Vous êtes sur la page 1sur 13

2003-01-1060

A New Quasi-Three Dimensional Combustion Model for Prediction of DI Diesel Engines Performance and Pollutant Emissions
E. G. Pariotis and D. T. Hountalas
Mechanical Engineering Department, National Technical University of Athens
Copyright 2003 Society of Automotive Engineers, Inc.

ABSTRACT
The fundamental understanding of mixture formation and combustion process taking place in a DI diesel engine cylinder is an important parameter for engine design since they affect engine performance and pollutant emissions. Multi-dimensional CFD models are used for detailed simulation of these processes, but suffer from complexity and require significant computational time. The purpose of our work is to develop a new quasi-dimensional 3D combustion model capable of describing the air fuel mixing, combustion and pollutant formation mechanisms, on an engine cycle by cycle basis, needing reasonably low computational time compared to CFD ones, while describing in a more fundamental way the various processes compared to existing multi-zone phenomenological models. As a result, a number of problems associated with the application of multi-zone models are resolved. The combustion chamber is divided into a number of computational cells and the equations of mass, energy and species conservation are solved to calculate temperature and species concentration at each node. The gas flow field is estimated using a newly developed semi-empirical gas motion model, based on the assumption that the in-cylinder pressure is uniform. The finite volume method is used for the solution of the conservation equations. The implicit temporal and hybrid central upwind spatial differencing scheme is used for the discretization of the conservation equations. Spray trajectory, fuel vaporization and combustion are simulated using simplified sub-models based on semiempirical correlations. A first application of this new model is made on a high speed DI diesel engine. The effect of operating conditions on the combustion and pollutant formation mechanism is examined. Information concerning local air fuel ratio, temperature distribution and species concentrations is derived. To validate the model experiments have been conducted at the authors' laboratory on a DI diesel engine. Comparing experimental with calculated results, a relatively good agreement is observed. This reveals that the proposed model produces qualitatively reliable predictions of the in-cylinder processes and engine performance at

reasonable computational time. Furthermore it appears that it can be used to study pollutant formation and the effect of various engine design parameters on the combustion mechanism, providing results that are a compromise between phenomenological models and detailed CFD ones.

INTRODUCTION
Diesel engines have been widely used extensively in heavy-duty vehicles while lately are increasingly being used for light-duty vehicles. The attractiveness of diesel engine lies on its high efficiency. On the other hand its main disadvantages compared to spark-ignition engines are lower power density and higher particulate pollutant emission levels. To increase power density, turbocharging systems can be used, while for the reduction of pollutant emissions various strategies have been proposed, focusing on the improvement of the combustion mechanism or the use of after-treatment techniques. With the anticipated extreme tightening of NO and PM emission standards by year 2010, an intensive research is taking place nowadays aiming to reduce diesel engine emissions without affecting seriously its bsfc. Some of the main strategies that have been proposed for the optimization of the combustion mechanism are improved combustion chamber design, use of advanced high-pressure fuel injection systems, and use of EGR [16]. The determination of the optimum strategy is very difficult due to the complexity of the combustion mechanism, and the contribution of computer modeling to the fundamental understanding of these processes is essential. Diesel engine computer simulation models are widely used to predict the effect of various parameters on engine performance and emissions. In this way a considerable amount of time and effort is saved by reducing the number of experiment, while it is possible to focus on the effect of individual parameters, which is difficult using experimental techniques. There exist various computer models and the final choice depends

on the purpose of the simulation and the level of accuracy required [7]. In the present work a new quasithree dimensional combustion model for the prediction of DI diesel engines performance and pollutant emissions has been developed. The main scope of this model is to describe the fuel air mixing mechanism and the combustion process in a more fundamental way compared to existing multi-zone phenomenological models, while being less time consuming and complicated compared to existing CFD models. To achieve this, some of the basic features of the sophisticated multi-zone phenomenological models and the multidimensional CFD ones have been combined. The cylinder is divided in finite volumes (cells) where the mass, energy and species conservation equations are numerically solved similarly to CFD models, while the flow field is estimated from a newly developed phenomenological model to avoid the solution of the momentum equation. On the other hand the fuel injection, spray trajectory, ignition delay and pollutant emissions (NO and Soot) are simulated using existing semi-empirical phenomenological models that are properly modified. As a result the spatial distribution of temperature, species concentration and pollutant emissions is derived at each crank angle. To validate the model, experiments have been conducted at the authors laboratory on a high speed DI diesel engine. Comparing experimental with calculated results at a wide range of engines speeds (1500-2500 rpm) and at engine loads ranging from 20% to 80% of full engine load, it is revealed that the model predicts adequately engine performance and with a reasonable good accuracy the exhaust pollutant emissions (NO and Soot). This is quite encouraging considering the complexity of the proposed model and the fact that it has been developed from scratch.

resulting heat absorption are computed in each spray package and treated as source terms in the conservation equations at the computational cells located at each spray package neighborhood. The combustion model computes the ignition delay of the mixture in each computational cell and the combustion products are treated as source terms in the species conservation equation. As a result the spatial distribution of enthalpy, fuel vapor, oxygen, combustion products and temperature is obtained at each crank angle, and the mean cylinder pressure, the heat release rate and pollutant emissions (NO, Soot) can be calculated. The model simulates only the closed part of the engine cycle, using initial conditions estimated at inlet valve closure.

