You are on page 1of 9

Applied Thermal Engineering 37 (2012) 344e352

Contents lists available at SciVerse ScienceDirect

Applied Thermal Engineering


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

Validation of a zero-dimensional model for prediction of NOx and engine


performance for electronically controlled marine two-stroke diesel engines
Fabio Scappin a, Sigurur H. Stefansson a, Fredrik Haglind a, *, Anders Andreasen b, Ulrik Larsen a
a
Technical University of Denmark, Department of Mechanical Engineering, DK-2800 Kgs. Lyngby, Denmark
b
MAN Diesel & Turbo, Engine Process Research, Process Development, Marine Low Speed Development, Teglholmsgade 41, DK-2450 Copenhagen SV, Denmark

a r t i c l e i n f o a b s t r a c t

Article history: The aim of this paper is to derive a methodology suitable for energy system analysis for predicting the
Received 24 May 2011 performance and NOx emissions of marine low speed diesel engines. The paper describes a zero-
Accepted 22 November 2011 dimensional model, evaluating the engine performance by means of an energy balance and a two
Available online 30 November 2011
zone combustion model using ideal gas law equations over a complete crank cycle. The combustion
process is divided into intervals, and the product composition and ame temperature are calculated in
Keywords:
each interval. The NOx emissions are predicted using the extended Zeldovich mechanism. The model is
Diesel engine
validated using experimental data from two MAN B&W engines; one case being data subject to engine
Modeling
Engine tuning
parameter changes corresponding to simulating an electronically controlled engine; the second case
NOx providing data covering almost all model input and output parameters. The rst case of validation
Performance suggests that the model can predict specic fuel oil consumption and NOx emissions within the 95%
Optimization condence intervals given by the experimental measurements. The second validation conrms the
Emissions capability of the model to match measured engine output parameters based on measured engine input
Zero-dimensional parameters with a maximum 5% deviation.
2011 Elsevier Ltd. All rights reserved.

1. Introduction a combined energy system, the focus is on the interaction among


components in the system rather than on component behavior.
The development of marine low speed diesel engines with lower Furthermore the newly adapted Engine Efciency Design Index
emissions is primarily driven by the MARPOL Annex VI regulation (EEDI), expected to enter into force on January 1, 2013 [2], may
[1] adopted by the International Maritime Organization (IMO). impose further constraints on the layout of the engine, auxiliaries,
MARPOL Annex VI contains regulations on NOx, SOx and Particulate etc. Thus, a fast, yet thermodynamically realistic engine model
Matter emissions. On July 1, 2010 the revised MARPOL Annex VI which can be integrated with an energy system analysis software is
entered in to force and it contains new limits for NOx emissions for highly desired for optimizing engine performance in combination
both new and existing ships as well as reduced SOx and PM emis- with waste heat recovery for minimal NOx as well as Green House
sions for all ships. Gas (GHG) emissions.
Allowable NOx emissions are reduced to 14.4 g/kWh for large The literature offers various modeling methodologies for the
marine low speed engines (130 rpm) installed on ships con- internal combustion engine. Scope of the application, accuracy and
structed from January 1, 2011 and onwards, according to the Tier II calculation time demand, are the determining parameters for the
standard, and to 3.4 g/kWh for engines installed on ships con- modeling approach [3e8]. Zero-dimensional models [9e13] with
structed from January 1, 2016 and onwards, according to the Tier III a few combustion zones are effective tools for providing reasonable
standard in designated emission control areas. estimations of NOx emissions and/or engine performance with low
The widely used low speed two-stroke diesel engine can be computational effort. Multi-zone combustion models [14e20]
combined with a waste heat recovery unit and potentially offer seem to offer more accurate NOx predictions, a more realistic
uniquely high fuel efciency and low specic emissions for diesel modeling of the fuel spray, as well as provide the modeling of other
engine ship propulsion. In designing and optimizing such emissions such as soot. These advantages come with the cost of
increased computing time, possibly without providing additional
information essential for energy system analysis.
* Corresponding author. Tel.: 45 45 25 41 13; fax: 45 45 93 52 15. Analytical models for cycle simulation of four-stroke engines
E-mail address: frh@mek.dtu.dk (F. Haglind). and HCCI engines have been described with some validation

1359-4311/$ e see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.applthermaleng.2011.11.047
F. Scappin et al. / Applied Thermal Engineering 37 (2012) 344e352 345

Nomenclature U internal energy (kJ)


V volume (m3)
v instantaneous piston speed (m/s)
Abbreviations and chemical formulas W mechanical work (kJ)
BDC Bottom Dead Center Xr residual gas fraction (e)
EVC Exhaust Valve Closing as heat transfer scaling factor
EVO Exhaust Valve Opening l air excess ratio
HCCI Homogeneous Charge Compression Ignition Dq duration of combustion period (rad)
SFC Specic fuel consumption q crank angle (rad)
SOI Start of injection n gas kinematic viscosity (m2/s)
VTG Variable turbine geometry s StefaneBoltzmann constant (5.670373$108 Wm2 K4)
g specic heat ratio
Notations u angular velocity (rad/s)
[] molar concentration
A surface area (m2) Subscripts
a crank radius (m) o outgoing
AF air-to-fuel ratio () HR heat release
AFT adiabatic Flame Temperature (K) 0 reference state
B bore (m) a air
c rod length (m) c compression
c1, c2 heat transfer coefcients ch charge (air fuel)
h specic enthalpy (kJ/kg) comp compression
k gas thermal conductivity (kW m1 K1) cyl cylinder
ki correlation factors for scavenging efciency (e) d diffusive combustion
M combustion shape factor (e) exh exhausts
m mass (kg) f fuel; forward
m_ mass ow (kg/s) fm formation
n a particular interval in which the combustion is i inlet; initial
divided (e) id ignition delay
N number of intervals in which the combustion is divided inj fuel injection
(e) max maximum
P absolute pressure (kPa) mot motoring
Q heat (kJ) osv opening scavenge valve
R universal gas constant (8.314 4621 kJ kmol1 K1) p pre-mixed combustion
rc compression ratio (e) P products
SEV scavenging efciency on volumetric basis (e) r reference, residual
Sp instantaneous piston speed (m/s) R reactants
SRV scavenging ratio (e) s swept
t time (s) t total
T temperature (K) W wall

