You are on page 1of 8

Journal of the Energy Institute xxx (2015) 1e8

Contents lists available at ScienceDirect

Journal of the Energy Institute


journal homepage: http://www.journals.elsevier.com/journal-of-the-energy-
institute

Combustion characteristics and performance of a methanol fueled


homogenous charge compression ignition (HCCI) engine
Chunhua Zhang, Han Wu*
School of Automobile, Chang'an University, Xi'an 710064, People's Republic of China

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

Article history: Under the pressure of energy crisis and environmental pollution, a new combustion mode, homogenous
Received 26 November 2014 charge compression ignition (HCCI) combustion, combined with a renewable alcohol fuel, methanol, was
Received in revised form studied on a four-stroke HCCI engine. Intake charge temperature, fueleair equivalence ratio and engine
8 March 2015
speed, were varied during the experiments. The results show that the intake charge temperature in-
Accepted 9 March 2015
uences both the combustion phasing and heat release rate signicantly, which is the most sensitive
Available online xxx
parameter among tested parameters for methanol HCCI combustion. Equivalence ratio has obvious in-
uence on IMEP and cyclical variation but has little inuence on thermal efciency. The engine speed
Keywords:
Methanol scopes are dominated by operation conditions and the optimized speed where highest thermal efciency
Homogenous charge compression ignition obtained increases gradually with equivalence ratio increasing. The maximum thermal efciency can be
(HCCI) obtained when CA50 locates near 7.5  CA and combustion duration is less than 11  CA on the experi-
Boundary condition mental setup.
Combustion characteristics 2015 Energy Institute. Published by Elsevier Ltd. All rights reserved.
Engine performance

1. Introduction

Since the energy crisis and environmental pollution have been severe problems all over the world, conventional internal combustion
engine will be difcult to meet the stricter emission and higher fuel economy regulations in the future. Therefore, a new engine combustion
mode, homogeneous charge compression ignition (HCCI) was discovered as a promising alternative combustion mode due to its ability to
reduce emissions and increase thermal efciency simultaneously.
HCCI combines the advantages of spark-ignition combustion engine, typically associated with gasoline engines, and compression-
ignition engine, typically associated with diesel engines. The soot and NOx could be reduced as the result of low local combustion tem-
perature in the chamber due to the lean premixed airefuel mixture. In addition, the feature of compression-ignition and less throttle loss
during the intake process favors a high thermal efciency for HCCI combustion. Even though HCCI has the great advantages, it also faces
challenges such as combustion phase controlling, operation range extension [1,2]. The HCCI engine lacks direct combustion phasing control
approaches such as spark on SI engines and fuel direct injection on diesel engines, its ignition timing is only dominated by the chemical
kinetic reactions of the airefuel mixture during the compression stroke [3,4]. In macroscopic view, it is mainly affected by properties of fuel
and the histories of in-cylinder pressure, and temperature and concentration of airefuel mixture. Therefore, the physical and chemical
properties of fuel and the intake temperature of engine are extremely important for HCCI combustion.
Ibrahim et al. [5] investigated the effect of intake temperature on a hydrogen HCCI engine. They found that with intake temperature
increasing the combustion phasing advanced signicantly. Xie et al. [6] regarded that the intake preheating could signicantly inuence the
economy and emission of HCCI engine, and found a supplementary of intake preheating by waste heat recovery is able to provide 8e12% fuel
economy improvement throughout the typical load range of HCCI combustion. In addition, the thermal stratication also inuences the
HCCI combustion process. Lim et al. [7] studied the effect of thermal stratication in combustion chamber on a rapid compression machine,
and found that under the thermal stratication, the ame luminosity started from high temperature region to low temperature region, and

* Corresponding author. Tel.: 86 13679124323.


E-mail address: whanzi@163.com (H. Wu).

http://dx.doi.org/10.1016/j.joei.2015.03.005
1743-9671/ 2015 Energy Institute. Published by Elsevier Ltd. All rights reserved.