COMPUTATIONAL DOMAIN
The engine treated in this study has a bowl in piston and a three-hole injector centered at the cylinder bore. Due to symmetry, the computational domain is restricted to one third of the cylinder volume Fig.1, to save computational time. The area inside the cylinder is divided into cylindrical computational cells as shown in Fig.2a.
r direction

z direction

OUTLINE OF COMPUTATIONAL MODEL


A quasi-three dimensional combustion model has been developed for four-stroke direct injection diesel engines. The cylinder is divided in finite volumes where the conservation equations for the fuel vapor, combustion products and mixture enthalpy are numerically solved by the finite difference technique. The flow field is estimated using a revised method proposed by the authors in the past [8], assuming that the in-cylinder pressure is almost uniform which is very close to reality [9]. In this way a three-dimensional vector of the velocity is estimated at each computational cell, avoiding the solution of the momentum equation, as it happens in a pure CFD code [10-13]. Although the velocity field obtained by this method is only a rough estimation, it is more realistic than the corresponding one used by existing phenomenological models. In this way the effect of combustion chamber geometry and operating conditions is considered at least qualitatively. As far as the spray model is concerned, the spray consists of many packages, each having its own velocity, which is computed using semi-empirical equations. The rates of fuel vapor generation and the
direction Cylinder walls Injection Direction

r direction Piston cavity

-r plane view

Fig.1 Computational domain

The number of cells in the r and direction is constant, in contrast to the number of cells outside the piston bowl in the z direction, which is variable depending on piston position. The number of cells inside the piston bowl in

r S

1 zpiston 1 ( u r ) 1 ( v) ( w) + + + t z zpiston r r r r r 1 z r 1 + S + + = z r r r

(2)

E z T
Fig.2a. Computational cell and coordinates

In this study the combustion products are defined considering dissociation, using the chemical equilibrium scheme proposed by Vickland et al. Thus in general at each computational cell, apart from fuel vapor and pollutant emissions (NO, Soot), the following eleven species are considered to be present in chemical equilibrium: O2, N2, CO2, H2O, H, H2, N, NO, O, OH, CO The velocities u, v, w are computed by the simplified gas motion model which will be described later. The diffusivity is defined by the following equation (4) and is assumed to be the same for all the dependent variables .
= cp

direction z is also constant. The grid follows the piston motion contracting and expanding [10,11,14]. The axial grid velocity inside the piston bowl is equal to the piston velocity, in contrast to the axial velocity of the grid in the area between the top of the piston and the cylinder head, which is obtained from EQ(1). In the radial direction the cylinder is divided into an annular and inner volume. Thus the number of nodes, referring to the r direction, in the annular volume may differ from the corresponding one in the inner volume.

(3)

w grid

w piston = z w piston z piston

if z > z piston

(1)
if z z piston

As far as the volumetric source rate S is concerned when the dependent variable is the enthalpy (=h) is defined as follows: Heat transfer Power due to rate through the + pressure = cylinder boundaries variation Heat Absorption by + Fuel Vaporization S = h = S heat + S pressure + S vaporization The heat transfer between cylinder walls and gas computational cells is defined by:

In the present study the grid used for all cases examined has in the r-direction, ten cells inside the piston bowl and ten cells in the outer volume. In the axial direction, there are ten cells inside the piston bowl and the number of cells between the piston and the cylinder head range from twelve to five, according to the piston position. Finally in the angular direction there are fourteen cells. The present resolution is found to give adequately gridindependent results, while it should be noticed that the results are found to be more sensitive to the grid resolution in the angular direction, which is consistent to what has been observed by other researchers [15]. Moreover reducing the number of cells in the axial direction during piston motion towards TDC a more uniform grid is obtained, since the computational cells in the piston bowl retain their shape in contrast to the ones outside the bowl which contract and expand [14,15].

S = h

(4)

CONSERVATION EQUATION
The conservation equation for the species concentration and gas enthalpy is described in terms of the cylindrical coordinates, which expand and contract following the piston motion EQ(2).

(T Tcell ) 4 4 S heat = A h wall + c Twall Tcell Vcell

(5)

The convection heat transfer coefficient is obtained from the following correlation:
h = c1 Re c2 Pr c3 k l char

(6) (7)

where: Re = w char

lchar

Pr =

cp k

(8)

scheme is only first-order in time, which is the reason for selecting a rather small time step. In this study the time increment is equivalent to 0.5 degree crank angle. Discretising the energy equation results to a system of linear algebraic equations that are solved by the tridiagonal matrix algorithm (TDMA), which is applied iteratively, in a line-by-line fashion. In this study the TopBottom sweep direction has been applied for the line-byline solution of the system.

and c1, c2, c3, c: constants. In this study c1=0.30, c2=0.80, c3=0.33, c=5.668E-8. In the Reynolds number expression, lchar is the characteristic length and wchar is the characteristic velocity. As far as the energy source rate due to pressure changes is concerned, it is defined as:
S pressure = P t

(9)

