You are on page 1of 10

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 8 5 2 5 e8 5 3 4

Available online at www.sciencedirect.com

ScienceDirect
journal homepage: www.elsevier.com/locate/he

An experimental investigation of hydrogenenriched gasoline in a Wankel rotary engine


F. Amrouche a,*, P. Erickson b, J. Park b, S. Varnhagen b
a

Renewable Energy Development Center, Hydrogen e Renewable Energy, B.P. 62, Observatory Street, Bouzareah,
Algiers 16340, Algeria
b
Mechanical and Aerospace Engineering Department, UC Davis, CA 95616, USA

article info

abstract

Article history:

The Wankel rotary engine is a potential alternative to the reciprocating engine in hybrid

Received 29 January 2014

applications because of its favorable energy to weight ratio. In this study, a Wankel rotary

Received in revised form

engine was modified to run on a hydrogenegasoline blend. Hydrogen enrichment

13 March 2014

improved the performance of a lean-burn spark-ignition rotary engine operating at high

Accepted 22 March 2014

speed and wide open throttle conditions with the original ignition timing, using 0%, %2, 4%,

Available online 19 April 2014

5%, 7%, and 10% hydrogen energy fractions at the intake. The experimental results showed
that adding hydrogen to gasoline in the engine improved the thermal efficiency and the

Keywords:

power output. Hydrocarbon and carbon monoxide emissions were reduced while nitrogen

Hydrogen

oxide emissions increased with the increase of hydrogen fraction.

Gasoline

Copyright 2014, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights
reserved.

Wankel rotary engine


Combustion
Emissions

Introduction
Pollution and the fossil fuel depletion are among the most
important concerns in the last decade. The transportation
sector that largely depends on fossil fuels and is responsible
for a large portion of pollution emissions has received
particular attention from scientific and political communities.
Hydrogen is considered an alternative for conventional fuels
because it can be produced from renewable energy sources
and also, when burned in air produces only NOx as harmful
emissions [1]. However, when hydrogen is used in the spark
ignition engine, it can lead to combustion problems such as
back firing, auto-ignition and pre-ignition. Also, the low
volumetric energy density of gaseous hydrogen makes storage

onboard the vehicle a great technical challenge for the


transportation sector. Therefore, blends of hydrogen with a
hydrocarbon fuel can be an alternative that offers the possibility of enhancing the mixture by taking advantage of properties from both fuels. Indeed, the use of hydrogen as
enrichment technique is shown to improve the engine emissions and performance for many fuels such as landfill gas [2],
natural gas [3] and methane [4], gasoline [5e8], ethanol [9], and
methanol [10]. Hydrogen blending can also partially resolve
the difficulties of onboard storage of hydrogen by requiring a
much smaller amount of hydrogen fuel to be stored and
transported. Furthermore, hydrogen blending could potentially serve as the first step toward widespread use of
hydrogen fuel as applied to the transportation sector.

* Corresponding author. Tel.: 213 771125072.


E-mail address: fethia.amrouche@gmail.com (F. Amrouche).
http://dx.doi.org/10.1016/j.ijhydene.2014.03.172
0360-3199/Copyright 2014, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved.

8526

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 8 5 2 5 e8 5 3 4

Hydrogen enrichment has been shown to extend the lean


operating limit of many fuel-air flames [11]. Lean burn has
been proven to increase thermal efficiency, and decrease
brake specific fuel consumption but can potentially decrease
engine flame stability [11]. The lean operating limit in an engine depends on combustion parameters such the flame speed
and the fuel flammability limit. A slow flame speed typically
will result in combustion instability, decreased heat release,
and increased HC emissions [12]. The lean flammability limit
of a fuel/air mixture also depends on the thermal diffusivity,
reaction rate, and specific heat of the reactants. Both the
thermal diffusivity and reaction rate are increased with the
addition of hydrogen. Moreover, hydrogen has a diffusion
coefficient of 0.61 cm3/s which is larger than most fuels. This
improved diffusion improves homogeneity of the mixture [13].
In the literature, reciprocating spark ignition engines
fueled by gasoline and hydrogen have been reported [5e8].
However, there are no reports of hydrogen enrichment with
the Wankel rotary engine which is an alternative to the
reciprocating engine [14,15], especially for hybrid vehicle applications [16].
The Wankel engine, by using an eccentric rotary design
instead of reciprocating pistons to produce motion is less
complicated and has an advantage of high power to weight
ratio as compared to similar reciprocating engines. The
Wankel is generally more compact, has less vibration, and
lower noise and power losses than a similar reciprocating
engine. The Wankel also generally has flat torque characteristics. In the Wankel engine three combustion events occur
with each full rotation of the rotor yielding high power especially at the higher end of the operating speed domain. This
results in higher performance at high speeds with a smooth
operation of the engine [14,15].
However, the Wankel engine also has some distinct disadvantages such as: Sealing problems [14] and low power
output at low speeds. Furthermore, Wankel engines typically
have lower thermodynamic efficiency because of the long
combustion-chamber shape. High emissions of NOx, CO and
especially HC are another problem in the rotary engine. These
drawbacks are mainly caused by the rotarys long geometry
and relatively larger quenching area. Additionally, unburned
gas leakage past seals to the exhaust chamber, creates further
emissions [15].
The experience gained through the years in operating the
reciprocating engine with hydrogen has led to the promise of
the Wankel engine being particularly well suited for hydrogen
enrichment. Theoretical reasons for this are as follows. The
presence of hydrogen can make a significant difference in
terms of emissions and performance by extending the lean
operating conditions because of the wide flammability limits
of hydrogen. Moreover, the absence of an exhaust valve which
is a common hot spot in reciprocating engines causing preignition and the physical separation of combustion and
exhaust chambers can prevent pre-ignition and/or backfire.
Also, the higher turbulence within the rotary combustion
chamber could allow a leaner operating limit to be reached.
Furthermore, the high hydrogen flame speed and relatively
small quenching distance potentially eliminate difficulties
with the rotary engine flame quenching. Brown et al. [1]
experimentally investigated the performance of a single

