Вы находитесь на странице: 1из 12

Downloaded from SAE International by Alessandro D'Adamo, Wednesday, July 29, 2015

Numerical Investigation on the Effects of Water/Methanol


Injection as Knock Suppressor to Increase the Fuel Efficiency
of a Highly Downsized GDI Engine

2015-24-2499
Published 09/06/2015

Fabio Berni, Sebastiano Breda, Alessandro D'Adamo, and Stefano Fontanesi


Universita degli Studi di Modena

Giuseppe Cantore

Univ of Modena & Reggio Emilia


CITATION: Berni, F., Breda, S., D'Adamo, A., Fontanesi, S. et al., "Numerical Investigation on the Effects of Water/Methanol
Injection as Knock Suppressor to Increase the Fuel Efficiency of a Highly Downsized GDI Engine," SAE Technical Paper 2015-242499, 2015, doi:10.4271/2015-24-2499.
Copyright 2015 SAE International

Abstract

Introduction

A new generation of highly downsized SI engines with specific power


output around or above 150 HP/liter is emerging in the sport car
market sector. Technologies such as high-boosting, direct injection
and downsizing are adopted to increase power density and reduce
fuel consumption. To counterbalance the increased risks of preignition, knock or mega-knock, currently made turbocharged SI
engines usually operate with high fuel enrichments and delayed
(sometimes negative) spark advances. The former is responsible for
high fuel consumption levels, while the latter induce an even lower
A/F ratio (below 11), to limit the turbine inlet temperature, with huge
negative effects on BSFC.

The recent rise in fuel prices, the need to reduce ground transportgenerated emissions (more and more constrained by legislation) and
to improve urban air quality have brought fuel-efficient, lowemissions powertrain technologies at the top of vehicle
manufacturers' and policy makers' agenda. One of the most promising
techniques adopted by engine designers in order to obtain a reduction
in pollutant formation and fuel consumption, without incurring into
lower engine performance, is the so-called downsizing approach,
meaning with this a renewed interest into the thermal efficiency
optimization of the whole chain of engine processes, from
combustion to engine cooling. Standardized practice for the
passenger car market, downsizing is a more recent approach on the
high performance engine side: a trend can now be seen in this specific
market sector, where a new generation of highly downsized engines
with specific power outputs around or above 150 HP/litre is
emerging. Consequences of this technique are the very high levels of
in-cylinder pressure and thermal loading, which are well-known
causes for the onset of abnormal combustion events in SI engines.
The most damaging and analyzed of these is engine knock [1, 2],
which is originated by the spontaneous ignition of the unburned
charge ahead of the flame front. To prevent, or at least limit, knock
onset, charge cooling by means of the adoption of fuel-rich mixtures
is usually applied. Despite the implied reduction of engine efficiency
(or related increase of specific fuel consumption), mixture enrichment
is preferable to the reduction of either the SA or the boost level, since
these last are responsible for the amount of indicated work. The
vaporization of the exceeding fuel enhances the heat subtraction and
it increases the air/fuel mixture specific heat. Both factors limit the
end-gas compression heating due both to the piston movement and to
the high-pressure combustion developed in the cylinder volume. Even
if the result is a reduction of the risk of knock onset, the poor
energetic efficiency intrinsic in the use of some gasoline just for
cooling purposes makes this approach diametrically opposite to the
legislation trends in terms of fuel efficiency. To avoid fuel-rich

A possible solution to increase knock resistance is investigated in the


paper by means of 3D-CFD analyses: water/methanol emulsion is
port-fuel injected to replace mixture enrichment while preserving, if
not improving, indicated mean effective pressure and knock safety
margins. The peak power engine operation of a currently made
turbocharged GDI engine is investigated comparing the adopted
fuel-only rich mixture with stoichiometric-to-lean mixtures, for
which water/methanol mixture is added in the intake port under
constant charge cooling in the combustion chamber and same air
consumption level. In order to find the optimum fuel/emulsion
balance analytic considerations are carried out. Different strategies
are evaluated in terms of percentage of methanol-water emulsion rate,
to assess the effects of different charge dilutions and mixture
compositions on knock tendency and combustion efficiency. Thanks
to the lower chemical reactivity of the diluted end gases and the faster
burn rate allowed by the methanol addition, the water/methanolinjected engine allows the spark advance (SA) to be increased; as a
consequence, engine power target is met, or even crossed, with a
simultaneous relevant reduction of fuel consumption.

Downloaded from SAE International by Alessandro D'Adamo, Wednesday, July 29, 2015

