The Effect of Grid Resolution and Oxygen Storage in a One-Dimensional Monolithic Three-Way Catalyst Model

Similar documents
Identification of catalytic converter kinetic model using a genetic algorithm approach

Mathematical Investigation and CFD Simulation of Monolith Reactors: Catalytic Combustion of Methane

Kinetic Parameters Estimation using Vehicle Data for Exhaust Aftertreatment Devices

Three-Way Catalytic Converter Modeling as a Modern Engineering Design Tool

Nonlinear dynamics of three-way catalyst with microkinetics and internal diffusion

INTRODUCTION TO CATALYTIC COMBUSTION

DARS Digital Analysis of Reactive Systems

Mathematical Investigation and CFD Simulation of Monolith Reactors: Catalytic Combustion of Methane

Explanation: They do this by providing an alternative route or mechanism with a lower activation energy

ADVANCED DES SIMULATIONS OF OXY-GAS BURNER LOCATED INTO MODEL OF REAL MELTING CHAMBER

ONE-DIMENSIONAL HIGH-FIDELITY AND REDUCED-ORDER MODELS FOR THREE- WAY CATALYTIC CONVERTER

Identification of Catalytic Converter Kinetic Model Using a Genetic Algorithm Approach

AUTOMOTIVE EXHAUST AFTERTREATMENT

DETAILED MODELLING OF SHORT-CONTACT-TIME REACTORS

Understanding NH 3 emissions over a three-way catalyst in lean/rich conditions.

Numerical simulation of fluid flow in a monolithic exchanger related to high temperature and high pressure operating conditions

A Numerical Study of Catalytic Light-Off Response THESIS

CFD study of gas mixing efficiency and comparisons with experimental data

Elements and Their Oxides

CHAPTER 7 NUMERICAL MODELLING OF A SPIRAL HEAT EXCHANGER USING CFD TECHNIQUE

A Machine Learning Approach to Modeling and Identification of Automotive Three-Way Catalytic Converters

Simulation of Selective Catalytic Reduction using DARS 1D Tool

Interactions between oxygen permeation and homogeneous-phase fuel conversion on the sweep side of an ion transport membrane

Kinetic Parameter Identification for a DOC Catalyst Using SGB test and Advanced Optimization Algorithms

Prediction of Catalytic Converter Efficiency in Natural Gas Engine for Nitrogen Monoxide and Carbon Monoxide Emission Control

This section develops numerically and analytically the geometric optimisation of

Simulation of Engine Exhaust Aftertreatment with CFD using Detailed Chemistry

University of Bath DOI: / Publication date: Link to publication. Sage

Modeling of Humidification in Comsol Multiphysics 4.4

Numerical Investigation of Wall Heat Conduction Effects in Catalytic Combustion Systems

Advanced Turbulence Models for Emission Modeling in Gas Combustion

The Influence of Geometry on the Performance of Catalytic Converter

Available online at ScienceDirect. IFAC-PapersOnLine (2015)

Three Way Catalyst and Lean NO x Trap Modeling for a Lean Burn SIDI Gasoline engine

Emissions Catalyst Design using GT-SUITE. Dr. Chaitanya Sampara Viridis Chemicals Private Limited

Let us consider the classical transport problem of a fluid

Predicting the Effect of Catalyst Axial Active Site Distributions on a Diesel Oxidation Catalyst Performance

Summary of Dimensionless Numbers of Fluid Mechanics and Heat Transfer

Model-based analysis of TWCcoated filters performance aspects

7.2 Sublimation. The following assumptions are made in order to solve the problem: Sublimation Over a Flat Plate in a Parallel Flow

ALE 1. Chemical Kinetics: Rates of Chemical Reactions

Direct pore level simulation of premixed gas combustion in porous inert media using detailed chemical kinetics

CURVE is the Institutional Repository for Coventry University

Three Way Catalyst Control using PI-style Controller with HEGO Sensor Feedback

Correlation between LNT NH 3 and N 2 O selectivities under fast cycling conditions

Surface Chemistry Tutorial

The Seeding of Methane Oxidation

Advanced Transport Phenomena Module 6 - Lecture 28

Catalytic bead sensors are used primarily to detect

* 1 K.Mohan Laxmi, 2 V.Ranjith Kumar, 3 Y.V.Hanumantha Rao

Zeolite HC Traps: Experiments and Simulations

Fuel Cell System Model: Auxiliary Components

DETAILED DYNAMIC MODELING OF AUTO-THERMAL GASOLINE FUEL-PROCESSORS

Overview of Turbulent Reacting Flows

COMBUSTION CHEMISTRY COMBUSTION AND FUELS

Modeling of Packed Bed Reactors: Hydrogen Production by the Steam Reforming of Methane and Glycerol

A Systematic Approach for Performance Comparisons of NO x Converter Designs. Michelle Bendrich. Master of Science. Chemical Engineering