Please cite this article in press as: C. Zhang, H. Wu, Combustion characteristics and performance of a methanol fueled homogenous charge
compression ignition (HCCI) engine, Journal of the Energy Institute (2015), http://dx.doi.org/10.1016/j.joei.2015.03.005
2 C. Zhang, H. Wu / Journal of the Energy Institute xxx (2015) 1e8

both the low temperature reaction starting point and high temperature reaction starting point were advanced. In addition, it was found the
effect of engine speed and wall temperature, intake strategies could affect the thermal stratication in the chamber [8,9].
Meanwhile, a variety of fuels also have been studied on HCCI engines, including traditional fossil fuels, such as natural gas, gasoline and
diesel [10,11], pure hydrocarbon fuels, such as n-heptane, iso-octane, hydrogen, acetylene [12e15], and dual fuels, such as n-heptane/natural
gas and gasoline/ethanol [16,17]. Even though HCCI engine displays a strong compatibility on fuel type, almost all the fuels used on con-
ventional engine can be used in the HCCI engines, meanwhile it was found that HCCI engines displayed different combustion and emission
properties when fueled with different fuels [18]. Tanaka et al. [19] found HCCI combustion process showed typical two stage reactions by
using heavy hydrocarbon, and the combustion was very sensitive to octane number of fuel, the ratio of rst stage of energy release decreased
and the ignition delay increased with octane number increasing. L et al. [20] found that the cycle to cycle variation could be improved by
decreasing octane number. Among the fuels, alcohols were regarded as one of promising alternative fuels which can take full advantage of
HCCI combustion. The relative high octane number of alcohols allows HCCI combustion to be carried out on a high compression ratio engine
to achieve higher thermal efciency. Meanwhile, with the use of alcohols the HCCI engines are able to run at a larger range of overall
equivalence ratio comparing with the use of gasoline [21]. In addition, Varol et al. [22] reported that the addition of alcohols (methanol,
ethanol and n-butanol) in traditional fossil fuels was effective to reduce CO and HC emission, and the methanol was the most effective one.
Furthermore, the methanol is found to be able to ease knocking phenomenon on SI engine due to its high latent heat value, which is also
helpful to extend the operational range of HCCI combustion [23].
Methanol can be produced either from biomass by fermentation method or from coal by the chemical reaction way. The latter gains the
most market occupancy currently in China because of the rich coal reserves, and both of them can be regarded as relative long term solutions
for the transportation sector. In previous studies [24e26], authors have found the alcohols are very effective to reduce the soot emission due
to their oxygen containing in the fuel.
In order to combine the advantages from engine side and the fuel side, experiments were carried out on an HCCI engine fueled with
methanol in the current work. And the important parameters that inuence HCCI performance including intake charge temperature, engine
speed and mixture concentration were considered in the experiments.

2. Test apparatus and procedure

2.1. Experimental setup

An HCCI engine was achieved based on a two-cylinder, four-stroke diesel engine, CT2100Q. One of the cylinders was remained as the
conventional diesel engine mode, while the other one was modied to run as HCCI mode. The specications of the engine are shown in Table
1.
The experimental setup is shown in Fig. 1. In order to load the engine, an eddy current dynamometer, with the rated power of 25 kW and
maximum speed of 10,000 r/min, was connected to the ywheel. An independent intake system and fuel supply system were used for each
cylinder, where an electric heating system is used to control the HCCI intake charge temperature. To ensure mixture formation, a port fuel
injector made by Bosch (which actuates only if the pulse width is greater than 2.5 ms) was installed in the intake manifold 60 cm upstream
of the intake valve, and the injection timing was 80  CA before intake TDC. An ECU is used to control injection timing and duration, then the
different fueleair equivalence ratios can be obtained by considering intake temperature. The injection duration varied from 4 ms to 8 ms
under different equivalence ratios. A piezo-electric pressure transducer (Kistler 6052A) is used to record in-cylinder pressure, with the
timing resolution of 0.25  CA. Then the signal was further analyzed and saved by combustion analyzer, CB566. The accuracy of the
equipment employed is shown in Table 2.