[24e26]. Zannis and Rakopoulos [27,28] developed a multi-zone emissions of low speed, two-stroke diesel engines and the effect of
combustion model for predicting performance and emissions in engine tuning. Additionally the methodology is to be validated
large-scale two-stroke engines with comparisons to experimental against actual measurements on two engines.
data and Kowalski [29,30] proposed a simple method for simu- The description of the diesel engine model is covered in section
lating a two-stroke ship engine with some NOx emission validation. 2; the description of the experimental data used as a basis for
Most recently, Payri et al. [31] published a zero-dimensional ther- validating the model is outlined in section 3; the results and their
modynamic model of a four-stroke direct injection diesel engine analysis are discussed in section 4. Lastly, in section 5, the conclu-
with similarities with the model described in the present paper, sions are outlined.
although not including emission modeling.
While the development and methodologies found in these 2. Model description
papers offer clear descriptions of the models with some specic
validations to experimental data, few or no non-dimensional The overall sequence of running the sub models corresponds to
models for electronically controlled marine two-stroke diesel the stages in a two-stroke engine cycle, i.e. it runs from compres-
engines have been developed and validated using experimental sion to scavenging. Scavenging is modeled in order to calculate the
results for engine tuning adjustments (e.g. changed injection content of remaining exhaust gasses in the cylinder at the start of
timing, timing of exhaust valve closing and scavenging pressure; compression, so the model has to be run typically three or four
see e.g. [32e35]). Being able to predict with a reasonable accuracy cycles to converge. In this section the equations and methods used
the effects of engine tuning on performance and NOx emissions is of in the model are described. Firstly the energy balance and heat loss
particular importance for modern marine diesel engines. models, then the methodology for modeling the combustion event
The aim of this paper is to derive a methodology suitable for and later the models for NOx formation, exhaust gas temperature
energy system analysis and for predicting the performance and NOx and scavenging are presented.
346 F. Scappin et al. / Applied Thermal Engineering 37 (2012) 344e352

2.1. Energy balance characteristic distance, and a velocity factor that is dependent on
pressure rise caused by combustion and mean piston speed. Thus,
The engine energy balance for a control volume consisting of the the inuence of combustion on the heat transfer is included and the
cylinder gasses is solved over a complete crank cycle: heat transfer process is divided throughout the engine cycle into
the relevant stages [8]. Generally the correlation is appropriate if
dU dW dQHR dQW experimental data are available for calibration of the engine specic
 m _ o ho
_ i hi  m (1)
dt dt dt dt heat transfer characteristics [6]. The instantaneous characteristic
where U is the internal energy, W is the external work, QHR is the velocity v is dened as follows:
fuel heat release and QW is the heat loss through the cylinder wall,
Vs Tr
while m _ o ho are the inlet and outlet enthalpy ows,
_ i hi and m v c1 Sp c2 P  Pmot (7)
respectively. The rate of work rate is calculated as follows:
Pr Vr
where Pmot Pr(Vr/V)g, and g is the specic heat ratio. The
dW dV constants c1 and c2 need to be calibrated according to each specic
P (2)
dt dt engine and are varying with each stage of the cycle. Typical values
where P is the pressure inside the cylinder and the trapped volume for the gas exchange process are c1 6.18 and c2 0 [5] during the
V is given as a function of the crank angle q: compression stroke, c1 2.28 and c2 0 during combustion and in
the expansion stage c1 2.28 and c2 3.24 e-3. Vs is the swept
 r
V 1 c  c 2 volume, Vr, Tr, and Pr are the volume, temperature and pressure
1 rc  1 1  cosq  sin2 q (3) evaluated at any reference condition, in this study dened at the
VC 2 a a
time of exhaust valve closing (EVC) [5].
VC is the compression volume, c/a the piston rod length/crank The cylinder scavenging entails a rapid outgoing mass ow
radius and rc the engine compression ratio. Burn fraction and heat when the exhaust valve opens. The pressure dependency on the
release rate in internal combustion engines are mostly governed by crank angle q is set by the following equation, which describes
functions based on the law of normal distribution of continuous a sudden pressure drop, from the pressure at exhaust valve opening
random variables. Best known of these functions is the Wiebe (PEVO) to the outgoing exhaust pressure Po:
function, [36,37]. While the Wiebe function by no means describes
the complex fuel air mixing in the diesel combustion process, it can PEVO  Po expq qEVO  1
P Po (8)
provide valuable thermodynamic input for the model in terms of exp1 q  qEVO 1
a realistic (shape of the) heat release. The variant used in the
present model is the double Wiebe function following the Miya- The equation cannot be described as realistic and serves only the
moto model [38] (see also [39]): purpose of avoiding problems with discontinuities in the modeled
engine cycle. The initial pressure for this equation is dened as


 earlier described and the nal pressure is an input value for the
1 dQHR Qp   q Mp q Mp 1
6:9 Mp 1 exp  6:9 model (e.g. from measurements). This approach is suitable for
u dt Dqp Dqp Dqp
predicting the engine performance with reasonable accuracy in the

