Pulkrabek, Willard W - Engineering Fundamentals of The Internal Combustion Engine-Pearson Education (2013 - 2014) 300-490
Pulkrabek, Willard W - Engineering Fundamentals of The Internal Combustion Engine-Pearson Education (2013 - 2014) 300-490
With flame travel distance the same, the time of flame propagation is
t = Dmax>Vf = 10.050 m2>134.7 m>sec2 = 0.00144 sec
Rotational angle during flame propagation is
angle = 13000>60 rev>sec21360°>rev210.00144 sec2 = 25.92°
Flame propagation starts at 13.92° bTDC, and spark plug firing is at 21.92° bTDC. Therefore,
ignition timing must be advanced 3.92°.
Variations in Combustion
Ideally, combustion in every cylinder of an engine would be exactly the same, and there
would be no cycle-to-cycle variation in any one cylinder. This does not happen due to
variations that occur in the intake system and within the cylinder. Even if no variations
occurred before combustion, the turbulence within the cylinder would cause statistical
variations to occur during combustion.
Differences in length and geometry of the intake manifold runners leading to the
different cylinders causes cylinder-to-cylinder variations in the volumetric efficiency
and air–fuel delivered. Temperature differences in the runners cause variations in the
evaporation rates, and this causes variations in the air–fuel ratio. More fuel vapor in a
hotter runner will displace more air and give a richer mixture and lower volumetric ef-
ficiency. Evaporative cooling causes temperature differences and, consequently, densi-
ty differences. Because gasoline is a mixture with components that evaporate at
different temperatures, the component mixture in each cylinder will not be exactly the
same. The vapor of components that evaporate early in the intake manifold will not
follow exactly the same paths and distribution as those of the still-liquid particles of the
components that evaporate later at higher temperatures. This is less of a problem in en-
gines with port injectors than in those with throttle body injectors or carburetors. Fuel
additives evaporate at different temperatures and so end up in different cylinder-to-
cylinder concentrations and even in different cycle-to-cycle concentrations for any single
cylinder. Time and spacial variations will occur when EGR is added to the intake system
(Fig. 7). Passage of air around the throttle plate breaks into two flows, causing vortices
and other variations that will then affect all downstream flow. Because of imperfect
quality control in the manufacturing of fuel injectors, each injector does not deliver ex-
actly the same quantity of fuel, and there will be cycle-to-cycle variations from any one
injector. The standard deviation of AF within a cylinder is typically on the order of
2–6% of the average (Fig. 8).
Within the cylinder, variations that already exist in the air–fuel ratio, amount of
air, fuel components, and temperature, along with normal turbulence, will cause slight
variations in swirl and squish, cylinder to cylinder and cycle to cycle. The variations in
turbulence and mass motion within the cylinder affect the flame that occurs, and this
results in substantial combustion variations, as shown in Fig. 9.
Local variations and incomplete mixing, especially near the spark plug, cause
the initial discharge across the electrodes to vary from the average, which then initi-
ates cycle-to-cycle combustion differently. Once there is a difference in the start of
295
Combustion
80
EGR
STABILITY
RATE
60 0% EXCELLENT
FREQUENCY, %
20 ACCEPTABLE
40 28 POOR
1400 rpm
3 kgm
20 MBT
0
0 100 200 300 400 500 600
imep (KPa)
FIGURE 7
Effect of EGR on the consistency of combustion in the cylinder of an SI
engine. Ideally, the value of the indicated mean effective pressure (X
axis) would be the same for all cycles (100%). With no EGR, the
frequency of average imep is very high, with some variation due to
inconsistency in turbulence, AF, etc. As EGR is added, more variation
in combustion occurs. This results in a larger spread of imep
experienced and average imep occurring less often. Reprinted with
permission from SAE Paper 780006 © 1978, SAE International [77].
19
Air–Fuel Ratio
18
17
16
Desired
AF
15
5 10 15 20 25
Cycle Number
FIGURE 8
Typical variation of air–fuel ratio that may be delivered to a single cylinder in 30
consecutive engine cycles. Adapted from [84].
combustion, the entire following combustion process will be changed. Figure 10 shows
how turbulence can change the way the same spark plug initiates combustion in two
different cycles. The kernel of combusting gases can even become detached from the
spark plug by high swirl or tumble action at the start of combustion. When this
296
Combustion
4
Cylinder Pressure, P (MPa)
FIGURE 9
Pressure as a function of time for 10
1
consecutive cycles in a single cylinder
of an SI engine, showing variation
that occurs due to inconsistency of
combustion. Similar variation would
be obtained if the pressure of the Y
TDC
coordinate were replaced with
Time temperature. Adapted from [42].
FIGURE 10
Schlieren photographs of the start of combustion for two different cycles in the same test engine
cylinder, using propane fuel at 1400 RPM and a spark plug gap of 0.8 mm. Variations result from
randomness of cylinder turbulence and cycle-to-cycle inconsistencies in swirl, squish, and tumble. Once
there is variation at the start of combustion, the entire combustion process will be different for the two
cycles. Reprinted by permission of Elsevier Science Inc. from “Flame Photographs in a Spark-Ignition
Engine,” by Gatowski, Heyward, and Deleplace, Combustion and Flame, vol. 56, pp. 71–81, copyright
1984 by The Combustion Institute, Ref. [43].
297
Combustion
happens, the entire combustion process is changed. If this kernel of startup combus-
tion is pushed against the combustion chamber wall, the added heat loss will cause
very slow reaction during that cycle. When two cycles have a difference in ignition, the
ensuing combustion process for those cycles will be quite different.
The fastest burn time within a cylinder is about twice as fast as the slowest burn
time within the same cylinder, the difference being due to random variations that
occur. The greatest percentage differences occur at light engine loads and low speeds,
with idle being the worst condition (Fig. 11).
As a compromise, the average burn time is used to set the engine operating condi-
tions (i.e., spark timing, AF, compression ratio, etc.). This lowers the output of the engine
from that which could be obtained if all cylinders and all cycles had exactly the same com-
1600
mbt
mbt–10
1400
Maximum Pressure (kPa)
mbt–20
1200
fast burn
1000
slow burn
800
600
10 15 20 25 30 35
Crank Angle of Maximum Pressure ( aTDC)
(a)
360
mbt
mbt–10
340
mbt–20
320
imep (kPa)
FIGURE 11
300
Cyclic pressure variation in same cylinder
for three different spark timings, showing
cycle-to-cycle inconsistency of combustion 280
in running engine. Spark plug was mounted
in center of combustion chamber and
260
AF = 20. (a) Maximum cycle pressure vs.
crank angle of maximum pressure, (b)
Indicated mean effective pressure vs. crank 240
10 15 20 25 30 35
angle of maximum pressure. Reprinted
Crank Angle of Maximum Pressure ( aTDC)
with permission from SAE Paper No.
830337 © 1983 SAE International [201]. (b)
298
Combustion
bustion process. A cycle in which fast burn occurs is like a cycle with an over-advanced
spark. This happens when there is a rich AF ratio, higher than average turbulence, and
good initial combustion start-up.The result of this condition is a temperature and pressure
rise too early in the cycle, with a good chance of knock occurring. This limits the compres-
sion ratio and fuel octane number that can be tolerated for a given engine. A cycle with a
slower-than-average burn time is like a cycle with a retarded spark. This occurs when
there is a lean mixture and higher-than-average EGR. The result of this condition is a
flame lasting well into the power stroke and thus hot exhaust and hot exhaust valves. This
is when partial burns and misfires occur (Fig. 12).There is also a power loss due to higher-
than-average heat loss during these cycles. Higher heat loss occurs because of the longer
combustion time and because the flame front is wider with the slower burning lean mix-
ture. Slow burn limits the EGR setting for an engine and the acceptable lean setting for
good fuel economy at cruise conditions. For smooth operation, engine conditions must be
set for the worst cyclic variations in the worst cylinder. If all cylinders had exactly the same
combustion process cycle after cycle, a higher engine compression ratio could be tolerat-
ed, and the air–fuel ratio could be set for higher power and greater fuel economy. Cheap-
er, lower octane fuel could be used.
6000
HC ppmc1
HC
4000
1400 rpm
2000 3 kgm FIGURE 12
MBT
10 A/F 14.5:1 Effect of EGR on quality of combustion
in an SI engine and hydrocarbon
FREQUENCY, %
299
Combustion
chemical, and combinations of these. The controlled variables include ignition timing,
valve timing, fuel injection duration, exhaust air pump actuation, air–fuel ratio, trans-
mission shifting, turning on of warning lights, repair diagnostic recording, reprogram-
ming of computer, etc.
On some engines, control of things such as ignition timing and injection duration
are adjusted for the entire engine. On other engines, these adjustments are made sepa-
rately for a bank of cylinders or even for a single cylinder. The fewer cylinders con-
trolled by a separate control unit, the more optimum engine operation can be made.
However, this requires more sensors, a larger control computer, potentially greater
maintenance, and higher cost.
FIGURE 13
Divided combustion chambers of CI
engines with fuel injector mounted in
secondary chamber. Reprinted with
permission from SAE Paper 710558 ©
1971 SAE International [62].
300
Combustion
ignition for the gas mixture there. This type of swirl chamber eliminates the need to
create a primary swirl in the main chamber. The intake manifold and valves can be de-
signed with a smoother straight-in flow, and higher volumetric efficiency is achieved.
Stratified charge engines often have divided chambers. These engines do not have
a homogeneous air–fuel mixture throughout the combustion chamber, but have a rich
mixture around the spark plug and a leaner mixture away from the plug. The rich mix-
ture around the spark plug assures a good start and early spread of combustion, while
the lean mixture in the rest of the combustion chamber gives good fuel economy. Often,
the air–fuel mixture occupying the greater part of the combustion chamber is too lean
to consistently ignite from a spark plug, but burns adequately when ignited by the small
rich mixture near the plug. Dual-chamber stratified charge engines have a rich mixture
in the small secondary chamber where the spark plug is located and a lean mixture in
the main chamber. Most of the power of the engine is generated in the large primary
chamber using an economical lean mixture. Some super economy lean-burn engines on
the market use an overall air–fuel ratio of 25:1. This mixture is leaner than that which
could be combusted in a homogeneous mixture engine. High swirl and squish, a rich
mixture around the spark plug, and a very high voltage spark plug with a larger than
normal electrode gap promote good starting of combustion. Experimental SI engines
that can operate on overall air–fuel ratios up to 50:1 have been developed.
Lean mixtures have low flame speeds, which, under normal conditions, would cre-
ate a potential knock problem. However, lean mixtures have lower combustion temper-
atures, caused by the excess nonreacting gases, and this eases the knock problem. Some
modern combustion technologies use very large amounts of recycled exhaust gases
(EGR) in lean-burn engines. The actual air and fuel combination used is then near stoi-
chiometric, which helps keep harmful emissions to a minimum.
Some engines have one intake valve in the main chamber and one in the secondary
chamber of each cylinder. These engines supply air and fuel at different air–fuel ratios,
with a rich mixture in the secondary chamber (Fig. 14). The extreme of this approach
RICH MIXTURE
INTAKE
LEAN MIXTURE
INTAKE
SLIGHTLY RICH
STRATIFIED MIXTURE
CHARGE
STRATIFIED
CHARGE
FIGURE 14
Divided–chamber stratified charge engine at end of intake stroke (left) and end of compression stroke
(right). Reprinted with permission from SAE Paper No. 740605 © 1974 SAE International [168].
301
Combustion
is when one intake valve supplies only air with no fuel added. Some engines have only
one intake valve per cylinder operating at low engine speeds and two valves operating
at higher engine speeds, supplying different air–fuel ratios. Some engines use a combi-
nation of intake valves and an in-cylinder fuel injector to create a stratified charge in
the combustion chamber. Some stratified charge engines do not use a divided chamber,
but have only normal, single open-chamber geometry.
A variation of stratified charge engines is dual-fuel engines. These engines use
two types of fuel simultaneously, one usually a less expensive fuel, with a lesser amount
of a better fuel used to assure ignition. These engines can be of a divided–chamber or
normal open-chamber design. Fuel is supplied, and various air–fuel ratios are ob-
tained, by a combination of multiple intake valves, fuel injectors, and/or proper con-
touring of the intake flow. Natural gas is often the main fuel used in dual-fuel engines.
This is especially true in underdeveloped third-world countries, where natural gas is
more available than other fuels.
Another type of divided–chamber engine is shown schematically in Fig. 15. This
engine essentially has a small passive secondary chamber off the side of the main com-
bustion chamber containing no intake, ignition, or special swirl. When combustion oc-
curs in the main chamber, high-pressure gases are forced through the very small orifice
into the secondary chamber. When the pressure in the main chamber falls during the
power stroke, these high-pressure gases flow slowly back into the main chamber, slight-
ly increasing the pressure pushing on the piston face and producing more work. De-
pending on the design, these backflowing gases may contain a combustible mixture and
extend combustion time (and, consequently, work output).
Many combinations and variations of divided chambers and stratified charge en-
gines have been tried, and a number of these exist in modern automobiles.
FIGURE 15
Schematic of a combustion chamber with a passive
secondary air chamber as is found on some SI and CI
engines. During the high pressure of combustion, gas
(air, air–fuel, and/or exhaust) is forced through the
small orifice and fills the secondary chamber. Then, as
cylinder pressure is reduced during the expansion
stroke, gas in the secondary chamber slowly flows
back into the main chamber, slightly extending
combustion and the power stroke process.
302
Combustion
Example Problem 3
A six-cylinder SI engine with a total displacement of 1.86 liters operates at 2400 RPM using
gasoline direct injection (GDI) with two injections per cycle in each cylinder. The spark plug is
fired at 19° bTDC, and there is an ignition delay of 0.0015 seconds before combustion is estab-
lished. During combustion, there is a rich air–fuel mixture around the spark plug of AF = 11:1
and a lean mixture in the rest of the combustion chamber of AF = 20:1. The rich zone can be
modeled as a 2-cm–diameter hemisphere around the spark plug with a flame speed of 32 m/sec.
The lean zone then fills the rest of the combustion chamber to the outer edge of the bore and a
flame speed of 19 m/sec. The engine has a compression ratio of 9.8, a stroke of 7.20 cm, and a con-
necting rod length of 13.3 cm, with the spark plug at the center of the combustion chamber.
Calculate:
1. crank angle position at the end of combustion
2. piston speed at the end of combustion
3. volume in combustion chamber of one cylinder at end of combustion
(1) Bore can be calculated as follows:
Time for flame to travel through fuel-lean zone to outer edge of bore is
303
Combustion
Power Operation
For maximum power at WOT (e.g., fast start-up, accelerating up a hill, an airplane tak-
ing off), fuel injectors and carburetors are adjusted to give a rich mixture, and ignition
systems are set with retarded spark (spark later in cycle). This gives maximum power at
a sacrifice of fuel economy. The rich mixture burns faster and allows the pressure peak
to be more concentrated near TDC, with the probable compromise of rougher opera-
tion. At high engine speeds, there is less time for heat transfer to occur from the cylin-
ders, and exhaust gases and exhaust valves will be hotter. To maximize flame speed at
WOT, no exhaust gas is recycled, resulting in higher levels of NOx.
Interestingly, another way of obtaining added power from an engine is to operate
with a lean mixture. Race cars are sometimes operated this way. In a lean mixture,
flame speed is slow and combustion lasts well past TDC. This keeps the pressure high
well into the power stroke, which produces a greater power output. This way of opera-
tion produces very hot exhaust gases due to the late combustion. This hot exhaust,
combined with the unused oxygen of the lean mixture, oxidizes the exhaust valves and
seats very quickly. This possibly requires changing of the exhaust valves quite often,
something unacceptable except for race cars. Ignition timing must be set specially for
this kind of operation.
Cruising Operation
For cruising operation such as steady freeway driving or long-distance airplane travel,
less power is needed and brake-specific fuel consumption becomes important. For this
type of operation, a lean mixture is supplied to the engine, high EGR is used, and igni-
tion timing is advanced to compensate for the resulting slower flame speed. Fuel usage
efficiency (miles/liter) will be high, but thermal efficiency of the engine will be lower.
This is because the engine will be operating at a lower speed, which gives more time
per cycle for heat losses from the combustion chamber.
304
Combustion
305
Combustion
Special starting fluids can be purchased for aiding engine start-up in extremely
cold temperatures. Substances such as diethyl ether with very high vapor pressures
evaporate more readily than gasoline and give a richer air–fuel vapor mixture for initi-
ating combustion. These fluids generally are obtained in pressurized containers and are
sprayed into the engine air intake before starting.
FIGURE 16
Combustion chamber of L head, valve-in-block engine.
For several decades from around 1910 to the 1950s, this
was the standard geometry of many engines. With a few
exceptions to the general design, this type of combustion
chamber generally did not promote high levels of swirl,
squish, or tumble, considered very desirable in modern
combustion philosophy. Flame travel distance was also
long compared with that of modern combustion
chambers. All of these factors restricted these early
engines to much lower compression ratios than those
common today.
306
Combustion
The spark plug is placed near the centerline of the cylinder, so the flame must
travel only about one-fourth of the bore diameter before most of the air–fuel mixture
is consumed. Some engines have two spark plugs per cylinder. This allows the air–fuel
mixture to be consumed by two flame fronts and, with proper placement, can almost
decrease combustion time by a factor of two. Dual plugs can be designed to be fired ei-
ther simultaneously or sequentially. At least one automobile manufacturer has experi-
mented with a prototype four-cylinder engine with four spark plugs per cylinder, one in
the center and three at the outer periphery. Most aircraft engines have two spark plugs
per cylinder. However, this is more a safety feature than a design to improve combus-
tion. Many aircraft systems have redundancy in case one spark plug fails.
In addition to fast combustion time, the combustion chambers would provide
smooth engine operation during the power stroke. With the spark plug located near the
center of the clearance volume, pressure buildup at the start of combustion will be slow
due to the large volume of surrounding gas that must be compressed. Placing the spark
plug near the edge of the combustion chamber would give a quicker early pressure rise,
because there would be less gas to compress in the immediate vicinity. This would create
a rougher engine cycle. Near the end of combustion, the flame front exists in the small
volume of gas at the edges of the combustion chamber. This allows the pressure rise to
die away slowly and contributes again to a smooth power stroke. If combustion were to
end with the flame front in a large part of the combustion chamber, pressure rise would
end abruptly and the end of the power stroke would be less smooth. If knock occurs dur-
ing combustion, it will occur in the last end gas to burn. If only a small amount of end gas
exists, any knock that occurs can be tolerated and probably will not be detected. A very
small amount of knock is even desirable with this kind of combustion chamber. It means
that operating temperature and pressure are at a maximum, and the small amount of
knock is not detectable. It can even raise power output very slightly by increasing the
pressure a little near the end of the combustion process.
In addition to being placed near the center of the clearance volume, the spark
plug should be positioned near both the intake and exhaust valves. It should be near
the intake valve to assure a richer mixture between the spark plug electrodes, combus-
tion being easier to ignite in a rich mixture. The gas mixture away from the intake valve
will have a greater amount of leftover exhaust residual and will consequently be lean-
er. The spark plug should also be near the exhaust valve. The exhaust valve and port
are the hottest parts of the combustion chamber, and this higher temperature will as-
sure good fuel vaporization near the spark plug. This also keeps the exhaust valve away
307
Combustion
from the hot end gas, where the higher surface temperature could cause surface igni-
tion and knock.
To keep the size of the combustion chamber at a minimum, most modern SI en-
gines have overhead valves. This requires overhead camshafts or a hydromechanical
linkage between the valves and a camshaft mounted in the engine block. Another way
of decreasing combustion chamber size is to have more cylinders for a given displace-
ment volume.
These combustion chambers offer less heat loss, less force on the head bolts, less
wall deposit buildup, and less exhaust emissions. There is less heat loss due to the small-
er wall surface area-to-volume ratio than that which existed in the earlier valve-in-
block engines. This, in turn, provides better thermal efficiency. There is less force on the
head bolts holding the head to the engine block because of the smaller head surface
area of the combustion chamber. For a given cylinder pressure, total force will be pro-
portional to the surface area on which that pressure is applied. There will be less wall
deposit buildup with time due to the hotter temperatures and high swirl motion, which
cleans the walls in this type of chamber. There will be less exhaust emissions because of
the smaller flame quench volume and fewer wall deposits.
Probably the greatest disadvantage of this type of combustion chamber is the
limited design flexibility it offers. Because of the limited wall surface area, it is ex-
tremely difficult to fit the needed valves, spark plugs, and fuel injectors. Often, valve
sizing and gas flow control contouring must be compromised because of space consid-
erations. Cylinders with multiple intake and exhaust valves decrease flow resistance,
but increase design complexity. Often, surface areas must be cut away to allow for
clearance between valves and piston face. This compromises the desire for minimum
corner spaces in the chamber. Mechanical strength cannot be compromised, and
enough surface material must be allowed between valves to assure structural stability.
Some modern engines have divided combustion chambers, as described earlier.
These offer high volumetric efficiency, good fuel economy, and cycle operation flexibil-
ity. Two of their main disadvantages are greater heat loss, due to high surface area, and
higher cost and difficulty in manufacturing.
Combustion chambers in older automobile engines, especially the flat-head
valve-in-block type shown in Fig. 16, had a much longer flame travel distance and com-
bustion time. Inlet systems were not designed to create swirl motion, and any inlet swirl
that might exist would be greatly dampened out near TDC, when the air–fuel mixture
is forced away from the cylinder centerline. Little squish motion is promoted. Some
mass motion and some turbulence are present, but at lower levels, because of the slow-
er engine speeds. Because of the much greater resulting combustion times, compres-
sion ratios had to be much lower. In the early years of this type of engine (the 1920s),
compression ratios were in the range of four or five, increasing to about seven in later
years (the 1950s).
Very large engines are almost always CI engines. Because of their large combus-
tion chambers and corresponding long flame travel distance, combined with slow en-
gine speed, they would require very high octane fuel and/or very low compression
ratios if operated as an SI engine. With the very long real time of combustion in a cylin-
der, it would be impossible to avoid serious knock problems.
308
Combustion
5 COMBUSTION IN CI ENGINES
Combustion in a compression ignition engine is quite different from that in an SI en-
gine. Whereas combustion in an SI engine is essentially a flame front moving through a
locally homogeneous mixture, combustion in a CI engine is an unsteady process occur-
ring simultaneously at many spots in a very nonhomogeneous mixture at a rate con-
trolled by fuel injection. Air intake into the engine is unthrottled, with engine torque
and power output controlled by the amount of fuel injected per cycle. Because the in-
coming air is not throttled, pressure in the intake manifold is consistently at a value
close to one atmosphere. This makes the pump work loop of the engine cycle very
small, with a corresponding better thermal efficiency compared with that of an SI en-
gine. This is especially true at low speeds and low loads, when an SI engine would be at
part throttle with a large pump work. For CI engines,
Only air is contained in the cylinder during the compression stroke, and much
higher compression ratios are used in CI engines. Compression ratios of modern CI en-
gines range from 12 to 24. Unlike thermal efficiencies in normal SI engines, in CI en-
gines high thermal efficiencies (fuel conversion efficiencies) are obtained when these
compression ratios are used. However, because the overall air–fuel ratio on which CI
engines operate is quite lean (equivalence ratio f L 0.8), less brake power output is
often obtained for a given engine displacement.
Fuel is injected into the cylinders late in the compression stroke by one or more
injectors located in each cylinder combustion chamber. Injection time is usually about
20° of crankshaft rotation, starting at about 15° bTDC and ending about 5° aTDC. Ig-
nition delay is fairly constant in real time, so at higher engine speeds fuel injection must
be started slightly earlier in the cycle.
In addition to the swirl and turbulence of the air, a high injection velocity is need-
ed to spread the fuel throughout the cylinder and cause it to mix with the air. After in-
jection, the fuel must go through the following series of events to assure the proper
combustion process:
1. Atomization. Fuel drops break into very small droplets. The smaller the original
drop size emitted by the injector, the quicker and more efficient will be this at-
omization process.
2. Vaporization. The small droplets of liquid fuel evaporate to vapor. This occurs
very quickly due to the hot air temperatures created by the high compression of
CI engines. High air temperature needed for this vaporization process requires a
minimum compression ratio in CI engines of about 12:1. About 90% of the fuel
injected into the cylinder has been vaporized within 0.001 seconds after injection.
As the first fuel evaporates, the immediate surroundings are cooled by evapora-
tive cooling. This greatly affects subsequent evaporation. Near the core of the
fuel jet, the combination of high fuel concentration and evaporative cooling will
cause adiabatic saturation of fuel to occur. Evaporation will stop in this region,
and only after additional mixing and heating will this fuel be evaporated.
309
Combustion
3. Mixing. After vaporization, the fuel vapor must mix with air to form a mixture
within the AF range that is combustible. This mixing comes about because of the
high fuel injection velocity added to the swirl and turbulence in the cylinder air.
Figure 17 shows the nonhomogeneous distribution of air–fuel ratio that develops
around the injected fuel jet. Combustion can occur within the equivalence ratio
limits of f = 1.8 (rich) and f = 0.8 (lean).
4. Self-Ignition. At about 8° bTDC, 6–8° after the start of injection, the air–fuel
mixture starts to self-ignite. Actual combustion is preceded by secondary reac-
tions, including breakdown of large hydrocarbon molecules into smaller species
and some oxidation. These reactions, caused by the high-temperature air, are
exothermic and further raise the air temperature in the immediate local vicinity.
This finally leads to an actual sustained combustion process.
5. Combustion. Combustion starts from self-ignition simultaneously at many loca-
tions in the slightly rich zone of the fuel jet, where the equivalence ratio is f = 1
to 1.5 (zone B in Fig. 17). At this time, somewhere between 70% and 95% of the
fuel in the combustion chamber is in the vapor state. When combustion starts,
multiple flame fronts spreading from the many self-ignition sites quickly con-
sume all the gas mixture that is in a correct combustible air–fuel ratio, even where
self-ignition wouldn’t occur. This produces a very quick rise in temperature and
pressure within the cylinder, shown in Fig. 18. The higher temperature and pres-
sure reduce the vaporization time and ignition delay time for additional fuel par-
ticles and cause more self-ignition points to further increase the combustion
process. Liquid fuel is still being injected into the cylinder after the first fuel is al-
ready burning. After the initial start of combustion, when all the air–fuel mixture
that is in a combustible state is quickly consumed, the rest of the combustion
Liquid
Injector
A
B
C
D
E
FIGURE 17
Fuel jet of a CI engine showing air–fuel vapor zones around the inner liquid core. The
liquid core is surrounded by successive zones of vapor that are (A) too rich to burn, (B)
rich combustible, (C) stoichiometric, (D) lean combustible, and (E) too lean to burn.
