Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 109

Title: Thematic on homogeneous charge compression ignition (HCCI) engine

Acknowledgments

Table of Contents
CHAPTER 1: INTRODUCTION...................................................................................6
1.1. The reason for choosing the thesis topic..............................................................6
1.2. The object and scope of the research...................................................................8
1.2.1. Object of the research....................................................................................8
1.2.2. Scope of the research......................................................................................8
1.3. Research method...................................................................................................8
1.4. Outline research....................................................................................................8
CHAPTER 2: OVERVIEW OF HCCI ENGINE........................................................10
2.1. History of HCCI engines.....................................................................................10
2.2. What is HCCI?....................................................................................................10
2.3. Advantages and disadvantages of HCCI engine...............................................12
2.3.1. Advantages....................................................................................................12
2.3.2. Disadvantages...............................................................................................15
2.4. Gasoline HCCI/CAI combustion engines..........................................................16
2.4.1. Introduction..................................................................................................16
2.4.2. Two-stroke Gasoline CAI engine................................................................19
2.4.3. Four-stroke Gasoline CAI engine................................................................23
2.5. Diesel HCCI/CAI combustion engines...............................................................28
2.5.1. Formation of NOx and soot.........................................................................28
2.5.2. Low-load combustion efficiency..................................................................29
2.5.3. High-load limits............................................................................................31
2.5.4. Fuel effect......................................................................................................32
CHAPTER 3: THEORETICAL BASIC OF HCCI ENGINE....................................35
3.1. HCCI engine principle and working cycle........................................................35
3.2. HCCI combustion control strategies..................................................................39
3.2.1. Intake thermal management........................................................................39
3.2.2. Exhaust gas recirculation.............................................................................41
3.2.3. Variable compression ratio..........................................................................43
3.2.4. Variable valve actuation..............................................................................46
3.3. Homogeneous charge preparation strategies....................................................49
3.4. The HCCI engine’s structural features.............................................................54
3.4.1. Performance of homogeneous charge compression ignition (HCCI) engine
with common rail fuel injection.............................................................................54
3.4.2. HCCI Fuel requirement...............................................................................71
3.5. HCCI engine uisng variable valve timing..........................................................84
3.6. HCCI using exhaust gas recirculation system...................................................89
CHAPTER 4: RESEARCH ON SKYACTIV-X ENGINE.........................................97
4.1. Introducing Mazda’s Skyactiv technology........................................................97
4.2. Advantages of Skyactiv-X engine.......................................................................98
4.3. Skyactiv-X with Spark Controlled Compression Ignition...............................98
4.3.1. Using compression effects created by flame propagation..........................99
4.3.2. Fuel density distribution within the air-fuel mixture................................99
4.3.3. Controlling the air-fuel mixture to prevent abnormal combustion........100
4.4. Value provided by Skyactiiv-X engine.............................................................101
4.4.1. Dramatically improved output performance and responsiveness..........101
4.4.2. Dramatic improvement in fuel economy...................................................102
CHAPTER 5: CONCLUSION AND RECOMMENDATIONS...............................103
5.1. Conclusion..........................................................................................................103
5.2. Recommendations...............................................................................................103
REFERENCES............................................................................................................. 103
General Abbreviations
Acronym Definition
ATAC Active thermo-atmosphere combustion
AVT Active valve train
BMEP Brake mean effective pressure
BS British standard
BSFC Brake specific fuel consumption
CA Crank angle
CAD Crank angle degree
CAI Combustion auto ignition
CI Compression ignition
CLD Chemical luminescence detector
CR Compression ratio
EGR Exhaust gas recirculation
EGT Exhaust gas trap
EHN Ethyl hexyl mitrate
EVC Exhaust valve closing
FTM Fast thermal management
FVVT Fully variable valve timing
HCCI Homogeneous charge compress ignition
HCCI Homogeneous charge with compression ignition
HCSI Homogeneous charge with spark ignition
HRR Heat release rate
HTRA High temperature reaction appearance
HTRE High temperature reaction end
IC Internal combustion
IEGR Internal exhaust gas recirculation
IMEP Indicated mean effective pressure
IQT Ignition quality tester
ISFC Indicated specific fuel consumption
IVC Intake valve closing
IVO Intake valve open
LTC Low temperature combustion
LTHR Low temperature heat release
LTRA Low temperature reaction appearance
LTRD Low temperature reaction duration
LTRE Low temperature reaction end
MK Modulated kinetics
NVO Negative valve overlap
PFI Port fuel injection
PM Particular matter
PRR Pressure rise rate
PVO Positive valve overlap
RCCI Reactivity controlled compression ignition
RCCI Reactivity controlled compression ignition
SFC Specific fuel consumption
SI Spark ignition
SOI Start of injection
SOR Start of reaction
SVC Saab variable compression
TDC Top dead center
TVC Throttle control valve
TRG Trapped residual gases
VCR Variation of compression ratio
VVA Variable valve actuation
VVLei Variable valve lift exhaust and intake
VVT Variable valve timing
VVTe Variable valve timing exhaust
VVTei Variable valve timing exhaust and intake
WOT Wide open throttle
ZNCT Zone of negative temperature coefficient

List of Figures
Figure 1.1 Share of source of global GHG emission in 2016 by main sector....................6
Figure 1.2 CO emission by sector, World.........................................................................7
Figure 1.3 The major emission standards in different countries and areas........................7

List of Tables
CHAPTER 1: INTRODUCTION
1.1. The reason for choosing the thesis topic
From ancient times to the present day, in every place where human beings exist in
the world, people have been always trying to find ways to transport themselves and goods
robustly and efficiently over long distances. The advent of the automobile has
significantly changed the world and the lifestyle of human beings. Internal combustion
engines, which are mainly powered by petroleum-derived liquid fuels, have been the
primary source of transportation capacity for the past century and are expected to
continue to be so in the future. Even though, it is difficult to image a world without
automobiles today, there is no doubt that the first car would have been impossible to
create without the contribution of the internal combustion engine. This engine is due to
the efforts of an inventor named Nikolaus Otto, who gave his name to the ‘Otto' cycle,
which has been used in hundreds of millions of internal combustion engines worldwide.
Growing fossil fuel consumption along with increases in fuel cost and a focus on the
reduction of carbon dioxide (CO2) emissions in the transportation sector motivate the
need for increased efficiency from the internal combustion engine (ICE). Figure 1.1
illustrates the breakdown of global greenhouse gas (GHG) emission in 2016.

Figure 1.1 Share of source of global GHG emission in 2016 by main sector.
(https://en.wikipedia.org/wiki/Greenhouse_gas_emissions)
As can be seen, the transportation sector account for 17% of global GHG
emission. This sector primarily involves fossil fuels burned for road, rail, air, and marine
transportation. Almost all (95%) of the world's transportation energy comes from
petroleum-based fuels, largely gasoline and diesel. Moreover, automotive engine and
fuels are facing challenges to lower pollution in order to improve local area air quality as
well as improve fuel conversion performance, which reduces CO2 emissions and
therefore the possibility of global warming, as shown in figure 1.2. The global
breakdown for CO2 is similar to that of total greenhouse gas – electricity and heat
production, followed by manufacturing, construction and transport. The grand challenges
for engine researchers are to develop technologies for maximizing engine efficiency and
reducing emissions.

Figure 1.2 CO emission by sector, World.

(https://ourworldindata.org/emissions-by-sector)

Figure 1.3 The major emission standards in different countries and areas.

(Pielecha J, Merkisz-Guranowska A, Jacyna-Gołda I. A new ecological research:


real driving emissions. J KONES 2014;21)
Figure 1.3 presents the main emissions standards in different countries and
different areas of the world. As can be seen from figure 1.3, the emissions standards have
become increasingly stringent. In order to meet strict emissions regulations and economic
demand. There are two types of internal combustion engines: Spark-ignition (SI) and
compression ignition (CI). Conventional CI diesel engines have higher fuel conversion
efficiency than SI engines. Diesel engines typically emit comparatively higher NOx and
particulate matter (PM). In order to further improve the fuel economy and reduce
emissions of the IC engine, alternative combustion systems have been proposed and
investigated. In particular, a new generation of combustion methodology known as
Controlled Auto Ignition (CAI) or Homogeneous Charge Compression Ignition (HCCI).
When the homogenous charge mixtures reach chemical activation energy, combustion
begins in multiple spots. Compared to conventional compression ignition (CI) and spark
ignition (SI) combustion methods. The HCCI mode means the compression ignition
(auto-ignition/self-ignition) of homogeneous (premixed) fuel and air mixture in the
engine cylinder. Hence, HCCI operating mode exhibits characteristics from both CI and
SI engines. The HCCI operating mode is similar to conventional SI engine in charge
preparation strategy (both modes use premixed charge) and similar to CI engines in
combustion initiation process (in both modes combustion starts with autoignition).
1.2. The object and scope of the research
1.2.1. Object of the research
The research object is homogeneous charge compression ignition (HCCI) engine.
1.2.2. Scope of the research
The aim of the project is to evaluate the potential of two stroke HCCI and four
stroke HCCI operations.
The specific objectives are:
1. To explore and research HCCI (homogeneous charge compression ignition)
combustion for stable, fuel efficient and low emission part-load operation in
gasoline engine and diesel engine.
2. To study some effective techniques and controlling strategies, such as fuel
management, homogeneous charge preparation, exhaust gas recirculation and
variable valve timing.
3. Overview on fuel properties and influence on HCCI combustion.
1.3. Research method
Researching previous scientific reports on this motive, mainly based on
information sources from foreign universities in the world. Collect information
related to this engine. Consult the teacher's opinion on the content of the study.
1.4. Outline research
The project structure consists of the following parts:
Chapter 1: Introduction
Chapter 2: Introduces HCCI with its benefits, drawbacks and ideal applications
and compare HCCI against spark-ignited and diesel engines.
Chapter 3: Describes the combustion control variables and strategies for
HCCI/CAI engines.
Chapter 4: Presents Skyactiv-X gasoline engine.
Chapter 5: Includes the conclusion and recommendations for further work.
CHAPTER 2: OVERVIEW OF HCCI ENGINE
2.1. History of HCCI engines
Amongst the numerous research papers published over the last decade, the
homogeneous charge compression ignition (HCCI) or controlled auto-ignition (CAI)
combustion has often been considered a new combustion process in reciprocating internal
combustion engines. However, it has been around perhaps as long as the spark ignition
(SI) combustion in gasoline engine and compression ignition (CI) combustion in diesel
engines. In the case of diesel engines, the hot-bulb 2-stroke or 4-stroke oil engines or
diesel engines were patented and developed over 100 years ago, wherein kerosene, or raw
oil was injected onto the surface of a heated chamber (hot-bulb), which was separated
from the main cylinder volume, very early in the compression stroke, giving plenty of
time for fuel to vaporize and mix with air, as shown figure 2.1.

Figure 2.4 Hot bub engine 2 stroke


1. Hot-bub; 2. Injection; 3. Piston; 4. Connecting rod; 5. Crankshaft; 6. Crankcase; 7. Air filter;
8. Exhaust pipe.
During the start-up, the hot bulb was heated on the outside by a torch or a burner.
Once the engine had started, the hot-bulb was kept hot by the burned gases within. The
bulb was so hot that the injected fuel vaporized immediately when it got in contact with
the surface. Later design placed injection through the connecting passage between the
hot-bulb and the main chamber so that a more homogeneous mixture could be formed,
resulting in auto-ignited homogeneous charge combustion. In the case of gasoline
engines, the auto-ignited homogeneous charge combustion had been observed and was
found responsible for the ‘after-run’/ ‘run-on’ phenomenon that many drivers had
experienced with their carburetor gasoline engines in the sixties and seventies, when a
spark ignition engine continued to run after the ignition was turned off. The same type of
combustion was also found to be the cause of ‘dieseling’ or hot starting problems
encountered in the early high compression gasoline engines.
2.2. What is HCCI?
Homogeneous Charge Compression Ignition (HCCI) is a form of internal
combustion in which well-mixed fuel and oxidizer (typically air) are compressed to the
point of auto-ignition. As in other forms of combustion, this exothermic reaction releases
energy that can be transformed in an engine into work and heat. HCCI combines
characteristics of conventional gasoline engine and diesel engines.

Figure 2.2. Different combustion modes in internal combustion engine.


Gasoline engines combine homogeneous charge (HC) with spark ignition (SI),
abbreviated as HCSI. Modern direct injection diesel engines combine stratified charge
(SC) with compression ignition (CI), abbreviated as SCCI. As in HCSI, HCCI injects fuel
during the intake stroke. However, rather than using an electric discharge (spark) to ignite
a portion of the mixture, HCCI raises density and temperature by compression until the
entire mixture reacts spontaneously. Stratified charge compression ignition also relies on
temperature and density increase resulting from compression. However, it injects fuel
later, during the compression stroke. Combustion occurs at the boundary of the fuel and
air, producing higher emissions, but allowing a leaner and higher compression burn,
producing greater efficiency. Controlling HCCI requires microprocessor control and
physical understanding of the ignition process. HCCI designs achieve gasoline engine-
like emissions with diesel engine-like efficiency. HCCI engines achieve extremely low
levels of oxides of nitrogen emissions (NOx) without a catalytic converter. Hydrocarbons
(unburnt fuels and oils) and carbon monoxide emissions still require treatment to meet
automobile emissions control regulations. Recent research has shown that the hybrid
fuels combining different re-activities (such as gasoline and diesel) can help in
controlling HCCI ignition and burn rates. RCCI (reactivity control compression ignition)
has been demonstrated to provide highly efficient, low emissions operation over wide
load and speed ranges.
2.3. Advantages and disadvantages of HCCI engine
2.3.1. Advantages
 Improved combustion stability
Figure 2.3 shows a comparison between SI and CAI combustion stability at
a light load operating point (about 2 bar IMEP). With SI combustion, a lot of
fluctuations of the maximum in-cylinder pressure can be observed in correlation
with corresponding IMEP fluctuations. This deterioration of the combustion
comes from cycles with poor or incomplete combustion followed by cycles
with higher IMEP.

Figure 2.3. Compared in-cylinder pressure traces and IMEP distribution between SI and CAI
combustion (Habert, 1993).
 Improved pollutant emissions and fuel efficiency
This deterioration of the combustion quality results in both high unburned
hydrocarbon (HC) emissions and high fuel consumption as shown respectively
Figure 2.4. Benefits of CAI/AR combustion (obtained by the AR exhaust
control valve) on 2-stroke engine brake specific HC emissions 4000 rpm
(Ishibashi, 1993).

Figure 2.5. Effect of CAI (obtained by transfer port throttling) on 2-stroke engine stability – 1.2
bar BMEP @ 2000 rpm (Duret, 1996).
On the contrary, a remarkably stable cycle to cycle combustion can be achieved
with CAI as shown by the in-cylinder peak pressure as well as by the IMEP distribution
(Fig 2.1). Thanks to CAI, the incomplete or poor combustion disappear and the
theoretical ideal specific fuel consumption (SFC) curve can be approached. For the
same reason, the HC emissions can be significantly improved as shown by below
Figure by Ishibashi, 1993. In this figure, the HC emissions are reduced for engine loads
below 4 bar BMEP and the reduction is maximum when the engine load goes down to 1
bar BMEP.

 Similar advantages when CAI is combined with DI


All the figures show that for this light load operating point (1.2 bar BMEP @
2000 rpm), the combustion becomes very stable above 60% of transfer throttling. The
standard deviation of IMEP representative of the cyclic fluctuations of the combustion
drops drastically to remarkably low values between 0.06 and 0.08 bar. A similar
significant benefit is obtained in BSFC which is reduced from about 420 g/kWh to a
minimum of 375 g/kWh. The improvement in HC emissions is even more spectacular
since the already rather low level of HC obtained with direct injection (32 g/kWh) is
divided by a factor of 4 for the best transfer throttling settings (between 60 and 80%
throttling rate) reaching a level similar to four-stroke engine HC emissions.