M d 

Q q q Md 1 case of lacking knowledge about the exhaust valve geometry and
6:9 d Md 1 exp  6:9 (4)
Dqd Dqd Dqd lift-law.
The gas is considered ideal, and the specic heat and enthalpy
Dqp and Dqd are the duration of the pre-mixed and diffusive are calculated as a function of the temperature and composition,
combustion periods, while Mp and Md are combustion shape using correlations suggested by Gyftopoulos and Beretta [42].
factors. Qp is the heat release due to the burning of the fuel during
the pre-mixed combustion, while Qd is the remaining heat release.
2.2. Combustion and mixing
Following Miyamoto [38], Mp and Dqp are given the values 3 and
7, respectively, while half of the fuel injected during the ignition
The combustion and NOx formation models are nested in a loop
delay is supposed to burn during the pre-mixed combustion. The
with step size equal to the time duration of the entire combustion
remaining parameters, Md and Dqd, are subject to model calibration
event divided by a chosen number of steps. 30e90 steps have been
to match the maximum pressure, power output and SFC as dis-
found to provide good accuracy and speed. Before the loop starts,
cussed in section 4. The ignition delay tid is described by the
combustion parameters are calculated; total heat release, global AF,
Arrhenius formula [40]:
ignition delay, Wiebe parameters and initial guess on AFT. In the
! loop two zones are considered; a zone with combustion and a zone
AF 6000
tid 1:8  105 exp (5) containing the remaining cylinder gasses. The loop starts by solving
P2 Tcyl
a system of ordinary differential equations by integration. This
system includes equations for heat losses by Woschni, energy and
Here, AF is the air-to-fuel ratio in the cylinder and Tcyl is the
mass balances, ideal gas law and heat release by Wiebe. The solu-
temperature in the cylinder when the ignition starts. Several
tion provides values of T, P, V, mass and HR over the time interval of
methodologies have been proposed on gas-to-wall heat transfer in
the present step. The mass of burnt fuel is now given and the mass
compression ignition engines. Among those is the widely used
of oxygen involved is found by multiplying a local lambda factor
Woschni correlation [41] which is implemented in the model, in all
with the mass of burnt fuel divided by the concentration (at this
the stages of the cycle:
step) of oxygen present in the cylinder gasses. This way the mass of
dQW h i gasses in the combustion zone is scaled to keep the oxygen
Acyl as B0:2 Pt0:8 Tt0:55 vt0:8 *Tt  Tw (6) concentration constant. The local lambda is constant for the 360
dt
engine cycle and is correlated with the load. This correlation is
as is a scaling factor used for calibration to match a specic engine subject to further investigation at the present time, but is showing
geometry. The correlation is based on a simple Nusselt and Rey- promise in the future work of validating the models part load
nolds number correlation by using the cylinder bore as the performance.
F. Scappin et al. / Applied Thermal Engineering 37 (2012) 344e352 347

