constants are far shorter for these processes
compared to the solid heat balance (Eq. (A7)) and variations in inlet
conditions.
The gas bulk mass balance is given by:

0=

Vk,0

dci,k,0
dt

= Ftot (yi,k1,0 yi,k,0 ) i,k,0 (ci,k,0 ci,k,1 ) (A1)

and the washcoat mass balance is given by:

0=

dci,k,n
Vk,n
dt

= i,k,n1 (ci,k,n1 ci,k,n )

i,k,n (ci,k,n ci,k,n+1 ) +

vi,j actq rj,k,n Am,k,n

(A2)

for n1
where rj,k,n is the reaction rate in mol/m2 Pt/s and Am,k,n is the catalyst active surface area in channel segment k and wall layer n,
Vk,n is the volume of channel segment k and layer n and is the
washcoat porosity. Index k indicates the segment number (axial
discretization) where k = 1 represents the rst segment in the direction of the ow and k = K represents the last. The index n indicates
the layer number (radial discretization) where n = 1 represents the
rst washcoat layer closest to the gas bulk, n = N represents the last
layer and n = 0 represents the gas bulk. The index i indicates the
species and index q represent catalyst conguration.
The mass transfer coefcients  i,k,n are given by:
i,k,0 =

Ak
(1/kc,i,k ) + (0.5x1 /Deff,i,k )

(A3)

i,k,n =

Deff,i,k Ak
0.5xn + 0.5xn+1

(A4)

for n = 1,. . ., N 1. And for n = N it is:


The method of parameter estimation evaluated in the current
paper is based on data screening by sensitivity analysis where
MVDA has been applied to a large transient data set to select a
subset on which parameter estimation is performed. The subset
was continuously updated as the parameter values developed. The

i,k,N = 0

(A5)

where Ak is the radial mass and heat transfer area in channel segment k. For simplicity, Ak was assumed to be constant for all wall
layers. Deff,i,k is the effective pore diffusion coefcient of component

154

B. Lundberg et al. / Computers and Chemical Engineering 74 (2015) 144157

i in channel segment k, kc,i,k is the lm transfer coefcient of component i in channel segment k and xn is the thickness of washcoat
layer n.
The gas energy balance is given by:

0=

Vk,0 cp,g

dTg,k

dt

Deff,i,k =
= Ftot cp,g (Tg,k1 Tg,k ) hk Ak (Tg,k Ts,k )
(A6)

where cp,g is the heat capacity for the gas, hk is the heat transfer
coefcient in channel segment k, Tg,k and Ts,k are the temperatures
in channel segment k in the gas bulk and of the catalyst respectively. Note that the solid temperature, Ts,k , was not discretized
radially since the high solid conductivity and short conduction distance will result in very small radial temperature gradients. The
main temperature transport resistance is in other words assumed
to be in the lm. The solid energy balance is given by:
dTs,k
dt

cp,s ms,k +

sheat
K


n

actq rj,k,n Am,k,n (Hj )

(A7)

where ms,k is the mass of catalyst material and substrate in channel segment k, cp,s is the mass weighted average of washcoat and
substrate heat capacity, As is the channel cross sectional area of the
substrate and washcoat, Hj is the heat of reaction j. sheat is a heat
sink originating from solid material not included in the catalyst
(e.g. catalyst canning and insulation). The heat loss to the environment is modeled by the lumped heat transfer parameter tot and
the effective environmental temperature T . K is the total number
of channel segments.
The solid axial heat ux was calculated as:
Ts,k Ts,k1
0.5zk + 0.5zk1

(A8)

for k = 2,. . ., K 1, K
qk = 0

(A9)

hk =

ShDi,k
d

Nug
d

Di,k = Dref,i

Ts,k
Tref

dp
3

8RTs,k
Mi

(A14)

where dp is the mean pore diameter.


