Anda di halaman 1dari 25

21st Aachen Colloquium Automobile and Engine Technology 2012 1

Low cost wet ethanol fuel: benefits and


challenges
Thompson D. Metzka Lanzanova, B. Eng., Dr. Horácio Antonio Vielmo , Federal
University of Rio Grande do Sul
Mr. Rafael Sari, Mario Eduardo Santos Martins Phd., Dr. Paulo Romeu Moreira
Machado, Federal University of Santa Maria

Summary

The search for renewable fuels, following the concept of “green energy”, and the
rising costs of oil has stimulated the search for cheaper and less polluting fuels. In
this situation, ethanol stands out and is widely used in various countries and can be
obtained from several sources. Considering the exponential increase of energy
consumption in the ethanol distillation process from over 80% of purity, the use of
ethanol with higher water content would lead to a low cost fuel. This paper evaluates
the efficiency parameters of a spark ignition engine fuelled with ethanol-in-water
mixtures through computational simulation. The engine used in this paper is a Diesel
modified mono-cylinder 0.668-L engine operating with several percentages of
ethanol-in-water mixtures. A detailed combustion analysis is carried out. Performance
and emissions are evaluated to assess the implications of using such a high water
fuel, presenting a summary of its benefits and challenges.

1 Introduction

Recent advances in internal combustion engines have been motivated by emission


legislation aiming at reducing greenhouse gases and pollutant emissions. In the
search for renewable and clean energy the “green energy” concept is standing out.
Ethanol, or ethyl alcohol, usually produced from cassava, beetroot, sugarcane and
corn fermentation is a biofuel extensively produced worldwide to complement and
even substitute fossil fuels.
In the ethanol life cycle, GHG (greenhouse gases) are reduced and energy balance
is positive when compared to fossil fuels [1],[2]. Walter et al studied the GHG
emissions of Brazilian gasohol (25% anhydrous ethanol and 75% gasoline) [3].
Comparing to pure gasoline, the emissions drop to 78%. According to the Directive
2009/30/EC, in the sugarcane ethanol production pathway the saving in typical GHG
emissions is 71% when compared to oil and gas combustion and total gCO2 eq/MJ
emission in around 24 [4]. This reduction can be attributed to photosynthesis of
biomass feedstock which fixes carbon dioxide during its growth phase. According to
Seabra et al (2010), biochemical conversion of ligno-cellulosic materials to ethanol is
a promising way to drastically reduce the energy spent in a near future, consequently
reducing GHG emission [5] .
Brazil sugarcane is one of the cheapest bioethanol feedstock [6]. The expansive
territory, geographical position, abundant water resources, and solar radiation are
2 21st Aachen Colloquium Automobile and Engine Technology 2012

some of the advantages presented in this country which for more than 30 years has
been investing in sugarcane ethanol production. In 2007, estimated production was
19 billion liters of ethanol, being the production similar to that of corn ethanol in the
USA [7].
The most commonly bioethanol production methods are basically constituted by
feedstock formation, mashing and cooking, fermentation, distillation and dehydration.
After fermentation phase, ethanol-in-water content conventionally varies from 6% to
12% [8]. As ethanol and water is an azeotropic mixture completely miscible,
distillation can separate ethanol from water until ethanol-in-water content of 95.57%.
To obtain even higher ethanol-in-water contents, dehydration through adsorption in
vapor phase with molecular sieves can be done. Recent works [9][10] show that a
large amount of energy is spent to extract ethanol from water in the distillation and
dehydration phases, and the use of 65% ethanol-in-water can reduce the net energy
spent from 37% (anhydrous ethanol) to 3% - for corn ethanol. Fig. 1 shows the net
energy gain considering sugarcane anhydrous ethanol (a) and 20% water-in-ethanol.
Fig. 2 shows the productions costs in each production phase to reach anhydrous
ethanol.
Transport
1.27 % Transport
Production Milling 2.01%
Net energy 7.17 % 2.95 % Production
gain 9.7 % 11.36 % Milling;
Fermentati 4.68 %
on 0.84 %
Fermentat
ion 1.34
%
Distilation,
19.83 %
Net
Hydrolysis/ energy
broth Distilation
54.85 % 25.77 %
treatment
58.23 %

(a) (b)

Fig. 1 Net energy balance in (a) Sugar cane anhydrous ethanol, (b) 20% of
hydration. The figure indicates energy consumption in all stages of the
Sugarcane ethanol production as percent of the heating values in MJ.L -1. [11]
21st Aachen Colloquium Automobile and Engine Technology 2012 3

Sugarcane ethanol production Energy production Energy production Energy Cost


phase cost (MJ/L) fraction(%) (€/litre)

Plantation 1.7 7.94 0.02

Transport 0.3 1.40 0.004

Distilation 4.7 21.96 0.06

Hydrolysis/saccharification/broth 13.8 64.49 0.18


treatment

Milling 0.7 3.27 0.01

Fermentation/maintenance 0.2 0.93 0.003

Total 21.4 100.00 0.28

Fig. 2 Production costs of sugarcane ethanol.

