You are on page 1of 28

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/329091874

The Quest for Improving Efficiency in Internal Combustion Engines

Conference Paper · September 2018

CITATIONS READS

0 10,825

4 authors:

Jorge Martins Francisco P. Brito


University of Minho University of Minho
142 PUBLICATIONS   1,354 CITATIONS    83 PUBLICATIONS   1,166 CITATIONS   

SEE PROFILE SEE PROFILE

Tiago Costa Joaquim da Costa


University of Minho Universidade Nacional Timor Lorosa'e
5 PUBLICATIONS   18 CITATIONS    11 PUBLICATIONS   6 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Exhaust2Energy - Automotive Exhaust Heat Recovery with Thermal Control View project

Prototipagem Rápida & Projecto para o Fabrico View project

All content following this page was uploaded by Jorge Martins on 21 November 2018.

The user has requested enhancement of the downloaded file.


The Quest for Improving Efficiency in Internal Combustion
Engines
Jorge Martins, F.P Brito, Tiago Costa, Joaquim da Costa

ABSTRACT

Traditional Internal Combustion Engines (ICE), particularly Spark Ignition (SI)


engines, have a maximum thermal efficiency which does not usually exceed
the 35% mark. Typically, only a third of the energy supplied within the fuel
reaches the crankshaft and the other two thirds are, more or less, equally
divided between the exhaust heat (enthalpy of the burned gases) and through
the engine cooling. Therefore, in order to improve engine efficiency, one
should look at these two waste heat sources and try to reduce them as far as
possible.

By the second law of thermodynamics a thermal engine receives heat from a


hot reservoir and looses heat to a cold reservoir. The cold reservoir is the
atmosphere and the heat that should be lost is in the form of exhaust gas
enthalpy. However, the extent of heat lost within the exhaust gases is much
higher than the one required from a second law perspective. In fact, the
exhaust gases leave the engine at a temperature much higher than the
ambient temperature, so we developed a way to reduce this wasted energy by
the use of over-expansion. Other supplementary methods of recovering
energy from the hot exhaust gases is by the use of Organic Rankine Cycles
(ORC) or using ThermoElectric Generators (TEG), which are devices that can
convert directly a portion of the exhaust heat into electricity using the so-
called Seebeck effect.

Unlike the heat lost through the exhaust, the heat lost through cooling is not
enforced by any thermodynamic law, it just occurs as the materials would
exceed their temperature limit if not properly cooled down. There are two
ways of tackling this problem: producing near adiabatic surfaces inside the
engine; or absorbing this heat but transforming it back to work, in a process
herein called "internal regeneration".

Adiabatic engines were developed some decades ago without any advantage
due to the technical problems associated to high temperature and loss of
volumetric efficiency. Recently there was another strategy for the
"adiabaticity" of the engine internal walls: the use of a material with extremely
low specific heat capacity and extremely low heat conductivity in a process
called Temperature Swing. These materials are mostly hollow with very fine,
but strong, cell structures. The idea is that the surface would follow
instantaneously the temperature of the burning gases therefore avoiding heat
transfer.
The internal regeneration consists on injecting sprays of liquid water onto the
internal surfaces right after they were heated by the combusting gases. The
liquid water vaporizes, removing heat from the surfaces and transforms itself
into high pressure steam, which presses down onto the piston, producing
crankshaft work. This way, the cooling heat is partially recovered.

Other ways of enhancing engine efficiency are the improved ignition, the use
of stratified charge, the Homogeneous Charge Compression Ignition (HCCI)
concept and engine downsizing and boosting.

This paper assesses all the above recovery concepts, as well as some other
promising efficiency-oriented technologies.

INTRODUCTION

Spark Ignition (SI) engines have been around since the end of the 19th
century and have been progressively and steadily improved mainly in terms of
efficiency and exhaust emissions. Exhaust emissions was an issue that came
to life at the end of the 60's with the problematic air pollution in downtown Los
Angeles. Soon after, emission legislation was in place across the USA and
Europe, initially restricting the emission of CO and HC but quickly the NOx
were also controlled. The best solution would be to use stoichiometric
mixtures and a three-way catalyst that would reduce the referred pollutants in
almost 99%. However, the reduction of NOx production in an engine requires
measures that would considerably increase fuel consumption. In fact, the
production of NOx is associated to high engine efficiency:
- high compression ratio;
- supercharging;
- lean operation;
- high temperature combustion.

Therefore, the development for low NOx producing engines involved


measures that did increase fuel consumption, such as Exhaust Gas
Recirculation (EGR) (reintroduction of burned gases in the combustion
chamber), low compression ratios and retarded ignition. However, the engine
efficiency had been improved by other methods, namely by the reduction of
friction in various components (pistons, piston rings, valve system control,
electric water and oil pumps, etc.) and by improvements of the mixture
formation and combustion.

THERMODYNAMIC ANALYSIS

If you look at an engine from a thermodynamics point of view, from the


amount of energy that is injected as fuel into an engine, roughly 1/3 (at best)
is used to produce mechanical power and the rest is lost within the exhaust
gases and through engine cooling. However, the same thermodynamics tells
us that not all the thermal power produced during fuel burning can be
transformed into useful work. There is a limit to how much thermal efficiency
an engine may have, which is related to the temperature of the burning fuel
and, specifically for SI engines, is directly related to the compression ratio.
The higher the compression ratio and the higher the maximum combustion
temperature, the higher the engine efficiency. But engines (Otto cycle) have,
at most, only half of the limiting value for efficiency, this limit being typically
2/3 (67%). So, if we may convert 1/3 of the energy into mechanical power, we
have a potential of another 1/3 of power that we are losing and may be
recoverable, according to the laws of thermodynamics.

WAYS OF IMPROVING EFFICIENCY

So, where can we get these 1/3 of energy that is lost? In fact we may recover
it from the losses or, even better, avoiding the losses in the exhaust and of the
cooling in the first place. As said before, part of the exhaust enthalpy cannot
be recovered as it is mandatory by the second law of thermodynamics. But
there is no thermodynamic need to waste the cooling energy. So why do we
do it? Just because without cooling the materials would increase enormously
their temperature and would seize. There is also another problem with SI
engines which is knock. Knock is a destructive combustion that takes place
when the "end gas" is hot enough to start auto-ignition. Higher wall
temperatures would trigger knock much earlier, forcing lower compression
ratios, therefore reducing the efficiency.

But we have already two ways of improving efficiency: reduction of exhaust


gas enthalpy and reduction of cooling losses. Let us see how we can tackle
each of these strategies.