rotor Wankel engine fueled on pure hydrogen. The engine was


operated at a part and open throttle at relative equivalence
ratios varying from 1 to 4. They found out that the maximum
power was decreased but the brake thermal efficiency increases at part load as compared to gasoline while the NOx
emissions change was found to be negligible. Salanki [17]
carried out an experimental study of the hydrogen fueled rotary engine for hybrid vehicle applications. They used a 2.2 kW
Wankel rotary engine at WOT and lean conditions. The preliminary results showed an absence of combustion difficulties
and NOx emissions are on the order of 5 ppm for power outputs up to 70% of the maximum attainable on hydrogen fuel.
Mazda has been developing hydrogen vehicles with
hydrogen Rotary Engines (RE) from the early 1990s. Recently,
the company has delivered the RX-8 Hydrogen RE powered. To
achieve high power density, injectors are installed on the top
of the housing for direct injection, and other injectors are
placed on air intake ports. The combination of the direct and
port injection has brought simultaneous achievement of high
power, NOx reduction and better thermal efficiency. For
further improvement in fuel efficiency, lean burn operation is
used [18].
Because all published studies found in literature are
focused on the performance of pure hydrogen in the Wankel
engine, we have a strong motivation to investigate the effect
of hydrogen addition to a gasoline fueled single rotor Wankel
engine. This study provides experimental data on the engine
performance and emissions at or near the Lower Operating
Limit as found with gasoline only operation. Because the best
operating of the Wankel engine is reached at high speed, a
speed of 3000 rpm and a wide open throttle was used during
these tests.

Experimental procedure
Experimental setup
The engine used in this test bench is a 0.530 L single rotor, air
cooled Wankel engine, using a single spark plug. The

Fig. 1 e Schematic of Wankel rotary engine.

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 8 5 2 5 e8 5 3 4

8527

Table 1 e OMC Wankel engine technical specifications.


Type of engine
Number of rotors
Cooling
Engine manufacturer
Ignition source
Fuel supply system
Displacement (cc)
Compression ratio
Power output
Weight
Fuel
Engine geometry
Eccentricity 3 generating radius 3 width of rotor housing (mm)

schematic of the Wankel engine used in this study is shown in


Fig. 1. This engine is manufactured by Outboard Marine Corporation, USA. The engine Technical Specifications are listed
in Table 1.
This OMC Wankel engine was initially built with a carburetor to prepare the intake air/fuel mixture. To achieve real
time control over the air/fuel mixture preparation as well as
hydrogen addition, two fuel injection systems were added,
one for gasoline and the other for hydrogen, the two injectors
were mounted into an aluminum block in the fueling system,
this block became a mixing chamber that allows the gasoline
vapor and fumigated hydrogen to mix in the intake air stream.
The aluminum injector block was mounted between the engine and the carburetor. The carburetor is used only as a
throttle body for the engine. The gasoline fuel flow was
metered by a calibrated Micromotion CMF010M Coriolis flow
meter with an RFT9739 transmitter with an accuracy of 0.10%
of rate for flow rates of 0e23 g/s. The hydrogen used in this
study was bottled industrial hydrogen (99.95% purity) that is
regulated down to 35 PSIG (2.41 bar) after two steps of pressure
reduction, and is regulated and metered via a calibrated Aalborg differential pressure mass flow controller, model GFC 47,
with an accuracy of 3% (0e20% full scale), and 1.5%
(20e100% full scale). The air mass flow was metered via a
specifically designed orifice plate, and a large air chamber was
added past the intake to damp cyclic variations in air flow
caused by the engine operation. The inlet temperature of air
was also measured. The schematic of the experimental set up
is shown in Fig. 2.
The spark timing, injection timing and pulse durations of
hydrogen and gasoline were controlled via a hybrid electronic
control unit (HECU) developed in the laboratory. That HECU
uses the original electromotive ECU, two different softwares
(Wintech Electromotive and National Instruments (Austin TX)
LabView version 7 software) and computers for the control of
the injection timing and pulse durations of gasoline and
hydrogen according to the desired air-to-fuel ratio and the
specified hydrogen energy fractions in the intake.
The exhaust emissions were filtered, and water vapor was
removed. The sample was then split between three calibrated
emissions measurement systems. These were:
1) a Horiba MEXA-574GE emissions analyzer, which analyzes
the HC, CO, CO2 and O2. The O2 and HC emissions were
determined by a hydrogen flame ionization detection