Self-ignition starts mainly in zone B. Solid carbon soot is generated mostly in zones A
and B.
310
Combustion
Cylinder Pressure, P
B
FIGURE 18
Cylinder pressure as a function of crank angle for a CI engine. Point A is where fuel
injection starts, A to B is ignition delay, and point C is the end of fuel injection. If the
cetane number of the fuel is too low, a greater amount of fuel will be injected during
ignition delay time. When combustion then starts, the additional fuel will cause the
pressure at point B to increase too fast, resulting in a rough engine cycle. Adapted
from [10].
process is controlled by the rate at which fuel can be injected, atomized, vapor-
ized, and mixed into the proper AF. This rate of combustion, now controlled by
injection rate, can be seen in Fig. 18 in the slower pressure rise that occurs after
the initial fast rise. Combustion lasts for about 40° to 50° of engine rotation, much
longer than the 20° of fuel injection. This is because some fuel particles take a
long time to mix into a combustible mixture with the air, and combustion there-
fore lasts well into the power stroke. This can be seen in Fig. 18, where the pres-
sure remains high until the piston is 30°–40° aTDC. About 60% of the fuel is
burned in the first third of combustion time. Burning rate increases with engine
speed, so the burn angle remains about constant. During the main part of the
combustion process, anywhere from 10% to 35% of the fuel vapor in the cylinder
will be in a combustible AF.
It has been theorized that average engine speed strongly correlates with the in-
verse of stroke length. This puts the average piston speed for all engines in the range of
about 5 to 20 m/sec. Large, slow engines have adequate real time to inject, atomize, va-
porize, and mix the fuel for combustion to occur in 40°–50° of engine rotation. These
direct injection (DI) engines have large open chambers without the need for high swirl.
They generally have very high injection pressures that give the fuel jets high velocity.
This assures that the penetration of the jet reaches across the large combustion cham-
ber and greatly assists in the mixing of the fuel and air. Large DI engines generally
have higher brake thermal efficiency because they operate slower, which reduces fric-
tion losses, and they have lower combustion chamber surface area-to-volume ratios,
which reduces heat losses.
311
Combustion
Small CI engines operate at much higher speeds and need high swirl to enhance
and speed the vaporization and mixing of the fuel. This must occur at speeds up to 10
times faster so that combustion can occur in the same desired 40°–50° of engine rota-
tion. Special intake and cylinder geometries are needed to generate this necessary high
swirl. These geometries can include special swirl chambers separate from the main
combustion chamber, as shown in Fig. 13. These indirect injection (IDI) engines with di-
vided chambers inject the fuel into the smaller secondary chamber and can use much
lower injection pressures. This type of engine gives lower fuel jet velocities, which are
adequate to penetrate across the smaller combustion chamber. The high swirl generat-
ed in the secondary chamber provides the needed mixing of fuel and air. As the gas
mixture in the secondary chamber combusts, it expands through the orifice into the
main chamber, carrying liquid fuel droplets with it and providing swirl in the main
chamber. Here, the main portion of combustion occurs much as it does in an SI engine.
The higher speeds at which IDI engines generally operate make them better automo-
bile engines. Because of the large surface area-to-volume ratio in the combustion
chambers, there is a greater heat loss, and this typically requires a higher compression
ratio. Starting a cold engine is also more difficult because of this factor.
Example Problem 4
The diesel engine has a compression ratio of 18:1 and operates on an air-standard Dual cycle. At
2400 RPM, combustion starts at 7° bTDC and lasts for 42° of engine rotation. The ratio of con-
necting rod length to crank offset is R = 3.8. Calculate:
1. ignition delay
2. cycle cutoff ratio
(1) Combustion starts at 7° bTDC and fuel injection starts at 20° bTDC. Ignition delay in
degrees of engine rotation is
Example Problem 5
It is desired to have combustion in a medium-size CI engine operating on a Diesel cycle start
at 1° aTDC, using fuel with a cetane number of 41. The engine is a straight six operating at
980 RPM, with a total displacement of 15.6 liters and bore and stroke related as S = 2.02 B.
The compression ratio is 16.5:1, and the temperature and pressure of air entering the cylin-
ders are 41° C and 0.98 bar.
312
Combustion
Calculate:
1. crank angle at which fuel injection should start
2. ignition delay in milliseconds
(1) Bore and stroke can be calculated as follows:
Vd = Nc1p>42B2S = 15.6 L = 0.0156 m3 = 16 cyl21p>42B212.02 B2
B = 0.1179 m S = 12.022B = 12.02210.1179 m2 = 0.2382 m
Average piston speed is
= 0.0000854
[121.22>1Pirck - 12.42]0.63 = 5121.22>[10.982116.521.35 - 112.42]60.63 = 0.7914
ID1ca2 = [10.362 + 10.22217.782]exp[19376210.0000854210.79142] = 3.91°
Crank angle at which fuel injection should start is
Fuel Injection
The nozzle diameter of a typical fuel injector is 0.2–1.0 mm. An injector may have one
nozzle or several.
The velocity of liquid fuel leaving a nozzle is usually about 100 to 200 m/sec. This
is quickly reduced by viscous drag, evaporation, and combustion chamber swirl. The
vapor jet extends past the liquid jet and, ideally, just reaches the far walls of the com-
bustion chambers. Evaporation occurs on the outside of the fuel jet while the center re-
mains liquid. Figure 17 shows how the inner liquid core is surrounded by the following
successive vapor zones of air–fuel:
(A) too rich to burn
(B) rich combustible
(C) stoichiometric
(D) lean combustible
(E) too lean to burn
313
Combustion
Liquid drop diameter size leaving the injector is on the order of 10 -5 m 110 -2 mm2
and smaller, generally with some normal distribution of sizes. Factors that affect droplet
size include pressure differential across the nozzle, nozzle size and geometry, fuel prop-
erties, and air temperature and turbulence. Higher nozzle pressure differentials produce
smaller droplets.
Injectors on some small engines with high swirl are designed to spray the fuel jet
against the cylinder wall. This speeds the evaporation process, but can be done only in
those engines that operate with very hot walls. This is necessary because of the limited
real time of each cycle in small engines that operate at high speeds. This practice is not
needed and should not be done with large engines operating at slower speeds. These
have low swirl and cooler walls, which would not evaporate the fuel efficiently. This
would lead to high specific fuel consumption and high HC emissions in the exhaust.
ID r 1>CN (2)
Cetane number and oetane number also have a strong inverse correlation for
most fuels:
CN r 1>ON (3)
Cetane number can be changed by blending small amounts of certain additives to
the fuel. Additives that accelerate ignition include nitrites, nitrates, organic peroxides,
and some sulfur compounds.
Alcohol, with its high octane number, is a poor fuel for CI engines.
314
Combustion
Soot
The flame in a CI engine is highly nonuniform. When self-ignition occurs, the flame will
quickly engulf all parts of the combustion chamber that have an air–fuel mixture in a
combustible ratio. Mixtures with an equivalence ratio in the range of 0.8 to 1.5 will sup-
port combustion, so some of the reaction will be in a lean mixture, some will be at or
close to stoichiometric, and some will be in a rich mixture. In the combustion zone,
where the mixture is rich, there is not enough oxygen to form stoichiometric CO2. In-
stead, some carbon monoxide and some solid carbon will be formed in the reaction
products, yielding
where
x = b + c + d
a = b + c>2 + e>2
e = y>2
These solid carbon particles are the black smoke seen in the exhaust of large
trucks and railroad locomotives.
In the very fuel-rich zones where AF is just marginally combustible, very large
amounts of solid carbon particles are generated. As the combustion process proceeds
and the air–fuel mixture in the combustion chamber is further mixed by swirl and tur-
bulence, most of the carbon particles further react, and only a very small percentage of
them eventually reach the surrounding environment. Solid carbon particles are a fuel
and react with oxygen when the proper mixture is obtained:
In that the overall air–fuel ratio is lean in a CI engine, most of the carbon will find
and react with the excess oxygen. Even after the mixture leaves the combustion cham-
ber, additional reactions take place in the exhaust system, further reducing the amount
of solid carbon. In addition, most CI engine exhaust systems have a particulate trap
315
Combustion
that filters out a large percentage of the remaining solid carbon. Only a small
percentage of the original solid carbon particles that were generated in the
combustion chamber is exhausted to the environment.
To keep exhaust smoke (soot) within tolerable limits. CI engines are op-
erated with an overall lean AF 1f 6 0.82. If these engines were to operate
with an overall stoichiometric AF, the amount of exhaust smoke would be un-
acceptable. Even with lean operation, many metropolitan areas are very con-
cerned with diesel exhaust from trucks and buses. In many locations, very
stringent laws are being imposed on bus and truck operation, and major im-
provements must be made to reduce exhaust emissions from these vehicles.
A CI engine operates with unthrottled intake air, controlling engine
power by the amount of fuel injected. When a truck or railroad locomotive is
under heavy load, such as when it accelerates from a stop or going up a hill,
more than the normal amount of fuel is injected into the cylinders. This results
in a richer mixture that generates a higher amount of solid carbon soot. A
large amount of exhaust smoke is very noticeable under these conditions.
Because CI engines operate lean overall, they have a high combustion
efficiency—generally around 98%. Of the 2% combustion inefficiency, about
half appears as HC emissions in the exhaust. This is in the form of solid carbon
and other HC components. Some HC components are absorbed on the carbon
particles and carried out in the exhaust. If 0.5% of the carbon in the fuel were
exhausted as solid particles, the resulting smoke would be unacceptable. This
means that the amount of solid carbon being exhausted must be kept well
below this 0.5%.
Because of the 98% fuel conversion efficiency and the high compression
ratios, the thermal efficiency of CI engines is high compared with that of Otto
cycle SI engines. However, because of their fuel-lean operation, their power
output per unit displacement volume is not as good.
Cold-Weather Problems
Starting a cold CI engine can be very difficult. The air and fuel are cold, so fuel
evaporation is very slow and ignition delay time is lengthened. Lubrication oil
is cold, its viscosity is high, and distribution is limited. The starter motor has to
turn the cold engine that is poorly lubricated with very high viscosity oil. This
results in a slower-than-normal turnover speed to start the engine. Because of
very low engine speed, a greater amount of blowby past the piston occurs, re-
ducing the effective compression. As the starter motor turns the engine to start
it, air within the cylinders must compress enough to raise the temperature well
above self-ignition temperature. This does not readily happen. The slower-
than-normal rotation of the engine, combined with the cold metal cylinder
walls, promotes a large heat loss to the walls and keeps the air temperature
below that needed to self-ignite the fuel.To overcome this problem, a glow plug
is used when starting some CI engines.A glow plug is a simple resistance heater
connected to a battery with the heated surface located within the combustion
chamber of the engine. For a short time, 10–15 seconds, before starting the en-
316
Combustion
gine, the glow plug is turned on and the resistor becomes red hot. Now, when the engine
is started, combustion in the first few cycles is not ignited by compressive heating, but by
surface ignition off the glow plug.After just a few cycles, the cylinder walls and lubricant
are warmed enough and more normal operation of the engine is possible. Then, the
glow plug is turned off and self-ignition caused by compressive heating occurs. Another
cold-weather method used to aid starting on some engines is an electrically heated in-
take manifold that heats the air entering the cylinders. Due to their larger wall surface
area, engines with divided combustion chambers have a greater heat loss problem than
those with single open chambers and are generally more difficult to start.
Because of the large amount of energy needed to rotate and start very large CI
engines that are cold, using an electric motor powered from a battery is sometimes not
practical. Instead, a small internal combustion engine can be used as the starting motor
for the larger engine. This pony engine, usually having two or four cylinders, is first
started and then used to turn over the large engine by engaging it to the flywheel of the
large engine. The pony engine is then disengaged when the large engine is started.
To aid in cold starting, many medium-size CI engines are built with a higher com-
pression ratio than would otherwise be needed. Some are also given a larger flywheel
for this purpose. Preheating the lubricating oil electrically is done on some engines to
help the starting process; some systems even distribute the oil throughout the engine
before starting by means of an electric oil pump. This oil not only lubricates the engine
parts and makes starting easier, but also reduces the high engine wear that occurs at
this time. Late injection and a richer air–fuel mixture are also used to aid starting.
It is not an uncommon practice to leave large CI engines running continuously
during cold weather to avoid the problem of restarting them. Truck engines are often
left running at highway truck stops in northern climates during winter. This is undesir-
able in that it wastes fuel and adds air pollution to the environment.
Another cold-weather problem encountered with cold CI engines in trucks and
automobiles is pumping the fuel from the fuel tank to the engine. Often the fuel tank is
located some distance from the engine, and the fuel supply lines run outside the warm
engine compartment. The high viscosity of cold fuel oil makes it very difficult to pump
it through the long, often small-diameter, fuel lines. Some diesel fuels will even gel in
the fuel tank in cold weather. Many vehicles overcome this problem with an electric
fuel tank heater and/or by recirculating the fuel through the warm engine compart-
ment. As much as twice the needed fuel is pumped to the engine compartment. The ex-
cess fuel, after being warmed in the engine compartment, is recirculated back to the
fuel tank, where it mixes with and warms the rest of the fuel. It is often necessary to
change to a higher grade fuel oil for winter operation of an automobile. The more cost-
ly high-grade fuel has lower viscosity and is more easily pumped. It also works better
through the fuel injectors.
Recent development work has been done on engines, using a new combustion philoso-
phy for CI engines. These engines add fuel with the input air much as an SI engine does.
This results in an almost homogeneous air–fuel mixture filling the combustion chamber
317
Combustion
before combustion. Compression ignition is still used, but the resulting combustion is a
combination of diffusion flame combustion and homogeneous mixture combustion.
Some fuel is added with intake port injectors, while ignition occurs from normal CI in-
jection. The engine can operate on dual fuel, with the homogeneous charge of fuel fill-
ing the combustion chamber being one fuel (e.g., natural gas, methanol) and diesel fuel
supplying ignition.
318
Combustion
from various sensors, can adjust the compression ratio anywhere between 8:1 and 14:1
to obtain the most efficient and cleanest operation for any condition. Many different
fuels can be used, with automatic adjustment being made for air–fuel ratio, octane
number, etc.
With the engine head and upper block constructed as one piece, better coolant
flow can be obtained, as there is no need for head bolts or gaskets. The small engine
also has less mass, and small engines in general have less frictional losses.
Another engine system with the capability of changing compression ratio as the
engine is running was introduced early in the 21st century [205]. With this system, there
is a pivoted lever arm between the crankshaft and connecting rod of each piston. The
lever arm is also connected to the engine block through an actuator. By pivoting the
lever arm as the engine is running, the circular rotation of the large end of the connect-
ing rod can be modified, changing the stroke length and thus changing the compression
ratio. In addition, this system can modify piston motion and give an elliptical path to
the large end of the connecting rod. This causes the piston to slow down immediately
after ignition, resulting in combustion being closer to constant volume, the most effi-
cient form of combustion. It is claimed that this system can be used with either four-
stroke or two-stroke cycles, SI or CI engines, with or without supercharging, and with
variable valve timing.
A third method of changing the compression ratio of a running engine is to use
an Alvar cycle engine [238]. This is an engine that uses small secondary pistons that
reciprocate in secondary chambers which are housed in the cylinder head and connect-
ed to the primary combustion chambers as shown in Fig. 19. By phasing the motion of
the secondary pistons relative to the motion of the primary pistons, the displacement
and the compression ratio of the cylinders can be changed. This can be done with a
movable idler pulley in the belt drive of the secondary pistons.
Example Problem 6
A small airplane has a modified supercharged 2.4–liter SI engine equipped with variable com-
pression ratio. The Otto cycle engine is operated in two different modes, high-load takeoff and
low-load cruising, with the following conditions:
Takeoff Cruising
Calculate:
1. indicated thermal efficiency at takeoff and at cruising
2. rate of fuel flow into engine at takeoff and at cruising
3. indicated power at takeoff and at cruising
4. indicated specific fuel consumption at takeoff and at cruising
319
Combustion
HIGH COMPRESSION
AT LOW LOAD
FIGURE 19
Alvar cycle engine system with small pistons in secondary combustion chamber of each
cylinder. Secondary pistons can be run in phase or out of phase with large primary
pistons by changing phase angle of secondary crankshaft relative to primary crankshaft.
This changes the clearance volume at TDC of primary pistons, which changes the
compression ratio. Reprinted with permission from SAE Paper No. 981027 © 1998 SAE
International [238].
320
Combustion
= 151 kW at takeoff
isfc = mf>Wi
# #
8 SUMMARY
Combustion in an SI engine consists of ignition and flame development, flame
propagation, and flame termination. Combustion is initiated with a spark plug that ig-
nites the air–fuel mixture in the immediate vicinity of the spark plug electrodes. There
is essentially no pressure rise or work done at first, and 5–10% of the air–fuel mixture
is consumed before the combustion process is fully developed. When the flame is fully
developed it, propagates very rapidly across the combustion chamber, accelerated and
spread by turbulence and mass motions within the cylinder. This raises the temperature
and pressure in the cylinder, and the piston is forced down in the power stroke. By the
time the flame front reaches the corners of the combustion chamber, only a small per-
centage of the air–fuel mixture remains, and the flame is terminated by heat transfer
and viscous drag with the wall.
Combustion in engines with divided combustion chambers is a two-step process:
fairly normal ignition and flame development in the secondary chamber, and main
flame propagation in the primary chamber ignited by a flame jet through the separat-
321
Combustion
ing orifice. Often the air–fuel ratio is fuel rich in the secondary chamber and lean in the
primary main chamber.
Combustion in a CI engine starts with injection of fuel late in the compression
stroke. The liquid fuel droplets atomize, evaporate, mix with air, and then, after an igni-
tion delay time, self-ignite simultaneously at many sites. The flame then consumes all
fuel that is in a combustible state and continues to do this as injection continues. Com-
bustion terminates when the last fuel droplets have reacted after evaporating and mix-
ing with air to form a combustible mixture.
PROBLEMS
1 An SI engine operating at 1200 RPM has a 10.2-cm bore with the spark plug offset by 6
mm from center. The spark plug is fired at 20° bTDC. It takes 6.5° of engine rotation for
combustion to develop and get into flame propagation mode, where the average flame
speed is 15.8 m/sec.
Calculate:
(a) Time of one combustion process after flame has developed (i.e., time for flame front
to reach the furthest cylinder wall). [sec]
(b) Crank angle position at the end of combustion.
2 It is desired that flame termination be at the same crank angle position when the speed of
the engine in Problem 1 is increased to 2000 RPM. In this range, flame development takes
the same amount of real time and flame speed is related to engine speed as Vf r 0.92 N.
Calculate:
(a) Flame speed at 2000 RPM. [m/sec]
(b) Crank angle position when the spark plug should be fired.
(c) Crank angle position when flame propagation starts.
3 A CI engine with a 3.2-inch bore and 3.9-inch stroke operates at 1850 RPM. In each cycle
fuel injection starts at 16° bTDC and lasts for 0.0019 second. Combustion starts at 8°
bTDC. Due to the higher temperature, the ignition delay of any fuel injected after com-
bustion starts is reduced by a factor of two from the original ID.
Calculate:
(a) ID of first fuel injected. [sec]
(b) ID of first fuel injected in degrees of engine rotation.
(c) Crank angle position when combustion starts on last fuel droplets injected.
4 A 3.2-liter SI engine is to be designed with bowl-in-piston combustion chambers. With a
central spark plug and combustion at TDC, this gives a flame travel distance of B/4. The
engine is to operate with an average piston speed of 8 m/sec and a burn angle of 25° of en-
gine crank rotation. Stroke and bore will be related by S = 0.95 B.
Calculate:
(a) Average flame speed if the design is for an in-line four-cylinder engine. [m/sec]
(b) Average flame speed if the design is for a V8 engine. [m/sec]
5 A large CI engine operating at 310 RPM has open combustion chambers and direct injec-
322
Combustion
tion, with 26-cm bores, a 73-cm stroke, and a compression ratio of 16.5:1. Fuel injection in
each cylinder starts at 21° bTDC and lasts for 0.019 second. ID is 0.0065 second.
Calculate:
(a) ID in degrees of engine rotation.
(b) Crank angle position when combustion starts.
(c) Crank angle position when injection stops.
6 Some Ford Thunderbird V8 engines have two spark plugs per cylinder. If everything else is
kept the same, list three advantages and three disadvantages this gives for modern engine
operation and design.
7 The divided combustion chambers of a two-liter, four-cylinder, lean-burn, stratified charge
SI engine have 22% of clearance volume in the secondary chamber of each cylinder. The
intake system delivers an air–gasoline mixture into the secondary chamber at an equiva-
lence ratio of 1.2 and into the main chamber at an equivalence ratio of 0.75.
Calculate:
(a) Overall air–fuel ratio.
(b) Overall equivalence ratio.
8 The engine in Problem 7 has a volumetric efficiency of 92%, an overall combustion effi-
ciency of 99%, an indicated thermal efficiency of 52%, and a mechanical efficiency of 86%
when operating at 3500 RPM.
Calculate:
323
Combustion
bore (i.e., flame travel distance during flame propagation = 12 cm). At 1200 RPM, the
spark plug is fired at 19° bTDC. Flame speed is proportional to engine speed as Vf r 0.80 N,
while ignition and flame development takes 0.00125 seconds at all speeds. It can be assumed
that all combustion takes place during flame propagation. The engine has a 35-cm stroke, a
compression ratio of 8.2, and a connecting–rod length of 74 cm.
Calculate:
Calculate:
(a) Indicated thermal efficiency at high load and at low load. [%]
(b) Rate of fuel flow into engine at high load and at low load. [%]
(c) Indicated power at high load and at low load. [hp]
(d) The isfc at high load and at low load. [lbm/hp-hr]
DESIGN PROBLEMS
1D It is desired to build an engine with a low height for a high-speed sports car with a very low
profile. Design the intake manifold and combustion chamber for a modern fast-burn
valve-in-block engine. The engine must have high turbulence, swirl, squish, and tumble,
with a short flame travel distance.
2D Design a fuel delivery system for a flexible-fuel automobile engine. The engine should be
able to use any mixture combination of gasoline, ethanol, and methanol. Tell how engine
variables will change for various fuel combinations (e.g., ignition timing, fuel injection,
etc.). State all assumptions you make.
324
Combustion
325
This page intentionally left blank
Exhaust Flow
Touring in Automobiles
After combustion is completed and the resulting high-pressure gases have been used
to transfer work to the crankshaft during the expansion stroke, these gases must be re-
moved from the cylinder to make room for the air–fuel charge of the next cycle. The
exhaust process that does this occurs in two steps, exhaust blowdown followed by the
exhaust stroke. The resulting flow out the exhaust pipe is a non-steady-state pulsing
flow that is often modeled as pseudo-steady-state.
1 BLOWDOWN
Exhaust blowdown occurs when the exhaust valve starts to open towards the end of
the power stroke, somewhere around 60° to 40° bBDC. At this time, pressure in the
cylinder is still at about 4–5 atmospheres and the temperature is upwards of 1000 K.
Pressure in the exhaust system is about one atmosphere, and when the valve is opened
the resulting pressure differential causes a rapid flow of exhaust gases from the cylin-
der through the valve into the exhaust system (i.e., exhaust blowdown).
Flow at first will be choked, and the outflow velocity will be sonic. This occurs
when the ratio of pressures across an orifice is greater than or equal to
From Chapter 8 of Engineering Fundamentals of the Internal Combustion Engine, Second Edition.
Willard W. Pulkrabek. Copyright © 2004 by Pearson Education, Inc. All rights reserved.
327
Exhaust Flow
where
P1 = upstream pressure
P2 = downstream pressure
k = ratio of specific heats
This ratio is equal to about 2 for most gases. P1>P2 = 1.86 for air with k = 1.35.
Sonic velocity is equal to
c = 1kRT (2)
where
R = gas constant
T = temperature
It is quite high at this point because of the high temperature of the gases in the
cylinder.
As the gas flows from the cylinder into the exhaust system, it experiences a pres-
sure drop and a corresponding temperature drop due to expansion cooling. A model
often used to calculate the temperature in the exhaust system is the ideal–gas isentrop-
ic expansion relationship between temperature and pressure, namely,
where
Although the gases are not truly ideal and the blowdown process is not isentropic
due to heat losses, irreversibility, and choked flow, Eq. (3) gives a fairly good approxi-
mation to gas temperature entering the exhaust system.
In addition, the first gas leaving the cylinder will have a high velocity and a cor-
respondingly high kinetic energy. This high kinetic energy will quickly be dissipated in
the exhaust system, and the kinetic energy will be changed to additional enthalpy,
raising the temperature above Tex of Eq. (3). As the pressure in the cylinder decreases
during the blowdown process, the gas leaving will have progressively lower velocity
and kinetic energy. The first gas leaving the cylinder will thus have the highest tem-
perature in the exhaust system, with any following gas having a lower temperature.
The last elements of exhaust leaving the cylinder during blowdown will have very lit-
tle velocity and kinetic energy and will be at a temperature about equal to Tex in Eq.
(3). If a turbocharger is located close to the engine near the exhaust valves, the kinet-
ic energy gained in blowdown can be utilized in the turbine of the turbocharger. Heat
transfer also contributes to the final pseudo-steady-state temperature found in the ex-
haust system.
In an ideal air-standard Otto cycle or Diesel cycle, the exhaust valve opens at BDC
and blowdown occurs instantaneously at constant volume (process 4–5 in Fig. 1). This
does not happen in a real engine, where blowdown takes a finite length of time. So that
328
Exhaust Flow
Cylinder Pressure, P
Po 7
6 1 5
TDC BDC
Specific Volume, v
FIGURE 1
Air-standard Otto cycle on P–v coordinates showing exhaust gas after
blowdown at hypothetical state 7.
the pressure in the cylinder has been fully reduced by BDC when the exhaust stroke
starts, the exhaust valve starts to open somewhere around 60° to 40° bBDC. When this
happens, the pressure is quickly reduced, and what would have been additional useful
work is lost during the last part of the expansion stroke. Because of the finite time re-
quired, the exhaust valve is not fully open until BDC or slightly before. The timing when
the exhaust valve is opened (in most engines using a camshaft) is critical. If the valve
opens too early, more than the necessary amount of work is lost in the latter stages of
the power stroke. If it opens late, there is still excess pressure in the cylinder at BDC.