2.2. Experimental procedure

In order to warm up engine and start HCCI combustion mode smoothly, the engine always starts with the diesel cylinder. Once the oil and
coolant temperature reached the preset values (95  C and 85  C), and the engine speed and intake charge temperature were adjusted to the
desired values, the combustion mode of the engine then shifted from diesel mode to HCCI mode by stopping diesel injection for diesel
cylinder and simultaneously starting methanol injection for HCCI cylinder. The data can be recorded when the engine run stably. It should be
noted that only one of the two cylinders can be provided fuel while engine running. The details of experimental setup and test procedures
can be referred to [27].

2.3. Denition of parameters

Based on the average of 60 continuous cycles in-cylinder pressure, the heat release rate (HRR) and accumulated heat release rate (AHRR)
are calculated by a zero-dimensional model.

Table 1
Specications of HCCI cylinder.

Item (units) Parameters Item (units) Parameters


Bore  stroke (mm) 100  105 Injection timing ( CA BTDC) 80
Displacement volume (mL) 824.7 Injection pressure (bar) 3.8
Compression ratio 17:1 IVO ( CA BTDC) 17
Combustion chamber u type IVC ( CA ABDC) 43
Ratio of crank radius to connecting rod 0.32 EVO ( CA BBDC) 47
Intake pressure ambient EVC ( CA ATDC) 17

Please cite this article in press as: C. Zhang, H. Wu, Combustion characteristics and performance of a methanol fueled homogenous charge
compression ignition (HCCI) engine, Journal of the Energy Institute (2015), http://dx.doi.org/10.1016/j.joei.2015.03.005
C. Zhang, H. Wu / Journal of the Energy Institute xxx (2015) 1e8 3

Fig. 1. Experimental setup. 1. Electric heater; 2. Intake charge temperature controller; 3. Temperature sensor; 4. Injector; 5. Injection ECU; 6. Hall sensor; 7. Injection pressure
regulator; 8. Fuel lter; 9. Fuel tank and pump; 10. Photo-electricity transducer; 11. Crank-angle encoder; 12. Computer 1; 13. Combustion analyzer; 14. Charge amplier; 15.
Pressure transducer; 16. Temperature sensor; 17. Exhaust gas analyzer; 18. Computer 2; 19. Dynamometer and control cabinet.

The timing (crank angle) of 10%, 50% and 90% of AHRR are referred to as CA10, CA50 and CA90, which represent start of combustion (SOC),
combustion location and burn out timing, respectively. Furthermore, the combustion duration dened as the period between CA10 and
CA90.
During the HCCI combustion process, the maximum in-cylinder pressure is signicantly sensitive to combustion phasing, heat release
prole etc. Thus, the stability of maximum in-cylinder pressure is able to indicate cycle-to-cycle variation level. The coefcient of variation,
COVPmax, is calculated by Formula (1).

COVPmax sPmax P  100% (1)
max

where, sPmax is the standard deviation of the maximum combustion pressure, P max is the mean maximum pressure.
The indicated thermal efciency (ITE) is calculated by Formula (2).

ITE P i Vh H  m   100% (2)
u cyc

where, Pi is the indicated mean effective pressure (IMEP), Vh is displacement volume, Hu is the lower heating value of tested fuel, and mcyc is
the fuel mass of each cycle.