REDUCTION OF WASTED EXHAUST ENTHALPY

When the exhaust valve opens the gases within the cylinder are still at high
pressure and high temperature. This is the reason why the exhaust gases are
very hot (they can surpass 1000ºC) and they produce a lot of noise (sonic
blow out when the exhaust valve opens). In fact, these high pressure gases
could actually expand over the piston for a much longer piston stroke until
they would reach atmospheric pressure if there would be a technical solution
to do it. This is the principle of the Atkinson cycle. In this cycle the expansion
is taken all the way to atmospheric pressure, leading to the maximum amount
of mechanical energy recovered from the full expansion. However, the
exhaust gases, even at atmospheric pressure, would still be hot. Therefore it
would be possible, in principle, to use that enthalpy and produce some
mechanical power using a thermal cycle, such as an ORC (Organic Rankine
Cycle). With this strategy, a theoretical SI engine with a compression ratio of
12:1 could increase its efficiency from the 63% of the Otto cycle (Fig. 1) to the
73% of the Atkinson cycle [1] and still improve to 87% using the enthalpy of
the exhaust gases. Even with the theoretic Atkinson cycle the gases would
exit at 830ºC. This enthalpy of the gases could then be used in an ORC and
produce extra work, leading to the referred total efficiency of 87% with the
help of an ORC, using the expression:
3
𝑊"#$,&'( = 𝑚 ∙ 𝑐- ∙ 𝑇 − 𝑇0 − 𝑇0 ∙ 𝑚 ∙ 𝑐- ∙ 𝑙𝑛
34
Fig. 1: Efficiency of the various cycles including Atkinson cycle (Adapted from [1])

We can argue that we could use the same strategy for the Otto cycle: as the
gases would exit the exhaust at much higher temperature, there would be a
higher potential for heat recovery and transformation into work with an ORC.
Doing the theoretical cycle calculations for the Otto cycle, the exhaust valve
would open when the gases were at 1592ºC and 6.25bar and they would
expand (irreversibly onto the atmosphere, therefore losing exergy), which
would result in 1 bar and 1144ºC. This would give a total efficiency of 85%,
therefore inferior to that when using the Atkinson cycle plus the ORC. The
difference (between 87% and 85%) is caused by the irreversibility of the
expansion over the atmospheric pressure.

So, the referred 87% for the Atkinson cycle plus ORC is much higher than the
maximum indicated efficiency of a SI engine. It should be noted that indicated
efficiency specifies the ratio of energy that can be "harvested" from the
burning of the fuel to the expansion of the burned gases over an expanded
boundary (piston head). As said before, the exhaust gases, although at
atmospheric pressure, are still at a high temperature level and it is possible to
harvest some of its enthalpy and convert it into useful work. This conversion is
limited by the second law of thermodynamics and uses mechanical heat
engines (such as ORC) or devices that convert directly heat into work
(electricity), the so-called ThermoElectric Generators (TEGs).

To enable the over-expanded cycle (sometimes referred to as Miller cycle, as


in Fig. 1) to work properly over different loads, the engine needs Variable
Valve Timing (VVT) and Variable Compression Ratio (VCR). VVT is now a
common technology used by different manufactures such as BMW (the first
with the Valvetronic), Nissan (VVEL) and Toyota (Valvematic). However, VCR
is a more complicated technology to be achieved. Saab [2] developed such a
system during the last decade of the 20th century, but it never went into
production. Also, if we do a find in "Google Patents" there are over 100
patents on that subject, but so far, only one car brand is scheduled to release
a commercial VCR engine. It is Nissan’s VC-Turbo engine, which is
scheduled to equip the upcoming 2019 Nissan Altima model.
OVEREXPANDED CYCLE

During part-load operation of a conventional S.I. engine the typical reduction


of intake pressure by the throttle reduces the indicated work, as the engine
works as a vacuum pump increasing the pumping mean effective pressure
(pmep) and reducing indicated mean effective pressure (imep). Also, it
reduces the pressure at the end of the compression stroke. VVT and VCR are
used to minimize these drawbacks. To avoid the throttle plate and its
shortcomings (engine working as a vacuum pump) the intake valve should do
the control of the charge by opening for a specified time. There are two ways
of controlling the charge by the intake valve opening: the EIVC (early intake
valve closure), where the valve opens only during the required time for the
right amount of air (fuel mixture) to enter the cylinder; and the LIVC (late
intake valve closure) where the valve remains open long after BDC (during
the up-stroke) letting some of the air already inside the cylinder to escape
back to the intake manifold.

There is a third way of producing the over-expansion, using a native over-


expansion engine. This type of engine [3, 4, 5] has different stroke lengths,
with the expansion much longer than the intake one (Fig. 2). The engine
depicted in [3, 5] has the added benefit of having the connecting rod almost
vertical during expansion (Fig. 2), further reducing mechanical losses.

Fig. 2: Native over-expanded engine (UMotor, from [5])

In relation to VCR, the first vehicle to be equipped with an engine with such a
system seems that will be the VC-Turbo 4-cylinder 2.5 L engine (Fig. 3) of the
2019 Nissan Altima [6].
.

Fig. 3: VC-Turbo, a VCR engine from Nissan (from [7])


FEV (and other companies including Porsche) has been developing a
connecting rod (Fig. 4 - based on a patent from Hilite [8]) with the potential of
altering its length, therefore varying the compression ratio of the engine. Its
length is varied by an eccentric bearing on the piston crankpin. It rotates with
the help of pressurized oil and two piston that maintain the angle (length) as
required.

Fig. 4: Connecting rod developed by FEV, enabling VCR (from [9])

As said before, Saab [2] developed a VCR system that received several
awards in 2000, and it was capable of changing the CR from 8:1 to 14:1 (Fig.
5). However, this system was not intended for use in over-expanded cycled
engines.

Fig. 5: Saab Variable Compression (SVC) engine (from [10])

THERMOELECTRIC GENERATORS

The exhaust gases have a high temperature level which makes them
especially suitable for recovery on a second law of thermodynamics
perspective [11, 12]. One way of recovering this enthalpy is through the direct
conversion of heat into electricity using thermoelectric generators (TEGs).
These devices operate under the Seebeck effect, where a temperature
difference across the junctions between a pair of dissimilar materials
originates a voltage (see Fig. 6a). In reality this is the same principle under
which thermocouples operate but unlike these, it is intended to generate
electric power.
Several couples of P and N type semiconductor materials are arranged
electrically in series and thermally in parallel constituting thermoelectric
modules (Fig. 6b) in order to achieve sufficiently high voltages and powers
[12].


(a) (b)
Fig. 6 – (a) Outline of the operation of a thermoelectric pair [13]; (b) outline of a conventional
thermoelectric module [14]

The enormous advantage of TEGs is their lack of moving parts, which makes
them virtually maintenance-free if correctly installed. They are also easily
scalable (modularity) in terms of output power. The main disadvantages of
currently available TEGs is their still low efficiency (around 5%), high cost per
unit power and sensitivity to thermal level (they are temperature limited and
their output is very sensible to the thermal level). New, more abundant
materials exploring phenomena such as quantum confinement and
technologies such as nano-structuring, as well as highly automated and
upscaled manufacturing processes promise to reduce cost and increase
efficiency. Materials such as magnesium/manganese silicides and
skutterudites are good candidates for increasing the affordability, efficiency
and temperature limit of thermoelectric modules above currently used bismuth
telluride and lead telluride, which are costly and toxic materials [12].

