You are on page 1of 15

Renewable Energy 30 (2005) 1733–1747

www.elsevier.com/locate/renene

Data bank

Experimental and theoretical investigation of using


gasoline–ethanol blends in spark-ignition engines
Hakan Bayraktar*
Naval Architecture Department, Faculty of Marine Science, Karadeniz Technical University,
Çamburnu, Sürmene, Trabzon 61530, Turkey
Received 20 September 2004; accepted 15 January 2005
Available online 2 March 2005

Abstract
The effects of ethanol addition to gasoline on an SI engine performance and exhaust emissions are
investigated experimentally and theoretically. In the theoretical study, a quasi-dimensional SI engine
cycle model, which was firstly developed for gasoline-fueled SI engines by author, has been adapted
for SI engines running on gasoline–ethanol blends. Experimental applications have been carried out
with the blends containing 1.5, 3, 4.5, 6, 7.5, 9, 10.5 and 12 vol% ethanol. Numerical applications
have been performed up to 21 vol% ethanol. Engine was operated with each blend at 1500 rpm for
compression ratios of 7.75 and 8.25 and at full throttle setting. Results obtained from both theoretical
and experimental studies are compared graphically. Experimental results have shown that among the
various blends, the blend of 7.5% ethanol was the most suitable one from the engine performance
and CO emissions points of view. However, theoretical comparisons have shown that the blend
containing 16.5% ethanol was the most suited blend for SI engines. Furthermore, it was
demonstrated that the proposed SI engine cycle model has an ability of computing SI engine cycles
when using ethanol and ethanol–gasoline blends and it can be used for further extensive parametric
studies.
q 2005 Elsevier Ltd. All rights reserved.

Keywords: SI engine cycle simulation; Alternative fuels; Ethanol; Gasoline–ethanol blends

* Tel.: C90 462 752 2805; fax: C90 462 752 2158.
E-mail address: hakanbay@ktu.edu.tr.

0960-1481/$ - see front matter q 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.renene.2005.01.006
1734 H. Bayraktar / Renewable Energy 30 (2005) 1733–1747

Nomenclature
Af area of spherical flame front, m2
cp specific heat at constant pressure, J kgK1 KK1
(F/A)s stoichiometric fuel to air ratio, dimensionless
h enthalpy, J kgK1
LHV lower heating value, MJ kgK1
lt characteristic length scale of turbulent flame, m
m mass, kg
me mass entrained by the flame front into the flame zone, kg
PRT purity of ethanol, %
p cylinder pressure, bar
Q_ w total heat transfer rate to the cylinder walls, W
Sl laminar flame speed, msK1
T temperature, K
Ut characteristic turbulent speed, msK1
V instantaneous cylinder volume, m3
X volume percentage, %
Greek letters
3 compression ratio, dimensionless
f fuel/air equivalence ratio, dimensionless
q crank angle, degrees
r density, kg mK3
tb characteristic reaction time to burn an eddy of size lt, s
j mass fraction of residual burned gases in the fuel–air–residual gas mixture,
%
Subscripts
b burned gases
bl blend
E ethanol
f flame
F conditions at the end of the flame propagation process
G gasoline
p pure
u unburned gases
w water

1. Introduction

Alcohols have been suggested as an engine fuel almost since automobile was invented
[1,2]. However, some shortcomings of alcohols such as difficulties in their production and
their some unsuitable properties for engines have limited their widespread usage [2,3].
In recent years, the reasons such as the possibility of depletion of the world’s crude oil
H. Bayraktar / Renewable Energy 30 (2005) 1733–1747 1735