In the ethanol distillation, the required energy increases exponentially to obtain


ethanol-in-water proportions above 80% in volume, leading to higher production
costs. Increasing the water volume fraction in the mixture, the lower heat of
combustion (LHC) of the mixture is degraded. But, as the LHC and production costs
behaviours differs, there is an optimum point where the relation between the energy
dispended in production and the LHC lead to the best cost of energy unit. To obtain
the net energy the cost to produce the ethanol-in-water mixture is discounted from
the total energy released by combustion. The Fig. 3 shows the cost of energy unit
in the production process of sugarcane ethanol. Can be noticed that minimal cost of
energy is achieved when ethanol-in-water concentration reaches around 80%, after
that point, mainly due to distillation process, the cost of energy grows exponentially.

0.14
Cost /Net energy ratio
0.12
Cost/Energy (€/MJ)

0.1
0.08
0.06
0.04
0.02
0
0.5 0.6 0.7 0.8 0.9 1
Ethanol fraction

Fig. 3 Cost /Net energy ratio. The cost for obtain the net energy in € /MJ.
4 21st Aachen Colloquium Automobile and Engine Technology 2012

There is a wide range of applications for ethanol as fuel and additive in internal
combustion engines. Tests using fuels containing 20% ethanol-in-diesel
demonstrated a very slow reduction in brake specific power at low charges and no
significant changes at medium and high charges [12]. Reductions in BSNOx and
submicron particles emissions can be noticed, mainly in, at the same time BSHC and
BSCO increase [13].
The water injection is also used to control hydrogen knock by reducing the exhaust
gas temperature. For the same reason, is very effective to control NOx emissions,
but an increase in HC is verified [14]. Nguyen and Wu, (2009) reported an
improvement in the thermal efficiency when the water-gasoline emulsions are used
but decreases the lean mixture due to dilution of charge [15]. Using a manifold water
injection on a SI engine running with liquefied petroleum gas, the water addiction
reduced the temperatures of charge and exhaust gases. The NOx emissions and
rate of fuel burning is reduced and engine thermal efficiency increases [16].
Ethanol is also used as anti-knock additive to gasoline, replacing more costly anti-
knocking additives. For the oxygenated nature of ethanol, ethanol gasoline blends
burns more cleanly reducing unburned hydrocarbon (HC), carbon monoxide (CO)
and particulates. Studies conducted in a spark ignition (SI) engine running with high
water content ethanol and gasoline blends shows that there is a limit of gasoline in
the mixtures due the miscibility issues [17].
It was found that the use of ethanol with high water content is somewhat difficult due
to flame degradation and consequently decrease in burning velocities. These facts
can be overcome with the advance of combustion techniques such as CAI and HCCI.
Due the nature of CAI combustion, the fuel can be burned in adverse conditions
where the SI combustion is impossible, showing good efficiency using highly
hydrated alcohols [18].In this way, recent studies shown that HCCI engines can
operate with water-in-ethanol percentages up to 40% and achieve brake engine
efficiencies around 40% [9][19][20].
Cordon [21] investigated the use of catalytic igniter to burn lean ethanol-water-air
mixtures in a 15 kW Yanmar 3-cylinder engine converted to operate with the
combustion starting with HCCI in a catalytic pre-chamber, followed by torch ignition in
the main chamber. Reduce in pollutants emissions and increase in power output
compared with the original diesel engine were achieved.
In the HCCI combustion, in a narrow range, the ignition timing and the rate of
combustion can be controlled by water injection with a low pressure fuel injector. This
was proven for Christensen et al utilizing a Volvo Td100 series Diesel 6-cylinder in
line engine [22]. The inlet air was preheated with an electrical heater to achieve high
enough temperature to propitiate the mixture auto ignition. Aiming at improve the
energy balance in favor of ethanol, studies were conduced using HCCI combustion in
a Volkswagen 1.9 L 4-cylinder engine running with anhydrous ethanol and wet
ethanol with 90%, 80%, 60 % and 40% ethanol-in-water mixtures controlling the auto
ignition temperature with an electrical heater and thus the timing of ignition. Stable
HCCI operation for mixtures with up 40% of water were achieved [0] .
21st Aachen Colloquium Automobile and Engine Technology 2012 5

This paper aims at studying the effects of using water-in-ethanol mixtures as fuel for
internal combustion engines. Indicated parameters and pollutant emissions are
compared for different water in ethanol contents. For evaluation porpoise, operation
cost is compared to conventional diesel and gasoline.

Fig. 4 Model interface snapshot.