A.2. Adjustable parameters

(A15)

where w is a weight factor for every parameter and p is the parameter that is changed by the gradient search method. The superscript
o in the equations above indicates that this is the original values
taken from literature or previous successful parameter tuning. The
reaction rate coefcients were centered around a reference temperature to decrease the parameter correlation (Box et al., 1973).
o
kj,ref
= Aoj e

(E o /RTref )

(A16)

A,j

The reaction rate coefcient kj,ref was scaled according to:


o
ln(kj,ref ) = ln(kj,ref
) + wa,j pa,j

(A17)

The tuned rate constant at a reference temperature was then


used to calculate the pre-exponential factor
(A18)

When Eq. (A18) is inserted into the expression for the reaction
rate coefcient in Eq. (6), the following expression for the tuned
reaction rate coefcient is obtained
kj = kj,ref e[(EA,j /R)(1/TS 1/Tref )]

(A19)

The mass transport was tuned by adjusting the effective diffusivities for the species taking part in the reactions.

(A10)

fDscl ,i = fDo

(A11)

This will change the expression for the effective diffusivity


according to

where Di,k is the diffusion coefcient for component i in channel


segment k, g is the gas heat conductivity and d is the channel
dimension. Sh and Nu are the Sherwood and the Nusselt numbers,
respectively. Only the asymptotic values for Sh and Nu were used
and thus entrance effects were neglected. Asymptotic values were
taken from Tronconi and Forzatti (1992).
The gas diffusivities for each component i at reference temperature (Tref ) were calculated using the FullerSchettletGiddins
equation (Murzin and Salmi, 2005). Their dependence on temperature was expressed as:

DKi,k =

Aj = kj,ref eEA,j /RTref

for k = 1, k = K + 1where s is the heat conductivity for the solid material (same conductivity assumed for the washcoat and the cordierite
support) and zk is the length of channel segment k (the total heat
transfer distance is half the distance of each segment).
The mass and heat transfer is described with the lm model and
the mass and heat transfer coefcients are given by:
kc,i,k =

(A13)

where fD is a factor that takes into consideration the porosity and


the tortuosity of the porous material. DKi,k is the Knudsen diffusivity
which was calculated as:

o
EA,j = EA,j
+ WEa,j pEa,j

tot
(Ts,k T )
K

fD
(1/Di,k ) + (1/DKi,k )

Both the pre-exponential factors, Aj , and the activation energies, EA,j , can be tuned. The scaling of EA,j was straightforward and
performed according to:

= hk Ak (Tg,k Ts,k ) As (qk+1 qk )


+

qk = s

The effective diffusivity for pore diffusion was initially estimated


based on an additive resistance, also known as the Bosanquet formula (Froment et al., 2011):

1.75
(A12)

scl ,i

Deff,i,k =

+ wDscl ,i pDscl ,i

fD
fD ,i
(1/Di,k ) + (1/DK,i,k ) scl

The value for fDo

scl ,i

(A20)

(A21)

is typically around 1.

The heat loss to the environment was modeled as a part of the


total heat balance in Eq. (A7) and the parameters sheat , tot and T
could be adjusted. The parameter scaling and tuning was performed
in the same way as the other parameters:
o
T = T
+ wT pT

(A22)

tot = otot + Wtot ptot

(A23)

o
sheat = sheat
+ Wsheat psheat

(A24)

B. Lundberg et al. / Computers and Chemical Engineering 74 (2015) 144157

A.3. PCA and parameter sensitivity


Principal Component Analysis, PCA, is a mathematical method
where large sets of observations of possibly correlated variables
are orthogonally transformed into linearly uncorrelated variables.
The number of uncorrelated variables, referred to as principal components, is usually chosen to be of a smaller magnitude compared
to the number of variables in the untreated data. The PCA can be
performed with several different purposes such as identication of
classes of data and outliers, simplication, data reduction, modeling, variable selection, and prediction (Martens and Naes, 1989). A
simple way of describing the method would be as a simplication
of a data matrix (X) where the more similarity within the objects
results in fewer terms needed for a satisfactory representation of
the data. The resulting linear model can be described by Eq. (A25):
X = TP + E