Predicting NO Formation with Flamelet Generated Manifolds

Chemical Reaction Engineering Prof. Jayant Modak Department of Chemical Engineering Indian Institute of Science, Bangalore

Rates of Reaction HL

Phone: , For Educational Use. SOFTbank E-Book Center, Tehran. Fundamentals of Heat Transfer. René Reyes Mazzoco

CFD STUDIES IN THE PREDICTION OF THERMAL STRIPING IN AN LMFBR

DARS overview, IISc Bangalore 18/03/2014

External Forced Convection :

Analysis of homogeneous combustion in Monolithic structures

GPC 2001 GLOBAL POWERTRAIN CONGRESS:

Rate of reaction refers to the amount of reactant used up or product created, per unit time. We can therefore define the rate of a reaction as:

AP Chem Chapter 14 Study Questions

Chemistry Assessment Unit AS 2

If there is convective heat transfer from outer surface to fluid maintained at T W.

Design of experiments and empirical models for up to date burners design for process industries

Overview of Reacting Flow

Thermal Energy Final Exam Fall 2002

CFD STUDY OF MASS TRANSFER IN SPACER FILLED MEMBRANE MODULE

Combustion Theory and Applications in CFD

MICRO SOLID OXIDE FUEL CELL ON THE CHIP

Balancing chemical reaction equations (stoichiometry)

Modelling of a Passive Catalytic Recombiner for Hydrogen Mitigation by CFD Methods

This chapter focuses on the study of the numerical approximation of threedimensional

The Combination of Detailed Kinetics and CFD in Automotive Applications

CFD Simulation of Catalytic Combustion of Benzene

NO x Emissions Prediction from a Turbulent Diffusion Flame

Ugur Pasaogullari, Chao-Yang Wang Electrochemical Engine Center The Pennsylvania State University University Park, PA, 16802

Mechanical Engineering. Postal Correspondence Course HEAT TRANSFER. GATE, IES & PSUs

Convective Mass Transfer

(g) 2NH 3. (g) ΔH = 92 kj mol 1

Axial profiles of heat transfer coefficients in a liquid film evaporator

Chemical Kinetics of HC Combustion

Sustainable Power Generation Applied Heat and Power Technology. Equations, diagrams and tables

PAPER No. 4: Environmental Chemistry MODULE No. 19: Photochemical reaction in atmosphere and smog formation

International Journal of Scientific & Engineering Research, Volume 6, Issue 5, May ISSN

D DAVID PUBLISHING. 1. Introduction. Akira Nishimura 1, Masashi Baba 1, Kotaro Osada 1, Takenori Fukuoka 1, Masafumi Hirota 1 and Eric Hu 2

Numerical Modeling of Laminar, Reactive Flows with Detailed Kinetic Mechanisms

Fundamentals of Combustion

Convective Heat and Mass Transfer Prof. A.W. Date Department of Mechanical Engineering Indian Institute of Technology, Bombay

Introduction to Heat and Mass Transfer. Week 12

Hydrogen addition to the Andrussow process for HCN synthesis

Laminar flow heat transfer studies in a twisted square duct for constant wall heat flux boundary condition

Chapter 1 Introduction and Basic Concepts

1. A. Define the term rate of reaction. The measure of the amount of reactants being converted into products per unit amount of time

Transcription:

The Effect of Grid Resolution and Oxygen Storage in a One-Dimensional Monolithic Three-Way Catalyst Model Denis I. Andrianov Robert J. Dingli Michael J. Brear Chris Manzie Department of Mechanical Engineering, The University of Melbourne, Victoria 31, Australia (E-mail: d.andrianov@pgrad.unimelb.edu.au, rdingli@unimelb.edu.au, mjbrear@unimelb.edu.au, manziec@unimelb.edu.au) Abstract: A one-dimensional model of a monolithic three-way catalyst is presented, featuring a reduced order chemical kinetic scheme comprised of ten reactions and including oxygen storage. The model behaviour is validated against measurements taken from air/fuel ratio sweeps as well as European emission drive cycle data, covering a broad range of flow velocities, gas composition and inlet temperatures. The effects of grid resolution and oxygen storage reactions are investigated to provide insight into the minimum modelling requirements to capture key phenomena such as light-off time and conversion efficiency of the catalyst. Keywords: Reduced-order models, automotive emissions, engine control, optimisation problems 1. INTRODUCTION The major pollutants from vehicles with internal combustion engines are unburned hydrocarbons (HC), carbon monoxide (CO) and oxides of nitrogen (NO X ). Tail-pipe NO X emissions are largely comprised of NO and not NO 2 (Twigg, 27). In the atmosphere, in the presence of various unburned hydrocarbons and CO, N O is converted to NO 2. Under the influence of light, a fraction of NO 2 decomposes back into NO, losing an oxygen atom, which quickly combines with atmospheric oxygen O 2, forming harmful ozone O 3. NO 2 is eventually transformed almost entirely into nitric acid (HNO 3 ) in the form of acid rain, acid fog, or dry deposition (Seinfeld, 24; Twigg, 27). The first attempt to tackle this problem was with the introduction of a catalytic converter in 1975, which helped to cut HC and CO emissions by roughly 96% and NO X emissions by 75% (Seinfeld, 24). Most of the early catalysts were in pellet form and were soon abandoned due to catalyst wear and subsequent deterioration in their performance. Newer catalytic converters feature a monolithic structure with a large number of long thin channels, the surface of which is coated with a catalytically active material, referred to as a washcoat. Monolithic substrates are often manufactured from cordierite, a ceramic with a very low coefficient of thermal expansion. This helps to prevent cracking of the material when it experiences high temperatures and spatial temperature gradients during normal operation. The washcoat is typically on the order of 2 15 µm thick (Twigg, 27) and contains a combination of two or more of the following metals: platinum (Pt), palladium (Pd) and rhodium (Rh). A catalyst that can simultaneously remove HC, NO X and CO emissions is called a three-way catalyst (TWC). For the conversion efficiency to be maximised, TWC s usually require combustion to take place at close to a stoichiometric air/fuel mixture. Since precisely stoichiometric conditions are difficult to maintain, the washcoat is fitted with some oxygen storage compounds such as ceria (cerium oxides), which absorb oxygen under lean conditions and release oxygen under rich conditions. This mechanism improves the performance of TWC s by helping to maintain the exhaust around stoichiometric conditions. During roughly the last 25 years, some significant contribution has been made towards the understanding of the processes within the TWC, its mathematical modeling and simulation. However, many of the available models in the literature are of too high order to be useful for control or optimisation studies. The development of models with appropriate complexity to capture the key characteristics, while remaining suitable for optimisation studies and real time control, is an existing challenge. It is well known that most pollutants are emitted during the first few minutes of driving, while the catalyst is below its activation or lightoff temperature. Appropriate models could potentially be used to help minimise tail-pipe emissions over the warm up period by optimally scheduling engine control variables such as the air/fuel ratio, spark timing, cam phasing and throttle angle. Voltz et al. (1973) was one of the first to formulate the reaction rates on the surface of a platinum catalyst. His work appears to have formed the basis for almost all future developments in this area of research. Voltz s basic model has since been extended by others with updated

