Está en la página 1de 33

Engine Design for Low Emissions [subscription] Page 1 of 33

DieselNet Technology Guide

www.DieselNet.com. Copyright Ecopoint Inc. Revision 2009.09

Engine Design for Low Emissions


W. Addy Majewski, Hannu Jskelinen

Abstract: Changes in diesel engine design contributed to some 10-fold decrease in emissions over the
period from the late 1980s to early 2000s. The most important of these engine technologies are advanced
fuel injection systems, air intake improvements, combustion chamber modifications, and electronic engine
control. Additionally, exhaust gas recirculation (EGR) was introduced on both light- and heavy-duty diesel
engines to control NOx emissions. Low emission engine designcombined with increased exhaust gas
aftertreatmentwill continue to play important role in future diesel engines.

Introduction

Engine Design Optimization

Emission Control Techniques

Improving Fuel Economy

Evolution of Engine Technology

Other Emission Control Technologies

Introduction

Since the introduction of the first diesel emission standards in the 1970s, and even more so after the
introduction of the first particulate matter emission standard in 1988 (0.6 g/bhp-hr) and the subsequent
increasingly stringent standards for all regulated pollutants, the need to lower emissions became the
prime driving factor in the development of the diesel engine. The evolution of NOx and PM standards for
heavy-duty engines in the USA and in European Union is shown in Figure 1.

Figure 1. Emission Standards for Heavy-Duty Diesel Engines

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 2 of 33

Note: The US 2007 limits shown reflect a common phase-in option of US 2010 emission standards.

As a result of these standards, diesel engine emission levels decreased dramatically in the period from
the late 1980s to the early 2000s. For example, PM emissions from heavy duty diesel engines in the USA
decreased by a factor of 10, from about 1 - 2 g/bhp-hr in the 1970s to below 0.1 g/bhp-hr in 1994.
Reductions of similar magnitude were achieved for emissions of nitrogen oxides and hydrocarbons. This
spectacular clean-up of the diesel engine was achieved through engine design methods in conjunction
with better fuels. The US 1994 PM standard of 0.1 g/bhp-hr was accompanied by the introduction of low
sulfur diesel fuel of 500 ppm (0.05%) S, down from the previous 0.5%. Similar sulfur levels were
legislated in the EU, with a 500 ppm S cap introduced in 1996 and a 350 ppm S limit in 2000.

Diesel emission reductions in this time periodup to and including the US 2004 standardswere
achieved without relying on exhaust gas aftertreatment (post-combustion) technologies. Only emission
standards that took effect later forced the use of advanced aftertreatment devices such as diesel
particulate filters and NOx reduction catalysts by setting extremely low emission limits. From this
perspective, the 1980s and 1990s represent a transitional pre-aftertreatment period in the evolution of
diesel emission control technology.

It should be noted that a number of diesel engines in this transitional, pre-aftertreatment period used
exhaust gas aftertreatment devices. However, aftertreatment was typically not the critical technology
used for compliance in these engines, especially for the most troublesome emissions of NOx and PM. For
example, many US 1994 mechanically controlled engines used diesel oxidation catalysts to support PM
emission reduction. Electronically controlled diesel engines were able to meet the 1994 0.1 g/bhp-hr PM
standard with no exhaust aftertreatment. As the mechanically controlled engines were phased out, so was
the use of catalytic converters on highway trucks in the USA.

The widespread use of diesel oxidation catalysts was seen on Euro 2/Euro 3 (1996/2000) diesel cars. An
important reason for introducing these catalysts was the required reduction of CO and HC emissions.
The PM emission reduction and in some cases the lean NOx functionality provided by these catalysts
was limited and cannot be considered prime strategies for meeting emission standards. In some cases car
manufacturers introduced aftertreatment voluntarily, to promote the clean image of the diesel engine or
for other political reasons. Two noteworthy examples of such voluntary use of aftertreatment include the
catalyst-equipped Volkswagen Golf Umwelt in 1989 and the particulate filter-equipped Peugeot 607 in
2000.

Advanced aftertreatment became an integral part of the diesel engine to satisfy stringent emission limits
imposed by aftertreatment-forcing standards. Examples include the US 2007/2010, the Euro V (2008)
and Euro VI (2013), as well as the US Tier 2 (2004-2009 phase-in), California LEV II and Euro 5 (2009)
legislation for light-duty vehicles. The stringency of these regulations may be illustrated by the US 2007
PM limit of 0.01 g/bhp-hr which required another 10-fold emission reduction from the 1994 levels.
Meeting these standards generally requires a combination of integrated engine and aftertreatment
strategies, the latter being enabled by ultra low sulfur (10-15 ppm S) diesel fuel.

It should be emphasized that in-cylinder emission control continues to play a very important role in
engines with aftertreatment. Aftertreatment technologies come with an added cost. Furthermore, some of
them involve fuel economy penalties, while others have servicing requirements, such as the need to
replenish urea in engines using selective catalytic reduction (SCR) for NOx control, that may
inconvenience vehicle users. Through in-cylinder emission control methods, the required emission
reductions may often be achieved at a lower cost. If aftertreatment is still necessary, cost savings may be
possible on cleaner engines, for example by enabling smaller catalyst sizes. Several manufacturers
continue to develop ultra clean diesel combustion systems that can meet the most stringent emission
standards without the use, or with only a limited use, of aftertreatment. Examples of such engines include

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 3 of 33

heavy-duty Euro V and US 2010 engines without NOx aftertreatment.

Fuel economy and greenhouse gas emissions, in addition to the conventional pollutant emissions, are
becoming major drivers in the development of future internal combustion engines. The first CO2
emission standards have already emerged for light-duty vehicles, Figure 2 [ICCT 2009]. Due to the large
engine and vehicle sizes, passenger cars in North America historically had the highest CO2 emissions.
However, compared to other regions, the United States is expected to require the highest CO2 emission
reductions starting from 2011-2012. Fuel economy and GHG regulations are also being developed for
heavy-duty engines. Hence, increased thermal efficiency while meeting near-zero emissions is now the
focus in engine development.

Figure 2. GHG Emission Standards for New Passenger Vehicles


Standards converted to the NEDC test. Dashed lines indicate proposed regulations.

New technologies, as well as improvements to existing technologies are required for the diesel engine to
meet the emissions and engine efficiency challenges. The diesel engine industry has implemented and is
further investigating several approaches to achieve the NOx and PM emission reductions required by
emission standards:

In-cylinder and engine-based design changes


Fuel and lubricant formulations that reduce engine-out emissions and are compatible with exhaust
aftertreatment systems.
Aftertreatment (post-combustion) exhaust emission control devices.

In-cylinder and engine-based design changesthe subject of this papercover a number of the engine
subsystems including the fuel injection, induction, combustion, valve train, engine cooling, the electrical
system, as well as accessories. With the advent of electronics and their massive use in the automotive
technology, more of the emerging technologies are based on the capabilities afforded by electronic
controls. From intelligent sensors to sophisticated control algorithms, controls technologies are opening
new opportunities for automotive systems.

Engine Design Optimization

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 4 of 33

In comparison to other types of internal combustion engines, the conventional diesel combustion process
produces low emissions of carbon monoxide and hydrocarbons. Limits for CO and HC emission, such as
those for many US diesel engines, have long been met without any special emission control measures. In
applications with more stringent CO/HC requirements, such as EU diesel cars, these emissions have been
relatively easy to control with diesel oxidation catalysts. However, trends in diesel engine development
have lead to increases in CO/HC emissions and reduced the efficiency of catalytic converters so that
control of these emissions has become much more challenging.

By far the most difficult challenge in the design of diesel engines continues to be compliance with NOx
and PM emission standards. Both of these regulated pollutants are generated in relatively high quantities
in the conventional diesel combustion process.
NOx-PM Trade-Off. Simultaneous NOx and PM control becomes even more complex because of a
trade-off between them from the diesel engine. In conventional diesel combustion (before the
development of combustion processes that allow engine operation at significantly lower combustion
temperatures), engine design techniques that lead to a reduction of PM increased NOx emissions and vice
versa. Diesel injection timing is a well known example of this behavior. Retarded injection timing is a
common approach to control NOx emissions. However, it can be used only to a certain point, beyond
which unacceptably high particulate emissions will occur.
The concept of NOx-PM trade-off is illustrated in Figure 3 [Needham 1991], which depicts the evolution of
emission technology in heavy-duty diesel engines in the 1990s. The thick navy-blue lines represent NOx-
PM emission levels that can be achieved by different engine technologies. The thin red lines reflect
selected emission standards for heavy-duty diesel truck engines.