Rotary Wankel engine


Single rotor
Air cooled
Outboard Marine Corps, USA
Spark plug
Fuel injection
530 cm3
8.75e1
26 kW @ 5500 rpm
27 kg
50 parts gasoline, 1 part oil
13.97 mm  92.202 mm  77.7748 mm

method, and the CO and CO2 were measured by a nondispersive infrared method. The Horiba has a time
response less than 10 s, a repeatability of 0.04%, 20 ppm,
0.1%, and 0.4%, respectively, for CO, HC, CO2 and O2.
2) a California Analytical model 400 Heated Chemiluminescence Photodiode Detector (HCLD) for measuring
NO and NOx. The resolution of this equipment is 0.1 ppm,
repeatability of 0.5% of full-scale, and sensitivity up to
5 ppm.
3) a California Analytical model 300 Non dispersive Infrared
(NDIR) Analyzer for measuring CO, and CO2, it has a linearity and repeatability of better than 1%, and a 90%
response time of in 2 s.
During the experiment, the measured emissions showed a
good consistency between the different systems.
The air to fuel ratio was determined in three ways. There
were:
1) Monitoring by the Horiba MEXA-574GE emissions analyzer;
2) Direct calculation of the ratio from the measured mass
flow rate of air, gasoline and hydrogen;
3) A Bosch Universal Exhaust Gas Oxygen (UEGO) sensor
installed in the exhaust stream. This sensor was controlled
by an Innovative Motorsports LC-1 wideband controller
which determined the real time oxygen content in the
exhaust stream and output a predicted air-to-fuel ratio.
This air-to-fuel ratio was used as a feedback signal ratio, in
a closed loop control system, by the ECU to achieve more
precise control of the air-to-fuel ratio.
The calculated and measured air to fuel ratios has shown a
good agreement between the three methods during all the
experiments.
The
specific
equivalence
ratio
of
the
gasolineehydrogeneair mixture was calculated by Eq. (1) [5]:
_ air
lm

.

m_ g $AFst;g m_H2 $AFst;H2

(1)

_ air , m_ g , m_H2 are respectively the measured air, gasoline


where m
and hydrogen mass flow rates (kg/h). AFst,g and AFst;H2 are the
stoichiometric air-to-fuel ratios of gasoline and hydrogen,
such as AFst,g 14.6 and AFst;H2 34:3.
A calibrated Kistler 6051B high temperature piezoelectric
pressure transducer was used to capture the working

8528

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 8 5 2 5 e8 5 3 4

Fig. 2 e Schematic of the experimental set up.

chamber pressure data. This transducer is mounted in the


engine housing that was machined and tapped for this purpose. The signal was amplified using a Kistler 5010B charge
amplifier. A 58-tooth (60-2) trigger wheel and a Hall effect
magnetic sensor were mounted on the engines output shaft
to obtain crank angle position data, the crank angle signal was
synchronized with the pressure trace.
The pressure and crank angle were recorded using National Instruments (Austin TX) LabView version 7 software.
Chamber pressure and crank angle signals for over 100
consecutive cycles were sampled and treated post process via
Matlab to obtain chamber pressure against crank angle
profiles.
A Telma CC100 eddy current dynamometer was coupled to
the engine to control and measure the engine speed and torque output.
The control of the dynamometer as well as the majority of
data acquisition in the engine test cell was accomplished
using components from the Superflow 742 product line. The
brain of the system was the Superflow ProSensor box, which
communicated with the Superflow ProConsole and Superflow
WinDyn software for real time data display.

or LOL is defined as the equivalence ratio at which first misfire


occurs. Also it is defined by Shiomoto [20] as the equivalence
ratio at which CO and HC emissions starts to increase. Ma [12]
defined the LOL as the excess air ratio (defined as the reciprocal of equivalence ratio) at which coefficient of variations
(COV) in IMEP reaches 10%. This is because it is generally
accepted that a COV above 10% will be perceived by a driver as
a poor running condition. However, in this study, LOL was
based the equivalence ratio at which first misfire occurs when
the engine was running on pure gasoline.
For this work, the hydrogen input was adjusted according
to the intake energy of the gasoline fuel. For each energy
fraction of hydrogen added to the fuel mixture, gasoline flow
rate was reduced through decreasing gasoline flow rate by the
ECU to keep the global excess air ratio of the hydrogen gasolineeair mixture at an approximately fixed air-to-fuel
equivalence ratio for gasoline.
The hydrogen energy fraction in the total fuel energy varies
from 0% to 10%. The lean operating limit of gasoline which is
equal to an equivalence ratio of 0.77 was used during these
tests. Since we wanted to investigate the effect of different
amount of hydrogen addition on engine performance and
emissions with all engine operation parameters unchanged,

Tests conditions
The tests were performed at an engine speed of 3000 rpm and
at wide open throttle. After a warm-up period of the engine on
gasoline, the appropriate air to fuel ratio and hydrogen energy
fraction was set. The engine was again allowed to come to
steady state at each operating condition and the data were
collected. Table 2 summarizes the operating conditions used
in the experiments.
There are different methods of defining engine lean operating limit (LOL). As defined by Badr [19] the lean misfire limit

Table 2 e Engine operating conditions during the


experiments.
Hydrogen energy fraction
Ignition timing
(degree Crank Angle BTDC)
Engine speed (rpm)
Equivalence ratio f
Throttle position