SIMPLIFIED GAS MOTION MODEL


The definition of the velocity field inside the cylinder is very important given that in the conservation equations the convection term depends on the magnitude and direction of the local velocity field. Multidimensional models use the momentum equation to calculate the velocity and pressure at each computational cell. However the solution of the momentum equation needs special treatment given that the convective terms contain non-linear quantities, and that the momentum equation (containing the local pressure) and the continuity equation are intricately coupled. Although there have been proposed many computational fluid dynamic methods for the calculation of the velocity field inside the cylinder of an internal combustion engine [1113,16], they all have in common that they are complicated and extremely time consuming. On the other hand they appear to be quite accurate. In this work we estimate the velocity field using a revised version of a simplified gas motion model developed by the authors [8]. The advantage of this method is that it is simple and describes the physical phenomenon of mass transfer between the computational cells in a realistic way, based on the assumption that the in-cylinder pressure must be practically uniform at each crank angle. The main improvement that has been made to the existing gas motion model is that at present is able to handle three-dimensional cases. The velocity is estimated at the boundaries of each computational cell, by calculating the mass that should be transferred through each boundary to the neighboring cell to achieve a uniform pressure field inside the cylinder. The velocity field is obtained following an iterative procedure. First, at each crank angle the energy conservation equation is solved, and a spatial distribution of the temperature is obtained assuming that the velocity of the gas relative to the grid is zero. Then the pressure distribution is obtained using the perfect gas state equation. Given that the pressure must practically be uniform, an amount of mass dmcell should be transferred to each computational cell through its boundaries from the neighboring cells to eliminate the pressure difference and make its pressure practically equal to the mean pressure of the cylinder. The required mass is defined from, dm cell = Pmean Vcell m cell R Tcell

It is assumed that the gas is ideal which means that the pressure at each computational cell can be defined by,

P = R T

(10)

The density of the gas at each computational cell is calculated from the gas motion model described later on. The source term Svaporization represents the energy that is absorbed by each computational cell located in the region where fuel vaporization occurs. It is calculated by weighting the distance between each fuel vapor package and the computational cells that exist in its neighborhood. Solving EQ(2), using the finite volume technique at each computational cell, the spatial distribution of enthalpy is obtained and using the specific heat of the gaseous components the local temperature is calculated at each crank angle. Consequently the local pressures can be computed from the perfect gas state equation EQ(10). The calculation of the spatial distribution of pressure is essential for the gas motion model, as explained later on. When the dependent variable is the fuel vapor concentration then in order to calculate the volumetric source rate S, the rate of fuel vapor generation calculated at each spray package is distributed to the computational cells that exist in its neighborhood similarly to what is done for the calculation of Svaporization. In this way a coupling between the mesh generated to solve the conservation equations and the fuel spray geometry is achieved.

COMPUTATIONAL METHOD
The conservation equations are solved using the fully implicit finite volume method. For the spatial differencing the hybrid-differencing scheme is used due to its ability to exploit the advantages of the upwind and the central differencing scheme. It switches to the upwind differencing when the central differencing produces inaccurate results at high Peclet numbers. The scheme is fully conservative and since the coefficients are always positive it is unconditionally bounded [16]. As far as the temporal differencing is concerned the fully implicit method is used due to its unconditionally stable behavior for any time step. However, the accuracy of the

(11)

The total transferred mass dmcell for all computational cells at each crank angle should be equal to zero, given that the total mass of the gas inside the cylinder is constant. Thus the mean cylinder pressure can be estimated from,

m
Pmean =
i, j,k

cell

Vcell T i, j, k cell

(12)

procedure where various assumptions are usually made [17]. The primary problem in analyzing fuel spray trajectory and dispersion lies in treating the coupling of mass, momentum and energy between the two phases (liquid and vapor) of the fuel and the gas mixture of the cylinder [18]. To overcome these difficulties and due to the different nature of the proposed model we describe the spray trajectory using semi empirical relations, as in most multi-zone phenomenological models [19-23]. As the fuel spray penetrates inside the combustion chamber, it is divided into packages (called zones) as follows [24]: in the injection direction it is divided at each time step, while in the radial direction it is divided into three packages (M-direction) and in the circumferential direction (N-direction) into eight packages (Fig.2b). Every group of new zones that enter the cylinder at each time step is called a parcel. The mass of fuel in each parcel entering the cylinder can be either specified as input, or calculated using a fuel injection sub-model. The fuel mass of a parcel is distributed evenly to each zone.
Frame 001 31 Aug 2002 | |

In this way the amount of mass dmcell transferred to each computational cell is estimated. This amount of mass is taken from the neighboring cells. The amount of mass transferred from the neighboring cells depends on the local pressure differences. For example if the examined computational cell requires an amount of mass to be added in order to increase its pressure and make it equal to the mean cylinder one, this mass has to be taken from each of its neighboring cells through their common boundary. The proportion of the total mass dmcell that each of the neighboring cells will offer is a function of its own pressure relative to the calculated mean cylinder pressure. This means that if a neighboring cell has a pressure that exceeds the mean cylinder pressure i.e DP>0 then its contribution to the total mass needed will be directly proportional to DP, otherwise this cell will not submit any mass. The opposite occurs if the examined computational cell needs to expel mass i.e. dmcell <0. Sweeping all the computational cells, the gas velocity at the boundaries of each cell is calculated and the same procedure is followed at the next iteration. At each iteration the total mass transferred to a computational cell from the beginning is computed from: dm total,cell = dm total,cell + dm cell

Fig.2b Spray division into zones (N=8, M=3) at a certain instance.

(13)
Each parcel entering the cylinder travels a distance at constant speed, equal to the fuel injection speed, called the break-up length. Fuel vaporization initiates after this time has passed. After break-up the injected fuel is distributed within a spray angle, which is unique for each spray parcel and is calculated using empirical correlations. In the present study the Hiroyasu and Arai [25] correlations are used as modified by Assanis et al. [26] to be applicable in cases where the nozzles discharge coefficient is not equal to 0.39 as in the original study. According to these correlations the spray penetration for each zone before and after breakup time is calculated as follows:
2P S(M ) = c D l
0.5

With the new velocity field the mass conservation equation is solved for each computational cell to determine the local densities: cell = m cell,old + dm total,cell Vcell

(14)

Then the energy conservation equation is solved, to determine a new temperature field. The final solution of the velocity and the temperature field is achieved when the temperature field has converged.