Some TEG prototypes have been tested by companies such as BMW [15],
Ford [16], or Honda [17] but so far no commercial application has been
released in any vehicle. One of the problems seems to be the relatively low
thermal power that these systems may absorb under the conditions of highly
variable real driving cycles. This variability makes it very difficult to optimize
heat exchangers under these conditions [11]. The authors have been
exploring ways to thermally optimize these systems by using a phase change
interface (a heat-pipe/thermosiphon based solution) which downgrades the
highly variable exhaust temperature down to the ideal operating temperature
[18, 19]. This temperature control is achieved by pre-regulating pressure of
the phase change interface with a non-condensable gas. Then, the boiling
temperature of the phase change fluid (which will be the heat transfer
temperature) will be a function of this pressure. The highly variable
temperature of the exhaust gases will then be downgraded to a nearly
constant temperature since all the thermal energy of the exhaust will be
transmitted to the TEGs through the vapour.
Fig. 7 displays some predictions of the thermal and electric powers of a TEG
system equipped with the thermal control philosophy described [19]. It is
worth noting that these are results concerning a light vehicle (1.6 Litre
atmospheric engine). Several interesting observations may be done. On one
hand, it may be seen that most of the available exhaust heat (in red) is
effectively absorbed by the system (evaporator power, in green). This high
heat exchanger effectiveness is only possible because there is a passive
protection against overheating, so the heat exchanger can have as low a
thermal resistance as desired. On the other hand, the heat which is effectively
channelled to the thermoelectric modules is condenser power, in light blue.
Note that it follows closely evaporator power but it has a ceiling which
depends on the size of the system. The less or more available thermal power
will activate a lower or a higher fraction of the system (because the system
will be filled with a lower or higher fraction of vapour/non-condensable gas).
So it can be seen that this system is in full power often during the driving
cycle. It can be seen that the resulting electric power production is
proportional to the condenser power. If more modules would be available, a
higher peak power ceiling would exist.
The advantage of a system like this one is that both thermal dilution under low
loads and overheating at high loads can be avoided. Therefore the modularity
of TEGs can be used to size the system to the desired degree without the fear
of thermal dilution, that is, all the active modules will be operating at top
efficiency conditions, since the available heat will be distributed to a
proportional number of modules, not to all modules all the time.
1000

Electric power [W]


500
Thermal Power (kW)

60 0
50
40
30
20
10
0
0 200 400 600 800 1000
Time (s)
Pevap Pavail Pcond Pe
Fig. 7 - Generator thermal and electric performance during a custom highway driving cycle -
available (Pavail) exhaust power, evaporator (Pevap) and condenser (Pcond) thermal powers
and electrical (Pe) power generated [14]

The savings (in terms of fuel energy) achieved with these systems are much
higher than the power they produce, since for each watt of electric power that
the engine alternator is no longer required to produce corresponds to an even
greater mechanical power that is saved (efficiency of the electricity production
on the alternator) and an even higher saving of fuel power (engine efficiency).
For instance, the average efficiency of the thermoelectric modules operating
under the custom highway cycle of Fig. 7 will be around 5%, which is close to
their maximum efficiency. The system efficiency will be slightly lower (around
4%) because not all the exhaust heat is effectively used by the system.
Nonetheless, the fuel and emissions savings obtained with the reduction of
alternator operation will actually be 6%.

ORGANIC RANKINE CYCLES (ORC)

The exhaust gas enthalpy may be partially recovered through the use of
bottoming thermodynamic cycles such as the Rankine cycle. This cycle uses
the exhaust heat to produce steam, which then expands in a turbine
producing mechanical work and ultimately electricity through an alternator.
Conventionally it is in the form of a closed cycle requiring an additional
condenser and pump to complete the cycle [20]. Water may be used as a
working fluid, as done by BMW prototype “Turbosteamer” (Fig. 8) [21], but the
high pressures involved and the need for substantial superheating of the
steam to avoid condensation at the turbine expansion makes the use of
organic fluids often more attractive. Many of these organic fluids used as
working fluids in bottoming cycles (ORCs) require little or no superheating
prior to expansion and they allow for lower operating pressures and lower
heat source temperatures than water.



Fig. 8 – BMW’s Turbosteamer Rankine Cycle exhaust heat recovery project [21]

Compared to TEGs, ORCs have substantially higher efficiency potential (often


exceeding 12%) and they typically have a higher power density. However,
they are mechanically complex and unlike TEGs, require maintenance, which
might be a critical parameter for commercial exploration. Nonetheless, it
seems that ORC systems are closer to the market than TEGs. An example of
a system being developed by an OEM is the one by BorgWarner (Fig. 9).
Fig. 9 – Outline of BorgWarner’s Organic Rankine Cycle Exhaust Heat Recovery system
prototype [22]

COOLING LOSSES

During combustion there is heat transferred from the hot burning gases to the
walls of the combustion chamber including the piston crown. As most engines
use aluminium for the piston and cylinder head, there is a limit for the
operating temperature of these parts, as solid aluminium cannot work at more
than 500 - 600ºC, even less for mechanically stressed parts. Therefore these
parts should be cooled down to more manageable temperatures, usually 150 -
180ºC. However, some parts inside the cylinder show higher temperatures,
such as the exhaust valves and the spark plugs, as these parts do not rely on
aluminium. The engine head and cylinder liners are cooled with the help of
water based coolants streaming around these parts and conveying that heat
to the outside of the engine, through a heat exchanger device named
"radiator". The heat gained by the piston is lost through the heat transfer to
the cylinder liners and through the lubricating oil that has the extended role of
cooling some of the internal engine parts (the underside of the pistons are
cooled by oil jets and the crankshaft and camshaft bearings are cooled by
advection). Then, this oil is cooled directly in an air-cooled or water-cooled
heat exchanger. Exhaust valves are cooled through the valve seat (and in a
smaller way through the valve guide) and intake valves generally do not
require cooling as the air-fuel mixture passing across them cools them to
acceptable temperatures.

As stated before, the cooling losses can exceed the 1/3 of the energy within
the injected fuel, so it is a major part of the engine energy balance, of the
same level as the mechanical work. So, what happens if we could reduce or
eliminate these cooling losses? Looking at the pressure trace of a firing
engine (Fig. 10), if these cooling losses (area A) do not occur, the area of the
cycle will increase (more mechanical power transferred to the piston) and
more enthalpy will be lost within the exhaust gases, as when the exhaust
valve opens (Fig. 10) the pressure and temperature of these gases will be
much higher.

Fig. 10: Pressure trace of a SI engine (the area A denotes heat transfer losses) (adapted from
[23])

So, if the heat transfer losses are minimized or eliminated, part of the energy
saved will be lost again within the exhaust gases. But again, this is another
benefit for the over-expanded concept. If it is possible to avoid most of the
heating losses, in a conventional engine the benefits of the "adiabaticity" are
mainly lost through the exhaust gases. Therefore the use of more extreme
over-expansion would be the best way to promote efficiency in these
"adiabatic" engines.