Figure 3. NOx-PM Trade-Off

Within each engine technology group, lower NOx emissions can be achieved at the cost of higher PM
emissions. More advanced engine technologies form lines gradually shifting toward the origin of the
plot. For example, lines for turbocharged diesel engines (TC in Figure 3) show NOx and PM emissions
that are lower than those for naturally aspirated (NA) engines. Further emission reductions are achieved
by the introduction of turbocharged and aftercooled diesel engines (TCA) and through electronic engine
control. More advanced engine technologies, not shown in Figure 3, continue the same trend producing

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 5 of 33

lines of similar shapes, but positioned closer to the beginning of the chart. Emission trade-off lines of a
similar character can be also drawn for light-duty diesel engines.
The NOx-PM emission trade-off is also an important consideration for the integration of aftertreatment
strategies with the diesel engine. In some applications, aftertreatment can be used to control only one of
the pollutants, while the other can be controlled in-cylinder. There are two possible approaches:

i. Low NOx engine calibration with PM emission aftertreatmentA diesel particulate filter system
can be designed to control increased PM emissions that result from low NOx engine design. The
NOx reduction potential in this case is determined by NOx-PM trade-off line of the particular
engine. Since the diesel particulate filter technology is more mature than NOx aftertreatment
technologies, this approach has been used in a number of light- and heavy-duty engine
applications.
ii. Low PM engine strategy with a NOx control deviceThis opposite concept involves a NOx
aftertreatment device combined with low PM/high NOx engine calibration. An example of this
approach is heavy-duty Euro V engines using a urea-SCR NOx control system without the use of
PM aftertreatment.

Operating an engine at different points along the NOx-PM trade-off line affects not only emissions, but
also other engine performance parameters. This includes fuel economy, which tends to improve with
higher engine-out NOx/lower PM.

It should be emphasized that the concept of a NOx-PM trade-off is based on work with conventional
diesel combustion processes that degrade rapidly if attempts are made to lower combustion temperatures
too low. Developments that allow lower combustion temperatures to lower engine-out NOx emissions
show that the nature of the NOx-PM trade-off can change dramatically at low temperature/low NOx
levels.

Not only did work with conventional diesel combustion lead to the assumption that there was always a
NOx-PM trade-off, it also seemed to point to a lower limit on NOx that could be practically achieved
with in-cylinder emission reduction approaches. In heavy-duty diesel engines, the majority of energy is
released in a diffusion flame, resulting in high flame temperatures and high NOx production. Cooling the
flame reduces the formation of NOx at the expense of increases in PM, HC, and CO emissions.

Studies have attempted to predict the NOx reduction limits that can be achieved by engine design
techniques. One such study attempted to determine the minimum flame temperature required to maintain
complete combustion in engines in order to estimate the minimum NOx emissions that could be achieved
through combustion temperature reductions. Selected results are presented in Figure 4, which depicts
fuel specific NOx emissions as a function of flame temperature [Flynn 2000a]. The minimum flame
temperatures are indicated by red rectangles. The authors concluded that conventional diesel combustion
must have peak flame temperatures above about 2300 K which would result in a minimum of 5.5 g
NOx/kg of fuel. This is equivalent to about 1.5 g/bhp-hr over the FTP Transient test. In other words,
NOx could be reduced to 1.5 g/bhp-hr by engine design methods; reductions below that level would
require the use of NOx aftertreatment. For comparison, the equivalent NOx limits for the spark ignited
engine are 2.5 g/kg-fuel and 0.75 g/bhp-hr, respectively.

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 6 of 33

Figure 4. Fuel Specific NOx as a Function of Flame Temperature

Fuel Economy. Most engine design techniques that are governed by the NOx-PM trade-off produce not
only increased PM levels when lowering NOx, but also exhibit a thermal efficiency loss and increased
fuel consumption. Fuel economy penalties may be minimized by careful selection of NOx control
strategies, as illustrated in Figure 5 which shows predicted BSFC increases for the conventional timing
retard and EGR-based NOx control techniques [Wall 1997].

Figure 5. Impact of NOx Reduction Technologies on Fuel Economy

IMT - intake manifold temperature; EGR - exhaust gas recirculation

One has to keep in mind, however, that the studies mentioned above [Flynn 2000a][Wall 1997] were carried
out on engines representative of the technology used in the 1990s. It has been demonstrated that diesel
engines can be operated even when peak combustion temperatures are far lower than 2300 K. One
approach that used high EGR and fuel injection pressures in the range of 200-300 MPa has shown that
diesel combustion can occur below 1970 K [Haugen 2004] and produce NOx less than 0.2 g/bhp-hr. Others
have demonstrated diesel engine operation at temperatures below 1600 K [Akihama 2001]. The term low
temperature combustion (LTC) is commonly used to refer to such approaches that allow engines to
successfully operate a combustion temperatures lower than those encountered in conventional diesel
combustion.

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 7 of 33

The lower limit on engine-out NOx emissions is technology specific and is largely determined by the
lowest combustion temperature at which an engine can operate without unacceptable increases in fuel
consumption or other emissions. Advances in fuel injection and engine air management technology have
allowed engines to operate at lower combustion temperatures by using higher levels of charge dilution.
As a result, engine out NOx emissions have come down to the point that engine manufacturers have been
targeting engine-out NOx on the order of 0.5 g/kg-fuel (approximately 0.1 g/bhp-hr) [Durrett 2008]about
ten times lower than the value shown in Figure 4.

Low Temperature Combustion and NOx-PM Trade-Off. Until engine-out NOx levels were pushed
well below the limits thought possible in the 1990s, it was commonly believed that little could be done to
overcome the NOx-PM trade-off illustrated in Figure 3. However, work with LTC has shown that,
provided the engine is properly designed, the nature of the NOx-PM characteristic as combustion
temperature is decreased can change dramatically at low NOx levels (Figure 6) and measures that reduce
NOx can also reduce PM emissions [Green 2002]. For engines that can successfully operate with lower
combustion temperatures, not only do NOx emissions decrease, but the combustion temperature becomes
too low for significant production of combustion generated soot. Under these conditions, there is no
longer a trade-off between NOx and combustion generated soot emissions.

Figure 6. Impact of Decreasing Combustion Temperature on NOx-PM Trade-Off

Low temperature combustion relies on high rates of cooled EGR to achieve low combustion
temperatures and low NOx and some combination of low temperature and higher rates of fuel and air
premixing to limit soot emissions. While in some cases it can eliminate the NOx-PM trade-off, it can
introduce several other trade-offs [Durrett 2008]:

Soot vs. Combustion NoisePremixed combustion produces increased noise levels compared to
conventional diesel combustion. This trade-off is one of the key factors that limit the use of some
LTC approaches to light loads.
NOx vs. HydrocarbonsOxidation of unburned HCs and CO becomes difficult as temperature
decreases and they tend to increase with low temperature combustion, primarily at light load

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 8 of 33

operation.
Exhaust Temperature vs. NOxHigh EGR rates increase the mass flow through the engine, which
at constant fuel energy causes the exhaust temperature to decrease. Thus, lower engine-out NOx
may be accompanied by reduced aftertreatment effectiveness.

Fuel economy is another important variable that may limit the NOx emission reductions available
through LTC, just as was the case with conventional diesel combustion. The introduction of these other
trade-offs introduces a six way optimization for engine design with the following variables: soot, NOx,
HCs, noise, exhaust temperature, and fuel economy.

Emission Control Techniques

Overview
The following sections discuss selected emission reduction techniques related to engine technology in
the period after 1988 (aftertreatment systems which play a critical role in meeting diesel emission
standards beginning from mid-2000s are discussed elsewhere). Some of these methods were used to
control NOx emissions, others to control PM emissions, still others to recover fuel economy losses due
to emission control. The most important engine measures taken to help meet emission standards can be
classified as follows:

Fuel injection system modifications


Injection timing

High injection pressure


Electronic injection control
Injector nozzles (small holes, low sac volume, converging nozzles)
Air intake improvements
Cooled charge air (inter- and aftercoolers)

Progress in turbocharger technology (improved performance, variable geometry


turbochargers, two-stage turbochargers)
New intake manifolds
Swirl ratio match with fuel injection characteristics
Combustion chamber improvements
Reentrant bowls

Higher top piston rings


Better air utilization (central injection, four valves)
Lower compression ratio
Variable valve actuation (Miller cycle)
Exhaust gas recirculation
Low temperature combustion
Electronic engine control

Fuel Injection
Injection Timing
Among the numerous fuel injection system improvements, ignition timing has been one of the basic tools
to control NOx emissions from diesel engines. Advanced injection favors premixed combustion, high

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 9 of 33

pressures, high combustion temperatures, and increased fuel economy. Without additional measures such
as charge dilution with EGR, this will increase NOx emissions. Retarded ignition reduces NOx
emissions and increases emissions of HC, PM, and smoke. It also negatively influences fuel economy
and has a negative impact on the contamination of lubricating oil with soot. The fuel economy and
smoke emissions penalty for retarded timing can be minimized by such measures as pilot injection and
optimization of the combustion bowl and swirl ratio. An example of the effect of ignition timing on
noise, emissions and fuel consumption from a prototype direct injection diesel engine based on the
Duramax V8D 6600 used in many GM pick-up trucks is shown in the Figure 7 [Ishikawa 2004].