mixtures, injection of alternative vaporizing liquids can be used to


achieve the same charge cooling effect with reduced gasoline
amounts.
The application of water cooling is not a novelty in internal
combustion engines, and the first studies on its potential on knock
inhibition can be traced back to the early 1930's studies of Sir Harry
Ricardo. Much later, a limited number of road vehicles with largedisplacement engines from manufacturers such as Chrysler have
included water injection. Saab offered water injection for the Saab 99
Turbo. With the introduction of the turbocharger intercooler, the
interest in water injection disappeared. More recently, it was used
simply as a method to enhance air cooling (in addition to the use of
intercooler and fuel-rich mixtures), in order to increase engine
performance. However, benefits of water injection stretch beyond: in
the recent years the potential gain deriving from water injection has
been investigated both on SI and CI engines, mainly to limit pollutant
emissions. Lanzafame [3] showed the remarkable potential in NOx
formation reduction in a SI engine with port-injected water.
Regarding CI units, the focus was again on the decreased particulate
matter (PM) and NOx formation [4, 5] and water injection was found
more effective than EGR (Exhaust Gas Recirculation) especially at
high rpms.
Combining the just mentioned trends in the recent downsized engine
design and the water adoption in turbocharged engines, a different
perspective was pursued by the authors in previous studies [6, 7],
where water injection confirmed to be a valid substitute of the excess
fuel in originally fuel-enriched engines: therefore, water was used to
substantially lean out the originally rich mixture (=1.21), and not a
simple addition with the aim of knock tendency gain as in previous
applications. Two levels of numerical investigation were followed in
[7]. The preliminary part of the study was developed in a 0D
framework, where the trade-off between the beneficial increase in
autoignition delay time and the undesired slowdown of the burning
velocity was identified as maximum for the stoichiometric condition
with a specific amount of injected water (status B), as recapped in
Figure 1.

Figure 1. Results from a 0D analysis of combustion duration elongation and


increased knock resistance for several mixture leaning (point A: =0.9, point
B: =1, point C: =1.1) and water addition levels.

Detailed 3D CFD analyses were then carried out to confirm this


prediction and were used to compare the fuel-only cases with the
fuel+water ones. Results for the peak power engine speed condition
showed that the fuel consumption was improved by a 25% thanks
the simultaneous mixture leaning and the water addition. However,
the same strategy was not found so beneficial for low rpm conditions
where a reduced intake charge temperature largely lengthened the
evaporation time needed for the injected water, strongly limiting the
effectiveness of this technology. Therefore, the numerical simulations
presented in [7] indicated a relevant field of fuel efficiency
improvement in such highly-turbocharged DISI engine, where rich
mixtures were necessarily employed as a knock-limiter actor.
The paper is structured as follows. In the next section the advantages
and drawbacks of methanol-water injection are outlined and the
constraints followed in the presented analyses are explained. Then,
numerical results are shown for the different combustion and knock
behavior for the proposed MW blends. Finally, the specific fuel
consumption is estimated for the simulated cases and the gain from
the adoption of MW injection is illustrated.

Methanol - Water Methodology


Blends of water with alcohols, in this study methanol and hereafter
indicated as methanol-water (MW) mixtures, are receiving analogous
interest, as they offer the additional advantage to have a portion of the
evaporating mixture which is a secondary fuel. Unlike the water
injection alone, the presence of a second fuel allows the engine to
meet higher performance. Moreover, the latent heat of vaporization of
methanol is three times that of gasoline: its evaporation can be
potentially exploited to further reduce the mixture temperature at
compression TDC (top dead center), resulting in a potentially
effective knock suppressor. In addition to this aspect, the octane
number of methanol is much higher than that of gasoline. When this
is added to the already high anti-knock quality of water-enriched
end-gases, the global result is that the engine is allowed to run with
further increased compression ratio or SA for even better fuel
economy. Another advantage of methanol is that its laminar flame
speed is higher than that of gasoline [19], thus burn rate is also
expected to be improved. The oxygen content of methanol will
promote the complete combustion of mixture and reduce soot
emission. Finally, a potential benefit of MW mixtures over water-only
injection is the faster evaporation rate that is expected from the
former, given the higher saturation pressure of methanol compared to
that of the latter. This is an interesting perspective in order to
overcome the mentioned evaporation difficulties experienced by
water injection at low rpm conditions.
Despite these many advantages, the substitution of a portion of water
with methanol leads to some drawbacks: higher performance implies
an increase of in-cylinder pressure level, potentially losing the
mentioned anti-knock benefits. In addition, methanol is an energy
source itself and a customer cost, so it has to be taken into account
for the calculation of specific consumption, which could increase
compared to the water-only case.
The purpose of this paper is still to substitute the fuel-enrichment of
the original engine as in [7], although a different strategy is evaluated
rather than the previously analyzed water-only injection. The target of
a global mixture stoichiometric ratio is maintained. However, the

Downloaded from SAE International by Alessandro D'Adamo, Wednesday, July 29, 2015

addition of a secondary fuel needs to be considered for a given


trapped air quantity, i.e. for a fixed oxidizer mass. Therefore, the
injected mass of methanol and water is calculated in order to maintain
a stoichiometric ratio and to provide the same charge cooling
supplied by fuel excess in the reference fuel-only rich case.
Three blends of methanol and water are considered for the presented
analyses: MW25, MW50 and MW75, i.e. 25%, 50% and 75% of
methanol mass fraction. Despite gasoline is direct injected, in order to
carry out a realistic evaluation blends are injected in the intake port as
in Figure 2, so that the cylinder dome is not modified, as for the
water-only case in [7].

Methanol/Water Ratio Condition


In this work, the MW blend composition XMW is expressed by the
mass fraction of methanol in the mixture as in Eq. 1.

(1)

with mM and mH2O the mass of methanol and water, respectively.


Their sum is the injected MW blend mass mMW.