kinetics and more chemical reactions (e.g. Subramaniam and Varma (1985); Dubien et al. (1998); Pontikakis and Stamatelos (24); Holder et al. (26)). One-dimensional TWC models (Heck et al., 1976; Oh and Cavendish, 1982; Groppi et al., 1995; Siemund et al., 1996; Holder et al., 26) predict the distribution of temperature and species concentration along the length of the catalyst as a function of time. Whilst these models contain a number of simplifying assumptions, some of which are the subject of debate, the use of more complex models would render many optimisation studies infeasible. One of the major drawbacks of 1D models is the need for Nusselt number (N u) and Sherwood number (Sh) correlations for calculation of the heat and mass transfer coefficients (Hayes and Kolaczkowski, 1994; Groppi et al., 1995; Wanker et al., 2). Two-dimensional (Hayes and Kolaczkowski, 1994; Groppi et al., 1995; Wanker et al., 2) and three-dimensional (Groppi et al., 1995; Jeong and Kim, 21) models also exist, but require a substantial amount of computation time. A relatively simple one-dimensional TWC model formulation is presented in this paper, the solutions of which are comparable to accurately resolved one-dimensional models. While still incorporating a chemical kinetic scheme, the model is adequate for many optimisation studies, demonstrating reasonable estimates of light-off time and tail-pipe emissions. 2. MODELLING A one-dimensional model is considered in this work. The model is based heavily on previous 1D models and separates the conditions inside the washcoat layer from the bulk fluid conditions inside the channels. 2.1 Key Model Characteristics No heat loss from the monolith to the environment is considered. All properties of the solid phase are constant, except the temperature. A uniform flow distribution is assumed across the face of the monolith block. Nonuniformities are therefore accounted for in the model optimisation which effectively averages across all channels. Reactions occur within the washcoat layer, and heat is generated at the interface between the solid and gas phases. The gas pressure along the length of the monolith is assumed to be at atmospheric pressure. The density of the exhaust is therefore a function of temperature only. Laminar flow is assumed since the hydraulic diameter of monolithic channels is small and the Reynolds number based on the diameter ranges from about 2 to 4. Axial thermal conduction in the solid phase is considered. Axial thermal conduction in the gas phase is ignored as convective heat transfer dominates in the stream-wise direction under mean velocity flow. The heat transfer between the gas and solid phases is governed by forced convection. Mass transfer between the gas and solid phases is governed by diffusion. The diffusion velocities of all species are expected to be much smaller than the average flow velocity and, hence, axial species diffusion in the gas phase is ignored. Table 1. Variables defining a TWC model Var. Description Value S geometric surf. area per reactor vol. 274 m 2 /m 3 α catalytic surf. area per reactor vol. 548 m 2 /m 3 ǫ reactor void fraction.757 ρ s density of the substrate 224 kg/m 3 k s thermal conduct. of the substrate 3. W/m K C P,s specif. heat capacity of the substrate 9 J/kg K Ψ oxygen storage capacity 15 mol/m 3 lw washcoat thickness 17 1 6 m L reactor length.1435 m D h hydraulic diameter of channels.11 m calibratable parameter 2.2 Model Equations The parameters defining the model are given in Table 1. Let h gs and h m,gs,i denote heat and mass transfer coefficients of species i, R c,i the consumption rate of species i, R r,j the rate and h r,j the enthalpy of reaction j in the washcoat for a total of N reactions. Also let M g, C P,g, ρ g and u be the molar mass, specific heat capacity, density and velocity of the exhaust gas respectively. The equations for the fluid temperature T g, the substrate temperature T s, the molar fractions of species i in the flow C g,i and inside the washcoat layer C s,i can be described as ǫρ g C P,g T g t ǫ C g,i t = ǫρ g C P,g u T g x + h gss(t s T g ), (1) C P,s ρ s (1 ǫ) T s = k s (1 ǫ) 2 T s t x 2 + N h gs S (T g T s ) l w α R r,j h r,j, (2) i=j = uǫ C g,i x h m,gs,is (C g,i C s,i ), (3) ρ g S C s,i S = h m,gs,i ρ g (C g,i C s,i ) αm g R c,i. (4) t l w The reactions considered and the corresponding indices j from 1 to 1 are (1) CO +.5O 2 CO 2, (2) H 2 +.5O 2 H 2 O, (3) CH fast 1.8 + 1.45O 2 CO 2 +.9H 2 O, (4) CH1.8 slow + 1.45O 2 CO 2 +.9H 2 O, (5) 2NO + 2CO 2CO 2 + N 2, (6) Ce 2 O 3 +.5O 2 2CeO 2, (7) Ce 2 O 3 + NO 2CeO 2 +.5N 2, (8) 2CeO 2 + CO Ce 2 O 3 + CO 2, (9) CH fast 1.8 + 3.8CeO 2 1.9Ce 2 O 3 + CO 2 +.9H 2 O, (1) CH1.8 slow + 3.8CeO 2 1.9Ce 2 O 3 + CO 2 +.9H 2 O. and have been taken from Pontikakis and Stamatelos (24) along with the corresponding reaction rates R r,j, which incorporate oxygen storage dynamics for some fixed oxygen storage capacity Ψ. The developers of the chemical kinetic scheme have used a catalyst with precious metal loading of roughly 177 g/m 3 and a Pt : Rh ratio of 5 : 1 in their experiments for tuning the reaction rates. However, our washcoat composition (Table 2) was different. The reactions rates R r,j have been taken directly from Pontikakis and Stamatelos (24) and have not been optimised for our catalyst.