At this light load operating point, the NOx emissions are already
remarkably low without CAI thanks to the two-stroke engine’s inherent
advantages of mainly the high dilution by internal EGR and the low IMEP per
cycle because of a combustion frequency every cycle (Duret, 1990).
Nevertheless, it is interesting to remark that running in CAI doesn’t increase the
NOx. A tendency towards reduction is even observed.
2.3.2. Disadvantages
 The light load limit of CAI range (irregular combustion and misfiring)
When the load becomes too low, the amount of energy released per combustion
cycle is too low to maintain a sufficiently high level of temperature of the internal EGR.
To extend the CAI range to the lowest load needs, therefore, to try to minimize as much
as possible the mixing between the internal EGR and the fresh mixture while keeping the
in-cylinder pressure at the beginning of the compression as high as possible. In this sense,
the engine speed has a major effect. When the engine speed increases, the time for
mixing between the fresh charge and the internal EGR decreases. This clearly explains
why it is rather easy to reach very light loads CAI at medium to high engine speeds while
it is much more difficult for engine speeds lower than 2000 rpm and especially at idle.
The transfer throttling is probably the most efficient to minimize the internal mixing and
increase the EGR stratification (Duret, 1996) and the exhaust throttling is more efficient
to combine this stratification effect with an increase of the in-cylinder pressure (Ishibashi,
1993). Using these solutions, it is generally possible in the best engine configurations to
reach light load auto-ignition below 2000 rpm for engine loads as low as 0.5 bar BMEP.
Below such loads, it is difficult to maintain the CAI auto-ignition. Poor combustion and
misfiring start to occur. It is a shame because to reach idle in CAI combustion would
probably solve all the part load emissions problems of the two-stroke engines. This is the
reason why a lot of researchers have worked in the direction of extending the CAI range
up to idle. Nevertheless, from what is known in the scientific knowledge, idle operation
in CAI has probably never been obtained until now. Even if in theory it seems feasible, it
would probably require a combination of solutions such as both transfer and exhaust
throttling combined with higher engine compression ratio and perhaps a more appropriate
fuel formulation.
 The high load limit (too strong uncontrolled combustion)
When the engine load increases in CAI operation, the combustion start
progressively earlier and earlier. Even if the engine can continue to run in auto-ignition,
the heat release rate becomes very high with more than 50% of the mass fraction burnt
before TDC. In these conditions, the combustion noise becomes very strong like severe
knock and the NOx emissions quickly rise. There is no interest in keeping the auto-
ignition mode (which is no longer ‘controlled’ as the CAI ‘controlled auto-ignition’ name
would suggest!) and it is generally better to switch to the SI combustion mode. The next
paragraph will show that this high load mode switch between CAI and SI can be rather
easily managed in a two-stroke engine.
 The management of the transition between SI and CAI combustion modes
It is clear that two types of combustion mode transition have to be distinguished,
firstly the transition between irregular SI to stable CAI at light load and secondly the
transition between CAI to higher load conventional SI. To simplify the engine
management system, the spark ignition is generally maintained for all engine speeds and
loads even if it has no effect inside the CAI range. As explained by Ishibashi, the
transition between auto-ignition combustion to higher load conventional SI is easily and
naturally achieved without any jump in engine torque. The main requirement is to make
the transition range as short as possible to avoid the occurrence of combustion cycles
with strong too early and noisy auto-ignition. To achieve this, when the load increases,
the control valve (for example, the exhaust throttling or the transfer throttling valve) has
to move quickly from the throttled position to the fully open position without staying in
an intermediate partially throttled position. On the other hand, the transition from CAI to
the light load less regular SI combustion is more difficult. In this light load transition
area, and for the same intake air delivery ratio (same intake throttle position), the engine
gives more torque in auto-ignition than in spark ignition.
2.4. Gasoline HCCI/CAI combustion engines
2.4.1. Introduction
A controlled auto ignition (CAI) gasoline engine is an attempt to combine the high
efficiency of a diesel engine with the low emission of an SI engine. In the 4-stroke spark-
ignition engines, controlled auto-ignition (CAI) or HCCI combustion has been the focus
of comprehensive research over the last decade for improved part-load fuel, fuel
consumption and low exhaust emissions [1]. CAI combustion is a process where a
premixed mixture of fuel, air and recirculated gases auto-ignites at multi points around
the piston top dead center when the charge temperature reaches auto-ignition
temperature. It typically features homogeneous charge, auto ignition and low temperature
combustion.

Figure 2.8. In-cylinder pressure traces of CAI and SI operation at the same operating condition.
Figure 2.8 compares P-V diagrams for SI and CAI engine operation. It may be
noted that the combustion CAI engine is much faster and it approaches close to constant
volume combustion.
 History background
The origins of CAI/HCCI combustion can be traced back to 1930s, when Russian
scientist Nikolai Semenov and his colleagues began researching a chemical-kinetic
controlled combustion process for IC engines, in order to overcome the limitations
imposed by the physical-dominating processes of SI and CI engines. Semenov and
Gussak discovered that the thermodynamic and chemical conditions of the entire cylinder
charge are similar to those of cool flames of hydrocarbon air mixtures, a more uniform
heat release process should be reached.
In 1979, Onishi et al. [2] and Noguchi et al. [3] were the first to successfully apply
CAI combustion technology to two-stroke gasoline engines. CAI combustion was found
to improve fuel economy, achieve incredibly low NOx emissions, and reduce unburned
hydrocarbon (HC) emissions in two-stroke gasoline engines. For two-stroke gasoline
engines, Onishi et al. discovered that the attainable CAI (or ATAC) area was constrained
by inadequate combustion temperature at low loads and knocking combustion at high
loads. Onishi et al. also found that the attainable CAI (or ATAC) region in their two-
stroke gasoline engine was limited by insufficient combustion temperatures at low loads,
and knocking combustion at high loads. Other factors such as heat rejection to chamber
walls, and degree of cycle-to-cycle variation in the gas exchange process were also found
to be important.
In 1983, Najt and Foster [4] were the first to use CAI combustion technology in a
four-stroke gasoline engine and they used experimentally obtained heatelease data to
develop empirical models of homogenous CAI combustion. The ignition process was
found to be controlled by low temperature (below 9500 K ) hydrocarbon oxidation kinetic.
In 1989, Thring [5] investigated the effect of the A/F ratio, EGR rate, fuel type,
and compression ratio on the attainable homogeneous CAI combustion region and engine
out emissions. Thring carried out the premixed compression ignition combustion of
diesel and introduce the definition of HCCI.
In 1999, Christensen et al. [6] achieved homogenous CAI with diesel fuel mucjh
smoke was generated in some cases. But with gasoline almost no smoke was generated at
all. The combustion efficiency for iso-octane and n-heptane decreased with increase
compression ratio.
In 2001, Oakley et al [7] expanded on the research by performing a detailed
analysis of the heat release characteristic of the CAI combustion process in a 4-stroke
gasoline engine. In order to study fueland dilution effects only, a lean-to-stoichiometric
operating region has been defined with regard to A/F ratio and EGR rate.

Figure 2.9. History of production and most advanced prototype CAI 2 stroke engines.
In summary, CAI combustion has been identified as a means of simultaneously
reducing NOx and CO2 emissions for 4-stroke engines. Significant improvements in fuel
economy can be achieved through un-throttled CAI operation at part-load. Since CAI
does not rely on conventional flame propagation, lean flammability limits that normally
apply are irrelevant. Extremely lean, low temperature (< 1800 K) combustion is possible,
leading to massive reductions in NOx formation.
According to the previous studies above, one (or a combination) of four methods
can be used to promote the CAI combustion process in 4 stroke gasoline engines:
- Direct intake heating
- Higher compression ratio
- Variation in fuel blends
- Exhaust gas recirculation (EGR)
 Region of CAI operation
The following results were obtained using a single-cylinder test engine under the
following conditions:
Engine speed: 1500rpm
Airflow: WOT
Inlet charge temperature: 320 ± 1°C
Figure 2.10 shows the CAI region attainable for gasoline engine. The vertical axis
represents the overall A/F ratio of the cylinder charge. The horizontal axis shows
the total gravimetric percentage of EGR in the cylinder [4].

Figure 2.10. Boundary regions for CAI operation for unleaded gasoline
The attainable CAI region is limited by three boundaries [8]:
 Partial burn
 Knock limit
 Misfire

Figure 2.7. Indicated mean effective pressure (imep) values (bar) in the Cai operation range.

2.4.2. Two-stroke Gasoline CAI engine


Controlled Auto Ignition combustion was first demonstrated and accomplished in
the late 1970s to solve the partial burn and misfire at part load of a spark-ignited mixture
in a 2-stroke cycle engine with intake and exhaust ports [2]. Two-stroke CAI combustion
was accomplished by means of residual gas trapping via reduced overlap between
exhaust and intake valve openings.
 Operating range of two-stroke gasoline CAI engine:
Firstly, Figure 2.11 shows a broader range of CAI operation could be
accomplished at engine speeds ranging from 800rpm to 3000rpm and engine loads
ranging from idle to 7.8bar IMEP, which is equal to 15.6bar IMEP on the four-stroke
operation. The NOx emission and real fuel consumption were also very poor in the CAI
operating range. The in-cylinder mixture was too rich as a result of the short-circuiting.
Figure 2.11. Operation range of two-stroke CAI operation with Exhaust Lambda 1.0 [9]
The figure 2.11 describe the operation range of CAI combustion is enveloped by 3 limits:
- The knock limit on the low speed high load boundary
- The gas exchange limit on the high speed and high load boundary
- The misfire limit on the low load boundary.
Secondly, CAI combustion of gasoline with ethanol was suited over a wide range
of speed and load when the exhaust lambda was set to 1.0. [10]

Figure 2.12. Operating range of Two-stroke CAI fuelled with Gasoline, E15 and E85
This figure represents the low load boundary of E85 occurred at higher IMEP
values than that of gasoline and E15 at 800rpm. When the engine speed increased, the
difference in the low load CAI limits between gasoline and its mixture with ethanol
disappeared because of the reduced heat loss and higher thermal loading at higher engine
speeds.
 Performance and emission at exhaust lambda 1.0
Figure 2.13. ISNOx, ISCO, ISHC, and Exhaust temperature over the operating range
As shown in figure 2.13 the NOx emission was very low and the specific fuel
consumption was also low in the CAI operation range. Due to the short-circuiting, the in-
cylinder mixture was too rich for complete combustion when the exhaust lambda was
controlled at 1.0 and hence higher CO and uHC (ISHC) emissions were observed [9].
 Emissions efficiency
In 2011, Y. Zhang et al [10] investigated the engine operation modes through
different combinations of valve timings and durations.
- Mode 1: 4-Stroke throttle-controlled SI mode.
- Mode 2: 4-Stroke intake valve throttle SI mode.
- Mode 3: 4-Stroke positive valve overlap SI mode.
- Mode 4: 4-Stroke negative valve overlap CAI mode.
- Mode 5: 4-Stroke exhaust rebreathe CAI mode.
- Mode 6: 2-Stroke CAI mode.
- Mode 7: 2-Stroke SI mode.
Figure 2.14 illustrates the emission of CO and HC. According t the emission
results, CAI combustion generates similar CO and HC emissions to those of SI operation.
High CO and HC emissions are primarily caused by the poor mixing or substantial
diffusion combustion from any liquid fuel film on the piston crown.
Figure 2.14. ISCO and ISHC of 7 Operation Modes [11]
According to the emissions data shown in Figure 2.15, it is clear from the bar chart
that conventional throttle-controlled SI combustion 4-stroke intake valve throttle SI
combustion produces slightly lower NOx emissions owing to the lower in-cylinder
temperature.

Figure 2.15. NOx emissions of the 7 operating modes


 Combustion efficiency
A comparison of combustion efficiency, calculated from fueling rat and emission
data, represents the quality of fuel mixing and atomization in the fuel in the cylinder
before ignition.
Figure 2.16. Combustion efficiency of the 7 operating modes [10]
It is clear that 4-stroke throttle-controlled SI, positive valve-overlap SI and
negative valve overlap CAI have the higher combustion efficiency. 2-stroke SI has the
lowest combustion efficiency.
2.4.3. Four-stroke Gasoline CAI engine
 Introduction
Controlled Auto-Ignition (CAI) combustion was achieved in a production style 4-
stroke 4-cylinder gasoline engine without intake charge heating or increasing
compression ratio. The CAI engine operation was achieved using substantially standard
components modified only in camshafts to restrict the gas exchange process. The residual
gas trapping method has been presented as one of the most effective and practical
approaches for four-stroke gasoline engines.
 Engine set-up and test procedure
The engine used in this work was a Ford 1.7 Litre Zetec SE 16 valves four-
cylinder port-fuel injection gasoline engine with a pent roof combustion chamber, which
details are given in Table 2.1 [11].

Table 2.1. The engine specification


The intake and exhaust camshafts were equipped with two independent VCT
systems, as shown in Figure 2.17 [12].
Figure 2.17. Ford 1.7L Zetec engine with dual VCTs
The fuel used was a commercial unleaded gasoline of RON 95 complying with the
British standard BS EN 228. During the experiments, the throttle was kept wide open,
and the airflow was varied by charging cam timings, which can be continuous changed
by up to 40 degrees crank angle. Although the engine could be run with slightly lean or
rich mixtures [11], most tests were conducted at lambda 1. To minimize the effect of
coolant temperature, all experiments were performance when the coolant temperature
reached 90 0 C∨higher . Figure 2.18 shows the speed and load range within which CAI
combustion was achieved in the engine.

Figure 2.18. CAI operation range with residual gas trapping


The overall BMEP values at each speed were primarily limited by the restricted
airflow caused by the camshaft reduced valve lifts. As speed increased, there was less
time for the gas exchange process and hence more residuals were trapped in the cylinders
and less fresh charge was inducted, resulting in lower maximum BMEP values. The CAI
operation ranged from 0.4 to 1.1 bar BMEP at 3500 rpm [8]. At a given speed, the load
was varied by changing the cam timings. There was a minimum BMEP at each speed,
which was limited by misfire. At these conditions a lot of residual gases were retained
but the exhaust gas temperature was too low, and the mixture failed to auto-ignite. At
higher speeds, the lower limit was shifted to lower load as the exhaust gas temperature
increased at higher speeds.
According to [11], to evaluate the potential benefit of the application of CAI
combustion to 4-stroke gasoline, a vehicle simulation program created by the Ford Motor
Company was used model a new Ford Mondeo car, with a hypothetical CAI/SI hybrid
engine or standard PFI (port fuel injection) SI engine drove through the European New
Emission Drive Cycle (NEDC). Table 2.2 shows the total percentage change of the
CAI/SI vehicle relative to standard SI engine vehicle in fuel consumption (FC) and
exhaust emissions.

Table 2.2. Comparison between SI/CAI and SI vehicles.


It is clear that there were moderate reductions in FC, CO and NOx emissions, but
higher unburned hydrocarbon over the entire NEDC driving cycle. The most important
reduction took place at 50 km/h and 70 km/h cruise point, when CAI combustion was
operative, but with the highest emission of unburned hydrocarbons.
 Analysis of the In-cylinder residual
The purpose off starting CAI combustion in this study is to advance exhaust valve
closing and residual trapping [12]. Knowing the volume of trapped residual is vital in
analyzing CAI combustion. In the present study, two methods were used to determine the
in-cylinder residual fraction.
In the first approach, the in-cylinder residual fractions were determined using a
one-dimensional gas dynamic code. It provided a thorough stud off the fluid dynamics in
the multi-cylinder engine’s intake and exhaust systems. The experimental heat release
data calculated from the in-cylinder pressure measurements were used to determine the
start and duration of the heat release process in the simulation code. The simulation
program was calibrated by comparing the simulated airflow and engine power with the
experimental values.
In the second approach, the in-cylinder pressure value used for the calculations
was measured at EVC (exhaust valve closing) using a pressure transducer. The cylinder
volume was calculated at EVC, based on the engine geometry and EVC timing. The
temperature was assumed to be the burnt gas temperature at EVC, measured by the
thermocouples installed in the exhaust manifold.
The mass off trapped residualsmr , at EVC was estimated using the ideal law.
P V =m r R T

Where P in-cylinder pressure


V cylinder volume
m r mass of trapped residual

R specific gas constant


T burnt gas temperature
 Analysis of the engine’s performance
The difference in IMEP and BMEP show in figure 2.19 represents the frictional
losses. As expected, the frictional losses increased slightly with engine speed.

Figure 2.19. Effect of residual fractions on MEP values


Figure shows there was a linear correlation between the residual fraction and engine’s
performance, independent of the engine speed. As the engine was operated at WOT, the
mass in the cylinder was more or less the same and only the concentration of the mixture
changing. The more residuals were trapped, the less air/fuel mixture the engine will
breathe in, resulting in lower torque [12].
 Combustion analysis