2 Engine Modelling Process

For the present study, a commercial engine simulation software GT-Power® has been
used in order to provide full thermodynamic cycle analysis. GT-Power® is one-
dimensional engine gas dynamic largely used by car and engine manufacturers
[23][24][25]. To evaluate the viability of high content water in ethanol as fuel for
internal combustion engines, spark ignition (SI) and homogeneous charge
compression ignition (HCCI) operation modes were chosen. While in SI engines the
start of combustion can be controlled by the spark timing, in the HCCI engine the
start of the combustion process happens when the charge reaches the auto ignition
temperature through compression. This mechanism is high chemical kinetic depend,
being almost impossible to exactly impose start of combustion.
To make possible auto ignition temperatures to be reached inside the cylinder before
the top dead center with no spark discharging, charge pre-heating can be
considered. High compression ratios could also permit to reach auto ignition
6 21st Aachen Colloquium Automobile and Engine Technology 2012

temperatures. Considering the hypothetical naturally aspirated spark ignition engine


used for this study (characteristics shown in Fig. 5), compression ratio is increased
from 12:1 to 16:1 and exhaust heat recovery is implemented to enable auto ignition
temperatures to be reached, enabling HCCI operation without necessity of external
heat source. Fig. 4 shows a snapshot of the GT-Power® model interface.

Bore Stroke Conecting rod Compression Displaced


Strokes
(mm) (mm) Length (mm) ratio volume (dm³)

4 90 105 160 12/16 0.668

Fig. 5 Engine characteristics.

Combustion in the SI operation is simulated through a turbulent flame speed model


(CombSITurb), which is well accepted and described in several articles [26][27][28].
The model simulates flame propagation as a spherically growing region with the
origin in the spark plug. As the sphere grows, its border intersects gases and so,
flame area is proportional to sphere area. Flame speed is calculated through an
entrainment model based in flow turbulence characteristics. During combustion
simulation, flow characteristics as swirl, tumble and turbulent kinetic intensity are
calculated through a spatially and time resolved flow model, which feeds the flame
speed model. The entrained mass rate of unburned gas is determined by the
entrained area 𝐴𝑒 , specific weight of the unburned gases 𝜌𝑢 and an entrainment
velocity determined as the sum of the turbulent and laminar flame speeds – 𝑆𝑡 and 𝑆𝑙 .
Equations can be expressed as:

𝑑𝑀𝑒
= 𝜌𝑢 𝐴𝑒 (𝑆𝑡 + 𝑆𝑙 )
𝑑𝑡
(1)

This way, entrainment model accounts for temperature, pressure and gas
composition, in the laminar burnup behind the flame front, and is assumed to be
proportional to the unburned mass behind the flame front.

𝑑𝑀𝑏 𝑀𝑒 − 𝑀𝑏
=
𝑑𝑡 𝜏
(2)

Where 𝑀𝑏 stands for burned mass and 𝑀𝑒 , entrained mass. As burnup takes place at
the flame front over a turbulence length scale 𝜆– Taylor microscale, calculated by the
in cylinder model–, at the laminar flame flame speed, the time to burnup an ignition
site of length 𝜆 can be expressed as:
21st Aachen Colloquium Automobile and Engine Technology 2012 7

𝜆
𝜏=
𝑆𝑙
(3)

𝑇𝑢 𝛼 𝑃 𝛽
𝑆𝑙 = 𝑆𝐿𝑜 𝐹𝑑 ( ) ( )
𝑇𝑜 𝑃𝑜
(4)

𝑆𝑙 is the laminar flame speed at 𝑇𝑢 and 𝑃, 𝑆𝐿𝑜 is the laminar flame speed at 𝑇𝑜 , 𝑃𝑜 . Also,
α and β, are experimental coefficients of the fuel and 𝐹𝑑 is the retardation factor due
to residual gases. The calculation of 𝑆𝑡 is much more dependent of the flow
characteristics, and won’t be described here, but can be said it accounts for early
flame development, where laminar flame speed dominates the combustion process,
with its contribution increased in the entrainment velocity after this period.
Considering coefficient 𝐹𝑑 in the software accounts only for burned residual gases,
and water addition to fuels reduce laminar flame speed new flame laminar flame
speeds of wet ethanol should be specified - 𝑆𝐿𝑜%𝐻2𝑂 (𝑇𝑜 , 𝑃, %𝐻2𝑂) [29][30]. To calculate
these velocities, Eisazadeh-Far et al formulation for laminar flame speed was used,
changing burned residual gases fraction to water fraction. This way, using
Eisazadeh-Far et al [31] correlation, laminar flame speed of ethanol at To and P, for
ϕ stoichiometric, as function of water fraction can be calculated as:
𝜃
𝑐𝑝𝐻2𝑂
𝑆𝐿𝑜%𝐻2𝑂 = 𝑆𝐿𝑜 (1 − 𝑥𝐻2𝑂 )
𝐶𝑝𝐸𝐺𝑅
θ θ
cpH2O
SLo%H2O = SLo (1-xH2O ) SLo%H2O = SLo (1-xH2O ) (5)
CpEGR

This way, laminar flame speed from software will be calculated as:

𝑇𝑢 𝛼 𝑃 𝛽
𝑆𝑙 = 𝑆𝐿𝑜%𝐻2𝑂 𝐹𝑑 ( ) ( )
𝑇𝑜 𝑃𝑜
(6)