Figure 7. Effect of Ignition Timing on Emissions


Light-duty turbocharged DI diesel engine; 0.825 l/cyl., 1500 rpm, 432 kPa BMEP, 100 MPa injection
pressure, 24% EGR, 1.3 swirl ratio

Injection Pressure
New fuel injection systems utilize increasingly higher injection pressures. The high pressures are needed,
especially in DI engines, to introduce fuel to the combustion chamber at sufficiently high velocity in
order to (1) promote the premixing of air and fuel in the near nozzle region by increasing the fuel jet lift-
off length and by allowing smaller injector hole sizes and (2) to ensure that the fuel penetrates the entire
chamber and fully utilizes the air charge. These are important to help limit PM emissions as shown in
Figure 8 [Su 1995].

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 10 of 33

Figure 8. Impact of Injection Pressure on PM Emissions


Caterpillar 3406 diesel engine

There is a widespread belief that fuel injection pressures continue to rise so that further improvements in
fuel atomization can be achieved. While atomization can improve with higher pressures and this may
have been the primary reason for raising pressures in older diesel engine fuel injection systems that used
relatively moderate injection pressures, pressures and injector hole sizes in many modern injection
systems produce a fuel spray that is very finely atomized. Evidence strongly supports the notion that
under most operating conditions, the injection pressures used by modern diesel injection systems are
capable of producing sprays that are so fine that liquid evaporation is governed predominantly by factors
that affect the bulk entrainment of air into the fuel spray [Siebers 1998]. Further refinements in fuel jet
break-up and atomization resulting from the higher injection pressures would be of only minor
importance relative to the improvements in bulk mixing of air and fuel that occur.

One of the primary drivers for increased injection pressures is the use of EGR to control NOx emissions.
As engines utilize higher levels of EGR, the relative oxygen content decreases and higher injection
pressures are required to achieve higher levels of mixing of fuel and air/EGR. The increased mixing
ensures a sufficient amount of oxygen is still supplied to the burning fuel spray to offset potential
increases in soot emissions.

For pre-1991 heavy-duty diesel engines, full load rated speed injection pressure was in the range of 70 -
75 MPa (approx. 10,000 - 11,000 psi or 700 - 750 bar). In 1991 engines, the pressure approached 100
MPa (15,000 psi or 1000 bar) [Khair 1992]. Common rail injection systems work at pressures of 160 MPa
(23,000 psi or 1600 bar) and higher with pressures of 200 MPa becoming increasingly common. Unit
injectors used on many heavy-duty diesel engines can reach pressures of 250 MPa.

Changes in the fuel injection pressure may also lead to changes in several other parameters, including
fuel injection rate and injection duration.

Fuel Injector Nozzles


The design of fuel injector nozzles has paramount importance for the performance of the engine and its
emissions. The evolution of fuel injection nozzles has involved reduced sac volumes, optimized injection
hole numbers and diameters, optimized length of the nozzle hole and optimized spray cone [Khair 1992].

Two common injector nozzle designs are sac-type (or blind-hole nozzles) and valve covered orifice
(VCO) nozzles (Figure 9). The sac-type seats upstream of the nozzle opening and upon closing, can

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 11 of 33

contain fuel in its sac that can later evaporate and contribute to HC emissions. The valve seat in a VCO
injector covers the nozzle hole upon closing and does not allow unburned fuel to escape from the injector
when the injector is closed. Because of the potential for lower hydrocarbon emissions, VCO type nozzles
have become quite popular. Sac-type nozzles remain common thanks to their advantages including
higher stress capacity and better spray dispersal. Any potential increase in hydrocarbon emissions from
sac-type nozzles are easily dealt with if the engine contains a diesel oxidation catalyst.

Figure 9. Injector Nozzle Types


(A) Sac-less VCO nozzle. (B) Cylindrical and conical blind hole sac-type nozzles.
(Source: Bosch)

Another important nozzle development is the use of converging nozzle holes that have a larger inlet
diameter that outlet diameter. This allows reduced cavitation in the spray hole and an increase in fuel and
air mixing. There is also a reduced variability in hole-to-hole flow at maximum needle lift [Mahr 2002].

Variable Injection Nozzle Geometry. Injector nozzles with variable geometry are an interesting
development. They can be used to better optimize injector flow area and fuel spray characteristics for
different engine operating conditions.

In a standard diesel fuel injector, the spray hole diameter is defined by the number of spray holes, the
injection duration and the injected fuel quantity at full load. In some cases, this can make the hole too
large for part load operation and result in excessive soot formation. Figure 10 shows an example of a
variable flow area nozzle by Bosch. This Vario nozzle allows for a reduced spay hole area at part load
[Mahr 2002][Busch 2004].

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 12 of 33

Figure 10. Bosch Vario Nozzle

A prototype mixed mode injector developed by Caterpillar allows two different included spray angles,
Figure 11. With a narrow spray angle, fuel impingement on the cylinder wall can be avoided when fuel is
injected early in the compression stroke (during LTC mode operation) or late in the expansion stroke (to
raise exhaust temperature for DPF regeneration). The wider spray angle is intended for more
conventional diesel injection near TDC of the compression stroke. This injector allows mixed mode
operation of the engine where different combustion regimes are employed over different parts of the
engine map [Duffy 2004].

Figure 11. Caterpillar Mixed Mode Injector

Electronic Injection Systems


Electronically controlled fuel injection systems have replaced traditional mechanical systems in both
heavy-duty and light-duty diesel engines. The benefits of electronic full authority fuel injection systems
include better control of all injection parameters resulting in lower emissions and increased performance,
integration with the engine ECU and other vehicle systems, self-diagnostics, and performance checks.

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 13 of 33

Examples of electronic fuel injection systems include:

Electronically Controlled Injection Pumps. Mechanical fuel injection pumps vary the
fueling and injection timing using governors and mechanical linkages. Electronic
distributor type pumps controlled using electrohydraulic devices were introduced by
several manufacturers. In the Lucas EPIC system the cam ring is rotated by a hydraulic
actuator to vary injection timing [Hawley 1998]. Fueling is varied by moving the rotor
mechanism axially using another actuator. The pressure of working fluid, which was diesel
fuel, was regulated by the ECU through a solenoid valve. In some distributor type pumps
by Bosch (VP30, VP44), a solenoid operated spill valve controls the injection quantity. The
injection timing is still set by rotating a cam ring.

Electronic Unit Injector (EUI) System. Unit injector systems do not employ the central
injection pump supplying individual injectors. Instead, both pump and injector are
combined into a single unit for each cylinder. Each pumping plunger is driven directly by
the camshaft. This allows for very high injection pressures, up to 250 MPa. In the EUI
system, the fueling is controlled by a solenoid valve. While EUIs have some capacity for
pilot injections to reduce NOx emissions and noise, post injections can be a challenge and
they are used less frequently in light-duty vehicles that incorporate diesel particulate filters.

HEUI System. Caterpillar developed a hydraulic electronic unit injector (HEUI) system,
which is powered by hydraulic pressure and controlled electronically [Stockner 1993]. The
features of the HEUI system include no mechanical actuation or mechanical control
devices, injection pressure independent from the engine speed and load, and flexible
injection timing. The system is controlled by the ECU through a solenoid control valve that
starts and ends the injection process. The unit injectors are powered by hydraulic oil
delivered by an engine driven high pressure oil pump. The pump raises the systems oil
pressure from typical engine operating levels to the actuation pressure required by the
injector. The high pressure oil is distributed to the injectors through a manifold.

Common Rail System. In the common rail system, fuel is distributed to the injectors
through a high pressure manifold, called the rail. The injectors are solenoid or piezoelectric
actuated nozzles, operated by the ECU. Common rail is a competitor of the EUI system.
The advantages of common rail, relative to the traditional pump-line-nozzle system,
include injection pressure independent of engine speed and load and the flexibility to
control injection timing, duration, and shape. The system is capable of pilot injections to
lower NOx and engine noise, as well as post-injections that can be utilized to increase
exhaust temperatures in order to regenerate diesel particulate filters.

Multiple Injections and Injection Rate Shaping


Electronic fuel injection systems allow not only the possibility of using multiple injection events during
an engine cycle, they also allow for varying the rate of fuel injection during injection An example
summarizing the possibilities is illustrated in Figure 12 [Mahr 2002]. The upper curve shows the pressure
at the nozzle needle seat and the lower curve the corresponding needle lift with multiple injections.
Reasons for the various injection events include:

One or two pilot injection are carried out at low pressure to reduce noise and NOx emissions
To further control NOx emissions, varying the pressure when the nozzle is completely open can
be used for injection rate shaping. In this example the main injection event is boot shaped and the
length of the boot can be varied or the boot shape can be replaced with a triangular (ramp) or
rectangular shaped injection event. A coupled post injection under high pressure can be used to

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 14 of 33

reduce soot emissions.