(A25)

where T is the scores matrix, P is the loading matrix, and E is the


error. If the dimension of X is N K then T has the dimension N A
and P has the dimension A K, where A is the number of components in the PCA.
The scores matrix can be used to interpret the relation between
the N observations (rows in X) and the loading matrix can be used
to interpret the relation between the K variables (columns in X).
In the present study the Jacobian (J), or the parameter sensitivity
matrix, was used as the X-matrix. The Jacobian is dened as:
J(f, ) =

f


155

data set and could even result in a sub-optimization if used for


parameter estimation. In a D-optimal onion design the data set is
divided into layers where every layer includes points in a specied
parameter sensitivity range. The so-called onion D-optimal design
is then applied to every onion layer and a number of data points will
thereby be selected in every parameter sensitivity range, resulting in a more balanced data point selection. The D-optimal onion
design method, as applied in the current study, can be divided into
a number of steps.
1. The scores matrix from the PCA analysis (T) were column-wise
centered and scaled. The new matrix is called Tcs .
2. The elements of Tcs were squared, summed row-wise and thereafter sorted from high to low. The resulting vector is denoted
Tcs2,sort and has the data points with the highest parameter sensitivity rst and the data points with the lowest sensitivity last.
3. The rows in Tcs were sorted in the same order as Tcs2,sort and is
denoted Tcs,sort
4. The number of onion layers (nlayers ) was calculated from the
number of data points (m) and number of components in the PCA
analysis (n): nlayers = 0.01m/(n + 1) 1, where nlayers was rounded
up to the closest integer.
5. The number of data points in every layer was calculated according to
npoints (j) = 100
npoints (j) = 10(n + 1)kj2

j=1
j = [2, 3, . . ., nlayers , nlayers + 1]

(A26)

where f is some function value which, in the case of the current


study, was the residuals for HC, CO and NO (as described in Eqs. (8)
and (9)) and  was the parameter vector in the nonlinear model.
The number of columns in the J matrix will then correspond to
the number of residual types (different components in this study)
times the number of parameters to be tuned. In the current study
the Jacobian was estimated by nite differences and the number of
rows was equal to the number of time points (observations). Before
the PCA was performed on J, all the rows with all components set to
zero were removed. This was either a result of the time points being
within the rst 120 s of an experiment or all conversions being close
to 100% according to the limits given in the previous section.
In the PCA model of the Jacobian matrix the resulting scores
matrix contained information about the relation between the time
points and the loadings matrix contained information about the
relation between the residual sensitivities. By analyzing the scores
matrix it was possible to identify the time points where the inuence from parameter changes on the different residuals were most
noticeable, that is the parameter sensitivity was high. These time
points were thereby good candidates to use if the number of time
points used for parameter tuning were to be reduced.
A.4. D-optimal onion design
For linear models traditional experimental designs such as full
factorial designs, fractional factorial designs, and response surface
designs are suitable when the factors are relatively unconstrained.
For nonlinear models less traditional experimental designs such as
D-optimal design are more favorable, especially when the data is
dened (in a so called Candidate set) and also when the number of
experimental runs is restricted (Martens and Naes, 1989).
The advantage with applying a D-optimal onion design, instead
of a regular D-optimal design, is that the selection of points from
the data set will be more balanced. A regular D-optimal design will
result in a selection where only the most extreme points are chosen for parameter estimation. These points will have the highest
parameter sensitivity but may not be representative for the whole

k
was
calculated
to
fulll
npoints
where

(1:nlayers ) m npoints (1:nlayers + 1)
6. A vector denoted int was created to divide the data set into
intervals:
int (j) = 0

npoints (1 : j 1)
int (j) =

j=1
j = [2, 3, . . ., nlayers , nlayers + 1]