ADIABATIC ENGINES AND TEMPERATURE SWING

During the 80's of the 20th century some researchers developed "adiabatic"
engines [24], diesel engines that were internally insulated, usually using
ceramic coatings. Obviously there would be no option of using adiabatic SI
engines, as the extremely high surface temperature of the internal
components would trigger knock. But even diesel engines did not perform well
under these so-called adiabatic conditions. There would be mainly three
problems. The engines required a very long period to get to steady-state
conditions, the heat transfer coefficient from the gases to the walls actually
increased (by 300% - [25]) and the volumetric efficiency was very low. As the
inner temperatures of the engines were extremely high, the air entering the
engine would heat up very quickly and loose density, therefore leading to low
volumetric efficiency and a significant decrease in maximum power and
torque. There were other problems involving lubricants for these ceramic
materials working at very high temperature [26] and the increased production
of NOx [27]. In the overall there was no gain in terms of engine overall
efficiency as the heat that could be saved from the cooling system would
show up in the exhaust [28] and only turbo-compounding was able to recover
some of the lost energy.

The adiabaticity of engines has lately been based on a different approach.


Instead of using a dense ceramic internal coating for the combustion chamber
and piston, the material to be deposited in those locations should have
extremely low conductivity and extremely low heat capacity. The purpose is
for the wall surface to be able to follow the gas temperatures during the
various phases of the engine operation. This concept is known as
"temperature swing" [29]. It implies a specific coating that, due to its extremely
low conductivity and heat capacity, creates a huge temperature variation on
the combustion wall surface (temperature swing). As its heat capacity is
extremely low, a very small amount of heat is enough to increase its surface
temperature to values similar to those of the burning gases. The effect of the
extremely low heat conductivity does not allow that the high temperature is
transmitted along the wall depth. However, the coating should be thin enough
for allowing overall heat transfer enabling the average surface temperature to
be within low levels, not at the extremely high level of the adiabatic engines.

The temperature swing (DT) is, in general, related (proportional) to material


properties in terms of density (r), specific heat (cP) capacity and thermal
conductivity (k) through a relation proposed by Assanis [30]:

1
ΔT ~
ρ × cP × k

The temperature swing is expressed in terms of temperature differential, and


the higher value the better for this kind of application. Inconel achieves a fairly
low value (25K), zirconia may achieve values around 100K, whereas a
material developed by Toyota (Thermo Swing Wall Insulation Technology -
TSWIN [31]) can lead to values around 200K. However, interesting values
should be in the region of 1000K (Fig. 11). This means that the surface of the
wall (combustion chamber or piston crown) fluctuates 1000K within one
engine cycle. The higher this differential the better is the "following" of the
temperature of the wall in respect to the gas temperature. Andruskiewicz [32]
predicted that the temperature swing will prevent approximately 1/3 of the
cooling heat to be avoided while improving volumetric efficiency (as the walls
cool down during intake).

The volumetric heat capacity (r cP) is a function of the density of the material,
so it is desirable to have an extremely low density, this being the more
important property of the material. This low density usually is created by the
introduction of voids (holes) in a low heat capacity and low thermal
conductivity material. Air is an exceptional material to be used, but it is not
solid. Air alone should produce a temperature swing above 1500K. Therefore
the objective is to create a solid barrier with as much air and less solid
material as possible. Air has a volumetric heat capacity in the region of one
thousandth lower than most solids and is one hundredth lower in terms of
conductivity, so its temperature swing is extremely high.
The material (TSWIN [31]) developed by Toyota allowed the surface wall
temperature to follow the gas temperature (Fig. 12) in a significant way. In that
engine they noticed that most of the "heat not lost" to the walls emerged in the
exhaust gas as enthalpy, so they tried to "harvest" it using a more developed
turbo-charger [31]. The diesel engine (2.8L 1GD-FTV) using his technology
showed a brake efficiency of 44%, a very high value for such an engine,
corresponding to a 15% increase over the base engine.

Fig. 11: Thermal conductivity vs volumetric heat capacity for various materials (from [28])

Fig. 12: Wall temperature during one engine cycle using Toyota SiRPA coating (from [33])

shows the result of the temperature swing concept on the surface


Fig. 13
temperature of a combustion chamber within one cycle. The surface
temperature drops to 800K during intake and then hikes to 1800K during
combustion.
Fig. 13: Temperature swing on the surface of the combustion chamber (from [28])

INTERNAL REGENERATION

Another possibility to reduce or eliminate the cooling losses could be the


injection of water right after the combustion. There would be heat transfer to
the various walls (combustion chamber, exhaust valve, piston) but this heat
would be used to vaporize the injected water over those surfaces. This would
eliminate the need for external cooling and the heat "removed" from the walls
would produce high pressure steam that would "push down" onto the piston,
creating more mechanical work, therefore increasing engine efficiency [4].
But, like for adiabatic engines, the most of this recovered energy would be
thrown out within the exhaust gases. However, the use of the over-extended
cycle would recover a considerable amount of this energy, greatly improving
engine efficiency. This is known as "internal regeneration", as the cooling
energy is transformed into useful work within the engine cycle.

However, there is a big problem with that approach. If we require that the
engine becomes adiabatic (no external cooling), the ratio water/fuel (in mass)
should be around 7:1. This is unrealistic for vehicles, unless we develop an
effective system to recover water from the exhaust gases. Otherwise it would
be silly to put 50L of petrol and 350L of water in a car... However, the process
is realistic for stationary engines, as the water can come directly from the tap.

It is easy to calculate the referred 7:1 water/fuel ratio. At 10 bar the latent heat
of water is 2 MJ/kgwater whereas the lower heating value (LHV) for petrol is
44 MJ/kg. As the heat transferred to the walls is a typically a third of this value
(14.5 MJ/kgpetrol) the water/fuel ratio should be 14.5 / 2 @ 7.

A recent paper [34] calculated the required water mass flowrate required to
render an engine adiabatic using "internal regeneration". It was proven that a
7:1 water to fuel ratio would be enough for eliminating the need for external
cooling and that the time available for water vaporization was enough for
engine speeds up to 6000 rpm. The combustion chamber walls (including
piston crown and exhaust valve face) were of the conventional materials,
aluminium and steel. The average surface temperature was 500K and it would
vary from 490 to 510K during one engine cycle (Fig. 14).

Fig. 14: Wall temperature as a function of crank angle and wall depth (WOT, 3000rpm)

Comparing Fig. 13 to Fig. 14 one can notice that the engine depicted on Fig. 14 is
adiabatic because the temperature trace is horizontal with the wall depth,
while on Fig. 13 the temperature fall from 1200K to 400K across the wall (from
the gas surface to the coolant surface). One of the problems is that while on
the temperature swing engine the mean wall temperature is 400K (127ºC) on
the "internal regeneration" engine the wall are at an average of 500K (227ºC).
This high temperature is required to enable the water vaporization at
appropriate pressure levels. The water boils at 500K at 25bar and boils at
400K at 2.5bar. As the exhaust valve of an over-expanded engine opens
when the gases are at more than 2bar, it would not make sense to try to inject
the water during engine expansion, as the water would not boil. If the walls
are at a much higher temperature, the pressure for water boiling required for
the internal regeneration is also much higher, so it is possible to implement
the internal regeneration.