3. Results and discussion

3.1. Combustion characteristics

3.1.1. In-cylinder pressure and HRR


Fig. 2 shows the cylinder pressure and HRR curves under various intake charge temperatures, fueleair equivalence ratios (ER) and engine
speeds in (a), (b) and (c), respectively. It can be seen that the intake temperature is the most sensitive parameter of them for combustion
pressure and HRR. Both of the maximum combustion pressure and HRR decrease with the intake temperature signicantly. Under the
identical speed and ER (1250, 0.4), the maximum combustion pressure decreases from 60 bar to 43 bar when the intake temperature drops
from 160  C to 140  C, the combustion pressure peak almost disappears when the intake temperature further drops to 135  C. At the same
time, the combustion phasing also retarded remarkably with the intake temperature reducing. This is because the combustion of the air-
efuel mixture becomes deteriorated with the intake temperature reducing. In HCCI combustion mode, there is neither the external energy
source to ignite the airefuel mixture like spark in gasoline engines, nor injection of high active fuel in high temperature chamber envi-
ronment, the mixture is only ignited by chemical reaction due to compression. If the initial temperature of the intake mixture is too low, the
temperature at the end of compression is still not high enough to trigger the fast combustion reaction. The reason why the HCCI combustion
is signicantly sensitive to the intake temperature is that the chemical reaction rates of the eair-fuel mixture, which controls the HCCI
combustion, increase exponentially with ambient temperature, so any little change of the ambient temperature could lead to a signicant
different result.
Fig. 2(b) shows the effects of ER on combustion pressure and heat release rate. It can be observed that the ER has slightly inuence on the
combustion phasing but has remarkably inuence on the maximum combustion pressure and HRR. With the ER increasing, combustion
phase almost does not change, while the maximum combustion pressure and maximum heat release rate increase signicantly. As known
that, the chemical reaction during the compression stroke of HCCI engine is mostly controlled by temperature near the end of compression
stroke. Thus, the combustion phasing can not be changed obviously once the intake temperature and compression ratio of HCCI engine are
determined. In HCCI combustion model, it is regarded almost all the airefuel mixture reaches the combatable condition at the same time
and burn-out within a short period. Hence, it makes sense that the richer airefuel mixture containing more energy and exhibits a higher
maximum combustion pressure and higher maximum heat release rate.

Please cite this article in press as: C. Zhang, H. Wu, Combustion characteristics and performance of a methanol fueled homogenous charge
compression ignition (HCCI) engine, Journal of the Energy Institute (2015), http://dx.doi.org/10.1016/j.joei.2015.03.005
4 C. Zhang, H. Wu / Journal of the Energy Institute xxx (2015) 1e8

Table 2
The accuracy of the equipment employed.

Equipment Accuracy
Intake temperature control system 1  C
Cylinder pressure sensor 0.5 bar
Crank angle sensor 0.25  CA
AVL 4000 for airefuel equivalence ratio 5%
Dynamometer for engine speed 3 r/min

Fig. 2. The impacts of Tin, ER and speed on cylinder pressure and HRR.

It can be noted in Fig. 2(c) that the combustion pressure and heat release rate are also inuenced by the engine speed. The in-cylinder
pressure peak and heat release peak increase with the engine speed increasing. The most important reason is heat loss through cylinder wall
during the compression stroke. With the engine speed increasing, the accumulated heat loss during the compression stroke reduced
signicantly because of the shorter cyclic period, which increases the in-cylinder temperature at the end of compression stroke. Therefore,
the chamber could achieve a higher maximum pressure at higher engine speed due to less heat loss through the cylinder wall. The effects of
engine speed are also conrmed by the combustion phasing, less heat loss under higher engine speed condition corresponds to an advanced
combustion phasing which resulted by relative higher in-cylinder temperature in the compression stroke. Furthermore, the delayed
combustion phasing under lower combustion speed condition also decreases maximum combustion pressure due to the larger in-cylinder
volume in combustion period.

3.1.2. SOC and combustion duration


Fig. 3 shows CA10, CA50 and the combustion duration of HCCI combustion under various intake charge temperatures, equivalence ratios
and engine speeds.
It can be observed from Fig. 3(a) and (b) that with the increase of ER, both CA10 and CA50 retard while the combustion duration de-
creases slightly. However, Dec et al. [28] found that the kinetic rates leading to ignition were almost unaffected by the concentration of
airefuel mixture, as indicated by the similar SOC of all fuel loads in his calculation result [29]. Thus the delay of CA10 and CA50 under higher
ER condition in current work is probably resulted by evaporative cooling. Even though the injector for the HCCI cylinder is mounted on the
intake pipe 60 cm away from the inlet valve and the intake pipe is equipped with heaters, it still cannot guarantee the injected fuel in the
intake pipe has been evaporated completely and heated up to the pre-set temperature before the intake valves especially for the richer
airefuel mixture cases. The combustion duration becomes shorter when ER increased is because the larger amount of released heat ac-
celerates the chemical reaction rates during the combustion progress.
Another observation in Fig. 3(a) and (b) is that at higher intake temperature (160  C), CA10 and CA50 are earlier and the combustion
duration is shorter than those at lower intake temperature condition (140  C). Increasing temperature is the most effective way of tested
parameters to increase the chemical reaction rates that signicantly enhance the HCCI combustion.
The effects of the engine speed on the combustion phasing are shown in Fig. 3(c) and (d). It can be seen that with the increase of engine
speed the CA10 and CA50 are slightly advanced, and combustion duration is also reduced. At higher engine speed condition, the heat loss
during compression stokes is reduced due to shorter cycle period, and airefuel mixing is enhanced due to the stronger turbulence in the
intake pipe. Those two facts promote the HCCI combustion, advance the combustion phasing and shorten combustion duration.