Cooling Condition
The benefit in heat subtraction given by the fuel excess in the original
fuel-rich mixture must be substituted by the evaporation of the MW
blend. This is expressed by Eq. 2, where Qev represents the removed
heat, while hv,G, hv,H2O and hv,M are the latent heat of evaporation of
gasoline, water and methanol, respectively.

(2)

Figure 2. Coordinate systems which represent injectors for port MW injection


and gasoline direct injection.

Water-methanol mixtures are injected by means of a simple lowpressure injection system whose characteristics are derived from
standard low pressure PFI gasoline. Unfortunately no experimental
data regarding water-methanol injection are available to the authors
on the investigated engine, although all the physical models adopted
are widely used in similar analyses. In this framework, the lowpressure injector in a region far away from the cylinder helps to
reduce uncertainties related to the spray.
The injected amount of mixture is calculated following three
conditions:
a.

Methanol/Water Ratio Condition: the composition of the


injected methanol-water mixture;

b.

Cooling Condition: the injected MW blend has to provide the


same cooling that was subtracted by the removed fuel excess
that was present in the original fuel-rich mixture;

c.

Air Consumption Condition: the mixture of air, gasoline and


methanol is at stoichiometry, in order to completely consume
the trapped air in the combustion chamber.

These conditions are imposed in order to compare the results with the
original fuel-only case.

In Eq. 2 the excess gasoline mG,ex is calculated as the difference


between the injected gasoline mG,O in the original rich mixture and
the current mG, as in Eq. 3.

(3)

Substituting Eq. 3 in Eq. 2, the relationship in Eq. 4 is the Cooling


Condition:

(4)

Values for latent heat of the involved components that are used in the
CFD analyses that will be presented are reported in Table 1.
Table 1. Latent heats of vaporization [26] of injected components used in the
analytical conditions.

Finally, it is important to underline that the cooling condition is


obtained by a simplified approach. In fact this relationship only
considers the effect of the heat subtraction by evaporation. Therefore,
the variation in the mixture specific heat that also contributes to a
different heating due to piston compression is neglected. However,
given the preliminary status of the analysis, this is considered as an
acceptable simplification.

Downloaded from SAE International by Alessandro D'Adamo, Wednesday, July 29, 2015

Air Consumption Condition


The stoichiometric air-to-fuel ratio of gasoline, st,G, in this work is
equal to 14.8. The following relationship holds between the gasoline
and air mass, mG and mA respectively at stoichiometry:

(5)

For a dual-fuel engine, i.e. gasoline and methanol (st,M =6.0), this
becomes (Eq. 6):

(6)

From Eq.6, an expression for methanol mass mM can be obtained


(Eq.7)

(7)

Figure 3. Computational grid (top) and spark plug region detail (bottom).

Injection Modelling

Calculation of Gasoline and MW quantities


Finally, it is possible to solve a linear system of the three conditions
for the three unknown terms mM, mH2O and mG. Considering Eq. 1,
Eq. 4 and Eq. 7, for a selected XMW blend, the solution of the linear
system allows to calculate the three variables mM, mH2O and mG.
through Eq. 8, Eq. 9 and Eq. 10.

(8)

(9)

(10)

3D CFD simulations
The 3D CFD analyses presented in this paper are carried out by
means of a customized version of Star-CD v4.22, licensed by
CD-adapco. Time varying pressure and temperature boundary
conditions (the same for all cases) derive from a tuned 1-D model
supplied by engine manufacturer. The same is valid for wall
temperatures, which are set to cycle-averaged/component-specific
values. The adopted turbulence model is the k- RNG for
compressible flows. The computational mesh used in the 3D
simulations covers half of the combustion system, thanks to the
geometrical symmetry and the RANS framework used in this study.
The grid is represented in Figure 3, together with a close-up of the
spark plug region with the fully meshed ground J-electrode. The total
number of fluid cells is about 620000 and 325000 at BDC and TDC,
respectively.

For the high-pressure fuel injection simulation, a pre-atomized


population of Lagrangian particles is assigned to each of the 7
injector nozzles. The secondary break-up is modelled by means of the
Reitz model [8] while for droplet wall interaction the Bai model [9] is
adopted. No liquid film model is used in the simulations as, despite
the wall guided configuration, the high-temperature full load peak
power operation is simulated. Preliminary evaluations clearly showed
that no liquid film is formed due to the high piston crown temperature
for such condition, hence the model is not used in the analyses in
order to save CPU cost. Vapor fuel is modelled as C7H13 with
temperature dependent physical properties provided by the fuel
supplier. For the simulation of the injected water-methanol solutions
an analogous strategy is used for both primary and secondary
breakup, while injection pressure for liquid water methanol solutions
is limited to 5 bar. This value is chosen in agreement with
experimental datasets found in literature for PFI systems [10, 11].
Water-methanol solutions used in the analyses are three, namely
MW25, MW50 and MW75 with increasing methanol ratios (25%,
50%,75% by mass, respectively) modelled as multicomponent liquid
with temperature dependent physical properties derived from the
NIST database. As known, due to the polar nature of water, it attracts
the OH molecular group, while it repels the carbon chain as nonpolar.
Solubility of alcohols is therefore determined by the strongest of the
two forces; because of the attraction of the OH group, methanol is
completely miscible in water in any amount. For miscible liquid
multicomponent droplets, the evaporation of each component
depends on the relative presence of the other components because its
vapor pressure is a function of the concentration in the mixture. To
account for these effects, the Raoult's law could be applied, although
it provides reasonable predictions only if the molecular structure of
the liquid mixture components are similar. In order to accurately
predict the behavior of such complex mixtures, the Raoult's law can
be improved by introducing the activity coefficient concept as in Eq.
11. The partial pressure of the i-th component at the liquid-vapor
interface is expressed by pvi,s, Xi,l is the component mole fraction in