SPARK IGNITION ENGINE POTENTIAL IMPROVEMENTS

The traditional S I Engine has a homogeneous and stoichiometric fuel-air


mixture and the combustion starts with the discharge of a high voltage spark
somewhere in the combustion chamber. This engine has some problems:
a) this combustion is slow, therefore incapable of producing the isochoric
combustion of the theoretical cycles;
b) the stoichiometric mixture is not the best for power (it should be rich)
neither for efficiency (it should be lean);
c) the flame front goes right to the walls, leading to wall quench (reducing
combustion efficiency - some of the fuel will not burn) and high heat
transfer (reducing engine efficiency);
d) the stoichiometric engine produces high levels of HC, CO and NOx,
although they can be significantly reduced within the 3-way catalyst;
e) the ratio of mechanical losses to the work produced is significant;
f) knock is a limiting factor for S I engines.

There are some improvements that can be made to this engine:


g) new methods for ignition (including HCCI) produce much faster burning
rates, therefore improving efficiency, mainly at high engine speeds;
h) the use of lean and extra lean mixtures produces higher efficiency and
lower levels of all pollutants, mainly NOx;
i) if only the centre of the cylinder is filled with a combustible mixture,
there would be no wall quench and heat transfer to the walls would be
greatly reduced;
j) reducing the engine capacity while increasing the intake pressure (by
turbo-charging or super-charging, called "downsizing") to produce
similar torque and power levels will reduce the friction losses and result
in a smaller and lighter engine.

The following sections introduce some of these possible improvements.

STRATIFIED CHARGE ENGINES

One of the reasons for the high efficiency of the diesel engines at low loads is
its highly heterogenic combustion. The diesel engine has no air control, so
under low load the cylinders are filled with air but there is only a small amount
of fuel that is injected at the centre of the combustion chamber. Therefore the
combustion is limited to the volume of the fuel spray, never touching the
combustion walls, with the exception of the piston bowl. This is the reason
diesel engines are slow to warm up and during winter the cabin heating is
limited.

Petrol SI engines are traditionally homogeneous burning engines. However,


there is the possibility of using the concept of "stratification", leading to the
advantages of the low heat loss of the diesel engines. The idea is to create a
heterogeneous mixture of fuel and air, where there is a stoichiometric mixture
near the centre of the combustion chamber near the spark plug and the rest of
the cylinder is filled with air or with a very lean mixture. Obviously this only
happens at low loads, because at higher loads it is necessary to use all the air
to burn as much fuel as possible. At WOT all the cylinder volume should be
filled with stoichiometric mixture to produce the highest torque and power.

The idea of the stratified charge engine was initially introduced by Brayton
(inventor of the Brayton or gas turbine cycle) but the concept was developed
by sir Harry Ricardo [35, 36], based on a Diesel Comet chamber (Fig. 15).
During the 70s of the 20th century Honda developed and produced stratified
charge engines, the CVCC [37, 38] initially on the Honda Civic and produced
up to 1987. This engine ( Fig. 16) had a small chamber where the spark plug
would ignite a rich fuel-air mixture produced by a carburettor connected to a
small intake valve leading to the small chamber. This small chamber was
connected to the major combustion chamber by a small passage. The rich
mixture would ignite in the small chamber and is projected into the major
combustion chamber while burning, creating strong turbulence and burning
efficiently the lean mixture in the main chamber.
Fig. 15: Ricardo stratified charge engine

Fig. 16: Honda CVCC engine (from [37])

As said before, the improvements of the stratified charge engine come with
the benefits of lean mixtures and reduced heat transfer to the walls. Lean
mixtures are beneficial in thermodynamic terms. As the mixture becomes
leaner, there is more air for the same amount of fuel, so the same heat (from
combustion) is supplied to a higher mass, therefore resulting in a lower
temperature. As the specific heat capacity of the air increases with
temperature, a lower temperature involves lower heat capacity for the gases,
which is beneficial for efficiency. Also, the increase in temperature rises the
air adiabatic coefficient. A lower adiabatic coefficient will produce more work
during expansion.

However, the stratified charge mode only works at partial loads. For full load
(WOT) these engines should revert to homogeneous, stoichiometric
conditions enabling the maximum torque and power production.

A modern approach to the old stratified charge concept is the Mahle Turbulent
Jet Ignition (TJI) [39, 40] (Fig. 17). This system, believed to be utilized by
Ferrari in their F1 engines, uses a very small chamber with a spark plug and
with near stoichiometric mixture. This diminutive chamber (only 3% of the total
fuel) is linked to the main combustion chamber by tiny holes. When the spark
takes place, the mixture starts to burn and is violently projected (called
"plasma jets") into the main chamber in burning high velocity jets. These
turbulent jets ignite the mixture in the main chamber, leading to a very high
speed combustion. Additionally, these high speed and turbulent jets are able
to ignite a very lean or even an extra-lean mixture (Fig. 18). In order to further
improve combustion efficiency and overall efficiency, the mixture in the main
chamber may also be stratified. It is believed that this type of ignition is the
basis of the very high efficiency of Ferrari (and Mercedes) F1 engines. The
Mahle TJI system is extremely small and was manufactured in the volume
occupied by the spark plug, enabling it to be retrofitted in older engines.
Fig. 17: Mahle TJI patent drawings (from [39])

Fig. 18: Mahle TJI lean operation (from [39])

ADVANCED IGNITION SYSTEMS

The advanced ignition systems have the potential for improving combustion
efficiency (burn a higher quantity of the injected fuel), produce a faster
combustion (higher dp/dt) and burn lean and extra-lean mixtures leading to
overall efficiency gains.

Engines working with lean mixtures and/or with high levels of EGR benefit
from traditional ignition systems with multiple sparks, as this may reduce the
occurrence of cycle variation. The problem lies with the inhomogeneous of the
charge, namely of the gases that are within the volume of the spark. If at the
time of the spark there is a high proportion of burned gases (from the EGR) in
that area, the flame development is very slow or inexistent. A system
providing multiple sparks has a better chance of one of these sparks occurs
during the "passage" of a richer mixture through the spark gap.

Corona Ignition or non-thermal transient plasma ignition (TPI) are systems,


such as those developed by BorgWarner [41] and Federal-Mogul [42], where
an electric discharge is brought by the ionization of the fluid. There is no need
for positive and negative electrodes, just a high voltage with a sufficient
potential gradient capable to form a conductive region around a conductor,
without enabling the formation of an arc to nearby objects. It forms naturally,
for example, in high voltage electricity lines (Fig. 19).
Fig. 19: Corona forming on high voltage electric lines

There are, at least, two major companies developing this type of ignition
systems (Fig. 20), BorgWarner [41] and Federal-Mogul [42]. Its use is based on
multiple streamers of electrons enabling multiple ignition sites and uses high
energy (one order of magnitude higher than sparks).

Fig. 20: Images for corona ignition systems by BornWarner (left - [41]) and Federal-Mogul
(right - [42])

According to a report by Roney & Gunderson [43], for a wide variety of loads
and engine speeds, a hike in peak pressure and imep (indicated mean
effective pressure) was observed when the ignition mode changed from spark
to corona in a special engine having both ignition systems available in each
combustion chamber. A much shorter burning duration was observed for the
advanced corona ignition and the imep (and consequently the indicated
efficiency) increased by 15 to 20% [43], during similar engine conditions
ignited by spark. Although the NOx production increased, the trade-off
between efficiency and NOx emissions improved with the use of corona
ignition.