reserves, increases in wholesale price of crude oil and increasing air pollution have
increased the interest in alcohols [2–4].
Among the various alcohols, ethanol is known as the most suited fuel for spark-ignition
(SI) engines [5,6]. The most attractive properties of ethanol as an SI engine fuel are that it
can be produced from renewable energy sources such as agricultural feedstock and it has
high octane number and flame speed [5–7]. Production methods and properties of ethanol
will be further explained and discussed in the following sections.
Ethanol can be used in SI engines as pure or by blending with gasoline [1,6–9]. Pure
ethanol necessitates some modifications on engine design and fuel system [10] whereas
it can be used in SI engines by blending with gasoline at low concentrations without
any modification [2]. It was reported that using gasoline–ethanol blends including
ethanol at low concentrations could improve engine performance and exhaust emissions
[2,5,6,9,11].
It is well understood from the above literature review that using ethanol in SI engines
by blending with gasoline is more practical than using it alone. If ethanol production can
meet the demand and the cost of blended fuels can compete with that of conventional
gasoline, widespread use of gasoline–ethanol blends can be possible. However, before
using these blends in engines, the whole effects on engine must be evaluated. For this
reason, the present study is focused on this topic. Here, the effects of ethanol addition to
gasoline in various concentrations on engine performance and exhaust emissions are
examined by conducting both theoretical and experimental studies.

2. Availability and usability of ethanol as an SI engine fuel

2.1. Production of ethanol

Any material which can be converted to fermentable sugars [1,7] and synthesis gas
which is mainly consist of CO and H2 [12], can be used to produce ethanol. Suitable
feedstocks for ethanol production are agricultural products such as sugar cane and grains
[1–4,7,13], agricultural solid wastes [14,15], cellulosic materials such as wood [16] and
coal [12]. Simple sugars can be obtained from agricultural raw materials by direct
treatment or by acid hydrolysis and then they can be fermented by yeast to yield raw
ethanol [3,14,16]. Furthermore, ethanol can also be produced by fermentation of the
synthesis gas, which can be manufactured by gasification of coal [12]. The purity of
ethanol produced via fermentation is approximately 96% by volume. To remove the water
in the alcohol or to improve the purity of ethanol requires additional costly distillation
processes [1,3,7,8,13]. Such processes further raise the cost of ethanol. For this reason,
although production of ethanol from renewable sources is an advantage, its higher cost
relative to gasoline is a disadvantage.

2.2. Properties of ethanol

Certain properties of ethanol and gasoline are given in Table 1. Following effects on
engine can be expected when ethanol is used as an engine fuel. Ethanol contains an
1736 H. Bayraktar / Renewable Energy 30 (2005) 1733–1747

Table 1
Fuel properties

Fuel Molecular Molecular Octane number LHV Stoichio- r (g cmK3)


formula weight (MJ kgK1) metric fuel/
(kg kmoleK1) air ratio
Gasoline C7H17 100–110 91–96 44 0.0685 0.72–0.78
Ethanol C2H5OH 46 106–110 27 0.1111 0.785

oxygen atom therefore, it can be considered as a partially oxidized fuel [1,2,17]. For
this reason, it has lower calorific value and stoichiometric air–fuel ratio than gasoline.
Consequently, much more fuel is needed to obtain same performance when ethanol or
ethanol–gasoline blends are used [1,5,7,8,17]. Ethanol has higher octane number than
gasoline thus it can lead to operation at higher compression ratios therefore
improvement in power output, efficiency and fuel consumption. Furthermore, it has
high latent heat of vaporization. As a consequence of both low calorific value and high
latent heat of vaporization, engine volumetric efficiency may increase. However,
vaporization of the intake mixture may be reduced. This problem can be avoided by
heating intake manifold. It was reported that although vapor pressure of pure ethanol is
low, Reid vapor pressure (RVP) of gasoline–ethanol blends rises depending on the
ethanol proportion in the blend [5,7]. Low RVP can cause cold starting problems,
therefore, volatile additives should be used when pure ethanol is used [18].
Furthermore, vapor lock may occur in the warm weathers [1,5,7,17,18]. Because of
the cooling effect on the intake charge and leaner operation, significant reductions in
CO and NOx emissions may be expected [2,5,7]. As known, ethanol has high affinity
for water, thus, it contains certain amount of water [1,2,7]. This is not a problem for
pure ethanol because it is fully miscible with water but some serious problems can arise
when gasoline–ethanol blends are used. Phase separation can occur in gasoline–ethanol
blends since gasoline and ethanol are immiscible. This problem can be prevented by
using semi-polar cosolvents (solubility improvers) such as isopropanol [9,17]. On the
other hand, water contamination of ethanol can cause corrosion on mechanical
components [1,2,7]. Another important problem relating to use of ethanol in engines is
that the emissions of formaldehyde, acetaldehyde and acetone are significantly
increased [2,5].
General conclusions deduced from the above literature review can be summarized as
follows. If ethanol can be produced abundantly and economically, it will be an attractive
alternative fuel for SI engines. It can be used either as a pure fuel or as a gasoline additive.
Both options can provide some advantages for engine performance, fuel economy and
exhaust emissions. Gasoline–ethanol blends including ethanol at low proportions can be
used without any engine modification but pure ethanol requires major modifications to the
engine design and fuel system. Consequently, the use of gasoline–ethanol blends in SI
engines is more practical than using ethanol alone. For this reason, presented theoretical
and experimental studies have been focused on the effects of using gasoline–ethanol
blends on SI engine performance and exhaust emissions.
H. Bayraktar / Renewable Energy 30 (2005) 1733–1747 1737