Please cite this article in press as: C. Zhang, H. Wu, Combustion characteristics and performance of a methanol fueled homogenous charge
compression ignition (HCCI) engine, Journal of the Energy Institute (2015), http://dx.doi.org/10.1016/j.joei.2015.03.005
C. Zhang, H. Wu / Journal of the Energy Institute xxx (2015) 1e8 5

Fig. 3. The impacts of Tin, ER and speed on the ignition timing and combustion duration.

Fig. 3(c) and (d) also exhibit that all CA10 and CA50 curves of lower ER (0.4) cases under different engine speeds are below the curves of
higher ER (0.5) cases, which means the combustion phasing is advanced by reducing the ER, and the combustion duration curve of lower ER
is above that of higher ER. The reason is the evaporation effect of injected fuel in the pipe at richer mixture case.

3.1.3. Cyclic variation


Fig. 4 displays the traces of COVPmax versus the engine speed and the equivalence ratio. As shown in Fig. 4(a), different airefuel con-
centrations correspond to different HCCI operation scopes.
In this paper, the engine speed where cyclic variation reaches the lowest point regarded as the optimized speed.
In the shown conditions, the optimized speed is 1200 r/min when the equivalence ratio is 0.37; the optimized speed increases to 1300 r/
min when ER is increased up to 0.4; and the optimized engine speed appears up to 1500 r/min when the ER is further increased up to 0.5.
Thus, the engine speed with the lowest cyclic variation increases gradually with the ER.
HCCI combustion is very sensitive to the engine operational boundary, any little cyclic change of airefuel mixture condition during the
compression stoke is able to lead to a signicant variation in HCCI combustion and engine performance.
A higher engine speed may cause a larger intake uctuation, inducing a greater variation of ER, which could deteriorate the cyclic
variation. However, a higher engine speed also brings a more intense airow into cylinder, benecial to the combustion process by
increasing airefuel mixing quality. For the richer mixture cases, the impact of the latter overweighs that of the former, so the optimized
engine speed tends to be higher. At the same time, the effect of injected fuel in the intake pipe of richer mixture could compensate the less
heat loss at higher engine speed, which is another important reason why optimized speed is higher for the rich mixture cycles.
As shown in Fig. 4(b), the cyclic variation under richer mixture conditions is much lesser than that under leaner mixture conditions. That
probably because lean mixture has a problem on combustion in the area near cylinder wall, which is a potential reason to bring an
inhomogeneous heat distribution in the chamber. The COVpmax curves versus ER under different intake temperature, 140  C and 160  C, are
very close to each other except in the very lower ER stage.

Fig. 4. The impacts of Tin, ER and speed on cyclic variation.

Please cite this article in press as: C. Zhang, H. Wu, Combustion characteristics and performance of a methanol fueled homogenous charge
compression ignition (HCCI) engine, Journal of the Energy Institute (2015), http://dx.doi.org/10.1016/j.joei.2015.03.005
6 C. Zhang, H. Wu / Journal of the Energy Institute xxx (2015) 1e8

Fig. 5. The impacts of Tin, ER and speed on IMEP.