The evaluation of the AFT requires knowledge of the combustion The NO formation rate in the interval (in the combustion calcula-
products composition. Thus the combustion products are now tion) is calculated, as a function of composition and temperature, by
determined, based on an initial guess of AFT, using the 11 species integrating the following equation [5]:
chemical equilibrium scheme described in Rakopoulos [43], except
for the formation of NO. Then the calculation of the exact AFT in the dNO
K1f N2 OK1b NONK2f O2 NK22b NOO
burning zone is carried out [47] and the combustion products are dt
re-calculated using the calculated AFT. Next NOx formation is K3f OHNK3b NOH (18)
modeled (see Section 2.3) and lastly the gasses from the two zones
are mixed to form the new composition of the cylinder gasses. This where all the concentrations (denoted [ ]), with the exception of
is done by adding the components (by mass weighted) of the two NO, are obtained from the chemical equilibrium calculation in the
zones together and dividing by the total mass of gas. burning zone. The numerical method used is described in ref. [5]
As described in Section 2.2, the combustion is divided into
intervals and complete mixing is supposed at the end of each nth
2.3. NOx formation interval between the burning and the non-reacting zone. The
hypothesis of obtaining a complete mixing of the two zones at the
The emission of nitrogen oxides is mainly in the form of NO and end of each interval is not fully realistic, and neither would be the
reactions that involve NO2 are negligible concerning NOx formation in assumption of a zero rate mixing modeleconditions that represents
low speed diesel engines [44]. Four contributions determine the total extremes in the mixing process in the combustion chamber [50].
NO emission due to combustion; thermal NO, prompt NO, fuel NO and Moreover, it is expected that the innite-rate mixing model would
NO arising from the N2O mechanism. Previous studies conrm the cause under-prediction of the NO formation, because of the lower
relevance of the thermal mechanism in the NO formation, consid- oxygen concentration in the reactants and of the lower tempera-
ering the contribution of the prompt mechanism negligible [21,45]. ture of the products, after they have been mixed with the non-
The N2O mechanism becomes important in fuel-lean pre-mixed reacted air. Conversely, a zero rate mixing model is expected to
combustion, which pertains more to gas turbines [46] and smaller over-predict the NO emissions.
four-stroke engines with larger pre-mixed burn fraction [47]. Fuel NO
contribution becomes relevant in the case the engine is operating
2.4. Exhaust gas temperature
with heavy fuel oil [21]. In fuel-lean combustion, all the fuel nitrogen
is converted to NO [48] and the fuel NO contribution to the total NO
The exhaust gas temperature is calculated by an overall energy
emission depends only on the fuel and not on the engine operation. A
balance of the engine over a complete cycle:
5e10% contribution from fuel-bound nitrogen to the total NOx
emission may be expected in the case of heavy fuel oil [49]. _  Q_
_ intake air CPairi Ti m
m _ f hf m
_ f LHV  W W
In the present model only the thermal mechanism is considered, Texh (19)
_
mexh cpexh
since no simple and applicable model is available and the effect is
isolated to the type of fuel used, not the engine process and the LHV is the lower heating value of the fuel. Qw is the total of heat
interplay with auxiliaries. The extended Zeldovich mechanism, losses found via the Woschni correlation in each of the stages of the
extensively described [21e23] and apparently an effective instru- cycle.
ment, which couples accuracy and low computing resources for the
calculation of thermal NOx, is implemented in the model:
2.5. Scavenging efciency and determination of initial temperature
at EVC
N2 O4N NO (9)
The initial temperature at EVC is inuenced by the amount of
O2 N4O NO (10) trapped hot residual gasses and the heat transfer from hot surfaces
to the gas. By evaluating the amount of trapped residual gasses after
OH N4H NO (11) each cycle, determined by residual gas fraction Xr, the initial
temperature can be evaluated using a simple initial temperature
The forward and backward reaction rate constants that dene
model [25,51].
reactions K1eK3 (Eqs. (9e11)) depend only on the temperature at
The scavenging efciency is used to specify the connection
which the reactions are occurring, here assumed to be the adiabatic
between the trapped residual gas and the fresh air at the start of the
ame temperature (AFT) of the burning zone. With the tempera-
compression stroke [52]. The relationship between the volumetric
ture T in Kelvin, the reaction rate constants, in cm3/(gmol s), used in
scavenge efciency, SEV, and the residual gas fraction is dened as:
reactions K1eK3 are [16]:
Xr 1  SEV (20)
k1f 7$1013 exp37800=T (12)
The scavenge efciency on volumetric basis is evaluated using
13 a numerical t to acquire the correlation between theory and
k1b 1:55$10 (13)
experimental data conducted for different scavenging characteris-
tics in two-stroke engines, Blair [52]:
k2f 1:33$1010 $T$exp3600=T (14)  
SEV 1  exp k0 k1 SRV k2 SR2V (21)
k2b 3:2$109 $T$exp19700=T (15)
where SRV is the scavenge ratio, or the ratio between the delivered
fresh air and the swept volume of the combustion chamber. The
k3f 7:1$1013 exp450=T (16) correlation factors k0, k1 and k2 are scavenging coefcients for
ported uniow scavenging in two-stroke engines, from Blair [52].
k3b 1:7$1014 exp24560=T (17) Once the scavenging ratio and the scavenging efciency are known
measures, the modied temperature at the time of EVC (Tmix) can
Fig. 1. Response surface of NOx as a function of EVC and SOI. All other independent
variables are kept at their center levels.

be determined by mixing of the fresh air charge with trapped


residual gas from the previous cycle:

cv;ch cv;r
Tmix i 1  Xr * *Ti i Xr * *Texh i  1 (22)
cv;t cv;t
where i is the cycle number index and subscripts r, ch and t
represents the residual gas, the charge (air fuel), and the total in-
cylinder gas respectively at EVC [25]. Assuming identical specic
heats for the intake air and exhaust gases, the expression for initial
temperature is simplied [5]:

Tmix i 1  Xr *Ti i Xr *Texh i  1 (23)


Hence, the trapped cylinder mass for the ith cycle at EVC is
determined using the ideal gas law:

PEVC VEVC
mc i (24)
RTmix i

3. Experimental and response surface model validation data

In order to validate the developed engine model, experimental


ndings are used from Goldsworthy [21,58] and Mayer [54] based
on measurements on two direct injection, turbocharged, uniow
scavenged, low speed two-stroke marine MAN B&W diesel engines,
models 4T50 ME-X and 7L70 MC.
The data from the 4T50 ME-X engine is comprehensive and
consists of 50 data points, and it contains simultaneous variations
in several pivotal parameters severely affecting both engine
performance and emissions. The parameters are: Exhaust valve
closing timing (EVC), start of injection (SOI), injection pressure
(Pinj), injection prole (InjProf), injection nozzle hole size (Nozz),
and turbine area (VTG), controlling mainly (but not only)
compression pressure ratio to the scavenging pressure, pressure
rise, injection length and intensity, as well as scavenging pressure.
The number of tests has been kept at a minimum by employing
principles from the theory of design of experiments [56] with the
test plan laid out according to a central composite face centered
design (response surface methodology), which contains three
levels for all six variables (factors), with several repetitions of the
center point. Since many parameters are varied at a time, the
experimental data are used to provide response surface models [57]
Fig. 2. aed. Predicted versus measured responses. Pcomp has been left out, since it is
of responses SFC, Pmax, Pcomp, Pscav, and NOx used for validating the
similar to Pmax in terms of R2.
F. Scappin et al. / Applied Thermal Engineering 37 (2012) 344e352 349

Table 1
Engine operative conditions.