The referred comparison was for typical spark ignition working conditions.
However, these advantages could be enhanced if the corona ignition should
be exploited, using leaner mixtures and using lower turbulence levels as the
system produces sufficient high burning rates on itself. Both these measures
would improve burning efficiency and would lower heat transfer to the walls,
while improving engine breathing. One of the major advantages of this system
is that the energy is supplied to a large volume (as wide as 10mm) of the
combustion chamber, not just the tiny location of the spark plug gap. This
leads to a much higher start of combustion (Fig. 21).
Fig. 21: Initial flame development after ignition point (AIP) (from [44]

However, corona ignition operation depends on the condition of the charge


inside the cylinder, as it works through the ionization of the gas. Therefore,
differences in load, mixture strength or engine speed will change the
behaviour of the ignition. For example, ignition is limited to operations at part
load, because the corona discharge scales inversely with gas pressure [45].

HCCI ENGINES

Homogeneous Charge Compression Ignition (HCCI) engines have been


around for decades but only in laboratories not on the road. HCCI engines
promise the best of both worlds: better efficiency that diesel engines and
power density of S I engines together with almost no NOx production. The
working principle is as follows. A pre-mixed blend of air and fuel is supplied to
a piston engine with sufficiently high compression ratio capable of auto-ignite
the mixture at TDC, producing a very high heat rate release. As the engine
could run very lean or with high levels of EGR, the production of NOx would
be very small. Also, as there is no fuel injection like in diesel engines, there is
no region with stoichiometric mixture, the lean mixture is ever-present. HCCI
combustion does not have flame propagation as it relies just on fuel auto-
ignition. On the overall the HCCI combustion was assumed to be divided into
two semi-independent processes, self ignition controlled by low temperature
chemistry (bellow 1000K, similar to knock) and bulk energy release controlled
by high temperature chemistry (above 1000K, led by CO oxidation).

However, HCCI engines are difficult to control, as the auto-ignition should


occur exactly at TDC. If happens before, a huge pressure acts onto the up-
stroke piston, and it would not happen after TDC (the temperature and
pressure do not increase after that point so misfire would occur). Different
factors such as load, engine speed, mixture strength, mixture temperature,
engine temperature, compression ratio, intake pressure, type of fuel, etc., all
play part in the reactivity and ignitability of the fuel within the engine. HCCI
ignition is difficult to occur at light and high loads and at low and high engine
speeds, so it would be very difficult to implement it in a "normal" engine. For
example, the use of higher octane fuel would allow higher loads, but the
"envelope would shrink, while the use of diesel-like fuels would promote lower
loads and larger working envelopes. A "full" envelope engine would require
variable valve and compression ratios, tight control of intake pressure, tight
control of EGR and temperature and the use of two fuels, one very reactive
(high cetane number) and another with low reactivity (high octane number).
But there are various types of engines "similar to" HCCI ones. During the 80s
Honda developed a 2-stroke ATAC (or controlled auto ignition - CAI) engine
which would not require spark induced ignition for a wide envelope of its use.
However, spark was necessary to induce ignition for high and light loads and
for low engine rpm. This engine raced in a motorcycle in the 1995 Grenada-
Dakar rally, producing very good fuel economy.

PCCI (premixed charge compression ignition) engines are conventional diesel


engines where part of the fuel injection occurs earlier during compression
stroke, leading to a higher charge dilution and much lower levels of NOx and
PM (particulate matter) production. The rest of the fuel is injected at the
normal timing, producing ignition. The "window" for this type of combustion is
very small.

Mazda is working on the SkyActiv-X, a form of HCCI engine that improves fuel
consumption significantly. This type of combustion is called SPCCI (spark
controlled compression ignition) and relies on a high compression ratio (16:1,
and the engine is supercharged) and lean and extra-lean mixtures. A spark
starts the process, increasing the charge pressure and temperature and the
remaining mixture auto-ignites on itself, as on knock. It is here that the lean
mixture is important: the auto-ignition does not become full blown knock
because the mixture is too lean and uses high levels of EGR. The engine is
over 20% more efficient than the previous engine and is more powerful as
well. The strange thing about this engine is that it performs better with the
"worst" petrol, which is the one with lower octane number. In fact, a high
octane fuel is bad for auto-ignition and this is the reason Mazda encourages
the drivers of these engines to use the lowest possible octane rating fuel.

Another concept is the so-called stratified-charge compression-ignition


(SCCI). In true HCCI engines (because of the high level of homogeneousity)
some of the mixture does not burn because it is at a lower temperature level
at the end of the compression stroke. What happens is that the mixture near
the cold walls remains colder than the rest, creating large quenching zones
near the cylinder walls. If these zones were occupied with just air there would
be no such problem. This is the principle of the SCCI, where the outer volume
does not have fuel, being the fuel relegated to the centre of the cylinder
creating a stratified charge mixture. The centre of the cylinder has a richer
mixture producing higher levels of NOx and PM, but there wont be
combustion quenching and the consequence high HC emissions.

Apparently (the secrecy does not allow for the technology to be known) the
Formula 1 engine world revolves around these concepts of HCCI, stratified
charge and jet ignition systems. These high performance engines (more than
700 hp from a 1.6 L V6 engine) have thermal efficiencies reaching the 50%
mark. So, how is it possible, knowing that powerful racing engines usually
work at high speeds at have huge mechanical and thermal losses? First of all,
these engines are well beyond their regulatory limit of 15 000 rpm, usually
working at 11 or 12 000 rpm, therefore reducing the mechanical losses. It is a
very long way from the 21 000 rpm of the bigger V8 2.4L engines of before
2007, when the rev limit was set at 19 000 rpm.

So, where are these engines getting the efficiency from? My view is that these
engines are capable of burning the fuel extremely quickly and use stratified
charge. The ignition may be of the kind of the referred jet turbulent ignition or
even spark assisted compression ignition. The near instant combustion
greatly improves efficiency. Then, the mixture is globally lean or very lean,
also improving efficiency. But what I think makes a huge difference is the
charge stratification, where the centre of the cylinder has a near stoichiometric
mixture (or even lean) and the outer of the combustion chamber has just air,
so the flame front never gets near the walls, therefore vastly reducing thermal
losses to the walls. This reduced thermal losses are evident when comparing
the cooling systems of old cars and new cars. Fig. 22 shows the huge cooling
intakes for the 1985 cars and compares them with the tiny cooling ducts of
2018 cars. It seems that the ducts area were reduced to less than 1/4 during
the 33 year span. All the cars show in the figure were turbo-charged, so part
of the cooling was necessary for the intercooling of the intake charge (1.5L in
1985 and 1.6L in 2018). This clearly shows the huge reduction in thermal
cooling required be the new engines and is a good indication of their huge
efficiency.