3.2. Engine performance

3.2.1. Indicated mean effective pressure (IMEP)


The impacts of the intake charge temperature, fueleair equivalence ratio and engine speed on IMEP can be found in Fig. 5.
As shown in Fig. 5(a), IMEP decreases slightly with engine speed under both tested equivalence ratios, 0.5 and 0.4, but the IMEP for higher
ER is signicant higher than that for lower ER. The main reason why the IMEP vary with engine speed is the variation of the combustion
phasing, which further affects the work quantity output from engine. The higher IMEP corresponding to higher ER is obviously because more
cyclic input energy.
As shown in Fig. 5(b), IMEP increases with ER under both 140  C and 160  C, but the IMEP under higher intake temperature 160  C is even
higher than that of lower intake temperature. In fact, the main reason for IMEP increasing is the more energy input to the cylinder, because
either more injected fuel under higher ER condition or the proper combustion phasing.

3.2.2. Indicated thermal efciency (ITE)


Fig. 6 exhibits ITE in the conditions of different engine speeds, equivalence ratios and intake charge temperatures. Apparently, ITE of HCCI
engine is observably higher than that of spark ignition engine even higher than that of diesel engine.
From Fig. 5(a), it can be found that ITE decrease with engine speed under both 0.4 and 0.5. The main reason is due to the variation of
combustion phasing, which is similar to the effect of engine speed on IMEP. Moreover, it also can be seen from the gure that there is no
obviously different on ITE between two tested equivalence ratios, which probably means that ITE is not as sensitive to engine speed as ER.
This may be conrmed by the data in Fig. 5(b), which shows that ER has little impact on ITE. However, similar to the impact of intake
temperature on IMEP, ITE of higher intake temperature, 160  C, is much higher than that of lower intake temperature, 140  C. The reason is
also the same, higher intake temperature means more energy to enter into the cylinder, and the proper combustion phasing means to
convert more energy into work. It should be noted that it takes energy to heat the intake charge into higher temperature. In current work,
heat cable mounted on the intake pipe is used to warm up airefuel mixture to the pre-set point.
Combustion phasing plays an important role on engine performance. Fig. 7 shows the relationship between CA50, combustion duration
and ITE.
In Fig. 7(a), under the engine speed of 1000 r/min, ITE rst slightly increases then decreases with CA50 slowly, while combustion
duration rst decreases and then increases with CA50. At the higher engine speed (1250 r/min) in Fig. 7(b), ITE decreases with CA50 while
combustion duration increases with CA50 continuously. It can be found that higher ITE mostly happen together with short combustion
duration. That make sense, short combustion duration make the HCCI combustion process closer to a high efciency combustion mode,
constant volume combustion. But, it should be noted that the precondition is combustion should be carried out in the proper timing.
Furthermore, the shorter combustion duration also reduce heat loss during the combustion process that will increase the ITE. In the
condition of 1100 r/min, the maximum ITE occurs when CA50 locates around 8  CA while the combustion duration is about 11  CA. Under
the condition of 1250 r/min, the maximum ITE happens when CA50 is around 7  CA while the combustion duration is around 11  CA.

Fig. 6. The impacts of Tin, ER and speed on ITE.

Please cite this article in press as: C. Zhang, H. Wu, Combustion characteristics and performance of a methanol fueled homogenous charge
compression ignition (HCCI) engine, Journal of the Energy Institute (2015), http://dx.doi.org/10.1016/j.joei.2015.03.005
C. Zhang, H. Wu / Journal of the Energy Institute xxx (2015) 1e8 7

Fig. 7. The impacts of CA50 and the combustion duration on ITE.

Therefore it can be concluded in some degree that high ITE can be reached at CA50 near 7.5  CA while combustion duration is less than
11  CA of the specic engine in the correct work.

4. Conclusions

(1) With the increase of the intake temperature, fueleair equivalence ratio, and engine speed, the maximum cylinder pressure and heat
release rate are increased, the combustion phases also are advanced. Among all of the tested parameters, intake temperature is the most
sensitive one to inuence the methanol combustion characteristics.
(2) With the increase of the intake charge temperature and engine speed, the ignition timing (CA10 and CA50) are advanced while the
combustion duration is shortened; while, with the increase of the equivalence ratio, the ignition timings are delayed and the com-
bustion duration is shortened.
(3) The engine speed scopes change with operation conditions and the optimized speed where lowest cyclic variation obtained increases
with equivalence ratio. Relative rich airefuel mixture is helpful to reduce the cyclic variation.
(4) The IMEP increases with the increase of intake charge temperature and fueleair equivalence ratio.
(5) Higher ITE usually corresponds to the short combustion duration. For the methanol HCCI engine in this work, high ITE can be obtained at
CA50 near 7.5  CA while combustion duration is less than 11  CA.