Engine model 4T50 ME-X 7L70 MC


Cylinders 5 7
Bore [mm] 500 700
Stroke [mm] 2200 2268
Compression volume [m3] 0.0252 N/A
Compression ratio 18.14 12.6 (effective)
Fuel LHV [kJ/kg] 42 700 42 700
Specic humidity of the charge air (ISO) 0.0107 0.0107
Load 75% 75%
Speed [Revs/min] 111.8 98.1
Maximum pressure [MPa] 16.0 12.6
Compression pressure [MPa] 14.1 10.1
Scavenge pressure [MPa] 0.29 0.28
Fuel per cylinder [kg/s] 0.06199 0.1009
Air per cylinder [kg/s] 3.26 5.885
Fig. 3. Heat release versus crank angle at 75% engine load. Scavenge temperature [ C] 30.6 34.0
Cooling load [kW] N/A 2400
Injection angle [degrees after BDC] N/A 178.5
present model. The response surface is illustrated for NOx in Fig. 1.
The agreement with the experimental ndings is excellent and the
model error is minimal as visualized in Fig. 2aed. the heat release is considered acceptable. The deviation at the peak
The sensitivity of NOx and SFC on compression and maximum of heat release rate shows that the model seems to predict more
pressures has been calculated effectively by changing SOI but energy released during early stages of the combustion as seen in
simultaneously allowing EVC to adjust in order to have unchanged Fig. 3. However, when looking at the cumulative heat release (burn
compression pressure for varying maximum pressure, and to have fraction) in Fig. 4, there is good agreement with the timing of the
constant maximum pressure for varying compression pressure. energy release.
Effectively scavenge pressure is changed by adjusting VTG, but In order to validate further the engine model, a simulation
again simultaneously allowing EVC and SOI to adjust in order to model of the 7L70 MC mk6 MAN B&W engine is built according to
have unchanged maximum and compression pressure. The sensi- the data provided in two articles by Goldsworthy [21,58]; see
tivity of NOx and SFC on pressure levels hereby provided is used for Table 1. The articles provide a large part of the model input
validating the present model. These sensitivities are referred to as parameters and results, and this data is thus suitable for validation
(surrogate) experimental, since the response surface models used purposes because it puts constraints on the model, i.e. there are
are close to a 1:1 match of the measured data. only a few parameters available for calibration adjustments to help
The experimental data for the 7L70 MC engine is obtained from match the measurements. Table 2 shows how the measured values
a paper by Goldsworthy [58]. and the model outputs are in good agreement, i.e. within 5% of
measured data values.
4. Results and discussions

Firstly the model validation results using 4T50 ME-X engine Table 2
data are shown and secondly validation results using the 7L70 MC The 7L70 MC engine model results compared to experimental data.

engine data. Experimental Engine model Deviation


values output
Maximum pressure (MPa abs) 12.6 12.1 3.9%
4.1. General validation by 4T50 ME-X and L70MC experimental
Compression pressure (MPa abs) 10.1 9.6 4.9%
data Specic fuel consumption (g/kWh) 171.1 172.1 0.6%
Power output per cylinder (MW) 2.123 2.111 0.5%
Data obtained from MAN Diesel & Turbo providing the NOx emissions (g/kWh) 17.6 17.6 0.0%
measured heat release rate is used to calibrate the injection timing
and the combustion shape parameters, see Figs. 3 and 4. The
agreement between the model results and the measured values for

Fig. 4. Burn fraction versus crank angle at 75% engine load. Fig. 5. Normalized SFC versus maximum pressure by varying injection timing.
350 F. Scappin et al. / Applied Thermal Engineering 37 (2012) 344e352

Fig. 6. NOx emissions versus maximum pressure by varying injection timing. Fig. 8. NOx emissions versus compression pressure by varying EVC timing keeping
maximum pressure constant.

Since the model is using the ideal gas equation of state (EOS),
somewhat lower calculated pressures, compared to the measure- its effect on the maximum cylinder pressure and AFT. Advancing
ments, are to be expected. Investigating a case of adiabatic the injection timing increases the pre-mixed combustion temper-
compression using the ideal gas and the real gas EOS developed by ature which results in higher NOx formation [55]. Here, the injec-
Lemmon et al. [59], respectively, resulted in about 5% lower tion timing is used as a control parameter to vary the maximum
compression pressure, when assuming ideal gas at the relevant pressure and the effect on SFC and NOx emissions are evaluated. As
conditions as compared to the real gas EOS. This explains the seen in Figs. 5 and 6 results suggest that the model is able to predict
deviation in compression pressure as well as maximum pressure the effect of maximum pressure on SFC and NOx emissions all
shown in Table 2. As also seen from the table, the modeled NOx within the 95% condence intervals (/2s) of the measured data,
emission and SFC are in excellent agreement with the experimental indicated by error bars to the mean value in the result in Figs. 5e10
ndings. [53,54].
As mentioned only a few parameters are available for calibration
since the measured parameter values provided are extensive. 4.3. Effect of compression pressure on SFC and NOx emissions
Provided parameters relevant for the presented methodology are:
Engine geometry, effective compression ratio, engine speed, intake Generally, when increasing the compression pressure, more
air composition and fuel mass ows, intake air temperature and work is needed for the compression stroke, resulting in higher SFC
pressure, pressure loss due to scavenging and start of injection. and specic NOx emissions [14]. Increasing the compression pres-
Parameters available for calibration are: Timing of end of injection, sure results in higher maximum pressure which could exceed the
timing of opening of the exhaust valve, and heat release function structural limits of the engine [45]; thus when calibrating the
parameters in the diffusive combustion phase (Eq. (4)). In the compression pressure by the timing of EVC, the maximum pressure
following sections the model is validated against parameter varia- is kept constant at a reference structural limit of 160 bar, by
tions from 4T50 ME-X. adjusting the injection timing accordingly. Figs. 7 and 8 illustrate
how SFC and NOx emissions both increase with increasing
4.2. Effect of maximum pressure on SFC and NOx emissions compression pressure. A comparison between model results and
measured data suggests that the model is also able to predict the
The injection timing is an important factor in the combustion SFC and NOx within the 95% condence although the model seems
process, and SFC and NOx emissions are strongly inuenced due to to over-predict NOx emissions at relatively low compression