Figure 2.20. Effect of residual fraction on the start and finish of CAI combustion.
Figure 2.20 shows the crank angles at which 10% and 90% of fuel had burnt,
respectively. Depend on speed and load, the autoignition began between 355 and 364 deg
CA. In general, higher loads resulted in higher residual temperature and an early start of
autoignition. The temperature of the residual before ignition was higher at higher engine
speeds, so the ignition started earlier as well.
 Analysis of emissions
According to the results shown in figure 2.21, The ultra-low NOx emissions at
most CAI operation conditions were confirmed to be associated with low gas temperature
[13]. It is important to maintain a high dilution rate in order to achieve high load activity
with minimal NOx emission. This is accomplished by the use of
supercharging/turbocharging.
Figure 2.21. In-cylinder gas temperature histories
2.5. Diesel HCCI/CAI combustion engines
2.5.1. Formation of NOx and soot.
To appreciate the fundamental of HCCI combustion. it’s firstly necessary to
understand how NOx and soot are formed. The regions of formation of NOx and soot
have been conceptualized in an equivalence ratio-temperature ( θ−T ) map as shown in
figure 2.22. NOx formation occurs at low equivalence ratios and high adiabatic,
equilibrium flame temperature. By lower the flame temperature to level less than 2200K,
suppression of NOx formation can be achieved. Moreover, soot formation occurs in
regions of high equivalence ratios or fuel rich mixture and moderate temperature. Net
soot emission which is a balance between formation and oxidation can be reduced either
by increasing mixing or increasing oxidation.
It is clear that HCCI combustion falls outside the soot and NOx islands. In CAI
combustion as the flame temperature is considerably lower than the conventional diesel
combustion due to lean or diluted mixture, the NOx emissions are low. In addition, the
well-premixed charge present in the cylinder leads to lower soot emissions as well. []
Figure 2.22 Equivalence ratio and local temperature with typical paths of different
combustion concepts.
Low Temperature Combustion (LTC) takes benefit of this fact by adjusting
combustion to take place anywhere in the gray shaded region, while making effort to
ensure that most of the fuel is mixed to θ ≤1 (the HCCI region) before the reactions are
quenched by expansion for maintaining good combustion efficiency. In a conventional
diesel combustion, the fuel and air charge under goes rich combustion of around θ=4 at
the end of the adiabatic mixing process during the ignition delay phases, and then
combustion moves to completion in a stoichiometric (θ=1) diffusion flame. Thermal
NOx is produced when the local in cylinder temperature are in excess of 1800-2000K and
there is enough oxygen available. Spark ignition (SI) combustion also generates
significant amount of NOx emissions, but they are removed by three-way catalysts.
2.5.2. Low-load combustion efficiency
For HCCI engines, the power output is controlled by the fueling rate. To maintain
high thermal efficiencies, this should be done without throttling. As the load is reduced,
either the mixture becomes leaner or the amount of dilution with EGR must be increased.
Figure 2. Depicts a plot of the emissions and combustion efficiency for an HCCI engine
as a function of ∅ for fully premixed operation with no EGR,
Figure 2.23 Combustion efficiency and emissions as a function of fuel loading for fully
premixed mixture off iso-octane and air.
The emissions are given as the percentage of total fuel carbon in each exhaust species, to
remove changes due solely to the amount of fuel supplied. The engine speed was 1200
rpm, and the 50% burn point (CA50) was held constant at top dead center (TDC) since
variations in combustion timing can themselves affect emissions []. As can be shown, for
moderate loads, ∅ ≥0.2, carbon monoxide (CO) and hydrocarbon (HC) emissions are low
and combustion efficiency is high. However, as ∅ is reduced below 0.2, CO emissions
rise dramatically, and the combustion efficiency falls. HC emissions also increase, but to
a lesser extent. Approximately 60% of the fuel carbon remains as CO in the exhaust at an
idle fueling rate of ∅=0.12 (for complete combustion) or lower, HC emissions climb to 10% or
more and combustion efficiencies drop to 62% or less.
Figure 2. illustrate the plot of the emissions and combustion efficiency for ∅ = 0.12 as the
injection timing is swept from 200 CA (crank angle) after TDC (aTDC) intake to 3250 CA,
which is well up the compression stroke. As fuel-injection timing is delayed, the CO
emissions fall, slowly at first and then more rapidly for start of injection (SOI) beyond
3250CA.
Figure 2.24 The effect of varying SOI on combustion efficiency and emissions, using an 8-hole
GDI with iso-octane fuel.
Fuel of iso-octane was kept constant at an intake ∅ = 0.12, and CA50 was fixed at TDC by
adjusting the intake temperature. Early injection essentially gives the same results as the
fully premixed case, with a low combustion efficiency 62% and high levels of CO and
HC emissions. As injection timing is delayed beyond 90 0 CA, combustion efficiency
starts to increase while CO and HC emissions start to decrease. Soot levels remain low.
By retarding the injection timing, the time for fuel/air mixing is reduced, which creates
local regions of higher in-cylinder ∅ that burn hotter, enabling the bulk-gas reactions to
reach completion.
[] Wontae Hwang, John E. Dec and Magnus Sjöberg, Fuel Stratification for Low-Load HCCI
Combustion: Performance & Fuel-PLIF Measurements, SAE Paper 2007-01-4130, 2007.
2.5.3. High-load limits
High-load HCCI is typically limited by an excessive rate of pressure rise during
combustion and the resulting engine knock. Figure 2. Depicts a series of cylinder
pressure curves for various fueling rates in an HCCI engine with CA50 held constant at
TDC.
At a fairly low load, corresponding to ∅ = 0.18, the pressure -rise rate (PRR)
associated with combustion is very moderate. However, the maximum PRR increase
substantially as fueling is progressively increase to ∅ = 0.3. Eventually, the PRR becomes
sufficiently rapid to excite an acoustic resonance causing the engine to knock. This
creates a distinct ripple on the pressure trace, as strongly evident in the ∅ = 0.3 curve.
Because excessive knock causes excessive noise and can lead to engine damage, the
maximum PRR must be kept to an acceptable level. Although the PRR increases
substantially with increasing ∅ , it does not increase nearly as rapidly as it would if the
charge were fully homogeneous. In a real engine, however, it is not possible to produce a
charge that is fully homogeneous in both mixture and temperature, so a single-zone
CHEMKIN (Senkin application) computation was used for the comparison in Figure 2.
Figure 2.25 Experimental cylinder -pressure traces as compared to a homogeneous charge
2.5.4. Fuel effect
The homogeneous charge compression ignition (HCCI) combustion process holds
the promise high fuel economy as well as very low NOx and soot emissions. The
characteristics of autoignition after top-dead-center (TDC) for both single and two-stage
ignition fuels were explored. The single stage ignition fuel was iso-octane and two-stage
ignition fuel was PRF80 (80% iso-octane and 20% n-heptane).
Figure presents a comparison of the cylinder pressure and mass averaged
temperature for iso-octane and PRF80 (a mixture of the gasoline primary reference fuels
consisting of 80% iso-octane and 20% n-heptane) []. As can be seen, PRF80 has low-
temperature heat-release (LTHR) beginning at about 3400CA, which increase its
temperature and pressure before the main hot ignition at about 3670 CA (two-stage
ignition). In contrast, iso-octane shows no significant heat release until the hot ignition at
about 3670 CA (single-stage ignition). Because of the LTHR, PRF80 requires a much
lower intake temperature than iso-octane for the same CA50, as reflected by the lower
temperature at the far left of the plot in Figure 2. (2600 CA). This lower temperature is
advantageous because it means that the intake charge density is higher, so more charge
mass is inducted into the cylinder. Therefore, PRF80 gives a significantly higher power
output for the same charge-mass/fuel (C/F) ratio.
Another benefit, a fuel with some LTHR is that it has a higher heat release rate
before the main hot ignition. This can be seen by a comparison of the temperature in
figure 2.26
Figure 2.26 Temperature and pressure traces for iso-octane and PRF80.
As a result of the greater temperature-rise rate prior to hot ignition, and the reduced
magnitude of cycle-to-cycle temperature fluctuations, PRF80 can tolerate a greater
combustion-phasing retard while maintaining an acceptable standard deviation of the
IMEPg, as shown in figure 2.

Figure 2.27 Standard deviation of IMEP g divided by ( IMEP g−IMEP gmotored as a function of
CA50
Both iso-octane and PRF80 experience increasing cycle-to-cycle variations as the
delay of CA50. However, the increase is small for PRF80 and much greater for iso-
octane. For this study, IMEP g variations above 2% are deemed unacceptable for both
fuels as this corresponds to the appearance of (>1% occurrence) partial-burn cycles. A
cycle is considered a partial-burn cycle if its total heat release is reduced 10% or more,
compared to a well-burning cycle.
The highest loads ( IMEP g) and thermal efficiencies achieved for iso-octane,
PRF80 and PRF60 (60% iso-octane, 40% n-heptane). All data are for a compression ratio
(CR) of 14, except the upper iso-octane dataset, which was taken with CR = 18. As can
be seen, a maximum IMEP g of about 650 kPa was reached for PRF80 and PRF60, while
for iso-octane, the load was limited to an IMEP of 480 or 525 kPa for CR=14 and
CR=18.

\
Figure 2.28 Thermal efficiency vs. IMEP g for iso-octane, PRF80 and PRF60.

For iso-octane, combustion phasing becomes very sensitive to small changes in the
charge temperature. This results in an inability to adequately control combustion phasing
(wandering CA50) for CR = 14, or for CR = 18, small changes in the wall temperature
can cause CA50 to advance to runaway knock or retard to misfire (designated T wall on
Figure 3). The two-stage primary reference fuels are less sensitive to changes in
temperature and are load-limited by other factors. For PRF80, the load can be increased
until the combustion becomes sufficiently hot to produce small amounts of NOx. For
PRF60, at the conditions studied, the high load was limited by the available oxygen. This
is because PRF60 has a high autoignition reactivity, so substantial amounts of EGR must
be used to retard the combustion phasing to maintain an acceptable PRR.

[] Magnus Sjöberg, John E. Dec, Comparing Late-cycle Autoignition Stability for Single-
and Two-Stage Ignition Fuels in HCCI Engines. Proceeding of the combustion Institute,
Vol. 31, pp. 2895-2902,2007.
[] John E. Dec. Advanced compression-ignition engines—understanding the in-cylinder
processes. Proceeding of the combustion Institute 32 (2009) 2727-2742.
CHAPTER 3: THEORETICAL BASIC OF HCCI ENGINE
3.1. HCCI engine principle and working cycle
 Working of HCCI/CAI engines
Similar to a conventional SI engine, in a HCCI/CAI engine the fuel and air are
mixed together either in the intake system or in the cylinder with direct injection. The
premixed fuel and air mixture is then compressed. Towards the end of the compression
stroke, combustion is initiated by auto-ignition in a similar way to the conventional CI
engine. The temperature of the charge at the beginning of the compression stroke has to
be increased to reach auto-ignition conditions at the end of the compression stroke. This
can be done by heating the intake air or by keeping part of the hot combustion products in
the cylinder. Both strategies result in a higher gas temperature throughout the
compression process, which in turn speeds up the chemical reactions that lead to the start
of combustion of homogeneously mixed fuel and air mixtures.

Figure 3.1. Different combustion modes in internal combustion engine


(a) The HCCI combustion process. (b) Combustion process comparison of SI, CI and HCCI
engines in four-stroke cycle.
Although the start of main heat release usually occurs when the temperature
reaches a value of 1050–1100K for gasoline or less than 800K for diesel, many
hydrocarbon components in gasoline and diesel undergo low temperature oxidation
reactions accompanied by a heat release that can account for up to 10% of the total
energy released. The contribution of the low temperature energy release to obtaining
auto-ignition and heat release rate from the HCCI/ CAI combustion depends not only on
the unique chemical kinetics of the fuel used and the dilution strategy, but also on the
thermal conditions or the temperature-pressure history that the mixture goes through
during compression.
 Principle of HCCI engines
In an idealized HCCI/CAI engine, the auto-ignition and combustion will take
place simultaneously throughout the combustion chamber, resulting in a rapid rate of heat
release. In order to prevent the runaway heat release rate associated with the simultaneous
burning of mixtures, HCCI/CAI engines have to run on lean or/and diluted fuel and air
mixtures with burned gases. The heat release characteristics of the HCCI/CAI
combustion can be compared with those of SI and CI combustion [8].

Figure 3.2. Heat release characteristic of Spark Ignition (SI) combustion


In the case of SI combustion, a thin reaction zone or flame front separates the
cylinder charge into burned and unburned regions and the heat release is confined to the
reaction zone. The cumulative heat released in a SI engine is therefore the sum of the heat
released by a certain mass, dm i, in the reaction zone and it can be expressed as
N
Q=∫ q . dmi
1

where q is the heating value per unit mass of fuel and air mixture, N is the number of
reaction zones.
Figure 3.3. Heat release characteristic CAI/HCCI combustion
In an idealized HCCI/CAI combustion process, combustion reactions take place
simultaneously in the cylinder and all the mixture participates in the heat release process
at any instant of the combustion process. The cumulative heat release in such an engine is
therefore the sum of the heat released from each combustion reaction, dq i, of the
complete mixture in the cylinder, m, i.e.

K
Q=∫ m. dqi
1

Where K is the total number of heat release reactions, and (q i) is the heat released
from the (ith) heat release reaction involving per unit mass of fuel and air mixture.
Whereas the entire heating value of each minute parcel of mixture must be released
during the finite duration spend in the reaction zone in a SI engine, heat release takes
place uniformly across the entire charge in an idealized HCCI/CAI combustion.
However, in practice, due to inhomogeneities in the mixture composition and temperature
distributions in a real engine, the heat release process will not be uniform throughout the
mixture. Faster heat release can take place in the less diluted mixture and/or high
temperature region, resulting in a non-uniform heat release pattern as indicated by the
dashed lines. In comparison, combustion in a diesel engine is more complicated.
Figure 3.4. Heat release characteristic CI combustion
In a typical direct injection diesel engine, soon after the start of fuel injection a small
amount of mixture is involved in the premixed charge compression ignition combustion
process similar to HCCI/CAI, but most of the heat is released during the mixing control
diffusion combustion process. The cumulative heat released may be expressed as a sum
of the two processes:
K N
Q=∫ mp . dqi +¿ ∫ m j dq j ¿
1 1

Where the first part of the expression represents the premixed burning phase and the
second is the diffusion burning, during which the heating value of each mixture varies
according to the local mixture strength. In the above equation, m P is the amount of
premixed mixture taking part in the premixed burning phase, m j and (dq j ) are the mass
and heating value of each parcel being burned during diffusion burning.
 HCCI engine challenges
Even though HCCI engine has benefits over conventional diesel combustion
strategy, there are several challenges which must be resolved before commercialization of
HCCI combustion engine for automobiles. The main challenges of HCCI combustion are
stated as follows [29]:
- Control of ignition timing and combustion rate over wide engine operating range.
- Higher level of CO and HC emissions particularly at lower engine loads.
- Higher level of combustion noise particularly at higher engine loads.
- Lower engine load operating range.
- Cold-start problems.
Figure 3.5. The HCCI engine advantages, major challenges and their proposed solutions.
3.2. HCCI combustion control strategies
One of the main challenges in lower temperature combustion (LTC) engines,
particularly in HCCI combustion engines, is controlling the combustion rate. To achieve
higher thermal efficiency, the desired phasing of combustion timings is essential even at
moderate combustion. The combustion phase in HCCI engines is controlled either by
alter time-temperature history or by altering the mixture reactivity [14].

Figure 3.6. Methods for controlling HCCI combustion phasing


Figure 3.6 illustrates the methods of controlling HCCI combustion. The first group
includes fuel injection timing, intake thermal temperature, variable valve timing (VVT)
and variation of compression ratio (VCR). The second group attempts to control the
reactivity of the charge by varying the properties of the fuel, the fuel-air ratio or the
amount of oxygen by EGR.
3.2.1. Intake thermal management
In order to achieve autoignition at the optimal combustion phasing, appropriate
thermodynamic and chemical in-cylinder conditions are needed in an HCCI combustion
close to TDC position. Moreover, in HCCI combustion, the air-fuel mixture must be
sufficiently dilute to keep combustion rates and maximum charge temperature low
enough to achieve acceptable ringing intensity and NOx emissions. Fuel typically used in
HCCI combustion is gasoline-like fuels, which have higher autoignition temperature. The
basic theory is to increase the pressure of the charge at the intake valve closing (IVC),
where compression of the charge starts. This temperature can be increased by increasing
intake temperature or trapping the hot residual gases in the cylinder by negative overlap
valve (NOV). Using NOV could heat the inlet mixtures and improve the spontaneous
ignition characteristic of the engine. The residual exhaust gas could slow down the heat
release rate, decrease the pressure rise rate and maximum combustion temperature and
reduce the NOx emission. Several techniques are used to increase the initial temperature
of the charge at compression start position such as electrical preheating, exhaust heat
exchanger and glow plugs.
Fast thermal management (FTM) can be used to control the temperature of the
mixture at the beginning of compression stroke. The inlet temperature strongly affects the
combustion phasing. Figure 3.7 shows the schematic of FTM system used for control of
HCCI combustion [15].

Figure 3.7. Schematic of the FTM system.


The left throttle controls the amount of hot inlet air and the right throttle controls
the amount of cold ambient air to the engine. A source of cold air and a source of hot air
(heated electrically or by exhaust heat recovery) are used to achieve fast thermal
management control of the inlet temperature. By controlling the valves of the hot and
cold airflows, cycle-to-cycle control of the inlet temperature can be achieved. The
advantages of this approach are that it does not require major engine modifications or the
use fuel additivities. The only disadvantages of FTM are that the engine system has to be
fitted with a heater (electrical or exhaust heat exchanger). The mixing of hot and cold air
can be inside or outside cylinder by independently controlling the valve.
The intake/exhaust system for the engine system uses the waste thermal energy in
the coolant and gases for intake air heating. Figure 3.8 represent the system used for
extracting heat from engine exhaust and coolant to preheat the intake air.
2 1
1

Figure 3.8. Schematic of air preheating system by using heat exchanger taking heat input from
both exhaust and coolant [16].
It includes an internal combustion engine. Engine coolant circulates through the
engine block of the engine. The coolant flows through the outlet at a temperature that
typically would be about 90°C. It is received by heat exchanger 1, which is a liquid-to-air
heat exchanger that forms a part of the engine radiator. A portion of the intake airflow 1
is distributed through to heat exchanger 1. The outlet side of the heat exchanger 1
distributes heated air, typically at a temperature of about 70°C through heat exchanger 2.
This increases the temperature of the air before it is delivered to TVC. The temperature
may be about 200°C.
A portion of the intake air is delivered from the intake air 2 through TCV. The
heated air before TCV is combined at the mixing point at TCV with the cooler air in the
intake air 2. The combined flow is distributed to the engine intake port. The presence of
the catalytic converter between the exhaust port and the heat exchanger 2 will use the
boosted temperature level by intake air heating, thereby making the removal of unburned
hydrocarbons and carbon monoxide in the exhaust more efficient.
The diference between exhaust gas temperature and intake air temperature can
become insufficience for intake preheating at lower engine load. Exhaust temperature is
aslo lower, and typically, higher intake temperature requirement is aslo higher at low
engine loads. To avoid detonation or knocking, the intake air temperature should be as
low as possible so that the effective compression ratio and the thermal efficiency can be
as high as possible.
3.2.2. Exhaust gas recirculation
The technology of EGR is commonly used in HCCI combustion due to ít high
potential of controlling the auto-ignition of time-temperature history and enhancement of
NOx emission reduction. The EGR is broadly divided into two categories (internal and
exhaust EGR) based on the induction method in the cylinder. Internal EGR (iEGR), traps
hot residual gases in the cylinder by changing the valve timing. In the case of external
EGR (eEGR), recirculated a portion of exhaust gases from the engine’s tailpipe through
the intake manifold after going through an EGR cooler.
Internal EGR is acquired by the exhaust gas trap (EGT) using the variable valve
timing and negative valve overlap (NVO). In the NVO method, the exhaust valve is
closed early and a fraction of the exhaust gas trapped in the cylinder undergo
“recompression” due motion of the piston towards TDC. After that, the trapped residual
gas mix with the cooler fresh incoming charge. When the fresh charge is mixed with the
hot residual gases, the charge temperature rises overall, making auto-ignition easier in
HCCI combustion with high octane fuel.
In contrast to NVO, the positive valve overlap (PVO) is commonly used where
both the exhaust and intake valves are opened simultaneously for a short period of time
near the intake TDC position. Fig 3.8 shows the PVO and NVO valve lift profiles
together with corresponding typical cylinder pressure.