This pressure resists the piston movement early in the exhaust stroke and increases the
negative pumping work of the engine cycle.
The ideal time to open the exhaust valve depends on engine speed. The finite real
time of blowdown is fairly constant, mainly because of the choked flow condition that
occurs at the start (i.e., sonic velocity is the same regardless of engine speed). The lobe
on a camshaft can be designed to open the valve at one given crankshaft angle, which
can be picked to be optimum at one engine speed. Once this compromise speed has
been decided in the design and manufacture of the camshaft, all other engine speeds
will have less-than-optimum timing for the opening of the exhaust valve. At higher
speeds the valve will be opening late, and at lower speeds the valve will be opening
early. Many automobile engines with variable valve control can lessen this problem
somewhat, depending on the sophistication of the control system.
329
Exhaust Flow
The exhaust valve should be as large as possible, considering all other demands in
the design of the combustion chamber. A larger valve gives a greater flow area and re-
duces the time of blowdown. This allows for a later exhaust valve opening and a longer
expansion stroke with less lost work. Many modern engines have two exhaust valves
per cylinder, with the flow area of the two smaller valves being greater than the flow
area of one larger valve. This gives added design flexibility (and complexity) for fitting
the exhaust valves into the existing combustion chamber space.
Some industrial and other constant-speed engines can be designed with valve
timing optimized for that speed. Vehicle engines that do not have variable control can
be designed for the most-used condition (e.g., cruising speed for trucks and airplanes,
red line maximum speed for a drag racer). Slow-speed engines can have very late ex-
haust valve opening.
2 EXHAUST STROKE
After exhaust blowdown, the piston passes BDC and starts towards TDC in the exhaust
stroke. The exhaust valve remains open. Pressure in the cylinder resisting the piston in
this motion is slightly above the atmospheric pressure of the exhaust system. The differ-
ence between cylinder pressure and exhaust pressure is the small pressure differential
caused by the flow through the exhaust valves as the piston pushes the gases out of the
cylinder. The exhaust valve is the greatest source of flow restriction in the entire exhaust
system and is the location of the only appreciable pressure drop during the exhaust
stroke.
The exhaust stroke can best be approximated by a constant-pressure process,
with gas properties remaining constant at the conditions of point 7 in Fig. 1. Pressure
remains about constant, slightly above atmospheric, with temperature and density con-
stant at values consistent with Eq. (3).
Ideally, at the end of the exhaust stroke when the piston reaches TDC, all the ex-
haust gases have been removed from the cylinder and the exhaust valve closes. One rea-
son this does not actually happen is the finite time it takes to close the exhaust valve.
The lobe on the camshaft is designed to give a smooth closing of the valve and to keep
wear at a minimum. One cost of doing this is a slightly longer time required to close the
valve. To have the valve totally closed at TDC requires the closing process to start at
least 20° bTDC. This is unacceptable in that the valve would be partially closed during
the last segment of the exhaust stroke. Closing can start only at or very close to TDC,
which means that total closing doesn’t occur until 8°–50° aTDC.
When the exhaust valve is finally closed, there is still a residual of exhaust gases
trapped in the clearance volume of the cylinder. The higher the compression ratio of
the engine, the less clearance volume exists to trap this exhaust residual.
The valve problem is compounded by the fact that the intake valve should be to-
tally open at TDC when the intake stroke starts. Because of the finite time required to
open this valve, it must start to open 10°–25° bTDC. There is, therefore, a period of
15°–50° of engine rotation when both intake and exhaust valves are open. This is called
valve overlap.
During valve overlap, there can be some reverse flow of exhaust gas back into
the intake system. When the intake process starts, this exhaust is drawn back into the
330
Exhaust Flow
cylinder along with the air–fuel charge. This results in a larger exhaust residual during
the rest of the cycle. This backflow of exhaust gases is a greater problem at low engine
speeds, being worst at idle conditions. At most low engine speeds, the intake throttle is
at least partially closed, creating low pressure in the intake manifold. This creates a
greater pressure differential, forcing exhaust gas back into the intake manifold. Cylin-
der pressure is about one atmosphere, while intake pressure can be quite low. In addi-
tion, real time of valve overlap is greater at low engine speed, allowing more backflow.
Some engines are designed to use this small backflow of hot exhaust gas to help va-
porize the fuel that has been injected directly behind the intake valve face.
Some engines have a one-way reed valve at the exhaust port to keep exhaust gas
from flowing back from the exhaust manifold into the cylinder and intake system dur-
ing valve overlap.
Engines equipped with turbochargers or superchargers often will have intake
pressures above one atmosphere and are not subject to exhaust backflow.
Another negative result of valve overlap is that some intake air–fuel mixture can
short-circuit through the cylinder when both valves are open, with some fuel ending up
as pollution in the exhaust system.
Variable valve timing, which is becoming common in automobile engines, de-
creases the problem of valve overlap. At low engine speeds, the exhaust valve can be
closed earlier and the intake valve can be opened later, resulting in less overlap.
If the exhaust valve is closed too early, an excess of exhaust gases is trapped in the
cylinder. Also, cylinder pressure will go up near the end of the exhaust stroke, causing
loss of net work from the engine cycle. If the exhaust valve is closed late, there is an ex-
cess of overlap, with more backflow of exhaust gas into the intake.
Figure 2 shows the flow of gases through the exhaust valve out of the cylinder.
When the valve is first opened, blowdown occurs with a very high flow rate due to the
large pressure differential. Choked flow will occur (sonic velocity) at first, limiting the
maximum flow rate. By the time the piston reaches BDC, blowdown is complete, and
flow out of the exhaust valve is now controlled by the piston during the exhaust stroke.
The piston reaches maximum speed about halfway through the exhaust stroke, and this
is reflected in the rate of exhaust flow. Towards the end of the exhaust stroke, near
TDC, the intake valve opens and valve overlap is experienced. Depending on engine
operating conditions, a momentary reverse flow of exhaust gas back into the cylinder
can occur at this point.
Example Problem 1
A 6.4-liter V8 engine with a compression ratio of 9:1 operates on an air-standard cycle and has
the exhaust process shown in Fig. 1. Maximum cycle temperature and pressure are 2550 K and
11,000 kPa when operating at 3600 RPM. The exhaust valve effectively opens at 52° bBDC.
Calculate:
331
Exhaust Flow
Exhaust
Stroke
Valve
Overlap
0
FIGURE 2
Exhaust gas flow out of cylinder through the exhaust valve(s), showing
blowdown and exhaust stroke. Possible reverse flow back into the cylinder may
occur during valve overlap. Adapted from [28].
(1) Blowdown will occur between 52° bBDC and BDC. This is 52>360 = 0.1444 of a
revolution.
Clearance volume is
332
Exhaust Flow
At the end of blowdown, gases are at hypothetical state 7, but volume is V4 or V1:
P7 = Po = 101 kPa
T7 = T31P7>P321k - 12>k = 12550 K21101>11,000211.35 - 12>1.35 = 756 K
VBDC = V4 = V1 = 1Vc + Vd2 = 10.0001 + 0.00082 = 0.0009 m3
m7 = PV>RT = 1101 kPa210.0009 m32>10.287 kJ>kg-K21756 K2
= 0.00042 kg
1¢m2blowdown = mEVO - m7 = 0.00150 - 0.00042 = 0.00108 kg
This is the same result as the one obtained previously. Calculation of the
amount of mass in the cylinder would be the same using any point along process line
3–4. This means that the same percentage of the exhaust flow must occur during blow-
down regardless of when the exhaust valve is opened. Size of the exhaust valve(s) and
the corresponding mass flow rate through the valve(s) then dictate when the valve
should be opened.
(3) If flow is choked at the start of blowdown, velocity will be sonic. Using Eq. (2)
yields,
VEVO = c = 1kRTEVO = 111.3521287 J>kg-K211244 K2 = 694 m>sec
3 EXHAUST VALVES
Exhaust valves are made smaller than intake valves, although the same amount of mass
must flow through each. The pressure differential across the intake valves of a natural-
ly aspirated engine is less than one atmosphere, while the pressure differential across
the exhaust valves during blowdown can be as high as three or four atmospheres. In ad-
dition, if and when choked flow is occuring, sonic velocity through the exhaust valve is
higher than sonic velocity through the intake valve. This can be seen in Eq. (2), with the
exhaust gas being much hotter than the intake air–fuel mixture. We have for intake
A i = 1constant2B21Up2max>ci (4)
where
A i = area of inlet valve1s2
1Up2max = average piston speed at maximum engine speed
333
Exhaust Flow
A ex = 1constant2B21Up2max>cex (5)
In actual engines, a usually has a value of about 0.8 to 0.9. To find the valve diam-
eters, we use the relationship
where
dv = valve diameter
x = number of intake valves or number of exhaust valves
Valves can be sized by making them as large as possible on the basis of Eq. (6).
There will probably not be enough room in a modern combustion chamber to make
them large enough to totally satisfy Eqs. (4) and (5). To reduce noise, some engines
with multiple exhaust valves have only one valve activated at low speeds.
4 EXHAUST TEMPERATURE
Both temperature and mass flow rate vary greatly with time in the exhaust system of
an engine. Cyclic variations will occur at engine cycle times (e.g., 0.04 second at 3000
RPM). A temperature sensor, such as a thermocouple, with a time constant much
greater than this will give a pseudo-steady-state temperature of the flow. This thermo-
couple temperature will be approximately an enthalpy average temperature and not
necessarily a true time average
# #
Tthermocouple = [ mcpT dt]>[ mcp dt] (8)
L L
where
#
m = mass flow rate of exhaust
t = time
334
Exhaust Flow
cp = specific heat
T = temperature
Temperature of the gases in the exhaust system of a typical SI engine will average
400°C to 600°C. This drops to about 300°C to 400°C at idle conditions and goes up to
about 900°C at maximum power. This is about 200°C to 300°C cooler than the exhaust
gases in the cylinder when the exhaust valve opens. The difference is because of expan-
sion cooling. All temperatures will be affected by the equivalence ratio of the original
combustion mixture.
The average temperature in the exhaust system of a typical CI engine will be
200°–500°C. This is lower than SI engine exhaust because of the larger expansion cool-
ing that occurs due to the higher compression ratios of CI engines. If the maximum
temperature in a CI engine is about the same as in an SI engine, the temperature when
the exhaust valve opens can be several hundred degrees less. The overall lean equiva-
lence ratio of a CI engine also lowers all cycle temperatures from combustion on.
Exhaust temperature of an engine will go up with higher engine speed or load,
with spark retardation, or with an increase in equivalence ratio. Things that are affect-
ed by exhaust temperature include turbochargers, catalytic converters, and particulate
traps.
Example Problem 2
When the exhaust valve opens and blowdown occurs, the first elements of flow have high sonic
velocity and high kinetic energy. The high velocity is quickly dissipated in the exhaust manifold,
where flow velocity is relatively low. The kinetic energy of the gas is converted to additional en-
thalpy with an increase in temperature. Calculate the theoretical maximum temperature in the
exhaust flow of Example Problem 1.
Applying conservation of energy, we have
¢KE = V2>2gc = ¢h = cp ¢T
Using values from Example Problem 1, we obtain
¢T = V2>2gccp = 1694 m>sec22>[211 kg-m>N-sec2211.108 kJ>kg-K2] = 217°
Tmax = Tex + ¢T = 756 + 217 = 973 K = 700°C
Actual maximum temperature would be less than this, due to heat losses and other irreversibili-
ties. Only a very small percentage of the exhaust flow would have maximum kinetic energy and
would reach maximum temperature. The time-averaged temperature of the exhaust would be
more consistent with T7 in Fig. 1 and Example Problem 1.
5 EXHAUST MANIFOLD
After leaving the cylinders by passing out of the exhaust valves, exhaust gases pass
through the exhaust manifold, a piping system that directs the flow into one or more
exhaust pipes. Exhaust manifolds are usually made of cast iron and are sometimes de-
signed to have close thermal contact with the intake manifold. This is to provide heat-
ing and vaporization in the intake manifold.
335
Exhaust Flow
Chemical reactions are still occurring in the exhaust flow as it enters the exhaust
manifold, with carbon monoxide and fuel components reacting with unreacted oxygen.
These reactions are greatly reduced because of heat losses and the lower temperature
after blowdown. Some modern engines have insulated exhaust manifolds that are de-
signed to operate at much hotter temperatures and act as a thermal converter to reduce
unwanted emissions in the exhaust gas. Some of these are equipped with electronically
controlled air intake to provide additional oxygen for reaction.
Modern smart engines have a number of sensors in the exhaust manifold to pro-
vide input to engine controls. These can be some combination of thermal, chemical,
electrical, and/or mechanical in nature and can supply information about levels of O2,
HC, NOx, CO, CO2, particulates, temperature, and knock. This information is then
used by the engine management system (EMS) to adjust engine parameters such as
AF, injection timing, ignition timing, and EGR rate.
From the exhaust manifold, the gases flow through an exhaust pipe to the emission
control system of the engine, which generally consists of thermal and/or catalytic convert-
ers. One argument says these converters should be as close to the engine as space allows
to minimize heat losses. On the other hand, this setup creates high–temperature problems
in the engine compartment. These converters promote reduction of emissions in the ex-
haust gases by additional chemical reaction.
336
Exhaust Flow
HISTORIC—EXHAUST REDUCTION
To reduce exhaust and emissions, and to save fuel, some automobiles have been designed
to automatically turn the engine off when the vehicle stops, such as at a stoplight. A light
touch on the accelerator pedal restarts the engine when the driver desires to move on. This
method of operation will become more efficient when 42-volt electrical systems become
standard and the starter is integrated with the engine flywheel. Some manufacturers have
developed engine drive systems that shift into neutral gear when the engine is at idle. By
thus reducing the speed and load of the engine, fuel use and exhaust emissions are reduced.
When engine speed is increased, the system automatically shifts back into drive gear.
6 TURBOCHARGERS
In turbocharged engines, exhaust gases leaving the exhaust manifold enter the turbine
of the turbocharger, which drives the compressor that compresses the incoming air.
Pressure of the exhaust gas entering the turbine is only slightly higher than atmospher-
ic, and only a very small pressure drop is possible through the turbine. In addition, this
non-steady-state pulsed flow varies widely in kinetic energy and enthalpy, due to the
velocity and temperature differences that occur during blow down and the following
exhaust stroke. A pseudo-steady-state flow is assumed, with
where
#
Wt = time-averaged turbine power
#
m = time-averaged exhaust mass flow rate
h = specific enthalpy
cp = specific heat
T = temperature
Because of the limited pressure drop through the turbine, it must operate at speeds
upward of 100,000 RPM to generate enough power to drive the compressor. These high
speeds, along with the high-temperature corrosive gases within which the turbine oper-
ates, create major mechanical and lubrication design challenges.
Turbochargers should be mounted as close as possible to the cylinder exhaust
ports so that turbine inlet pressure, temperature, and kinetic energy can be as high as
possible.
One problem associated with turbocharging is the slow response time experi-
enced when the throttle is opened quickly. It takes several engine cycles before the in-
creased exhaust flow can accelerate the turbine rotor and give the desired pressure
boost to the inlet air–fuel mixture. To minimize this turbo lag, lightweight ceramic ro-
tors with small rotational moments of inertia that can be accelerated more quickly are
used. Ceramic is also an ideal material because of the high temperatures.
337
Exhaust Flow
Many turbochargers have a bypass that allows exhaust gas to be routed around
the turbine. This is to keep intake flow from being overcompressed when engine oper-
ating conditions are less demanding. The amount of gas passing through the turbine is
controlled according to engine needs.
Some experimental exhaust turbines have been used to drive small high-speed
generators instead of intake compressors. The electrical energy output from these sys-
tems can be utilized in various ways, such as driving the engine cooling fan.
338
Exhaust Flow
the exhaust valve opens or when the exhaust slot is uncovered near the end of the
power stroke. This is immediately followed with an intake process of compressed air or
air–fuel mixture. As the air enters the cylinder at a pressure usually between 1.2 and 1.8
atmospheres, it pushes the remaining lower pressure exhaust gas out the still-open ex-
haust port in a scavenging process. There is some mixing of intake and exhaust, with
some exhaust residual staying in the cylinder and some intake gas passing into the ex-
haust system. For those engines which use direct fuel injectors (CI engines and larger
modern SI engines) and have only air in the intake system, this intake gas that gets into
the exhaust system during valve overlap does not contribute to emission problems.
However, it does reduce engine volumetric efficiency and/or trapping efficiency. For
those engines which intake an air–fuel mixture, any intake gas that gets into the ex-
haust system adds to hydrocarbon pollution and reduces fuel economy. Some two-
stroke cycle engines have a one-way reed valve at the exhaust port to stop gases from
flowing back into the cylinder from the exhaust system.
Example Problem 3
The experimental two-stroke cycle automobile engine 5-8 burns stoichiometric gasoline mixed
with oil at a ratio of 15 to 1. The engine has a combustion efficiency of 98% for gasoline, but only
82% for the oil. Peak temperature and pressure in the cycle are 2550°C and 9610 kPa.
Calculate:
1. total mass flow rate of exhaust
2. mass flow rate of unburned gasoline in the exhaust
3. mass flow rate of unburned oil in the exhaust
4. approximate exhaust temperature
We have the following:
engine speed N = 3700 RPM
displacement volume Vd = 0.0004095 m3 1one cylinder2
delivery ratio ldr = 0.880
charging efficiency lce = 0.641
trapping efficiency lte = 0.728
(1) Mass flow rate of air into engine is
#
ma = ldrVdraN>n
= [10.880210.0004095 m3>cycle211.181 kg>m3213700>60 rev>sec2>11 rev>cycle2]
* (6 cylinders2
= 0.1575 kg>sec
Mass flow rate of gasoline into engine is
mf = ma>AF = 10.1575 kg>sec2>114.62 = 0.0108 kg>sec
# #
339
Exhaust Flow
Total mass flow rate of exhaust equals total mass flow rate in:
(2) Of the mass flow rates in, 72.8% gets trapped in the cylinder and 27.2% passes
through the cylinder during valve overlap.
Mass flow rate of fuel not trapped is
Mass flow rate of trapped fuel that does not get burned is
Mass flow rate of trapped oil that does not get burned is
1mo2nb = 1mo2inlte11 - hc2
# #
340
Exhaust Flow
the resulting increase in specific enthalpy. The exhaust valve must open soon enough
for blowdown to be complete when the piston reaches BDC. At this point, the cylin-
der is still filled with exhaust gas at about atmospheric pressure, and most of this is
now expelled during the exhaust stroke.
Two-stroke cycle engines experience exhaust blowdown, but have no exhaust
stroke. Most of the gas that fills the cylinder after blowdown is expelled by a scaveng-
ing process when inlet air enters at elevated pressure.
To reduce the generation of nitrogen oxides, many engines have exhaust gas re-
cycling, with some of the exhaust flow ducted back into the intake system. Those en-
gines equipped with turbochargers use the exhaust flow to drive the turbine, which in
turn drives the inlet compressor.
PROBLEMS
1 A six-cylinder SI engine, with a compression ratio of rc = 8.5, operates on an air-standard
Otto cycle at WOT. Cylinder temperature and pressure when the exhaust valve opens are
1000 K and 520 kPa. Exhaust pressure is 100 kPa, and air temperature in the intake mani-
fold is 35°C.
Calculate:
341
Exhaust Flow
5 Give two reasons that exhaust valves are smaller than intake valves.
6 A 1.8-liter, three-cylinder SI engine produces brake power of 42 kW at 4500 RPM, with a
compression ratio rc = 10.1:1 and bore and stroke related by S = 0.85 B. Connecting–rod
length is 16.4 cm. Maximum temperature in the cycle is 2700 K, and maximum pressure is
8200 kPa. Exhaust pressure is 98 kPa. The exhaust valve effectively opens at 56° bBDC.
Calculate:
(a) Time of exhaust blowdown. [sec]
(b) Percent of exhaust gas that exits cylinder during blowdown. [%]
(c) Exit velocity at the start of blowdown, assuming choked flow occurs. [m/sec]
7 Exhaust manifold pressure of the engine in Problem 6 is 98 kPa. In the manifold, the high
kinetic energy of the exhaust flow during blowdown is quickly dissipated and converted to
an increase in specific enthalpy.
Calculate:
342
Exhaust Flow
DESIGN PROBLEMS
1D Design a variable valve-timing system to be used on an in-line four-cylinder SI engine.
2D Design a turbine–generator system that can be driven by the exhaust flow of a four-cylinder
SI farm tractor engine. The output of the generator can be used to power the engine cool-
ing fan and other accessories.
343
This page intentionally left blank
Emissions and Air Pollution
From Chapter 9 of Engineering Fundamentals of the Internal Combustion Engine, Second Edition.
Willard W. Pulkrabek. Copyright © 2004 by Pearson Education, Inc. All rights reserved.
345
Emissions and Air Pollution
by E. R. Hewitt (1913)
This chapter explores the undesirable emissions generated in the combustion process
of automobile and other IC engines. These emissions pollute the environment and con-
tribute to global warming, acid rain, smog, odors, and respiratory and other health
problems. The major causes of these emissions are nonstoichiometric combustion, dis-
sociation of nitrogen, and impurities in the fuel and air. The emissions of concern are
hydrocarbons (HC), carbon monoxide (CO), oxides of nitrogen (NOx), sulfur, and
solid carbon particulates (part). Ideally, engines and fuels could be developed such that
very few harmful emissions are generated, and these could be exhausted to the sur-
roundings without a major impact on the environment. With present technology, this is
not possible, and aftertreatment of the exhaust gases to reduce emissions is very im-
portant. This consists mainly of the use of thermal or catalytic converters and particu-
late traps.
1 AIR POLLUTION
Until the middle of the 20th century, the number of IC engines in the world was small
enough that the pollution they emitted was tolerable, and the environment, with the
help of sunlight, stayed relatively clean. As world population grew, power plants, facto-
ries, and an ever-increasing number of automobiles began to pollute the air to the extent
that it was no longer acceptable. During the 1940s, air pollution was first recognized as a
problem in the Los Angeles basin in California. Two causes of this were the large popu-
lation density and the natural weather conditions of the area. The large population cre-
ated many factories and power plants, as well as one of the largest automobile densities
346
Emissions and Air Pollution
in the world. Smoke and other pollutants from the many factories and automobiles
combined with the fog that was common in this ocean area, and smog resulted. During
the 1950s, the smog problem increased along with the population density and automo-
bile density. It was recognized that the automobile was one of the major contributors to
the problem, and by the 1960s emission standards were beginning to be enforced in Cal-
ifornia. During the next decades, emission standards were adopted in the rest of the
United States and in Europe and Japan. By making engines more fuel efficient, and with
the use of exhaust aftertreatment, emissions per vehicle of HC, CO, and NOx were re-
duced by about 95% during the 1970s and 1980s. Lead, one of the major air pollutants,
was phased out as a fuel additive during the 1980s. More fuel-efficient engines were de-
veloped, and by the 1990s the average automobile consumed less than half the fuel used
in 1970. However, during this time, the number of automobiles greatly increased, result-
ing in no overall decrease in fuel usage. In 1999, petroleum consumption in the United
States amounted to 16,500 L/sec (4350 gal/sec), a large percentage of which was fuel for
internal combustion engines [214].
Additional reduction will be difficult and costly. As world population grows,
emission standards become more stringent out of necessity. The strictest laws are gen-
erally initiated in California, with the rest of the United States and the world following.
Although air pollution is a global problem, some regions of the world still have no
emission standards or laws.
Table 1 gives some of the emission standards established by the Environmental
Protection Agency (EPA). Tier 1 consists of Standards that were phased in during the
1990s and that apply in the early years of the 21st century.Tier 2 consists of the Standards
Passenger Cars 0.25 3.4 0.4 1.0 0.08 0.31 4.2 0.6 1.25 0.10
Light-Light-Duty 0.25 3.4 0.4 1.0 0.08 0.31 4.2 0.6 1.25 0.10
Trucks
Heavy-Light-Duty 0.32 4.4 0.7 0.08 0.40 5.5 0.97 0.97 0.10
Trucks
All Light 0.075–0.125 3.4 0.05–0.4 — 0.015 0.01–0.156 4.2 0.02–0.9 0.01–0.12 0.004–0.032
Vehicles
347
Emissions and Air Pollution
phased in during 2004 to 2009. During the Tier 1 period, light vehicles are divided into
(1) passenger cars, (2) light-light-duty trucks, and (3) heavy-light-duty trucks, depending
on weight. There are also separate standards for NOx levels for gasoline-fueled vehicles
and diesel oil–fueled vehicles. In Tier 2 Standards, there is only one light–vehicle classifi-
cation, and all fuels (gasoline, diesel oil, or others) have the same NOx requirements.
Tier 2 Standards have several levels of compliance under which certain vehicles can be
certified. During the phase-in period the entire fleet average of any manufacturer must
meet a NOx standard of 0.30 gm/mile. After the phase-in period, the fleet average must
then meet the NOx standard of 0.07 gm/mile. In that the standards are given in units of
gm/mile, they become more difficult to comply with as the size of the vehicle and engine
increases. In the United States, all states accept these Standards except California. Cali-
fornia has its own Standards, which are generally more stringent then those of the EPA
or any other country’s Standards. One-tenth of new vehicle sales in the United States oc-
curs in California, constituting about 100 billion dollars per year early in the 21st century.
Because of this, automobile manufacturers must give major consideration to California
emission Standards in designing their engine and vehicle products [171, 175].
2 HYDROCARBONS (HC)
Exhaust gases leaving the combustion chamber of an SI engine contain up to 6000 ppm
of hydrocarbon components, the equivalent of 1–1.5% of the fuel. About 40% of this is
unburned gasoline fuel components. The other 60% consists of partially reacted com-
ponents that were not present in the original fuel. These consist of small nonequilibri-
um molecules that are formed when large fuel molecules break up (thermal cracking)
during the combustion reaction. It is often convenient to treat these molecules as if
they contained one carbon atom, as CH 1.
The makeup of HC emissions will be different for each gasoline blend, depending
on the original fuel components. Combustion chamber geometry and engine operating
parameters also influence the HC component spectrum.
When hydrocarbon emissions get into the atmosphere, they act as irritants and
odorants; some are carcinogenic. All components except CH 4 react with atmospheric
gases to form photochemical smog.