Downloaded from SAE International by Alessandro D'Adamo, Wednesday, July 29, 2015

the liquid,
the saturation pressure of the pure substance that
constitutes the i-th component and i is the activity coefficient of the
i-th component.

(11)

Activity coefficients account for the interaction between the different


components in a mixture and are simply set to unity for Raoult's law.
Since the vapor pressure at the liquid surface is the driving force for
the evaporation process, their accurate calculation is essential for a
detailed modelling and therefore they play a central role in the
presented cases. The UNIFAC (Unique Functional-group Activity
Coefficients) method [12] turns out to be a simple and accurate way
for the calculation of the activity coefficients. The basic idea behind
UNIFAC is to regard a molecule as an aggregate of functional groups
and to assume that certain thermodynamic properties can be
calculated by summing the group contributions. Model parameters
are obtained from UNIFAC database. It has been however observed
in [13] that for infinite dilution or for systems with components that
are very different in size, the aforementioned UNIFAC model often
gives poor results, instead better results can be obtain through the
adoption of a modified UNIFAC model [13] used by the authors to
model the completely miscible water-methanol solutions.

Combustion Modelling
The adopted combustion model is the ECFM-3Z [14], which has
been widely used in previous publications by the authors [15, 16, 17]
and whose capability to simulate partially premixed as well as
autoigniting combustion realizations is fundamental for such an
engine. As for spark ignition, a relatively simple algebraic model
based on a Flame Surface Density (FSD) deposition after a kernel
growth delay is used [18]. Following a user-coded approach
purposely developed within the research group, during the cold
portion of the engine cycle (i.e., before spark discharge), gasoline and
methanol are separately handled, i.e. each injected liquid evaporates
in a dedicated transported scalar representing its vapor concentration.
For the two vapors, temperature dependent properties from the NIST
database are used.

Figure 4. Outline of the gasoline+methanol combustion modelling strategy.

In order to account for multi-fuel combustion, at spark onset the two


mixed fuels are lumped into a unique scalar, whose properties are
those of an equivalent fuel blend. The same strategy is adopted for
the two unmixed fuels of the ECFM-3Z formalism. The process is
highlighted in the flowchart showed in Figure 4 for the sake of clarity.
Given the low percentage of methanol (which always represents less
than 5% of the overall fuel mass) and considering its almost
homogeneous distribution an assumption of linear-by-volume
properties are assumed for the equivalent fuel. These are computed as
the mass-weighted average of the properties of each fuel component,
as in Eq. 12 for lower heating value:

(12)

In Eq. 12 LHV are the lower heating values of gasoline, methanol


and equivalent fuel. The same approach is adopted for the evaluation
of the equivalent fuel temperature dependent specific heat and of the
air-to-fuel stoichiometric ratio, which must be recomputed given the
largely different stoichiometric ratios of the two fuel components.
The st for the equivalent fuel is given as in Eq. 13:

(13)

The one-step reaction used by the combustion model is then redefined


by changing the carbon-to-hydrogen ratio in order to comply with the
computed st.
Finally, the laminar flame speed of the equivalent fuel is also
modified accordingly, using the same mass average concept and the
correlations for pure iso-octane and pure methanol proposed by
Metghalchi and Keck [19]. This simplification is considered
acceptable thanks to the very low methanol level in the fuel mixture.

Knock Modelling and Knock Index


An in-house procedure is applied in this paper also for the modelling
of knock. The methodology is extensively described in previous
publications [20, 21, 22] and it is briefly recalled here in its
fundamentals for the sake of brevity and comprehension. It
constitutes an internal development of a relatively simple yet
effective knock model proposed by Lafossas et al. [23]. Autoignition
is cell-wise predicted in the CFD code by a 2-species model: one is a
knock precursor (YIG), the second is the fuel tracer as a reference
concentration (YTF). The autoignition delay is locally calculated at
each time-step, as it will be later described. Based on the resulting ,
the YIG species is increased until the YTF value is reached. When this
condition is met, knock is locally triggered. This technique allows a
cell-wise prediction of knock occurrence with a CPU-saving
approach based on just two species (transport equations for YIG and
YTF only are solved), given the availability of an accurate AI delay
from chemical kinetics simulations.