Fig. 7. Normalized SFC versus compression pressure by varying EVC timing keeping Fig. 9. Normalized SFC versus scavenge air pressure while keeping maximum pressure
maximum pressure constant. and compression pressure constant.
F. Scappin et al. / Applied Thermal Engineering 37 (2012) 344e352 351

due to the use of the ideal gas law, in the range of less than 5%
deviation. It is recommended though that future work should
include implementation of more accurate equations of state in the
methodology and faster optimization methods.
In conclusion, the model derived here features the desired
rapidity of execution and provides a sufcient accuracy, with
regards to essential engine performance characteristics, for being
used for future energy system analysis. The model has demon-
strated capability of responding well to tuning engine parameters
regarding SFC and NOx emissions, thus enabling the possibility to
use the model to explore different scenarios, e.g. SFC minimization
within the IMO NOx emission limits either as a stand-alone engine
or combined with a waste heat recovery system.

Acknowledgements

Fig. 10. NOx emissions versus scavenge air pressure keeping maximum pressure and The authors would like to thank Brian Elmegaard, Michael Vin-
compression pressure constant.
cent Jensen and Kim Rene Hansen from Technical University of
Denmark, Department of Mechanical Engineering for the advice and
pressures. This could be explained by the fact that no detailed discussions during the work. A special acknowledgement is given to
information about the scavenging process is involved in the model. Prof. Spencer C. Sorenson from Technical University of Denmark,
Therefore, there might be some differences in the actual cylinder Department of Mechanical Engineering who provided many helpful
properties at the time of EVC compared to calculated properties. suggestions during the model development and for the manuscript.
Another factor possibly causing the deviations is the use of the ideal
gas law which underestimates the actual pressure at the end of the References
compression stroke. This might be improved by implementing real
gas behavior. [1] Revised MARPOL Annex VI, Regulations for the Prevention of Air Pollution
from Ships and NOx Technical Code 2008, International Maritime Organiza-
tion, London, UK, 2009.
4.4. Effect of scavenging pressure on SFC and NOx emissions [2] Mandatory energy efciency measures for international shipping adopted at
IMO environment meeting, Marine Environment Protection Committee
(MEPC) e 62nd session: 11 to 15 July 2011. http://www.imo.org/MediaCentre/
Figs. 9 and 10 illustrate how increasing scavenge pressures PressBriengs/Pages/42-mepc-ghg.aspx.
result in low specic NOx emissions and SFC [7,54]. Here, the [3] E. Pedersen, H. Valland, H. Engja, Modelling and simulation of Diesel engine
compression pressure and the ring ratio (Pmax/Pcomp) were kept processes, Paper 34, CIMAC Congress, Interlaken, Switzerland, 1995.
[4] H. Grimmelius, E. Mesbahi, P. Schulten, D. Stapersma, The use of diesel engine
constant by adjusting the timing of EVC and the injection timing. simulation models in ship propulsion plant design and operation, Paper 227,
The results suggest that the model also predicts this effect within CIMAC conference, Wien, Austria, 21e24 May 2007.
the 95% condence interval limits of the measured data. [5] S.C. Sorenson, Engine Principles and Vehicles, Technical University of Denmark,
Lyngby, Denmark, 2008.http://books.google.com/books?idVFVtQwAACAAJ.
[6] J.B. Heywood, E. Sher, The Two-Stroke Cycle Engine, Taylor & Francis, London,
5. Conclusions UK, 1999.
[7] J.B. Heywood, Internal Combustion Engine Fundamentals, McGraw-Hill, New
York, USA, 1988.
A zero-dimensional model for electronically controlled two-
[8] S.H. Chan, J. Zhu, Modelling of engine in-cylinder thermodynamics under high
stroke low speed diesel engines was derived. A two zone values of ignition retard, International Journal of Thermal Sciences 40 (2001)
combustion model, an energy balance and the ideal gas law 94e103.
[9] G. P. Merker, B. Hohlbaum, M. Rauscher, Two-zone model for calculation of
equations were applied and coupled with a combustion and NOx
Nitrogen-Oxide formation in direct-injection diesel engines, SAE paper
formation calculation subroutine using the extended Zeldovich 932454, 1993.
mechanism. Initial properties of the trapped gas in the cylinder at [10] R. J. Asay, K. I. Svensson, D. R. Tree, An Empirical, Mixing-Limited, Zero-
the start of compression were determined from the mixture of fresh Dimensional Model for Diesel Combustion, SAE paper 2004-01-0924, 2004.
[11] D.T. Hountalas, Prediction of marine diesel engine performance under fault
air introduced in the cylinder and trapped residual gasses evaluated conditions, Applied Thermal Engineering 20 (2000) 1753e1783.
from the scavenging efciency and residual gas fraction. The [12] D. Descieux, M. Feidt, One zone thermodynamic model simulation of an ignition
modeled heat release was calibrated to match measured data, by compression engine, Applied Thermal Engineering 27 (2007) 1457e1466.
[13] X. Tauzia, A. Maiboom, P. Chesse, N. Thouvenel, A new phenomenological heat
selecting the correct timing of fuel injection and adjusting release model for thermodynamical simulation of modern turbocharged
combustion shape parameters. heavy duty, Diesel Engines 26 (2006) 1851e1857.
The effects of engine tuning parameters on SFC and NOx emission [14] G. Stiesh, G.P. Merker, A Phenomenological Heat Release Model for Direct
Injection Diesel Engines, 22nd CIMAC Congress, Copenhagen, 1998.
at 75% engine load were validated by data from experimental tests [15] A. Maiboom, X. Tauzia, S.R. Shah, J.F. Hetet, New phenomenological six-zone
on the MAN B&W 4T50 ME-X test engine. The results indicate that combustion model for direct injection diesel engines, Energy Fuels 23
the trends on SFC and NOx emissions of varied maximum pressure, (2009) 690e703.
[16] O. Durgun, Z. Sahin, Theoretical investigation on heat balance in DI diesel
attained by changed injection timing, can be captured within the
engines for neat diesel fuel and gasoline fumigation, Energy Convers. Manage.
95% condence interval of the measurements. Also the effects on SFC 50 (2009) 43e51.
and NOx emissions of adjusting the compression pressure, by [17] A.M. Kulkarni, G.M. Shaver, S.S. Popuri, T.R. Frazier, D.W. Stanton, Computa-
tionally efcient whole-engine model of a Cummins 2007 turbocharged diesel
changing the timing of EVC, and of scavenging pressure, respec-
engine, J. Eng. Gas Turbines Power 132 (2010) 0228031.
tively, were predicted within limits of the measure values. [18] M. Andersson, B. Johansson, A. Hultquist, C. Nhre, A Real Time NOx Model for
The model was also validated using data from the MAN B&W Conventional and Partially Premixed Diesel Combustion, SAE paper 2006-01-
7L70 engine comprising both engine operating parameters and 0195, 2006.
[19] D. Jumg, D. A. Assanis, Multi-Zone DI Diesel Spray Combustion Model for Cycle
measured engine outputs. There was good agreement between Simulation Studies of Engine Performance and Emissions, SAE paper 2001-01-
model and measurements, although modeled pressures were low 1246, 2001.
352 F. Scappin et al. / Applied Thermal Engineering 37 (2012) 344e352