0%, %2, 4%, 5%, 7%, 10%


15
3000
LOL of gasoline
Wide open

8529

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 8 5 2 5 e8 5 3 4

%H2 

#

_ H2  LHVH2
m
 
  100
_ g  LHVg m
_ H2  LHVH2
m


(2)

_ g are respectively the mass flow rate of hydrogen


_ H2 , m
where m
and gasoline (g/s).
LHVH2 , LHVg are the lower heating value, respectively, of
hydrogen and gasoline (MJ/kg), such as LHVH2 120:1,
LHVg 43.5.
The maximum working chamber temperature was calculated from the pressure in this paper using the ideal gas law
[27]:

Tmax

Pmax  V
mmix  RMix

26

24

22

20

18

16

14

12

10


(3)

where Rmix, mmix, Pmax, V are respectively the gas constant, the
mass of the mixture, the maximum working chamber pressure and working chamber volume.
The error bars shown represent the standard deviation.

Results and discussion


Working chamber pressure trace
Fig. 3 shows the variations of the working chamber pressure
according to the crank angle with hydrogen enrichment levels
of 0%, 2%, 4%, 5%, 7%, and 10% energy fraction, at a constant

10

11

12

Hydrogen Energy Fraction (% )

(b)
30

Crank Angle for Maximum Working Chamber Pressure(DATDC)

"

(a)

Maximum Working Chamber Pressure (Bar)

so the spark timing was kept at 15 CA BTDC of the original


engine. Such a fixed ignition timing control strategy was also
adopted by other researchers for exploring the performance of
hydrogen-enriched engines [5,21e25].
The energy fraction of hydrogen in the total intake gas is
defined as energy fraction, calculated as follow [26]:

27.5
25
22.5
20
17.5
15
12.5
10
7.5
5
2.5
0

10

11

12

Hydrogen Energ Fraction(% )

22

Fig. 4 e a) Maximum working chamber pressure at


different hydrogen energy fraction levels; b) the relevant
Crank angle of maximum working chamber pressure at
different hydrogen energy fractions.

20
18

Pressure(bar)

16
14

0% H2
2% H2
4% H2
5% H2
7% H2
10% H2

12
10
8
6
4

2
-70

TDC

+90

BDC

-90

Crank Angle(Degree)

Fig. 3 e Working chamber pressure trace for 0%, 2%, 4%, 5%,
7% and 10% hydrogen energy fractions.

spark timing of 15 DBTDC and at LOL. According to the pressure trace data, the working chamber pressure increases as
the hydrogen fraction increases. Because of high flame speed
(six times faster than that of gasoline), a larger heating value
and diffusivity of hydrogen, with hydrogen addition, the peak
working chamber pressure is raised earlier and is much closer
to TDC than the gasoline fueled engine pressure trace, but
after reaching its peak value, pressure drops more quickly
than the pure gasoline fueled engine, which means the combustion duration and post-ignition temperature is reduced
with hydrogen addition. This yields a more isochoric heat
release theoretically yielding better efficiency in the energy
conversion process. The sudden jump in pressure around
90 (Fig. 3) is due to the sensor coming into contact with the
following working chamber.

8530

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 8 5 2 5 e8 5 3 4

Brake thermal efficiency


The thermal efficiency of an engine is an important factor in
determining the performance of an automobile [15] and its
economics [5].
30
28

Brake Thermal Efficiency (%)

26
24
22
20
18
16
14
12
10

10

11

Hydrogen Energy Fraction (% )

Fig. 5 e Brake thermal efficiency at different hydrogen


energy fractions.

12

15
14
13

SAE Corrected Brake Power (Kw)

At lean conditions, the fuel is harder to ignite than at


stoichiometric condition which is why the spark timing has to
be advanced at LOL [5]. However, the spark timing in this
experiment is fixed at 15 DBTDC, which makes the combustion for pure gasoline suffer. The high diffusivity, the low
flammability limit and the low ignition energy of the hydrogen
improve the combustion of the gasolineehydrogen fuel
mixture. This hydrogen addition makes the air/fuel mixture
more homogenous and increases laminar burning velocity
that results in increased maximum pressure. This increases
the maximum pressure in the working chamber as shown in
Fig. 4a. For pure gasoline, at fixed spark timing, the combustion duration is longer when running close to LOL and the
gasoline in the working chamber is incompletely burnt
resulting in a reduced working chamber combustion temperature and pressure. However, when hydrogen is added the
combustion is shorter and that phenomenon continues as the
hydrogen fraction increases [6].
As is shown in Fig. 4b, when hydrogen is added at fixed
spark timing, it advances the relevant crank angle for the peak
working chamber pressure. Indeed, since the same ignition
timing has been adopted, the increase of peak pressure value
and the shift of its position toward TDC for the blends prove
that combustion speed increases as the hydrogen content in
the blend increases. Furthermore, it was demonstrated by
Kahraman et al., [28] and Ma et al., [22] the engine performance
peaks generally when the crank angle for the maximum cylinder pressure appears around 10e15 DCA ATDC. Hydrogen
appears to shift the maximum peak pressure around this
location in DCA when the hydrogen energy fraction is over 4%.
This is likely due to the low ignition energy and fast flame
speed of hydrogen that improves the combustion by shortening the flame development and propagation durations,
which offer the same effect as increasing spark advance [5].