SPRAY MODEL
SPRAY TRAJECTORY AND DISPERSION In multi-dimensional models the spray trajectory is calculated solving the momentum equation in the computational domain. This is a very complicated

, 0 < t < t b (M)

(15)

P S(M ) = 2.95

0.25

d 0.5 (t + t b (1) t b (M ))0.5 , t b (M) t n

(16)

FUEL VAPORIZATION After break-up a group of drops is generated in each zone. In this study the size of the drops is assumed to be uniform and equal to the Sauter Mean Diameter given by the following empirical correlation [24]. D 32 = 23.9(P )0.135 ( )0.121 (Vf )0.131

where P is the pressure drop across the nozzle hole, l and is the density of the liquid fuel and the ambient gas respectively, and dn the nozzle hole diameter. The velocity of each zone V(M) is derived from the temporal differentiation of EQ(15) and EQ(16). Since the break-up time of the zones at the edge of the spray is shorter than the inner zones, the radial variation of the break-up time in the spray is incorporated as follows: t b (M ) = 4.351

(22)

c2 D

( P )

ld n

0.5

NM + 1 M NM

(17)

where Vf is the volume of fuel injected per fuel injection in (mm3). The number of droplets N within each zone is calculated from the following expression (assuming that the droplets are spherical): N zone = 6 m l,zone D 32 l

(23)

where NM is the number of zones in the radial direction of the spray, which in the present study is NM= 3. The direction of each parcel (M) is computed from the empirical relation of the spray angle [24] as follows: Pd 2 (M ) = 0.05 2 n
0.25

M NM

(18)

where ml,zone is the zone fuel mass. For evaporation the model of Borman and Johnson [27] is followed. After break-up the fuel evaporation rate (dmfg/dt) and the heat supplied from the hot ambient gas to the liquid fuel droplets for vaporization is calculated using the following expressions: dm fg dt = dm l dt

In this study fuel spray impingement on the cylinder walls is considered, according to the Hiroyasu et al. model [23]. After impingement on the cylinder walls the zones are assumed to penetrate radially from the impingement point along the wall, retaining their width. Given that no zone mixing is allowed, if a zone near the edge of the spray has not yet reached the cylinder walls, but reaches inner zones that have already impinged and changed their direction, then this zone follows the radial direction too. Spray wall impingement affects the penetrating velocity of each zone, which is given by the following correlation: P S(M ) = 2.95 INJECTION RATE To estimate the injection rate of fuel and the initial conditions at the nozzle exit, it is assumed that the flow through the nozzle hole is quasi-steady, one dimensional and incompressible so the fuel mass flow rate is calculated by the following equation [7]: & m = c D A n 2 l P
0.25

(24) (25) (26)

dm l = D l gO D fO ln(1 + M )Sh dt dD l 2 = dt D l2 l dm l D 3 d l dTl l dt 6 dTl dt + ) ln(1 T ) Nu


T

q = D l K gO Tg Tl dTl 1 = dt m l C pl

(27) (28)

d 0.5 (t + t b (1) t b (M ))0.25 n

(19)

dm l q + L dt

The mean temperature at the interface between the gas and liquid phase of the droplet, is estimated by the following expression: Tm,int erface = Tg + 2Tl 3

(29)

(20)
COMBUSTION MODEL
IGNITION In each computational cell combustion initiates after an ignition delay period tign calculated from the local pressure and temperature using the following relation [28]:

where An is the nozzle hole area and P the local pressure difference. The fuel injection velocity at the nozzle tip is obtained from: u = cD 2P l

(21)

ign

1 c ign P
1.19

(4650 e

Tg

) =1

(30)

where P is the mean cylinder pressure and cign is a constant. After ignition, evaporated fuel (which in this study is assumed to be normal dodecane) and air react at a rate give by the following Arrehnious type equation [7]:
m n R fv = A 2 Yfv Yox exp( mix

where b=[NO]/[NO]e and Vcell is the volume of each computational cell. In the previous relations, index e denotes equilibrium. Integrating the previous differential equation, we obtain the NO concentration in each computational cell. Soot Formation and Oxidation As far as the soot formation is concerned, the semiempirical two-rate equation model proposed by Hiroyasu et al [24] is used. Although the detailed mechanism of soot formation in internal combustion engines is not fully understood, it is well established that the net soot formation is primarily affected by pressure, temperature and equivalence ratio. Based on this model the net soot formation rate is given by: dm s dm sf dm sb = dt dt dt

E ) Tg

(31)

where mix is the local density of the mixture, A is the frequency factor, E the reduced activation energy, m, n constants, and Yfv, Yox are the local fuel vapor and oxygen mass concentrations respectively. POLLUTANT EMISSION FORMATION Nitric Oxide Formation In order to calculate the formation of nitric oxide inside each computational cell the chemical equilibrium scheme proposed by Vicland et al. [29] is used to calculate the concentration of each of the following eleven species. O2, N2, CO2, H2O, H, H2, N, NO, O, OH, CO The formation of nitric oxide is controlled by chemical kinetics [7]. In the present work, the extended Zeldovich mechanism is used to calculate the NO concentration at each computational cell. According to the extended Zeldovich mechanism the principal reactions that govern the formation of NO are: N + NO N 2 + O k 1f = 1.6x1010

(35)

where dmsf/dt, dmsb/dt are the soot formation and oxidation rates respectively defined by the following equations: dm sf = A f m f ,ev P 0.5 exp[ E sf / (R mol T )] dt PO dm sb = A bms 2 P dt 1.8 P exp[ E sb / (R mol T )]

(36) (37)

where mev is the mass of evaporated fuel inside the computational cell and PO2 is the partial pressure of oxygen. The activation energies of soot formation and combustion Esf and Esb, are 1.25x104 kcal/kmol and 1.40x104 kcal/kmol respectively while Af and Ab are constants.