Downloaded from SAE International by Alessandro D'Adamo, Wednesday, July 29, 2015

As far as the calculation of the autoignition delay is concerned, a set


of look-up tables is created based on a semi-detailed chemical
mechanism supplied by the fuel manufacturer. The mechanism is the
138-species TRF (Toluene Reference Fuel) scheme by Andrae et al.
[24]. It was originally calibrated for the actual anti-knock quality of
the RON-98 European gasoline used in the experimental campaign.
Small methanol quantities are progressively added to the main fuel to
account for its presence in the combustion chamber. Six different
indexed tables of autoignition delays are then extracted from constant
pressure reactors, ranging from 0% to 5% of methanol mass fraction,
since 5% is the maximum methanol concentration reached in the
unburnt mixture. Look-up tables are created with 1% discretization
step in the methanol mass fraction YM. For each look-up table, the
initial conditions in terms of p, T, and EGR are also discretely
varied in order to cover the entire design space of in-cylinder
conditions. The input for the knock model is a 5-variable vector =
(p, T, , EGR, YM) of the local and instantaneous physical state of
each fluid cell and the methanol concentration; a cell-wise delay is
then interpolated from the look-up tables. As aforementioned and in
coherence with the methodology presented in [7], water is considered
as an EGR species in the analyses.

As stated earlier, the engine is analyzed at peak power engine speed,


W.O.T. For this engine condition and for the fuel-only case, the CFD
models for each of the lagrangian spray, the combustion heat release
and the knock proximity were deeply investigated and calibrated by
the authors in previous publications [20, 21]. As visible in Figure 5,
the simulated in-cylinder pressure trace (dashed black) is well within
the measured range.
Starting from this baseline, the substitution of a portion of gasoline
with different MW blends is examined. For each of them, the
quantities of water and methanol, as well as the final quantity of
gasoline, are calculated based on the linear system presented before.
Injection phasing for the presented analyses are SOIGasoline= 294 CA
BTDCF and SOIMW= 470 CA BTDCF. Both values derive from an
optimization analysis outlined in [6] and they are not recalled in this
paper for the sake of brevity.

Since the investigated engine operation is defined by the engine


manufacturer as knock limited, the Knock Severity Index (KSI) as
proposed by Klimstra [25] is adopted to compare the analyzed cases
in terms of knock intensity. Such index is based on thermodynamics
analyses and easily measurable variables. The fundamental
hypothesis of the KSI is that at knock onset the residual unburnt
mixture is instantaneously consumed at constant volume and assumes
that knock-related engine damage is more serious if a large fraction
of residual fuel is still present at knock onset. The KSI is defined as in
Eq. 14:

(14)

being mf,KO the mass of residual fuel at knock onset, mf,tot the total
fuel mass, LHV the fuel lower heating value, R the universal gas
constant, cp the isobaric specific heat, pKO the in-cylinder pressure at
knock onset and TKO the end-gas temperature at knock onset.

Investigated Engine and Operating Conditions


The investigated engine is a currently produced 8 cylinder V-shaped
direct injection spark ignition (DISI) turbocharged unit, whose main
characteristics are reported in the Table 2.
Table 2. Main characteristics of the investigated engine.

Figure 5. In-cylinder pressure trace. RANS (dashed, black), experiments


(grey), experimental bounds (red).

Results
Combustion Analysis
The injection of a secondary fuel (i.e. methanol) leads to a reduction
of the available gasoline: this is due to the Air Consumption
Condition (Eq. 7), i.e. a portion of air is needed for the oxidation of
the methanol itself. The trapped air is assumed not to be altered when
moving from fuel only to MW cases. However, the Cooling
Condition (Eq. 4) dictates that the same heat as the excess fuel is
subtracted by the evaporation of the secondary injected blend: due to
the lower latent heat for methanol with respect to that of water, the
total injected mass is maximum for the MW75 mixture. The
calculated quantities for the analyzed MW blends are in Table 3.

Downloaded from SAE International by Alessandro D'Adamo, Wednesday, July 29, 2015
Table 3. Gasoline and Methanol-Water mixture injected quantities resulting by
cooling and air consumption conditions.

One of the mentioned benefits of methanol-enriched mixtures is the


increase in the evaporation rate. In Figure 6 the percentage of vapor
mass for the three analyzed MW blends is reported.
In Figures 7, 8, 9 the MW liquid phase is represented together with
the methanol concentration. In Figure 9 it is visible that the MW75
case is the one presenting the highest injected mass, as anticipated in
Table 3. The residual mass of liquid droplets in the cylinder is
estimated to be the lowest for MW75 case as well, thanks to the
higher evaporation rate given by methanol-biased composition.

Figure 6. Evaporated methanol-water mixture percentage for the three


different methanol/water ratio over Crank Angle.

Figure 7. MW PFI mixture spray and methanol mass concentration for MW25 case. Left at 170CA Right at 260 CA After SOI-MW (SOI-MW=470 CA BTDCF).

Figure 8. MW PFI mixture spray and methanol mass concentration for MW50 case. Left at 170CA Right at 260 CA After SOI-MW (SOI-MW=470 CA BTDCF).

Downloaded from SAE International by Alessandro D'Adamo, Wednesday, July 29, 2015

Figure 9. MW PFI mixture spray and methanol mass concentration for MW75 case. Left at 170CA Right at 260 CA After SOI-MW (SOI-MW=470 CA BTDCF)

Despite the different evaporation rate for the three MW blends at the
end of the compression stroke, negligible differences are estimated in
the average in-cylinder temperature between the cases. These are
reported in Figure 10. This is due to the procedure outlined in the
Methanol-water methodology section for the calculation of the fuel,
water and methanol mass, where the different specific heat due to
mixture composition variation was not considered. A further
improvement of the proposed methodology including this factor is
currently under development. However, from the point of view of the
end-gas reaction rate, this poses the MW cases in a worse condition
than the fuel-only case, hence they are considered valid for the
evaluation of the knock tendency.