Fig. 22: Left: Ferrari and Williams of 1985; right: Ferrari and Williams of 2018

DOWNSIZED BOOSTED ENGINES

So far we have discussed ways of improving the indicated efficiency, which


means, the energy stream from the fuel to the top of the piston. But there is a
lot of work to be done in the so called mechanical efficiency, the energy
stream from the pressure on the piston crown to the crankshaft. I am not
going to present the various "mechanical" improvements, such as shorter
piston skirts, anti-friction deposition on piston skirts, improved piston rings,
optimized crankshaft journals or improved roller follower based cam-valve
connections or even improved lub oils. I am presenting concepts, so the better
concept for improving mechanical efficiency is to use the so called downsized
(and boosted) engines. As mechanical efficiency is the ratio between brake
(crankshaft) power and indicated power, one can improve this efficiency either
by reducing the mechanical losses (maintaining the indicated power) or
maintaining the mechanical losses and enlarging the indicated power. The
first option is used on the downsizing concept. If we start from a standard 2.0L
4-cylinder naturally aspirated S.I engine, we can develop the same levels of
torque and power using a much smaller (1.0L 3-cylinder) turbo-charged or
super-charged engine. In energy terms, turbo-charging is a better proposition,
as the power required to charge the intake is taken from the exhaust, which is
mainly "free". Super-charging requires power to be taken from the crankshaft,
so there is a penalty for using it. Further, as most of the time the engine runs
at part-load, it is a nonsense to take power from the crankshaft to pressurize
the intake only to throttle it to a pressure lower than atmospheric.

So, why do some manufacturers choose super-charging rather than turbo-


charging? The reason is "instant torque response". As the intake system of a
super-charged engine is pressurize upstream of the throttle plate, when you
open the throttle suddenly the engine response is instantaneous. The problem
is the mechanical power required to run the supercharger is of the same level
of the mechanical power saved by reducing the size of the engine. Therefore,
one end up with a similar engine in terms of efficiency, only gaining in terms of
weight and volume for the standard larger naturally aspirated engine, but
loosing in terms of price and complexity.

Using a turbo-charger implies having "turbo-lag". The intake system of a


"charged" engine needs an inter-cooler. When the air is pressurized it
increases in temperature. If its temperature can be reduced, its density is
boosted, supplying more mass flow to the engine, therefore producing more
power. The inter-cooler is traditionally an air-air heat exchanger placed in front
of the car requiring long and wide tubes to connect it to the engine. During a
sudden opening of the throttle there is an enormous torque "lag" because:
• the intake (which was at low pressure) needs to be filled with high
pressure air; the wide pipes and intercooler represent a large volume of
air that takes some time to fill them;
• the turbine of the turbo-charger only pressurizes the entering air after
the exhaust has "power" enough to turn the turbine at high speed; but
the exhaust only has enthalpy enough sometimes after the accelerator
pedal is pressed;
• the turbo-charger has mechanical inertia.

Present-day downsized engines can produce instant torque when demanded,


so there is almost no difference between larger and downsized engines.
Some cars have a second "cold" cooling system working at much lower
temperatures than the normal 90ºC of the engine cooling system. This "cold"
cooling system is an important part of the fast response downsized engines,
because it cools down the inter-cooler. As the intercooler is water-cooled it is
smaller and it may be placed near the intake, largely reducing the intake
volume and reducing turbo-lag significantly.

The turbo-charger produces inherently lag because of its inertia and because
it uses the exhaust gases that require output of the engine to spin the turbine.
Recent cars use an electric compressor (e-booster, Fig. 23) downstream of the
turbo-charger. The high voltage (48V) of modern cars enables a fast
accelerating centrifugal compressor to boost intake pressure in less than
1/3 s. Also, the 48V system also enables a mechanical boost such as the
alternator converted into electric motor. The cold cooling system is also
required to cool down these electric motors.

Fig. 23: E-booster from BorgWarner (source: [46])

List of References

[1] MARTINS, Jorge J.G., UZUNEANU, Krisztina, RIBEIRO, Bernardo & JASANSKY, Ondrej
"Thermodynamic Analysis of an Over-Expanded Engine", in 'Modeling of Spark Ignition Engines',
edited by SAE (ISBN Number: 0-7680-1366-6), 2004

[2] Larsen Gregory, "Reciprocating piston engine with a varying compression ratio", US Patent
No. US5025757, 1991

[3] MARTINS, J.J.G., JASANSKY, O., RIBEIRO, B.
"Development of a Small 4-Stroke EPI-Cycloid Miller Cycle Engine", SAE paper No. 2004-01-0195,
apresentado no SAE World Congress, Cobo Hall, Detroit, Michigan, USA, em 8 Março, 2004

[4] BON, Bernard, MARTINS, Jorge, BRITO, Francisco, COSTA, Tiago
"Double Cam Axial Engine with Over-expansion, Variable Compression, Constant Volume
Combustion, Rotary Valves and Water Injection for Regenerative Cooling", US patent 9 194 287,
issued 2015.11.24, Publication number: US9194287 B1, 2015

[5] Tiago COSTA, Jorge MARTINS, F. P. BRITO
"UMotor - Performance Evaluation of a Small Over-Expanded Engine using 3D-CFD and 1D
Engine Models", CONEM2018 (Conferência Nacional de Engenharia Mecânica), Salvador, Bahia,
Brasil, 23 Maio, 2018

[6] Kojima, S., Kiga, S., Moteki, K., Takahashi, E., "Development of a New 2L Gasoline VC-Turbo
Engine with the World’s First Variable Compression Ratio Technology", SAE Technical Paper
2018-01-0371, 2018

[7] Nissan Motor Corporation, "Two new engines, including advanced Variable Compression
Turbo, set to power the all-new 2019 Nissan Altima" (press release), available:
https://newsroom.nissan-global.com/releases/release-
487297034c80023008bd9722aa010086-two-new-engines-including-advanced-variable-
compression-turbo-set-to-power-the-all-new-2019-nissan-altima

[8] Falk Müller, Dietmar Schulze, Christian Scheibe, Stefanie Hutzelmann, Tobias Matschine,
rManfred Balling (Hilite), "Connecting rods for a two-step variable compression", German Patent
Application No. DE102013107127A1, 2015

[9] FEV, "Variable Compression Ratio", 2018 (retrieved from http://vcr.fev.com)

[10] The Saab Network, "Saab Variable", retrieved from
http://www.saabnet.com/tsn/press/000318.html)

[11] Patowary R., Baruah D.C., Thermoelectric conversion of waste heat from IC engine‐driven
vehicles: A review of its application, issues, and solutions, Int J Energy Res. 2018;1–20., doi:
10.1002/er.4021

[12] Champier, Daniel, Thermoelectric Generators: A Review of Applications, Energy Convers.
Manag 2017, 140, 167–81

[13] Wikimedia commons, Thermoelectric Generator Diagram, available:
https://commons.wikimedia.org/wiki/File:Thermoelectric_Generator_Diagram.svg (last
accessed: 28 July 2018)

[14] Wikimedia commons, Peltier element, available:
https://commons.wikimedia.org/wiki/File:Peltierelement.png, (last accessed: 28 July 2018)

[15] LaGrandeur, J., Crane, D., Hung, S., Mazar, B., Eder, A., Automotive Waste Heat Conversion
to Electric Power using Skutterudite, TAGS, PbTe and BiTe, Thermoelectrics, 2006. ICT'06. 25th
International Conference on. IEEE, 2006, doi: 10.1109/ICT.2006.331220

[16] Zervos, H., Waste heat recovery systems in vehicles, Energy Harvesting Journal – Available
at: <http://www.energyharvestingjournal.com/articles/waste-heat-recovery-systems-in-
vehicles-00003754.asp> [accessed Feb 2018]

[17] Mori, M., Yamagami, T., Sorazawa, M., Miyabe, T. et al., Simulation of Fuel Economy
Effectiveness of Exhaust Heat Recovery System Using Thermoelectric Generator in a Series
Hybrid, SAE Int. J. Mater. Manuf. 4(1):1268-1276, 2011, doi: https://doi.org/10.4271/2011-01-
1335

[18] Brito, F., Martins, J., Hançer, E., Antunes, N., Goncalves, L.M., Thermoelectric Exhaust Heat
Recovery with Heat Pipe-based Thermal Control, J Elect Mat 2015, 44(6), 1984-1997

[19] Brito, F., Alves, A., M. Pires, J., Martins, L., Martins, J., Oliveira, J., Teixeira, J., Goncalves, L.M.,
Hall, M., Analysis of a Temperature Controlled Exhaust Thermoelectric Generator during a
Driving Cycle, J Elect Mat 2016, 45(3), 1846-1870.