N + O 2 NO + O k 2f = 6.4 x10 6 T exp( 3125 / T ) N + OH NO + H k 3f = 4.2 x1010 The variation of [NO] concentration computational cell is expressed as follows: 2 1 R1 1 d([NO]Vcell ) = Vcell dt R1 1 + R2 + R3
2

(32)
RESULTS AND DISCUSSION

in

each

TEST ENGINE SPECIFICATIONS To validate the model, it has been applied on a naturally aspirated air-cooled four stroke DI Diesel engine, with a bowl-in-piston combustion chamber and a three-hole injector nozzle located at the center of the cylinder bore. The main engine specifications are given in Table 1.

(33)

where

MODEL CALIBRATION

R 1 = k 1f [N ]e [NO]e R 2 = k 2 f [N ]e [O 2 ]e R 3 = k 3f [N ]e [OH ]e (34)

A single operating point, which lies in the middle of the engine speed range (2000 rpm) and at 80% of full engine load, has been selected as the reference point for model calibration. Once the constants of the model

80 70

Cylinder Pressure 2000 rpm, 80% Load Calculated Experimental

Table 1. Engine Specifications Type Single Cylinder, 4-Stroke,DI Bore 85.73mm Stroke 82.55mm Connecting Rod 148.59mm Length Compression Ratio 18 Inlet Valve Opening 15oCA before TDC Inlet Valve Closure 41oCA after BDC Exhaust Valve 41oCA before Opening BDC Exhaust Valve 15oCA after TDC Closure

Cylinder Pressure (Bar)

60 50 40 30 20 10 0

40 60 80 100 120 140 160 180 200 220 240 260 280

Crank Angle (degree)

Fig. 3b. Comparison of predicted and measured cylinder pressure traces at 2000 rpm engine speed, 80% load.

Cylinder Pressure (Bar)

are determined, they retain their value during the entire range of the operating conditions examined. Model calibration is based on the proper estimation of the cylinder pressure diagram and the prediction of the ignition delay period and the Soot tailpipe value. MODEL VALIDATION Performance A wide range of operating conditions is selected to validate the models ability to predict engine performance. In Figs. 3a-d the predicted cylinder pressure traces are compared with the experimental ones at 2000 and 2500 rpm engine speed only for the sake of space, for 40% and 80% of full engine load.

80 70 60 50 40 30 20 10 0

Cylinder Pressure 2500 rpm, 40% Load Calculated Experimental

40 60 80 100 120 140 160 180 200 220 240 260 280

Crank Angle (degree)

Fig. 3c. Comparison of predicted and measured cylinder pressure traces at 2500 rpm engine speed, 40% load.

80 70

Cylinder Pressure 2000 rpm, 40% Load Calculated

80 70

Cylinder Pressure 2500 rpm, 80% Load Calculated Experimental

Cylinder Pressure (Bar)

50 40 30 20 10 0 40 60 80 100 120 140 160 180 200 220 240 260 280

Cylinder Pressure (Bar)

60

Experimental

60 50 40 30 20 10 0

Crank Angle (degree)

40 60 80 100 120 140 160 180 200 220 240 260 280

Crank Angle (degree)

Fig. 3a. Comparison of predicted and measured cylinder pressure traces at 2000 rpm engine speed, 40% load.

Fig. 3d. Comparison of predicted and measured cylinder pressure traces at 2500 rpm engine speed, 80% load.

At all operating conditions examined a good agreement between calculated and experimental values is observed during the whole part of the closed engine cycle, which means that the model is able to describe the effect of engine speed and load on engine performance. The previous finding is enhanced comparing the predicted and measured values of peak cylinder pressure for all cases examined i.e. 1500, 2000 and 2500 rpm engine speed and 20% to 80% engine load (Fig. 4). The maximum cylinder pressure is well-predicted at all operating conditions. The highest differences between calculated and measured values are observed at part load, being always lower than 5% of the maximum cylinder pressure.
Calculated Max. Cyl. Pressure
80 70

70 60 50 40 30 20 10 0

Heat Release Rate 2000 rpm, 40% Load Calculated Apparent HRR Experimental Net HRR

Heat Release Rate (Joule/deg)

160 170 180 190 200 210 220 230 240 250 260

Crank Angle (degree)


Experimental Max. Cyl. Pressure

1500 rpm

Fig. 5a. Comparison of heat release rates at 2000 rpm engine speed and 40% load.

Max. Cylinder Pressure (Bar)

60 50 80 70 60 50 80 70 60 50

70

Heat Release Rate 2000 rpm, 80% Load Calculated Apparent HRR Experimental Net HRR

Heat Release Rate (Joule/deg)

2000 rpm

60 50 40 30 20 10 0

2500 rpm

Engine Load (% of full Engine Load)

20

40

60

80

Fig. 4. Comparison of predicted and measured maximum cylinder pressure at 1500, 2000, 2500 rpm engine speed and 20%, 40%, 60%, 80% load.