3. Theoretical study

In this section, a quasi-dimensional SI engine cycle model used here is described. This
model is based on the First Law of Thermodynamics. Here, combustion is simulated as a
turbulent flame propagation process, therefore, geometries of flame and combustion
chamber are taken into account. Presented cycle model of which some details will be given
below, can compute SI engine cycle for the cases of using gasoline, ethanol and gasoline–
ethanol blends.

3.1. SI engine cycle model

Presented SI engine cycle model is originally based on the model previously developed
by the author [9]. Broad descriptions about mathematical model can be found in Refs.
[9,19,20]. Here, only governing equations and modifications performed on this model for
ethanol and gasoline–ethanol blends are presented. This model is governed by the
following first-order ordinary differential equations derived for cylinder pressure and
temperature
        
B m_ h V_ 1
T_ i Z 1K i K C ðKQ_ wi C m_ i hu Þ (1)
A i m i Bi V i ðBmÞi
       
r V_ vr=vT _ m_
p_ Z 10 K5
K K Ti C (2)
vr=vp i V i r i m i
where the dots denote the differentiation with respect to crank angle and coefficients Ai and
Bi are expressed for unburned and burned gases as follows
   
ðvr=vTÞu 1 ðvr=vTÞb 1 vh
Au Z cpu C ; Ab Z cpb C K 10K5
ðvr=vpÞu ru ðvr=vpÞb rb vp b
   
1 1 vh
Bu Z ; Bb Z 1 K 10K5 r
ðvr=vpÞu ðvr=vpÞb vp b
If Eqs. (1) and (2) are arranged for each engine process under suitable assumptions,
then thermodynamic state of cylinder content at any instant of the cycle can be determined
by numerical integration of these equations. The terms included in the above equations are
specified simultaneously as follows.
Heat transfer between the gases and cylinder walls is calculated by using the empirical
correlation developed by Annand [21]. During intake and compression, cylinder charge
consists of only unburned gases (iZu) which are considered as a nonreacting ideal gas
mixture of air, fuel vapor and residual burned gases at low temperatures. Composition and
thermodynamic properties of unburned mixture and their partial derivatives with respect to
crank angle are determined by using the method firstly developed by Komiyama and
Heywood [22] and later extended for various fuels by Bayraktar [9]. Throughout the
combustion period, both unburned and burned gas regions, which are assumed to be
separated by a spherical flame front, exist in the combustion chamber. For this reason,
1738 H. Bayraktar / Renewable Energy 30 (2005) 1733–1747

Eqs. (1) and (2) are arranged for both unburned (iZu) and burned gas regions (iZb).
Combustion model will be further described in Section 3.2. During the expansion process,
there are only burned gases in the cylinder. It is assumed that burned gases are in chemical
equilibrium and their composition and thermodynamic properties are computed by means
of the method developed by Olikara and Borman [23].