12
11
10
9
8
7
6
5

10

11

12

Hydrogen Energy Fraction (% )

Fig. 6 e SAE corrected indicated brake power at different


hydrogen energy fractions.

Fig. 5 shows the change in brake thermal efficiency with


variation of the hydrogen energy fraction added to the gasoline at 3000 rpm, at the LOL. As it can be seen in Fig. 5, for pure
gasoline, the brake thermal efficiency of this engine is relatively low. However, when hydrogen is added the brake
thermal efficiency increases with the increase of the hydrogen
fraction to about 28.8% with 10% hydrogen energy fraction.
When pure gasoline is used in the engine, the highest thermal
efficiency is found close stoichiometric conditions near 22%,
and with the increase of excess air found at the LOL, the
combustion becomes incomplete, resulting in a lower power
output and brake thermal efficiency [5].
Therefore, because of hydrogens high diffusion, heating
value (hydrogen is 120.97 MJ/kg, while gasoline is 44.4 MJ/kg),
flame speed (six times higher than that of gasoline), and the
wider flammability range of hydrogen, higher charge homogeneity and shortened burn duration and a more complete
and faster combustion of hydrogen enriched gasolineeair
mixtures at LOL is observed. This contributes to decreased
cooling and exhausts losses, leading to a higher brake thermal
efficiency than the baseline. Indeed, adding hydrogen to the
air/fuel mixture has a significant effect on reducing the flame
quenching distance, and also reduces the combustion duration during expansion, reducing losses during the combustion
by more closely approximating an ideal constant volume heat
release, which leads to an increase in thermal efficiency.
As it was noted by Salanki et al. [17], who ran a rotary engine at 3300 rpm with pure hydrogen, use of hydrogen leads to
an increase in thermal efficiency as compared to gasoline
especially in a lean condition where the equivalence ratio is
ranging between 0.3 and 1.
Compared to the reciprocating engine, the Wankel rotary
engine is known to have higher cooling losses because of a
larger surface to volume ratio of the working chamber and a
longer time required for one-cycle (1. 5 times longer than that
of the equivalent reciprocating engine). Hence, for higher
thermal efficiency it is effective to raise the burning speed
which allows a more isochoric combustion. The hydrogen

8531

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 8 5 2 5 e8 5 3 4

(a)

(a) 600

100

575
550

Brake Specific Fuel Consumption(g/kw-h)

90

Volumetric Efficiency (%)

80

70

60

50

525
500
475
450
425
400
375
350
325
300
275
250

40

225
200

30

10

11

Hydrogen Energy Fraction (% )

(b)

10

11

12

Hydrogen Energy Fraction (% )

12

(b) 30
29

100

28
95

27
26

Brake Thermal Efficiency (%)

90

Volumetric Efficiency (%)

85
80
75
70

25
24
23
22
21
20
19

65

18
60

17
16

55
50
70

15
300
75

80

85

90

95

100

325

350

375
400
425
450
475
Brake Specific Fuel Consumption(g/kw-h)

105

500

525

550

Rotor Temparature(Degree Celcius)

SAE corrected indicated brake power


To insure that repeatable data could be collected on a day with
different ambient conditions, the power output was corrected
using SAE standard J1349. Fig. 6 shows the SAE corrected brake
power. As expected, running the engine on gasoline at the
LOL, reduces the power produced by the engine because less
fuel is burned completely. Therefore, when hydrogen is added
there is a slight improvement at these operating conditions.
The brake power of a Wankel rotary engine fueled on
hydrogen enriched gasoline showed an improvement in performance as compared to pure gasoline. Compared to reciprocating engines, at high speed the Wankel rotary engine

840

1800

Exhaust Temperature
Maximum Working Chamber Temperature

820

1700

800

1600

780

1500

760
-2

10

Hydrogen Energy Fraction (%)

Fig. 9 e Exhaust gas temperature and the maximum


working chamber temperature for different hydrogen
energy fractions.

1400
12

Maximum Working Chamber Temperature (Degree Celcius)

addition contributes to raising the burn velocity, which is a


solution to obtaining stronger squish flow and better output
performance which becomes especially true at high speed.
This confirms the fact that even a small amount of hydrogen
addition effectively improves the engine lean burn capability
in the Wankel engine as shown by the data collected.

Fig. 8 e a) Brake specific fuel consumption at different


hydrogen energy fractions; b) thermal efficiency
dependence on brake specific fuel consumption.

Exhaust Temperature (Degree Celcius)

Fig. 7 e a) Volumetric efficiency at different hydrogen


energy fractions; b) volumetric efficiency dependence on
rotor temperature.

8532

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 8 5 2 5 e8 5 3 4

produces more overall power due to less parasitic power losses moving air in and out of the engine and smaller frictional
losses because of the absence of a reciprocating mass. Further
the Wankel has a high charging efficiency due to longer intake
and exhaust strokes, and the absence of intake and exhaust
valves causing gas flow resistance. Also, since hydrogen has
fast burning characteristics, it is expected that better torque
results at high speed operation because in this condition the
engine needs a fuel which burns more quickly. Furthermore,
the engine brake torque increases with the increase of the
hydrogen energy fraction. Contributing to this torque increase
is the complete burn up of fuel as enabled by the smaller
quenching distance of the hydrogen fuel. The Wankel engine
has historically suffered from a large relative area and related
quenching by its rotor housing interface resulting in higher
squish flow. Decreasing the quenching distance by adding
hydrogen enables previously unburnt fuel to attain combustion adding to the torque and the related efficiency.