Causes of HC Emissions
Nonstoichiometric Air–Fuel Ratio. Figure 1 shows that HC emission levels are
a strong function of AF. With a fuel-rich mixture, there is not enough oxygen to react
with all the carbon, resulting in high levels of HC and CO in the exhaust products. This
is particularly true in engine start-up, when the air–fuel mixture is purposely made very
rich. It is also true to a lesser extent during rapid acceleration under load. If AF is too
lean, poorer combustion occurs, again resulting in HC emissions. The extreme of poor
combustion for a cycle is total misfire. This occurs more often as AF is made more lean.
One misfire out of 1000 cycles gives exhaust emissions of 1 gm/kg of fuel used.
Incomplete Combustion. Even when the fuel and air entering an engine are at
the ideal stoichiometric mixture, perfect combustion does not occur and some HC ends
348
Emissions and Air Pollution
HC
Amount of Emissions in Exhaust Flow
NOx
FIGURE 1
up in the exhaust. There are several causes of this. Incomplete mixing of the air and fuel
results in some fuel particles not finding oxygen to react with. Flame quenching at the
walls leaves a small volume of unreacted air-and-fuel mixture. The thickness of this un-
burned layer is on the order of tenths of a mm. Some of this mixture, near the wall that
does not originally get burned as the flame front passes, will burn later in the combus-
tion process as additional mixing occurs due to swirl and turbulence.
Another cause of flame quenching is the expansion that occurs during combus-
tion and power stroke. As the piston moves away from TDC, expansion of the gases
lowers both temperature and pressure within the cylinder. This slows combustion and
finally quenches the flame somewhere late in the power stroke. This leaves some fuel
particles unreacted.
High exhaust residual causes poor combustion and a greater likelihood of expan-
sion quenching. This is experienced at low load and idle conditions. High levels of EGR
will also cause this.
349
Emissions and Air Pollution
It has been found that HC emissions can be reduced if a second spark plug is added
to an engine combustion chamber. By starting combustion at two points, the flame travel
distance and total reaction time are both reduced, and less expansion quenching results.
Crevice Volumes. During the compression stroke and early part of the combus-
tion process, air and fuel are compressed into the crevice volume of the combustion
chamber at high pressure. As much as 3% of the fuel in the chamber can be forced into
this crevice volume. Later in the cycle during the expansion stroke, pressure in the cylin-
der is reduced below crevice volume pressure, and reverse blowby occurs. Fuel and air
flow back into the combustion chamber, where most of the mixture is consumed in the
flame reaction. However, by the time the last elements of reverse blowby flow occur,
flame reaction has been quenched and unreacted fuel particles remain in the exhaust.
Location of the spark plug relative to the top compression ring gap will affect the amount
of HC in engine exhaust, the ring gap being a large percent of crevice volume. The farther
the spark plug is from the ring gap, the greater is the HC in the exhaust. This is because
more fuel will be forced into the gap before the flame front passes.
Crevice volume around the piston rings is greatest when the engine is cold, due to
the differences in thermal expansion of the various materials. Up to 80% of all HC emis-
sions can come from this source.
Leak Past the Exhaust Valve. As pressure increases during compression and com-
bustion, some air–fuel mixture is forced into the crevice volume around the edges of the
exhaust valve and between the valve and valve seat. A small amount even leaks past the
valve into the exhaust manifold.When the exhaust valve opens, the air–fuel mixture that is
6 2.16
NOx
4 Fuel
2.08
3 2.04
2 2.00
FIGURE 2
Fuel consumption, and HC and NOx HC
emissions, as a function of air–fuel ratio for a 1
5.7-liter, eight-cylinder SI engine running
with maximum torque (MBT) at 1400 RPM,
using a homogeneous air–fuel mixture. 0
Reprinted with permission from SAE Paper 18 19 20 21 22 23
No. 760288 © 1976 SAE International [180]. Air–Fuel Ratio
350
Emissions and Air Pollution
still in this crevice volume gets carried into the exhaust manifold, and there is a mo-
mentary peak in HC concentration at the start of blowdown.
Valve Overlap. During valve overlap, both the exhaust and intake valves are
open, creating a path where air–fuel intake can flow directly into the exhaust. A well-
designed engine minimizes this flow, but a small amount can occur. The worst condition
for this is at idle and low speed, when real time of overlap is greatest.
Oil on Combustion Chamber Walls. A very thin layer of oil is deposited on the
cylinder walls of an engine to provide lubrication between them and the moving pis-
ton. During the intake and compression strokes, the incoming air and fuel come in con-
tact with this oil film. In much the same way as wall deposits build up, this oil film
absorbs and desorbs gas particles, depending on gas pressure. During compression and
combustion, when cylinder pressure is high, gas particles, including fuel vapor, are ab-
sorbed into the oil film. When pressure is later reduced during expansion and blow-
down, the absorption capability of the oil is reduced and fuel particles are desorbed
back into the cylinder. Some of this fuel ends up in the exhaust.
Propane is not soluble in oil, so in propane-fueled engines the absorption–des-
orption mechanism adds very little to HC emissions.
As an engine ages, the clearance between piston rings and cylinder walls becomes
greater, and a thicker film of oil is left on the walls. Some of this oil film is scraped off
the walls during the compression stroke and ends up being burned during combustion.
Oil is a high-molecular-weight hydrocarbon compound that does not burn as readily as
gasoline. Some of it ends up as HC emissions. This happens at a very slow rate with a
new engine, but increases with engine age and wear. Oil consumption also increases as
the piston rings and cylinder walls wear. In older engines, oil being burned in the com-
bustion chamber is a major source of HC emissions. Figure 3 shows how HC emissions
go up as oil consumption increases.
351
Emissions and Air Pollution
HC in Exhaust (gm/hr)
FIGURE 3
HC exhaust emissions as a function of engine 2
oil consumption. Often, as an engine ages,
clearance between the pistons and cylinder
walls increases due to wear. This increases oil
consumption and contributes to an increase in 1
HC emissions in three ways: There is added
crevice volume, there is added
absorption–desorption of fuel in the thicker
oil film on cylinder walls, and there is more oil
burned in the combustion process. Adapted 0.001 0.002 0.003 0.004
from [138]. Oil Consumption Rate (L/hr)
In addition to oil consumption going up as piston rings wear, blowby and reverse
blowby also increase. The increase in HC emissions is, therefore, due to both combus-
tion of oil and the added crevice volume flow.
Two-Stroke Cycle Engines. Older two-stroke cycle SI engines and many mod-
ern small two-stroke cycle SI engines add HC emissions to the exhaust during the scav-
enging process. The air–fuel intake mixture is used to push exhaust residual out the
open exhaust port. When this is done, some of the air and fuel mixes with the exhaust
gases and is expelled out of the cylinder before the exhaust port closes. This can be a
major source of HC in the exhaust and is one of the major reasons that there have been
no modern two-stroke cycle automobile engines. They could not meet antipollution re-
quirements. Some experimental automobile two-stroke cycle engines and just about all
small engines use crankcase compression, and this is a second source of hydrocarbon
emissions. The crankcase area and pistons of these engines are lubricated by adding oil
to the inlet air–fuel mixture. The oil is vaporized with the fuel and lubricates the sur-
faces that come in contact with the air–fuel–oil mixture. Some of the oil vapor is car-
ried into the combustion chamber and burned with the air–fuel mixture. Lubricating
oil is composed mostly of hydrocarbon components and acts like additional fuel. How-
ever, due to the high molecular weight of its components, oil does not fully combust as
readily as fuel, and this adds to HC emissions in the exhaust.
Modern experimental two-stroke cycle automobile engines do not add fuel to the
intake air, but scavenge the cylinders with pure air, avoiding the placement of HC into
the exhaust. After the exhaust port closes, fuel is added by fuel injection directly into
the cylinder. This creates a need for very fast and efficient vaporization and mixing of
the air and fuel, but it eliminates a major source of HC emissions. Some automobile en-
gines use superchargers instead of crankcase compression, and this eliminates HC pol-
lution from that source.
352
Emissions and Air Pollution
Until recently, most small engines, such as those used in lawn mowers and boats,
were not regulated for pollution control. Many of these engines are still being manu-
factured with uncontrolled scavenging and oil vapor lubrication, contributing to seri-
ous HC (and other) pollution. This problem is starting to be addressed, and in some
parts of the world (starting in California) emission laws and standards are starting to
be applied to lawn mowers, boats, and other small engines. This will probably phase
out, or at least greatly reduce, the number of small two-stroke cycle engines. Low cost
is a major requirement for small engines, and fuel injection systems are much more
costly than the very simple carburetors found on older engines. Many small engines
now operate on a cleaner four-stroke cycle, but still use a less costly carburetor for fuel
input.
In the 1990s, there were an estimated 83 million lawn mowers in the United
States, producing as much air pollution as 3.5 million automobiles. Government studies
of equipment using small engines give the following pollution comparison between
that equipment and automobiles (the numbers represent one hour of operation in
terms of miles traveled in an average automobile):
353
Emissions and Air Pollution
fuel-lean zones, combustion is limited and some fuel does not get burned. With over-
mixing, some fuel particles will be mixed with already burned gas and will therefore
not combust totally.
It is important that injectors be constructed such that when injection stops, there
is a minimum of dribble from the nozzle. A small amount of liquid fuel will be trapped
on the tip of the nozzle, however. This very small volume of fuel is called sac volume, its
size depending on the nozzle design. This sac volume of liquid fuel evaporates very
slowly because it is surrounded by a fuel-rich environment, and, once the injector noz-
zle closes, there is no pressure pushing it into the cylinder. Some of this fuel does not
evaporate until combustion has stopped, and this results in added HC particles in the
exhaust.
CI engines also have HC emissions for some of the same reasons as SI engines do
(i.e., wall deposit absorption, oil film absorption, crevice volume, etc.).
Example Problem 1
As the flame front reaches the wall of a combustion chamber, reaction stops due to the closeness
of the wall, which dampens out all fluid motion and conducts heat away. This unburned bound-
ary layer can be considered a volume 0.1 mm thick along the entire combustion chamber surface.
The combustion chamber consists mainly of a bowl in the face of the piston, which can be ap-
proximated as a 3-cm-diameter hemisphere. Fuel is originally distributed equally throughout the
chamber. Calculate the percentage of fuel that does not get burned due to being trapped in the
surface boundary layer.
Volume of the combustion chamber is
354
Emissions and Air Pollution
carbon monoxide. Not only is CO considered an undesirable emission, but it also rep-
resents lost chemical energy that was not fully utilized in the engine. CO is a fuel that
can be combusted to supply additional thermal energy:
O + N2 : NO + N (2)
N + O2 : NO + O (3)
N + OH : NO + H (4)
NO, in turn, can then further react to form NO2 by various means, including the
following:
NO + H 2O : NO2 + H 2 (5)
NO + O2 : NO2 + O (6)
N2 : 2 N (7)
Table 3 in the Appendix shows that the chemical equilibrium constant for Eq.
(7) is highly dependent on temperature, with a much more significant amount of N
355
Emissions and Air Pollution
generated in the 2500–3000 K temperature range that can exist in an engine. Other
gases that are stable at low temperatures, but become reactive and contribute to the
formation of NOx at high temperatures, include oxygen and water vapor, which break
down as follows:
O2 : 2 O (8)
1
H 2O : OH + 2 H2 (9)
O + O2 : O3 (11)
356
Emissions and Air Pollution
⫽ 0.95
2
⫽ 1.00
⫽ 0.70
0
0 0.002 0.004 0.006 0.008 0.010
Time (sec)
FIGURE 4
Generation of NOx in an engine as a function of combustion time. Many modern engines
produce lower NOx emissions due to fast-burn combustion chamber design. Adapted
from [92].
atmospheric reactions with other engine emissions such as HC, aldehydes, and other
oxides of nitrogen.
Example Problem 2
To reduce the amount of reactive monatomic nitrogen in the engine, EGR is added to the engine
operation, which reduces cylinder temperature at the end of combustion from 3500 K to 2500 K.
Calculate the percent reduction of the approximate amount of N, assuming pressure remains the
same.
Chemical equilibrium constant Ke is obtained from Table 3 in the Appendix at
T = 2500 K:
then
357
Emissions and Air Pollution
0.3
0.2
0.1
FIGURE 5 Stoichiometric
⫽ 1.0
Generation of NOx in an SI engine as a
function of spark timing. Earlier spark
ignition creates a higher combustion
temperature, which generates higher 50 40 30 20 10 0
levels of NOx. Adapted from [108]. Spark Timing (degrees bTDC)
or
5 PARTICULATES
The exhaust of CI engines contains solid carbon soot particles that are generated in the
fuel-rich zones within the cylinder during combustion. These are seen as exhaust
smoke and are an undesirable odorous pollution. Maximum density of particulate
emissions occurs when the engine is under load at WOT. At this condition, maximum
fuel is injected to supply maximum power, resulting in a rich mixture and poor fuel
economy. This can be seen in the heavy exhaust smoke emitted when a truck or rail-
road locomotive accelerates up a hill or from a stop.
358
Emissions and Air Pollution
Soot particles are clusters of solid carbon spheres. These spheres have diameters
from 10 nm to 80 nm 11 nm = 10 -9 m2, with most within the range of 15–30 nm. The
spheres are solid carbon with HC and traces of other components absorbed on the sur-
face. A single soot particle will contain up to 4000 carbon spheres [58].
Carbon spheres are generated in the combustion chamber in the fuel-rich zones
where there is not enough oxygen to convert all carbon to CO2:
C xH y + z O2 : a CO2 + b H 2O + c CO + d C1s2 (12)
Then, as turbulence and mass motion continue to mix the components in the
combustion chamber, most of these carbon particles find sufficient oxygen to further
react and are consumed to CO2:
C1s2 + O2 : CO2 (13)
Over 90% of carbon particles originally generated within an engine are thus con-
sumed and never get exhausted (Fig. 6). If CI engines were to operate with an overall
stoichiometric air–fuel mixture, instead of the overall lean mixture they do operate
with, particulate emissions in the exhaust would far exceed acceptable levels.
Up to about 25% of the carbon in soot comes from lubricating–oil components,
which vaporize and then react during combustion. The rest comes from the fuel and
amounts to 0.2–0.5% of the fuel. Because of the high compression ratios of CI engines,
a large expansion occurs during the power stroke, and the gases within the cylinder are
cooled by expansion cooling to a relatively low temperature. This causes the remaining
high-boiling-point components found in the fuel and lubricating oil to condense on the
surface of the carbon soot particles. This absorbed portion of the soot particles is called
the soluble organic fraction (SOF), and the amount is highly dependent on cylinder
temperature. At light loads, cylinder temperatures are reduced and can drop to as low
as 200°C during final expansion and exhaust blowdown. Under these conditions, SOF
can be as high as 50% of the total mass of soot. Under other operating conditions when
temperatures are not so low, very little condensing occurs and SOF can be as low as 3%
of total soot mass. SOF consists mostly of hydrocarbon components with some hydro-
gen, SO2, NO, NO2, and trace amounts of sulfur, zinc, phosphorus, calcium, iron, silicon,
and chromium. Diesel fuel contains sulfur, calcium, iron, silicon, and chromium, while
lubricating–oil additives contain zinc, phosphorus, and calcium.
359
Emissions and Air Pollution
10
1 35
Exhaust level
FIGURE 6
Particulate concentrations at various
locations in combustion chamber during 0.5
combustion and power stroke of large 23
direct injection diesel engine. A large
percentage of the particulates generated
early in the combustion process is
consumed later during the latter part of
17
the combustion and power stroke. Only a
small percentage of the original 0.1 29
particulates is exhausted from the
combustion chamber; the level is shown at
the right of the figure. Engine running at 0.05
500 RPM with compression ratio of 12.9.
Reprinted with permission from SAE
Paper No. 820464 © 1982 SAE TDC 20 40 60 80 100
International [202]. Crank angle, deg.
360
Emissions and Air Pollution
0
0 200 400 600 800 1000 1200
NOx (ppm)
FIGURE 7
Nitrogen oxide (NOx)–smoke (particulates) trade-off at various engine operating
conditions. Reprinted with permission from SAE Paper No. 811234 © 1981 SAE
International [240].
6 OTHER EMISSIONS
Carbon Dioxide (CO2)
At moderate levels of concentration, carbon dioxide is not considered an air pollutant.
However, it is considered a major greenhouse gas and, at higher concentrations, is a major
contributor to global warming. CO2 is a major component of the exhaust in the combus-
tion of any hydrocarbon fuel. Because of the growing number of motor vehicles, along
361
Emissions and Air Pollution
with more factories and other sources, the amount of carbon dioxide in the atmosphere
continues to grow. At upper elevations in the atmosphere, this higher concentration of
carbon dioxide, along with other greenhouse gases, creates a thermal radiation shield.
This shield reduces the amount of thermal radiation energy allowed to escape from the
earth, raising slightly the average earth temperature. The most efficient way of reducing
the amount of CO2 is to burn less fuel (i.e., use engines with higher thermal efficiency).
Aldehydes
A major emission problem when alcohol fuel is used is the generation of aldehydes, an
eye and respiratory irritant. These have the chemical formula
H
ƒ
R¬ C “O
Sulfur
Many fuels used in CI engines contain small amounts of sulfur, which, when exhausted,
contributes to the acid rain problem of the world. Unleaded gasoline generally con-
tains 150–600 ppm sulfur by weight. Some diesel fuels contain up to 5000 ppm by
weight, but in the United States and some other countries sulfur content is restricted
by law to a tenth of this value or less.
At high temperatures, sulfur combines with hydrogen to form H 2S and with oxy-
gen to form SO2:
H 2 + S : H 2S (14)
O2 + S : SO2 (15)
Engine exhaust can contain up to 20 ppm of SO2. SO2 then combines with oxygen
in the air to form SO3:
These molecules combine with water vapor in the atmosphere to form sulfuric
acid 1H 2SO42 and sulfurous acid 1H 2SO32, which are ingredients in acid rain:
Many countries have laws restricting the amount of sulfur allowed in fuel, and
these laws are continuously being made more stringent. During the 1990s, the United
States reduced acceptable sulfur levels in diesel fuel from 0.05% by weight to 0.01%.
362
Emissions and Air Pollution
The amount of sulfur in natural gas can range from small (sweet) to large (sour)
amounts. This can be a major emissions problem when this fuel is used in an IC engine
or any other combustion system.
When the allowable sulfur level in diesel fuel was lowered, a new problem sur-
faced in CI engines. It was found that fuel with very low levels of sulfur lost its lubri-
cating ability, resulting in sticking fuel pumps and injectors [184]. In addition, there was
abnormal wear on cylinder surfaces and rapid pressure buildup in some particulate
traps. To overcome these problems, additives are put into low–sulfur fuels. These addi-
tives include aliphatic ester derivatives and carboxylic acids.
A more serious effect of sulfur, in addition to being a harmful emission, is that it
poisons most emissions aftertreatment systems. Catalyst materials in catalytic convert-
ers and regenerating particulate traps deteriorate in the presence of sulfur, lead, or
phosphorus.
Lead
Lead was a major gasoline additive from its introduction in 1923 to when it was phased
out in the 1980s. The additive TEL (tetraethyl lead) was effectively used to increase
gasoline octane number, which allowed higher compression ratios and more efficient
engines. However, the resulting lead in the engine exhaust was a highly poisonous pol-
lutant. During the first half of the 1900s, due to the lower number of automobiles and
other engines, the atmosphere was able to absorb these emissions of lead without no-
ticeable problems. As population and automobile density increased, the awareness of
air pollution and its danger also increased. The danger of lead emissions was recog-
nized, and a phaseout occurred during the 1970s and 1980s.
The use of lead could not be stopped immediately, but had to be phased out over
a number of years. First, low-lead gasoline was introduced, followed, years later, by no-
lead gasoline. Lead was still the major additive to raise the octane number of gasoline,
and alternate octane raisers had to be developed as lead was phased out. Millions of
modern high-compression engines could not use low-octane fuel. Metals used in en-
gines also had to be changed as lead in gasoline was phased out. When leaded fuel is
burned, it hardens the surfaces in the combustion chamber and those of the valves and
valve seats. Engines designed to use leaded fuel had softer metal surfaces initially and
relied on surface hardening effects that occurred during use. If these engines are used
with unleaded fuel, surface hardening is not realized and serious wear is quickly expe-
rienced. Catastrophic failures of valve seats or piston faces are common in a short peri-
od of time (i.e., 10,000–20,000 miles in an automobile). Harder metals and added
surface treatments are used for engines designed to use unleaded fuel. It was necessary
to phase out leaded gasoline over a period of time as older automobiles wore out and
were taken out of operation.
Leaded gasoline contains about 0.15 gm/liter of lead in the fuel. Between 10%
and 50% of this gets exhausted out with the other combustion products. The remaining
lead gets deposited on the walls of the engine and exhaust system. The hardened com-
bustion chamber surfaces that resulted from the burning of leaded gasoline were quite
impervious to the absorption of gases such as fuel vapor. HC emissions were also,
therefore, slightly reduced in these engines.
363
Emissions and Air Pollution
Phosphorus
Small amounts of phosphorus are emitted in engine exhaust. These come from impuri-
ties in the air and small amounts found in some fuel blends and lubricating oil. In addi-
tion to being an air pollutant, phosphorus poisons the catalytic materials in catalytic
converters.
7 AFTERTREATMENT
After the combustion process stops, those components in the cylinder gas mixture that
have not fully burned continue to react during the expansion stroke, during exhaust
blowdown, and into the exhaust process. Up to 90% of the HC remaining after com-
bustion reacts during this time either in the cylinder, near the exhaust port, or in the
upstream part of the exhaust manifold. CO and small component hydrocarbons react
with oxygen to form CO2 and H 2O and reduce undesirable emissions. The higher the
exhaust temperature, the more these secondary reactions occur and the lower are the
engine emissions. Higher exhaust temperature can be caused by stoichiometric air–fuel
combustion, high engine speed, retarded spark, and/or a low expansion ratio.
Thermal Converters
Secondary reactions occur much more readily and completely if the temperature is high,
so some engines are equipped with thermal converters as a means of lowering emis-
sions. Thermal converters are high-temperature chambers through which the exhaust
gas flows. They promote oxidation of the CO and HC which remain in the exhaust. For
the CO, we have
CO + 12 O2 : CO2 (19)
For this reaction to occur at a useful rate, the temperature must be held above 700°C
[58]. For HC, we have
C xH y + z O2 : x CO2 + 12 y H 2O (20)
1
where z = x + 4 y.This reaction needs a temperature above 600°C for at least 50 msec
to substantially reduce HC. It is therefore necessary for a thermal converter not only to
operate at a high temperature, but to be large enough to provide adequate dwell time
to promote the occurrence of these secondary reactions. Most thermal converters are
essentially enlarged exhaust manifolds connected to the engine immediately outside the
364
Emissions and Air Pollution
exhaust ports. This location is necessary to minimize heat losses and keep the exhaust
gases from cooling to nonreacting temperatures. However, in automobiles, this creates
two very serious problems for the engine compartment. In modern, low-profile, aero-
dynamic automobiles, space in the engine compartment is very limited, and fitting in a
large, usually insulated, thermal converter chamber is almost impossible. Secondly, be-
cause the converter must operate above 700°C to be efficient, even if it is insulated the
heat losses create a serious temperature problem in the engine compartment.
Some thermal converter systems include an air intake that provides additional
oxygen to react with the CO and HC. This increases the complexity, cost, and size of the
system. Flow rate of air is controlled by the EMS as needed. Air addition is especially
necessary during rich operating conditions such as start-up. Because exhaust from en-
gines is often at a lower temperature than is needed for efficient operation of a thermal
converter, it is necessary to sustain the high temperatures by the reactions within the
system. Adding outside air, which is at a lower temperature, compounds this problem
of maintaining the necessary operating temperature.
NOx emissions cannot be reduced with a thermal converter alone.
8 CATALYTIC CONVERTERS
The most effective aftertreatment system for reducing engine emissions is the catalytic
converter found on most automobiles and other modern engines of medium or large
size. HC and CO can be oxidized to H 2O and CO2 in exhaust systems and thermal con-
verters if the temperature is held at 600°–700°C. If certain catalysts are present, the
temperature needed to sustain these oxidation processes is reduced to 250°–300°C,
making for a much more attractive system. A catalyst is a substance that accelerates a
chemical reaction by lowering the energy needed for it to proceed. The catalyst is not
consumed in the reaction and so functions indefinitely unless degraded by heat, age,
contaminants, or other factors. Catalytic converters are chambers mounted in the flow
system through which the exhaust gases flow. These chambers contain catalytic mater-
ial, which promotes the oxidation of the emissions contained in the exhaust flow.
Generally, catalytic converters are called three-way converters because they pro-
mote the reduction of CO, HC, and NOx. There are basically two types of construction.
Some consist of a stainless steel container mounted somewhere along the exhaust pipe
of the engine. Inside the container is a porous ceramic structure through which the gas
flows. In most converters, the ceramic is a single honeycomb structure with many
flow passages (see Fig. 8). Some converters use loose granular ceramic with the gas
passing between the packed spheres. The volume of the ceramic structure of a convert-
er is generally about half the displacement volume of the engine. This results in a volu-
metric flow rate of exhaust gas such that there are 5 to 30 changeovers of gas each
second through the converter. Catalytic converters for CI engines need larger flow pas-
sages because of the solid soot in the exhaust gases.
The surface of the ceramic passages contains small embedded particles of cat-
alytic material that promote the oxidation reactions in the exhaust gas as it passes. Alu-
minum oxide (alumina) is the base ceramic material used for most catalytic converters.
Alumina can withstand the high temperatures, it remains chemically neutral, it has
very low thermal expansion, and it does not thermally degrade with age. The catalyst
365
Emissions and Air Pollution
INSULATION
CONVERTER SHELL
OUTER WRAP
RETAINER
OUTER WRAP
FILL PLUG INSULATION
CATALYST
METAL MESH
SEAL SHELL
FIGURE 8 CATALYST
Catalytic converters for SI engines:
(a) packed spheres; (b) honeycomb
structure. Reprinted with permission from
SAE Paper 801440 © 1980 SAE
International [4]. (b)
materials most commonly used are platinum and rhodium, with other noble metals
(e.g., palladium and iridium) sometimes used as well.
Palladium and platinum promote the oxidation of CO and HC as in Eqs. (19) and
(20), with platinum especially active in the hydrocarbon reaction. Rhodium promotes
the reaction of NOx in one or more of the following reactions:
NO + CO : 12 N2 + CO2 (21)
2 NO + 5 CO + 3 H 2O : 2 NH 3 + 5 CO2 (22)
2 NO + CO : N2O + CO2 (23)
1
NO + H 2 : 2 N2 + H 2O (24)
2 NO + 5 H 2 : 2 NH 3 + 2 H 2O (25)
2 NO + H 2 : N2O + H 2O (26)
366
Emissions and Air Pollution
FIGURE 9
Cutaway drawing of three-way catalytic converter system, showing main components. This system was
used on Ford automobiles during the 1990s. Courtesy Ford Motor Company.