Even though wet ethanol laminar flame speeds are calculated, this model does not
take in account flame quenching due to chemical kinetics and in some cases, where
high water in ethanol is used, flame quenching could make not possible real engine
operation under the same conditions.
Also, to evaluate high water in ethanol content as fuel for internal combustion
engines, HCCI operation was simulated through Wiebe functions, to cover the areas
where SI operation could not be achieved. The Wiebe functions were used to
represent heat release profiles typical of HCCI operation according to technical data
[9][18][32]. To reach HCCI operation, induced air was heated through a simulated
exhaust heat recovery strategy. As chemical kinetics could not be simulated, it is
assumed that the auto ignition condition is reached when charge reaches
temperatures around 1100 K. This temperature is slightly controlled through the
8 21st Aachen Colloquium Automobile and Engine Technology 2012

amount of energy added to induced air and the heat exchange was controlled to
deliver enough energy to make the charge reach around 1100 K at ~ 2 CA BTDC.
To find the best operation conditions, for SI and HCCI, maximum Ringing Intensity
(RI) was used, as a knock indicative parameter. Ringing intensity compares the
pressure rise rate to peak cylinder pressure and engine speed and therefore gives a
good indication of the combustion noise. Border of knock operation is assumed at
ringing intensity around 5 MW/m² [32][33]. Can be calculated as:
2
𝑑𝑃
1 (𝛽 )
𝑑𝑡 𝑚𝑎𝑥
𝑅𝑖𝑛𝑔𝑖𝑛𝑔 𝐼𝑛𝑡𝑒𝑛𝑠𝑖𝑡𝑦 = ( ) √𝛾𝑅𝑇𝑚𝑎𝑥
2𝛾 𝑃1
(7)

Table 3 shows equivalence air ratio ϕ used in each simulation. For all simulations,
intake and exhaust ambient pressure and temperatue is set to 100 kPa and 300 K.
Classic Woschni model was used to calculate in-cylinder heat transfer.

Exhaust heat
10-90 RI máx
Operation Combustion Compression recovery
Combustion PHI (kW/m²
mode model ratio variaton range
duration (CA) )
(kW)*
Calculated by
SI SITurb 12:1 1 5 0
the model
10 0.25 5 0-1
20 0.75 5 0-4
HCCI Wiebe 16:1
30 1 5 0-5
40 1 5 0-5
*Auto ignition heat source
Fig. 6 Operation conditions simulated

3 Simulation Results

3.1 Spark Ignited Simulation

The first part of this study is about spark-ignited operation in which a turbulent
combustion mode was used to predict the influence of increased water content on
engine performance, combustion and pollutant emissions. It was studied the use of
up to 40% water by volume on ethanol.

The first study was performed to find MBT spark timing for each water percentage. It
was established, however, a ringing intensity (RI) limit of 5 MW/m 2 above which the
engine was deemed as knocking.
21st Aachen Colloquium Automobile and Engine Technology 2012 9

As Fig. 7 shows, MBT ignition timing would yield IMEP results that would come up
with heavy knocking, as shown by the large values of ringing intensity, reaching up to
20 MW/m2. To reduce ringing intensity to values under 5MW/m2, hereby deemed as
the knock threshold, ignition timing had to be retarded and IMEP was obviously
reduced. It remained, however, varying from 1250 up to 1350kPa with the highest
water fraction, showing the interesting fact that if combustion can take place with
such amount of water in ethanol, the power figures would actually increase due to
better combustion phasing, as will become evident in the following discussions.

Fig. 7 Indicated Mean Effective Pressure and Ringing Intensity as a function of water
fraction, for MBT ignition timing and retarded ignition timing to keep RI under 5
MW/m2

Fig. 8 Volumetric and indicated efficiencies as a function of water fraction


10 21st Aachen Colloquium Automobile and Engine Technology 2012

It can be seen from Fig. 8 that volumetric efficiency has a slight increase with
increased water content, reaching a maximum at the fraction of 0.2 and decreasing
once again thereafter. The initial increase can be attributed to the increased charge
cooling caused by water evaporation, which compensates for the volume displaced
by water. This trend changes with higher water content because the water addition
starts to take a larger volume of the cylinder, leaving less space available to be filled
by the air/fuel mixture. The increased volumetric efficiency contributes to the higher
in-cylinder pressure found when using high water-in-ethanol content, as is evident
according to Fig. 9 .

Fig. 9 Cylinder pressure traces for 5% and 40% water-in-ethanol volume fraction

Indicated efficiency, on the other hand, increases with water content. This is due to
better combustion phasing, since with increased water content the engine is able to
run closer to MBT spark timing due to the anti-knock properties of high water content
mixtures. This is evident when one looks at Fig. 10, where the point of 50% mass
fraction burn is shown to be advanced as water fraction increases, also as a
consequence of longer combustion duration. Ignition delay, usually defined as the
period between 0-2% mass fraction burn, is shown to increase with the water
content, all in accordance with the literature, showing that the turbulent combustion
model was able to represent correctly the effects of water addition.
21st Aachen Colloquium Automobile and Engine Technology 2012 11

Fig. 10 Ignition delay (0-2%MFB), burn duration (10-90%MFB) and combustion


phasing (50%MFB) as a function of water fraction

In-cylinder pressure and temperature are affected by the water content of the fuel, as
Fig. 11 shows. The reduced in-cylinder temperatures are advantageous from the
point of view of NOx emissions, which can be substantially reduced, as Fig. 12
shows. This is a very positive aspect of high-water content mixtures since they
enable engine operation without the need of complicated NO x aftertreatment
systems.