As a consequence, the MW25 case is the one with the highest EGR
level, as reported in Figure 12. Combustion results for the same SA
are presented in Figure 13, and the MW75 case is the one with the
highest in-cylinder pressure throughout the power stroke. The
baseline SA is increased with respect to experimental KLSA in order
to see difference in terms of knock tendency. The best performance
given by the MW75 case is motivated by two main factors:

Reduced EGR quantity: the laminar flame speed is little affected


by the residual presence;

Higher combustion heat: for a given air quantity, methanol


combustion liberates more heat than gasoline. Despite the
reduced lower heating value of the first, which is referred to
the unity of fuel, this results is motivated by the constraint of
a fixed oxidizer mass. Therefore, from the point of view of the
global released heat, the progressive increase of methanol mass
is more beneficial than the gasoline removal, hence leading to a
net torque benefit.

Figure 10. In cylinder temperature during the second half of compression


stroke for the three MW cases in comparison with gasoline-only case.

For increasing methanol presence, a twofold effect is observed. On


one hand the gasoline consumption decreases, following the Air
Consumption Condition; on the other hand the Cooling Condition
needs a further increase in MW quantity to be injected to reach the
same mixture cooling. This is clarified by Figure 11, where the
portion of heat subtracted by each component (gasoline, water and
methanol) is measured for all the analyzed blends.

Figure 11. Percentage of heat subtracted by each liquid component (Gasoline,


Methanol and Water) for the three MW cases.

Downloaded from SAE International by Alessandro D'Adamo, Wednesday, July 29, 2015

Figure 12. In cylinder EGR (CO2 + H2O) mass for the three MW cases in
comparison with the gasoline-only case.
Figure 14. In cylinder Auto Ignition heat release rate for the three MW cases
whit same SA=15 CA BTDCF (SA gasoline-only case 12CA BTDCF).

Fuel Consumption Analysis


The Indicated Specific Fuel Consumption (ISFC) is evaluated for the
three MW blends. To this aim, the SA is independently varied for
each condition until the same KSI is measured as the fuel-only
reference case. Thanks to the reduced knock tendency for all the MW
blends with respect to the original case (reported in Figure 15),
significant SA increases can be evaluated. This is also beneficial in
order to recover the combustion speed slowdown deriving for
globally leaner and EGR-diluted mixtures.

Figure 13. In cylinder pressure for the three MW cases with same SA=15 CA
BTDCF (SA gasoline-only case 12CA BTDCF).

When the knock tendency is observed, an opposed trend is verified.


Results show that the less knock-prone cases are those with the lower
methanol presence. Reasons for this are:

Reduced end-gas reactivity: a beneficial anti-knock effect is


given by water-rich secondary injection, which is maximized for
these cases;

Reduced thermal loading: this is an opposite consequence of the


aforementioned higher combustion heat and peak pressure levels
given by methanol-rich cases.

It is worthwhile to remind that the present SA increase is limited by


the aforementioned higher temperature measured for the MW cases at
the end of the compression stroke.
The results from this analysis show that the MW25 case is the one
allowing the highest SA for the same knock intensity as indicated by
the KSI. It is remarkable to recall that the MW25 case is the one with
the lowest pressure trace, when the SA is fixed and equal for all the
cases (see Figure 13).

A confirmation of these observations is given in Figure 14, where the


thermal power generated by autoignition is reported for the three MW
blends. It is null for MW25 case (i.e. knock-free condition), while it
is moderate for MW50 case and much higher for the MW75
condition. Still, it is worthwhile to underline that the total heat
released by autoignition is negligible for all the three cases, so that all
the conditions can be considered as knock-safe.

Figure 15. Knock Severity Index over SA increase for the three MW cases in
comparison with gasoline-only case.

Downloaded from SAE International by Alessandro D'Adamo, Wednesday, July 29, 2015

In Figure 16 the resulting Indicated Mean Effective Pressure (IMEP)


is reported for all the evaluated SA for the three MW blends. For
reduced SA values the MW25 produces a lower IMEP than the
analogous MW75, although the first is also the safest from knock.
When SA is increased, a benefit in IMEP is measured for all the
blends. However, for a given KSI the cases with reduced methanol
presence (i.e. MW50 and MW25, respectively) are those with the
highest output performance.

However, the ISFC benefits reported in Figure 17 highlight the


remarkable benefits given by the adoption of MW blends. The most
efficient strategy is confirmed to be the one with the lowest methanol
presence, i.e. MW25. The ISFC calculation is carried out considering
both the gasoline and the methanol.
This results is motivated by the Air Consumption Condition. Since
the low value of methanol AFRst (equal to 6.0), the same oxidizer
quantity originally consumed by the gasoline-only reaction needs
approx. 2.5 times more methanol mass to be fully burnt. For this
reason the use of methanol-poor mixtures is advantageous both in
terms of ISFC and of overall MW mixture consumption.
Conversely, if the observation was focused on the gasoline only the
best ISFC consumption would be given by the MW75 blend since it
is the one allowing the least gasoline use. However, since methanol
itself is an energy source and a customer cost, it is included in the
efficiency analysis.