Also often used is cerium oxide, which promotes the so-called water–gas shift:
CO + H 2O : CO2 + H 2 (27)
This reduces CO by using water vapor as an oxidant instead of O2, which is very
important when the engine is running in a rich manner.
Instead of ceramics, the interior chamber of some catalytic converters is filled
with a very thin corrugated metal foil wrapped into a monolith structure. The exhaust
gases pass between the rolls of foil, the surface of which is embedded with catalytic ma-
terial. The active surface area of a larger converter of this type can be as great as
70,000 m2.
Figure 10 shows that the efficiency of a catalytic converter is very dependent on
temperature. When a converter in good working order is operating at a fully warmed
temperature of 400°C or above, it will remove 98–99% of CO, 95% of NOx, and more
than 95% of HC from exhaust flow emissions. Figure 11 shows that it is also necessary
to be operating at the proper equivalence ratio to get high converter efficiency. Effec-
tive control of HC and CO occurs with stoichiometric or lean mixtures, while control of
NOx requires near–stoichiometric conditions. Very poor NOx control occurs with lean
mixtures.
Because an engine has a number of cyclic variations occurring—including AF—
the exhaust flow will also show variation. It has been found that this cyclic variation
lowers the peak efficiency of a catalytic converter, but spreads the width of the equiva-
lence–ratio envelope of operation in which there is acceptable emissions reduction.
367
Emissions and Air Pollution
100
NOx
80
60 CO
40
FIGURE 10
Conversion efficiency of catalytic converters Light-Off
as a function of converter temperature. Temperature
A converter in good condition will generally 20
reduce emissions by over 90% if it is at
normal operating temperature. When cold, a
converter is very inefficient. The
temperature at which a converter becomes
50% efficient is often called light-off 100 200 300 400
temperature. Adapted from [4]. Temperature (˚C)
100
Catalytic Converter Efficiency (%)
HC
80
60
NOx CO
40
Lean Rich
20
0
0 0.96 0.97 0.98 0.99 1.00 1.01 1.02 1.03 1.04 1.05
Equivalence Ratio,
FIGURE 11
Conversion efficiency of catalytic converters as a function of fuel equivalence ratio.
Greatest efficiency occurs when engines operate near stoichiometric conditions.
Converters are very inefficient for NOx conversion when an engine operates in a lean
manner. This creates a greater problem for modern CI engines and stratified charge SI
engines, which generally operate in a very lean manner overall. Adapted from [76].
368
Emissions and Air Pollution
Sulfur
Sulfur offers unique problems for catalytic converters. Some catalysts promote the con-
version of SO2 to SO3, which eventually gets converted to sulfuric acid. This degrades
100
Catalytic Converter Efficiency for HC (%)
80
60
FIGURE 12
Reduction of catalytic converter efficiency
due to contamination by lead. Some of the
40
lead contained in fuels gets deposited on
the catalyst material in a converter, greatly
reducing converter efficiency. It is
imperative (and legally required) that
20 leaded gasoline not be used in automobiles
equipped with catalytic converters. To
reduce the chances of accidently using
leaded gasoline with a catalytic converter,
the fuel pump nozzle size and the diameter
0 2 4 6 8 10 of the fuel tank inlet are made smaller for
Lead Contaminant on Catalyst (mass %) nonleaded gasoline. Adapted from [76].
369
Emissions and Air Pollution
the catalytic converter and contributes to acid rain. New catalysts are being developed
that promote the oxidation of HC and CO, but do not change SO2 to SO3. Some of
these create almost no SO3 if the temperature of the converter is kept at 400°C or lower.
Cold Start-Ups
Figure 10 shows that catalytic converters are very inefficient when they are cold. When
an engine is started after not being operated for several hours, it takes several minutes
for the converter to reach an efficient operating temperature. The temperature at
which a converter becomes 50% efficient is defined as the light-off temperature, and
this is in the range of about 250°–300°C. A large percentage of automobile travel is for
short distances where the catalytic converter never reaches efficient operating temper-
ature, and therefore emissions are high. Some studies suggest that half of the fuel used
by automobiles in the United States is on trips of less than 10 miles distance. Unfortu-
nately, most short trips occur in cities where high emissions are more harmful. Add to
this the fact that most engines use a rich mixture when starting, and it can be seen that
cold start-ups pose a major problem. It is estimated that cold start-ups are the source of
70–90% of all HC emissions. A major reduction in emissions would therefore be possi-
ble if catalytic converters could be preheated, at least to light-off temperature, before
engine start-up. Preheating to full steady-state operating temperature would be even
better. Several methods of preheating have been tried with varying success. Because of
the time involved and amount of energy needed, most of these methods preheat only a
small portion of the total converter volume. This small section is large enough to treat
the low exhaust flow rate that usually occurs at start-up and immediately following. By
the time higher engine speeds are used, more of the catalytic converter has been heat-
ed by the hot exhaust gas, and the higher flow rates are fully treated. Methods of cat-
alytic converter preheating include those discussed in the subsections that follow.
Locate Converter Close to the Engine. One method used to heat a converter as
quickly as possible is to locate it in the engine compartment very close to the exhaust
ports. This method does not actually preheat the converter, but does heat it as quickly
as possible after the engine is started. It eliminates the large heat loss from the exhaust
pipe that occurs between the engine and a converter in more common systems where
the converter is located away from the engine. These converters can also be insulated
to reduce early heat loss. This method does reduce overall emissions by quick heat-up
of the converter, but there is still a short period of time before light-off temperature is
reached. In addition, the same problems described for thermal converters mounted in
the engine compartment are encountered with this type of converter. Adequate cool-
ing of the engine compartment because of the high temperatures and restricted flow
rate of air caused by the converter is a serious problem. If located in the hot engine
compartment, a catalytic converter will also have a higher steady-state temperature,
and this will cause a greater long-term thermal–degrading problem.
Some automobiles use a small secondary catalytic converter mounted in the en-
gine compartment close to the engine. Because of its small size and location, it heats
up very quickly and is sufficient to oxidize the emissions in the low flow rates at engine
start-up. There is also a normal full-size catalytic converter mounted away from the
370
Emissions and Air Pollution
engine compartment that supplies the catalytic action for the larger flow rates of nor-
mal operation. This converter is heated by the first exhaust flow and ideally reaches
efficient operating temperature before the engine is speeded up and higher flow rates
are experienced. These small preconverters restrict flow in the exhaust manifold and
add some back pressure to the engine. This results in a slight reduction in engine
power output.
Flame Heating. A catalytic converter can be heated with a flame from a burner
nozzle mounted within the structure of the converter [57]. Before the engine is started
(for instance, when the ignition key is inserted), a flame is initiated in the burner, using
fuel and air pumped from external sources. Concern must be given to what emissions this
flame would contribute to the overall air pollution problem. A fuel like propane burned
with the correct amount of air would create very little pollution. However, this would re-
quire an axillary propane fuel tank on the automobile, something that would be undesir-
able unless the automobile engine were also fueled with propane. In a gasoline-fueled
engine, it would be logical to use gasoline in the converter preheater. However, it would
371
Emissions and Air Pollution
be more difficult to get clean burning with gasoline. Cost, complexity, and some time
delay are disadvantages of this type of system.
A variation of this system used by at least one major automobile manufacturer is
an afterburner mounted directly before the catalytic converter. A very rich air–fuel
mixture is used at start-up, which leaves excess fuel in the first exhaust flow. Air is
added to this exhaust by an electric pump, and the resulting mixture is combusted in
the afterburner, preheating the catalytic converter.
Thermal Battery. Energy from a thermal storage system can be used to preheat
a catalytic converter if the engine is started within about three days of last being used.
With present technology, only partial preheating to a temperature around 60°C is pos-
sible, which is still below light-off temperature and well below normal operating tem-
peratures. In addition, the limited amount of energy available in a thermal battery is
often distributed between preheating the engine, warming the passenger compart-
ment, and preheating the catalytic converter.
Dual-Fuel Engines
Some engines are made to run on a combination of gasoline and methanol, with the
percent volume of methanol ranging from 0% to 85%. The engine control systems on
these engines are capable of adjusting the air and fuel flow to give optimum combus-
tion and minimum emissions with any combination of these fuels. However, this cre-
ates a unique problem for a catalytic converter. Each of these fuels requires separate
catalysts. Incomplete combustion of methanol produces formaldehyde, which must be
removed from the exhaust. To effectively reduce the formaldehyde and any remaining
methanol, a catalytic converter must be operated above 300°C. Preheating of the con-
verter on these systems is very important.
372
Emissions and Air Pollution
Lean-Burn Engines
A large number of automobiles on the market obtain high fuel efficiency by use of lean-
burn engines. Lean-burn combustion is becoming a common philosophy for reducing
fuel consumption, but it creates special problems in reducing NOx in catalytic convert-
ers. By using a stratified charge, engines obtain efficient combustion with overall
air–fuel ratios of 20 or 21 1f L 0.72, with some operating as high as AF = 40 1f L 0.42.
Figure 11 shows that normal catalytic converters will work in reducing HC and CO at
these lean conditions, but are very inefficient at reducing NOx.
Some automobiles using lean-burn engines have special converters with interior
surfaces that absorb the NOx which does not get treated when operating lean. Then,
when the vehicle accelerates, works under load, or operates in any mode requiring
stoichiometric combustion, the absorbed NOx desorbs off the surface and is treated
with the high converter efficiency that occurs under these conditions. When the auto-
mobile does not accelerate for several minutes, the EMS is programed to periodically
inject a few seconds of rich combustion to desorb the catalytic converter. For instance,
when the Toyota Opa is traveling at a steady 60 km/hr (37 mph), one to two seconds of
rich injection occurs every two minutes [239]. Another method used to eliminate this
lean-burn problem is to use high levels of EGR, which helps in two ways. It dilutes the
air–fuel mixture and lowers the combustion temperature, which is what normal lean-
burn does. The lower combustion temperature then generates less NOx. In addition, it
allows the actual air and fuel to be added in a near-stoichiometric ratio, which allows
the catalytic converter to operate efficiently. Catalysts using platinum, rhodium, palla-
dium, iridium, and other noble metals, combined with alkaline rare earths, have been
developed for lean-burn engines.
373
Emissions and Air Pollution
9 CI ENGINES
Catalytic converters are used with CI engines, but are not efficient at reducing NOx,
due to their overall lean operation. HC and CO can be adequately reduced, although
there is greater difficulty because of the cooler exhaust gases of a CI engine (because
of the larger expansion ratio). This is counterbalanced by the fact that less HC and CO
are generated in the lean burn of the CI engine. NOx is reduced in a CI engine by the
use of EGR, which keeps the maximum temperature down. EGR and lower combus-
tion temperatures, however, contribute to an increase in solid soot.
Platinum and palladium are two main catalyst materials used for converters on
CI engines. They promote the removal of 30–80% of the gaseous HC and 40–90% of
the CO in the exhaust. The catalysts have little effect on solid carbon soot, but do re-
move 30–60% of the total particulate mass by oxidizing a large percent, of the HC ab-
sorbed on the carbon particles. Diesel fuel contains sulfur impurities, and this leads to
poisoning of the catalyst materials. However, this problem is being reduced as legal lev-
els of sulfur in diesel fuels continue to be lowered.
Particulate Traps
Compression ignition engine systems are equipped with particulate traps in their ex-
haust flow to reduce the amount of particulates released to the atmosphere. Traps are
filterlike systems often made of ceramic in the form of a monolith or mat, or some-
times made of metal wire mesh. Traps typically remove 60–90% of particulates in the
exhaust flow. As traps catch the soot particles, they slowly fill up with the particulates.
This restricts exhaust gas flow and raises the back pressure of the engine. Higher back
pressure causes the engine to run hotter, the exhaust temperature to rise, and fuel con-
sumption to increase. To reduce this flow restriction, particulate traps are regenerated
when they begin to become saturated. Regeneration consists of combusting the partic-
ulates in the excess oxygen contained in the exhaust of the lean-operating CI engine.
Carbon soot ignites at about 550°–650°C, while CI engine exhaust temperature is
150°–350°C at normal operating conditions. As the particulate trap fills with soot and
restricts flow, the exhaust temperature rises, but is still not high enough to ignite the
soot and regenerate the trap. In some systems, automatic flame igniters are used. These
igniters start combustion in the carbon when the pressure drop across the trap reaches
a predetermined value. The igniters can be electric heaters or flame nozzles that use
diesel fuel. If catalyst material is installed in the traps, the temperature needed to ignite
the carbon soot is reduced to the 350°–450°C range. Some such traps can automatical-
ly regenerate by self-igniting when the exhaust temperature rises from increased back
pressure. Other catalyst systems use flame igniters.
Another way of lowering the ignition temperature of the carbon soot and pro-
moting self-regeneration in traps is to use catalyst additives in the diesel fuel. These ad-
ditives generally consist of copper compounds or iron compounds, with about seven
grams of additive in 1000 liters of fuel being normal.
To keep the temperatures high enough for self-regeneration in a catalytic system,
traps can be mounted as close to the engine as possible, even before the turbocharger.
On some larger stationary engines and on some construction equipment and large
trucks, the particulate trap is replaced when it becomes nearly filled. The removed trap
374
Emissions and Air Pollution
is then regenerated externally, with the carbon being burned off in a furnace. The regen-
erated trap can then be used again.
Various methods are used to determine when soot buildup becomes excessive
and regeneration is necessary. The most common method is to measure pressure drop
in the exhaust flow as it passes through the trap. When a predetermined ¢P is reached,
regeneration is initiated. Pressure drop is also a function of exhaust flow rate, and this
must be programmed into the regeneration controls. Another method used to sense
soot buildup is to transmit radio frequency waves through the trap and determine the
percentage of the signal that is absorbed. Carbon soot absorbs radio waves, while the
ceramic structure does not. The amount of soot buildup can therefore be determined
by the percent decrease in radio signal. This method does not readily detect soluble or-
ganic fraction (SOF).
Modern particulate traps are not totally satisfactory, especially for automobiles.
They are costly and complex when equipped for regeneration, and long-term dura-
bility does not exist. An ideal catalytic trap would be simple, economical, and reli-
able; it would be self-regenerating; and it would impose a minimum increase in fuel
consumption.
375
Emissions and Air Pollution
determined over a range of operating variables, including AF, temperature, flow veloc-
ity, and zeolite structure. At present, durability is a serious limitation with this method.
Various chemical absorbers, molecular sieves, and traps are being tested to re-
duce HC emissions. HC is collected during engine start-up time, when the catalytic
converter is cold, and then later released back into the exhaust flow when the convert-
er is hot. The converter then efficiently burns the HC to H 2O and CO2. A 35% reduc-
tion of cold-start HC has been achieved.
H 2S emissions occur under rich operating conditions. Chemical systems are being
developed that trap and store H 2S when an engine operates under rich conditions and
then convert this to SO2 when operation is lean and excess oxygen exists. The reaction
equation is
H 2S + O2 : SO2 + H 2 (28)
4 NH 3 + 4 NO + O2 : 4 N2 + 6 H 2O (29)
6 NO2 + 8 NH 3 : 7 N2 + 12 H 2O (30)
376
Emissions and Air Pollution
0.20
Mole Fraction of NOx in Exhaust (%)
0.15
0.10
Argon
0 5 10 15 20 25 30
Diluent in Intake (mole %)
FIGURE 13
Reduction of NOx generation by adding non–combustible diluent gas to intake
mixture. Adding any nonreacting neutral gas to the inlet air–fuel mixture reduces
flame temperature and NOx generation. Exhaust gas is the one gas that is readily
available for engine use. Adapted from [99].
377
Emissions and Air Pollution
After EGR combines with the exhaust residual left from the previous cycle, the
total fraction of exhaust in the cylinder during the compression stroke is
Example Problem 3
The theoretical maximum combustion temperature in an engine burning isooctane at an equiva-
lence ratio of 0.833 is 2419 K. To reduce formation of NOx, it is desired to reduce this maximum
temperature to 2200 K. This is done by exhaust gas recycling (EGR). Calculate the amount of
EGR needed to reduce maximum combustion temperature to 2200 K.
Exhaust gas, which consists mostly of N2, CO2, and H 2O, will be approximated as all ni-
trogen at a temperature of 1000 K. Enthalpy values can be obtained from most thermodynamics
textbooks [90].
The combustion equation is
378
Emissions and Air Pollution
12 NONEXHAUST EMISSIONS
Engines and fuel supply systems also have sources of emissions other than exhaust flow.
Historically, these were considered minor and were just released to the surrounding air.
A major source of HC emissions was the crankcase breather tube that was vent-
ed to the air in older automobiles. Blowby flow past the pistons ended up in the
crankcase, and due to the higher pressure it created, it was then pushed out the
breather vent tube. Blowby gas is very high in HCs, especially in SI engines. Also, in
older engines with greater clearance between the piston and cylinder wall, blowby flow
was much higher. As much as 1% of the fuel was vented to the atmosphere through the
crankcase breather in some automobiles. This accounted for up to 20% of total emis-
sions. A simple solution to this problem, which is used on all modern engines, is to vent
the crankcase breather back into the intake system. This not only reduces emissions,
but also increases fuel economy.
To keep the pressure at one atmosphere in the fuel tank and in the fuel reservoir
of a carburetor, these systems are vented to the surroundings. Historically, these vents
were an additional source of HC emissions when fuel evaporated from these fuel
reservoirs. To eliminate these emissions, fuel vents now include some form of filter or
absorption system that stops the HC vapor from escaping. One such system absorbs
the HCs onto the surface of a carbon filter element. Then, when the engine is operat-
ing, the element is back flushed and the HC is desorbed off the surface. The recovered
HC is ducted into the engine intake with no resulting emissions.
Many modern gasoline pumps and other fuel-dispensing systems are equipped with
vapor-collecting nozzles that reduce HC vapor lost to the atmosphere during refueling.
379
Emissions and Air Pollution
13 NOISE POLLUTION
Since the 1990s, noise generated by engines and other systems has been considered a
pollution. Noise is sometimes defined as “undesirable sound,” and at high levels is rec-
ognized as a possible health hazard. Large engines produce high levels of sound, and
many countries now have laws governing acceptable levels of noise allowed in closed
engine rooms of ships and in stationary applications. Engine noise from automobiles is
less of a problem, and existing technologies are adequate for controlling it.
Sound is caused by pressure waves in an elastic medium (i.e., air) generated by vi-
brations of the engine components. These vibrations cause pressure pulses in the air,
and these pulses transfer energy to the human ear; the greater the energy, the louder is
the sound. Fortunately, the ear is not very sensitive, and so the quantifying scale is a
logarithmic scale with units of decibels (dB). Normal conversation has a sound level of
about 55 dB, while the ear begins to feel pain at about 120 dB. Many engine room codes
allow noise up to 110 dB. The sensitivity of the human ear is closely related to sound
frequency, being less sensitive at low frequency. For this reason, international standards
are divided into three categories—A, B, and C—each related to a range of frequencies.
In the United States the EPA has an acceptable level of drive-by noise for vehicles of
74 dB (A). This standard often must be considered in automobile design in areas such
as tailpipe and muffler placement.
Noise is generated in many ways in an operating IC engine: pressure pulsations in
the gaseous flow of intake and exhaust, fuel injectors, superchargers, chain and belt
drives, and vibration by many engine components. If exhaust systems did not include
resonators and mufflers, the noise generated would be a serious pollution.
Reduction in engine and exhaust noise can be done in one of three ways: passive,
semiactive, or active. Noise reduction is accomplished passively by correct design and
use of proper materials. The use of ribs and stiffeners, composite materials, and sand-
wich construction is now routine. This type of construction reduces noise vibrations in
the various engine components. Mufflers and resonators in the exhaust system dampen
out the majority of exhaust noise. Engine mounts, which connect the engine to the ve-
hicle body, are designed to dampen out vibrations that would transmit sound to the
passenger compartment.
Hydraulics are often used in semiactive noise abatement systems (e.g., some en-
gines are equipped with flywheels that have hydraulic passages through which fluid
can flow). The flywheels are designed such that, at different speeds, the fluid flows to
the specific locations that help to provide proper stiffness for absorbing engine vibra-
tions at the frequencies present at that speed. Some automobiles have hydraulic engine
mounts connecting the engine to the automobile body. Fluids in these mounts act to
absorb and dampen engine vibrations and isolate them from the passenger compart-
ment. Engine mounts using electrorheological fluids, which will allow better vibration
dampening at all frequencies, are being developed. The viscosity of these fluids can be
changed by as much as a factor of 50:1 with the application of an external voltage. En-
gine noise (vibration) is sensed by accelerometers that feed this information into the
engine management system (EMS). Here, the vibration frequency is analyzed, and the
proper voltage is applied to the engine mounts to best dampen that frequency [38]. Re-
sponse time is on the order of 0.005 seconds.
380
Emissions and Air Pollution
PROBLEMS
1 A diesel truck uses 100 grams of light diesel fuel (assume C 12H 22) per mile of travel. One-
half percent of the carbon in the fuel ends up as exhaust smoke. If the truck travels 15,000
miles per year, how much carbon is put into the atmosphere each year as smoke? [kg/year]
2 (a) Why isn’t a normal three-way catalytic converter, as used with SI engines, as useful when
used with a CI engine? (b) What main method is used to limit NOx emissions on a modern
diesel truck or automobile? (c) Give at least three disadvantages to using this method.
381
Emissions and Air Pollution
3 (a) List five reasons that there are HC emissions in the exhaust of an automobile. (b) To
reduce emissions from an SI engine, should AF be set at rich, lean, or stoichiometric? Ex-
plain the advantages and disadvantages of each. (c) Why is it good to place a catalytic con-
verter as close to the engine as possible? Why is this bad?
4 A four-cylinder, 2.8-liter, four-stroke cycle SI engine operates at 2300 RPM with a volu-
metric efficiency of 88.5%. The fuel used is methyl alcohol at an equivalence ratio of
f = 1.25. During combustion, all hydrogen is converted to water, and all carbon is con-
verted to CO2 and CO.
Calculate:
(a) Mole fraction of CO in the exhaust. [%]
(b) Energy lost in the exhaust due to CO. [kW]
5 The combustion chambers of a V8 Otto cycle engine with a 7.8:1 compression ratio, bore of
3.98 inches, and 410-cubic-inch displacement can be approximated as right circular cylinders.
The engine operates at 3000 RPM using gasoline at an AF = 15.2 and a volumetric efficien-
cy of 90%. When combustion occurs, the flame is dampened out near the walls and a bound-
ary layer of air–fuel mixture does not get burned. Combustion is at constant volume at TDC,
and the unburned boundary layer can be considered to be 0.004 inch thick over the entire
combustion chamber surface. Fuel is originally distributed equally throughout the chamber.
Calculate:
(a) Percent of fuel that does not get burned due to being trapped in the surface boundary
layer. [%]
(b) Amount of fuel lost in the exhaust due to this boundary layer. [lbm/hr]
(c) Chemical power of the fuel lost in the exhaust. [hp]
6 An older automobile using leaded gasoline gets 16 mpg fuel economy at 55 mph. The lead
in the gasoline amounts to 0.15 gm/L. Forty-five percent of the lead in the fuel gets ex-
hausted to the environment. Calculate the amount of lead exhausted to the environment
in lbm/mile and lbm/day if the automobile is driven continuously.
7 A small truck has a four-cylinder, 2.2-liter CI engine that operates on an air-standard Dual
cycle using light diesel fuel at an average AF = 21. At a speed of 2500 RPM, the engine
has a volumetric efficiency hv = 92%. At this operating condition, 0.4% of the carbon in
the fuel ends up as soot in the exhaust. In addition, there is 20% additional carbon soot
from the lubricating oil. The amount of soot is then increased by 25% due to other com-
ponents condensing on the carbon. Carbon density rc = 1400 kg>m3.
Calculate:
(a) Rate of soot put into the environment. [kg/hr]
(b) Chemical power lost in the soot. (Consider the entire mass of soot as carbon.) [kW]
(c) Number of soot clusters exhausted per hour. Assume that an average cluster contains
2000 spherical carbon particles, and each particle has a diameter of 20 nm.
8 The engine in Problem 7 operating at 2500 RPM has an indicated thermal efficiency of
61%, combustion efficiency of 98%, and mechanical efficiency of 71%.
Calculate:
(a) Brake specific fuel consumption. [gm/kW-hr]
(b) Specific emissions of soot particulates. [gm/kW-hr]
(c) Emissions index of soot particulates. [gm/kg]
382
Emissions and Air Pollution
9 The clearance volume in the cylinder of a large stationary SI engine can be approximated
as a right circular cylinder, with a diameter of 6 cm and a depth of 2 cm. The gas mixture in
the combustion chamber is a homogeneous mixture of gasoline vapor and air at an
AF = 16. It can be assumed that complete combustion occurs at TDC except for a thin
boundary layer 0.10 mm thick on all clearance volume surfaces where the closeness of the
walls damps out combustion and the air–fuel mixture does not get burned. Fuel flow into
the entire engine equals 0.040 kg/sec.
Calculate:
(a) Percent of fuel that is in the boundary layer and so does not get burned. [%]
(b) Chemical energy lost in exhaust due to unburned fuel from the boundary layer. [kW]
(c) Emissions index of HC due to this unburned fuel. [gm HC>kgf]
10 A turbocharged, 6.4-liter, V8 SI engine operates on an air-standard Otto cycle at WOT
with an engine speed of 5500 RPM. The compression ratio is rc = 10.4:1, and conditions in
the cylinders at the start of compression are 65°C and 120 kPa. Crevice volume is equal to
2.8% of clearance volume, pressure is equal to cylinder pressure, and temperature equal to
185°C.
Calculate:
(a) Total engine crevice volume. [cm3]
(b) Percent of fuel that is trapped in the crevice volume at the start of combustion at
TDC. [%]
11 The engine in Problem 10 has a volumetric efficiency of 89% and uses isooctane as fuel at
an air–fuel ratio AF = 14.2. Sixty percent of the fuel that is trapped in the crevice volume
at the start of combustion is later burned due to additional cylinder motion.