Figure 3.9. Illustration of PVO and NVO valve lift profiles along with typical cylinder pressure
curve [17].
The PVO represent in figure has negligible overlap between intake and exhaust events,
but it can vary depending on engine speed. This figure also illustrate that advance in
exhaust valve closing (EVC) timing is complemented by a corresponding and equal
retard in intake valve open (IVO) timing in case of NVO, and this approach is called
“symmetric” NVO. This technique is used reducing pumping losses caused by
recompression.
3.2.3. Variable compression ratio
Variable compression ratio (VCR) is another promising technology in the HCCI
engines that allows for the control of auto-ignition timing and expand the low-and high-
load limitation due to the high effectiveness in changing the temperature and pressure at
the end of the compression stroke in the combustion chamber prior the combustion.
One strategy is to tilt the upper part of the engine block, the mono head. A
hydraulic motor, controlled by an electronic valve turns the eccentric shaft that changes
the engine configuration to vary the compression ratio [18]. A five-cylinder 1.6L Saab
Variable Compression (SVC) prototype engine was used. It is based on a downsized
highly boosted SI engine concept. The compression ratio in the SI version can be varied
between 8 and 14 by tilting the upper part of the engine up to four degrees. A higher CR
than in the original engine is required to achieve HCCI combustion without excessive
inlet air heating or excessive amounts of hot residual.

Figure 3.10. Saab variable compression (SVC) engine.


The engine lowered its compression ratio from 14.0:1 to 8.0:1 by pivoting the top
of the block, including the cylinders and the head, around a beefy hinge on the intake
side. A set of mini connecting rods riding on an eccentric shaft on the exhaust side did
the lifting. Since the crankshaft position is fixed, tilting the block’s top changed the
combustion-chamber volume and thus the compression ratio. Saab claimed 225
horsepower and a fuel-economy bump of 30 percent over an engine with similar output.
Figure 3.11. The compression ratio is altered by tilting the mono-head in relation to the
crankcase by means of hydraulic actuator [19].
Its secret was a two-part engine block that was hinged on one side, with a
hydraulic actuator that could pivot the upper portion of the block by a small number of
degrees, moving the combustion chambers closer to or farther from the tops of the pistons
at top-dead-center. When the driver wants fast acceleration, the engine responds by
changing the compression ratio to avoid knock at the same time as it boosts the engine
using the compressor.
Haraldsson et al. [18] used a multi-cylinder equipped with the VCR to investigate
the influence of the inlet air temperature and compression ratio for controlling the
compression temperature and subsequently the combustion phase. The most common
way is to control the inlet air temperature by preheating for the low reactively fuel, and
thereby increasing the in-cylinder temperature at the end of the compression stroke. They
indicated that the higher compression ratio will replace the inlet air preheating and easily
control by cycle-by-cycle. The result of using a variable compression ratio to control the
ignition timing and combustion phase of HCCI engine are summarized in figure 3.11.
Figure 3.12. Overview of the variable compression ratio used in HCCI engine to control the
ignition timing and combustion phasing [18].
The major disadvantages of the VCR system are that it does not currently allow
for individual cylinder control, which is needed for good combustion timing control. The
VCR systems are also costly and complicated [8].
Japanese automaker Nissan has developed its own solution to the riddle of how to
vary the compression ratio of an internal combustion engine called the “Infiniti VC-T” –
or “Variable-Compression Turbo” – engine. The Nissan design is a turbocharged 2.0-liter
inline four producing 268 hp. It’s seen as replacing the company’s 3.5-liter V-6, with a
55-lb. reduction in weight and 27-percent reduction in fuel consumption.
The VC-Turbo replaces the traditional trade-offs imposed on conventional engines
caused by having a fixed compression ratio. Fuel consumption and exhaust emissions
performance are all enhanced to an unprecedentedly high level by changing the
compression ratio according to engine operation conditions [20].
Figure 3.13. The VC-Turbo engine offers any compression ratio between 8:1 and 14:1.
3.2.4. Variable valve actuation
Variable valve actuation (VVA)is a very fast method of varying the breathing
process of the engine. For HCCI control approach, there are two major methodologies;
internal EGR possibilities and effective compression ratio reduction.
 Internal EGR
Internal gas recirculation (called IGR) is an effective way to reduce HC and CO
emissions at low load in HCCI combustion. IGR is composed of residual hot exhaust
gases which remain in the cylinder after the exhaust phase. IGR allows improving the
homogenization (with a better evaporation) after the injection and increasing the global
temperature during the combustion. This has a positive effect on HC and CO emissions.
Figure 3.14 illustrates different VVA configurations in respect of effectiveness in
reducing HC emissions versus increase in indicated specific fuel consumption (ISFC)
along with cost/complexity of the system with reference to fixed valve actuation
(commercial configuration) [21].

Figure 3.14. Trade off effectiveness/cost at very low load for each solution.
VVTe (Variable valve timing exhaust): The exhaust re-compression is
accomplished with a modification of the exhaust valve timing an enhanced exhaust valve
lift time keeps valve opening time and lift constant, as seen in figure 3.15 [20].

Figure 3.15. Internal EGR with exhaust valve timing modification.


VVTei (Variable valve timing exhaust and intake): To compensate the LP IMEP
drawback and recover part of the energy used to re-compress the exhaust gases, the
exhaust valve lift timing change could be combined with a symmetric intake valve lift
timing modification, as show in Figure 3.16.

Figure 3.16. Internal EGR with exhaust and intake valve timing symmetric modification.
VVLei (Variable valve lift exhaust and intake): An alternate approach is fixing the
IVC (Intake Valve Closing) and the EVO (Exhaust Valve Opening) as seen in Figure
3.17 in order to prevent the negative effect on off expansion stroke and effective
compression reduction.
Fixing IVC prevents decreasing the effective compression ratio which a
drawback for HC and CO emission.
Fixing EVO prevents opening too soon the exhaust valve and reducing the
expansion phase of the gases.

Figure 3.17. Internal EGR with EVO (Exhaust valve opening) and IVC (Intake valve closing)
fixed.
To sum up, Walter B et al [20] investigated VVTe (Variable valve timing exhaust)
configuration reduces HC emissions by 50% and increase the ISFC (>15%) because of the early
opening of exhaust valve leading to reduction in expansion work. VVTei (Variable valve lift
exhaust and intake) is considerably better for ISFC increase due to the opening of the intake
valves after expansion of the residual gases, but relatively lower HC emission in this strategy.
 Effective compression ratio using VVA
A Variable Valve Actuation (VVA) system has the potential of changing the
compression ratio as well as controlling the amount of trapped residual gas. Further
possibilities are for example rebreathing of exhaust gases during the inlet stroke
through an extra exhaust valve opening. A disadvantage with controlling the effective
compression ratio using VVA is that it also reduces the effective displacement
volume, see Example 3.1.
Example 3.1 (Early/late intake valve closing). An automotive engine cylinder has
a displacement volume of 0.5 (l) and a compression ratio of 11. Find the effective
compression ratio and the effective displacement volume if the intake valve is closed
when the combustion chamber volume is 90% of the maximum volume.
Solution:
V d =0.5 ( l )

r c =11

The clearance volumes:


V d+ V c Vd
rc= =1+
Vc Vc
Vd 0.5
→ V c= = =0.05(l)
r c −1 11−1

The maximum volumes:


Vmax= V c +V d =¿0.05 + 0.5= 0.55 (l)
The volume when intake valve closed when the combustion chamber volume is 90% of
the maximum volume
V IVC =90 % .V max =0.9 . 0.55=0.495(l)

The effective compression ratio is then given as the ratio between V IVC and V C
V IVC 0.495
r c , eff = = =9.9
V C 0.055

Which is 90% of the geometric compression ratio.


The effective displacement volume is given as the between V IVC ∧V C
V d ,eff =V IVC −V c =0.495−0.05=0.445 ( l )

Which is 89% of the geometric displacement volume.


3.3. Homogeneous charge preparation strategies
The preparation of the homogeneous mixture is the most important in reducing the
local fuel rich region to minimize oxides of nitrogen and the particular matter (soot)
emissions. The effective preparation of the mixture for HCCI combustion includes both
fuel-air homogeneous and temperature control over combustion. The mixture preparation
strategies are either in-cylinder direct injection, or external mixture which in shown
figure 3.18.

Figure 3.18. Strategies for mixture preparation [22]


 External mixture formation
The homogeneous mixture which is prepared external to the engine cylinder is the
most effective because more mixing time availability, before the start of combustion.
The approach of preparation is more suitable for high volatile fuel such as gasoline
and alcohol. Port fuel injection (PFI), fumigation, wide open throttle (WOT)
carburation and manifold induction etc.
The port fuel injection (PFI) is the simplest method of external mixture
preparation, in which injector is mounted in the intake manifold, very close to the intake
valve. This method takes advantage of the turbulence generated as the intake flow rushes
past the intake valves to promote mixing. However, port fuel inject of heavy fuels can
result in high HC and CO emission, increased fuel consumption and oil dilution. The
disadvantage of this strategy is injection timing cannot influence the start of ignition.
This injection is mainly attractive for gaseous and liquid fuels with high volatility, but not
for diesel fuel.
 Internal mixture preparation
Internal mixture preparation can be achieved by early and late direct injection of
fuel inside the combustion chamber. The injection timing for early direct injection was
set during compression stroke, for late direct injection was set after TDC.
Early direct injection
The fuel injection process in HCCI combustion is charge homogeneity, which is
affected by injection timing. Early injection strategy is mostly utilized strategy of
achieving HCCI diesel combustion. It permits a longer ignition delay along with the low
temperature to homogenize the diesel-air mixture. A part or even the total amount of the
fuel is injected noticeably before top dead center, see Figure 3.19 [23].

Figure 3.19. Pulsed injection strategy for early in-cylinder injection.


In the case of diesel injection, the low gas density of the air inside the cylinder and
the poor volatility of the fuel can result in considerable wall wetting. A suitable injection
system must have a high degree of adaptability to adjust the injection strategy to the
varying boundary conditions during the infusion. High-pressure injection in combination
with a huge number of little nozzle holes is for the most part utilized to increase spray
disintegration and include the complete cylinder charge in the mixture formation process
while avoiding wall film formation. Another way to improve spray penetration is to split
the injection event into several pulses having different durations.
The shorter the pulse duration, the less the momentum of the liquid, resulting in
reduced penetration. The area below the curves represents the fuel mass belonging to
each pulse. The low gas density at the beginning of injection requires short pulses with
reduced injection velocities, and the time interval between the pulses is relatively large.
When the piston moves up, the density and temperature in the cylinder increase
and the penetration reduced. The duration of the pulse can be extended, while the interval
between subsequent pulses is reduced. When the end of pulse injection, the distance
between the nozzle and piston reduces significantly, and the mass injected per pulse must
be reduced again in order to prevent fuel deposition on the piston.
To prevent fuel deposition on the cold cylinder linear, reduce the angle between
the sprays. Walter and Gatellier [24] proposed narrow angle direct injection (NADI) to
keep the fuel target within the piston bowl and prevent the interaction of the spray with
the liner at advanced injection timing. Myung et al [25] investigated a narrow fuel spray
angle and dual injection. To avoid early combustion of the combination, the fuel injection
angle was reduced from1560 of a traditional diesel engine to 600 , and the compression
ratio was reduced from 17.8:1 to 15:1. Figure 3.20 shows the spray shape of the injection
spray according to the injection angle and piston position corresponding to the crank
angle.

Figure 3.20. Schematic diagrams of the (a) Conventional diesel engine. (b) NADI for an early
injection.
Figure 3.20 a shown the configurations of the conventional diesel engine, the
spray missing the piston bowl when the fuel was injected at early timing. However, the
employment of a narrow spray cone angle injector shows that the spray injected at the
early timing was not missing the combustion bowl as shown figure 3.20 b. The results
showed that the NOx emissions were greatly reduced as the injection timing was
advanced beyond 300 BTDC and the IMEP indicated a modest decrease although the
injection timing advanced to 500 −600 BTDC in the case of the narrow spray angle
configuration. This reveals that the narrow-angle concept was effective in maintaining the
high ISFC and IMEP when the fuel was injected at an early timing for HCCI combustion.

Figure 3.21. Effect on injection angle on the NOx emissions.

Figure 3.22. Effect of injection timing in the ISFC and IMEP.


Late direct injection
Another direct injection strategy is late in-cylinder injection, the development of
diesel-fuel late DI HCCI system is the modulated kinetics (MK) combustion system
developed by Nissan Motor Co, Ltd [26]. This system includes two mutually independent
intake ports, one of which is a helical port for generating an ultra-high swirl ratio and the
other is a tangential port for generating a low swirl ratio. The tangential port incorporates
a swirl control valve that controls the swirl ratio (3.5–10) by varying the flow rate.
To achieved the premixed combustion, the fuel–air mixture homogeneity before
ignition is required in MK combustion that can be achieved by increasing the ignition
delay longer and rapid mixing with a high swirl. The main specifications of the single-
cylinder test engine used in the experiments are given in Table 3.2 [27].

\
Table 3.2 The specification of the single-cylinder test engine
In the MK system, there are three characteristic: (i) late fuel injection timing starts
from 7 BTDC to 3 ATDC, (ii) high level of EGR and (iii) a high swirl ratio.
The system consists of independent intake ports, one for generating a high swirl
ratio and the other for generating a low swirl ratio. A flow control valve is installed at the
inlet the low swirl port. Closing this valve results in a high swirl ratio and opening it
produces a low swirl ratio. This means that the maximum controllable swirl ratio is
dependent on the high-swirl port, and the minimum swirl ratio is the combined swirl
obtained when the two ports are used together, as shown in Figure 3.23 [28].
Figure 3.23 Configuration of variable intake port system.
The effect of each combustion factor in the MK concept on exhaust emissions and
thermal efficiency is shown individually in Figure 3.24. With a lower oxygen
concentration, NOx emissions were dramatically reduced by approximately 90%, and
retarded injection timing reduced the smoke level markedly. Moreover, it is seen that the
higher swirl ratio played a part in reducing smoke and unburned HC emissions [8].