Volumetric efficiency
Volumetric efficiency is a parameter related to the load and
temperature on the Wankel. An increase in the temperature
was indicative of a decrease in the density of the mixture inlet
into the combustion chamber, reducing volumetric efficiency.
Since this engine is air-cooled, the operating temperature inside the housing is dependent on load rather than being
thermostatically controlled to a set value as in liquid-cooled
engines. As the operating temperature increases in the
housing there is less liquid fuel vaporized in the working
chamber, and the charge cooling decreases, thus reducing the
volumetric efficiency. As can be seen in Fig. 7a, as the
hydrogen energy fraction increases, there is less liquid fuel
vaporized inside the working chamber, which leads to less
charge cooling, resulting in lower volumetric efficiency.
Additionally, adding hydrogen increases the combustion
temperature, resulting in higher temperature of the rotor and
working chamber causing a lower volumetric efficiency
compared to the same equivalence ratio as shown in Fig. 7b.
Previous studies made on pure hydrogen fueled rotary engine
air cooled, have shown the same behavior of the engine
regarding the effect of the volumetric efficiency and the rotor
temperature [17]. It should also be noted that as a low-density
gaseous fuel, hydrogen displaces a significant amount of air
further lowering the volumetric efficiency.

Brake specific fuel consumption (BSFC)


Fig. 8a shows the brake specific fuel consumption for the
Wankel engine running on pure gasoline and hydrogen
enriched 0%, 2%, 4%, 5%, 7%, and 10% energy fraction of gasoline at LOL and 3000 rpm. According to Fig. 8a, the BFSC decreases from 506 g/kWh for pure gasoline to 312 g/kWh with
an addition of 10% by energy fraction of hydrogen which
represents a drop of about 38%, and this is attributed mainly to
an increased thermal efficiency (see Fig. 8b) and the higher
specific energy value of hydrogen. The higher the mean
effective pressure is, the better the specific fuel consumption
will be. The main reason for this is that a higher mean effective pressure relatively reduces the effect of more constant

frictional losses. Also, the increase of combustion duration


which occurs at low equivalence ratios increases the BSFC.

Exhaust temperature and the working chamber maximum


temperature
As can be seen in Fig. 9, at the LOL, when hydrogen enrichment increases, the exhaust temperature dropped and is
proportional with the increase of hydrogen energy fraction. At
the LOL, because of the increased efficiency due to shorter
burn duration, the additions of hydrogen lead to a fast and
complete combustion of the fuel/air mixtures, less fuel is
burnt during the expansion stroke which decreases the postignition combustion duration and decreases the exhaust losses as manifested by exhaust temperature. This phenomenon
was also noted by other researchers [5]. This proves that
hydrogen is effective for reducing the exhaust losses, especially at lean conditions [29].
The calculated maximum working chamber temperature
increases as the hydrogen energy fraction increases as shown
in Fig. 9 and the maximum working chamber temperature
affects the NOx emissions. These results were also seen by Ji
et al. [29], where the in-cylinder temperature increased with
hydrogen addition at the same manifold pressure as
compared to pure gasoline engine operation.

Emissions
Fig. 10a illustrates the brake specific emissions of NOx at the
LOL, 3000 rpm and WOT. As is shown, the brake specific NOx
emissions increase with hydrogen blending levels. Moreover,
the brake specific NOx emissions rise from 0.29 g/kWh for
pure gasoline to 0.69 g/kWh for a hydrogen energy fraction of
10%, which represents an increase of over 100% of brake
specific emissions of NOx. As it was seen previously, this is
due to the increase of the working chamber pressure and
temperature that is activated by the fast burning velocity and
high flame temperature of hydrogen which tends to stimulate
the formation of thermal NOx.
However, as it can be seen in Fig. 10b, the HC emissions
decrease from 0.0309 g/kWh for pure gasoline to 0.0046 g/kWh
for hydrogen enriched at 10% hydrogen energy. This is a drop
of 85% in HC brake specific emissions. Indeed, when the pure
gasoline is burnt, as the fuel/air mixture approaches the LOL
there is less fuel burnt and more air which leads to an
incomplete combustion and some partial misfiring which increases the production of a hydrocarbon emissions for pure
gasoline. On the other hand, because hydrogen has a high
flame speed, low ignition energy, wide flammability and
diffusivity and small quenching distance, HC emissions
caused by incomplete combustion decrease. As explained
previously flame quenching occurs when a flame propagates
close to the chamber wall or a crevice in the engine. The cooler
wall acts as a heat sink, preventing complete combustion in
the region close to the wall or within the crevice, increasing
the emissions of HC and CO. This phenomenon is more pronounced in the case of the Wankel engine because of its geometry. Adding hydrogen has a significant effect on reducing
the flame quenching distance of the fuel mixture and the
flame can propagate much closer to the chamber wall than