[20] Feng Zhou, Shailesh N. Joshi, Raphael Rhote-Vaney, Ercan M. Dede, A review and future
application of Rankine Cycle to passenger vehicles for waste heat recovery, Renewable and
Sustainable Energy Reviews, Volume 75, 2017, 1008-102

[21] BMW Group, Looking for the next gram (press release), available:
https://www.press.bmwgroup.com/global/article/detail/T0119738EN (last accessed: 28 July
2018)

[22] BorgWarner, ORC Waste Heat Recovery System,
https://www.youtube.com/watch?v=rEH18JnPvqA (last accessed 28 July 2018)

[23] Jorge Martins, "Internal Combustion Engines - 6th edition" (in Portuguese), Publindustria,
2016

[24] Kamo, R., & Bryzik, W. (1978). Adiabatic Turbocompound Engine Performance Prediction.
SAE Technical Paper 780068, 1978. (adiabatic Cummins engine)

[25] Woschni, G., Spindler, W., & Kolesa, K. (1987). Heat Insulation of Combustion Chamber Walls -
A Measure to Decrease the Fuel Consumption of IC Engines? SAE Technical Paper 870339, 1987

[26] Timoney, S., & Flynn, G. (1983). A Low Friction , Unlubricated SiC Diesel Engine. SAE
Technical Paper 830313, 1983 (unlubricated diesel engine)

[27] Bryzik, W., & Kamo, R. (1983). TACOM/Cummins Adiabatic Engine Program. SAE Technical
Paper 830314, 1983

[28] Peter Andruskiewicz, "Analytical and Experimental Investigation of Temperature-Swing
Insulation on Engine Performance", PhD thesis, University of Valencia, Spain, 2017

[29] Kosaka, Hidemasa, Wakisaka, Yoshifumi, Nomura, Yoshihiro, Hotta, Yoshihiro, Koike,
Makoto, Nakakita, Kiyomi, Kawaguchi, Akio. (2013). "Concept of “Temperature Swing Heat
Insulation” in Combustion Chamber Walls, and Appropriate Thermo-Physical Properties for Heat
Insulation Coat", SAE International Journal of Engines. SAE paper No.2013-01-0274, 2013

[30] Assanis, D., & Badillo, E. (1987). Transient Heat Conduction in Low-Heat-Rejection Engine
Combustion Chambers. SAE Technical Paper 870156

[31] Hidemasa KOSAKA, Yoshifumi WAKISAKA, Yoshihiro NOMURA, Yoshihiro HOTTA, Makoto
KOIKE, Kiyomi NAKAKITA, Akio KAWAGUCHI, "Concept of 'Temperature Swing Heat Insulation'
in Combustion Chamber Walls and Appropriate Thermo-physical Properties for Heat Insulation
Coat", SAE paper No.2013-01-0274, 2013.

[32] Peter Andruskiewicz, Paul Najt, Russell Durrett, Raul Payri, " Analysis of the effects of wall
temperature swing on reciprocating internal combustion engine processes", International
Journal of Engine Research 19(4), 2017

[33] Fukui, K., Wakisaka, Y., Nishikawa, K., Hattori, Y., "Development of Instantaneous
Temperature Measurement Technique for Combustion Chamber Surface and Verification of
Temperature Swing Concept," SAE Technical Paper 2016-01-0675, 2016

[34] Tiago Costa, Jorge Martins and F.P. Brito, "Assessment of the Potential for Wall Heat
Removal by In-Cylinder Water Injection in Internal Combustion Engines", ECOS 2018 - The 31st
International Conference on Efficiency, Cost, Optimization, Simulation and Environmental Impact
of Energy Systems, June 17-22, Guimares, Portugal , 2018

[35] Harry Ricardo, "Memories and Machines: The Pattern of My Life", Constable & Co, 1968

[36] Weaving, J.H., "Stratified Charge Engines", in 'Internal Combustion Engines: Science &
Technology', edited by Weaving, J.H., Springer, 1990

[37] Honda, "Introducing the CVCC. Beating the World: Developing a Low Emission Engine",
Honda World Links, retrieved from
http://world.honda.com/history/challenge/1972introducingthecvcc/index.html

[38] Date, T., Yagi, S., Ishizuya, A. & Fujii, I., "Research and Development of the Honda CVCC
Engine", SAE paper 740605, 1974

[39] William Attard (Mahle Powertrain LLC), "Turbulent jet ignition pre-chamber combustion
system for spark ignition engines", US patent No. US20120103302A1, 2010

[40] Bunce, M., & Blaxill, H. (2016). Sub-200 g/kWhr BSFC on a Light Duty Gasoline Engine. SAE
Technical Paper 2016-01-0709, 2016. (Jet Ignition)

[41] BorgWarner - High Frequency Ignition System Based on Corona Discharge, retrieved from
https://cdn.borgwarner.com/docs/default-source/default-document-
library/2014_whitepaper_high-frequency-ignition-system-based-on-corona-
discharge_en.pdf?sfvrsn=b30acd3c_10

[42] Burrows, J., Lykowski, J. & Mixell, K., "Corona Ignition System for Highly Efficient Gasoline
Engines " MTZ Worldwide, Vol.74, Issue 6, pp 38-41, 2013

[43] Paul D. Ronney, Martin Gundersen, "Transient plasma ignition for clean, fuel- efficient,
transportation vehicle engines", Final Report METRANS project 07-20, University of Southern
California Department of Aerospace and Mechanical Engineering, 2008

[44] M Suess, M Guenthner, M Schenk, and H-S Rottengruber, "Investigation of the potential of
corona ignition to control gasoline homogeneous charge compression ignition combustion", Proc.
IMechE Part D: J. Automobile Engineering, 2011

[45] Rixecker, G., Bohne, S., Adolf, M., Becker, M., Trump, M., & Bargende, M., "The High Frequency
Ignition System EcoFlash". Proceedings of the 1st International Conference on Advanced Ignition
Systems for Gasoline Engines (IAV-Tagung), pp 65-81, Berlin,
2012

[46] BorgWarner, "Press Releases", 2018 (retrieved from https://www.borgwarner.com/news-
media/press-releases)



View publication stats

You might also like