Calculate:
(a) HC emissions in the exhaust due to the 40% of crevice volume fuel that does not get
burned. [kg/hr]
(b) Chemical power lost in these HC emissions of the exhaust. [kW]
12 A large supercharged, two-stroke cycle, diesel ship engine with a displacement of 196 liters
operates at 220 RPM. The engine has a delivery ratio of ldr = 0.95 and uses fuel oil that
can be approximated as C 12H 22, at an air–fuel ratio of AF = 22. The ship is equipped with
an ammonia injection system to remove NOx from the exhaust.
Calculate:
(a) Amount of NO entering the exhaust system if 0.1% of the nitrogen in the air is con-
verted to NO. (Assume no other forms of NOx are produced.) [kg/hr]
(b) Amount of ammonia to be injected to remove all NO in the exhaust by the reaction
given in Eq. (29). [kg/hr]
13 It is desired to reduce NOx generation in an engine that burns stoichiometric ethanol by
using exhaust gas recycling (EGR) to lower the peak combustion temperature. The tem-
perature of the air and fuel at the start of combustion is 700 K, and the exhaust gas can be
approximated as N2 at 1000 K. The enthalpy of ethanol at 700 K is -199,000 kJ>kgmole.
Calculate:
(a) Theoretical maximum temperature with stoichiometric ethanol and no EGR. [K]
(b) Percent EGR needed to reduce maximum temperature to 2400 K. [%]
383
Emissions and Air Pollution
(a) Electrical energy needed to heat the preheat zone from 25°C to a light-off tempera-
ture of 150°C. [kJ]
(b) Time needed to supply this amount of energy. [sec]
15 A 0.02-liter, two-stroke cycle SI lawn mower engine runs at 900 RPM, using gasoline at an
equivalence ratio f = 1.08 and a fuel-to-oil ratio of 60:1 by mass. The engine is crankcase
compressed and has a delivery ratio ldr = 0.88 and a charging efficiency lce = 0.72. Com-
bustion efficiency 1hc2gasoline = 0.94 for the gasoline trapped in the cylinder, but 1hc2oil is
only 0.72 for the oil trapped in the cylinder. There is no catalytic converter.
Calculate:
(a) HC from the fuel and oil exhausted to the environment due to valve overlap during
scavenging. [kg/hr]
(b) HC in the exhaust from unburned fuel and oil due to combustion inefficiency. [kg/hr]
(c) Total HC in exhaust. [kg/hr]
16 A 3.2–liter, V6, four-stroke cycle, CI truck engine with a volumetric efficiency of 93% pro-
duces 92 kW of brake power at 3600 RPM, using light diesel fuel at AF = 22. The fuel
contains 450 ppm of sulfur by mass, which is exhausted to the environment. In the sur-
roundings the sulfur is converted to sulfurous acid by reacting with atmospheric oxygen
and water vapor as given in Eqs. (15) and (18).
Calculate:
384
Emissions and Air Pollution
Calculate:
(a) Specific emissions of CO upstream of catalytic converter. [gm/kW-hr]
(b) Overall average of specific emissions of CO (cold and warmed) downstream of catalytic
converter. [gm/kW-hr]
(c) Percentage of total CO emissions to the environment that occurs when converter is cold. [%]
19 A modern six-cylinder automobile CI engine is adjusted to operate properly using diesel
fuel with a cetane number of 52. The vehicle is accidently fueled with a diesel fuel having a
cetane number of 42. Would more or less exhaust smoke be expected? Explain.
20 In 1972, about 2.33 * 109 barrels of gasoline were consumed in the United States as fuel
for internal combustion engines. The average automobile traveled 16,000 km, using gaso-
line at a rate of 15 L/100 km. The gasoline, on average, contained 0.15 gm/liter of lead, 35%
of which was exhausted to the environment. One barrel = 160 liters.
Calculate:
(a) The amount of lead put into the atmosphere yearly by the average automobile. [kg]
(b) Total amount of lead put into the atmosphere in 1972. [kg]
21 A man wants to work on his automobile in his garage on a winter day. Having no heating
system in the garage, he runs the automobile in the closed building to heat it. At idle speed,
the engine burns 5 lbm of stoichiometric gasoline per hour, with 0.6% of the exhaust being
carbon monoxide. The inside dimensions of the garage are 20 ft by 20 ft by 8 ft, and the
temperature is 40°F. It can be assumed that 10 parts per million (ppm) of CO in the air is
dangerous to health. Calculate the time to when the CO concentration in the garage is
dangerous. [min]
22 An SI automobile engine produces 32 kW of brake power while using, on average, 6 kg of
stoichiometric gasoline per 100 km traveled at 100 km/hr. Average emissions from the en-
gine upstream of the catalytic converter are 1.1 gm/km of NO2, 12.0 gm/km of CO, and 1.4
gm/km of HC. A catalytic converter removes 95% of the exhaust emissions when it is at
steady-state temperature. However, 10% of the time, the catalytic converter is cold at
start-up and removes no emissions.
Calculate:
(a) Specific emissions of HC upstream of the catalytic converter. [gm/kW-hr]
(b) Specific emissions of CO downstream of the catalytic converter, with the converter
warmed. [gm/kW-hr]
(c) Concentration of NOx in the exhaust upstream of the catalytic converter. [ppm]
(d) Overall average (cold and warmed) of HC emissions to the atmosphere. [gm/km]
(e) Percentage of total HC emissions occurring when the converter is cold. [%]
DESIGN PROBLEMS
1D Design a catalytic converter preheater using solar energy. Decide if solar collectors should
be on the vehicle or on a battery-recharging station. Calculate the sizes needed for the
main components (e.g., battery, collector). Draw a simple schematic of the system.
2D Design a system to absorb the fuel vapors escaping from the vent on an automobile fuel
tank. The system should have a method of regeneration, with all fuel eventually being
input into the engine.
385
Emissions and Air Pollution
386
Heat Transfer in Engines
From Chapter 10 of Engineering Fundamentals of the Internal Combustion Engine, Second Edition.
Willard W. Pulkrabek. Copyright © 2004 by Pearson Education, Inc. All rights reserved.
387
Heat Transfer in Engines
This chapter examines the heat transfer that occurs within an IC engine, this being
extremely important for proper operation. About 35% percent of the total chemical
energy that enters an engine in the fuel is converted to useful crankshaft work, and
about 30% of the fuel energy is carried away from the engine in the exhaust flow in
the form of enthalpy and chemical energy. This leaves about one-third of the total
energy that must be dissipated to the surroundings by some mode of heat transfer.
Temperatures within the combustion chamber of an engine reach values on the
order of 2700 K and up. Materials in the engine cannot tolerate this kind of temper-
ature and would quickly fail if proper heat transfer did not occur. Removing heat is
highly critical in keeping an engine and engine lubricant from thermal failure. On
the other hand, it is desirable to operate an engine as hot as possible to maximize
thermal efficiency.
Two general methods are used to cool combustion chambers of engines. The en-
gine block of a water-cooled engine is surrounded with a water jacket that contains a
coolant fluid which is circulated through the engine. An air-cooled engine has a finned
outer surface on the block over which a flow of air is directed.
388
Heat Transfer in Engines
1 ENERGY DISTRIBUTION
The amount of energy (power) available for use in an engine is
# #
W = mf Q HV (1)
where
#
mf = fuel flow rate into the engine
QHV = heating value of the fuel
The mass flow of fuel is limited by the mass flow of air that is needed to react with
the fuel. Brake thermal efficiency gives the percentage of this total energy that is con-
verted to useful output at the crankshaft:
1ht2brake = Wb>mf QHV hc
# #
(2)
where
ht = thermal efficiency
hc = combustion efficiency
#
Wb = brake power
The rest of the energy can be divided into heat losses, parasitic loads, and energy
that is lost in the exhaust flow. Figure 1 shows a typical distribution of energy use in an
IC engine given as a percentage of total fuel energy. The total sums to over 100% be-
cause friction losses are counted twice, first as the original loss and then again in the re-
sulting heat losses. For any engine,
# # # #
Power generated = Wshaft + Qexhaust + Qloss + Wacc (3)
where
#
Wshaft = brake output power off of the crankshaft
#
Qexhaust = energy lost in the exhaust flow
#
Qloss = all other energy lost to the surroundings by heat transfer
#
Wacc = power to run engine accessories
Depending on the size and geometry of an engine, as well as on how it is being
operated, the shaft power output is
#
Wshaft L 25–40%
CI engines are generally on the high end of this range, and SI engines are on the
lower end. Energy lost in exhaust flow is
#
Qexhaust L 20–45%
A greater percentage of energy is lost in the exhaust of SI engines because of their
higher exhaust temperatures. Lost exhaust energy is made up of two parts: enthalpy
(heat) and chemical energy. When the engine is running rich at full load, chemical ener-
gy makes up about half of the exhaust loss. Under many operating conditions, lost ex-
haust energy exceeds the brake power output of the engine. Other heat losses are
#
Qloss L 10–35%
389
Heat Transfer in Engines
100 Friction
Heat Loss to Surroundings
Percent of Intake Fuel Energy (%)
80 Exhaust Enthalpy
60
Coolant Load
40
20
Brake Power
0
1000 2000 3000
Engine Speed, N (RPM)
FIGURE 1
Distribution of energy in a typical SI engine as a function of engine speed. Friction
losses, which are generally on the order of 10%, add to other heat losses and make
the total energy distribution greater than 100%.
390
Heat Transfer in Engines
2 ENGINE TEMPERATURES
Figure 2 shows a typical temperature distribution that would be found in an IC engine
operating at steady state. Three of the hottest points are around the spark plug, the ex-
haust valve and port, and the face of the piston. Not only are these places exposed to
the high-temperature combustion gases, but they are difficult places to cool.
The highest gas temperatures during combustion occur around the spark plug.
This creates a critical heat transfer problem area. The spark plug fastened through the
combustion chamber wall creates a disruption in the surrounding water jacket, causing
a local cooling problem. On air-cooled engines, the spark plug disrupts the cooling fin
pattern, but the problem may not be as severe.
The exhaust valve and port operate hot because they are located in the pseudo-
steady flow of hot exhaust gases and create a difficulty in cooling similar to the one the
spark plug creates. The valve mechanism and connecting exhaust manifold make it
very difficult to route coolant or allow a finned surface to give effective cooling.
The piston face is difficult to cool because it is separated from the water jacket or
outer finned cooling surfaces.
Engine Warm-up
As a cold engine heats up to steady-state temperature, thermal expansion occurs in all
components. The magnitude of this expansion will be different for each component, de-
pending on its temperature and the material from which it is made. Engine bore limits
the thermal expansion of the pistons, and at operating temperatures of a newer engine,
Exhaust Intake
Valve Valve
650 Spark Plug 250
600
Intake
Exhaust Manifold
Flow 60
450
Piston
Face
300
Cylinder
Coolant
Wall
105
185
Piston
Ring
220
Piston
Skirt FIGURE 2
190 Typical temperature values found in
Oil an SI engine operating at normal
70 steady state conditions. Temperatures
are in degrees C.
391
Heat Transfer in Engines
221
232 220
205
177 193
228 226
(196)
221 176
(224)
240
228 201
172 (198)
201.8
201.9
210 201
(207) 162
(196)
FIGURE 3
Steady-state temperatures of piston in plane of piston pin and in thrust plane. Engine is a
four-cylinder, 2.5 liter, SI engine running at 4600 RPM with WOT. Temperatures are in °C.
Reprinted with permission from SAE Paper No. 820086 © 1982 SAE International, [194].
there can be very high resulting forces between the piston rings and skirt and the walls
of the cylinder. This causes high viscous heating in the oil film on the cylinder walls dur-
ing engine operation.
Figure 5 shows how the temperature of various automobile components increas-
es with time after a cold engine is started. In cold weather, the start-up time to reach
steady-state conditions can be as high as 20–30 minutes. Some parts of the automobile
reach steady state much sooner than this, but some do not. Fairly normal operating
conditions may be experienced within a few minutes, but it can take as long as an hour
to reach optimum fuel consumption rates. Engines are built to operate best at steady-
state conditions, and full power and optimum fuel economy may not be realized until
this state is reached. It would be poor practice to take off with an airplane, when full
power is needed, before the engine is fully warmed up. This is not as critical with an au-
tomobile. A large percentage of automobile use is for short trips with engines that are
not fully warmed up.
392
Heat Transfer in Engines
Measured
Center of Crown 1
Top Ring Land 2
2nd Ring Land 3
Middle of Skirt 4
300
250
Piston Temperatures, ⬚C
3
200
4
FIGURE 4
150
Piston temperatures as a function of engine
speed at full load. Engine is a four-cylinder,
2.5 liter, SI engine running at 4600 RPM
100 with WOT. Reprinted with permission from
0 1000 2000 3000 4000 5000 SAE Paper No. 820086 © 1982 SAE
Engine Speed, r/min International, [194].
600
Spark Plug
Temperature, T( C)
400
Exhaust
Valve
Piston
200
Coolant
0 10 20 30 40 50 60
Time (seconds)
FIGURE 5
Temperatures of engine components of a typical SI engine as a function of time
after cold start-up.
393
Heat Transfer in Engines
Q = hA1Twall - Tgas2
#
(4)
where
T = temperature
h = convection heat transfer coefficient
A = inside surface area of intake manifold
The manifold is hot, either by design on some engines or just as a result of its lo-
cation close to other hot components in the engine compartment. Carbureted engines
and those with throttle body injection that introduce fuel early in the flow process pur-
posely have heated intake manifolds to assist in the evaporation of the fuel. Various
methods are used to heat these manifolds. Some are designed such that the flow pas-
sages of the runners come in close thermal contact with the hot exhaust manifold. Oth-
ers use hot coolant flow through a surrounding water jacket. Electricity is used to heat
some intake manifolds. Some systems have special localized hot surfaces, called hot
spots, in optimum locations, such as immediately after fuel addition or at a tee where
maximum convection occurs (Fig. 6).
There are several consequences from convective heating in the intake manifold,
some good and some bad. The earlier that the fuel gets vaporized, the longer it is mixed
with air, resulting in a more homogeneous mixture. However, increasing the tempera-
ture reduces the volumetric efficiency of the engine by two mechanisms. Higher tem-
perature reduces the air density and added fuel vapor displaces some of the air, both
reducing the mass of air reaching the cylinders. A compromise is to vaporize some of
the fuel in the intake system and to vaporize the rest in the cylinder during compres-
sion, or even during combustion. With older carbureted engines, it was desirable to va-
porize about 60% of the fuel in the intake manifold. As fuel vaporizes in the intake
manifold, it cools the surrounding flow by evaporative cooling, counteracting the con-
vective heating. When the intake charge of air and fuel enters the cylinder, it is further
heated by the hot cylinder walls. This, in turn, helps to cool the cylinder walls and to
keep them from overheating.
Another reason to limit the heating of inlet air is to keep temperature to a mini-
mum at the start of the compression stroke. The higher the temperature at the start of
compression, the higher will be all temperatures throughout the rest of the cycle, and
the greater is the potential problem of engine knock.
394
Heat Transfer in Engines
Air
Liquid
Fuel
FIGURE 6
Hot spot in intake manifold to accelerate fuel
evaporation. Localized sections of wall surface,
called hot spots, can be heated by engine coolant,
by conduction from the exhaust manifold, or by
electrical heating. Heated sections are generally
placed close after fuel addition, or at a tee where
Hot Spot high convection occurs.
Engine systems using multipoint port injectors have less need for heating the in-
take manifold, relying on finer fuel droplets and higher temperature around the intake
valve to assure necessary fuel evaporation. This results in higher volumetric efficiency
for these engines. Often, the fuel is sprayed directly onto the back of the intake valve
face. This not only speeds evaporation, but cools the intake valve, which can reach
cyclic temperatures up to 400°C. Steady-state temperature of intake valves generally is
in the 200° -300°C range.
If an engine is supercharged or turbocharged, the temperature of the inlet air is
also affected by the resulting compressive heating. To avoid this, many of these systems
are equipped with aftercooling, which again lowers the temperature. Aftercoolers are
heat exchangers through which the compressed inlet air flows, using either engine
coolant or external air flow as the cooling fluid.
395
Heat Transfer in Engines
The air–fuel mixture entering a cylinder during the intake stroke may be hotter
or cooler than the cylinder walls, with the resulting heat transfer being possible in ei-
ther direction. During the compression stroke, the temperature of the gas increases,
and by the time combustion starts, there is already a convective heat transfer to the
cylinder walls. Some of this compressive heating is lessened by the evaporative cooling
that occurs when the remaining liquid fuel droplets vaporize.
During combustion, peak gas temperatures on the order of 3000 K occur within
the cylinders, and effective heat transfer is needed to keep the cylinder walls from
overheating. Convection and conduction are the main heat transfer modes to remove
energy from the combustion chamber and keep the cylinder walls from melting.
Figure 7 shows heat transfer through a cylinder wall. Heat transfer per unit sur-
face area will be
Heat transfer in Eq. (5) is cyclic. Gas temperature Tg in the combustion chamber
varies greatly over an engine cycle, ranging from maximum values during combustion to
Cylinder Cylinder
Wall Wall
Tg ⫽ 1000 Tg ⫽ 1000
Air
Coolant
Tw ⫽ 190 Tw ⫽ 190
Fins
Tc ⫽ 105
hg hg
hc Ta ⫽ 25
Combustion Combustion
Chamber Chamber
⌬X Air Cooled Engine
Liquid Cooled Engine
FIGURE 7
Heat transfer through the combustion chamber cylinder wall of an IC engine. The cylinder gas
temperature Tg and convection heat transfer coefficient hg vary over large ranges for each
engine cycle, while the coolant temperature Tc (or air temperature Ta) and heat transfer
coefficient hc are fairly constant. As a result of this, heat conduction is cyclic for a small depth
into the cylinder wall on the combustion chamber side. Temperatures are in degrees C.
396
Heat Transfer in Engines
minimum during intake. It can even be less than wall temperature early in the intake
stroke, momentarily reversing heat transfer direction. Coolant temperature Tc is fairly
constant, with any changes occurring over much longer cycle times. The coolant is air for
air-cooled engines and antifreeze solution for water-cooled engines. The convection heat
transfer coefficient hg on the cylinder gas side of the wall varies greatly during an engine
cycle due to changes in gas motion, turbulence, swirl, velocity, etc. This coefficient will
also have large spatial variation within the cylinder for the same reasons. The convection
heat transfer coefficient on the coolant side of the wall will be fairly constant, being de-
pendant on coolant velocity. Thermal conductivity k of the cylinder wall is a function of
wall temperature and will be fairly constant.
Convection heat transfer on the inside surface of the cylinder is
Wall temperature Tw should not exceed 180°–200°C to assure thermal stability of the
lubricating oil and structural strength of the wall.
There are a number of ways of identifying a Reynolds number to use for compar-
ing flow characteristics and heat transfer in engines of different sizes, speeds, and
geometries. Choosing the best characteristic length and velocity is sometimes difficult
[40, 120]. One way of defining a Reynolds number for engines [120] which correlates
data fairly well is
where
#
ma = mass flow rate of air into the cylinder
#
mf = mass flow rate of fuel into the cylinder
B = bore
Ap = area of piston face
mg = dynamic viscosity of gas in the cylinder
From this Reynolds number, a Nusselt number for the inside of the combustion
chamber can be defined as
where
C1 and C2 = constants
kg = thermal conductivity of cylinder gas
hg = average value of the convection heat transfer coefficient to be
used in Eqs. (5) and (6)
The Nusselt number and convection heat transfer coefficient on the coolant side
of the cylinder walls can be approximated by conventional methods of forced convec-
tion heat transfer.
397
Heat Transfer in Engines
Radiation heat transfer between cylinder gas and the combustion chamber walls is
q = Q>A = [s1T4g - T4w2]>5[11 - Pg2>Pg] + [1>F1 - 2] + [11 - Pw2>Pw]6
# #
(9)
where
Tg = gas temperature
Tw = wall temperature
s = Stefan–Boltzmann constant
Pg = emissivity of gas
Pw = emissivity of wall
F1 - 2 = view factor between gas and wall
Even though gas temperatures are very high, radiation to the walls amounts to
only about 10% of the total heat transfer in SI engines. This is due to the poor emitting
properties of gases, which emit only at specific wavelengths. N2 and O2, which make up
the majority of the gases before combustion, radiate very little, while the CO2 and H 2O
of the combustion products do contribute more to radiation heat transfer.
The solid carbon particles that are generated in the combustion products of a CI
engine are good radiators at all wavelengths, and radiation heat transfer to the walls in
these engines is in the range of 20–35% of the total. A large percentage of radiation
heat transfer to the walls occurs early in the power stroke. At this point, the combus-
tion temperature is maximum, and with thermal radiation potential equal to T4, a very
large heat flux is generated. This is also the time when there is a maximum amount of
carbon soot in CI engines, which further increases radiative heat flow. Instantaneous
heat fluxes as high as 10 MW>m2 can be experienced in a CI engine at this point of the
cycle.
Because an engine operates on a cycle, the gas temperature Tg within the cylinder
in Fig. 7 and Eq. (5) is pseudo-steady state. This cyclic temperature causes a cyclic heat
transfer to occur in the cylinder walls. However, due to the very short cycle times, this
cyclic heat transfer occurs only to a very small surface depth. At normal speeds, 90% of
these heat transfer oscillations are dampened out within a depth of about 1 mm of the
surface in engines with cast-iron cylinder walls. In engines with aluminum cylinders,
this depth of 90% dampening is slightly over 2 mm, and in ceramic walls it is on the
order of 0.7 mm. At surface depths greater than these, the oscillations in heat transfer
are almost undetectable and conduction can be treated as steady state [40].
Heat transfer to cylinder walls continues during the expansion stroke, but the
rate quickly decreases. Expansion cooling and heat losses reduce the gas temperature
within the cylinder during this stroke from a maximum temperature on the order of
2700 K to an exhaust temperature of about 800 K. During the exhaust stroke, heat
transfer to the cylinder walls continues but at a greatly reduced rate. At this time, cylin-
der gas temperature is much lower, as is the convection heat transfer coefficient. There
is no swirl or squish motion at this time, and turbulence is greatly reduced, resulting in
a much lower convection heat transfer coefficient.
Cycle-to-cycle variations in combustion that occur within a cylinder result in
cycle-to-cycle variations in cylinder wall temperature (Fig. 8) and cylinder wall heat
398
Heat Transfer in Engines
415
Average
(198 Cycles)
Cycle 4
Temperature, K 410
Cycle 5
405
Cycle 2
Cycle 3
Cycle 1
400
340 360 380 400 420 440
Crank angle, degrees
FIGURE 8
Temperature variation of five consecutive cycles recorded by sensor in wall of
combustion chamber of an SI engine. Also shown is the average temperature for
198 cycles as a function of crank angle, 360° being top-dead-center. The engine has
bore of 10.47 cm and stroke of 9.53 cm, and was operating at 1500 RPM with an
equivalence ratio of 0.87. Reprinted with permission from Journal of Heat Transfer
© ASME, [2].
flux (Fig. 9). Not only are there time variations at any given point, but there are spacial
variations from one point to another on the combustion chamber wall.
Heat transfer occurs in all four strokes of a cycle, ranging from very high fluxes to
low fluxes and even to zero or heat flow in the reverse direction (i.e., heat flow from
the walls to the gas mixture in the cylinder). In a naturally aspirated engine, heat flow
can be in either direction during the intake stroke, at a given cylinder location. During
the compression stroke, the gases heat up and a heat flux to the walls results. Maximum
temperature and maximum heat flux occur during combustion and then decrease dur-
ing the power and exhaust strokes. For engines equipped with a supercharger or tur-
bocharger, the intake gases are at a higher temperature, and the corresponding heat
flux during intake is greater and into the walls.
Cooling difficulties caused by the protrusions of the spark plug, fuel injectors, and
valves through the cylinder walls have been discussed. Another major cooling problem
is the face of the piston. This surface is exposed to the hot combustion process but can-
not be cooled by the coolant in the engine water jacket or an external finned surface.
For this reason, the piston crown is one of the hotter points in an engine. One method
used to cool the piston is by splashing or spraying lubricating oil on the back surface of
the piston crown. In addition to being a lubricant, the oil then also serves as a coolant
399
Heat Transfer in Engines
Cycle 3
Cycle 4
1.5
Surface Heat Flux, MW/m2
Cycle 5
Average
(198 Cycles)
1
Cycle 2
0.5 Cycle 1
0
340 360 380 400 420 440
Crank Angle, Degrees
FIGURE 9
Variation in measured heat transfer rate at combustion chamber wall for five
consecutive cycles of an SI engine operating at 1500 RPM with an equivalence ratio of
0.87. Also shown is the average value of 198 cycles as a function of crank angle.
Reprinted with permission from Journal of Heat Transfer © ASME, [2].
fluid. After absorbing energy from the piston, the oil flows back into the oil reservoir in
the crankcase, where it is again cooled. Heat is also conducted from the piston face, but
thermal resistance for this is quite high. The two conduction paths available are (1)
down the connecting rod to the oil reservoir, and (2) through the piston rings to the
cylinder walls and into the coolant in the surrounding water jacket (Fig. 10). Thermal
resistance through the piston body and connecting rod is low because they are made of
metal. However, there is high resistance where these connect together at the wrist pin
because of the lubricant film between the surfaces. This is also true where the connect-
ing rod fastens to the crankshaft through lubricated surfaces. The oil film between sur-
faces needed for lubrication and wear reduction constitutes a large thermal resistance
and a poor conduction path.
Aluminum pistons generally operate 30°–80°C cooler than cast-iron pistons due
to their higher thermal conductivity. This reduces knock problems in these engines, but
can cause greater thermal expansion problems between dissimilar materials. Many
modern pistons have a ceramic face and operate at a higher steady-state temperature.
Ceramic has poor heat conduction properties but can tolerate much higher tempera-
tures. Some very large engines have water-cooled pistons.
To avoid thermal breakdown of the lubricating oil, it is necessary to keep the
cylinder wall temperatures from exceeding 180°–200°C. As lubrication technology im-
proves the quality of oils, this maximum allowable wall temperature is being raised. As
400
Heat Transfer in Engines
Water
Jacket
A
q X
Y
FIGURE 10
Cooling of piston. The face of a piston (A) is one of the hotter
surfaces in a combustion chamber. Cooling is done mainly by
Y convection to the lubricating oil on the back side of the piston
face, by conduction through the piston face, by conduction
through the piston rings in contact with the cylinder walls, and
by conduction down the connecting rod to the oil reservoir.