160 170 180 190 200 210 220 230 240 250 260

Crank Angle (degree)

To validate the models ability to describe the combustion mechanism, the predicted heat release rates for 2000 rpm engine speed and at 40% and 80% engine load are compared with the corresponding experimental ones (Fig. 5a,b). It should be stated that the heat release rate derived from the measured cylinder pressure is the net one, whereas the calculated one is the apparent gross, thus only a qualitative comparison is made between them. Examining Figs. 5a,b it is obvious that the model predicts adequately at both engine loads examined, the initiation of combustion and moreover the duration of the premixed and mixing controlled phases. The highest differences between measured and predicted heat release rates occur during the premixed combustion phase, due to the high heat transfer rate through the cylinder walls (since the calculated one is the gross while the experimental is the net one). The ability of the model to predict the cylinder pressure and heat release rate at various operating conditions shown above, is a strong indicator that the various sub-models describe adequately the fuel spray trajectory, fuel evaporation, air-fuel mixing and combustion mechanisms.

Fig. 5b. Comparison of heat release rates at 2000 rpm engine speed and 80% load.

Pollutant Emissions As far as the pollutant emissions are concerned, in Figs 6a-c the calculated concentrations of NO and Soot emissions at the engine exhaust are compared with the measured ones, at all test cases examined. The effect of engine speed and load on pollutant emissions is well predicted and the predicted values of emission (NO, Soot) concentrations are close to the measured ones. A better agreement between calculated and measured values is observed as far as the NO and Soot concentrations is concerned, at 2000 rpm which is the engine speed selected for model calibration, while at 2500 rpm occur the highest differences. In general it seems that, the NO emission model gives more accurate predictions compared to the soot emission model. However taking into account that no adjustment of the calibration constants is made with the variation of engine speed and load, the calculated results are promising and indicate that the pollutant formation

400 350

1500 rpm Engine Speed Calculated Experimental

Soot (mgr/m3)

300 250 200 150 100 50 0 1700 1500

mechanism is simulated reliably, since trends are predicted with good accuracy. Spatial Distribution of Temperature Detailed information on the pollutant formation and combustion mechanism can be derived from the spatial distribution of temperature inside the combustion chamber, predicted by the proposed model. The predicted temperature distribution is presented for the case of 2000 rpm engine speed and 80% load, at certain time instants after injection. Even though no experimental data are available to validate the predicted distributions, qualitative annotations can be made.

NOx (ppm)

1300 1100 900 700 500 300 100 20 40 60 80

Engine Load (% of full Engine Load)

Fig. 6a. Comparison of predicted and measured NOx and Soot concentrations at 1500 rpm engine speed, 20%, 40%, 60% and 80% engine load.
400 350 2000 rpm Engine Speed Calculated Experimental

Soot (mgr/m3)

300 250 200 150 100 50 0 1700 1500

850 800 750 700 650 600 550 500 450 400 350 300

Fig. 7a. Temperature in a vertical plane through the center of the fuel spray at -12 CA degrees ATDC, 2000 rpm engine speed, 80% load.

NOx (ppm)

1300 1100 900 700 500 300 100 20 40 60 80

Engine Load (% of full Engine Load)

Fig. 6b. Comparison of predicted and measured NOx and Soot concentrations at 2000 rpm engine speed, 20%, 40%, 60% and 80% engine load.
400 350 2500 rpm Engine Speed Calculated Experimental

In Fig. 7a the predicted in-cylinder temperature field at 12 CA degrees ATDC in a vertical plane through the center of the fuel spray is presented. As observed, there is a cold region near the nozzle exit, which is attributed to fuel evaporation. Gas temperatures near the cylinder boundaries are lower compared to the ones towards the center of the cylinder due to the heat transfer through the cylinder walls.
2600 2400 2200 2000 1800 1600 1400 1200 1000 800 600 400

Soot (mgr/m3)

300 250 200 150 100 50 0 1700 1500

NOx (ppm)

1300 1100 900 700 500 300 100 20 40 60 80

Fig. 7b. Temperature field in a vertical plane through the center of the fuel spray, at 10 CA degrees ATDC, 2000 rpm, 80% load.

Engine Load (% of full Engine Load)

Fig. 6c. Comparison of predicted and measured NOx and Soot concentrations at 2500 rpm engine speed, 20%, 40%, 60% and 80% engine load.

In Fig.7b is given the temperature field at 10 CA degrees ATDC in a vertical plane through the center of the fuel spray. It is obvious that higher temperatures lie in the regions where combustion has initiated, whereas in the regions where no combustible fuel-air mixture exists low temperatures are prevalent. Moreover it is shown that

combustion takes place mainly inside the piston bowl, as it is expected [7]. Fuel-Air Equivalence Ratio Spatial Distribution In Fig. 8 the local fuel-air equivalence ratio is shown at 10 CA degrees ATDC in a vertical plane through the center of the fuel spray. As observed the main portion of the fuel vapour is restricted inside the piston bowl. Moreover comparing Fig. 8 with Fig. 7b, the highest temperatures are observed in regions where the fuel-air equivalence ratio is near the stoichiometric value.
4.0 3.6 3.2 2.8 2.4 2.0 1.6 1.2 0.8 0.4 0.0

Max

Min

Fig. 9b. Soot in-cylinder distribution in a vertical plane through the center of the fuel spray at 10 CA degrees ATDC, 2000 rpm engine speed, 80% load.