8533

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 8 5 2 5 e8 5 3 4

that of gasoline, which reduces HC emissions caused by the


crevice effect [7].
As presented in Fig. 10c, the CO brake specific emissions
decrease with the increase of hydrogen energy fraction. It can
be seen in that Fig. 10c that the CO brake specific emissions
drop from 2.409 g/kWh for pure gasoline to 0.86 g/kWh for 10%
hydrogen which represents a drop of about 64% in CO brake
specific emissions. That is explained by the fact that because
the maximum working chamber temperature increased with
hydrogen addition the oxidation reaction of CO into CO2 is
stimulated. Moreover, because the gasoline is partly replaced
by hydrogen that does not contain a carbon atom, the
hydrogen enriched gasoline exhausts less carbonaceous
emissions than the original gasoline engine [7,5].
Fig. 10d shows the variation of the CO2 brake specific
emissions against the hydrogen energy fraction of 0%, 2%, 4%,
5%, 7%, and 10% at WOT, 3000 rpm and LOL. It can be found
from Fig. 10d that the CO2 brake specific emissions is reduced
from 116 g/kWh for pure gasoline to 74 g/kWh with the 10%
energy fraction of hydrogen added fuel mixture, which represents a drop of about 36%. This is due to both the fact that
hydrogen is carbon free and the fact that efficiency increases.

(a)

Conclusion
The Wankel rotary engine is a promising alternative to the
reciprocating engine, especially for hybrid applications.
Because of the geometry of the Wankel engine, the concept of
hydrogen enrichment is well suited for the improvement of its
performance. It has been conclusively proven that a small
amount of hydrogen by energy fraction improves the performance and emissions with exception of NOx for a Gasoline
Wankel Rotary Engine at LOL, 3000 rpm, fixed spark timing
and WOT. The main conclusions resulting from this paper are
summarized as follows:
1) At fixed spark timing and LOL, the pressure trace indicated
an increase in the peak working chamber pressure with
advancing the relevant crank angle for peak pressure of a
gasoline Wankel engine with the increase of hydrogen
energy fraction added to the fuel mixture, which increases
the maximum working chamber temperature.
2) The brake thermal efficiency of the original gasoline
Wankel rotary engine was enhanced by about 28% over the

(c)

2.75
0.9

2.5

Brake Specific Emissions CO (g/kWh)

Brake Specific Emissions NOx(g/kw-h)

0.8
0.7
0.6
0.5
0.4
0.3
0.2

2.25
2
1.75
1.5
1.25
1
0.75
0.5

0.1

0.25

0
-2

10

12

Hydrogen Energy Fraction (% )

0.04

10

11

12

10

11

12

(d) 130

0.035

120

0.03

0.025

0.02

0.015

0.01

0.005

Hydrogen Energy Fraction (% )

Brake Specific Emissions CO2 (g/kWh)

Brake Specific Emissions HC(n-hexan Equiv)(g/kw-h)

(b)

110

100

90

80

70

60

Hydrogen Energy Fraction (% )

10

11

12

50

Hydrogen Energy Fraction (% )

Fig. 10 e Brake specific emissions at different hydrogen energy fractions: a) NOx; b) HC; c) CO; d) CO2.

8534

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 8 5 2 5 e8 5 3 4

baseline by adding 10% energy fraction of hydrogen in the


gasoline fuel mixture, decreasing the brake specific fuel
consumption and the exhaust temperature of the engine.
3) Because of the air cooling system of the engine, the volumetric efficiency is proved to be highly sensitive to the
operating temperature of the engine (the rotor) and volumetric efficiency decreased with hydrogen enrichment.
This is also due to the low volumetric energy density of
hydrogen displacing intake air.
4) The performance of the original gasoline Wankel engine as
indicated by the Bmep, Torque and Power are improved as
the energy fraction of hydrogen increases in the fuel
mixture at the conditions tested.
5) Because the increase of the peak working chamber pressure
and temperature, the NOx brake specific emissions are
raised by 137% as the energy fraction of hydrogen increased
from 0% to 10%. However, reductions in the brake specific
emissions of HC by 85%, CO by 64% and CO2 by 36% occurred
with increasing hydrogen fraction due to better combustion
of the air/fuel mixture at the conditions tested.

Acknowledgments
The authors thank Moller International for their donation of
the Wankel research engines. The University of California,
Green Transportation Laboratory and the Hydrogen Production and Utilization Laboratory and all their associated
members made this work possible.