High conduction resistance occurs because of lubricated
surfaces at cylinder walls (X) and rod bearings (Y).
an engine ages, deposits slowly build up on the walls of the cylinders, due to impurities
in the air and fuel, imperfect combustion, and lubricating oil in the combustion cham-
ber. These deposits create a thermal resistance and cause higher wall temperatures. Ex-
cessive wall deposits also slightly decrease the clearance volume of the cylinder and
cause a rise in the compression ratio.
Some modern engines use heat pipes to help cool internal hot regions that are in-
accessible to normal cooling by conduction or coolant flow. With one end of the heat
pipe in the hot interior of the engine, the other end can be in contact with the circulat-
ing coolant or exposed to external air flow.
Example Problem 1
A 3.0 liter, 5-cylinder, 4-stroke cycle SI engine, with a volumetric efficiency of 82%, operates at 3000
RPM using gasoline with a lambda value of 0.91. Bore and stroke are related as S = 1.08 B. At a
certain point in the engine cycle, the gas temperature in the combustion chamber is Tg = 2100°C
while the cylinder wall temperature is Tw = 190°C. Calculate the approximate convection heat
transfer rate to the cylinder wall at this instant.
Bore can be calculated as follows:
401
Heat Transfer in Engines
The mass flow rate of air into one cylinder of the engine is
#
ma = hvraVdN>n
= 10.82211.181 kg>m32[10.003 m3>cycle2>15 cylinders2]13000>60 rev>sec2>12 rev>cycle2
= 0.01453 kg>sec
The mass flow rate of fuel into one cylinder of the engine is
Equation (8) gives the Nusselt number and convection heat transfer coefficient (ref. [40] sug-
gests C1 = 0.035 and C2 = 0.80):
Equation (6) gives convection heat transfer rate at combustion chamber wall at this instant:
402
Heat Transfer in Engines
60
3500 rpm
3000 rpm
2500 rpm
40 2000 rpm
1400 rpm
20
NuD ⫽ 0.0483ReD
0.783
NuD ⫽ hD
k
10
Nu ⫽ 0.0246ReD
0.8
7
FIGURE 11
Average Nusselt number as a function of Reynolds number in the exhaust flow of a
reciprocating IC engine (top curve). The cyclic pulsing that occurs in the exhaust
increases the Nusselt number and convection heat transfer in the exhaust pipe by a factor
of about two over steady-state flow conditions of equal mass flow rate in the same pipe
(bottom curve). Reprinted with permission from SAE Paper No. 790309 © 1979 SAE
International, [199].
working fluid because of its thermal properties, and its melting point of about 98°C
(208°F) [223].
Example Problem 2
On the 6.8 liter, V8 engine with a compression ratio of 9:1, the exhaust manifold and pipe lead-
ing from the engine to the catalytic converter can be approximated as a 1.8-m length of pipe with
ID = 6.0 cm and OD = 6.5 cm. Volumetric efficiency of the engine at 3600 RPM is hv = 93%,
the air–fuel ratio AF = 15:1, and the average wall temperature of the exhaust pipe is 200°C. Cal-
culate the approximate temperature of the exhaust gas entering the catalytic converter.
403
Heat Transfer in Engines
Heat transfer equations from any standard textbook can be used. [63] The temperature of
the exhaust gas leaving the engine is T1 = 756 K = 483°C. As a first approximation, it will be as-
sumed that the temperature loss in the exhaust pipe is ¢T = 100 K, or T2 = 656 K = 383°C. As
in air-standard analysis, air property values are used to approximate exhaust gas.
Average bulk temperature of gas is
mex = [hvraVdN>n]116>152
#
= 159.5 m>sec
Using the Dittus–Boelter equation for the Nusselt number of interior turbulent flow in a pipe,
Nu = 12212722 = 544
= 37,709 W.
This gives a temperature drop in the exhaust flow between the engine and catalytic converter of
404
Heat Transfer in Engines
Engine Size
If two geometrically similar engines of different size (displacement) are run at the
same speed, and all other variables (temperature, AF, fuel, etc.) are kept as close to the
same as possible, the larger engine will have a greater absolute heat loss but will be
more thermally efficient. If the temperatures and materials of both engines are the
same, heat loss fluxes to the surroundings per unit area will be about the same, but the
absolute heat loss of the larger engine will be greater due to its larger surface areas.
A larger engine will generate more output power and will do this at a higher ther-
mal efficiency. As linear size goes up, volume increases on the order of linear dimen-
sion cubed. If one engine is 50% larger in linear size, its displacement will be on the
order of 11.523 = 3.375 larger. With similar mixture properties, the larger engine will
therefore combust about 3.375 times the fuel of the smaller engine and will release
3.375 times the amount of thermal energy. Surface area, on the other hand, is propor-
tional to length squared, and the larger engine will have only 2.25 times the surface
area and consequent heat loss of the smaller engine. Energy generated goes up with
length cubed, while heat losses go up with length squared. This makes the larger engine
more efficient if everything else is the same.
This reasoning can be extended to more than just absolute size when designing
an engine. What is desirable for good thermal efficiency is a combustion chamber with
a high volume–to-surface area ratio. This is one reason why a modern overhead valve
engine is more efficient than older valve-in-block L head engines that had large com-
bustion chamber surface areas. This also says that a cylinder with a single, simple, open
combustion chamber will have less percentage heat loss than one with a split dual
chamber that has a large surface area.
Engine Speed
As engine speed is increased, gas flow velocity into and out of the engine goes up, with
a resulting rise in turbulence and convection heat transfer coefficients. This increases
heat transfer occurring during the intake and exhaust strokes and even during the
early part of the compression stroke.
405
Heat Transfer in Engines
During combustion and power stroke, gas velocities within the cylinder are fairly
independent of engine speed, being controlled instead by swirl, squish, and combustion
motion. The convection heat transfer coefficient and, thus, convection are therefore
fairly independent of engine speed at this time. Radiation, which is important only dur-
ing this portion of the cycle, is also independent of speed. Rate of heat transfer (kW)
during this part of the cycle is therefore constant, but because the time of the cycle is
less at higher speed, less heat transfer per cycle (kJ/cycle) occurs. This gives the engine
a higher thermal efficiency at higher speed. At higher speeds, more cycles per unit time
occur, but each cycle lasts less time. The net result is a slight rise in heat transfer loss
with time (kW) from the engine. This is partly due to the higher heat losses for part of
the cycle, but is mostly due to the higher steady-state (pseudo-steady-state) losses that
the engine establishes at higher speeds. Mass flow of gas through an engine increases
with speed, with a net result of less heat loss per unit mass (kJ/kg) (i.e., higher thermal
efficiency).
All steady-state temperatures within an engine go up as engine speed increases,
as shown in Fig. 12.
Heat transfer to the engine coolant increases with higher speed is
Q = hA1Tw - Tc2
#
(10)
where
h = convection heat transfer coefficient, which remains about constant
A = surface area, which remains constant
1000
800
Exhaust Gas
Temperature, T( C)
600
Exhaust Valve
400
Piston Face
200
Cylinder Wall
FIGURE 12
Engine temperatures as a function of engine speed for a typical SI engine.
406
Heat Transfer in Engines
Load
As the load on an engine is increased (going uphill, pulling a trailer), the throttle must
be opened further to keep the engine speed constant. This causes less pressure drop
across the throttle and higher pressure and density in the intake system. Mass flow rate
of air and fuel, therefore, goes up with load at a given engine speed. Heat transfer with-
in the engine also goes up by
#
Q = hA¢T (11)
where
h = convection heat transfer coefficient
A = surface area at any point
¢T = temperature difference at that point
The heat transfer coefficient is related to Reynolds number by
h r ReC (12)
where C is a constant, usually on the order of 0.8. Reynolds number is proportional to
# #
the mass flow rate m, so the time rate of heat transfer increases with m0.8. Density of
# #
fuel into the engine increases as m, so energy into the engine increases as m.
The percentage of heat loss goes down slightly as engine load increases (kJ/cycle).
This quite often is offset by engine knock, which occurs most often in an engine under
load. The result of knock is localized high temperature and high heat transfer. Engine
temperatures increase with load. Figure 12 would be very similar if the X coordinate of
speed at constant load were replaced with load at constant speed.
CI engines are run unthrottled, and total mass flow is almost independent of
load. When speed or load is increased and more power is needed, the amount of fuel
407
Heat Transfer in Engines
injected is increased. This increases the total mass flow in the latter part of each cycle
only very slightly, on the order of 5%. This means that the convection heat transfer co-
efficient within the engine is fairly independent of engine load.
At light loads, less fuel is injected and burned, creating a cooler steady-state tem-
perature. This decreases the corresponding heat transfer. At heavy load, more fuel is
injected and burned, and the resulting steady-state temperature is higher. This causes a
greater convective heat transfer. Combustion of the richer mixture at heavy load also
creates a larger amount of solid carbon soot. This, in turn, further increases heat trans-
fer by radiation, solid carbon being a good radiator. The amount of fuel and, conse-
quently, the amount of energy released per cycle goes up with load. The percent of heat
loss, therefore, changes very little with load in a CI engine.
Spark Timing
More power and higher temperatures are generated when the spark setting is set to
give maximum pressure and temperature at about 5° to 10° aTDC. These higher peak
temperatures will create a higher momentary heat loss, but this will occur over a short-
er length of time. With spark timing set either too early or too late, combustion effi-
ciency and average temperatures will be lower. These lower temperatures will give less
peak heat loss, but the heat losses will last over a longer length of time and the overall
energy loss will be greater. Higher power output is thus gained with correct ignition
timing. Late ignition timing extends the combustion process longer into the expansion
stroke, resulting in higher exhaust temperature and hotter exhaust valves and ports.
408
Heat Transfer in Engines
added by one of three methods: (1) injection of water into the incoming air, either in
the intake system or directly into the combustion chamber; (2) emulsifying water with
the fuel; or (3) using high humidity inlet air. Of these methods, the direct injection of
water seems to be the most practical, and is used in a number of existing systems. Water
storage and concern of the water freezing are possible problems, especially with road
vehicles. Mixing water with the fuel in an emulsion can cause possible mixing, storage,
and/or injector problems. Large volumes of high humidity air can be difficult to pro-
vide, and can cause corrosion problems in the intake system.
Saab Automobile Company has been experimenting with water injection to im-
prove fuel economy at high speed and fast acceleration. To avoid the need for an addi-
tional water source, fluid is taken from the windshield washer reservoir. Antifreeze
solution and washer additives do not seem to harm the engine. High-speed fuel con-
sumption has been reduced by 20–30% [70].
Example Problem 3
A large supercharged aircraft engine generates 900 kW at 3600 RPM when operating with air
and gasoline at a fuel equivalence ratio f = 1.05. After supercharging and fuel addition, air en-
ters the engine at 65°C. Gasoline can be approximated as isooctane.
Calculate how much the air is cooled by evaporative cooling when the fuel vaporizes.
The reaction for stoichiometric combustion is
C 8H 18 + 112.5>1.052 O2 + 112.5>1.05213.762 N2 :
18>1.052 CO2 + 19>1.052 H 2O + 0.05 C 8 H 18 + 112.5>1.05213.762 N2
409
Heat Transfer in Engines
Example Problem 4
Water injection is added to the engine in Example Problem 3, with 0.25 kg of water injected for
each kg of fuel used. Let the heat of vaporization of water hfg = 2350 KJ>kg.
Calculate:
1. approximate inlet air temperature when water injection is used
2. approximate engine power with water injection
(1) For one mole of fuel, the mass of fuel will be
= 1024 kW
This value, however, is reduced somewhat when some of the inlet air is displaced by
water vapor. For each mole of inlet fuel, there are 112.5>1.05214.762 = 56.67 moles of
inlet air. With water injection, there are also 10.25211142>1182 = 1.583 moles of water
vapor. Engine output power is then
1W2output = 1W2with[1Na2>1Na + Nvapor2]
# #
410
Heat Transfer in Engines
Coolant Temperature
Increasing the coolant temperature of an engine (hotter thermostat) results in higher
temperatures of all cooled components. There is little change in the temperatures of
the spark plugs and exhaust valves. Indicated thermal efficiency would be higher, but
there is a potential for a greater knock problem in hotter engines.
Engine Materials
Different materials used in the manufacture of cylinder and piston components result
in different operating temperatures. Aluminum pistons, with their higher thermal con-
ductivity, generally operate about 30°–80°C cooler than equivalent cast-iron pistons.
Ceramic-faced pistons have poor thermal conductivity, resulting in very high tempera-
tures. This is by design, with the ceramic being able to tolerate the higher temperature.
Ceramic exhaust valves are sometimes used because of their lower mass inertia and
high temperature tolerance.
Compression Ratio
Changing the compression ratio of an engine changes the heat transfer to the coolant
very little. Increasing the compression ratio decreases heat transfer slightly up to about
rc = 10. Increasing the compression ratio above this value increases heat transfer
slightly [58]. There is about a 10% decrease in heat transfer as the compression ratio is
raised from 7 to 10. These changes in heat transfer occur mainly because of the com-
bustion characteristics that change as the compression ratio is raised (e.g., flame speed,
gas motion, etc.). The higher the compression ratio, the more expansion cooling will
occur during the power stroke, resulting in cooler exhaust. CI engines, with their high
compression ratios, generally have lower exhaust temperatures than SI engines. Piston
temperatures generally increase slightly with increasing compression ratio.
Knock
When knock occurs, the temperature and pressure are raised in very localized spots
within the combustion chamber. This rise in local temperature can be very severe and,
in extreme cases, can cause surface damage to pistons and valves.
7 AIR-COOLED ENGINES
Many small engines and some medium-sized engines are air cooled. This includes most
small-engine tools and toys like lawn mowers, chain saws, model airplanes, etc. This al-
lows both the weight and price of these engines to be kept low. Some motorcycles, au-
tomobiles, and aircraft have air-cooled engines, also benefitting from lower weight.
Air-cooled engines rely on a flow of air across their external surfaces to remove the
necessary heat to keep them from overheating. On vehicles like motorcycles and aircraft,
the forward motion of the vehicle supplies the air flow across the surface. Deflectors and
411
Heat Transfer in Engines
ductwork are often added to direct the flow to critical locations.The outer surfaces of the
engine are made of good heat-conducting metals and are finned to promote maximum
heat transfer. Automobile engines usually have fans to increase the air-flow rate and di-
rect it in the desired direction. Lawn mowers and chain saws rely on free convection from
their finned surfaces. Some small engines have exposed flywheels with air deflectors fas-
tened to the surface. When the engine is in operation, these deflectors create air motion
that increases heat transfer on the finned surfaces.
It is more difficult to get uniform cooling of cylinders on air-cooled engines than
on liquid-cooled engines. The flow of liquid coolants can be better controlled and duct-
ed to the hot spots where maximum cooling is needed. Liquid coolants also have better
thermal properties than air (e.g., higher convection coefficients, specific heats, etc.).
Figure 13 shows how cooling needs are not the same at all locations on an engine sur-
face. Hotter areas, such as around the exhaust valve and manifold, need greater cooling
and a larger finned surface area. Cooling the front of an air-cooled engine which faces
the forward motion of the vehicle is often much easier and more efficient than cooling
the back surface of the engine. This can result in temperature differences and thermal
expansion problems.
When compared with liquid-cooled engines, air-cooled engines have the following
advantages: (1) lighter weight, (2) less costly, (3) no coolant system failures (e.g., water
pump, hoses), (4) no engine freeze-ups, and (5) faster engine warm-up. Disadvantages of
air-cooled engines are that they (1) are less efficient, (2) are noisier, with greater air flow
requirements and no water jacket to dampen noise, and (3) need a directed air flow and
finned surfaces.
Standard heat transfer equations for finned surfaces can be used to calculate the
heat transfer off of these engine surfaces.
FIGURE 13
Variation of heat losses from the fins of an
air-cooled aircraft engine. Seventy-one
percent of the heat losses occur on the hotter
side of the cylinder, containing the exhaust
valve. The engine shown was used on a
number of different aircraft, including the
six-engine B-36 bomber. Reprinted with
permission from SAE Paper 500197 © 1950
SAE International, [106].
412
Heat Transfer in Engines
8 LIQUID-COOLED ENGINES
The engine block of a water-cooled engine is surrounded with a water jacket through
which coolant liquid flows (Fig. 14). This allows for a much better control of heat re-
moval at a cost of added weight and the need for a water pump. The cost, weight, and
complexity of a liquid coolant system makes this type of cooling very rare on small
and/or low-cost engines.
Very few water-cooled engines use just water as the coolant fluid in the water
jacket. The physical properties of water make it a very good heat transfer fluid, but it
has some drawbacks. Used as a pure fluid it has a freezing point of 0°C, unacceptable in
northern winter climates. Its boiling temperature, even in a pressurized cooling system,
is lower than desired, and without additives it promotes rust and corrosion in many
materials. Most engines use a mixture of water and ethylene glycol, which has the heat
transfer advantages of water but improves on some of the physical properties. Ethyl-
ene glycol 1C 2H 6O22, often called antifreeze, acts as a rust inhibitor and a lubricant for
the water pump, two properties not present when water is used alone. When added to
water, it lowers the freezing temperature and raises the boiling temperature, both de-
sirable consequences.This is true for mixtures with ethylene glycol concentrations from a
very small amount up to about 70%. Due to a unique temperature–concentration–phase
relationship, the freezing temperature again rises at high concentrations. The desirable
heat transfer properties of water are also lost at high concentrations. Pure ethylene gly-
col should not be used as an engine coolant.
FIGURE 14
Schematic of cooling system of water-cooled 1982 1.8 liter Chevrolet engine. Reprinted with permission
from SAE Paper No. 820111 © 1982 SAE International, [135].
413
Heat Transfer in Engines
Ethylene glycol is water soluble and has a boiling temperature of 197°C and a
freezing temperature of -11°C in pure form at atmospheric pressure. Table 1 gives
properties of ethylene glycol–water mixtures. When ethylene glycol is used as an en-
gine coolant, the concentration with water is usually determined by the coldest weath-
er temperature that is expected to be experienced.
Engine coolant cannot be allowed to freeze. If it does, it will not circulate through
the radiator of the cooling system and the engine will overheat. A more serious conse-
quence is caused when the water in the coolant expands on freezing and cracks the walls
of the water jacket or water pump. This destroys the engine. Even in climates where
there is no danger of freezing water, some ethylene glycol should be used because of its
°C °F °C °F
°C °F °C °F
414
Heat Transfer in Engines
Most commercial antifreezes satisfy these requirements. Many of them are basically
ethylene glycol with small amounts of additives.
A hydrometer is used to determine the concentration of ethylene glycol when it is
mixed with water.The specific gravity of the mixture is determined by the height at which
the calibrated hydrometer floats. Charts such as Table 1 can be used to determine the
concentration needed. Most of these hydrometers are used by service station attendants
who have no engineering training. For this reason they are usually not calibrated in con-
centration, but only in freezing temperature of the total water–ethylene glycol mixture.
Most commercial antifreeze products (Prestone, Zerex, Dex-Cool, etc.) are basically eth-
ylene glycol, and the same calibrated hydrometer can be used for all of these.
Some commercial engine coolants (Sierra, etc.) use propylene glycol 1C 4H 8O2 as
the base ingredient. It is argued that, when coolant systems leak or when the coolant
becomes aged and is discarded, these products are less harmful to the environment
than ethylene glycol. A far lesser amount of these products is sold in the United States
than those containing ethylene glycol.
HISTORIC—ANTIFREEZE
There have been some serious mistakes made when the few antifreeze products using
something other than ethylene glycol as the base ingredient have appeared on the market.
With a specific gravity different from that of ethylene glycol, a special hydrometer with a
different calibration must be used when testing concentrations of these antifreezes. Most
service stations use hydrometers calibrated for ethylene glycol. If these are used by mis-
take to test antifreezes made of other base materials, an incorrect freezing temperature is
obtained. This can lead to unexpected freezing of the coolant mixture and a destroyed
engine with a cracked block. This has happened a number of times.
The coolant system of a typical automobile engine is shown in Fig. 15. Fluid en-
ters the water jacket of the engine, usually at the bottom of the engine. It flows through
the engine block, where it absorbs energy from the hot cylinder walls. The flow pas-
sages in the water jacket are designed to direct the flow around the outer surfaces of
the cylinder walls and past any other surface that needs cooling. The flow is also direct-
ed through any other component that may need heating or cooling (e.g., heating of the
415
Heat Transfer in Engines
intake manifold or cooling of the oil reservoir). The flow leaves the engine block con-
taining a high specific enthalpy because of the energy it absorbed in engine cooling.
Exit is usually at the top of the engine block.
Enthalpy must now be removed from the coolant flow so that the circulation
loop can be closed and the coolant can again be used to cool the engine. This is done by
the use of a heat exchanger in the flow loop called a radiator. The radiator is a honey-
comb heat exchanger with hot coolant flowing from top to bottom exchanging energy
with cooler air flowing from front to back, as shown in Fig. 15. Air flow occurs because
of the forward motion of the automobile, assisted by a fan located behind the radiator
and either driven electrically or off the engine crankshaft. The cooled engine coolant
exits the bottom of the radiator and reenters the water jacket of the engine, completing
a closed loop. A water pump that drives the flow of the coolant loop is usually located
between the radiator exit and engine block entrance. This pump is either electric or
mechanically driven off the engine. Some early automobiles had no water pump and
relied on a natural convection thermal flow loop.
Air leaving the automobile radiator is further used to cool the engine by being di-
rected through the engine compartment and across the exterior surfaces of the engine.
Because of the modern aerodynamic shape of automobiles and the great emphasis on
cosmetics, it is much more difficult to duct cooling air through the radiator and engine
compartment. Much greater efficiency is needed in rejecting energy with the modern
FIGURE 15
Radiator of a liquid-cooled engine used to remove heat from the coolant loop of the engine. A
radiator is a liquid-to-air heat exchanger generally mounted in front of the engine on an
automobile. Liquid flow is supplied by the engine water pump, while air flow is a result of the
forward motion of the automobile, assisted by one or more fans. Adapted from [81].
416
Heat Transfer in Engines
radiator heat exchanger. Modern engines are designed to run hotter and thus can tol-
erate a lower cooling air-flow rate. Steady-state temperature of the air within the en-
gine compartment of a modern automobile is on the order of 125°C.
To keep the coolant fluid temperature from dropping below some minimum
value, and thus keeping the engine operating at a higher temperature and efficiency, a
thermostat is installed in the coolant loop, usually at the engine flow entrance. A ther-
mostat is a thermally activated go–no go valve. When the thermostat is cold, it is closed
and allows no fluid flow through the main circulation channel. As the engine warms up,
the thermostat also warms up, and thermal expansion opens the flow passage and al-
lows coolant circulation. The higher the temperature, the greater the flow passage
opening, with the greater resulting coolant flow. The coolant temperature is, therefore,
controlled fairly accurately by the opening and closing of the thermostat. Thermostats
are manufactured for different coolant temperatures, depending on engine use and cli-
mate conditions. They generally come in ratings from cold (140°F) to hot (240°F).
Coolant loops of older automobiles operated at atmospheric pressure using most-
ly water. This limited overall coolant temperatures to about 180°F (83°C), allowing for a
safety margin to avoid boiling. In order to increase engine operating temperature for
better efficiency, it was necessary to increase coolant temperature. This was done by
pressurizing the coolant loop and adding ethylene glycol to the water. The ethylene gly-
col raised the boiling temperature of the fluid as shown in Table 1. Pressurizing the sys-
tem further raises the boiling temperature of the fluid regardless of the concentration of
ethylene glycol. Normal coolant system pressures are about 200 kPa absolute.
It is desirable for the coolant to remain mostly liquid throughout the flow loop. If
boiling occurs, a small mass of liquid becomes a large volume of vapor, and steady-state
mass flow becomes almost impossible to sustain. By using ethylene glycol in a pressur-
ized system, high temperatures can be achieved without large-scale boiling. Localized
boiling in small hot spots does occur within the engine water jacket. This is good. The
very hottest spots within the engine (either momentary or almost steady state) require
the greatest heat removal and cooling. The phase change that is experienced when boil-
ing occurs at these local hot spots absorbs a large amount of energy and supplies the
necessary large cooling at these spots. The circulating convection flow carries the re-
sulting vapor bubbles away from the hot spots back into the main stream of the
coolant. Here they condense back into liquid due to the cooler fluid temperature, and
bulk flow is not interrupted.
As hot engine coolant leaves the engine block, it can be used to heat the passen-
ger compartment of an automobile, when desired. This is done by routing a portion of
the coolant flow through an auxiliary system that supplies the hot side of a small liquid-
to-air heat exchanger. Outside or recirculated air is heated as it passes through the
other half of the heat exchanger and is ducted into the passenger compartment and/or
onto the cold windows for defrosting. Various manual and automatic controls deter-
mine the flow rates of the air and coolant to supply the desired warming results.
The small diesel engines of some vehicles are so efficient that they do not supply
enough waste heat to adequately heat the passenger compartment under some oper-
ating conditions (e.g., idling at a stoplight). These vehicles sometimes have an electric
resistance auxiliary heater to use at these times. These heaters will become more prac-
tical with 42-volt electrical systems. Another auxiliary method used on some of these
417
Heat Transfer in Engines
Contessa 900
(⫽Renault 4CV) Renault R8
(⫽Benz 170H)
Direction of travel Direction of travel
Radiator
Exhaust
pipe
FIGURE 16
Various methods used to cool rear-mounted liquid-cooled engines. Reprinted with
permission from The Romance of Engines by T. Suzuki, © 1997 SAE International, [227].
vehicles is a viscous heater. A viscous heater is a pump that churns a fluid to generate
heat through friction, one such system using silicon gel [161]. The pump can be driven
electrically or belted to the crankshaft and controlled on–off, as needed, by the EMS.
HISTORIC—ENGINE COOLANTS
Very early automobile engines were either air cooled or water cooled. At first, it was com-
mon practice to drain the water from water-cooled engines in cold weather and to store
the car through the winter. Two of the first liquid antifreezes that were used were alcohols
and kerosene. These allowed use of the automobile in cold weather, but great care was
needed to avoid coolant system leaks. These liquids are combustible, and many automo-
biles burned up when leaking antifreeze came in contact with the hot engine and exhaust
system.
418
Heat Transfer in Engines
9 OIL AS A COOLANT
The oil used to lubricate an engine in operation also helps to cool the engine. Because
of its location, a piston gets very little cooling from the coolant in the water jacket or
the external finned surface of an engine. To help cool the piston face, one of the hottest
surfaces in the engine, the back surface of the piston crown, is subjected to a flow of oil.