Figure 3.24. Effects of Each Combustion Factor on Exhaust Emissions and Thermal Efficiency
To sum up, the formation of NOx emissions can be suppressed by high EGR rates
(reduces oxygen concentration from 21% to 15%) and low temperature combustion. The
ignition delay was increased by decreasing the compression ratio to 16:1. The advantage
in MK combustion concept combustion control, without any fuel wall impingements.
3.4. The HCCI engine’s structural features
3.4.1. Performance of homogeneous charge compression ignition (HCCI) engine
with common rail fuel injection
Although diesel engines have high thermal efficiencies than that for petrol
engines, they release higher amounts of NOx, smoke and particulate matter (PM). To
defeat these drawbacks, a relatively new combustion concept known as HCCI
(Homogeneous charge compression ignition) is being explored all over the world. HCCI
engines can operate with reduced NOx, smoke and PM emissions simultaneously [30-
32]. HCCI was developed as a consequence of current progress in electronic engine
control, variable valve timing and lift systems, so the implementation of the HCCI
combustion concept in the manufacture of engines is now considered to be practical
[33,34]. HCCI uses the advantages of spark ignition and compression ignition engines.
They have better thermal efficiency than that for Spark Ignition engines and come
up with the efficiency of a Compression Ignition engine [35-36]. Homogeneous Charge
Compression Ignition (HCCI) uses a lean homogeneous air-fuel mixture as a means to
decrease the combustion temperature. This mixture can be carried out by prolonging the
ignition delay period with extremely advanced injection timing. Most of the fuel is
injected, evaporated and mixed in the premixed mode of combustion, such that the
mixture at the moment of ignition is near homogeneous. Therefore, the premixed mode of
combustion is the predominant phase in HCCI [37] In recent times, the multiple injection
approach was useful in commercial common rail Direct Injection HCCI diesel engines to
accomplish an organized air-fuel mixture and evade wall wetting. The common rail
system provides higher injection pressure for better fuel spray atomization and
evaporation, improved air entrainment and mixing in the fuel jet, which is advantageous
for lowering soot emissions [38,39]. Several Direct Injection HCCI concepts have used
advanced injection circumstances to sufficiently prolong the ignition delay period so that
full fuel vaporization event preceding to combustion. The electronically controlled
common rail injection permits diesel engines to vary the injection timing broadly and also
gives the ability to achieve multi- injection strategies which result in improved charge
mixing and homogeneity [40]. During HCCI engine running, a homogeneous fuel and air
mixture is set throughout the compression process and the mixture auto ignition takes
place by compression only at different in-cylinder positions when reaching the chemical
activation energy. The combustion process is controlled entirely by chemical kinetics like
Compression ignition engines [41] instead of ignition by a spark followed by a flame
front propagation such as Spark Ignition engine [42]. The two-step heat release processes
with low-temperature kinetic (cool) reactions occurring firstly followed by high-
temperature kinetic reactions are noticed in HCCI engine combustion. Many researchers
have studied the effect of engine speed on the combustion behavior and performance
characteristics of a HCCI engine. Kong et al. [43] investigated the consequence variations
of ignition start moment on HCCI engine combustion at different speeds. They concluded
that the ignition start would be an efficient and practical required parameter to manage
timing of ignition for the dynamic engine regimes. Iida et al. [44] concluded that the start
of heat release is less sensitive to engine speed variations than that of compression ratio,
suction air, and coolant water temperatures. Aroonsrisopon et al. [45] experimentally
examined different reference fuels with an octane number of 70, which reveal significant
heat release at low-temperature, at different engine speeds from 600 to 2000 rpm with
constant suction air temperature. It was found that the low temperature heat release
period decreases as engine speed increases. Furthermore, decreasing low temperature
heat release duration causes shifting the high-temperature heat release period to later
timings, consequently, a long low-temperature (cool) heat release period resulted in
maintain high temperature heat release. Iida et al. [46] used n-butane as a fuel to
investigate the effect of inlet air temperature, compression ratio and engine speed on
HCCI operating ranges in a research engine. they found that operation of HCCI at a
relatively high engine speed was achieved of 2000 rpm, with a high compression ratio
and a high temperature inlet air of 16.55 and 400 K, respectively. Szybist and Bunting
[47] concluded that the amount of energy released during the low-temperature heat
release interval increased suddenly as engine speed reduced. Besides, a nearly constant
temperature rise which occurred during the low-temperature heat release period was
noticed for all engine operating conditions. Garcia et al. [48] and Genzale et al. [49]
reported that the angle of the start of combustion increases as the engine speed increases.
Agarwal [50] concluded that the maximum pressure rise rate and coefficient of indicated
mean effective pressure [IMEP] variation are the critical factors, that can be used to
characterize HCCI operating range.
Ebrahimi and Desmet [51] experimentally studied the effect of engine speed and
cyclic variations in an HCCI engine. Khandal et al [34] presented an exhaustive review
on the performance characteristics, exhaust emission and combustion behavior of
Common Rail Direct Injection and HCCI engines operated using with different fuels.
Hadia et al [52] numerically investigated the combined effects on efficiency of the
compression ratios and steam injection, characteristics of combustion and emissions of an
HCCI engine. The contaminant emission is restricted by the dilution of the reactant by
using steam injection. Combustion is performed by using engine modeling. Calam et al
[53] parametrically investigated the effects of compression ratio on HCCI combustion,
performance, and emissions. The engine operational BSFC maps were the results of
running the engine at 169 experimental data points. The experiments were carried out at
different equivalence ratios with predetermined intake air temperature (353 K) and by
using RON20 and RON40 fuels. El Shenawy et al [54] reported that one of the essential
aspects to reduce engine emissions is how to prepare a homogenous charge for HCCI
engines. One strategy is the external mixture formation, or port fuel injection (PFI). PFI
used only for high volatile fuel such as alcohol [55-57]. The second is the in-cylinder
mixture preparation, which can be classified according to injection timing early injection
in which fuel is injected in the compression stroke and late injection. in which fuel
injection is delayed behind the top dead center [58].
One of the main challenges that the HCCI combustion technique has to face is the
occurrence of knocking at high loads due to high combustion pressure which causes the
engine to knock. This may limit this technique commercially practical [59]. However,
there are very few studies investigating the effects of the external load on HCCI engines
parametrically at different engine speeds. The objective of this work experimentally
investigates the performance of an HCCI Diesel engine with a Common Rail Direct
Injection (DI) fuel system. The engine performance is evaluated and presented at
different loads and speeds. The Apparent Heat Release Rate during combustion is
evaluated from the measured cylinder pressure-crank angle data. The pressure in the fuel
line is measured at the entry to different fuel injectors to investigate the effect of pressure
wave propagation on the fuel line pressure at different operating conditions.
 The Test Facility (Experimental Setup)
Experimental investigations were implemented on a transport diesel engine of type
Mercedes-Benz with an open chamber. A four in-line cylinders, water-cooled direct
injection diesel engine with overhead valve gear and “common rail” fuel injection
system. (available in the laboratory of mechanical power and energy at the Military
technical college) The main engine design parameters are listed in table 1. The test rig
includes the engine and all the instrumentation necessary for measuring and recording the
operating parameters. An on-line data acquisition system is furnished to improve the
speed and accuracy of data collection and recording. is used. Detailed engine data are
given in table below

Table 3.4 Engine data


 Instrumentations
The parameters measured are classified into three main groups. a -External parameters:
Engine speed, engine load, fuel consumption, airflow, crank angle, ambient conditions
and real-time. B-Internal parameters: pressure inside the cylinder, inlet and exhaust
manifolds and fuel line,and turbocharger speed. C-Safety parameters: Lubricating oil
pressure and temperature, and cooling water, temperature. Figure 3.25 gives a general
scheme of installed test rig showing numbered locations where important

Figure 3.25. Scheme of the engine test facility.

An incremental digital quadrature encoder (type WDG 58B-360-ABN-G24-K3) is used


for engine speed measurement. The encoder gives 2 trains of pulses (A and B), each has
360 pulses per encoder shaft revolution. The two trains are phased by ¼ pulse, and a third
index train (N) with one pulse per each revolution is also produced. The Encoder is
mounted on a special bracket fixed at the free end of the dynamometer shaft, figure 3.25.
A standard orifice (40 mm) diameter is mounted at the air surge tank entrance with a U-
tube water manometer for measuring the air pressure drop across the orifice. The airflow
rate is calculated from the orifice area and the manometer reading. This technique is
considered as one of the most accurate methods for measuring the fluid flow rate. The
unsteady pressure in the exhaust manifold is measured using a strain gauge transducer
(model PT124BG1/4) which comes complete with its amplifier). The transducer is
positioned such as direct exposure to the exhaust blows from cylinder outlet ports are
avoided. There is a list of measured parameters and the corresponding locations relevant
to figure 3.24.
Table 3.5. A list of measured parameters and the corresponding location relevant to figure 3.24
A strain gauge pressure transducer (model AST4000AV0015P6C0000, complete with an
amplifier, is used for measuring the pressure inside the intake manifold. Thermocouples
type K (Nicr-Ni), are used to measure the exhaust gasses and inlet air temperature
 Test procedure
Measured parameters included engine speed, engine power (calculated from the
dynamometer reading), fuel consumption, inducted air mass flow rate, fuel line pressure,
and cylinder pressure. The measurements were carried in the range from 1000 to 3000
rpm with 250 rpm increment. At each speed, the engine external load was adjusted to 0,
50, 100, 150, 200 and 250 N.m which corresponds to 0, 2.25, and 4.5, 6.75 and 9 bar
brake mean effective pressure (BMEP). The test procedure is required to fulfill the
following main tasks in the prescribed order: first, preoperational checks and
preparations. Then, running the engine and sustaining the operating conditions as
requested, (speed and load). Finally, Safety monitoring, e.g. overheating over-speed and
overload.
 Results and Discussions
In-cylinder and Fuel Line Pressures
In-cylinder pressure records investigation is the mainly efficient method to
examine the engine combustion performance due to the direct effects of in-cylinder
pressure history on the engine output power and combustion behavior characteristics. In
this paper, in-cylinder pressure is recorded versus crank angle. Cycle to cycle variation
was tested with the number of cycles. It was notice that the cycle to cycle variation is
constant for the number of cycles more than 30 cycles. Average pressure values for 42
successive cycles is calculated to evade the cycle to cycle variations and to report for
measurement uncertainty. In HCCI engines The combustion is divided into two
subsequent processes namely; low temperature process (cold flame) and high-
temperature process with relatively small ignition delay period which strongly depends
on the energy released during the low-temperature process [60]. Ebrahimi and Desmet
[51] introduced some fundamental combustion parameters of HCCI engine as shown in
figure 3.26. HRRHmax and HRRLmax are the maximum rates of heat release of high and
low temperature processes (reaction). In the first stage, the following parameters are
defined. LTRA and LTRE are the low-temperature reaction appearance and end which
are the points of intersection of the horizontal axis and the tangent line of the point where
the gradient is maximum on the rising side and falling sides of the curve respectively.
LTRD is the low-temperature reaction duration which is the duration between LTRA and
LTRE. HTRA and HTRE are the high-temperature reaction appearance and end which
are the points of intersection of the horizontal axis and the tangent line of the point where
the gradient is maximum on the rising side and falling sides of the curve respectively.
HTRD is the high-temperature reaction duration which is the duration between LTRA
and LTRE. but for the second-stage combustion. The zone of negative temperature
coefficient (ZNCT) is the period between FRBT and ARHT is named

Figure 3.26. Definition of symbols and combustion parameters [51]


The apparent (net) heat release rate AHRR generally used in the literature is calculated by
[32,33]

Pressure-crank
angle (P-φ) diagram provides information about combustion start, rate of pressure rise
(pressure roughness), and maximum cylinder pressure. The variation in in-cylinder
pressure and apparent Heat Release rate are represented versus engine speed in figures
3.27 and 3.28. The speed changes from 1000 to 3000 rpm at a constant load of 200 N.m.
It is noticed that there is more than one peak in P-φ history. This is a characteristic of
multiple injection strategies of HCCI engines. The figures show that the auto-ignition
(during HCCI combustion) takes place in all cases. It is seen that heat is released in
waves. These waves follow the injection pattern (pilot and main injections). The number
of burning waves, however, is not the same in all cases, but changes with engine speed. It
is directly obvious that as the engine speed increased the phasing of the main energy
release shifts to later in the engine cycle. Also as the engine speed increases the
magnitude of the low-temperature flame energy release becomes greater with fewer
burning waves. The crank angle delay between the low temperature and the main flame
energy release becomes longer for higher speeds. It is hypothesized that the HCCI auto
ignition process is largely kinetically controlled, which is to say it is time-based [45]. The
crank angle is directly proportional to the engine speed. Figure 3.28 shows that LTRD,
ZNTC, and HTRD based on a crank angle are observed to increase with increasing the
engine speed. As a result, the angle-scale interval of chemical reactions increases. The
engine speed has a remarkable influence on LTRD in spite of the other parameters.
Figures 3.29 and 3.30 dedicate the in-cylinder pressure variation recorded versus the
crank angle for an engine speed of 3000 rpm and different loads of 50, 100, 150 and 200
N.m. The high-pressure rise line is shifted lower crank angle as the load increases
(amount of injected fuel increases). As the mixture becomes richer (higher engine load),
the start of combustion shifts towards earlier times (crank angle) due to the presence of a
satisfactory amount of fuel at higher temperatures and pressure which leads to an earlier
auto-ignition of charge. The maximum pressure is shown to increase as engine load
increases.

Figure 3.27. Cylinder pressure variation versus to crank angle for a constant load of 200 N.m
and different engine speed.
Figure 3.28. Apparent Heat Release rate versus to crank angle for a constant load of 200 N.m
and different engine speed.
Initial pressure distribution in the rail superimposes onto the pressure distribution that is
generated by the next injection [63]. Figure 7shows the fuel line pressure variation versus
crank angle for an engine speed of 3000 rpm and different loads of 50, 150 and 200 N.m.
The average value of fuel line pressure increased from 1408 bar at 50 N.m brake load to
1584 bar at 150 N.m then increased to 1610 bar at 200 N.m which may be attributed to
higher wave amplitude encountered due to the increase of the injector opening duration
which promote the building up of pressure wave and retard the reverse of wave travel
direction which helps in damping of the wave generated at the moment of injector
opening.
Figure 3.29. In-cylinder pressure variation versus crank angle for an engine speed of 3000 rpm
and different loads.

Figure 3.30. Apparent Heat Release rate variation versus crank angle for an engine speed of
3000 rpm and different loads.

Figure 3.28 and figure 3.29 indicate that the crank angle interval at any in-cylinder
pressure becomes larger as engine speed and load increases so for the same injected fuel
the maximum pressure and so the maximum temperature may be decreased at high loads
due to the distribution of high-pressure values on larger crank angle duration. This effect
may be attributed to multiple injection strategies used in HCCI engines.
The measured fuel line pressure showed a repetitive pattern which is related to the
pressure wave initiated at the beginning of each injection, it depends on the lengths of
fuel pipes and a common rail in addition to the fuel bulks' modulus at each condition. The
disturbances in the pressure distribution, due to injector and pump flow, propagate back
and forth inside of the rail, reflecting off of the ends. Initial pressure distribution in the
rail superimposes onto the pressure distribution that is generated by the next injection
[63]. Figure 3.31 shows the fuel line pressure variation versus crank angle for an engine
speed of 3000 rpm and different loads of 50, 150 and 200 N.m. The average value of fuel
line pressure increased from 1408 bar at 50 N.m brake load to 1584 bar at 150 N.m then
increased to 1610 bar at 200 N.m which may be attributed to higher wave amplitude
encountered due to the increase of the injector opening duration which promote the
building up of pressure wave and retard the reverse of wave travel direction which helps
in damping of the wave generated at the moment of injector opening.

Figure 3.31. Fuel Line pressure variation versus crank angle for an engine speed of 2000 rpm
and different loads.

Figure 3.32. Fuel line pressure variation versus crank angle for a constant load of 200 N.m and
different engine speed.
Figure 3.32 shows the fuel line pressure variation versus crank angle for a constant load
of 200 N.m and different engine speeds of 1000, 2000 and 3000 rpm. The average value
is found to increase from 750bar at 1000 rpm to 1150 bar at 2000 rpm to 1610 bar at
3000 rpm. This is may be attributed to the higher flow rate of the high-pressure pump at
higher speed and the higher wave amplitude encountered due to the increased number of
injectors opening at a higher speed. However, the fuel line pressure needs more analysis
in the crank angle domain and time domain which is the matter of future work. The effect
of engine speed on pressure fluctuation is more dominant than that the effect of engine
load. Alzahabi[63] concluded that fluctuation in the fuel line pressure promotes the
fluctuation in fuel delivery. The pressure fluctuation is not enough to cause emissions
problems at engine part loads but is enough to cause problems at full load.
 Engine Performance
Figures (3.32a to 3.17a) show the speed characteristics of the turbocharged
common rail diesel engine at different speeds from 1000 to 3000 rpm. Performance
curves are presented at 6 fixed external loads, namely 0.0, 50, 100, 150, 200 and 250 N.m
corresponding to 0, 2.25,4.5, 6.75, 9 and 11.25 bar BMEP in preference of brake torque.
The results at each load are plotted against engine speed. For a sake of preference, the
speed characteristics of a turbocharged diesel engine with a conventional direct injection
system are presented in figures (3.32b-3.17b). The technical specifications of the
conventional engine are attached in appendix A. Performance curves are presented at 5
fixed external loads, namely 0.0, 70, 140, 210, 200 and 280 N.m corresponding to 0,
1.56, 3.11, 4.67 and 6.22 bar BMEP in preference of brake torque [35]. The engine power
shown in figure 3.31(a) is a direct function of engine speed and brake mean effective
pressure (BMEP). This is clear from the linear dependence on speed drawn at fixed
BMEP. Comparing figure 3.32(a) and figure 3.32(b) on can find that the maximum
presented power is 79.57 kW at 3000 rpm ( Specific power of 28.8 kW/liter swept
volume) for Common rail injection engine and 82.38 kW at 2800 rpm ( Specific power of
14.5 kW/liter swept volume) for the conventional injection diesel engine which means
that the two engines produce almost the same power at the same speed but at completely
different BMEP which is 11.25 bar for common rail and 6.22 bar for the conventional
engines so it is clear that the specific power of common rail engines is much greater than
that for conventional injection engines.

Figure 3.32a. Engine power for diesel engine equipped with a common rail.
Figure 3.32b. Engine power for diesel engine equipped with a conventional injection system.
The equivalence ratio for the two engines is presented in figures 3.32(a) and 3.32(b). The
mixture equivalence ratio increases with the brake mean effective pressure. This is due to
the greater amount of fuel injected when higher loads are required. The effect of higher
speed is attributed to better pump efficiency and fuel line expansion behavior. The figures
reveal comparable results at the maximum load (BMEP). At lower external loads
comparable results could be obtained at the same rations of BMEP/BMEPmax. The brake
specific fuel consumption (BSFC) presented in figure 10 is shown to have an increasing
trend with the decrease in engine speed due to the colder engine and bad charge mixing at
low speeds. Then it is shown to increase again with the speed increase. This may be
attributed to the higher friction losses at higher engine speed especially at part loads. The
specific fuel consumption of the common rail injection engine (minimum value 210
gm/kW. hr at 2000 rpm and 11.25 bar BMEP) is much lower than that for the
conventional engines (minimum value 292 gm/kW. hr at 2000 rpm and 6.22 bar BMEP).
This because the displacement volume of the conventional engine is higher than that of
the common rail injection engine and both working with the same fuel and equivalence
ratios and the better utilization of fuel attained by high-pressure injection. The thermal
efficiency is inversely proportional to the brake specific fuel consumption which is clear
by comparisons of figures. 3.33 and 3.34.
Figure 3.32a. Equivalence ratio for diesel engine equipped with a common rail.

Figure 3.32b. Equivalence ratio for diesel engine equipped with a conventional injection
Figure 3.33a. BSFC for diesel engine equipped with a common rail injection system.

Figure 3.33b. BSFC for diesel engine equipped with a common conventional injection system
Figure 3.34a. Thermal efficiency for diesel engine equipped with a common rail.