Acknowledgment

This study was supported by the China Special Fund for Basic Scientic Research of Central Colleges (No. CHD2012TD007) and Doctor
Postgraduate Technical Project of Chang'an University (No. 2014G5220007).

References

[1] F. Zhano, T. Asmus, D. Assanis, J. Dec, J. Eng, P. Najt, Homogeneous Charge Compression Ignitions (HCCI) Engines: Key Research and Development Issues, SAE Paper PT-94,
2003.
[2] C. Zhang, H. Wu, The simulation based on CHEMKIN for homogeneous charge compression ignition combustion with on-board fuel reformation in the chamber, Int. J.
Hydrogen Energy 37 (5) (2012) 4467e4475.
[3] T. Shudo, Y. Shima, T. Fujii, Production of dimethyl ether and hydrogen by methanol reforming for an HCCI engine system with waste heat recovery - continuous control
of fuel ignitability and utilization of exhaust gas heat, Int. J. Hydrogen Energy 34 (2009) 7638e7647.
[4] M. Yao, Z. Zheng, H. Liu, Progress and recent trends in homogeneous charge compression ignition (HCCI) engines, Prog. Energy Combust. Sci. 35 (2009) 398e437.
[5] M.M. Ibrahim, A. Ramesh, Investigations on the effects of intake temperature and charge dilution in a hydrogen fueled HCCI engine, Int. J. Hydrogen Energy 39 (2014)
14097e14108.
[6] H. Xie, L. Li, T. Chen, H. Zhao, Investigation on gasoline homogeneous charge compression ignition (HCCI) combustion implemented by residual gas trapping combined
with intake preheating through waste heat recovery, Energy Convers. Manag. 86 (2014) 8e19.
[7] O.T. Lim, N. Iida, The investigation about the effects of thermal stratication in combustion chamber on HCCI combustion fueled with DME/n-butane using rapid
compression machine, Exp. Therm. Fluid Sci. 39 (2012) 123e133.
[8] N.P. Komninos, The effect of thermal stratication on HCCI combustion: a numerical investigation, Appl. Energy 139 (2015) 291e302.
[9] S.A. Kaiser, M. Schild, C. Schulz, Thermal stratication in an internal combustion engine due to wall heat transfer measured by laser-induced uorescence, Proc.
Combust. Inst. 34 (2013) 2911e2919.
[10] A. Sen, G. Litak, K. Edwards, C. Finney, C. Daw, R. Wang, Characteristics of cyclic heat release variability in the transition from spark ignition to HCCI in a gasoline engine,
Appl. Energy 88 (2011) 1649e1655.
[11] D. Ganesh, G. Nagarajan, Homogeneous charge compression ignition (HCCI) combustion of diesel fuel with external mixture formation, Energy 35 (2010) 148e157.
[12] J. Andrae, R. Head, HCCI experiments with gasoline surrogate fuels modeled by a semi-detailed chemical kinetic model, Combust. Flame 156 (2009) 842e851.
[13] J. Gomes Antunes, R. Mikalesen, A. Roskilly, An investigation of hydrogen-fuelled HCCI engine performance and operation, Int. J. Hydrogen Energy 33 (2008) 5823e5828.
[14] K. Kobayashi, T. Sako, Y. Sakaguchi, S. Morimoto, S. Kanematsu, K. Suzuki, T. Nakazono, H. Ohtsubo, Development of HCCI natural gas engines, J. Nat. Gas Sci. Eng. 3 (2011)
651e656.
[15] S. Swami Nathan, J. Mallikarjuna, A. Ramesh, Effects of charge temperature and exhaust gas re-circulation on combustion and emission characteristics of an acetylene
fuelled HCCI engine, Fuel 89 (2010) 515e521.
[16] M. Fathi, R. Khoshbakhti Saray, M. David Checkel, The inuence of exhaust gas recirculation (EGR) on combustion and emissions of n-heptane/natural gas fueled
homogeneous charge compression ignition (HCCI) engines, Appl. Energy 88 (2011) 4719e4724.