3.2. Combustion model

During combustion, in addition to thermodynamic state of each zone, instantaneous


mass of burned gas mixture must also be determined. A turbulent flame propagation model
firstly developed by Blizard and Keck [24] and later extended by Keck [25], Bayraktar [9]
and Bayraktar and Durgun [19] is used for this purpose. Detailed information about this
model can be found in Refs. [9,19,20]. Here, only brief descriptions about this model are
given. Governing equations of combustion model are as follows

m_ e Z ru Af ðUt C Sl Þ (3)

ðme K mb Þ
m_ b Z ru Af Sl C (4)
tb

where tb is taken as lt/Sl. The solution method applied here is identical to that given in
Refs. [9,19,20] and can be summarized as follows. Combustion is assumed to occur in
three basic steps: initial burning phase during which flame develops in laminar manner
(UtZ0; meZmb), faster burning phase which is a turbulent flame propagation process and
final burning phase which starts after flame propagation is terminated that is the
combustion chamber is completely enveloped by the flame. This process is considered by
the following exponential burning equation given by Keck [25]:
m_ b
Z eKðqKqF Þ=tb (5)
m_ bF

Instantaneous values of Af, Sl, Ut and lt must be determined to solve the above
equations. Flame front area corresponding to enflamed volume is calculated by
interpolation. Empirical correlations given by Keck [25] are used for determination of
Ut and lt. Laminar flame speed is computed from the following empirical correlations
given by Gülder [26]

Sl ðf; T; pÞ Z Slo ðfÞ½Tu =To a ½p=po b ð1 K f jÞ (6)

where f is a constant and given as 2.5 for 0%j%0.3. Laminar flame speed at the reference
conditions (ToZ300 K and poZ1 atm) is determined as

Slo ðfÞ Z ZWfh exp ½Kx ðf K 1:075Þ2  (7)

Constants for gasoline, ethanol and ethanol–gasoline blends are given in Table 2. Here,
laminar flame speed of gasoline is calculated by using data given for isooctane (C8H18).
H. Bayraktar / Renewable Energy 30 (2005) 1733–1747 1739

Table 2
Coefficients for laminar flame speed

PoZ1 bar, Z W (m sK h x a b
1
ToZ300 K ) f%1 fR1
C8H18 1 0.4658 K0.326 4.48 1.56 K0.22
C2H5OH 1 0.4650 0.250 6.34 1.75 K0.17/Of K0.17Of
C8H18C 1C0.07X0.35
E 0.4658 K0.326 4.48 1.56C0.23X0.46
E XGbGCXEbE
C2H5OH

3.3. Solution technique and computation of the engine cycle

Properties of pure fuels and blended fuels are determined at the beginning of the cycle
computation. If the properties of pure gasoline and ethanol are known, properties of the
blended fuels are calculated as follows:
P
Xj r j
rbl Z (8)
100
P
Xj rj ðF=AÞsj
ðF=AÞsbl Z P (9)
Xj r j
P
X r LHVj
LHVbl Z Pj j (10)
Xj r j
where subscript j refers to gasoline (G) or ethanol (E).
Fuel–air equivalence ratio of the blended fuels varies with varying ethanol ratio.
Following formula given by Bayraktar [9] and Bayraktar [11] is used to calculate the fuel–
air equivalence ratio of the blended fuels:
rffiffiffiffiffiffi
ðF=AÞsG rbl
fbl Z fG (11)
ðF=AÞsbl rG
Intake conditions are determined by means of an approximate method developed by
Durgun [27] and given by Bayraktar and Durgun [19]. The total mass within the engine
cylinder is determined depending on the intake conditions and taken as constant
throughout the cycle. The cylinder is treated as spatially uniform in pressure. At any
instant during the cycle, instantaneous composition and thermodynamic properties of the
gas mixtures are first determined for each of pure fuels and then thermodynamic properties
of the blends are calculated based on the volume fraction of each fuel in the blend [9]. At
any computation step, after determining all terms in differential equations, thermodynamic
state of the cylinder charge is predicted by numerical integration of Eqs. (1) and (2). This is
achieved by using Euler-predictor–corrector integration scheme taking the crank angle
increments of 18.
Computation of compression process is terminated at spark angle qs prior to top dead
center and then combustion calculation is started. When calculating the combustion,
Eqs. (1) and (2) are solved in conjunction with Eqs. (3)–(5), therefore, mass fractions
1740 H. Bayraktar / Renewable Energy 30 (2005) 1733–1747