Consumption rates R c,i are calculated from these reaction rates. An additional differential equation is included with the reaction scheme, dψ dt = R r,6 + R r,7 Ψ R r,8 + R r,9 + R r,1, (5) Ψ capturing the current fraction of oxygen saturation in the washcoat, ψ, at a particular time instant. Two lumped categories of hydrocarbons are considered: quickly oxidising ( fast HC) and slowly oxidising ( slow HC), which is common practice in many reduced order chemical reaction schemes (Oh and Cavendish, 1982; Zygourakis and Aris, 1983; Dubien et al., 1998). The relative amount of engineout fast and slow HC is difficult to predict or measure. This ratio is therefore fixed. The enthalpies of reaction h r,j are calculated from the enthalpies of formation of relevant species. Cioslowski et al. (2) and Lide (27) provide all of the necessary data. The heat and mass transfer coefficients are obtained from h gs = k g D h Nu Dh, h m,gs,i = D i,n 2 D h Sh Dh,i, where k g is the thermal conductivity of the exhaust gas, and Nu and Sh are Nusselt and Sherwood numbers. The diffusion coefficients D i,j have been obtained from Fuller et al. (1966), and are approximated for a binary mixture of nitrogen and the species in question. Since the exhaust is comprised largely of nitrogen, the thermal conductivity of the exhaust mixture is approximated by k g = 2.269 1 4 T.832 g (6) for nitrogen from Oh and Cavendish (1982). The Nu and Sh correlations proposed by Hawthorn (1974) are used in this work: ( Nu Dh = 3.66 1 +.95 D ).45 h L Pe h,d h, (7) ( Sh Dh,i = 3.66 1 +.95 D ).45 h L Pe m,d h,i, (8) where Pe are Peclet numbers, defined as Pe h,dh = Re Dh Pr, Pe m,dh,i = Re Dh Sc i. The terms Pr, Sc and Re are Prandtl, Schmidt and Reynolds numbers respectively, formulated as Pr = C P,gµ k g, Sc i = µ, Re Dh = ρ gud h, ρ g D i,n2 µ where µ is the dynamic viscosity of the exhaust. 2.3 Boundary Conditions Given that Tg in is the gas temperature and Cg,i in are the molar fractions of species i at the inlet of a TWC, the boundary conditions are T g (, t) = T in g, (9) C g,o2 (, t) = C in g,o 2, (1) C g,co (, t) = C in g,co, (11) C g,h2 (, t) = 1 3 Cin g,co, (12) C g,no (, t) = C in g,no, (13) C g,hcfast (, t) =.85 C in g,hc, (14) C g,hcslow (, t) =.15 Cg,HC in, (15) T g = (16) x x=l C g,i =. (17) x x=l According to Pontikakis and Stamatelos (24) the dynamics of the water gas shift reaction CO + H 2 O CO 2 + H 2 (18) result in the content of H 2 being roughly equal to a third of CO. This approximation was used as H 2 concentrations at the inlet could not be measured with the available lab equipment. The split between fast and slow-oxidising hydrocarbons (14) (15) was also adopted from Pontikakis and Stamatelos (24). 2.4 Solution Scheme The partial differential equations (1) (4) were discretised using central differences. The spatial domain was divided into a uniform grid of resolution x. The variables T g, T s, C g and C s were resolved using 51 ( x = 2.81 mm), 2 ( x = 71.75 mm) and 1 node ( x = 35.875 mm). With such discretisation, equations (1) (4) are reduced to a system of ODE s, ẋ(t) = f(x(t), t), with 15 differential equations associated with each node. The presence of reaction kinetics that are highly sensitive to temperature and reactant concentrations introduces numerical stiffness. A stiff implicit solver, DASSL, developed in Petzold (1982) was chosen for the development of solutions. If the lumped parameter model (1 node) is allowed to span the entire length of the catalyst, such that x = 143.5 mm, the thermal mass of the element becomes so large that the substrate temperature can be affected considerably during warm-up. This in turn can lead to significant errors in the prediction of light-off time, and hence, overall tail-pipe emissions. It has been empirically determined that by constraining x to 1/4 of the brick s length, the performance of the lumped parameter model (1 node) was closest to that of the models with a higher spatial resolution. This was possible as most CO, NO and HC molecules are consumed in front portion of the monolith. The amount of computation time spent on a 1.8 GHz Pentium 4 computer was roughly 2 hours for a 51 node, 3 min 3 sec for a 2 node and 45 sec for a 1 node discretisation scheme over 6 seconds of the NEDC test conditions. Note that whilst the accuracy of the model appears to be reduced only slightly, as the spatial grid resolution is dropped from 51 to 2 nodes, the amount of computational effort is reduced substantially. Further