7. The matrixes on which D-optimal design was applied then


became
Tonion,j = [Tcs,sort (int(j) + 1 : int(j + 1)),
2
(Tcs,sort (int(j) + 1 : int(j + 1))) ]
Tonion,j = Tcs,sort (int(j) + 1 : int(j + 1))

j=1
j = [2, 3, . . ., nlayers ]

8. D-optimal design was applied to different layers of the onion


design according to
D-optimal (Tonion,j , 2n + 2)
D-optimal (Tonion,j , n + 1)

j=1
j = [2, 3, . . ., nlayers ]

where the rst input to D-optimal design was the rows in the
selected interval and the second input was the number of points
to be selected.
The method results in a selection of 1% of the total number of
data points on which PCA was applied. The exponentially increasing
number of points in every onion layer resulted in the majority of
the points being selected at high sensitivity since an equal number
of points (number of components in the PCA +1) was selected from
every layer except the rst. An example of the sectioning of the
data set into layers and the relative parameter sensitivity of the
data points in the layers is shown in Fig. A1.
In the rst onion layer, where the parameter sensitivity is highest, the selected rows in the centered and scaled scores matrix
was expanded by additional columns containing the same columns
again but squared. The reason for this was to capture the nonlinear
behavior in the parameter sensitivity.

156

B. Lundberg et al. / Computers and Chemical Engineering 74 (2015) 144157

Fig. A1. D-optimal onion layer thickness colored by relative sensitivity. The relative sensitivity is an indication of how sensitive the data point is to changes in parameter
value relative to the rest of the data set. The color for one bar is set to the average relative sensitivity of the data points in that onion layer.

The D-optimal design algorithm selects rows (time points) from


the different onion layers that maximizes the parameter volume
(e.g. the determinant of (Tonion,j )T Tonion,j ). The D-optimal algorithm
used here was that contained in the Matlab function candexch
(included in Matlab Statistics Toolbox). By this algorithm the time
points with the most inuence on the change in determinant (highest sensitivity) could be identied. Together with the scaling of the
scores matrix (T), this ensured that the data points selected in every
layer widely spread the sensitivity for the different parameters. In
other words, the D-optimal design algorithm aided the selection
of data to provide a balanced inuence from parameters and the
onion layer method provided a balanced inuence from different
parts of the data set. The fact that different parts of the data set had
different parameter sensitivity becomes apparent when one set of
selected data points is studied as was illustrated in Section 4.1.

References
Ansell GP, Bennett PS, Cox JP, Frost JC, Gray PG, Jones AM, et al. The development
of a model capable of predicting diesel lean NOx catalyst performance under
transient conditions. Appl Catal B: Environ 1996;10:183201.
Berger RJ, Kapteijn F, Moulijn JA, Marin GB, De Wilde J, Olea M, et al. Dynamic
methods for catalytic kinetics. Appl Catal A: Gen 2008;342:328.
Bernaerts K, Versyck KJ, Van Impe JF. On the design of optimal dynamic experiments
for parameter estimation of a Ratkowsky-type growth kinetics at suboptimal
temperatures. Int J Food Microbiol 2000;54:2738.
Blockley R. Encyclopedia of aerospace engineering. Chichester, West Sussex,
UK/Hoboken, NJ: John Wiley & Sons; 2010. p. 522938.
Box MJ. The occurrence of replications in optimal designs of experiments to
estimate parameters in non-linear models. J R Stat Soc Ser B (Methodol)
1968;30:290302.
Box GEP, Hunter WG. The experimental study of physical mechanisms. Technometrics 1965;7:2342.
Box GEP, Lucas HL. Design of experiments in non-linear situations. Biometrika
1959;46:7790.
Box GEP, Hunter WG, MacGregor JF, Erjavec J. Some problems associated with the
analysis of multiresponse data. Technometrics 1973;15:3351.
Chen DKS, Bissett EJ, Oh SH, Van Ostrom DL. A three-dimensional model for the analysis of transient thermal and conversion characteristics of monolithic catalytic
converters. SAE, international technical paper number: 880282; 1988.
Coleman TF, Li YY. An interior trust region approach for nonlinear minimization
subject to bounds. Siam J Optim 1996;6:41845.
Crocoll M, Kureti S, Weisweiler W. Mean eld modeling of NO oxidation over
Pt/Al2 O3 catalyst under oxygen-rich conditions. J Catal 2005;229:4809.
Draper NR, Hunter WG. Design of experiments for parameter estimation in multiresponse situations. Biometrika 1966;53:52533.
Draper NR, Hunter WG. The use of prior distributions in the design of experiments for
parameter estimation in non-linear situations. Biometrika 1967a;54:14753.