Figure 3.34b. Thermal efficiency for diesel engine equipped with a conventional injection
system.
Exhaust gas temperature shown in figure 3.35 is shown to increase with engine
speed and load. The effect is explained by the less amount of heat transferred to cylinder
walls as engine speed increases, and hence more heat is carried away with the exhaust.
The effect of the increased load is a direct result of the more fuel injected and
consequently more heat evolved during the combustion processes. The exhaust
temperature of the common rail injection engine is much lower than that for the
conventional engines. This may be attributed to the lean mixture that HCCI uses to lower
the maximum combustion temperature resulting in reducing the engine exhaust
temperature. The volumetric efficiency calculated based on the air density achieved with
turbocharging is shown in Figure 3.36. The volumetric efficiency of the common rail
injection engine and conventional injection engine are almost similar except the engine
speed at maximum volumetric efficiency which corresponds to the speed at which
maximum torque occurs for both engines.

Figure 3.35a. Exhaust temperature for diesel engine equipped with a common rail.

Figure 3.35b. Exhaust temperature for diesel engine equipped with a conventional injection
system.
Figure 3.36a. Volumetric efficiency for diesel engine equipped with a common rail injection
system.

Figure 3.36b. Volumetric efficiency for diesel engine equipped with a conventional injection
system.
 Conclusions
In the absence of studying the performance of HCCI engines at different loads, the
objective of this work is to experimentally investigates the performance of an HCCI
Diesel engines Common Rail Direct Injection (DI) fuel system. The engine performance
is evaluated and presented at different loads and speeds. The analysis of in-cylinder
pressure, heat release rate and fuel line pressure were performed to investigate the
combustion behavior characteristics. The major conclusions from these investigations are
listed below:
- The power density of the HCCI engine is much higher than of the conventional
engine. The HCCI engine has higher thermal efficiency and lowers exhaust
temperature than that for the conventional engine at all speed and loads
- The maximum in-cylinder pressure and temperature may be decreased at high
loads due to the distribution of high-pressure values on larger crank angle
duration. This effect may be attributed to multiple injection strategies used in
HCCI engines. This may be advantageous in decreasing Nox or may give the
chance to increase the engine compression ratio.
- LTRD, ZNTC, and HTRD based on the crank angle-scale increase with the
increase of engine speed. So the chemical reaction intervals increase.
- As engine load increases (the mixture becomes richer), the start of combustion
shifts towards BTDC side. Also as the engine speed increases the magnitude of the
low-temperature flame energy release becomes greater with fewer burning waves.
The crank angle delay between the low temperature and the main flame energy
release becomes longer for higher speeds.
- The pressure wave amplitude generated in the high-pressure line due to the
opening of the injector is shown to increase as engine speed and load increase. The
effect of engine speed on pressure fluctuation is more dominant than that the effect
of engine load. The fuel line pressure needs more analysis in crank angle domain
and time domain which is the matter of future work.
3.4.2. HCCI Fuel requirement
 Diesel fuel HCCI engine
Early HCCI engine developments were based primarily on diesel engines using diesel
fuel. This was, in fact, a logical approach at the time, due to the significant difficulty of
catalytically reducing NOx in lean combustion systems and also due to the existing mode
of compression ignition used in both diesel and HCCI mode. This work continues, but
more recently has focused on limiting the required compromises in both the engine
operating range and the potential HCCI emissions and fuel consumption opportunities.
Some of the earliest diesel fuel based HCCI demonstrated the limits of using diesel fuel
in the ‘pure HCCI’ mode. Diesel fuels typically have 90% distillation points in the range
of 340°C. The volatility, or the distillation characteristics of the fuel, dominates the
evaporation of the fuel in the combustion chamber. For true HCCI engine operation, all
of the fuel must be evaporated, and at least partially mixed, prior to the start of reaction.
It has been shown that if liquid fuel survives through the start of reaction, these fuel
droplets, or packets, will burn as diffusion flames with dramatically elevated production
of soot and NOx. It can be observed that an increase in soot formation, as indicated by an
increase in BSN (Bosch Smoke Number) from zero, signals the onset of non-HCCI
operation. It appears that this onset occurs in the range of intake air temperatures from
130 to 150°C, for diesel fuel and blends of gasoline and diesel fuel. These trends are
demonstrated in Fig. 3.5, where the BSN is plotted versus intake air temperature. In
general, the presence of gasoline tends to reduce the temperature, but the limit is still
above 130°C. Droplet evaporation calculations were performed to determine the fate of
the fuel during compression. These calculations indicated that the intake temperature had
to be greater than 120°C for all of the diesel fuel blends. This temperature is of concern
because of the potential impact on the volumetric efficiency. In addition, it was observed
that this intake temperature.

Figure 3.37. BSN versus intake air temperature for various gasoline and diesel fuel blends.
CR = 14:1, A/F = 40:1, no EGR.

leads to early start of reaction (SOR) for diesel fuel and all of the blends that contain
diesel fuel. Acceptable HCCI operation (close to ‘pure HCCI’) on diesel fuel was only
possible at intake manifold temperatures of 130°C or higher, and compression ratios of
8:1 or less. The SOR appears to be dominated by the auto ignition temperature of the fuel
and the temperature history of the fuel air mixture. In an effort to control the SOR, a
variety of different fuels have been considered. One interesting fuel tested at Southwest
Research Institute was naphtha from Fischer Tropsch processing of natural gas. The
results of these experiments indicated that this particular fuel addressed the volatility, or
mixture preparation problem, but it did not totally address the SOR issues. The results of
engine SOR analysis for gasoline-diesel fuel blends and for Fischer Tropsch naphtha are
shown in Fig 3.38. It should be noted that the results in Fig 3.38 were based on tests
designed to achieve the maximum load at each condition, so that air-fuel ratio did vary
from test to test. The EGR level was zero in all cases. The lowest intake manifold
temperatures that could be used with the gasoline and diesel fuel blends were in the range
of 110 to 130°C. Attempts to use lower temperatures resulted in the onset of sooty
diffusion burning. The exception was gasoline. Gasoline would not react at the lower
temperatures, even when the compression ratio was increased to 16:1. There were no
temperature limits on the Fischer Tropsch naphtha. In fact, it can be seen that the use of
the naphtha actually improved the performance in two ways. First, it did not revert to
sooty diffusion burning, even as the intake temperature was reduced to ambient levels
(the BSN was zero for all of the Fischer Tropsch naphtha runs). Second, the SOR was
retarded due to the reduced intake temperature.
Figure 3.38 Intake manifold temperature versus start of reaction.

Figure 3.39. Cylinder pressure comparison (same conditions as Fig. 3.38)

The results for the naphtha fuel used in testing were encouraging with respect to
intake temperature and combustion characteristics. Naphtha has reaction timing similar to
that of DF-2 fuels at similar temperatures; however, it also provides the HCCI reaction at
ambient intake air temperatures. There is no comparison of the smoke levels of the two
fuels because, as indicated above, naphtha fueled HCCI produced zero smoke at all tested
conditions, indicating that total fuel evaporation was possible at low intake manifold
temperature when using the naphtha. Review of the combustion data reveals that the peak
cylinder pressure is greater for naphtha fueled HCCI (Fig 3.39), the peak value of the
main heat release rate is less (Fig. 3.40), and the cumulative heat release is greater (Fig
3.41). These results were obtained at approximately the same IMEP, but at
Figure 3.40. Heat release rate comparison (same conditions as Fig 3.37).

Figure 3.41. Cumulative heat release comparison (same conditions as Fig.3.5)


different MATs, with the diesel tested at approximately 130°C and the Fischer Tropsch
naphtha tested at approximate 28°C MAT. The reaction characteristics of HCCI are very
interesting. As indicated above, the reaction of hydrocarbon fuels in an HCCI engine
typically demonstrate a two-phase type of reaction. This is demonstrated in Fig. 3.40,
where the heat release rate diagrams for diesel fuel and Fischer Tropsch naphtha are
compared. It can be seen that the naphtha, in addition to having a larger pre-reaction heat
release, also had a significantly delayed SOR (start of reaction), based on the pre-reaction
heat release. It should also be noted, however, that the main heat release occurred at the
same timing as that of the diesel fuel. Crank angle degrees (CAD) 135 150 165 180 195
210 225 –0.02 0 0.02 0.04 0.06 0.08 0.1 0.12 Heat release rate (kJ/CAD) DF-2, A/F =
43.4, 6.72 kW(i), MAT = 109C, BSN = 0.6 Naphtha, A/F = 51.2, 7.25 kW(i), MAT =
31C, BSN = 0 14.5 Heat release rate comparison (same conditions as Fig. 3.38). Naphtha
DF-2 Crank angle degrees (CAD) 90 180 225 270 135 0.6 0.5 0.4 0.3 0.2 0.1 0 –0.1
Cumulative heat release (kJ) 14.6 Cumulative heat release comparison (same conditions
as Fig. 3.38) HCCI fuel. The most noticeable difference in the shape of the heat releases
shown in the Fig. 3.41 is that the initial heat release of the naphtha run is greater than that
of the diesel fuel run. The SOR occurs later in the engine cycle and a greater fraction of
the fuel is consumed during the initial heat release. This has the effect of moving the peak
cylinder pressure closer to TDC. The cumulative heat release rate is different than typical
diesel combustion, most likely due to completion of combustion prior to TDC. Prior to
testing, it was expected that the cetane number (CN) of the naphtha would be
significantly lower than that of the diesel fuel. In fact, it was found to be the same as the
diesel fuel, 46CN. This observation raised the question as to the application of CN as a
fuel rating parameter for SOR in HCCI engines. The results presented in Fig. 3.40
indicate very clearly that the CN is not a universal indicator of the fuels SOR
characteristics in an HCCI engine. In this case the CN variation is negligible but the
engine results indicate that the SOR does have a large variation. These results suggested
that some measure of the auto ignition temperature might provide a better indication of
the fuels propensity for reaction in an HCCI engine. The results presented in Fig 3.42
clearly demonstrate the validity of this assumption, where the temperature histories for a
number of HCCI runs on the same diesel fuel are compared. These temperature histories
were computed from the measured cylinder pressures, using the ideal gas assumptions.
The differences in the runs are intake temperatures, compression ratio, and load. In all
cases the temperature at SOR (based on the timing of the pre-reaction heat release) was
approximately the same, as indicated by the end points (approximately 400°C) in the
plots. The absolute value of 400°C may not be
Figure 3.42. Bulk gas temperature during compression and SOR temperature
accurate due to the assumptions used in the calculation of the compression temperature
histories. The largest errors are probably introduced through the assumed initial
temperature, where heat transfer during the intake process was not accounted for in the
calculation. Fundamentally, the start of reaction is related to the time-temperature history
of the fuel-air mixture, the ratio of fuel to air, and the activation energy of the fuel. The
thermodynamic history is relatively easy to address through calculation, however, the
activation energy is very difficult to determine as a fuel specification property and even
more difficult to apply in the changing thermodynamic environment of the engine. The
SwRI engine experiments provided some indication that the start of reaction can be
related to the auto ignition temperature of the fuel. Current standard test methods for
determining the auto ignition are generally related to atmospheric condition safety issues
and do not provide data representative of the conditions encountered in engines.
 HCCI fuel ignition quality
The ignition and reaction characteristics presented in the heat release rate diagrams
presented in Figure 3. 43, are typical of fuels with relatively large levels of low
temperature reaction. As indicated above, this is typical of diesel-like fuels, or fuels
which auto ignite easily at relatively low temperatures. The effects of fuel composition on
the low temperature reactions are discussed in another section. The approach taken in the
Southwest Research Institute work [5] was to determine the ignition delay time as a
function of the initial air temperature in a commercially available constant volume
combustion bomb apparatus currently used in diesel fuel cetane rating (ASTM D6890),
called the Ignition Quality Tester, or IQT. The goal was to develop relationships between
the ignition delay time and the test temperature, or conversely, ignition temperature (test
temperature) as a function of the delay time. Four different sets of fuels were examined in
this effort. The sets consisted of the following:
1. Hexadecane and heptamethylnonane blends (primary reference fuels for cetane number
rating).
2. Isooctane and n-Heptane blends (primary reference fuel for octane number rating).
3. Diesel fuel and gasoline blends
4. Fischer Tropsch naphtha (FTN) and ethyl hexyl nitrate (EHN) (additized fuels to test
alternative method for cetane number variation).
All of the fuels were tested in the IQT at approximately the same initial temperatures,
460, 480, 500, and 530°C. A fuel residence time, or ignition delay time, typically
encountered in an HCCI engine was selected, and using this ignition delay time and the
IQT data it was possible to determine the temperature required to give this ignition delay
time. This temperature represents the point at which each fuel starts to react in the same
time period in the HCCI engine, and it is called the elevated pressure auto ignition
temperature, or EPAIT. In addition to the ignition temperature versus the ignition delay
time relationships, the cetane numbers of each test fuel were also either defined by the
blend ratio (primary reference fuels) or determined in the IQT. It was possible to plot the
EPAIT versus the cetane number. These data are presented in Fig. 3.43, where the EPAIT
is plotted for all test fuels versus the cetane number. The results demonstrate what
appears to be some relationship between the EPAIT and the cetane number, however, the
relationship is not universal and there is some scatter about the line. The un-additized FT
fell on the curve, but the data for the additized FT fuels deviated from the curve. It is not
suggested that CN be used as a measure of ignition quality for HCCI engine fuels. The
correlation is presented only to demonstrate the problem associated with the use of CN
for this purpose. As will be demonstrated, octane number (ON) is even a worse indicator
of ignition quality in these engines. One of the test fuel sets consisted of blends of the
primary reference fuels for octane rating of gasolines. Since the cetane numbers of these
reference fuel blends were also determined in the IQT experiments, it was possible with
the IQT data to relate the cetane and the octane numbers.

Figure 3.43. Elevated pressure auto ignition temperature versus cetane number.
Figure 3.44. Relationship between octane and cetane numbers using isooctane and heptane fuels

Figure 3.45. Elevated pressure auto ignition temperature versus octane number
These data are plotted in Fig. 3.44. The regression equation, relating the ON to the
CN is also presented in Fig. 3.44. While the expression presented in Fig. 3.44 may not be
universally applicable it was used to estimate the ON of all of the other test fuels. (In fact,
extrapolation over the entire cetane number range (15 to 100) results in extremely
negative octane numbers.) It was then possible to plot the EPAIT versus the ON, as
presented in Fig. 3.45. As can be seen, there is a relationship between the EPAIT and
ON, but it requires significant extrapolation of the ON scale into the negative ON range,
a meaningless extrapolation. In addition, in the range of application of the ON scale the
slope is very steep. It should be noted that all of the test fuels have been successfully
tested in the HCCI engine, meaning that the fuel actually reacted and produced work in
the engine. Under these circumstances reactions were initiated, thus indicating that there
is some fuel property that represents their ignition characteristic in the engine. It is clear
that ON is not a good parameter for this purpose. The results of these experiments
indicate that the IQT can be used to determine the EPAIT of candidate HCCI fuels. It is
also apparent that this parameter is somewhat related to both the ON and CN. The results
also do indicate, however, that ON and CN are not necessarily good rating parameters for
all of the potential HCCI fuel candidates because of the limitations on the ranges of these
two methods that are imposed by the use of the specific reference fuels. In other words,
neither ON nor CN have adequate range to provide the needed scale for all possible
HCCI fuel candidates. The EPAIT, on the other hand, is a fundamentally based
measurement, not dependent on specific reference fuel behavior, and not limited in range.
It has also been observed that fuels can have the same EPAIT and significantly different
locations of the main, or high temperature reactions. This phenomenon is presented in
Fig. 3.46, where the cumulative heat release rate diagrams are presented for three fuels
that have similar EPAIT and ON. Note that the EPAIT does accurately predict the start of
the low temperature reactions (all fuels have the same start of reaction). Note, also that
the magnitudes of the cool flame reactions are significantly different and that the start of
the main reactions appear to be related to the magnitude of the low temperature reactions.
Figure 3.47 is a plot showing the relationship between the magnitude of the low
temperature reactions and the timing, or phasing between the start of the low temperature
reactions and the start of the high temperature reactions.

Figure 3.46. Cumulative heat release rate diagrams for three fuels with the same EPAIT
Figure 3.47. Phasing between the low temperature reactions and the high temperature
reactions for fuels that support low temperature reactions
It is clear in Fig. 3.47 that the onset of the main reaction is dependent on the
magnitude of the low temperature reactions, for fuels that demonstrate low temperature
reaction. Basically, the low temperature reaction increases the temperatures in cylinder,
leading to earlier onset of the high temperature, or main reactions. This is clearly
demonstrated in experiments with increasing load (richer air-fuel ratios) as shown in Fig.
3.48, where it can be seen that the magnitude of the low temperature reactions increasing
with richer mixtures (higher loads), due simply to the fact that there are larger quantities
of the fuel components that produce low temperature reactions in these richer mixtures.
 Gasoline HCCI
As pointed out above, mixture preparation in HCCI engines is dramatically easier with
gasoline-like fuels than with diesel-like fuels. It was also pointed above, however, that it
is very difficult to initiate reaction in gasoline HCCI engines due to the high auto ignition
temperature of gasoline. The EPAIT for a representative gasoline and a representative
diesel fuel, and other fuel blends are presented in Fig. 14.14. The plot is similar to Fig.
14.8, but with the addition of the data for the representative gasoline, diesel fuel, and the
Fischer Tropsch naphtha. As can be seen, the diesel fuel falls in the lowest
Figure 3.48. Heat release rate diagrams for two different equivalence ratio tests using a fuel that
demonstrates low temperature reactions
range, with EPAIT values in the order of 450°C, while gasoline falls in the highest
regions, with EPAIT values in the range of 650°C. Cycle simulation analysis has been
used to compute the compression temperature histories as a function of compression ratio
(11.5–17.5:1), intake manifold temperature (20–100°C), and engine speed (1200–2400
rpm). These predictions are presented in Figs 14.15–14.17, where the gasoline and diesel
fuel EPAITs are also plotted and annotated. As can be seen, the start of reaction is very
early for fuels with ignition characteristics like diesel fuel, or fuels with EPAIT values in
the range of 400–450°C (723–773 K). On the other hand, fuels with EPAIT values in the
range of gasoline, 600–650°C (923–973 K), may not react under any of the engine
operating conditions examined in the figures. These results indicate that neither gasoline
nor diesel fuel have the correct EPAIT for operation in practical conventional engines.
Researchers have worked diligently to increase the compression temperatures in gasoline-
fueled HCCI engines to overcome the high EPAIT with gasoline.
Figure 3.49. EPAIT versus cetane number for a variety of fuels.