[20] I. Arsie, F. Di Genova, A. Mogavero, C. Pianese, G. Rizzo, A. Caraceni, P. Ciof, G. [38] N. Miyamoto, T. Chikahisa, T. Murayama, R. Sawyer, Description and Analysis
Flauti, Multi-Zone Predictive Modeling of Common Rail Multi-Injection Diesel of a Diesel Engine Rate of Combustion and Performance Using Wiebes
Engines, SAE paper 2006-01-1384, 2006. functions, SAE Paper 850107, SAE Trans. vol. 94, 1985.
[21] L. Goldsworthy, Reduced kinetic schemes for oxides of nitrogen emissions [39] H. Yasar, H.S. Soyhan, H. Walmsley, B. Head, C. Sorusbay, Double-Wiebe
from a slow speed marine diesel engine, Energy Fuels 17 (2003) 450e456. function: an approach for single-zone HCCI engine modelling, Applied
[22] T. Chikahisa, M. Konno, T. Murayama, T. Kumagai, Analysis of NO formation Thermal Engineering 28 (2008) 1284e1290.
characteristic and its control concepts in diesel engines from NO reaction [40] G.P. Merker, C. Shwarz, G. Stiesh, F. Otto, Simulating Combustion, Springer-
kinetics, JSAE Review 15 (1994) 297e303. Verlag, Berlin Heidelberg, Germany, 2006, p. 108.
[23] K. Kikuta, T. Chikahisa, Y. Hishinuma, Study on predicting combustion and [41] G. Woschni, A Universally Applicable Eqution for the Instantaneous Heat
NOx formation in diesel engines from scale model experiments, Transactions Transfer Coefcient in the Internal Combustion Engine, SAE Paper 670931,
of the Japan Society of Mechanical Engineers 43 (2000) 89e96. 1967.
[24] L. Eriksson, I. Andersson, An analytical Model for Cylinder Pressure in a Four [42] E.P. Gyftopoulos, G.P. Beretta, Thermodynamics: Foundations and Applica-
Stroke SI engine, SAE paper2002-01-0371, 2002. tions, Macmillan, New York, 1991.
[25] M. Shahbakhti, C.R. Koch, Physics based control oriented model for HCCI [43] C.D. Rakopoulos, D.T. Hountalas, E.I. Tzanos, G.N. Taklis, A fast algorithm for
combustion timing, Journal of Dynamic Systems, Measurement and Control calculating the composition of diesel combustion products using 11 species
132 (2010) 021010. chemical equilibrium scheme, Adv. Eng. Software 19 (1994) 109e119.
[26] D.J. Rausen, A.G. Stefanopoulou, A mean-value model for control of homo- [44] E.C. Zabetta, P. Kilpinen, Improved NOx submodels for in-cylinder CFD simu-
geneous charge compression ignition (HCCI) engines, Journal of Dynamic lation of low e and medium speed compression ignition engines, Energy Fuels
Systems, Measurement, and Control 17 (2005) 355e362. 15 (2001) 1425e1433.
[27] T. C. Zannis, V. T. Lamaris, D. T. Hountalas, S. E. Glaros, Development and [45] J. E. Dec, R. E. Canaan, PLIF imaging of NO formation in a DI diesel engine, SAE
validation of a multi-zone combustion model for predicting performance Technical Paper 980147, 1998.
characteristics and NOx emission in large scale two-stroke diesel engines, [46] S.R. Turns, An Introduction to Combustion, second ed. McGraw-Hill, Singa-
ASME, 2009e11382, 2009. pore, 2000.
[28] C.D. Rakopoulos, K.A. Antonopoulos, D.C. Rakopoulos, Development and [47] S.M. Correa, A review of NOx formation under gas-turbine combustion
application of multi-zone model for combustion and pollutants formation in conditions, Combust. Sci. Technol 87 (1993) 329e362.
direct injection diesel engine running with vegetable oil or its bio-diesel, [48] Z. Bazari, DI diesel combustion and emission predictive capability for use in
Energy Conversion and Management 48 (2007) 1881e1901. cycle simulation, SAE Technical Paper 920462, 1992.