Conclusions

Figure 16. Indicated Mean Effective Pressure over Knock Severity Index for
the three MW cases in comparison with gasoline-only case.

For the MW25 case, an increase of 6 CA in SA is possible for the


same knock-damage. This moves the operating point closer to the
Maximum Brake Torque Timing (MBTT), since IMEP gain is
progressively reduced for spark time advancements. This is due to the
lower benefit in terms of output torque as the peak pressure phasing is
anticipated towards the TDC, thus further limiting spark advance
increase.

In the paper, the potential of port-injected methanol-water mixtures is


numerically evaluated as knock suppressor on a currently produced
highly-turbocharged DISI engine, whose knock tendency was
extensively studied by the authors in the past years. Three different
MW mixtures with different methanol-water ratios are injected at
knock limited, peak power operating condition (7000 rpm) originally
characterized by large fuel enrichment to limit knock tendency.
The objective of the numerical analysis is to replace the fuelenrichment with methanol-water mixtures able to simultaneously
subtract the same amount of heat as the removed fuel, and to
completely consume the air mass trapped, i.e. the resulting mixture is
stoichiometric. Three blends of methanol and water are considered
for the presented analyses: MW25, MW50 and MW75, i.e. 25%, 50%
and 75% of methanol mass fraction.
For increasing methanol ratios, the evaporation rate benefits of the
higher vapor pressure with respect to that of water. This is an
interesting perspective in order to overcome the evaporation
difficulties experienced by water injection at low rpm conditions and
further analyses are currently in progress.
If the engine is operated at the same spark timing as the experiments,
the MW75 blend is the case resulting in the highest output torque,
since the different AFRst between gasoline and methanol leads to an
increase in the heat released by combustion.

Figure 17. Indicated Specific Fuel Consumption for the three MW cases in
comparison with the gasoline-only case.

However, when the SA is increased, the gain in specific fuel


consumption and torque increase emerges more relevant for the
MW25 case for the same knock intensity, as resumed by Figure 18.
This is motivated by the burn rate for the MW50 and MW75 cases,
which moves the peak pressure phasing closer to the TDC for
relevant SA advances and resulting in proportionally little IMEP
increase, and to the higher knock resistance given by the waterenriched end-gases.
Given the knock tendency reduction for MW blends, a different
strategy rather than SA increase could be the increase of the boost
pressure, in order to benefit of a higher trapped mass.

Downloaded from SAE International by Alessandro D'Adamo, Wednesday, July 29, 2015

9.

Bai, C. and Gosman, A., Development of Methodology for


Spray Impingement Simulation, SAE Technical Paper 950283,
1995, doi:10.4271/950283.

10. Bianchi, G., Brusiani, F., Postrioti, L., Grimaldi, C. et al., CFD
Analysis of Injection Timing Influence on Mixture Preparation
in a PFI Motorcycle Engine, SAE Technical Paper 2006-320022, 2006, doi:10.4271/2006-32-0022.
11. Bianchi, G., Brusiani, F., Postrioti, L., Grimaldi, C. et al., CFD
Analysis of Injection Timing and Injector Geometry Influences
on Mixture Preparation at Idle in a PFI Motorcycle Engine,
SAE Technical Paper 2007-24-0041, 2007, doi:10.4271/200724-0041.
12. Reid, R.C., Prausnitz, J.M. and Poling, B.E. The Properties of
Gases and Liquids. 4th edition, McGraw-Hill, Inc., New York
(1987).
13. Gmehling, J., Li, J. and Schiller, A modified UNIFAC
model. 2. present parameter matrix and results for different
thermodynamic properties, Ind. Eng. Chem. Res. 32, pp.178193 (1993).
Figure 18. IMEP and BSFC for the three analysed MW blends.

References

14. Colin, O., Benkenida, A., The 3-Zones Extended Coherent


Flame Model (ECFM3Z) for Computing Premixed/Diffusion
Combustion, Oil & Gas Science and Technology - Rev. IFP,
Vol. 59 (2004), No. 6, pp. 593-609.

1.

Kalghatgi, G., Auto-Ignition Quality of Practical Fuels


and Implications for Fuel Requirements of Future SI and
HCCI Engines, SAE Technical Paper 2005-01-0239, 2005,
doi:10.4271/2005-01-0239.

15. Fontanesi, S., Paltrinieri, S., D'Adamo, A., Cantore, G. et al.,


Knock Tendency Prediction in a High Performance Engine
Using LES and Tabulated Chemistry, SAE Int. J. Fuels Lubr.
6(1):98-118, 2013, doi:10.4271/2013-01-1082.

2.

Attard, W., Toulson, E., Watson, H., and Hamori, F., Abnormal
Combustion including Mega Knock in a 60% Downsized
Highly Turbocharged PFI Engine, SAE Technical Paper 201001-1456, 2010, doi:10.4271/2010-01-1456.

16. Fontanesi, S., Cicalese, G., d'Adamo, A., Cantore, G., A


Methodology to Improve Knock Tendency Prediction in High
Performance Engines, Energy Procedia, Volume 45, 2014,
Pages 769-778.