A late post injection at moderate pressure allows exhaust gas temperature to be raised for the
regeneration of a diesel particulate filter or to provide hydrocarbons for NOx adsorber catalysts.

Figure 12. Multiple Injections and Injector Rate Shaping for Controlling Emissions and Combustion Noise

Air Management Improvements


Charge Air Cooling
Practically all modern automotive diesel engines are turbocharged. Turbochargers are devices
incorporating a turbine and a compressor on one rotor shaft. The exhaust gases expand in the turbine,
which drives the compressor placed in the intake stream to compress the intake air.

Turbochargers can be of fixed or variable geometry. Fixed geometry turbochargers are controlled with a
wastegate while variable geometry turbochargers are controlled by varying the turbocharger geometry.
The control is necessary to maintain the required intake manifold (boost) pressure at all operating
conditions. In wastegated turbochargers, the wastegate allows some of the exhaust gas to bypass the
turbine at high engine speeds and prevent excessive boost pressure. In variable geometry turbochargers,
the geometry, usually of the turbine, is varied not only to limit boost pressure at high engine speeds but
to increase boost pressure at low engine speeds and to control exhaust manifold pressure.

Most modern turbocharged engines use charge air cooling. The charge air coolers, called aftercoolers or
intercoolers, are heat exchangers that cool the intake air that was heated during compression in the
turbocharger. A schematic of a turbocharged and aftercooled diesel engine is shown in Figure 13 [Khair
1993].

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 15 of 33

Figure 13. Turbocharging and Aftercooling

Charge air cooling reduces the intake manifold temperature (IMT), thus increasing the density of the
charge air and therefore the engine specific power output. The benefits of charge air cooling include
improved fuel economy and emissions. The first generation of aftercooled engines in the late 1980s
utilized water-cooled intercoolers, as shown above. In the 1990s, the use of air-to-air coolers becomes
increasingly popular for both heavy- and medium-duty diesel engines. While engine coolant-cooled
intercoolers are more compact, they can increase the heat rejected to the engine coolant and consequently
the radiator. Air cooled intercoolers though larger, reject heat to the ambient air thereby reducing the
heat rejection requirements of the coolant and radiator. Air cooled intercoolers can also lower the charge
air to much lower temperatures than those that use engine coolant.

Another development in charge air cooling that has emerged is so called indirect cooling. In this
approach, instead of using ambient air to directly cool charge air in an air-to-air heat exchanger, a low
temperature liquid cooling circuit, separate from the engine coolant (Figure 14), is used. It allows charge
air to be cooled to temperatures comparable to those achieved with conventional air-to-air coolers. While
this approach increases the cost and complexity of the charge air cooling circuit, it does offer reduced
pressure losses in the charge air cooling circuit, faster transient response, more flexible vehicle front end
packaging and better control over charge air temperature [Pantow 2001][Kruger 2008].

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 16 of 33

Figure 14. Engine Cooling System Incorporating Indirect Charge Air Cooling Circuit
(Source: Behr GmbH)

Variable Geometry Turbocharging


Due to a nonlinear flow versus pressure ratio characteristic, a fixed geometry turbocharger is generally
unable to provide adequate air flow at low engine speeds. This inability to provide sufficient air flow at
low speeds results in reduced torque, reduced drivability, and smoke emissions. This problem is
traditionally solved by matching the turbine at medium engine speeds (i.e. using a smaller turbine) and
either (1) accepting the penalty of high boost pressure at high engine speed, or (2) using a wastegate to
reduce the turbine power at high engine speeds. In either case, boost pressure at low engine speeds is
usually lower than desirable.

To overcome these inherent limitations of the fixed geometry turbine, several variable geometry
concepts have been developed [Charlton 1998]. In the most common vaned nozzle approach, the turbine
nozzles are able to rotate along an axis parallel to the turbine rotor. The effective flow area of the nozzle
can be increased or decreased, as dictated by the engine needs. Variable geometry turbochargers may be
used to reduce emissions, notably transient smoke emissions, as well as a number of engine parameters,
such as low speed torque, without compromising high speed fuel consumption.

The major driver that made VGTs common on commercial heavy-duty diesel engines was the use of
high-pressure EGR to meet US EPA 2004 onroad heavy-duty emission standards [Dollmeyer 2007]. The
required EGR levels could not be easily achieved with fixed geometry turbochargers. However, the VGT
allows turbine inlet pressure to be raised to higher values than are possible with fixed geometry
turbochargers to ensure that there is a higher pressure in the exhaust manifold than the intake manifold to
drive EGR flow.

Although variable geometry compressors are possible to improve the operating range of the compressor,
in most cases a loss of speed variation and efficiency occurs that outweighs any potential benefits.

Two-Stage and Sequential Turbochargers


In order to overcome some of the limitations of single turbochargers, the use of two turbochargers
arranged in various arrangements is becoming more common. These two-turbocharger arrangements
offer one or more of the following advantages: higher boost pressure, higher airflow and/or faster
transient response. Two turbochargers can be arranged is series (sequential arrangement) or parallel

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 17 of 33

(two-stage turbocharging) arrangements (Figure 15). Within in these two general configurations,
numerous refinements are possible through the selection of turbocharger sizes and bypass arrangements.

When two turbochargers are connected in series, higher boost pressures are possible than with a single
turbocharger. When two turbochargers are connected in parallel, while there is no increase in the
maximum pressure ratio available, it is possible to maximize boost over a wider range of engine
operating conditions.

Figure 15. Three possible arrangements of systems incorporating two turbochargers

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 18 of 33

(a) Two-stage system; (b) Sequential system; (c) Modulated two-stage system
(Source: Holset)

Figure 15 shows three examples of two-turbocharger arrangements:

The two-stage arrangement employs two turbochargers working in series at all times.
The sequential arrangement uses a smaller turbocharger to achieve higher boost at low engine
speeds than would be possible with a single turbocharger arrangement. A larger turbocharger
connected in parallel is phased in at higher engine speeds to provide sufficient boost and airflow
over the entire engine operating regime.
A modulated two-stage system combines some of the benefits of the two-stage and sequential
arrangements. At low engine speeds both turbochargers ensure high boost pressure. At high
engine speeds, the smaller, high pressure (HP) turbocharger is bypassed to allow the larger low
pressure (LP) turbocharger to deliver the higher air flows required.

As an alternative to using two separate turbochargers in a sequential arrangement, a single sequential


turbocharger that has a compressor wheel comprising two impellers mounted on the same shaft in a
back-to-back arrangement configuration is available. This arrangement allows the compressor flow range
to be extended and the wheel diameter to be reduced relative to a single compressor [Arnold 2008]. A
similar approach that combines an axial and radial compressor on the same shaft can provide two-stage
turbocharging in a single compressor housing [Nikpour 2008].

Other Turbocharging Developments


Aluminum is the most common material used for manufacturing turbocharger compressor wheels. The
maximum pressure ratio available from aluminum compressor wheels is limited because of temperature
related material effects and the need to limit compressor failure due to low cycle fatigue. Alternative
materials such as titanium that allow much longer compressor lifetime have been available since the
1990s but their use has been limited to a few limited applications that justify their high cost with their
much better ability to withstand cyclical loads. With increasingly complex turbocharger arrangements
such as two-stage turbocharging being required to meet the emissions and performance demands of
many light-duty and heavy-duty diesel engines, single stage compressors with a titanium compressor
wheels offer a competitive alternative for some applications. Not only are the compressors more durable
but higher pressure ratios can be achieved than with single compressors that employ aluminum
compressor wheels.

Turbocompounding is a technology that has seen commercial use for many decades but whose
application appears set to grow significantly as focus shifts to improving fuel economy. In
turbocompounding, useful exhaust energy not needed for intake air compression is captured with an
exhaust turbine. The exhaust turbine can be mechanically connected to the crankshaft to provide
additional engine output or it can be connected to a generator that can provide power to the engines
electrical system. In many designs, the turbocompounding turbine is separate from the turbocharger [Vuk
2005][Vuk 2006].

Electrically assisted turbochargers include a motor/generator designed into the shaft of the turbocharger.
The electric motor can be used to accelerate the turbocharger much faster during a transient than would
be possible without it. This can provide a significant reduction in the time taken to reach the final boost
pressure and engine torque response is thus improved. The generator capability can allow the
turbocharger to be used for turbocompounding and avoid the need to have a separate turbocompounding
turbine [Hopmann 2002][Hopmann 2004][Algrain 2003].

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 19 of 33

Combustion Chamber Improvements


Piston Bowl Design. The objective in the combustion chamber design is to provide good mixing of fuel
and air before the start of combustion. Turbulence in the air motion in the combustion bowl (crater) in
the piston crown improves the mixing process. Proper design of the combustion bowl can enhance the air
swirl created by the intake port and increase turbulence.