CO emissions are in general strongly related to the local air/fuel ratio, being very
sensitive to mixture inhomogeneity. However, in the current spark-ignition model,
air/fuel ratio was always kept constant and the only source of variation in CO
concentration was the in-cylinder temperature history, which affected the CO reaction
rate constant, known to be dependent also on temperature [34].

Unburned fuel components such as HC emissions are not a common source of


pollutants from combustible blends with only alcohols as fuels. Instead, these are
better classified among the so called volatile organic compounds (VOC). In the case
of ethanol, the most important VOC emissions are those of acetaldehydes (C 2H4O),
which are linked to toxic and carcinogenic effects. The emissions models used in this
study are not able to predict VOC production and therefore will not be covered in this
study. It must be pointed out, however, that VOC’s are easily converted by simple
three-way catalyst systems, whereas NOx emissions are far more difficult and
troublesome to be treated and are the main interest in this study.
12 21st Aachen Colloquium Automobile and Engine Technology 2012

Fig. 11 In-cylinder maximum pressure (MPa) and temperature (K) as a function of


water fraction

Fig. 12 Pollutant emissions concentration (ppm) as a function of water volume


fraction
21st Aachen Colloquium Automobile and Engine Technology 2012 13

3.2 Homogeneous Charge Compression Ignition Simulation

The results shown in the above simulation with the turbulent combustion mode
assume that combustion would take place in all situations. Nevertheless, it is
expected that ignition would become more difficult and large cycle-to-cycle variations
would occur with high water fractions, possibly leading to misfire. In such a situation,
spark-ignition combustion might become difficult or maybe impossible due to its flame
propagating ignition nature. HCCI or CAI combustion could be a possible solution for
ignition in these adverse conditions, i.e., with high water dilution.

HCCI combustion was simulated by using a Wiebe function adapted to represent the
heat release rate typical of this combustion mode, for each of the 3 cases studies,
with equivalence ratios of =1, =0.75 and =0.25 and combustion durations of 30°,
20° and 10°, respectively.

Ringing intensity determined the upper load limit of operation whereas the lower limit
was set by the auto-ignition temperature. The minimum auto-ignition temperature
was set to be at least 1100K as this is a common value for ethanol combustion
according to the literature [35][36][37][38]. The original compression ratio would
enable maximum compression temperatures of only up to 800 K, below the minimum
temperature for enabling auto-ignition. Therefore, the compression ratio was
increased to 16:1 and additional intake heating was promoted by means of exhaust
heat recovery. This was accomplished placing a heat exchanger linking the intake
and exhaust manifolds.

Simulation was conducted for each equivalence ratio at the minimum combustion
durations that would yield ringing intensities under the proposed limit of 5 MW/m2
with enough heat recovery in order to reach the minimum auto-ignition temperatures.
The combustion duration values were rounded up to 10, 20 and 30°CA.

As ringing intensity is highly dependent on the maximum rate of pressure rise and
this parameter depends primarily on combustion duration and equivalence ratio, high
equivalence ratios (richer mixtures) could only be obtained with longer combustion
durations in order not to reach the knock limit (maximum ringing intensity).

Fig. 13 shows the maximum rate of pressure rise for the 3 case studies. These
values were deliberately controlled in order to keep ringing intensity low.
14 21st Aachen Colloquium Automobile and Engine Technology 2012

Fig. 13 Maximum Rate of Pressure Rise for the 3 case studies

Fig. 14 Indicated Mean Effective Pressure for HCCI operation for the 3 case studies

The indicated mean effective pressure for HCCI operation is shown in Fig. 14, in
which it can be seen that water fraction affects engine load by lowering IMEP as the
water fraction increases. This effect is more dependant, however, on equivalence
ratio as one could expect. With richer mixtures, more energy is available and more
power is delivered.

More activation energy for auto-ignition was necessary for richer mixtures due to a
higher specific heat and latent heat of vaporization of its contents (more water and
ethanol). This demanded higher exhaust heat recovery, as presented in Fig. 15.
21st Aachen Colloquium Automobile and Engine Technology 2012 15

Fig. 15 Exhaust recovered energy (kW)

Fig. 16 Volumetric Efficiency for the HCCI model and several combustion duration
values

In Fig. 16 it can be seen that the water fraction has little influence on volumetric
efficiency. The latter is more dependent on equivalence ratio, since with leaner
mixtures less air was displaced by the water content and less heat was needed from
exhaust recovery, having, therefore, less impact on charge density.
16 21st Aachen Colloquium Automobile and Engine Technology 2012

Fig. 17 Indicated efficency for different combustion durations and equivalence ratios
(PHI) for increasing water fractions.