CONCLUSIONS
Under the present work a new quasi-three dimensional combustion model has been developed. The main objective of the proposed model is to simulate in a simple and reliable way the air-fuel mixing, combustion and pollutant formation mechanism and the spatial distribution of the in-cylinder temperature and gas mixture concentration, in a reasonable computational time. In this model the basic features of multi-zone phenomenological models are embodied in the computational procedure followed by CFD models. Thus, although the results may be less detailed than those that a pure CFD model would give, it seems that the newly proposed model can be used to investigate the parameters affecting the fuel-air mixing and combustion mechanism and provide results on a cycle basis. To validate the model an extended experimental investigation is conducted at the authors laboratory on a high speed DI Diesel engine. Cylinder pressure and pollutant emissions are measured at various engine speeds and loads covering the entire engine operating range. Comparing measured and predicted values of cylinder pressure traces a good agreement is observed for all operating conditions examined without the need to vary the model constants. Similar results are observed for pollutant emissions (NO and Soot). The proposed model predicts the effect of operating parameters on NO and Soot quite well, which is the most important role of modelling, although differences between calculated and experimental emission concentrations have been observed when comparing absolute values. Moreover the predicted in-cylinder spatial distribution of temperature, fuel-air equivalence ratio and pollutant emission (NO, Soot) concentrations seems to be reliable and in accordance to the existing conceptual models for the fuel spray trajectory, fuel evaporation, combustion, and pollutant emission formation [7, 30]. Examining the key features of the model, it manages to describe in a more reliable way the air-fuel mixing mechanism compared to existing multi-zone models, where the air entrainment into each zone is estimated only based on

Fig. 8. Fuel-air equivalence ratio distribution in a vertical plane through the center of the fuel spray at 10 CA degrees ATDC, 2000 rpm engine speed, 80% load.

Spatial Distribution of NO and Soot Concentration As far as in-cylinder local pollutant emission concentrations are concerned, in Figs 9a,b the distribution of NO and Soot at 10 CA degrees ATDC in a vertical plane through the center of the fuel spray is shown. As revealed NO formation is predominant in regions where high temperatures are observed (Fig. 7b) and the gas mixture is stoichiometric to lean (Fig. 8), whereas soot concentration is higher in regions where fuel vapour concentration is higher and temperature is relatively high. Thus the model achieves to describe adequately the controversial mechanism of NO and Soot formation (i.e. formed in different areas), which is the main reason why it is difficult to find techniques that would reduce both of them.
Max

Min

Fig. 9a. NO in-cylinder distribution in a vertical plane through the center of the fuel spray at 10 CA degrees ATDC, 2000 rpm engine speed, 80% load.

empirical relations. Moreover in the proposed model the in-cylinder local conditions (temperature and fuel equivalence ratio), that affect combustion mechanism, are better estimated compared to multi-zone models, thus the effect of operating conditions on engine performance and pollutant emissions is simulated in a more detailed and fundamental way. Concluding, the proposed model manages to describe reliably the various processes taking place inside the cylinder during the close part of the engine cycle, in a more fundamental way, compared to existing multi-zone models while being less time consuming and complicated compared to existing CFD models.

u v Vcell Vf w wchar wgrid wpiston Y z zpiston

:Radial component of gas velocity :Circumferential component of gas velocity :Volume of a computational cell :Volume of fuel injected per fuel injection :Axial component of gas velocity :Characteristic velocity :Axial velocity of grid lines :Axial velocity of the piston :Mass concentration :Axial direction :Distance between the gas face of the cylinder head and the piston top

GREEK SYMBOLS NOMENCLATURE


:Area :Nozzle hole area :Transfer number for mass :Transfer number for heat :Nozzle Discharge coefficient :Specific heat of the gas :Mass which need to be transferred to a computational cell in order to have a pressure equal to the mean pressure :Nozzle hole diameter dn D32 :Sauter mean Diameter DfO :Diffusivity of fuel vapor at the interface between gas and liquid :Droplet diameter Dl h :Specific enthalpy of the gas k :Conduction heat transfer coefficient :Characteristic length lchar L :Heat of evaporation :Mass of the gas which is contained in a mcell computational cell :Mass of fuel vapor mfg ml :Liquid droplet mass NM :Number of zones in the radial direction of the fuel spray P :Pressure Pmean :Mean pressure of the gas in the cylinder q :Convective heat transfer from the hot gas to liquid fuel droplets r :Radial direction R :Gas constant :Universal gas constant Rmol S :Volumetric source rate Sconvection:Volumetric source rate due to heat transfer through the cylinder walls Spressure :Volumetric source rate due to the change of the pressure with respect to time Svaporization:Volumetric source rate due to heat absorption for liquid fuel evaporation S(M) :Penetration of fuel spray package M t :Time :Spray Break-up time tb T :Gas Temperature :Gas Temperature of a computational cell Tcell Tl :Droplet Temperature Twall :Temperature of the cylinder boundaries A An BM BT cD cp dmcell :Diffusion coefficient :Circumferential direction :Dynamic viscocity :Density :Crank angle degree :Spray Angle :Angular crankshaft speed

SUBSCRIPTS
fv gO l mix ox s sf sb :Ambient gas mixture :Fuel vapor :Interface between gas and liquid :Liquid :Mixture :Oxygen :Soot :Soot formation :Soot burnt

DIMENSIONLESS GROUPS
Re Pe Pr :Reynolds number :Peclet number :Prandtl number

ABBREVIATIONS
ATDC bsfc CA DI NO ppm TDC :After Top Dead Centre :Brake Specific Fuel Consumption :Crank Angle :Direct Injection :Nitric Oxide :Parts per million (volume) :Top Dead Centre

REFERENCES
1. Tow, T.C., Pierpont, D.A., Reitz, Reducing Particulate and NOx Emissions by Using Multiple Injections in a Heavy Duty D.I. Diesel Engine, SAE Transactions paper No 940897