Figure 16. Combustion Bowl Designs

There has been a general tendency to replace the traditional straight-sided bowl design with reentrant
bowls (Figure 16). The location of the bowl, which used to be eccentric with 2-valve per cylinder
engines, is now commonly located in the center of the piston crown as 4 valves per cylinder have
become more common. The position of the injector is central relative to the bowl, in both concentric and
eccentric bowl design. However, in eccentric bowls, the injector is positioned at an angle, while it is
vertical in concentric bowls.

Another feature of new combustion chambers is a lower compression ratio. In the past, compression
ratios tended to be high to offset degradation in cold starting performance resulting from measures such
as injection timing retard and charge air cooling used to control NOx emissions. This left the
compression ratio higher than the optimum that balances engine efficiency, emissions and friction losses.
Technologies that improved cold starting ability has allowed compression ratio to be reduced to values
that are more optimum.

Variable Compression Ratio.. Variable compression ratio (VCR) engines [Shaik 2007] have existed for
decades but have never been commercially available on production vehicles. The added complexity and
cost have been hard to justify. By varying compression ratio, it is possible to affect engine load
capability and fuel consumption. They are also considered as an option for combustion control in low
temperature combustion engines. One such prototype engine has been developed by Caterpillar for diesel
applications [Lawrence 2008][Duffy 2006][Gehrke 2007].

Oil Control. The introduction of PM emissions standards for diesel engines made lubricating oil
consumption control essential to directly reduce engine-out PM emissions. In some older engine designs,
unburned lubricating oil emissions exceeded the PM emission limits even in engines that were in good
repair. Unburned .lubricating oil can contribute to PM emissions through several possible pathways
including: the piston-ring-liner system, intake and exhaust valve stems, the turbocharger and vented
crankcase gases.

In the 1990s, oil control was one of several important measures taken to meet the US EPA onroad heavy-
duty diesel PM standards. The fitting of catalyzed aftertreatment devices and DPFs to diesel engines has
once again put pressure to further reduce oil consumption. In this case, the focus is not on reducing
engine out emissions but on minimizing the potential build-up of oil derived ash and poisons in DPFs
and other catalyzed aftertreatment devices.

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 20 of 33

In the case of some of the oil consumption pathways, namely the turbocharger and crankcase ventilation
gas, the goal is to eliminate oil loss completely. While the complete elimination of oil loss may not be
entirely possible, it can be made very low through judicious component selection and design.

In the case of other oil consumption pathways, namely the intake and exhaust valve stems and the piston-
ring-liner system, oil loss cannot be entirely eliminated. A certain flow of oil supply must be maintained
to lubricate valve stems, pistons and rings. The design objective for low oil consumption in this case is to
provide the minimum amount of oil that is practically possible for these components in order to ensure
their life time is acceptable and/or to design them to minimize their lubricant consumption needs.

Some lubricating oil properties can also affect oil consumption:

The volatility of the lighter oil fractions can have an impact on the evaporation of oil from the
piston-ring-liner system. This evaporated oil either burns in the combustion process or is expelled
out the exhaust valve.
Lubricating oil viscosity can impact the amount of oil that passes the piston ring pack into the
combustion chamber and the thickness of the oil layer on the cylinder liner. While higher
viscosities can decrease the amount of oil lost from the ring pack, it can increase the thickness of
the oil layer on the liner where it can evaporate when temperature conditions are appropriate.
While the amount of sulfated ash, phosphorus and sulfur (SAPS) in the lubricating oil does not
have a direct impact on lubricating oil consumption, it can impact the rate of ash accumulation in
and poisoning of exhaust aftertreatment devices. The trend is toward lower levels of these
componentsso called low SAPS oils.

Variable Valve Actuation


Variable valve actuation (VVA) allows control of valve timing and/or valve motion. Simpler approaches
select between two options for a valve timing event, e.g., the intake valve closing timing. More
sophisticated approaches allow a wide range of freedom in selecting not only the timing of valve events
but also valve lift and duration. The first commercial applications of VVA were reported on light-duty
gasoline engines [Borgmann 2001a]. Possible benefits include improved fuel economy, emissions, low
speed torque and transient response (reduced turbo lag).

Delaying or advancing intake valve timing allows for effective expansion ratios to be higher than
compression ratios. This allows improvements in engine efficiency without increasing peak temperatures
during the combustion process to exacerbate NOx formation. The resulting engine is often referred to as
a Miller cycle engine after an early developer of this approach.

In his patents, Miller referred to changes in intake valve closing timing to reduce TDC temperatures and
provide a higher expansion ratio than compression ratio in forced induction engines [Miller 1954][Miller
1956]. While he mentions both early and late intake valve closings, he seemed to prefer closing the intake
valve early while the cylinder volume was still increasing because additional expansion after intake
valve closing could further cool the intake charge. Modern engine design approaches referred to as using
the Miller cycle include both early [Codan 2007] and late intake valve closings [Cornell 2006].

Occasionally, engines with late intake valve closings are referred to as Atkinson cycle engines. Some
prefer to restrict reference to Atkinson cycle engines as those that are naturally aspirated and have a late
intake valve closing. However, the original patents by Atkinson do not refer to valve closing timing but
to an engine in which one engine cycle is completed in a single revolution of the crankshaft and with a
crankshaft mechanism that allowed a higher expansion ratio than compression ratio [Atkinson 1886]
[Atkinson 1887].

Exhaust Gas Recirculation

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 21 of 33

Exhaust gas recirculation (EGR) has been known for a long time and is a well-proven and effective
means of reducing NOx emissions from internal combustion engines [Ropke 1995]. It has been used on
both spark ignited and diesel engines. Most manufacturers of heavy-duty truck and bus engines in the
USA chose cooled EGR as the prime NOx control technology to meet the 2004 (2002) NOx+HC
standard of 2.5 g/bhp-hr and to certify 2007 engines to NOx levels as low as 0.85 g/bhp-hr. Cooled EGR
is also expected to be important for meeting US 2010 emissions even though most manufacturers will
also employ NOx aftertreatment to meet the 0.2 g/bhp-hr limit. In Europe, it is one of two options used
to meet Euro IV and V NOx limits in heavy-duty engines [Verbeek 2001]; urea-SCR aftertreatment is the
other technology commonly used to meet these EURO NOx limits.

A schematic of an EGR system is shown in Figure 17. A part of the exhaust gas stream is recycled and
mixed with the engine intake air. The system shown is a high pressure loop EGR system, meaning that
the recirculated gas is taken from upstream of the turbocharger and mixed with compressed intake air
downstream of the aftercooler. A low pressure loop EGR system, where the EGR stream is taken from
downstream of the turbo and mixed with low pressure air, is also used.

Figure 17. High Pressure Loop EGR

EGR systems can be implemented with either open or closed loop control systems. In open loop systems,
EGR valve position is set based on engine speed and load. While extensive calibration is needed to
determine what the correct valve positions should be, there is no feedback signal to determine if, over
time, the commanded valve opening is continuing to provide the correct amount of EGR. Closed loop
systems use a feedback signal, commonly from a sensor that determines the actual valve position or from
a flowmeter in the EGR line, to provide an error signal that allows corrections to be made to the signal
that drives the EGR valve so that any unexpected changes in EGR flow can be corrected. Closed loop
systems are critical for the most stringent emission standards where even small changes in EGR flow can
cause NOx emissions to exceed regulated limits.

In order to provide sufficient pressure difference across the EGR valve to ensure sufficient flow at all
engine operating conditions, several options are possible. One option is to supply EGR to the throat of a
venturi placed in the intake duct on the discharge side of the compressor [Baert 1999]. Venturi assisted
EGR was used by Cummins on diesel engines certified to the voluntary US TLEV standard [Charlton
1998]. It was later combined with a variable geometry turbocharger (VGT) in 2002 engines. Partially
closing the turbine nozzles in a VGT at some engine operating conditions can raise the exhaust manifold
pressure relative to intake manifold pressure well beyond what can be achieved with fixed geometry
turbines. Many US diesel engine manufacturers used VGTs in 2002 and later engines to achieve
sufficient EGR rates.

The dominant factor responsible for the NOx reduction effect of EGR is a dilution of the intake air with

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 22 of 33

inert gases, leading to a decrease of oxygen concentration. The increased heat absorbing capacity of CO2
and H2O (thermal effect) and the effect of CO2 dissociation (chemical effect) also play a less important
role [Hawley 1998]. The dilution effect lowers the oxygen concentration and causes the diffusion flame to
spread out over a larger area; for a given rate of fuel injection, the same amount of oxygen must be
consumed regardless of the in-cylinder oxygen concentration. This spreading of the flame entrains more
inert gas into the combustion zone and significantly lowers flame temperatures. This has a dramatic
effect on NOx. EGR rates as low as 10% can produce NOx reductions on the order of 40% [Baert 1999].
With conventional diesel combustion systems, the penalty is an increase in smoke and PM emissions. At
higher EGR rates (20-30%), increased HC and CO are also seen. EGR can also increase the amount of
soot entrained in the engine oil and make engine wear protection more challenging.