Please cite this article in press as: C. Zhang, H. Wu, Combustion characteristics and performance of a methanol fueled homogenous charge
compression ignition (HCCI) engine, Journal of the Energy Institute (2015), http://dx.doi.org/10.1016/j.joei.2015.03.005
8 C. Zhang, H. Wu / Journal of the Energy Institute xxx (2015) 1e8

[17] H. Wu, R. Wang, D. Ou, Y. Chen, T. Chen, Reduction of smoke and nitrogen oxides of a partial HCCI engine using premixed gasoline and ethanol with air, Appl. Energy 88
(2011) 3882e3890.
[18] L. Starck, B. Lecointe, L. Fortil, N. Jeuland, Impact of fuel characteristics on HCCI combustion: performances and emissions, Fuel 89 (2010) 3069e3077.
[19] S. Tanaka, F. Ayala, J. Keck, A reduced chemical kinetic model for HCCI combustion of primary reference fuels in a rapid compression machine, Combust. Flame 133
(2003) 467e481.
[20] Lyu Xi, W. Chen, Z. Huang, A fundamental study on the control of the HCCI combustion and emissions by fuel design concept combined with controllable EGR. Part 1.
The basic characteristics of HCCI combustion, Fuel 84 (2005) 1074e4083.
[21] R. Maurya, A. Agarwal, Experimental investigations of performance, combustion and emission characteristics of ethanol and methanol fueled HCCI engine, Fuel Process.
Technol. 126 (2014) 30e48.

[22] Y. Varol, C. Oner,
H. Oztop, S. Altun, Comparison of methanol, ethanol, or nebutanol blending with unleaded gasoline on exhaust emissions of an SI engine, Energy
Sourc. Part A 36 (2014) 938e948.
[23] J. Vancoillie, L. Sileghem, S. Verhelst. Development and Validation of a Knock Prediction Model for Methanol-fuelled SI Engines. SAE Paper 2013-01-1312.
[24] H. Wu, K. Nithyanandan, N. Zhou, T. Lee, C.F. Lee, C. Zhang, Impacts of acetone on the spray combustion of acetone-butanol-ethanol (ABE)-diesel blends under low
ambient temperature, Fuel 142 (2015) 109e116.
[25] H. Wu, K. Nithyanandan, N. Zhou, T. Lee, C.F. Lee, C. Zhang, Spray and combustion characteristics of neat acetone-butanol-ethanol, n-butanol, and diesel in a constant
volume chamber, Energy Fuels 28 (2014) 6380e6391.
[26] Han Wu, Chunhua Zhang, Boqi Li, Timothy H. Lee, Chia-Fon Lee, Investigation on spray and ame lift-off length of acetone-butanol-ethanol-diesel blend in a constant
volume chamber, J. Eng. Gas Turbines Power 137 (9) (2015) 091501e0915018.
[27] C.H. Zhang, L. Xue, J. Wang, Experimental study of the inuence of l and intake temperature on combustion characteristics in an HCCI engine fueled with n-heptane, J.
Energy Inst. 87 (2014) 175e182.
[28] P. Kelly-Zion, J.A. Dec, Computational study of the effect of fuel-type on ignition time in HCCI engines, Proc. Combust. Inst. 28 (2000) 1187e1194.
[29] J. Dec, M. Sjo berg. A parametric study of HCCI combustion- the sources of emissions at low loads and the effects of GDI fuel injection. SAE Paper 2003-01-0752.

Please cite this article in press as: C. Zhang, H. Wu, Combustion characteristics and performance of a methanol fueled homogenous charge
compression ignition (HCCI) engine, Journal of the Energy Institute (2015), http://dx.doi.org/10.1016/j.joei.2015.03.005

You might also like