of unburned and burned gases are determined in addition to thermodynamic state of each
zone. Once the entire charge is fully burned, expansion calculation is started and continued
until bottom dead center. Exhaust conditions are determined depending on the pressure
and temperature at the end of the expansion by using a simple method developed by
Durgun [27] and given by Bayraktar and Durgun [19]. After cycle computation is
completed, engine performance parameters such as mean indicated pressure, effective
power, effective efficiency and specific fuel consumption are calculated based on the
indicated work predicted throughout the cycle.

3.4. Numerical applications

Presented cycle model has been arranged as a computer code and various numerical
applications have been performed with this code to predict the SI engine performance and
exhaust emissions in the cases of using gasoline and various gasoline–ethanol blends.
Knowledge about the engine geometry, operating conditions and properties of gasoline
and ethanol used here are given in Section 4. In the numerical applications, volume
percentage of ethanol in the blend has been changed from 1.5 to 21% with the increments
of 1.5%. Fuel/air equivalence ratios measured for the case of fueling with pure gasoline
(fG) are used in Eq. (11) to calculate the fuel/air equivalence ratios of the blended fuels
(fbl) at the same operating conditions.

4. Experimental study

4.1. Fuel properties

Density of the unleaded gasoline used in the experiments was measured as 0.74 g cmK3
and its approximate molecular formula was taken as C7H17. Density of ethanol was
measured as 0.80 g cmK3. The purity of ethanol corresponding to this density is calculated
as 93% by using the densities of pure ethanol having purity of 99.99% and water as follows
rw _rE
PRT Z (12)
rw _rEP
where rEP is the density of pure ethanol and taken as 0.785 g cmK3; rw is the density of
water and taken as 1 g cmK3. Lower heating values of gasoline and ethanol were
calculated approximately based on the elementary composition of each fuel by using the
Mendeleyev’s formula given by Khovakh [28]. These properties for gasoline and ethanol
used here were determined as 45.69 and 26.62 MJ kgK1, respectively. Properties of
blended fuels are calculated as explained in Section 3.3.

4.2. Experimental setup and procedure

The engine and instrumentation are identical to those used in the author’s earlier work
[11]. Here, brief descriptions about the experimental setup are given. The engine used in
the experiments is a single-cylinder, four-stroke, water-cooled, variable compression
H. Bayraktar / Renewable Energy 30 (2005) 1733–1747 1741

engine having swept volume of 763 cm3. Dimensions of the engine are: the bore DZ
90 mm and stroke HZ120 mm. Combustion chamber has a simple cylindrical shape and
the distance of ignition point from the edge of the combustion chamber is 30 mm. Engine
is coupled to an electric dynamometer, which is used to load engine and measure the
output torque. Carbon monoxide (CO) emission was measured by means of an exhaust gas
analyzer connected to exhaust pipe outlet. A calibrated burette and a stopwatch were used
to measure the engine fuel consumption. Gasoline–ethanol blends have been prepared by
blending ethanol having a purity of 93%, with gasoline in concentrations of 1.5, 3, 4.5, 6,
7.5, 9, 10.5, and 12% (by volume). Above the 12% ethanol, engine could not run smoothly,
therefore, experimental results obtained up to this percentage of ethanol will be presented.
Properties of the blends have been determined by using Eqs. (8)–(10) given in Section 3.3.
All the experiments have been carried out at full throttle setting. Spark advance was
adjusted as 108 before top dead center and held constant. Engine performance parameters
and CO emissions have been determined for each fuel at the compression ratios of 7.75
and 8.25. At each compression ratio, engine was operated at the speed of 1500 rpm.