The increase in total PM emissions with EGR occurs due to increased emissions of solid carbon
particles. The SOF fraction of diesel particulates remains constant or decreases with increasing EGR
rates. EGR increases the mean particle size in diesel exhaust. The number of nuclei mode particles has
been reported to decrease and the number of accumulation mode particles to increase with increasing
EGR [Kreso 1998].

Until the late 1990s and early 2000s, diesel engines were capable of running at about 50% maximum
EGR rate without affecting the engine stability (about 20% for gasoline engines). Advances in
turbocharger and fuel injection system design since then allow engines to operate at even higher EGR
rates. Since EGR reduces the oxygen concentration in the intake air, the highest EGR rates can only be
used at lower engine loads. Engine speed/load maps are used to control EGR with increasing EGR rates
towards lower engine loads.

The full potential of EGR is realized in the cooled EGR technique. A heat exchanger is added to cool the
EGR before it is mixed with intake air. Cooling of the EGR decreases the charge air temperature
providing further NOx reduction. It also increases the inlet air charge and the oxygen-to-fuel ratio
relative to uncooled EGR. The increase in oxygen may increase the flame temperature and raise NOx,
but has a beneficial impact on PM. Cooled EGR will also reduce the pumping work, due to increased
charge density and better volumetric efficiency, providing increased specific fuel consumption and better
engine efficiency.

Similar to charge air cooling, EGR cooling may be carried out with engine coolant or air cooled heat
exchangers. The choice depends mainly on intake air temperature objectives. Air cooled EGR coolers
can be used in combination with engine coolant heat exchangers to reduce the necessary heat exchanger
surface area required for some EGR rate applications.

Low Temperature Combustion


Since the introduction of aftertreatment forcing exhaust emission standards, engine designers have
attempted to introduce changes to the combustion process in an attempt to lessen the reliance on
aftertreatment systems to meet these very stringent emission limits. Combined emissions of PM and NOx
from engines employing conventional diesel combustion are too high to meet US 2007 and US 2010
without expensive aftertreatment systems. While the ultimate objective would be to completely eliminate
the need for aftertreatment altogether, engine system development needs to still evolve before this can
occur on a commercial scale. However, developments is diesel combustion have reached the point where
some aftertreatment system components can be simplified or even eliminated.

In general, developments in these low temperature combustion (LTC) systems have focused on avoiding
in-cylinder conditions that produce significant amounts of NOx and soot (Figure 18) [Maier 2003]. The
coordinates in Figure 18 are local air-fuel equivalence ratio () and temperature. More common
coordinates are local fuel-air equivalence ratio ( or 1/) and temperature (a representation known as the
-T map).

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 23 of 33

Figure 18. Selected Low Temperature Combustion (LTC) Strategies in -T Coordinates


HCCI - Homogeneous charge compression ignition
DCCS - Dilution controlled combustion system
HPLI - Highly premixed late injection
HCLI - Homogeneous charge late injection

The LTC approaches typically involve increasing the amount of premixing of fuel, air and EGR before
ignition to avoid excursions into the high soot formation region and using EGR to keep peak combustion
temperatures out of the high NOx formation region. This is illustrated in Figure 19 with one of the LTC
approaches, premixed charge compression ignition (PCCI) combustion. Reducing NOx to 0.5 g/kg-fuel
(about 0.1 g/bhp-hr) requires high cooled EGR rates of over 50% and intake oxygen concentrations of
about 10-12% [Durrett 2008]. To achieve the premixing required to control soot, injection occurs at a lower
temperature to increase the ignition delay period. This extended ignition delay allows more time for
mixing of fuel, air and EGR and a leaner equivalence ratio before the start of ignition. Combustion can
then occur in the presence of higher concentrations of oxidative species and at temperatures below that
required for significant soot formation.

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 24 of 33

Figure 19. NOx And PM Emission Reductions Through PCCI Combustion

Another example of how the high soot and NOx formation regions can be avoided is lifted flame
combustion (Figure 20 and Figure 21). In this approach, the engine is designed to increase the lift-off
length and decrease the liquid penetration of the fuel spray when compared to a more conventional diesel
spray. This provides a larger jet surface area in the near nozzle region where air can be entrained into the
fuel jet and produce a leaner (though still fuel rich) mixture that enters the burning region of the spray
(compare B and B in Figure 20). At the end of the premixed burning phase, less soot and soot precursors
form as a result (compare C and C). By using higher amounts of EGR, the maximum combustion
temperatures can be controlled and NOx formation in the diffusion flame kept to a minimum (compare D
and D) [Stanton 2005][Frazier 2008][Gehrke 2008].

Figure 20. NOx And PM Control Through Lifted Flame Combustion

Figure 21. Spray Visualization in Lifted Flame Combustion

Many LTC approaches benefit greatly from fuels that are significantly different from commercial diesel
fuel. Over some parts of the load range, these approaches benefit greatly from high volatility and low

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 25 of 33

cetane number fuelssuch as gasolinewhile at other parts of the load range, commercial diesel fuel is
perfectly acceptable. Varying fuel properties in this way can improve emissions and expand the load
range over which these combustion approaches can be applied. By using two different fuels that can be
blended in various proportions, fuel properties can be tailored to the engine operating condition. Such
dual fuel approached are being studied [Gehrke 2009][Reitz 2009] and have yet to appear in commercial
applications.

Engines can be operated in any particular LTC combustion mode either full time or only in some parts of
the engine map. In the latter case, engines that use conventional diesel combustion at some operating
conditions and non-conventional combustionfor instance PCCIat others are sometimes referred to as
mixed-mode combustion engines.

Electronic Engine Control


The use of digital engine control for on-road heavy- and light-duty engines has grown exponentially
since the late 1980s. Digital technology has also entered the off-road and industrial engine markets.
Electronic control is a powerful tool to solve many traditional diesel engine control problems, such as
cold start, load response, governing, or transient smoke emission. As the scope of control broadened to
include emission control systems, fuel systems, and air handling systems, quite spectacular reductions of
all regulated diesel emissions have been realized.

Electronic engine control plays a vital role in the exhaust emission control from todays advanced diesel
engines. From the emission perspective, the goal of the engine control system is to provide the demanded
quantity of fuel, air, and EGR (if any) at the required time and in the required temperature and pressure
state. This control is executed over the engine lifetime, compensating for engine wear and deterioration.
Additionally, as required in an increasing number of applications, engine emission controls should be
supported by on-board diagnostic (OBD) systems, which would activate a malfunction indicator light on
the vehicles dashboard when an emission fault is detected.

An electronic control system for diesel engines includes a set of sensors, a microprocessor, and a set of
actuators. The sensors measure physical variables and pass the information in the form of electrical
signals to the controller. Examples include crank speed, boost pressure, intake manifold temperature and
pressure. The actuators perform mechanical actions as directed by signals from the microprocessor.
Examples of actuators are EGR valves or variable geometry turbochargers.

The collection of electronic hardware, including the microprocessor and a set of electronic sub-systems
for data storage, input/output handling, power supply, and other functions, is called an electronic control
unit (ECU). The ECU is usually mounted on a single board and enclosed in a protective housing. Many
functions can be incorporated into a single chip called a microcontroller.

Some of the important diesel engine control functions include [Charlton 1998]:

Fuel quantity and fuel timing control,


Boost pressure control, and
EGR control.

Fuel Quantity. Fuel quantity is controlled by a governor or a series of governors for different regimes
(cold start, low speed idle, high speed idle). There are two types of governors: variable speed (all-speed)
governor, and minimum-maximum speed governor.

In the variable speed governor, the engine speed is allowed to fall as load is applied at constant throttle
position. It is a form of simple proportional control in which the fueling is proportional to a speed
error. This type of governor is used in applications where load can change very rapidly, e.g. in
agricultural tractors or earthmoving machinery, or in applications where an approximately constant speed

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 26 of 33

should be maintained, e.g. cruise control. Variable speed governors make the engine more responsive to
rapid load changes and prevent stalling.

The min-max governor provides automatic control at low idle speed and regulates the engine at high
speed to prevent damage. At intermediate speeds the fueling is proportional to the pedal position, i.e., the
driver controls the engine torque output directly. This type of governor is used in automotive
applications, especially passenger car diesel engines.

With each type of control, the governor characteristic is stored by the ECU as a three-dimensional fuel
map versus engine speed and throttle position. The engine controller will first calculate the fuel rate from
the map. This demanded fuel rate is then compared with other computed fuel rates which are needed to
satisfy other engine conditions, for example the air flow limited fuel rate. The lowest fuel rate value is
selected and passed to the fueling system.

Fuel Injection Timing. Injection timing is one of the most important factors influencing combustion and
the resulting emissions. All important engine performance parameters, including specific fuel
consumption, emissions of NOx, PM, HC, and peak cylinder pressure, are a strong function of injection
timing.