references

[1] Brow RK, K Green R. An investigation of a hydrogen fuelled


Wankel engine. In: Hydrogen energy progress V2 (conf 11);
1996. pp. 1601e10.
[2] Bade Shrestha SO, Narayanan G. Landfill gas with hydrogen
addition e a fuel for SI engines. Fuel 2008;87:3616e26.
[3] Mariani Antonio, Prati Maria Vittoria, Unich Andrea,
Morrone Biagio. Combustion analysis of a spark ignition i c.
engine fuelled alternatively with natural gas and hydrogennatural gas blends. Int J Hydrogen Energy 2013;38:1616e23.
[4] Moreno F, Munoz M, Arroyo J, Magen O, Monne C, Suelves I.
Efficiency and emissions in a vehicle spark ignition engine
fueled with hydrogen and methane blends. Int J Hydrogen
Energy 2012;37:11495e503.
[5] Ji Changwei, Wang Shuofeng. Effect of hydrogen addition on
combustion and emissions performance of a spark ignition
gasoline engine at lean conditions. Int J Hydrogen Energy
2009;34:7823e34.
[6] DAndrea T, Henshaw PF, Ting DS-K. The addition of
hydrogen to a gasoline-fuelled SI engine. Int J Hydrogen
Energy 2004;29:1541e52.
[7] Wang Shuofeng, Ji Changwei, Zhang Minyue, Zhang Bo.
Reducing the idle speed of a spark-ignited gasoline engine
with hydrogen addition. Int J Hydrogen Energy
2010;35:10580e8.
[8] Ji Changwei, Wang Shuofeng. Experimental study on
combustion and emissions performance of a hybrid
hydrogenegasoline engine at lean burn limits. Int J Hydrogen
Energy 2010;35:1453e62.

[9] Wang Shuofeng, Ji Changwei, Zhang Bo. Effect of hydrogen


addition on combustion and emissions performance of a
spark-ignited ethanol engine at idle and stoichiometric
conditions. Int J Hydrogen Energy 2010;35:9205e13.
[10] Lindner B, Sjostrom K. Operation of an internal combustion
engine: lean conditions with hydrogen produced in an
onboard methanol reforming unit. Fuel; 1984:1485e90.
[11] Dunn-Rankin D. Lean combustion: technology and control.
Elsevier; 2008.
[12] Ma Fanhua, Wang Yu. Study on the extension of lean
operation limit through hydrogen enrichment in a natural gas
spark-ignition engine. Int J Hydrogen Energy 2008;33:1416e24.
[13] Bain A, Barclay JA, Bose TK, Edeskuty FJ, Fairlie MJ, Hansel JG,
Hay DR, Swain MR. Sourcebook for hydrogen applications.
Montreal, Canada: TISEC Inc.; 1998.
[14] Heywood J. Internal combustion engine fundamentals.
McGraw Hill; 1988.
[15] Yamamoto K. Rotary engine. Hiroshima: Sankaido Co. Ltd.; 1981.
[16] Ribau Joao, Silva Carla, Brito Francisco P, Martins Jorge.
Analysis of four-stroke, Wankel, and microturbine based
range extenders for electric vehicles. Energy Convers Manage
2012;58:120e33.
[17] Salanki PA, Wallace James S. Evaluation of the hydrogen
fuelled rotary engine for hybrid vehicle application. SAE
paper 960232; 1996.
[18] Wakayama Norihira, Morimoto Kenji, Kashiwagi Akihiro,
Saito Tomoaki. Development of hydrogen rotary engine
vehicle. In: WHEC 16/13e16 June 2006 e Lyon France; 2006.
[19] Badr O, Alsayed N, Manaf M. A parametric study on the lean
misfiring limit and knocking limits of gas-fuelled spark
ignition engines. Appl Therm Eng 1998;18:579e94.
[20] Shiomoto GH, Sawyer R, Kelley B. Characterization of the
lean misfire limit; 1978. SAE paper 780234.
[21] DAndrea T, Henshaw PF, Ting DS-K. The addition of
hydrogen to a gasoline-fuelled SI engine. Int J Hydrog Energy
2004;29:1541e52.
[22] Ma F, Wang Y, Liu H, Li Y, Wang J, Zhao S. Experimental
study on thermal efficiency and emission characteristics of a
lean burn hydrogen enriched natural gas engine. Int J Hydrog
energy 2007;32:5067e75.
[23] Ma F, Wang J, Wang Y, Wang Y, Li Y, Liu H. Influence of different
volume percent hydrogen/natural gas mixtures on idle
performance of a CNG engine. Energy Fuels 2008;22:1880e7.
[24] Ji Changwei, Wang Shuofeng. Effect of hydrogen addition on
lean burn performance of a spark-ignited gasoline engine at
800 rpm and low loads. Fuel 2011;90:1301e4.
[25] Mariani Antonio, Prati Maria Vittoria, Unich Andrea,
Morrone Biagio. Combustion analysis of a spark ignition i.c.
engine fuelled alternatively with natural gas and
hydrogenenatural gas blends. Int J Hydrogen Energy
2013;38:1616e23.
[26] Wang Shuofeng, Ji Changwei, Zhang Bo. Effects of hydrogen
addition and cylinder cutoff on combustion and emissions
performance of a spark-ignited gasoline engine under a low
operating condition. Energy 2010;35:4754e60.
[27] Ma F, Wang Y, Wang J, Ding S, Wang Y, Zhao S. Effect of
combustion phasing, combustion duration and their cyclic
variations on spark e ignition engine efficiency energy fuels.
Energy Fuels 2008;22:3022e8.
[28] Kahraman N, Ceper B, Akansu SO, Aydin K. Investigation of
combustion characteristics and emissions in a spark-ignition
engine fueled with natural gasehydrogen blends. Int J
Hydrogen Energy 2009;34:1026e34.
[29] Ji Changwei, Wang Shuofeng, Zhang Bo. Combustion and
emissions characteristics of a hybrid hydrogen-gasoline
engine under various loads and lean conditions. Int J
Hydrogen Energy 2010;35:5714e22.

You might also like