Draper NR, Hunter WG. The use of prior distributions in the design of experiments for
parameter estimation in non-linear situations: multiresponse case. Biometrika
1967b;54:6625.
Dubien C, Schweich D, Mabilon G, Martin B, Prigent M. Three-way catalytic converter modelling: fast- and slow-oxidizing hydrocarbons, inhibiting species, and
steam-reforming reaction. Chem Eng Sci 1998;53:47181.
AA. The microkinetics of
Dumesic JA, Rudd DF, Aparicio LM, Rekoske JE, Trevino
heterogeneous catalysis. Washington, DC: American Chemical Society; 1993.
Ericson C, Westerberg B, Odenbrand I. A state-space simplied SCR catalyst model
for real time applications. SAE number: 2008-01-0616; 2008.
Franceschini G, Macchietto S. Model-based design of experiments for parameter
precision: state of the art. Chem Eng Sci 2008a;63:484672.
Franceschini G, Macchietto S. Novel anticorrelation criteria for model-based experiment design: theory and formulations. AIChE J 2008b;54:100924.
Froment GF, De Wilde J, Bischoff KB. Chemical reactor analysis and design. Hoboken,
NJ: Wiley; 2011.
Glielmo L, Santini S. A two-time-scale innite-adsorption model of three way
catalytic converters during the warm-up phase. J Dyn Syst Meas Control
1999;123:6270.
Goodwin GC. Dynamic system identication: experiment design and data analysis/Graham C. Goodwin and Robert L. Payne. New York: Academic Press; 1977.
Guojiang W, Song T. CFD simulation of the effect of upstream ow distribution
on the light-off performance of a catalytic converter. Energy Convers Manag
2005;46:201031.
Gthenke A, Chatterjee D, Weibel M, Krutzsch B, Koc P, Marek M, et al. Current status of modeling lean exhaust gas aftertreatment catalysts. In: Marin GB, editor.
Advances in chemical engineering. Academic Press; 2007. p. 103283.
Hill WJ, Hunter WG, Wichern DW. A joint design criterion for the dual
problem of model discrimination and parameter estimation. Technometrics
1968;10:14560.
Hunter WG, Hill WJ, Henson TL. Designing experiments for precise estimation of all
or some of the constants in a mechanistic model. Can J Chem Eng 1969;47:7680.
Kandylas IP, Koltsakis GC. NO2 -assisted regeneration of diesel particulate lters: a
modeling study. Ind Eng Chem Res 2002;41:211523.
Koltsakis GC, Stamatelos AM. Modeling dynamic phenomena in 3-way catalytic
converters. Chem Eng Sci 1999;54:456778.
Koltsakis GC, Konstantinidis PA, Stamatelos AM. Development and application range
of mathematical models for 3-way catalytic converters. Appl Catal B: Environ
1997;12:16191.
Lafossas F, Matsuda Y, Mohammadi A, Morishima A, Inoue M, Kalogirou M, et al.
Calibration and validation of a diesel oxidation catalyst model: from synthetic
gas testing to driving cycle applications; 2011.
Lundberg B, Sjoblom J, Johansson , Westerberg B, Creaser D. Parameter estimation of a DOC from engine rig experiments with a discretized catalyst washcoat
model. SAE Int J Engines 2014;7:1093112.
Martens H, Naes T. Multivariate calibration. John Wiley & Sons; 1989.
Massman WJ. A review of the molecular diffusivities of H2 O, CO2 , CH4 , CO, O3 ,
SO2 , NH3 , N2 O, NO, and NO2 in air, O2 and N2 near STP. Atmos Environ
1998;32:111127.
Mladenov N, Koop J, Tischer S, Deutschmann O. Modeling of transport and
chemistry in channel ows of automotive catalytic converters. Chem Eng Sci
2010;65:81226.
Murzin D, Salmi T. Mass transfer and catalytic reactions. Catalytic kinetics. Amsterdam: Elsevier Science; 2005. p. 341418.