5. Results, comparisons and discussions

Experiments and numerical applications have been performed at the same operating
conditions and results obtained for gasoline and various gasoline–ethanol blended fuels
are compared graphically in the following figures.
Ethanol is an oxygenated fuel; therefore, it has high stoichiometric fuel/air ratio. For
this reason, ethanol addition to gasoline leads to leaner operation. Fig. 1 indicates the
variation of f with ethanol concentration in the blend. As shown, fuel/air equivalence
ratio decreases as the volume percentage of ethanol in the blended fuel increases.
In the theoretical study, equivalence ratio of the blended fuel at any condition has been
determined from Eq. (11) by using equivalence ratio measured for gasoline-fueled engine

Fig. 1. Variation of fuel/air equivalence ratio with ethanol concentration in the blend.
1742 H. Bayraktar / Renewable Energy 30 (2005) 1733–1747

Fig. 2. Variation of maximum cylinder pressure with ethanol concentration in the blend.

(fG) at the same operating condition. As can be seen from Fig. 1, acceptable agreement
between theoretical and experimental equivalence ratios was obtained.
As explained in Section 2.2, ethanol addition to gasoline raises engine volumetric
efficiency and causes leaner operation. As a result, combustion becomes more complete or
more stoichiometric, therefore, flame temperature and cylinder pressure rise to their
maximum values as f approaches 1. The effects on cylinder pressure and temperature have
been determined by only theoretically since these combustion properties cannot be
measured in our experimental setup. As shown from Fig. 2, theoretical maximum cylinder
pressure increases with increasing ethanol proportion. At both compression ratios of 7.75
and 8.25, increase in theoretical maximum cylinder pressure rises up to 11%. Fig. 3
indicates that predicted maximum cylinder temperatures increase by about 2% with
increasing ethanol percentage. Increases in both cylinder pressure and temperature reach
their maximum values at 16.5% ethanol. Theoretical combustion durations are compared

Fig. 3. Variation of maximum cylinder temperature with ethanol concentration in the blend.
H. Bayraktar / Renewable Energy 30 (2005) 1733–1747 1743

Fig. 4. Variation of combustion duration with ethanol concentration in the blend.

in Fig. 4. Ethanol addition to gasoline reduces the combustion duration by about 17% (at
3Z7.75) and 22% (at 3Z8.25) owing to improved combustion.
Mean indicated work and mean indicated pressure increase because of the increases in
cylinder pressures. Therefore, engine power output and thermal efficiency increases with
increasing ethanol content of the blended fuel, although heating values of blended fuels are
lower than that of gasoline. Theoretical and experimental engine effective powers obtained
with gasoline and various gasoline–ethanol blends are compared in Fig. 5. An agreement
of about 6% between measured and predicted values was obtained. As shown from the
figure, ethanol addition to gasoline enhances the engine power output. At the compression
ratio of 7.75, measured and predicted engine powers increase by about 2.34 and 2.70%,
respectively. Improvement in power is more evident at higher compression ratios owing to
increasing knock resistance of the blended fuel with increasing amount of ethanol.
Increases in measured and predicted engine powers at 3Z8.25 are about 4.37 and 2.86%,
respectively. Variation of effective efficiency with volume percentage of ethanol in

Fig. 5. Variation of effective power with ethanol concentration in the blend.


1744 H. Bayraktar / Renewable Energy 30 (2005) 1733–1747

Fig. 6. Variation of effective efficiency with ethanol concentration in the blend.