Mechanical fuel systems have fairly limited injection timing capabilities. They usually require hardware
changes in order to modify injection schedules. The timing schedule of mechanical systems has very few
degrees of freedom, so that handling such anomalies as cold start advance or low coolant temperature
advance requires additional mechanical elements. Furthermore, cam-driven fuel systems typically exhibit
a relationship between injection timing and fuel injection rate, which can influence injection pressure and
duration.

In contrast, electronic control allows the freedom to command injection timing in response to engine
load, engine speed, or ambient conditions. Exhaust emissions can be optimized over an emissions test
cycle, which samples a wide range of engine speeds and loads, through implementing injection timings
that optimize emissions and fuel consumption over local regions of the operating spectrum.

Electronic injection control also improves the accuracy of injection timing. Some engines, e.g. the AUDI
HSDI engine [Bauder 1990], use an injector needle lift sensor to provide feedback. Injection timing
accuracy within 1 degree of crankshaft rotation was claimed.

Boost Pressure. In engines with a fixed geometry, mechanically controlled turbocharger, the wastegate
is controlled by boost pressure feedback to a pneumatic diaphragm actuator that operates the wastegate
valve through a linkage. This type of boost pressure control is only as accurate as the setup procedure
during assembly. There is no compensation for changes in the characteristics of the pneumatic actuator
or the wastegate valve. Therefore, errors in the boost pressure can occur.

In electronically controlled engines, the turbocharger can be either a fixed or variable geometry. Fixed
geometry turbochargers can be equipped with electronic wastegate control, where the boost can be
controlled as part of an integrated engine control strategy. The air pressure fed to the pneumatic
wastegate actuator is modulated by a pressure control valve. Variable geometry turbochargers usually
contain an electronic actuator that is used to vary the turbine geometry. Electronic control allows for true
closed-loop control of the boost pressure, compensating for manufacturing variability and changes
during engine lifetime as well as varying environmental conditions such as altitude.

Another advantage of electronic control is the possibility to include several other factors, besides boost
pressure, in the wastegate control algorithm. For example, the boost pressure can be higher during
transients than it is at steady state to improve acceleration.

EGR Control. The amount of EGR is electronically controlled in response to various parameters,
including the throttle position and ambient conditions. Early implementations used open-loop EGR.

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 27 of 33

Different degrees of EGR control are possible. In the most simple open-loop system, the EGR valve is
closed by the ECU during start-up and when the engine coolant is below prescribed temperature. A
further elaboration of the open-loop EGR control involves modulating the EGR valve according to a
valve position schedule stored as a look-up table in the ECU.

In a closed-loop control system, the amount of EGR would be controlled based on a feedback signal.
Some systems us a direct flow measurement of the EGR as the feedback signal. Others use the position
of the EGR valve.

Serial Data Links. In the first generation of electronic controllers the function of the ECU was limited
to control the engine. The ECU was wired to the sensors and actuators through a complicated wiring
harness. Gradually, more electronic controllers were added to the engine. These included controllers of
transmission, traction, antilock braking systems, emission aftertreatment, and more. Additional electronic
devices, such as on-board diagnostics or computerized service or management tools were introduced.
The emergence of all these electronic sub-controllers and devices created a need for a communication
network to exchange information between these systems.

Industry standards are being developed for such electronic communication networksalso known as
serial data linksthat would eventually allow manufacturers to design their electronic components to a
single interface standard. Development of standards for communication networks began in the mid-
1980s. Among the layers of a communication networkwhich include the physical layer, network layer,
and protocol layerthe physical layer (as well as diagnostics) achieved the highest level of
standardization. Serial data links standards include the CAN (control area network) standard developed
by Bosch and SAE standards J1709, J1587, J1922, and J1939 in the USA (the last one based on the
CAN). Examples of newer standards are the MOST (media oriented system transport) and the FlexRay
protocol.

Serial data links can utilize a pair of wires, a single wire, or fiber optics. The high speed version of CAN
for example, runs on a pair of wires where one wire is at ground (0 V) and the other carries a binary
signal (0 - 5 V) used to transmit digital information. The low speed CAN utilizes a single wire vs.
vehicle ground at 12 V potential. MOST can be implemented on wire, but is also available on fiber optic.

On-Board Diagnostics. The on-board diagnostics (OBD) system is designed to monitor the performance
of vehicles emission controls, such as catalytic converter, diesel particulate filter, evaporative emission
controls, engine misfire detection, etc., and to detect emission faults. Such faults are signalled to the
vehicle operator by lighting a malfunction indicator light (MIL). The list of countries requiring OBD
systems on vehicles continues to grow. In the USA, OBD systems are currently required on light and
heavy duty vehicles fueled by gasoline, diesel, or alternative fuels. In Europe, OBD requirements
became effective in 2000 (phase-in schedule until 2005) for light-duty vehicles and in 2005 for heavy-
duty vehicles.

The exact requirements of the OBD system are specified in pertinent regulations. Generally, there are
also differences between requirements for diesel and gasoline fueled vehicles. The following components
or systems are typically subject to OBD monitoring:

Emission aftertreatment devices (catalysts, particulate filters)


Engine misfire
Fuel system
EGR system
Air conditioning system (CFC refridgerants)
Evaporative system, secondary air system, oxygen sensor (gasoline engines)

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 28 of 33

Improving Fuel Economy

Vehicle fuel efficiency standards and GHG limits have created opportunities for a wide range of
technologies to be incorporated into engines and vehicles. The search for increased fuel efficiency is
focusing attention on at least three key areas:

powertrain efficiency,
vehicle technology and
operational parameters.

The impact of powertrain efficiency on fuel consumption is direct and is an obvious choice for
improving fuel efficiency. The approaches that have received significant attention are: (1) kinetic energy
recovery and (2) waste heat recovery. The potential of the latter approach is illustrated in Figure 22.

Figure 22. Waste Heat Streams in Heavy-Duty Truck Engine

Kinetic energy can be recovered with hybrid drivetrains that utilize regenerative braking. Waste heat
recovery from the engines exhaust gas can be recovered with technologies such as turbocompounding
and thermoelectric devices. So called bottoming cycles, such as the organic Rankine cycle, can be used to
recover heat from several of the waste heat streams in the diesel engine. Addition powertrain efficiency
improvements can be realized through lubricant and material technologies to reduce friction, start-stop
systems, auxiliary power units and other devices that eliminate the need to idle the engine unnecessarily
and reductions in parasitic losses from essential auxiliary devices such as pumps.

Advances in vehicle technology are another option for achieving fuel economy improvements. Improved
vehicle aerodynamics and reduced rolling friction are two obvious factors that affect fuel economy.
Other important factors include vehicle weight and power used by non-powertrain auxiliaries such as air-
conditioning.

Vehicle operational parameters such as driving patterns and route selection can also be used to gain
significant improvements in fuel economy. For example, it has been established that fuel economy
improvements of 5-20% can be realized in truck fleets by simply improving the driving habits of the

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 29 of 33

vehicle operators [Takada 2007].

Evolution of Engine Technology

Following the increased stringency of emission standards, diesel engines gradually adopted a number of
emission control techniques. The integration of the various technologies required several design choices
by engine manufacturers to accommodate emission and efficiency requirements while controlling
production cost increases in a competitive business environment. Figure 22 provides a compact
illustration of the technologies implemented by the US heavy-duty diesel engine industry during the
decades of the 1980s and 1990s [Flynn 2000].

Figure 23. Evolution of Diesel Engine Emission Control (1974-2004)

As shown in the chart, injection timing retard and charge air cooling (reduced intake manifold
temperature, IMT) were important measures for achieving NOx reductions in the late 1980s and early
1990s. To offset increased cold-start emissions of HC and white smoke, compression ratios were raised
to increase TDC temperatures after a cold start. To maintain engine efficiency, heat release rates needed
to be increased to shorten the burn duration and avoid having it extend too far into the expansion stroke.
This was achieved by increasing fuel injection pressures. Raising intake manifold pressures helped offset
increasing in particulates caused by the retarded injection timing. Details on the effects of these
parameters on engines from this era can be found in the literature [Uchida 1992]. Lower friction has always
been a goal of the engine designer as part of the continuous engine design improvement. These
technologies were partially implemented in the 1980s and by 1991 heavy-duty engines generally featured
retarded injection timing, low IMT, and high intake manifold pressure. Some of these engines also
adopted simple injection timing control algorithms that provided some injection timing flexibility to
allow fuel consumption optimization.

Combustion system research in the early to mid-1990s led to more reliance on swirl-supported intake
ports and combustion bowl designs. This technology helped medium heavy-duty engines (1.0 to 1.5
liter/cylinder) achieve the 1991/1994 emission standards without exhaust gas aftertreatment. However,
some higher speed medium-duty engines (e.g., Navistar 7.3 L, Caterpillar 3116 and Cummins B-series)
were equipped with an oxidation catalyst for particulate control to provide a margin of safety against
production variability and engine aging. Continued improvements in in-cylinder technology coupled

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 30 of 33

with advances in electronic controls allowed some of the same engines to shed their oxidation catalysts
and still meet the 1998 heavy-duty emission standards.