3.

Lanzafame, R., Water Injection Effects In A Single-Cylinder


CFR Engine, SAE Technical Paper 1999-01-0568, 1999,
doi:10.4271/1999-01-0568.

4.

Tauzia, X., Maiboom, A., Shas, S., Experimental study of inlet


manifold water injection on combustion and emissions of an
automotive direct injection Diesel engine, Energy 35 (2010)
3628-3639.

17. Fontanesi, S., Cicalese, G., Cantore, G., and D'Adamo, A.,
Integrated In-Cylinder/CHT Analysis for the Prediction of
Abnormal Combustion Occurrence in Gasoline Engines, SAE
Technical Paper 2014-01-1151, 2014, doi:10.4271/2014-011151.

5.

Tesfa, B., Mishra, R., Gu, F., Ball, A.D., Water injection effects
on the performance and emission characteristics of a CI engine
operating with biodiesel, Renewable Energy 37 (2012) 333344.

6.

Berni, F., Breda, S., Lugli, M, Cantore, G., A numerical


investigation on the potentials of water injection to increase
knock resistance and reduce fuel consumption in highly
downsized GDI engines, 69th Conference of the Italian
Thermal Machines Engineering Association, ATI2014.

7.

d'Adamo, A., Berni, F., Breda, S., Lugli, M. et al., A Numerical


Investigation on the Potentials of Water Injection as a Fuel
Efficiency Enhancer in Highly Downsized GDI Engines, SAE
Technical Paper 2015-01-0393, 2015, doi:10.4271/2015-010393.

8.

Reitz, R. and Diwakar, R., Effect of Drop Breakup


on Fuel Sprays, SAE Technical Paper 860469, 1986,
doi:10.4271/860469.

18. Boudier, P., Henriot, S., Poinsot, T., Baritaud, T., A Model for
Turbulent Flame Ignition and Propagation in Spark Ignition
Engines, Twenty-Fourth Symposium (International) on
Combustion/The Combustion Institute, 1992/pp. 503-510.
19. Metghalchi, M., Keck, J.C., Burning Velocities of Mixtures of
Air with Methanol, Isooctane, and Indolene at High Pressure
and Temperature, Combustion And Flame 48:191-210 (1982).
20. Fontanesi, S., Paltrinieri, S., D'Adamo, A., Cantore, G. et al.,
Knock Tendency Prediction in a High Performance Engine
Using LES and Tabulated Chemistry, SAE Int. J. Fuels Lubr.
6(1):98-118, 2013, doi:10.4271/2013-01-1082.
21. Fontanesi, S., Cicalese, G., d'Adamo, A., Cantore, G., A
Methodology to Improve Knock Tendency Prediction in High
Performance Engines, Energy Procedia, Volume 45, 2014,
Pages 769-778.
22. Fontanesi, S., Cicalese, G., Cantore, G., and D'Adamo, A.,
Integrated In-Cylinder/CHT Analysis for the Prediction of
Abnormal Combustion Occurrence in Gasoline Engines, SAE
Technical Paper 2014-01-1151, 2014, doi:10.4271/2014-011151.

Downloaded from SAE International by Alessandro D'Adamo, Wednesday, July 29, 2015

23. Lafossas, F., Castagne, M., Dumas, J., and Henriot, S.,
Development and Validation of a Knock Model in Spark
Ignition Engines Using a CFD code, SAE Technical Paper
2002-01-2701, 2002, doi:10.4271/2002-01-2701.
24. Andrae, J., Brinck, T. and Kalghatgi, G.T. HCCI experiments
with toluene reference fuels modeled by a semi-detailed
chemical kinetic model, Combustion and Flame, vol 155, pp
696-712, 2008.
25. Klimstra, J., The Knock Severity Index - A Proposal for a
Knock Classification Method, SAE Technical Paper 841335,
1984, doi:10.4271/841335.
26. Yaws, Carl L., Thermophysical properties of chemicals and
hydrocarbons, William Andrew Inc., Norwich, NY (2008)

Definitions/Abbreviations
AI - Autoignition
AFRst - Air-to-Fuel Ratio at Stoichiometry
aTDC - After Top Dead Center
BTDCF - Before Top Dead Center (Firing)
CA - Crank Angle
CI - Compression Ignition
DISI - Direct Injection Spark Ignition
EGR - Exhaust Gas Recirculation
FSD - Flame Surface Density
IMEP - Indicated Mean Effective Pressure
ISFC - Indicated Specific Fuel Consumption
KLSA - Knock Limited Spark Advance
LHV - Lower Heating Value
MFB - Mass Fraction Burnt
MW - Methanol-Water
SA - Spark Advance
SI - Spark Ignition
SOI - Start of Injection
TRF - Toluene Reference Fuel
WOT - Wide-Open Throttle

The Engineering Meetings Board has approved this paper for publication. It has successfully completed SAEs peer review process under the supervision of the session organizer. The process
requires a minimum of three (3) reviews by industry experts.
All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or
otherwise, without the prior written permission of SAE International.
Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE International. The author is solely responsible for the content of the paper.
ISSN 0148-7191
http://papers.sae.org/2015-24-2499

Вам также может понравиться