As Error! Reference source not found. shows, increased water fraction has little
effect on indicated efficiency when HCCI combustion takes place with =1. This,
however, was achieved with combustion durations of 30°CA, in order to keep ringing
intensity below the knock threshold of 5 MW/m2. These values of burn duration are
not necessarily realistic for HCCI, a combustion mode known for being very fast.

Shorter burn durations were possible only for leaner mixtures such as =0.75 (burn
duration of 20°) and =0.25 (burn duration of 10°). It must be pointed out that
indicated efficiency was a combination of effects, namely the amount of energy
recovered from exhaust, changes in volumetric efficiency as a function of
equivalence ratio and combustion phasing influence due to different burn durations.
The ratio of specific heats (γ) also influences indicated efficiency as shown in Fig.
18, which would lead to higher indicated efficiency for =0.25.

This, nonetheless, is not the case since such a lean mixture is very sensitive to water
fraction because it needs a lot of energy to vaporize and impairs heat release. Also,
combustion phasing for =0.25 was not the best since it provided maximum
pressures away from the best location, i.e., around 13°CA, as shown in Fig. 19. Due
to all of these factors, the mixture with the highest indicated efficiency was found at
=0.75.
21st Aachen Colloquium Automobile and Engine Technology 2012 17

Fig. 18 Ratio of specific heats as a function of water fraction for the HCCI operation

Fig. 19 Crank angle at maximum pressure rise

Pollutant emissions have a different behaviour in the HCCI simulation also due to the
fact that it was run for non-stoichiometric mixtures. In the HCCI operation, NOx
emissions (Fig. 20) were overall reduced with increased water fraction. Again,
similarly to what happened in the SI simulation, this was caused by the reduction of
in-cylinder temperatures with increased water fraction. Values next to zero were
found, nonetheless, for the leanest case of =0.25, when in-cylinder temperatures
are overall very low, staying always under 1800K (Fig. 21), the temperature
threshold for nitrogen oxide formation.
18 21st Aachen Colloquium Automobile and Engine Technology 2012

Fig. 20 NOx emissions (ppm) for the HCCI operation

Fig. 21 In-cylinder temperature histories

CO emissions are strongly linked to air/fuel ratio (Fig. 22). This fact explains why for
=1 (stoichiometric mixture) values are much higher than with leaner mixtures
(=0.25 and 0.75). The general trend is to have lower values with increased water
fraction, as it happens also with =0.75. For =0.25 CO emissions are negligible due
to the ultra-lean condition.
21st Aachen Colloquium Automobile and Engine Technology 2012 19

Fig. 22 CO emissions (ppm) for the HCCI operation

4 Conclusions

This paper evaluated the use of water-ethanol mixtures as fuel for internal
combustion engines. Spark-Ignited (SI) and Homogeneous Charge Compression
Ignition (HCCI) operation modes were evaluated. Operation was only kept under the
knock limit defined by a Ringing Intensity (RI) parameter.

Spark-Ignited operation mode was simulated with a turbulent flame speed


combustion model. It could be noticed that as the water fraction increased, higher
IMEP values (from 1350 to 1450 MPa) were possible. CO emissions were slightly
lower whereas NOx emissions were dramatically reduced due to lower in-cylinder
temperatures. Indicated efficiency remained more or less constant at a value of 42%
regardless of changes in water fraction.

Simulation of HCCI operation was performed for 3 different situations: equivalence


ratio of =1, =0.75 and =0.25 and combustion durations of 30°, 20° and 10°,
respectively. It was noticed a slight reduction on IMEP with increased water fraction
for all cases. Emissions of NOx were overall much reduced in comparison to SI
operation and reduced steadily with increased water fraction when on HCCI
combustion. For one of the cases (=0.25) values were negligible due to very low in-
cylinder temperatures. CO emissions are highly dependent on air/fuel ratio and
followed this trend closely. However, for every equivalence ratio, values were lower
as water fraction increased.
20 21st Aachen Colloquium Automobile and Engine Technology 2012

Indicated efficiency with HCCI was shown to be higher than with SI operation for
leaner mixtures and reasonably the same when in stoichiometric operation (=1),
with values within 42-45%. Increasing water fraction did not show to have any
important effect on efficiency, with values remaining soundly constant.

All in all it is noticeable that the use of ethanol with high water fractions is a promising
alternative to enable the use of a low cost renewable fuel which offers good
improvement chances if its properties are properly explored.