2. Montajir, R.M., Trunemoto, H., Ishitani, H., Minami, T., Fuel Spray Behavior in a Small DI Diesel Engine : Effect of Combustion Chamber Geometry, SAE Transactions paper No 2000-01-0946 3. Assanis D.N., Heywood, J.B., Development and Use of a Computer Simulation of the Turbocompounded Diesel System for Engine Performance and Component Heat Transfer Studies, SAE Transactions paper No 860329 4. Hountalas D.T., Available Strategies for Improving the Efficiency of DI Diesel Engines-A Theoretical Investigation, SAE paper No 2000-01-1176 5. Kouremenos D.A., Hountalas, D.T., Binder K.B., The Effect of EGR on the Performance and Pollutant Emissions of Heavy-Duty Diesel Engines Using Constant and Variable AFR, SAE Transactions paper No 2001-01-0198 6. Greeves, G., Response of Diesel Combustion Systems to Increase of Fuel Injection Rate, SAE Transactions paper No 790037 7. Heywood, J.B., Internal Combustion Engine Fundamentals, McGraw-Hill, New York, 1988 8. Hountalas, D., Pariotis, E., A Simplified Model for the Spatial Distribution of Temperature in a Motored DI Diesel Engine, SAE Transactions paper No 2001-01-1235 9. Fitzgeorge, D. and Allison, J.L., Air Swirl in a RoadVehicle Diesel Engine, Proc. Instn. Mech. Engrs. (A.D.), No 4,pp. 151, 1962-63 10. Griffin, M.D., et al., Computational Fluid Dynamics Applied to Flows in an Internal Combustion Engine, AIAA, paper 78-57. 11. Ahmadi-Befrui, B., Gosman, A.D., Issa, R.I., Watkins, A.P., EPISO Implicit non-iterative solution procedure for the calculation of flows in reciprocating engine chambers, Computer Methods in Applied Mechanics and Engineering, Vol. 79, pp.249-279, 1990. 12. Gosman, A.D., et al., Axisymmetric Flow in a Motored Reciprocating Engine, Proc. Inst. Mech. Engrs, Vol. 192, No. 11, pp. 213-223, 1978. 13. Ramos, J.I., Humphrey. A.C., and Sirignano, W.A., Numerical Prediction of Axisymmetric Laminar and Turbulent Flows in Motored, Reciprocating Internal Combustion Engines, SAE Transactions, paper No 790356, 1979 14. Theodorakakos, A., Bergeles, G., Numerical investigation of the flow inside a 4-X IC model diesel engine, Journal Entropie, No 200, pp 53-63, 1996. 15. Patterson, M., Kong, S., Hampson, G., Reitz, R., Modeling the Effects of Fuel Injection Characteristics on Diesel Engine Soot and NOx Emissions, SAE Transactions paper No 940523 16. Patankar, S. V., Numerical Heat Transfer and Fluid Flow, Hemisphere Publishing Corporation, Taylor & Francis Group, New York 17. Bo, T., Clerides, D, Gosman, A., Theodossopoulos, P., Prediction of the Flow and Spray Process in an Automobile DI Diesel Engine, SAE paper No 970882

18. Crowe, C., Sharama, M., Stock, D., The ParticleSource-in-Cell (PSI-Cell) Model for Gas-Droplet Flows, ASME 75-WA/HT-25 19. Kouremenos, D.A., Rakopoulos, C.D., and Hountalas, D.T., Multi-Zone Combustion Modeling for the Prediction of Pollutants Emissions and Performance of DI Diesel Engines, SAE Transactions, paper No 970635, 1997. 20. Rakopoulos, C.D., and Hountalas, D.T., Development and validation of a 3-D multi-zone combustion model for the prediction of DI diesel engines performance and pollutants emissions, SAE Transactions, paper No 981021, 1998 21. Payri, F., Benajes, J., and Tinaut, F.V., A Phenomenological Combustion Model for Direct Injection, Compression Ignition Engines, Applied Math. Modeling, Vol. 12, pp.293-304, 1988 22. Bazarri, Z., A DI Diesel Combustion and Emission Predictive Capability for Use in Cycle Simulation, SAE paper 920462 23. Long, Y., Gakuma, H., Hiroyasu, H., The Simulation of the Distribution of Temperature and Mass of Liquid and Vapor Fuels, and the Wall Impingement Spray Pattern in a Diesel Combustion Champer, SAE paper No 2000-01-1887 24. Nishida, K., and Hiroyasu, H., Simplified ThreeDimensional Modeling of Mixture Formation and Combustion in a D.I. Diesel Engine, SAE Transactions, paper No 890269, 1989 25. Hiroyasu, H., Arai, M., Fuel Spray Penetration and Spray Angle of Diesel Engines, Transactions of JSAE, Vol. 21, pp.5-11, 1980 26. Jung, D., Assanis, D., Multi-Zone DI Diesel Spray Combustion Model for Cycle Simulation Studies of Engines Performance and Emissions, SAE paper 2001-01-1246 27. Borman, G.L., Johnson, J.K., Unsteady Vaporization Histories and Trajectories of Fuel Drops injected into Swirling Air, SAE paper 598C,1962 28. Wolfer, H.H., Ignition Lag in Diesel Engines, VDIForschungsheft 392, 1938; Translated by Royal Aircraft Establishment, Farnborough Library No. 358, UDC 621-436.047, August 1959 29. Vickland, C.W., Strange, F.M., Bell, R.A., Starkman, E.S., A consideration of the high temperature thermodynamics of internal combustion engines, SAE Transactions, 70, 785-793. 30. Dec, J., A Conceptual Model of DI Diesel Combustion Based on Laser-Sheet Imaging, SAE Transactions paper No 970873

Vous aimerez peut-être aussi