the blend is plotted in Fig. 6. Engine effective efficiency improves depending on the
ethanol amount in the blended fuel. This also resulted from more complete and perfect
combustion. Maximum increases in experimental and theoretical efficiencies at 3Z7.75
are about 7.56 and 5.93%, respectively. At 3Z8.25, experimental and theoretical effective
efficiencies increase by about 6.10 and 5.88%, respectively.
As indicated in Figs. 5 and 6, the most suitable blend was determined experimentally as
the blend of 7.5% ethonol while theoretically as the blend of 16.5% ethanol. This
difference between experimental and theoretical results can be attributed to water content
of ethanol: as known, in actual conditions, water may cause phase separation, therefore,
power loss and this can be observed experimentally, however, this negative effect of water
has not been considered in theoretical model. In this connection, it may be concluded that
the ethanol content of the blended fuel can be increased up to 16.5% by enhancing the
purity of ethanol and preventing phase separation.
Ethanol addition reduces the heating value of the gasoline–ethanol blends, therefore,
more fuel is needed (by mass) to obtain same power when blended fuels are used instead of
gasoline. However, as mentioned previously, ethanol addition to gasoline makes the
engine operation leaner and improves engine combustion and performance. For these
reasons, as shown in Fig. 7, specific fuel consumptions measured at 3Z7.75 and 8.25 when
operating with gasoline–ethanol blends are lower than those of gasoline-fueled engine.
The blend of 7.5% ethanol gives the reductions up to 5.59 and 4.94% in specific fuel
consumptions measured at 3Z7.75 and 8.25, respectively. However, results obtained from
theoretical model have shown that specific fuel consumption has slightly increased by
increasing ethanol amount in the blend.
Fuel–air equivalence ratio approaches 1 as the ethanol content of the blended fuel
increases (Fig. 1), consequently combustion becomes more complete and flame
temperature rises due to stoichiometric combustion (Fig. 3). Furthermore, carbon content
of ethanol is lower than that of gasoline. Because of the mentioned reasons, CO
concentration decreases with increasing ethanol concentration (Fig. 8). Experimental
results have shown that CO mole fractions could be reduced by 44.26% (at 3Z7.75) and
H. Bayraktar / Renewable Energy 30 (2005) 1733–1747 1745

Fig. 7. Variation of specific fuel consumption with ethanol concentration in the blend.

41.67% (at 3Z8.25). Reductions in theoretical CO emission are 61.82% (at 3Z7.75) and
68.16% (at 3Z7.75).
Here, the effects on NO concentration have been determined theoretically from the
equilibrium calculation because of the lack of equipment to measure NO emission. As
indicated in Fig. 9, NO mole fraction increases depending on the ethanol proportion. This
resulted from rising combustion temperature. NO mole fraction slightly increases up to
16.5% ethanol and then rapidly increased.

6. Conclusions

In the presented paper, experimental and theoretical studies were performed on the use
of gasoline–ethanol blends in SI engines. General results concluded from this study can be
summarized as follows:

Fig. 8. Variation of CO mole fraction with ethanol concentration in the blend.


1746 H. Bayraktar / Renewable Energy 30 (2005) 1733–1747

Fig. 9. Variation of NO mole fraction with ethanol concentration in the blend.

1. The results obtained with presented mathematical cycle model are in acceptable agreement
with those measured ones. In general, an agreement of 6% was determined between
theoretical and experimental results. This means that presented model can be used for more
extensive parametric studies concerning the using of gasoline–ethanol blends in SI engines.
2. Ethanol addition to gasoline leads to leaner operation and improves combustion.
Consequently, cylinder pressure and temperature increase and combustion duration
decreases.
3. Engine performance parameters such as effective power and effective efficiency increase
with increasing ethanol amount in the blended fuel as a result of improved combustion.
4. Using the gasoline–ethanol blends in SI engines dramatically reduces the CO
concentrations. However, NO concentrations are adversely affected due to rising
cylinder temperature with increasing ethanol proportion in the blend.
5. In general, most suitable blend for SI engines has been specified theoretically as the
blend of 16.5% ethanol and experimentally as the blend of 7.5% ethanol. This means
that if some problems, which arise in SI engines when fueling with gasoline–ethanol
blends, will be removed, the blends including ethanol up to 16.5% by volume can be
used in SI engines without any modification to the engine design and fuel system.