5 Definitions, acronyms, abreviations

 Equivalence ratio
Ae Entrained area
Fd Retardation factor due to residual gases
Mb Burned mass
Me Entrained mass
Po Initial state Pressure
SLo%H2O Maximum laminar flame speed at T_o and P_o– water diluted
SLo Laminar flame speed at T_o, P_o
Sl Laminar flame speed
St Turbulent flame speed
To Initial state temperature
Tu Unburned zone temperature
dMe Rate of unburned gas
dt
dP Pressure rise rate
dt
xH2O Molar water fraction
ρu Specific weight of the unburned gases
BSCO Brake specific CO
BSNOx Brake specific NOx
CA Crank angle
CAI Controlled Auto-Ignition
Cp Specific heat for constant pressusre
Dur. Duration
EGR Exhaust gas recirculation
HC Hydrocarbons
21st Aachen Colloquium Automobile and Engine Technology 2012 21

HCCI Homogeneous charge compression ignition


IMEP Indicated medium effective pressure
K Kelvin
L Liter
MBT Maximum brake torque
MFB Mass fraction burned
Ppm parts per million
R Gas constant
RI Ringing Intensity
SI Spark ignition
SITurbcomb Spark ignition turbulent flame speed combustion model
T Temperature
Td Turbo Diesel
VOC Volatile organic compounds
α Experimental coefficient of the fuel
β Experimental coefficient of the fuel
θ Degree
P Instantaneous pressure
γ Polytropic coefficient
λ Turbulence length scale
τ Taylor microscale
S Sulfur
GHG Green house gases
LHC Lower heat of combustion

6 References

[1] USDA (United States Department of energy).


Ethanol: the complete life cycle
picture.2007.Access:http://www1.eere.energy.gov/vehiclesandfuels/pdfs/program/
ethanol_brochure_color.pdf
[2] Foteinis S., Kouloumpis V. and Tsoutsos T.,
Life cycle analysis for bioethanol production from sugar beet crops in Greece.
Energy Policy 39
Technical University of Crete, Greece2011
[3] Walter A., Dolzan P., Quilodra O., de Oliveira J. G., Silva C., Piacente F. and
Segerstedt A.
Sustainability assessment of bio-ethanol production in Brazil considering land use
change, GHG emissions and socio-economic aspects.
22 21st Aachen Colloquium Automobile and Engine Technology 2012

Energy Policy 39, 5703–5716.


University of Campinas, Brazil,2011
[4] The European Parliament and of the Council of April 2009.
Directive 2009/30/EC. Official Journal of the European Union
L 140/88.
(2009)
[5] Seabra J. E. A., Tao L., Chum H. L. and Macedo I. C.,
A techno-economic evaluation of the effects of centralized cellulosic ethanol and
co-products refinery options with sugarcane mill clustering.
Biomass and Bioenergy 34, 1065–1078.
National Renewable Energy Laboratory, USA, 2010
[6] Walker M. G.,
Bioethanol: Science and technology of fuel alcohol.
Graeme M. Walker & Ventus Publishing ApS
University of Albertay, Scotland, August 2010
[7] Martinelli L.A., Filoso S.,
Expansion of Sugarcane Ethanol Production in Brazil: Environmental and social
Challenges.
Ecological Applications18 885-898.
Piracicaba-SP, Brazil, 2008
[8] Ladisch, M. R. and Dyck, K.,
Dehydration of Ethanol: New Approach Gives Positive Energy Balance. Science
205, 878–900.
USA,1979
[9] Mack J. H., Aceves S. M. and Dibble R. W.,
Demonstrating direct use of wet ethanol in a homogeneous charge compression
ignition (HCCI) engine.
Energy 34 782–787
California, USA, 2009
[10] Martinez-Frias J., Aceves S. M. and Flowers D. L.,
Improving Ethanol Life Cycle Energy Efficiency by Direct Utilization of Wet
Ethanol in HCCI Engines.
Transactions of ASME 129, 332-337
California,USA, 2007
[11] Salla A D., Furlanetto B. P. F., Cabello C., Kantach D. R. A.,
Energy Evaluetion of the Etanol Production Using as raw-material the Sugar
Cane.
Ciência Rural, 0103 – 8478
Santa Maria, 2009
[12] ChunDe Y., ZhiHui Z., ChenShun C., GuangLan X.,
Experimental study on the effect of gaseous and particulate emission from an
ethanol fumigated diesel engine.
Science China 53, 3294-3301
China, 2010
21st Aachen Colloquium Automobile and Engine Technology 2012 23

[13] Caroa, de S., Moulounguia Z., VAitilingomb G., Bergec J. Ch.