Table 2. Physical Properties of the TWC Fig. 1. Exhaust system model simplification with the omission of oxygen storage reactions gave a simulation speed improvement of roughly a third. 3. EXPERIMENTAL WORK Experiments were conducted on a 4 litre Ford BF engine and a Horiba-Schenck Titan 46 kw transient dynamometer. The equipment was controlled remotely using ATI Vision and STARS software packages. Experimental data was acquired using the same software. The exhaust system was comprised of a cast iron exhaust manifold, front pipe, an adaptor joining the front pipe to the exhaust manifold and slightly modified inter and rear pipes to suit the test cell s spatial constraints. The front, inter and rear pipes were standard Ford components. The set up is shown in Figure 1 along with locations of various sensors. The TWC was aged on a 75 hour Ford 4-mode schedule, equivalent to 8 km. Its properties are outlined in Table 2. Three equally spaced thermocouples were inserted into the monolith. Two more thermocouples were placed directly before and after the catalyst to measure inlet and exit gas temperatures. A Cambustion HFR4 Fast Flame Ionisation Detector (FFID) measured pre-catalyst and post-catalyst HC emissions. This device has a time constant of less than 2 ms and is specifically designed for taking measurements in the exhaust system where significant pressure fluctuations can exist. Two Autodiagnostics ADS9 gas analysers were employed to measure pre-catalyst O 2, CO, CO 2, NO and HC, and post-catalyst O 2, CO, CO 2 and HC. Tail-pipe NO emissions were not measured. The ADS9 uses nondispersive infrared (NDIR) techniques to measure CO, CO 2 and HC and an electro-chemical cell method to measure O 2 and NO X emissions. It samples at a rate of roughly 1.4 Hz and has a time constant on the order of a few seconds. Normalised air/fuel ratio (λ) measurements were taken using a Bosch LSU4.9 wideband oxygen sensor (UEGO) and the fuel flow rates were measured using an AVL KMA 4 fuel conditioning system. Mass flow rates in the exhaust were then calculated from these two measurements. This paper presents results from two experimental test cases. One is based on the first 6 seconds of the New European Drive Cycle (NEDC) operating conditions (engine speed and torque), logged in a vehicle chassis dynamometer test and implemented on the engine dynamometer (see Figure 2). The engine operated using an existing control strategy, and thus for the majority of the time λ 1. A zoomed in view of λ around light-off is given in Figure 3. Parameter Value Substrate material Cordierite Channel shape Square Number of bricks 1 Length 143.5 mm Channel hydraulic diameter 1.15 mm Thickness of channel walls.165 mm Cross-sectional area.119 m 2 Pt : Pd : Rh composition 12::5 Loading 6 g/m 3 The second test was a custom cycle, performed over 6 seconds from a cold start with a constant pedal position and engine speed fixed at 15 RPM. The torque produced was in the range 7 15 Nm, depending on the λ setpoint. During the first 3 seconds λ 1, while the remainder of the test is identified by a pulsating λ as shown in Figure 4. This alters the feed gas composition and provides the model with an extended validation case. The mass flow rates in this test were up to 5 times higher than in the NEDC test. Engine speed (rev/min) Vehicle speed (km/h) Engine torque (N m) 8 7 6 5 4 3 2 1 1 2 3 4 5 6 7 8 9 18 16 14 12 1 8 6 4 1 2 3 4 5 6 7 8 9 15 1 5-5 1 2 3 4 5 6 7 8 9 Fig. 2. NEDC drive cycle Most of parameters from Table 1 have been calculated from the known dimensions and the knowledge of the substrate material. The variables l w, α and Ψ, however, were difficult to measure directly. Their values were determined in an optimisation procedure using sequential quadratic programming (SQP), minimising a weighted squared difference between the measured and estimated instantaneous tail-pipe emissions over the NEDC test. The parameters l w and Ψ appear to be approximately equal to a third and a quarter respectively of those reported Normalised air/fuel ratio 1.4 1.3 1.2 1.1 1.9.8.7.6 13 14 15 16 17 18 Fig. 3. λ shortly before and after light-off (NEDC)

1.4 1.3.45.4 1.2.35 Normalised air/fuel ratio 1.1 1.9.8.3.25.2.15.1.7.5.6 1 2 3 4 5 6 Fig. 4. λ in the 2nd test in Pontikakis and Stamatelos (24). These values seem reasonable, given that our aged catalyst (Table 2) has a precious metal loading of roughly a third of that of Pontikakis and Stamatelos (24). The parameter α is expected to be a fraction of S as it takes into account catalyst three-dimensionality and the reduction in the active surface area due to ageing. 4. RESULTS A comparison between the measured and the estimated instantaneous tail-pipe emissions is made in Figures 5 and 6. Here the model was discretised using 51 nodes or a grid resolution of x = 2.81 mm. It can be seen that measured and estimated post-catalyst concentrations do not always agree, but most trends are well captured. This is typical behaviour of 1D catalyst models, as Pontikakis and Stamatelos (24) showed. Consider the NEDC test results of Figure 5. Early CO and HC light-off is predicted by the model roughly 85 seconds into the drive cycle. At this instant the model predicts a substrate temperature of 545 K at 1/4 length from the inlet, while the temperature measurement is 56 K for the same position (Figure 7). Modelled and measured temperatures compare well before the predicted light-off. However, actual light-off occurs only when the measured temperature approaches 62 K at 147 seconds. It may be argued that the light-off temperature could have been lower than this, as a rich discharge occurred shortly before light-off, between 143 and 147 seconds, with λ dropping to as low as.8 (see Figure 3), inhibiting possible oxidation reactions. This part of the cycle corresponded to a rapid throttle closing event and poor mixture control. The TWC temperature climbed from 59 K to 62 K. Thus, the true light-off temperature for λ = 1 is expected to lie within this range. The discrepancy between the predicted and the measured light-off temperature for λ = 1 is expected to be due to either catalyst aging, the differing Pt : Rh ratio used in this work and that of Pontikakis and Stamatelos (24) or both. Consider instantaneous emission results from the second test case in Figure 6. The light-off time is observed 43 seconds into the test and is predicted well. By t = 43 s the modelled temperature at 1/4 length position from the inlet was 65 K, coinciding with the measured temperature (Figure 8). This light-off temperature is significantly higher than what was observed in the NEDC cycle. Concentration (ppm C1) 12 1 8 6 4 2 1 2 3 4 5 6 1 2 3 4 5 6 Fig. 5. Measured and modelled tail-pipe CO, HC molar fractions over the NEDC Concentration (ppm C1).6.5.4.3.2.1 1 2 3 4 5 6 1 9 8 7 6 5 4 3 2 1 1 2 3 4 5 6 Fig. 6. Measured and modelled tail-pipe CO, HC molar fractions in the 2nd test