References

[1] Wagner TO, Gray DS, Zarah BY, Kozinski AA. Practicality of alcohols as motor fuel. SAE 1979, Paper no.
790429: 1591–1607.
[2] Hsieh WD, Chen RH, Wu TL, Lin TH. Engine performance and pollutant emission of an SI engine using
ethanol-gasoline blended fuels. Atmos Environ 2002;36:403–10.
[3] Mielenz JR. Ethanol production from biomass: technology and commercialization status. Curr Opin
Microbiol 2001;4:324–9.
[4] Al-Baghdadi MAS. Hydrogen–ethanol blending as an alternative fuel of spark ignition engines. Renew
Energy 2003;28:1471–8.
[5] He BQ, Wang JX, Hao JM, Yan XG, Xiao JH. A study on emission characteristics of an EFI engine with
ethanol blended gasoline fuels. Atmos Environ 2003;37:949–57.
H. Bayraktar / Renewable Energy 30 (2005) 1733–1747 1747

[6] Al-Hasan M. Effect of ethanol–unleaded gasoline blends on engine performance and exhaust emissions.
Energy Convers Manage 2003;44:1547–61.
[7] Thring RH. Alternative fuels for spark-ignition engines. SAE 1983, Paper no. 831685: 4715–25.
[8] Clancy JS, Dunn PD, Chawawa B. Ethanol as fuel in small stationary spark ignition engines for use in
developing countries. IMechE 1988;C67(88):191–4.
[9] Bayraktar H. Theoretical investigation of using ethanol–gasoline blends on SI engine combustion and
performance. PhD Thesis, Karadeniz Technical University, Trabzon, Turkey; 1997.
[10] Kreith F. Mechanical engineering handbook. Boca Raton, FL: CRC Press LLC; 1999.
[11] Bayraktar H. Using gasoline–ethanol–izopropanol blends in engines. MS Thesis, Karadeniz Technical
University, Trabzon, Turkey; 1991.
[12] Worden RM, Grethlein AJ, Jain MK, Datta R. Production of butanol and ethanol from synthesis gas via
fermentation. Fuel 1991;70:615–9.
[13] Wheals AE, Basso LC, Alves DMG, Amorim HV. Fuel ethanol after 25 years. TIBTECH 1999;17:482–7.
[14] Jiménez L, Bonilla JL, Ferrer JL. Exploitation of agricultural residues as a possible fuel sources. Fuel 1991;
70:223–6.
[15] Farina GE, Barrier JW, Forsythe ML. Fuel alcohol production from agricultural lignocellulosic feedstocks.
Energy Sour 1988;10:231–7.
[16] Lynd LR, Cushman JH, Nichols RJ, Wyman CE. Fuel ethanol from cellulosic biomass. Science 1991;251:
1318–23.
[17] Norton TS. The combustion chemistry of simple alcohol fuels. PhD Thesis, Princeton University; 1990.
[18] Furey RL, Perry KL. Volatility characteristics of blends of gasoline with ethyl tertiary-butyl ether (ETBE).
SAE 1990, Paper no. 901114: 416–29.
[19] Bayraktar H, Durgun O. Mathematical modeling of spark-ignition engine cycles. Energy Sour 2003;25(5):
439–55.
[20] Bayraktar H, Durgun O. Development of an empirical correlation for combustion durations in spark ignition
engines. Energy Convers Manage 2004;45:1419–31.
[21] Annand WJD. Heat transfer in the cylinders of reciprocating internal combustion engines. Proc Inst Mech
Eng 1963;177:973–90.
[22] Komiyama K, Heywood JB. Predicting NOx emissions and effects of exhaust gas recirculation in spark-
ignition engines. SAE 1973, Paper no. 730475: 1458–76.
[23] Olikara C, Borman GL. A computer program for calculating properties of equilibrium combustion products
with some applications to I.C. engines. SAE 1975, Paper no. 750468: 1–21.
[24] Blizard NC, Keck JC. Experimental and theoretical investigation of turbulent burning model for internal
combustion engines. SAE 1974, Paper no. 740191: 846–64.
[25] Keck JC. Turbulent flame structure and speed in spark-ignition engines. Nineteenth (international)
symposium on combustion. The Combustion Institute; 1982 p. 1451–66.
[26] Gülder ÖL. Correlations of laminar combustion data for alternative SI engine fuels. SAE 1984, Paper no.
841000: 1–23.
[27] Durgun O. A practical method for engine cycles. Society of Mechanical Engineers of Turkey. J Eng Mach
1991;383:19–28 [in Turkish].
[28] Khovakh M. Motor vehicle engines. Moscow: MIR Publishers; 1979.

You might also like