B. Lundberg et al. / Computers and Chemical Engineering 74 (2015) 144157


Oh SH, Cavendish JC. Transients of monolithic catalytic converters. Response to
step changes in feedstream temperature as related to controlling automobile
emissions. Ind Eng Chem Prod Res Dev 1982;21:2937.
Olsson L, Fridell E, Skoglundh M, Andersson B. Mean eld modelling of NOx storage
on Pt/BaO/Al2 O3 . Catal Today 2002;73:26370.
Olsson I-M, Gottfries J, Wold S. D-optimal onion designs in statistical molecular
design. Chemom Intell Lab Syst 2004;73:3746.
Pandya A, Mmbaga J, Hayes RE, Hauptmann W, Votsmeier M. Global kinetic
model and parameter optimization for a diesel oxidation catalyst. Top Catal
2009;52:192933.
Pinto JC, Lobo MW, Monteiro JL. Sequential experimental design for parameter
estimation: a different approach. Chem Eng Sci 1990;45:88392.
Ramanathan K, Sharma CS. Kinetic parameters estimation for three way catalyst
modeling. Ind Eng Chem Res 2011;50:996079.
Salomons S, Votsmeier M, Hayes RE, Drochner A, Vogel H, Gieshof J. CO and
H2 oxidation on a platinum monolith diesel oxidation catalyst. Catal Today
2006;117:4917.
Sjoblom J, Creaser D. Latent variable projections of sensitivity data for experimental
screening and kinetic modeling. Comput Chem Eng 2008;32:31219.

157

Stamatelos AM, Koltsakis GC, Kandylas IP, Pontikakis GN. Computer aided engineering in diesel exhaust aftertreatment systems design. Proc Inst Mech Engin Part
D: J Automob Eng 1999;213:54560.
epnek J, Koc P, Marek M, Kubcek M. Catalyst simulations based on coupling of
St
3D CFD tool with effective 1D channel models. Catal Today 2012;188:8793.
Tronconi E, Forzatti P. Adequacy of lumped parameter models for SCR reactors with
monolith structure. AIChE J 1992;38:20110.
Tsinoglou DN, Koltsakis GC. Effect of perturbations in the exhaust gas composition
on three-way catalyst light off. Chem Eng Sci 2003;58:17992.
Voltz SE, Morgan CR, Liederman D, Jacob SM. Kinetic study of carbon monoxide and
propylene oxidation on platinum catalysts. Prod R&D 1973;12:294301.
Wang TJ, Baek SW, Lee JH. Kinetic parameter estimation of a diesel oxidation catalyst
under actual vehicle operating conditions. Ind Eng Chem Res 2008;47:252837.
Watling TC, Ahmadinejad M, Tutuianu M, Johansson , Paterson MAJ. Development
and validation of a PtPd diesel oxidation catalyst model. SAE technical paper
number: 2012-01-1286; 2012.
Zullo LC. (PhD thesis) Computer aided design of experiments: an engineering
approach (PhD thesis). University of London; 1991.

You might also like