Interest of combining an additive with diesel–ethanol blends for use in diesel
engines.
Fuel 80,565-574
France,2001
[14] Subramanian, V., Mallikarjuna, J.M., Ramesh, A.
Effect of water injection and spark timing on the nitric oxide emission and
combustion parameters of a hydrogen fuelled spark ignition engine. International
Journal of Hydrogen Energy, vol. 32, pp. 1159-1173.
Indian Institute of Technology Madras, India, 2006.
[15] Nguyen, Q. and Wu, Y.
Experimental investigations of using water-gasoline emulsions as a NOx
treatment and its effects on performance and emissions of lean-burn spark-
ignition engine.
Proceeding of the International conference on Power Engineering
Japan,2009.
[16] Özcan H., and Söylemez M.S., 2005.
Experimental Investigation of the Effects of Water Addition on the Exhaust
Emissions of a Naturally Aspirated, Liquefied-Petroleum-Gas-Fueled Engine.
Energy and Fuels, vol. 19
Faculty of Engineering, University of Gaziantep, Turkey, 2005
[17] Rajan, S., Saniee, F.F.,
Water-ethanol-gasoline blends as spark ignition engine fuels.
Fuel, vol. 62, issue 1,pp.
Southern Illinois University, United States, 1983
[18] Zhao H.
Advanced direct injection combustion engine technologies and development.
Volume 1: Gasoline and gas engines
Woodhead Publishing
Cambridge 2010
[19] Flowers D.L., Aceves S.M., and Frias J.M., 2007.
Improving Ethanol Life Cycle Energy Efficiency by Direct utilization of Wet Ethanol
in HCCI Engines.
SAE Paper 2007011867, 2007
[20] Boretti, A.,
Towards 40% efficiency with BMEP exceeding 30 bar in directly injected,
turbocharged, spark ignition ethanol engines.
Energy Conversion and Management 57
Missouri University of Science and Technology, United States, 2012.
[21] Cordon D., Clarke E., Beyerlein S., Steciak J.,
Catalytic Igniter to Support Combustion of Ethanol- Water/air Mixtures in Internal
Combustion Engines.
SAE
University of Idaho, Moscow,2002
24 21st Aachen Colloquium Automobile and Engine Technology 2012

[22] Christensen M., Johansson B.,


Homogeneous Charge Compression Ignition with Water Injection.
SAE Technical papers series
Lund Institute of Technology,1999
[23] Neumann, J., Andrei, S. and Banischewski B.
Sensitivity Study of a Turbo- Charged SI-Engine at Rated Power.
GT-SUITE User’s Conference.
Frankfurt, Oktober 08, 2007.
[24] Birckett, A. and Keidel, S.
Modeling Combined Superchad/Turbocharged Engine Concepts in GT-POWER.
GT-SUITE User’s Conference.
Birmingham, November 7, 2011.
[25] Bromberg, L. and Blumber, P.
Estimates of DI hydrous Ethanol Utilization for Knock Avoidance and Comparison
to a Measuired and Simulated DI E85 Baseline.
Plasma Science and Fusion Center, 09-33.
[26] Morel, T. and Keribar, R.
A model for predicting spatially and time resolved convective heat transfer in
bowl-in-piston combustion chambers.
SAE Paper 850204, 1985
[27] Morel, T., Rackmil, C. I., Keribar, R., and M. J. Jennings.
Model for heat transfer and combustion in sparl ignited engines and its
comparison with experiments.
SAE Paper 880198, 1988
[28] Wahiduzzaman S., and Morel T.
Comparison of measured and predicted combustion characteristics of a four-valve
S.I. Engine.
SAE Paper 930613, 1993
[29] Parag S., and Raghavan V.
Experimental investigation of burning velocities of pure ethanol and ethanol
blended fuels.
Combustion and Flame: Vol 156. pp 997-1005. 2009.
[30] Zhang Z., Li g., Ouyoung L., Pan Z., You F., and Gao X.
Experimental determination of laminar burning velocities and Markstein lengths
for 75% hydrous-ethanol, hydrogen and air gaseous mixtures.
International Journal of Hydrogen Energy: Vol. 36. pp. 13194-13206. 2011
[31] Eisazadeh-Far K., Moghaddas A., Al-Mulki J., and Metghalchi H.
Laminar burning speeds of ethanol/air/diluent mixtures.
Proceedings of the Combustion Institu: Vol. 33. pp. 1021-1027. 2011
[32] Saxena S., Schneider S., Aceves S., and Dibble.
Wet ethanol in HCCI engines with exhaust heat recovery to improve the energy
balance of wet ethanol fuels.
Applied Energy: Vol. 98. pp. 448-457. 2012
[33] Johansson T., Johansson B., Tunestal P., and Aulin H.
21st Aachen Colloquium Automobile and Engine Technology 2012 25

HCCI Operating Range in a Turbo-charged Multi Cylinder Engine with VVT and
Spray-Guided DI.
SAE Paper 2009010494 (2009)
[34] Heywood J. B.
Internal Combustion Engines Fundamentals
McGraw-Hill
New York, 1988
[35] Aceves, S. M., Flowers, D. L., Westbrook, C. K., Pitz, W., Smith, J. R.,
Dibble, R. W., Christensen, M., and Johansson, B.
A Multi-Zone Model for Prediction of HCCI Combustion and Emissions
SAE Paper No. 2000-01-0327.
[36] Siebers, D.L. and Edward, C.F.
Autoignition of Metanol and Ethanol Sprays under Diesel Engine Conditions,
SAE 870588,
[37] Daeyup Lee, Simone Hochgreb and James C. Keck,
Autoignition of Alcohols and Ethers in a Rapid Compression Machine
SAE 932755
[38] Munsin R., Laoonual Y., Jugjai S., Matsuki M., Kosaka H.
Investigation of Effects of Ignition Improver on Ignition Delay Time of Ethanol
Combustion with Rapid Compression-Expansion Machine
The Second TSME International Conference on Mechanical Engineering
Krabi ,19-20 October, 2011

Anda mungkin juga menyukai