Figure 3.50. Predicted compression temperature histories at three compression ratios.


The most popular approach is to use internal EGR, this is exhaust gas that is retained in-
cylinder usually through early exhaust valve closing. Promising work in this area,
generally referred to as negative valve overlap, incorporates direct injection, with some
fuel being injected during re-compression resulting from the lack of overlap between the
exhaust and intake processes. This is demonstrated in Fig. 3.53 and 3.54, where various
injection timings are annotated on a valve actuation schedule in Fig. 3.53 and the
resulting heat release rate diagrams are presented in Fig. 3.54. Note that DI HCCI is
Figure 3.51. Predicted compression temperature histories with three different intake air
temperatures.

Figure 3.52. Predicted compression temperature histories at three different engine speeds.

Figure 3.53. Cylinder pressure diagram showing valve timing and DI fuel injection strategies to
achieve acceptable ignition in gasoline HCCI.
Figure 3.54. Heat release rate diagram for different fuel injection strategies
possible with gasoline because gasoline will atomize and evaporate with a characteristic
time equivalent to or less than the ignition delay time of the fuel. This relates to the fact
that gasoline typically exhibits little or no low temperature reaction.
3.5. HCCI engine uisng variable valve timing
To achieve a successful control of the HCCI process, the temperature, composition
and pressure of the charge mixture at IVC point have to be controlled. Using of the
variable valve timing strategy that enables quick changes in the amount of trapped hot
exhaust shows the potential for the control of HCCI combustion.
The influence of the variable valve timing strategy on the gas exchange process,
the process between the first valve open event (EVO) and the last valve closing event
(IVC), in a HCCI engine fuelled with standard gasoline fuel (RON 95). Another strategy
has the potential to control the start of auto ignition and the heat release rate of the HCCI
combustion, together with their effectives and practical feasibility. Trapping hot residual
gases into cylinder, accomplished by fully variable valve timing (FVVT) system, the
most promising and practicable method for achieving HCCI combustion power in the
certain load range [63].
The effect of the variable valve timing strategy on the engine parameters (such as
the trapped RG rate, pumping losses, load, volumetric efficiency and trapped gas
temperature) and cylinder charge properties (such as composition, temperature and
pressure) were investigated.
 Set up engine
The engine employed in this research is a simple cylinder, 4 stroke-engine based
on the GM Family One, 1.8litre, see figure 3.55. A standard 4-cylinder head is mounted
on top of a water-cooled barrel, with a custom made a bottom end. A fully variable valve
timing strategy named Active Valve Train (AVT): produced by Lotus Engineering, see
figure 2. Variable quantities of Trapped Residual Gases (TRG) can be captured in this
way. For this investigation, the CR was set at 10.5:1 [64].
Figure 3.55. Single-cylinder research engine with AVT
For the production AVT, Lotus and Eaton (a tier-1 valvetrain supplier) have
formed a close collaboration. The hydraulic system determined to offer the functionality
targeted in Table 3.3

Table 3.3. production AVT performance targets


Figure 3.56. Valve block off AVT hydraulic actuator.
The method used to initiate and control the HCCI combustion is based on the
trapping of a pre-determined quantity of residual gas by closing the exhaust valves early
in the exhaust stroke and opening the inlet valves late in the intake stroke. Figure 3.57
shows the general principle.

Figure 3.57 The method of sequential valve events.


This has referred as recompression. When the piston down ward on the intake
stroke, the intake valves are opened late and fresh charge is drawn into the cylinder which
is already partially filled with exhaust gases. At the end of the intake stroke the intake
valves are closed and the mixture off fresh charge and residual gas siss then compress in
the next compression stroke. HCCI combustion process as the mixture temperature rises
in the final stage off the compression stroke.
Rebreathing is another method to achieve HCCI combustion. The exhaust gases
are discharged from the cylinder when the piston approaches BDC from the power
stroke. On the intake stroke, when the piston approaches TDC overlap, both inlet and
exhaust valves open simultaneously, allowing fresh charge and exhaust gas to be drawn
into the cylinder simultaneously. Again, HCCI combustion occurs ass the mixture
increases in the final stage off compression stroke. This method is named simultaneous
method, see figure 3.58.

Figure 3.58. Schematic showing operating principle of HCCI combustion.


Law Don et al [65] investigated SI combustion mode with HCCI combustion of
operation because of the greater pumping work associated with running conventional
throttled SI combustion. In CAI modes the engine was run unthrottled and so the load
was controlled entirely using AVT system, as shown in figure 3.59.
Figure 3.59. Cylinder pressure trace comparisons of conventional spark ignition with HCCI
method 1 and 2 at 2000 rpm, 4.2 bar IMEP.
Figure 3.59 illustrates the resultant pressure profiles for method 1 and method 2
compared with a pressure profile obtained for a conventional SI combustion event at a
constant speed, load and AFR. In CAI combustion method 1, there is some in taking
work associated with the compression and combustion of the trapped residual gas when
the exhaust valve is closed early in the exhaust stroke during CAI modes of operation,
however, this is relatively small.
The most important advantage of HCCI combustion has been the reduction in specific
NOx emissions. The reduction in NOx emissions from pure HCCI is greater than 90%
across the whole range of air: fuel ratio investigated [65]. As shown in figure 3.60, It is
clear that spark-assisted HCCI offers no further benefit to NOx reduction than pure HCCI

Figure 3.60 Effect on air: fuel ratio on specific NOx emission for conventional SI and HCCI and
Spark-assisted HCCI.

Figure 3.61. Experimental cylinder pressure curves for various amounts of TRG obtained by
increasing the negative valve overlap
The result obtained using the recompression method by early IVC paired with a
symmetrically late IVO) is shown in figure 3.61. Demonstrate that HCCI mode is
achieved at 32 percent IEGR (high load) and is maintained up to 59 percent (low load) at
the given engine conditions.
At trap residual gas (TRG) quantities less than 32 percent, the HCCI combustion
cannot generated and it is necessary to use the spark-plug. The area below 32 percent
TRG is not “pure” HCCI operating mode, but it is also not the “pure” SI mode, since the
spark is only used to ignite the charge mixture and the turbulent flame propagation in the
rest of unburned mixture is not observed. Instead of this, the unburned mixture is auto-
ignited and sustained combustion in uniform and simultaneous auto-ignition process, as
in the HCCI combustion.
3.6. HCCI using exhaust gas recirculation system
In this case of HCCI diesel applications, the EGR technique can be divided into
internal EGR and external EGR. The internal EGR rate can be obtained by changing
valve overlap period and the external EGR rate can be adjusted by a backpressure valve
of the exhaust and an EGR valve. Lei Shi et al () investigated diesel HCCI combustion by
combined internal exhaust gas recirculation (IEGR) and cooled external EGR.
First, the high temperature of the remained exhaust gas in cylinder promotes
vaporization of fuel which was injected into the cylinder before the top center (TC).
Second, the cooled external EGR was used reduce the mixture temperature during
the compression stroke, which was a main factor to control the ignition timing in HCCI
engine. This experiment studied the effects of the IEGR and the cooled external EGR on
diesel HCCI combustion performance and emission performance. The results showed that
the IEGR could reduce the smoke emission but advanced the start of combustion (SOC)
in HCCI mode and the cooled external EGR could delay the SOC effectively, which
could expand the HCCI operation range.
Figure 3.62 shows the experimental setup. It used a single-cylinder, four stroke,
four-valve, water cooled and naturally aspirated diesel engine.

Figure 3.62. Schematic of experimental engine.


Table 1 shows the engine specifications. NOx emission was measured by AVL
DIGAS4000, which used the chemical luminescence detector (CLD) method. Smoke
emission was measured by AVL 439 opacimeter. A water-cooled AVL 12QP pressure
transducer was used to measure in-cylinder pressure. It is saved for 100 consecutive
cycles with a resolution of 0.5 crank angle degree (CAD). This is done by use of a PC
connected to a 12 bit multiplexed A/D converter. Table 3.4 shows the measurement
accuracies for each experimental parameter.

Table 3.4 Specifications of test engine

Table 3.5 Measurement accuracy


HCCI mode is hard to finish in engine warm-up and cold start measure because of the
low temperature. In order to achieve engine cold-start and warm-up, an auxiliary injector
was installed in the cylinder head in addition to the original injector and used to ignite the
premixed fuel injected by the original injector near the TC (top center) of intake stroke.
When the engine warmed up, the engine was changed to HCCI mode, which used only
one pulsed fuel injection by the main injector before the TC (top center) of exhaust
stroke. Figure 3.63 shown the injection mode.

Figure 3.63. Fuel injection mode.


The variable valve timing technique was applied to change the quantity and
temperature of the EGR. The amount of cooled external was regulated by the back
pressure valve and EGR valve, and the temperature is controlled by the cooler in the
EGR pipe entrance. Early closing of the exhaust valve and late opening of the inlet valve
can change the amount of exhaust gas in the cylinder. When the exhaust gas in the
cylinder is compressed by the piston, its temperature has increased. This is beneficial for
the vaporization of the injected fuel to form a homogeneous gas mixture. The IGER rate
was determined using the valve overlap in this study. Three valve overlaps were
investigated and are depicted below figure 3.64.

Figure 3.64. Schematic of valve overlap.

The cooled EGR prevent the heating of EGR. The temperature of EGR was maintained
by adjusting the water flux in EGR cooler. The external EGR rate is determined by
measuring concentration of CO2 in inlet and exhaust pipes:
[CO ¿¿ 2]inlet
EGR= ¿
[CO¿¿ 2] exhaust ×100 % ¿

Figure 12. The definition of combustion parameters.


The low temperature reaction (LTR) duration was the crank angle interval
between the start of LTR and the point which presented the endings of quickly decreasing
ò LTR. The HTR (heat temperature reaction) duration was defined using the same point
of start of HTR, because HTR was the main combustion and the LTR was smaller
compared with the HTR, see in figure 3.65.
Effect of IEGR on HCCI combustion
Figure 12 illustrate the heat release rate curves of HCCI combustion of different
valve overlaps.

Figure 3.66. Effect on IEGR on HRR (heat release rate)


When the negative valve overlap (NVO) is increased, the begins of LTR and HTR
are both advanced, the combustion duration has been reduced and the maximal heat
release rate has risen. More high-temperature residual gas is trapped in the cylinder with
rising the NVO and results in a higher in-cylinder temperature. The combustion rate of
HCCI engine is very sensitive to the temperature and it increased as the temperature rises.

Figure 3.67. Effect of IEGR on smoke emission of HCCI


Figure 3.67 present the smoke emission of HCCI combustion of various NVO. As
the NVO increase, smoke emissions lowers because the mixture is more homogenous.
The amount of heat required for fuel evaporation increase as the amount of fuel increase.
Therefore, it can be seen that the smoke emission of 10 CAD VO increase with increase
load because there is not enough heat to vaporize the injected fuel. When the NVO is
raised, in-cylinder residual gas temperature rises, which can boost fuel evaporation and
aid the formation of homogenous mixture gas. As a result, the smoke emission is reduced
rapidly.
Comparison of NOx emission of various NVO is depict in figure 3.68. NOx
emission of large NVO is higher and more sensitive to load. When the load is increased
to 100 CAD VO, NOx emission changes minimally. When the load is increased at −200
and −400 CAD VO, however, NOx emissions worsen. When the IMEP reaches 0.38
MPa, the NOx emission of −400 CAD VO rises fast. This is mostly due to the advanced
SOC caused by the significant NVO, which raises the in-cylinder temperature, which
causes NOx emission to rapidly grow.

Figure 3.68. Effect of IEGR on NOx emission off HCCI


Effect of cooled external EGR rate on HCCI combustion
The large NVO can decrease smoke emission of HCCI combustion, but not benefit
to expand the high load limit and increase the NOx emission. Figure 15 illustratre the
influence of EGR rate on the starts of LTR and HTR under different fuel quantities,
0.054, 0.058, and 0.062 g/cyl. The starts of HTR are advanced with the increase of fuel
quantity, since the incremental fuel can provide more heat to increase the in-cylinder gas
temperature and enrich the in-cylinder mixture. The starts of LTR are very close each
other. This is because the fuel does not release heat before the start of LTR and the
conflict between the absorption of heat due to high engine body temperature and the
consumption of heat due to more fuel vaporization. With the increase of EGR rate, both
the starts of LTR and HTR are retarded.
Figure 3.69. The effect of EGR on interval of LTR and HTR.

The higher specific heat of the exhaust gas, which absorbs more of the energy,
reduces the in-cylinder temperature. At the same time, exhaust gas dilutes oxygen
concentration in the inlet gas and thus decreases the speed of reaction. So, increased EGR
rate may delay both the starts of LTR and HTR.

Figure 3.70 The effect of EGR on emission of HCCI.


The effect on EGR rate on exhaust emission each fuel quantity of conventional
diesel mode and HCCI mode, see in figure 3.70. At HCCI mode, when the fuel amount
per cycle is relatively small, NOx emission, which remained at a low level, was slightly
influenced by the EGR rate increase. At this moment, boosting the EGR rate can
effectively reduce NOx and avoid knock. The influence of the EGR rate on conventional
diesel engine shows the same tendency.
There is still a certain amount of smoke emission though it is relatively small
compared with the conventional diesel engine. The influence of EGR on HCCI
combustion is relatively small in contrast with the conventional diesel engine. With the
increase of EGR rate, the smoke of conventional diesel engine increases quickly while
smoke in HCCI engine changes slightly. So cooled external EGR can effectively manage
HCCI combustion and not worsen the smoke emission. Furthermore, a more
homogeneous mixture can reduce smoke emission. CO emissions rise steadily as the
EGR rate rises in both HCCI and conventional diesel engines. The reaction speed and in-
cylinder temperature drop as the EGR rate rises. At the same time, more exhaust gas
reduces the O2 concentration. All these reasons weaken the oxidation reaction and
produce more CO emission.
Effect of internal and external EGR on specific fuel consumption
The effect of IEGR and external cooled EGR on specific fuel consumption is seen
in figure 3.71. The indicated specific fuel consumption of HCCI combustion is higher
than that of a conventional diesel engine.

Figure 3.70 The effect of internal and external EGR on fuel consumption.
With the increase of NVO, the specific fuel consumption decreased, because the
larger NVO benefits the fuel evaporation. The 35% external cooled EGR does not affect
the specific fuel consumption very much. This may be because of the conflict of delayed
SOC and the incomplete combustion caused by the increased EGR.
CHAPTER 4: RESEARCH ON SKYACTIV-X ENGINE
4.1. Introducing Mazda’s Skyactiv technology
Skyactiv (also known as SKYACTIV) is a brand name for a set of technologies
developed by Mazda to improve fuel efficiency and engine power. The original Skyactiv
technologies announcement featured new engines, gearboxes, bodies, and chassis, which
featured in Mazda products beginning in 2011.

Mazda's SKYACTIV-X engine, which features Spark-Controlled Compression


Ignition, a never-before-seen combustion process, is the second phase in Mazda's quest to
produce a gasoline engine with the optimal internal combustion mechanism. Engineers
have long desired to develop compression ignition for gasoline engines. Spark plug
ignition is employed to govern compression ignition in the SKYACTIV-X, delivering in
remarkable gains across a variety of crucial performance measures.
The SKYACTIV-X is a ground-breaking new engine exclusive to Mazda that
combines the benefits of a spark-ignition gasoline engine (expansion at high rpms and
cleaner exhaust emissions) with those of a compression-ignition diesel engine (superior
initial response and fuel economy) to create a crossover engine that delivers the best of
both worlds. Following Mazda's SKYACTIV-G gasoline engine and SKYACTIV-D
diesel engine, this third SKYACTIV engine has been renamed "X" to reflect its dual
purpose, as shown in figure 4.
Mazda believes that there is still plenty of opportunity for future improvement of
the internal combustion engine and that this technology has the potential to make a
significant contribution to global environmental conservation. Based on Mazda's business
objective of conserving our beautiful world while enhancing people's lives via the "joy of
driving," we intend to pursue the perfect combustion engine indefinitely.
Figure 4. Mazda/s roadmap toward the ideal ICE
4.2. Advantages of Skyactiv-X engine

Figure 4. SkyActiv-X technology features on the Mazda 3.


Following the path outlined above, we have conducted a comprehensive
reexamination of the nature of combustion in order to make significant improvements in
the efficiency of the internal combustion process. The SKYACTIV-G improves
combustion efficiency by increasing the compression ratio, while cooling losses from the
zone of heat transfer to the chamber wall are decreased by temperature regulation of the
cooling water. Using the Miller cycle reduces pumping losses and mechanical resistance.
The newest SKYACTIV engine, the SKYACTIV-X, was developed to improve the air-
fuel ratio. To do this, needed to allow lean burn, which involves combusting higher
amounts of air. The optimal air-fuel ratio (stoichiometric) is 14.7:1. Increasing the air-to-
fuel ratio by more than doubling the volume of air increases the specific heat ratio while
decreasing the combustion gas temperature. As a result, cooling losses are reduced.
Meanwhile, a design that admits more air decreases the losses caused by throttle closure,
resulting in better fuel economy. Better Fuel Economy, lower costs of motoring due to
increased engine efficiency (including low tailpipe emissions for CO2 and NOx) -
without compromising any power loss, as shown in figure 4.