973 873 measured inlet gas temperature measured temperature at 1/4 length modelled temperature at 1/4 length.45.4.35 Temperature (K) 773 673 573 473.3.25.2.15.1 373.5 273 1 2 3 4 5 6 Fig. 7. Substrate temperature over the NEDC 1 2 3 4 5 6 Temperature (K) 973 873 773 673 573 473 measured inlet gas temperature measured temperature at 1/4 length modelled temperature at 1/4 length Concentration (ppm C1) 12 1 8 6 4 2 373 273 1 2 3 4 5 6 Fig. 8. Substrate temperature in the 2nd test High mass flow rates prevail in this test case and at the previously modelled light-off temperature of 545 K the consumption rates are not fast enough to remove the additional inflow of CO and HC molecules. As a result a large proportion of CO and HC pass the front of the monolith unreacted. Substrate temperatures downstream are much lower and CO and HC react much less there. The light-off is thus postponed until higher substrate temperatures, and thus, faster reaction rates are achieved. A better prediction of the light-off time is a consequence of a much quicker substrate temperature rise compared to the NEDC test case. To compare the performances of the catalyst models of various grid resolutions, consider the results from the NEDC test in Figure 9 and the custom cycle in Figure 1. A very good agreement is demonstrated between the solutions using a fine (51 node), and an extremely sparse (2 node) grid. This observation was at first surprising to the authors, but the distribution of consumption rates R c,i [mol/(m 3 s)] in the 2 node model under many conditions appear to sensibly approximate those in the 51 node model. The lumped parameter model performs notably worse, but most trends are still followed. Figures 11 and 12 demonstrate measured and simulated, using various number of nodes, cumulative emissions over the two test cycles. It can be seen that cumulative emissions compare much better than instantaneous emissions from Figures 5 and 6 for all model discretisation schemes. The effect of removing all oxygen storage reactions from the chemical kinetic scheme on instantaneous post-catalyst 1 2 3 4 5 6 Fig. 9. Effect of grid resolution on the modelled tail-pipe CO, HC molar fractions over the NEDC emissions is demonstrated in Figure 13 for the NEDC test case (modelled with 51 nodes). Comparisons of cumulative emissions are shown in Figure 14. It can be seen that the conversion efficiency of CO is almost unaffected, while that of HC is considerably degraded. This is because under this kinetic scheme and test conditions, almost all CO and fastoxidising HC are oxidised via reaction 1 and 3 mechanisms respectively, while oxidation via reaction 8 and 9 is limited. However, slow-oxidising HC are primarily oxidised by reaction 1, and the removal of oxygen storage reactions 6 1 heavily restricts their oxidation in simulations. Similar observations have been reported (Shamim et al., 22; Shamim, 28) with other kinetic schemes. 5. CONCLUSION One of the major requirements for solving dynamic optimisation problems are models with the minimum number of state variables. A representation of a TWC model with a small ODE system was identified in this paper, making such optimisation studies possible. This paper demonstrated that with reasonable accuracy a one-dimensional TWC model may model the light-off time and tail-pipe emissions. The performance of the models with 2 and 51 nodes is very similar, whilst the results of simulations using the lumped parameter model are comparable. A simplification of the chemical kinetic scheme via the removal of oxygen storage reactions leads to an underestimation of HC conversion efficiency. However, the light-off time can still be adequately predicted.

.6.5.4.3.2 Mass (g) 16 14 12 1 8 6 4.1 2 1 2 3 4 5 6 1 2 3 4 5 6 Concentration (ppm C1) 1 9 8 7 6 5 4 3 2 Mass (g C) 16 14 12 1 8 6 4 1 2 1 2 3 4 5 6 1 2 3 4 5 6 Fig. 1. Effect of grid resolution on the modelled tail-pipe CO, HC molar fractions in the 2nd test Fig. 12. Measured and modelled tail-pipe CO, HC cumulative emissions in the 2nd test 45 4 35.45.4.35 modelled at outlet with o.s. reactions modelled at outlet without o.s. reactions Mass (g) 3 25 2 15.3.25.2.15 1.1 5.5 1 2 3 4 5 6 1 2 3 4 5 6 9 8 7.12.1 modelled at outlet with o.s. reactions modelled at outlet without o.s. reactions Mass (g C) 6 5 4 3.8.6.4 2 1.2 1 2 3 4 5 6 Fig. 11. Measured and modelled tail-pipe CO, HC cumulative emissions over the NEDC 1 2 3 4 5 6 Fig. 13. Effect of oxygen storage reactions on post-catalyst CO, HC molar fractions over the NEDC