As indicated in Figure 23, the 2004 (2002) emission standards involved the widespread use of exhaust
gas recirculation on medium- and heavy-duty diesel engines. This technology had been used in light-duty
diesels for sometime since these engines rarely operate at full load and can use large amounts of EGR at
part load where A/F ratios are rather high (greater than 50:1). This direction is in agreement with Wall
[Wall 1997] where he predicted that to obtain NO emission below 4.0 g/bhp-hr and maintain good fuel
x
economy, EGR and improved fuel injection systems would be employed (see Figure 5). In order to
successfully employ EGR on these engines, a number of other technologies were required. This included
fuel injection systems capable of higher fuel injection pressures to offset some of the potential increases
in particulates and the widespread adoption of variable geometry turbochargers to ensure sufficient
exhaust manifold pressures were available to drive the required EGR flows.

Post-2004 emission standards required a wide scale deployment of aftertreatment technologies on heavy-
duty engines. Diesel particulate filters were introduced on all US 2007 engines, while urea-SCR
technology is the primary NOx aftertreatment method in most US 2010 engines. However, in parallel
with the development of aftertreatment technologies, further progress was achieved in in-cylinder
emission control through refinement of the key engine subsystems, including fuel injection, EGR, air
handling, sensors, and engine control. Indeed, it became possible to meet the 2010 NOx emission
requirements without NOx aftertreatmentan approach chosen by Navistar. The 2010 Navistar
engineswhich use PCCI combustionare expected to be certified at around 0.5 g/bhp-hr NOx, with
the difference from the emission standard of 0.2 g/bhp-hr to be covered by emission credits accumulated
by the manufacturer.

Emissions and fuel economy in 2010 engine prototypes are illustrated in Figure 24 [Stanton 2008]. While it
is technically possible to control NOx to levels below 0.2 g/bhp-hr without aftertreatment, such strategies
cause a deterioration in fuel economy and increased PM emissions. For this reason, most manufacturers
adopted urea-SCR aftertreatment in US 2010 engines. An improvement in thermal efficiency of low
temperature combustion would be necessary to eliminate the NOx aftertreatment in heavy-duty engines.

Figure 24. Emissions And Fuel Economy in US2010 Engine Prototypes


Note: BSFC and PM emission scales were deduced and were not included in the original source
publication.

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 31 of 33

It is worth noting that engines produced in the late 1980s and during the 1990s were excellent
performers. Emissions were gradually reduced, the characteristic black smoke of pre-1991 had all but
been eliminated, and drivability and engine response were improved. At the same time fuel consumption
had decreased, with peak thermal efficiency of about 44% reached around 2000, Figure 25 [Stanton 2008].

Figure 25. Thermal Efficiency in US Heavy-Duty Truck Engines

Unfortunately, the thermal efficiency deteriorated in 2002, due to the rather hastily launch of EGR
engines. Since that time, the efficiency has steadily increased. It is anticipated that efficiency levels of
42-43% will be reached in 2010 engines with urea-SCR technology [Aneja 2009][Nelson 2009]. The US
Department of Energy (DOE) has a 55% efficiency target in heavy-duty engines by 2015.

Other Emission Control Technologies

Our discussion so far has focused on emission control approaches that were adopted on new highway
diesel engines. More engine emission control technologies have been developed that, although not
always practical for the mainstream diesel engine applications, have been commercialized in certain
specific niche markets.
Water Addition. The introduction of water into the combustion chamber is an effective method of NOx
control. The mechanisms of NOx control by water are similar to those seen in EGR. Water acts as an
inert gas and also absorbs heat in the process of evaporation. Both effects cause a decrease in the
combustion temperature. Water can be introduced to the engine cylinder through one of the following
methods:

Direct injection
Fumigation into intake air
Emulsified fuel

All three types of systems have been commercialized, with direct injection and charge air fumigation
methods offered primarily for selected large marine engine applications. Emulsions were common at one
time for both stationary engines and centrally fueled vehicles fleets. An additional advantage of using the
emulsion method is the simultaneous reduction of PM emission due to increased mixing due to

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 32 of 33

evaporation of water which is incorporated in the fuel spray droplets. Water-fuel emulsions are one of
the few emission control methods which are capable of reducing both the NOx and PM emission.

The main disadvantage of water addition methods is their high water requirement. As a rule of thumb,
reduction of NOx by one percent requires one percent of water in the water-fuel system. In other words,
achieving a 50% NOx reduction requires running the engine using a 1:1 mix of water and diesel fuel.
Direct injection and air fumigation systems require dedicated hardware, separate from the fuel injection
system, for water delivery. Once the hardware is in place the engine can be operated both in the water-
fuel and in diesel-only mode. Emulsions, on the other hand, require an increased capacity of the fuel
injection system to deliver a high proportion of water in the fuel. If the injection system is modified, the
engine may have to be operated in the emulsion mode at all times. In a more practical approach,
emulsions can be used for fueling of existing engines without modifications. In this case, however, the
maximum amount of water has to be limited to about 20%, resulting in a rather limited NOx reduction
potential. Even at the reduced water content, engines have to be derated due to the fuel system capacity
limitations.

Regardless of the water delivery method, the system has to be designed to prevent direct contact of water
with the cylinder surface to eliminate the possibility of a negative impact on engine durability.

Membranes/Oxygen Enrichment. Enriching the intake air with either oxygen or nitrogen is made
possible with relatively low cost polymeric semi-permeable membranes [Poola 1999]. Oxygen enrichment
can reduce emissions of the incomplete combustion products and increase power [Wartinbee 1971]. To
offset increased NOx emissions with modest oxygen enrichment to 23-25% by volume, fuel injection
timing can be retarded and fuel flow to the engine increased [Stork 1998]. In the heavy-duty engines for the
1990s, (i.e., before EGR was needed to meet NOx limits) this approach was able to reduce particulate
emissions by 60%, NOx emissions by 15% and increase gross engine power by 18%.

In order to meet the needs of much stricter emission standards that use charge dilution with an inert gas,
this approach can also be configured to reduce the oxygen concentration in the intake air by providing
nitrogen enriched air. Potentially this could avoid some of the challenges of introducing EGR into the
engine.

Reformer Gas. On-board reforming of a hydrocarbon fuel such as diesel fuel can produce a gas
consisting mainly of hydrogen and carbon monoxide. Intake fumigation of this reformer gas has been
applied to premixed SI engines and can enhance the laminar flame speed of fuel air mixture to provide
some performance and emissions benefit. A similar approach for conventional diesel combustion would
require significantly higher quantities to avoid incomplete combustion in overlean regions and a
degradation in combustion efficiency. It has however been studied as an option for controlling diesel
combustion strategies that rely on premixing of air and fuel such as some LTC concepts [Hosseini 2008]. It
is also used for enhancing the performance of diesel exhaust aftertreatment devices.

Engine Materials and Coatings. A number of different engine materials and coating technologies can
enhance engine performance and durability.

The increased performance demand with engine downsizing has created the need to use different
materials for major engine components such as the cylinder head and block in some applications.
Compacted graphite iron (CGI) offers a much higher fatigue limit than conventional grey iron and
aluminum alloys and can provide increase strength without increasing engine size or weight. The
strength and stiffness of CGI can also improve the dimensional stability of the cylinder bore to reduce
piston slap, bore wear, oil consumption and blow-by [Dawson 2007].

At one time, ceramic coatings, applied in-cylinder, were considered an important potential technology to
significantly reduce heat losses from the combustion chamber of internal combustion engines and
increase their thermal efficiency. This approach has also been demonstrated to control PM emissions in

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010
Engine Design for Low Emissions [subscription] Page 33 of 33

older diesel engines [EPA 1997]. Arguably a more interesting application of this technology is as a means
to thermally manage engine component temperatures for component life extensionthis is why they
are used in aircraft engines. Due to the high levels of thermal cycling in piston engines for ground
vehicle applications, however, the durability of in-cylinder ceramic coatings remains a significant
challenge. Ceramic coatings are also considered as a means of insulating exhaust system components to
enhance exhaust heat recovery efforts [Kass 2009].

Metal based coatings are not subject to the durability challenges as those composed of ceramics and have
seen some commercial applications. They have been used to coat cylinder bore surfaces of aluminum
diesel engine blocks to avoid the need for cast-iron liners. Reduced engine friction and oil consumption
are possible with this approach [Ernst 2008]. Some metal based coatings have some insulating properties
and have been proposed for component thermal management applications [Marr 2009].

Diamond-like-carbon (DLC) coatings are a mixture of diamond and graphite that can be used to reduce
friction and wear in valve train components, piston pins, fuel injectors and other engine components.
This can result in reduced fuel consumption, lower oil consumption and help ensure that emissions
remain low over the life of the engine [Lawes 2007][Ghlin 2001].

mhtml:file://D:\Meus Documentos\Textos para ler\Engine Design for Low Emissions [... 17/3/2010

También podría gustarte