Figure 4. SkyActiv-X technology on mazda 3 compares with the C-Segment petrol


(https://www.nortonway.com/mazda/about-us/why-buy-mazda/skyactiv-x-
technology/)
However, when this type of lean combination of air and gasoline is burnt utilizing
flame propagation-based combustion, as occurs when a spark plug is utilized, combustion
becomes unstable. To address this issue, high-temperature, high-pressure compression
combustion must be used. This means that such an engine will have to utilize
compression ignition, which is common in diesel engines. As a result, while designing
the SKYACTIV-X, we enhanced the seven parameters that must be regulated for
compression ignition of a low air-fuel combination. These include the compression ratio
(which must be increased to provide the needed high-temperature, high-pressure
circumstances), combustion timing at top dead center (as seen in compression ignition),
and a combustion phase during which all of the fuel burns concurrently.
4.3. Skyactiv-X with Spark Controlled Compression Ignition
“Compression ignition doesn’t require a spark plug, but a spark plug will still be
needed in the rpm and load ranges where compression ignition cannot take place.
Unfortunately, switching between these two modes is extremely difficult.” This is the
“received wisdom” about HCCI, setting out the main issue which has prevented HCCI
technology from being fully commercialized. Mazda’s breakthrough has been achieved
by questioning the conventional idea that no spark plug is needed for compression
ignition and suggesting a different approach instead: “If switching between different
combustion modes is difficult, do we really need to switch in the first place?” This idea is
the foundation of Mazda's unique combustion process, Spark-Controlled Compression
Ignition (SPCCI). Using SPCCI implies that the compression ignition range (in terms of
engine load and rpm) now encompasses the whole combustion range. That is to say, the
possible application of compression ignition has now been greatly enlarged, allowing this
technology to be utilised in practically every driving situation. In other words, because a
spark plug is being employed at all times, the engine can easily convert between
compression ignition and spark ignition combustion.

4.3.1. Using compression effects created by flame propagation


The SPCCI mechanism is a system in which the compression effect of spark-
ignited localized combustion is exploited to create the needed pressure and temperature to
cause compression ignition. In other words, the geometric compression ratio is increased
to the point where the air-fuel combination at top dead center is on the edge of igniting
(owing to compression). An expanding flame formed by spark ignition gives the last push
that allows the entire mixture to burn at this stage. The time and amount of pressure
necessary are continually changing due to continually changing driving conditions.
Because the SPCCI system can adjust the spark plug ignition timing, pressure and
temperature within the combustion chamber may be optimized at all times. Because a
spark plug is always employed, the system can effortlessly switch to spark ignition
combustion at rpm or load ranges where compression ignition would be impossible. As a
result, the system ensures that the compression ratio is never excessively high, while
allowing for a basic design that does not necessitate sophisticated features like variable
valve timing or variable compression ratio.
4.3.2. Fuel density distribution within the air-fuel mixture
Using the SPCCI mechanism, the SKYACTIV-X adjusts the distribution of the
air-fuel mixture to enable lean burn. First, a lean air-fuel mixture is spread throughout the
combustion chamber for compression ignition. Then, precision fuel injection and swirl
are used to generate a zone of richer air-fuel mixture surrounding the spark plug, rich
enough to be ignited with a spark while minimizing nitrous oxide emission. SPCCI
ensures steady combustion by employing these approaches.

4.3.3. Controlling the air-fuel mixture to prevent abnormal combustion


 Split fuel injection
To avoid the abnormal combustion that can occur when rich air-fuel mixtures are
compressed for extended periods of time—a long-standing issue for HCCI—SPCCI uses
a split fuel injection system, in which part of the fuel is injected during the air intake
process and part during the compression process. During the air intake phase, the low-
density lean mixture for the lean burn is injected; then, during the compression stroke, a
second injection provides the richer air-fuel combination that is ignited around the spark
plug. This not only distributes the density of the air-fuel mixture, allowing SPCCI to
occur, but also reduces the time lag before the air-fuel combination ignites under
compression, effectively limiting aberrant combustion.
 Super-high-pressure fuel injection system
To reduce compression time and maximize compression ignition efficiency, the fuel must
be vaporized and atomized quickly and then disseminated evenly throughout the cylinder.
As a result, the SKYACTIV-X has a system capable of injecting super-high-pressure fuel
from a multi-hole fuel injector located in the center of the combustion chamber. This
quickly vaporizes and atomizes the fuel while simultaneously creating tremendous
turbulence, considerably enhancing ignition stability and combustion speed. Super-high-
pressure fuel injection enables SPCCI, which suppresses abnormal combustion even at
full throttle/low rpms where traditional gasoline engines have to retard ignition and thus
sacrifice efficiency and power. SPCCI is enabled by super-high-pressure fuel injection,
which suppresses aberrant combustion even at full throttle/low rpms, when typical
gasoline engines must delay ignition and so forfeit efficiency and output.
 Adoption of the in-cylinder pressure sensor
In addition to the above-mentioned technologies for preventing abnormal combustion, an
in-cylinder sensor has been introduced as a monitoring control; it ensures continuously
optimized combustion by continuously observing whether the above controls are bringing
about proper combustion and compensating in real time for any deviations from intended
outcomes. Based on the methodologies described above, SPCCI has extended the zone of
compression ignition all the way to the whole throttle range, allowing for smooth
transitions between SPCCI and spark ignition combustion.

This novel combustion approach provides an all-encompassing combustion control


system that includes control of in-cylinder temperature and pressure, as well as control of
the fuel injection's air-fuel mixture distribution density and exhaust gas recirculation
(EGR).
4.4. Value provided by Skyactiiv-X engine
4.4.1. Dramatically improved output performance and responsiveness
With a 2.0L engine, the SKYACTIV-X produces at least 10% more torque than the
current SKYACTIV-G, and up to 30% more at certain rpms (data as of August 2017,
during the development process). Furthermore, because the throttle valve is open the
majority of the time, it has the greater initial acceleration response found in diesel
engines that lack a throttle valve. The SKYACTIV-X, on the other hand, spins up to
higher rpm ranges as smoothly and effortlessly as a regular gasoline engine.
4.4.2. Dramatic improvement in fuel economy
In a car with a 2.0L engine displacement, the SKYACTIV-X improves fuel
efficiency by 20% over the SKYACTIV-G, which is a significant improvement.
Furthermore, in places where low vehicle speeds are regularly used, super lean
combustion can enhance fuel economy by up to 30 percent. Fuel economy is enhanced by
35-40% when compared to the MZR engine from 2008, and SKYACTIV-X equals or
exceeds the fuel efficiency of Mazda's latest diesel engine, SKYACTIV-D. With
significant increases in the light engine load range, this engine debunks the frequently
held belief that a large engine displacement equates to low fuel economy. With the use of
the SKYACTIV-X, the engine's ability to deliver excellent fuel economy has been
dramatically expanded, implying that this system can now deliver lower fuel
consumption than ever before in a wide range of driving scenarios, including city driving,
long-distance driving on expressways, and more.
The SKYACTIV-X, which is unique to Mazda, is a new type of combustion
engine that blends the benefits of gasoline and diesel engines to provide great
environmental performance as well as unrivaled power and acceleration performance.
This ground-breaking technology marks the beginning of an exciting new chapter in our
quest to create the perfect internal combustion engine.
CHAPTER 5: CONCLUSION AND RECOMMENDATIONS
5.1. Conclusion
5.2. Recommendations

REFERENCES
[1] Zhao H, editor, HCCI and CAI engines for the automotive industry. Cambridge:
Woodhead Publishing; 2007.
[2] Onishi, S., Hong Jo, S., Shoda, K., Do Jo, P., and Kato, S., Active Thermo-
Atmosphere Combustion (ATAC) - A New Combustion Process for Internal Combustion
Engines. SAE Paper 790501, 1979.
[3] Noguchi, M., Tanaka, Y., Tanaka, T., and Takeuchi, Y., A Study on Gasoline Engine
Combustion by Observation of Intermediate Reactive Products during Combustion, SAE
Paper 790840, 1979.
[4] Najt, P. M., and Foster, D. E., Compression-Ignited Homogenous Charge
Combustion. SAE Paper 830264, 1983.
[5] Thring, R. H., Homogenous Charge CompressionIgnition (HCCI) Engines. SAE
Paper 892068, 1989
[6] Christensen, M., Hultqvist, A., and Johansson, B., Demonstrating the Multi Fuel
Capability of a Homogenous Charge Compression Ignition Engine with Variable
Compression Ratio. SAE Paper 1999-01-3679, 1999.
[7] Oakley, A. Zhao, H., Ma, T. and Ladommatos, N., Experimental Studies on
Controlled Autoignition (CAI) Combustion of Gasoline in a 4-Stroke Engine. SAE Paper
2001-01-1030.
[8] Hua Zhao (2007), HCCI and CAI engines for the automotive industry. Woodead
publishing limited, Cambridge England.
[9] Zhang, Y., Ojapah, M., Cairns, A., and Zhao, H., 2-Stroke CAI Combustion
Operation in a GDI Engine with Poppet Valves. SAE Technical Paper 2012-01- 1118,
2012.
[10] Zhang Y, Zhao, H et al, Experiment and analysis of a direct injection gasoline
engine operating with 2-stroke and 4-stroke cycles of spark ignition and controlled auto-
ignition combustion. SAE paper 2011–01-1774; 2011
[11] Yan Zhang, Hua Zhao, Mohammed Ojapah, Alasdair Cairns, CAI combustion of
gasoline and its mixture with ethanol in a 2-stroke poppet valve DI gasoline engine.
[12] Li, J., Zhao, H., Ma, T., and Ladommatos, N., Research and development of
controlled autoignition (CAI) combustion in a 4-stroke multicylinder gasoline
engine.SAE paper 2001-01-3608
[13] Hua Zhao, Jian Li, Tom Ma and Nicos Ladommatos, Performance and Analysis of a
4-Stroke Multi Cylinder Gasoline Engine with CAI Combustion, SAE paper 2002-01-
0420
[14] Stranglmaier RH, Roberts CE, Homogeneous Charge Compression Ignition
(HCCI): benefits, compromises, and future engine applications. SAE 1999-01-3682;
1999
[15] Martinez-Frias J, Aceves SM, Flowers D, Smith JR, Dibble R (2000), HCCI engine
control by thermal management (No. 2000-01-2869). SAE technical paper
[16] .Yang J, Culp T, Kenney T (2002), Development of a gasoline engine system using
HCCI technology-the concept and the test results. (No. 2002-01-2832). SAE technical
paper.
[17] Kodavasal J, Lavoie GA, Assanis DN, Martz JB (2015), The effects of thermal and
compositional stratification on the ignition and duration of homogeneous charge
compression ignition combustion. Combust Flame 162(2):451–461.
[18] Haraldsson G, Tunesta˚l P, Johansson B, Hyv€ onen J (2002), HCCI combustion
phasing in a multi cylinder engine using variable compression ratio (No. 2002-01-2858).
SAE technical paper.
[19] The Saab Network, “Saab variable”, retrieved from
(http://www.saabnet.com/tsn/press/000318.html).
[20] 2019 Nissan Altima features VC-Turbo variable compression ratio engine from
(https://www.greencarcongress.com/2018/03/20180329-altima.html)
[21] Walter B, Pacaud P, Gatellier B (2008), Variable valve actuation systems for
homogeneous diesel combustion: how interesting are they? Oil Gas Sci Technol-Revuede
l’IFP 63 (4):517–534.
[22] C.Baumgarten(2005),Mixture formation in Internal Combustion Engines.
Springer-Verlag Berlin Heidelberg New York.
[24] Walter B, Gatellier B. Near zero NOx emissions and high fuel efficiency
diesel combustion: the NADITM concept using dual mode combustion. Oil Gas
Sci Technol 2003.
[25] Kim MY, Lee CS. Effect of a narrow fuel spray angle and a dual injection
configuration on the improvement of exhaust emissions in a HCCI diesel engine.
Fuel 2007;86(17–18):2871–80
[26] Kimura, S, Aoki O, Ogawa H, Muranaka S et al. New combustion concept for
ultraclean and high-efficiency small DI diesel engines. SAE 1999-01-3681; 1999.
[27] Kimura S, Matsui Y, Itoh T. Effects of combustion chamber insulation on the
heat rejection and thermal efficiency of diesel engines. SAE 920543; 1992.
[28] Jun-ichi Kawashima, Hiroshi Ogawa and Yoshiyuki Tsuru. Research on a
Variable Swirl Intake Port for 4-Valve High-Speed DI Diesel Engines. SAE
982680; 1998.
[29] Maurya RK (2018) Characteristics and control of low temperature
combustion engines: employing gasoline, ethanol and methanol. Springer. ISBN
978-3-319 68507-6.
[30] Zhao H 2007 HCCI and CAI engines for the automotive industry (Cambridge
England: Wood head Publishing Limited).
[31] Mathivanan K, Mallikarjuna J M and Ramesh A (2016). Experimental Thermal and
Fluid Science. 77, pp. 337-346
[32] Yoon, S H, Kim H J and Park S (2018). Applied Thermal Engineering. 144, pp.
1081-1090.
[33] Han Z, Uludogan A, Hampson G and Reitz R (1996). SAE 960633.
[34] Khandal S, Banapurmath N,Gaitonde V and Hiremath S (2017). Renewable and
Sustainable Energy Reviews. 70, pp. 369–384
[35] Epping K, Aceves S, Bechtold R and Dec J (2002). SAE Technical Paper 2002-01-
1923.
[36] Agarwal K, SinghA P and Maurya R K (2017). Progress in Energy and Combustion
Science. 61, pp. 1-56
[37] Han S, JaeheunKim J, and Bae C (2014). Applied Energy. 119, pp. 454–466
[38] Wang X, Huang Z, Zhang W, Kuti O and Nishida K (2011). Applied Energy 88,
pp.1620–1628
[39] Nishida K, Zhu J, Leng X and He Z (2017). International J of Engine Research.
18(1-2), pp. 51– 65.
[40] Nathan S, Mallikarjuna S and Ramesh A (2009). SAE 2009-26-028.
[41] Najt P and Foster D (1983). SAE 830264.
[42] Kong S and Reitz R (2002). J. Eng. Gas Turb. Power ASME. 124, pp. 702–7.
[43] Kong S, Marriott C, Reitz R and Christensen M (2001). SAE 2001-01-1026.
[44] Iida M, Aroonsrisopon T, Hayashi M, Foster D and Martin J (2001). SAE 2001-01-
1880/4278
[45] Aroonsrisopon T, Foster DE, Morikawa K and Lida M (2002). SAE 2002-01-1924;
2002.
[46] Iida M, Hayashi M, Foster D and Martin J (2003). J. Eng. Gas Turb. Power ASME.
125, pp. 427– 78.
[47] Szybist J and Bunting B (2005). SAE 2005-01-3723.
[48] Garcia M, Aguilar F and Lencero T (2009). Energy. 34, pp. 159–71.
[49] Genzale C, Kong S and Reitz R (2008). J. Eng. Gas Turb. Power ASME. 130, p. 8.
[50] Maurya R and Agarwal A (2009). SAE 2009-01-1345.
[51] Ebrahimi R and Bernard Desmet B (2010). Fuel. 89, pp. 2149–2156
[52] Hadia F, Wadhah S, Ammar H and Ahmed O (2017). Case studies in thermal
engineering. 10, pp. 262-271.
[53] Alper Calam A, Solmaz H, Icingur Y and Yilmaz E (2018). Energy. 168, pp. 1208-
1216.
[54] Shenawy E, Elkelawy M, Bastawissi H A, Panchal H and Shams M (2019). Fuel.
249, pp. 277– 285.
[55] Ganesh D and Nagarajan G (2010). Energy. 35(1), pp. 148–57.
[56] Yu S, Li L and Zheng M (2019). Fuel 242, pp. 243–54.
[57] Liu H, et al (2018). Fuel. 231, pp. 26–36.
[58] Lu X, Han D and Huang Z (2011). Prog. Energy Combust Sci. 37(6), pp. 741–83.
[59] An Y et al (2018) Energy. 158, pp. 181–91.
[60] Tanaka S, Ayala F, Keck J and Heywood J (2003). Combustion and Flame. 132, pp.
219–239.
[61] Asad U, Zheng M (2008). International Journal of Thermal Sciences. 47, pp. 1688–
1700.
[62] Fathi M, Saray R K, Checkel M D (2010). Fuel. 89, pp. 2323–2330.
[63] Law D., Kemp D., Allen J., Kirkpartick G., and Copland T. Controlled
Combustion in an IC-Engine with a Fully Variable Valve Train. SAE Paper 2000-
01-0251, 2000.
[64] Nebojsa Milovanovic and Rui Chen. Influence of the Variable Valve Timing
Strategy on the Control of a Homogeneous Charge Compression (HCCI) Engine.
SAE Paper 2004-01-1899, 2004.
[65] Law D., Kemp D., Allen J., Kirkpartick G., and Copland T. Controlled
Combustion in an IC-Engine with a Fully Variable Valve Train. SAE Paper 2000-
01-0251, 2000.
Appendices

You might also like