Mass (g) Mass (g) 45 4 35 3 25 2 15 1 5 modelled at outlet with o.s. reactions modelled at outlet without o.s. reactions 1 2 3 4 5 6 9 8 7 6 5 4 3 2 1 modelled at outlet with o.s. reactions modelled at outlet without o.s. reactions 1 2 3 4 5 6 Fig. 14. Effect of oxygen storage reactions on post-catalyst CO, HC cumulative emissions over the NEDC ACKNOWLEDGEMENTS This research was supported by the Advanced Centre for Automotive Research and Testing (ACART, http://www.acart.com.au). REFERENCES Cioslowski, J., Schimeczek, M., Liu, G., and Stoyanov, V. (2). A set of standard enthalpies of formation for benchmarking, calibration, and parametrization of electronic structure methods. Journal of Chemical Physics, 113(21), 9377 9389. Dubien, C., Schweich, D., Mabilon, G., Martin, B., and Prigent, M. (1998). Three-way catalytic converter modelling: fast- and slow-oxidizing hydrocarbons, inhibiting species, and steam-reforming reaction. Chemical Engineering Science, 53(3), 471 481. Fuller, E., Schettler, P., and Giddings, J. (1966). A new method for prediction of binary gas-phase diffusion coefficients. Industrial and Engineering Chemistry, 58(5). Groppi, G., Belloli, A., Tronconi, E., and Forzatti, P. (1995). A comparison of lumped and distributed models of monolith catalytic combustors. Chemical Engineering Science, 5(17), 275 2715. Hawthorn, R. (1974). Afterburner catalysts effects of heat and mass transfer between gas and catalyst surface. In American Institute of Chemical Engineers Symposium Series, volume 7, 428 438. Hayes, R. and Kolaczkowski, S. (1994). Mass and heat transfer effects in catalytic monolith reactors. Chemical Engineering Science, 49(21), 3587 3599. Heck, R., Wei, J., and Katzer, J. (1976). Mathematical modeling of monolithic catalysts. AIChE Journal, 22(3). Holder, R., Bollig, M., Anderson, D., and Hochmuth, J. (26). A discussion on transport phenomena and three-way kinetics of monolithic converters. Chemical Engineering Science, 61, 81 827. Jeong, S. and Kim, W. (21). A new strategy for improving the warm-up performance of a light-off autocatalyst for reducing cold-start emissions. Proceedings Institution of Mechanical Engineers, 215, 1179 1196. Lide, D. (27). CRC handbook of chemistry and physics. CRC Press. Internet version: http://www.hbcpnetbase.com/. Oh, S. and Cavendish, J. (1982). Transients of monolithic catalytic converters: response to step changes in feedstream temperature as related to controlling automobile emissions. Industrial & Engineering Chemistry Product Research and Development, 21(1), 29 37. Petzold, L. (1982). A description of DASSL: a differential/algebraic system solver. IMACS World Congress. Pontikakis, G. and Stamatelos, A. (24). Identification of catalytic converter kinetic model using a genetic algorithm approach. Proceedings Institution of Mechanical Engineers, 218, 1455 1472. Seinfeld, J. (24). Air pollution: a half century progress. Environmental and Energy Engineering, 5(6), 196 118. Shamim, T. (28). The Effect of Oxygen Storage Capacity on the Dynamic Characteristics of an Automotive Catalytic Converter. Energy Conversion and Management, 49, 3292 33. Shamim, T., Shen, H., Sengupta, S., Son, S., and Adamczyk, A. (22). A Comprehensive Model to Predict Three-Way Catalytic Converter Performance. Journal of Engineering for Gas Turbines and Power, 124, 421 428. Siemund, S., Leclerc, J., Schweich, D., Prigent, M., and Castagna, F. (1996). Three-way monolithic converter: simulations versus experiments. Chemical Engineering Science, 51(15), 379 372. Subramaniam, B. and Varma, A. (1985). Reaction kinetics on a commercial three-way catalyst: the CO NO O 2 H 2 O system. Industrial & Engineering Chemistry Product Research and Development, 24(4), 512 516. Twigg, M. (27). Progress and future challenges in controlling automotive exhaust gas emissions. Applied Catalysis B: Environmental, 7, 2 15. Voltz, S., Morgan, C., Liederman, D., and Jacob, S. (1973). Kinetic study of carbon monoxide and propylene oxidation on platinum catalysts. Industrial & Engineering Chemistry Product Research and Development, 12(4). Wanker, R., Raupenstrauch, H., and Staudinger, G. (2). A fully distributed model for the simulation of a catalytic combustor. Chemical Engineering Science, 55, 479 4718. Zygourakis, K. and Aris, R. (1983). Multiple oxidation reactions and diffusion in the catalytic layer of monolith reactors. Chemical Engineering Science, 38(5), 733 744.