[29] J. Kowalski, W. Tarelko, NOx emission from a two-stroke ship engine. Part 1: [49] J.W. Miller, H. Agrawal, W.A. Welch, Criteria Emissions from the Main Pro-
modeling aspect, Applied Thermal Engineering 29 (2009) 2153e2159. pulsion Engine of a Post-Panamax Class Container Vessel Using Distillate and
[30] J. Kowalski, W. Tarelko, NOx emission from a two-stroke ship engine. Part 2: Residual Fuels (CARB Report), University of California, USA, 2009.
laboratory test, Applied Thermal Engineering 29 (2009) 2160e2165. [50] R. C. Peterson, K. J. Wu, The Effect of Operating Conditions on Flame
[31] F. Payri, P. Olmeda, J. Martn, A. Garca, A complete 0D thermodynamic Temperature in a Diesel Engine, SAE Technical Paper 861565, 1986.
predictive model for direct injection diesel engines, Applied Energy 88 (2011) [51] L. Eriksson, I. Andersson, An Analytical Model for Cylinder Pressure in a Four
4632e4641. doi:10.1016/j.physletb.2003.10.071. Stroke SI Engine, SAE Technical Paper 2002-01-0371, 2002.
[32] A. Parlak, H. Yasar, C. Hasimog lu, A. Kolip, The effects of injection timing on [52] G.P. Blair, Design and Simulation of Two-Stroke Engines, Society of Automo-
NOx emissions of a low heat rejection indirect diesel injection engine, Applied tive Engineers, Inc., Warrendale, USA, 1996.
Thermal Engineering 25 (2005) 3042e3052. [53] M. Finch Pedersen, A. Andreasen, S. Mayer, Two Stroke Engine Emission
[33] Cenk Sayin, Metin Gumus, Impact of compression ratio and injection Reduction Technology: State-of-the-Art, Paper 85, CIMAC Congress, Bergen,
parameters on the performance and emissions of a DI diesel engine fueled Norway, 14e17 June 2010.
with biodiesel-blended diesel fuel, Applied Thermal Engineering 31 (16) [54] S. Mayer, T. A. E. Tuner, A. Andreasen, Design of experiments analysis of the
(November 2011) 3182e3188. NOx-SFOC trade-off in two-stroke marine engine, Paper 38, CIMAC Congress,
[34] J. Benajes, S. Molina, J. Martn, R. Novella, Effect of advancing the closing angle Bergen, Norway, 2010.
of the intake valves on diffusion-controlled combustion in a HD diesel engine, [55] A. Sarvi, R. Zevenhoven, Large-scale diesel engine emission control parame-
Applied Thermal Engineering 29 (2009) 1947e1954. ters, Energy 35 (2010) 1139e1145.
[35] J.M. Desantes, J. Benajes, S. Molina, C.A. Gonzlez, The modication of the fuel [56] G.E.P. Box, W.G. Hunter, J.S. Hunter, Statistics for Experimenters, John Wiley &
injection rate in heavy-duty diesel engines. Part 1: effects on engine perfor- Sons, Hoboken, New Jersey, USA, 1978.
mance and missions, Applied Thermal Engineering 24 (2004) 2701e2714. [57] R.H. Myers, D.C. Montgomery, C.M. Anderson-Cook, Response Surface Meth-
[36] J.I. Ghojel, Review of the development and applications of the Wiebe function: odology, third ed. John Wiley & Sons, Hoboken, New Jersey, USA, 2009.
a tribute to the contribution of Ivan Wiebe to engine research, International [58] L. Goldsworthy, Real time model for oxides of nitrogen emissions from a slow
Journal of Engine Research 11 (2010) 297e312. speed marine diesel, Journal of Marine Engineering and Technology (2002) No. A2.
[37] J. Galindo, H. Climent, B. Pl, V.D. Jimnez, Correlations for Wiebe function [59] E.W. Lemmon, R.T. Jacobsen, S.G. Penoncello, Daniel Friend, Thermodynamic
parameters for combustion simulation in two-stroke small engines, Applied properties of air and mixtures of nitrogen, argon, and oxygen from 60 to 2000
Thermal Engineering 31 (2011) 1190e1199. K at pressures to 2000 MPa, J. Phys. Chem. Ref. Data 29 (2000) 331e385.

You might also like