This is done by spraying the oil in pressurized systems or by splash in nonpressurized
systems. The crankcase of many engines also serves as the oil reservoir, and the move-
ment of the crankshaft and connecting rods splash oil over all exposed surfaces. The oil
acts as a coolant on the back face of the piston crown as it absorbs energy and then
runs back into the larger reservoir. Here, it mixes with the cooler oil and dissipates this
energy into the other engine parts. This splash oil cooling of the piston is very impor-
tant in small air-cooled engines as well as in automobile engines.
Other components are also cooled by oil circulation, either by splash or by pres-
surized flow from the oil pump. Oil passages through internal components like the
camshaft and connecting rods offer the only major cooling these parts are subjected to.
As the oil cools the various components, it absorbs energy and its temperature rises.
This energy is then dissipated to the rest of the engine by circulation and eventually
gets absorbed in the engine coolant flow.
Some high-performance engines have an oil cooler in their lubricant circulation
system. The energy absorbed by the oil as it cools the engine components is dissipated
in the oil cooler, which is a heat exchanger cooled by either engine coolant flow or ex-
ternal air flow.
10 ADIABATIC ENGINES
A small increase in brake power output can be gained by decreasing the heat losses
from an engine cylinder. About 30% of available energy is converted to useful work
(thermal efficiency), and this is done near TDC during combustion and the following
expansion stroke, which encompasses about one-fourth of the total engine cycle. On
the other hand, heat transfer occurs over the entire 720° of the cycle. Therefore only
about one-fourth of the saved energy is available when output work is being generated,
and only about 30% of this is utilized. If a 10% decrease of heat loss energy were ac-
complished over the cycle (a major accomplishment), only a fraction of this would ap-
pear as added crankshaft output:
419
Heat Transfer in Engines
heat treating and alloying of metals and advancements in ceramics and composites. The
development of flexible ceramic materials which could withstand the mechanical and
thermal shocks that occured within an engine was a major breakthrough in the 1980s.
These materials are now commonly found in modern engines, especially at the highest
temperature spots, such as piston face and exhaust port. A common material found in
adiabatic engines is silicon nitride 1Si3N42. Because they have no cooling system (water
pump, water jacket, finned surfaces, etc.), adiabatic engines can be made smaller and
lighter than conventional engines.Vehicles can be made more aerodynamic with a lower
drag coefficient because there is no radiator. This also gives greater flexibility in engine
location and positioning.
All engine components of an adiabatic engine, including the cylinder walls, oper-
ate at higher temperatures. These heat the incoming air mixture quicker than in a con-
ventional engine. This reduces the volumetric efficiency of the engine, which, in turn,
deletes some of the brake power increase gained from less heat loss. The higher cylin-
der temperature during the compression stroke also raises the pressure and reduces
the net work output of the cycle by increasing the compression work input.
Adiabatic engines are all compression ignition. They cannot be used as spark ig-
nition engines, because the hot cylinder walls would heat the air–fuel mixture too
quickly and knock would be a major problem. Another problem created by the hot
cylinder walls, which are on the order of 800 K, is thermal breakdown of the lubricating
oil. Better oils have been developed that can tolerate the conditions in present-day en-
gines, but lubrication technology will need to continue to advance to keep up with in-
creasing engine demands. One solution that is being considered and developed is the
use of solid lubricants.
420
Heat Transfer in Engines
at moderate speed with no coolant in the cooling system. This is possible by firing only
four of the eight cylinders at any given time. The four cylinders that do not get fuel, and
do not fire, continue to pump air, which cools the engine enough to keep it from over-
heating. Each set of four cylinders cycles between firing for a period of time and pump-
ing cooling air at other times.
12 THERMAL STORAGE
Some vehicles are equipped with a thermal battery that can be used to preheat an en-
gine and automobile. A thermal battery takes waste heat from the engine coolant dur-
ing operation and stores about 500 to 1000 W-hr (1800–3600 kJ) for as long as three
days or more. Various methods to do this have been tried and used. The most common
system stores energy by use of a liquid–solid phase change occurring in a water–salt
crystal mixture. The stored energy can then be used in cold-weather starting to preheat
the engine, preheat the catalytic converter, and/or heat the passenger compartment
and defrost the car windows. Preheating can commence within a few seconds.
A number of different systems and materials have been tried with various suc-
cess. One early system uses about 10 kg of Ba1OH228H 2O as the base material. This
has a latent heat (solid-to-liquid) of 89 W-hr/kg and a melting point of 78°C. The
salt–water mixture of this system is contained within hollow fins fastened on the interi-
or of a cylindrical flow chamber through which engine coolant flows. The exterior wall
of the cylinder is insulated with super-high vacuum insulation that restricts heat losses
to three watts or less when the surrounding temperature is at -20°C [79].
When the engine is running under normal conditions, hot coolant is ducted
through the thermal battery and liquifies the salt–water mixture. This is done with en-
ergy that would otherwise be rejected in the automobile radiator, so there is no oper-
ating cost for the system. When the engine is turned off and coolant flow stops, the
liquid–salt solution will very slowly change phase to solid as it cools. This phase change
will take about three days because of the superinsulated wall of the container. As the
salt solution changes phase, the temperature in the container remains at 78°C. To later
recover the energy for use in preheating, an electric pump circulates the now cold
coolant through the thermal battery, where it initiates a liquid-back-to-solid phase
change in the salt–water mixture. When this occurs, the latent heat is absorbed by the
coolant, which leaves the battery at a temperature near 78°C. The coolant can then be
ducted to the engine or the catalytic converter for preheating or can be used to heat
the passenger compartment. One or all of these are possible with proper piping and
controls.
A preheated engine starts quicker with less wear and wasted fuel. The warmed
cylinder walls and intake manifold promote better fuel evaporation, and combustion
starts quicker. Also, the overrich intake mixture used to start a cold engine can be less-
ened, saving fuel and causing less emissions. The engine lubricating oil also gets pre-
heated, which greatly reduces its viscosity. This allows for faster engine turnover with
the starting motor and quicker engine starting. This also allows for quicker and better
early oil distribution, which reduces engine wear. Preheating an alcohol-fueled engine is
especially important. Because of its high latent heat, it is difficult to evaporate enough
421
Heat Transfer in Engines
alcohol to get a cold engine started. This is one of the serious disadvantages of alcohol
fuel, and engine preheating reduces the problem.
Preheating the catalytic converter is one very effective way of reducing emis-
sions. This can be done to a limited extent with a thermal battery.
Another way the stored energy of the thermal battery can be used is to pipe the
heated coolant through the heat exchanger of the passenger compartment heater sys-
tem. The heater can then immediately be turned on, with the air flow directed into the
passenger compartment and/or onto the windows for deicing.
For the first 10 seconds of operation, a thermal battery can supply 50–100 kW.
The system can be started when the ignition key is inserted or even when the car door
is opened. Effective preheating occurs within 20–30 seconds. This compares to several
minutes for effective heating of the engine, catalytic converter, and passenger compart-
ment on nonpreheated automobiles.
Various systems supply different percentages of the stored energy, and in differ-
ent sequences, to the various uses. Some systems have greater flexibility and change-
ability than others.
Even if all stored energy is delivered to the passenger compartment, the engine
will warm up quicker. This is because no engine heat will be diverted to the cabin
heater system. This may be important in large trucks as laws are being considered in
some areas that require driver compartment heating before the driver is allowed to get
into the vehicle. When the thermal battery is being recharged after the engine has heat-
ed up, there is no loss of cabin heater efficiency, since the engine supplies more than
enough energy for both purposes.
Most thermal battery systems have a mass of about 10 kg and can supply
500–1000 W-hr as their temperature drops from 78°C to 50°C. Total energy depletion
will normally take 20–30 minutes, depending on the engine and surrounding tempera-
tures. The battery can be located either in the engine compartment or somewhere else
in the automobile, such as the trunk. Mounting it in the engine compartment results in
the least piping and greatest efficiency, but space limitations may not allow this.
The greatest benefit from thermal batteries will probably be in those automobiles
that are used for city driving. Many city trips are short enough that the catalytic con-
verter never reaches operating temperature, resulting in high exhaust emissions. Re-
ducing startup emissions in densely populated areas is most important because of the
large number of automobiles and other polluting systems, as well as the large number
of people affected. Emissions from limited-range dual-powered automobiles that are
being developed for city commuting will be greatly reduced with thermal storage.
These automobiles, which are propelled by an electric motor most of the time and only
use their small IC engine when extra power or range is needed, operate their engines in
an on–off mode. The engine is started only when needed and therefore operates most
often only for short periods of time after start-up, a highly polluting method. With a
thermal battery, energy can be stored while the engine operates and then used to pre-
heat the engine before the next start-up.
Although there is no direct running cost in charging a thermal battery, there will
be a very slight cost in a higher fuel consumption due to the added mass of the vehicle
422
Heat Transfer in Engines
contributed by the battery. A 10-kg battery is only 1% of the mass in a 1000-kg auto-
mobile. This, however, will cause a very slight steady-state increase in fuel consumption
even on long-distance travel, which gets no benefit from the battery.
Thermal batteries are designed to last the lifetime of the automobile.
Example Problem 5
The fluid in a thermal battery system of an automobile consists of 6.2 lbm of salt solution which
changes liquid–solid phase at 175°F. When the automobile engine is operating, steady-state en-
gine coolant temperature is 215°F, and the coolant flow rate through the thermal battery is 0.025
lbm/sec. Because it is superinsulated, there is a steady heat loss from the thermal battery of only
11 BTU/hr. After the automobile has sat for some time, the engine coolant is at a temperature of
75°F, and the salt solution in the thermal battery is 60% liquid and 40% solid by mass. Immedi-
ately after engine start-up, coolant enters the thermal battery at a temperature of 75°F and exits
at 165°F. Water property values can be used for engine coolant. Property values of the salt solu-
tion are:
1. how long thermal battery can supply coolant at 165°F after engine start-up
2. time to cool thermal battery from 215°F to 75°F after engine has stopped
(1) Battery can supply coolant at 165°F until all liquid salt has turned to solid: energy to
change phase = Qcp = [16.2 lbm210.602]1125 BTU>lbm2 = 465.0 BTU
Rate of energy transfer to heat coolant flow is
423
Heat Transfer in Engines
13 SUMMARY
Combustion temperatures in the cylinders of IC engines can reach values of 2700 K
and higher. Without adequate cooling, temperatures of this magnitude would quickly
destroy engine components and lubricants. If cylinder walls are allowed to exceed tem-
peratures above 200°C, material failures will occur and most lubricating oils will break
down. To keep the cylinders from overheating, they are surrounded with a water jack-
et on liquid-cooled engines or a finned surface on air-cooled engines. On the other
hand, to obtain maximum efficiency from an engine, it is desirable to operate it as hot
as possible. With improvements in materials and lubrication technology, modern en-
gines can operate much hotter than engines of a few years ago.
Heat removed from engine cylinders is eventually rejected to the surroundings.
Unfortunately, by keeping the engine from overheating with heat transfer to the sur-
roundings, a large percent of the energy generated within the engine is wasted, and the
brake thermal efficiency of most engines is on the order of 30–40%.
Because of the low profile of a modern automobile, cooling air flow is much more
restricted and greater heat transfer efficiency is necessary.
Innovative cooling systems are being developed, but at present most automobile
engines are liquid cooled using a water–ethylene glycol solution. Most small engines
are air cooled because of the requirements of weight, cost, and simplicity.
PROBLEMS
1 An in-line, six-cylinder, 6.6-liter, four-stroke cycle SI engine with multipoint port fuel in-
jection operates at 3000 RPM, with a volumetric efficiency of hv = 89%. The intake man-
ifold runners can be approximated as round pipes with an inside diameter of 4.0 cm. Inlet
temperature to the manifold is 27°C.
Calculate:
(a) Average velocity and mass flow rate of air to each cylinder, using inlet temperature to
evaluate properties. [m/sec, kg/sec]
(b) Reynolds number in the runner to cylinder #1, which is 40 cm long (use standard inte-
rior pipe flow equations).
(c) Temperature of the air entering cylinder #1 if the runner wall temperature is constant
at 67°C. [°C]
(d) Wall temperature needed for the runner to cylinder #3, such that the cylinder air
inlet temperature is the same as that in cylinder #1. Runner length for cylinder #3 is
15 cm. [°C]
2 The engine in Problem 1 is converted to throttle body fuel injection with the fuel injected
at the inlet end of the manifold. Forty percent of the fuel evaporates in the intake manifold
runners, which cools the air by evaporative cooling. Wall temperatures remain the same.
Calculate:
(a) Temperature of the air entering cylinder #1 if the fuel is stoichiometric gasoline. [°C]
(b) Temperature of the air entering cylinder #1 if the fuel is stoichiometric ethanol. [°C]
424
Heat Transfer in Engines
3 The engine in Problem 2b has bore and stroke related by S = 0.90 B. Using inlet condi-
tions to evaluate properties, calculate the Reynolds number as defined by Eq. (7).
4 The engine in Problem 3 has an exhaust pipe between the engine and catalytic converter
that can be approximated as a round pipe, 1.5 m long and 6.5 cm inside diameter. Exhaust
temperature leaving the engine is 477°C, and the average wall temperature of the exhaust
pipe is 227°C. Calculate the exhaust temperature entering the catalytic converter. [°C]
5 An automobile cruises at 55 mph using a brake power of 20 kW. The engine, which is rep-
resented by Fig. 1, runs at a speed of 2000 RPM.
Approximate:
(a) Power lost in the exhaust flow. [kW]
(b) Power lost to friction. [kW]
(c) Power dissipated in the coolant system. [kW]
6 An engine represented by Fig. 1 has a coolant system with a flow rate of 25 gal/min and a
thermostat that controls coolant flow temperature into the engine at 220°F. The engine
produces 30 bhp at 2500 RPM as the automobile travels at 30 MPH. Frontal area of the ra-
diator is 4.5 ft2, and a fan increases air flow velocity through the radiator by a factor of 1.1.
Calculate:
(a) Coolant temperature as it exits the engine. [°F]
(b) Air temperature leaving the radiator if the ambient temperature is 75°F. [°F]
7 A certain automobile model is offered with two engine options. The two engines are iden-
tical V8s with different displacements; one is 320 in.3 and the other is 290 in.3. The engines
are run at identical speeds, temperatures, and operating conditions.
Calculate:
(a) Rough approximation of what percent greater (or less) will be the indicated thermal
efficiency of the larger engine versus the smaller. [%]
(b) Approximation of what percent greater (or less) will be the total heat transfer to the
coolant fluid of the larger engine versus the smaller. [%]
8 A service station attendant mixes a water–antifreeze solution so that the mixture will have
a freezing point of -30°C. The antifreeze used is ethylene glycol, but by mistake the at-
tendant uses a hydrometer calibrated for propylene glycol. Calculate the actual freezing
temperature of the mixture. [°C]
9 A thermal storage battery consists of 10 kg of a salt solution that changes phase at 80°C
and has the following properties:
425
Heat Transfer in Engines
(c) How long until the battery cools to an ambient temperature of 10°C. [hr]
(d) How long the battery can supply coolant at 80°C when the engine is started at a tem-
perature of 20°C. Assume the battery solution starts as all liquid at 80°C, and the en-
gine coolant enters at 20°C and leaves at 80°C. [min.]
10 Calculate the length of time the automobile in Example Problem 5 has been inoperative in
order to reach the condition described (i.e., salt solution in thermal battery is 60% liquid
and 40% solid). Surroundings are at 75°F. [hr]
11 Coolant (water) flow through an automobile engine has a flow rate of 20 gallons per
minute and removes 1000 BTUs per minute from the engine. The water enters the engine
at a temperature of 200°F. The radiator on the automobile has a frontal area of 4 ft2 and an
air flow velocity of 50 ft/sec through it. (for water 1 gal = 8.4 lbm)
Calculate:
(a) Energy made available to generate steam if the exhaust temperature decreases from
577°C to 227°C as it passes through the heat exchanger. [kW]
(b) Saturated steam vapor that can be generated if steam enters the heat exchanger as
saturated liquid at 101 kPa. The heat exchanger efficiency is 98% and, for water at 101
kPa, hfg = 2257 kJ>kg. [kg/hr]
13 During combustion, there is a momentary heat flux through the wall of the combustion
chamber at a certain spot equal to 67,000 BTU/hr-ft2. Gas temperature in the cylinder at
this time is 3800°R, and the convection heat transfer coefficient within the cylinder is
22 BTU/hr-ft2-°R. Coolant temperature is 185°F. Thermal conductivity of the 0.4-inch-
thick cast-iron cylinder wall is 34 BTU/hr-ft-°R.
Calculate:
426
Heat Transfer in Engines
DESIGN PROBLEMS
1D Design a thermal storage system for an automobile. The system is to be used to preheat
the oil and the catalytic converter, and to warm the passenger compartment. Determine
the size and materials of a thermal battery. Determine flow rates versus time, and draw a
flow diagram schematic. Explain the sequence of events when the automobile is started,
using approximate energy flows and temperatures.
2D Design an engine cooling system that uses two separate water jackets. Give the fluids used,
flow rates, temperatures, and pressures. Show the flow diagram and pumps on a schematic
drawing of the engine.
427
This page intentionally left blank
Friction and Lubrication
From Chapter 11 of Engineering Fundamentals of the Internal Combustion Engine, Second Edition.
Willard W. Pulkrabek. Copyright © 2004 by Pearson Education, Inc. All rights reserved.
429
Friction and Lubrication
“The motors and water coolers are placed up front, usually over or
nearly over the front axle, the gears and fuel in the rear and the
weight of the passengers in the middle. The net results of this
change are long wheel bases, low center of gravity, angle-iron
frames, plain springs, running gears without reaches, the comfort-
able tonneau body, freedom from vibration, good traction, great
hill-climbing qualities, almost total absence of slip, and easy access
to all parts, the motor being covered only by a detachable metal
hood or bonnet.”
This chapter examines the friction that occurs in an engine and the lubrication needed
to minimize this friction. Friction refers to the forces acting between mechanical com-
ponents due to their relative motion and to the forces on and by fluids when they move
through the engine. A percentage of the power generated within the engine cylinders is
lost to friction, with a reduction in the resulting brake power obtained off the crank-
shaft. Accessories that are run off the engine also reduce crankshaft output and are
often classified as part of the engine friction load.
430
Friction and Lubrication
(a)
Lubricating
Oil
FIGURE 1
Motion between engine components, highly
magnified to show surface roughness. (a) Dry or
nonlubricated surface showing friction caused
by high spots. (b) Lubricated surface showing
(b)
reduction of friction by hydraulic floating.
Bearings offer a unique lubrication problem because one surface (race) sur-
rounds the other surface (shaft). When an engine is not in operation, gravity pulls the
shaft in any bearing (crankshaft, connecting rod, etc.) down and squeezes out the oil
film between the two surfaces (Fig. 2a). In operation, the combination of a rotating
431
Friction and Lubrication
(a) Not Turning (b) Lubricated
FIGURE 2
Lubrication of bearings. (a) Nonrotating, lubricating oil is squeezed out and
surface contacts surface. (b) Rotating, oil film is dragged by moving surface
and surfaces are separated by thin layer of fluid.
shaft, viscous effects, and dynamic forces from various directions results in hydraulic
floating of the shaft, offset slightly from center (Fig. 2b). The position and thickness of
the minimum oil film in the bearing will depend on tolerances, load, speed, and oil vis-
cosity. It will be on the order of 2 m11 m = 10 -6 m2 for the main bearings in an engine.
For additional analysis, the reader is referred to the many books available on dynamic
lubrication of bearings.
2 ENGINE FRICTION
Friction can be classified as a loss using power terms:
Wf = 1Wi2net - Wb
# # #
(1)
subscript
f = friction
i = indicated
b = brake
Friction can also be classified in terms of work:
wf = 1wi2net - wb (2)
432
Friction and Lubrication
losses is in terms of mean effective pressure. Mean effective pressure can be related to
any work or power term
work W = 1mep2Vd (4)
power W = 1mep2Vd1N/n2
#
(5)
where
Vd = displacement volume
N = engine speed
n = number of revolutions per cycle
Using frictional work or frictional power and rearranging, these become:
fmep = Wf/Vd (6)
fmep = Wf/[Vd1N/n2]
#
(7)
fmep = imep - bmep (8)
In some analyses [e.g., 40], the mep concept is expanded to include all work and
power inputs and outputs of an engine. Various mep terms and the work they corre-
spond to include:
amep—work to drive auxiliaries such as the power steering pump
bmep—work done by the engine crankshaft
cmep—work to power the supercharger or turbocharger compressor
fmep—work lost to internal friction and to drive necessary engine equipment
such as the oil pump
gmep—gross work—indicated work of compression and expansion strokes
imep—net work generated in the combustion chambers
mmep—work needed to motor an engine
pmep—pumping work—indicated work of the exhaust and intake strokes
tmep—work recovered from the exhaust gas in a turbocharger turbine
These are related as follows:
fmep = imep - bmep - amep - cmep + tmep (9)
imep = gmep - pmep (10)
Assuming amep = 0 and cmep = tmep, it follows that
fmep = imep - bmep (11)
Friction mep can quite accurately be related to engine speed by the empirical
equation
fmep = A + BN + CN 2 (12)
where
N = engine speed
A, B, C = empirical constants related to a specific engine
433
Friction and Lubrication
The first term on the right side of Eq. (12) 1A = constant2 is sometimes called
boundary friction. It occurs between components of the engine where there is not
enough lubrication to hydraulically separate totally the motion of one surface from an-
other. Metal-to-metal contact occurs between the piston rings and cylinder walls at
TDC and BDC, and in heavily loaded bearings of the crankshaft. Periodic metal-to-
metal contact occurs when heavily loaded surfaces move at low speeds or undergo sud-
den acceleration and direction changes. When this happens, the lubrication is squeezed
out and there is a momentary lack of hydraulic floating. Places where this happens
include bearings of the crankshaft and connecting rods, the piston ring–cylinder wall
interface at TDC and BDC, and in most components at start-up.
The second term on the right side of Eq. (12) is proportional to engine speed and
relates to the hydraulic shear that occurs between many lubricated engine components.
Shear force per unit surface area is given as
ts = m1dU/dy2 = m1¢U/¢y2 (13)
where
m = dynamic viscosity of lubricating oil
1dU/dy2 = velocity gradient between surfaces
¢U = velocity difference between adjacent surfaces
¢y = distance between adjacent surfaces
For a given viscosity (temperature) and geometry, the velocity term ¢U is pro-
portional to the speed N.
The third term of Eq. (12) is related to engine speed squared. This term accounts
for the losses from turbulent dissipation in the intake and exhaust flows. Dissipation
equates to the square of mass velocity, which, in turn, relates to engine speed. Constants
A, B, and C must be determined for the operating conditions of a given engine.
An empirical equation similar to Eq. (12), but replacing engine speed with aver-
age piston speed Up, can be written as
434
Friction and Lubrication
of a cycle from an indicator diagram generated from pressure sensors in the combus-
tion chamber. Brake power is directly measured by connecting the crankshaft output
to a dynamometer.
It is much more difficult and less accurate to divide total friction into parts to de-
termine the percentage of the total contributed by various engine components. One of
the best ways to do this is to motor the engine (i.e., drive an unfired engine with an ex-
ternal electric motor connected to the crankshaft). Many electric dynamometers are
capable of doing this, making them an attractive type of dynamometer. Engine power
output is measured with an electric dynamometer by running a generator off the en-
gine crankshaft and measuring the electric load imposed on the generator. The genera-
tor is designed as a dual system, which also allows it to be used as an electric motor that
can drive the connected IC engine. When an engine is motored, the ignition is turned
off and no combustion takes place. Engine rotation is provided and controlled by the
connected electric motor, with the resulting cycle much like that shown in Fig. 3. Un-
like a fired engine, both the compression–expansion loop and the exhaust–intake loop
of this cycle represent negative work on the cylinder gases. This work is provided
through the crankshaft from the electric motor.
The power needed to be supplied by the electric motor to rotate the engine is
# # # #
Wm = Wf + Wg + Wp (16)
Pressure, P
Compression
and Expansion
FIGURE 3
P-V diagram of motored engine.
Intake and compression strokes are
Pumping similar to those in the cycle of a fired
engine. With no combustion to raise
pressure, the expansion stroke
TDC BDC
reverses the intake stroke and there is
Volume, V very little blowdown.
435
Friction and Lubrication
where
#
Wm = power to motor the engine
#
Wf = friction power
= gross indicated power 1compression and expansion2
#
Wg
= pumping indicated power 1exhaust and intake2
#
Wp
1gmep2m = 0 (18)
If the motored engine is operated at WOT, then the pump work loop is almost
zero, and
1pmep2m = 0 (19)
If all parameters such as speed and temperature, are kept consistent in the mo-
tored engine as would be found in the fired engine, then
and
and
# #
Wm L Wf (23)
Thus, by measuring the electric power input to the motor driving the engine, a
good approximation is obtained of the friction power lost in normal engine operation.
It is imperative that all conditions of the motored engine be kept as close as pos-
sible to the conditions of a fired engine, especially temperature. Temperature greatly
affects the viscosity of the engine fluids (lubricating oil, coolant, and air) and the ther-
mal expansion and contraction of the various components, both of which have a major
436
Friction and Lubrication
influence on engine friction. The oil must be circulated at the same rate and tempera-
ture (viscosity) as in a fired engine. Air and engine coolant flow should be kept as con-
sistent as possible (i.e., with the same throttle setting and pump rates).
The normal way of measuring friction in an engine by the motoring method is to
first run the engine in a normal fired mode. When the engine has reached a steady-state
condition with all temperatures, it is turned off and immediately tested using the elec-
tric motor. For a very brief period of time, the engine temperatures will be almost the
same as with a fired engine. This will quickly change because no combustion is occur-
ring, and the engine will start to cool off. There will be some differences immediately.
Even with all other temperatures being correct, the exhaust flow will be quite differ-
ent. Hot combustion products that make up the exhaust flow in a fired engine are ap-
proximated with much cooler air in a motored engine. At best, motored engine test
results give a close approximation of engine friction.
As friction in an engine is being tested by the motoring method, engine com-
ponents can be removed to determine how much they individually contribute to
total friction. For instance, the engine can be motored with and without the valves
connected. The difference in power required gives an approximation to the friction
of the valves. A problem with this is the difficulty of keeping engine temperatures
near normal operating temperatures when the engine is partially dismantled. Figure
4 gives typical results for the percent of friction contributed by various engine com-
ponents.
The components that contribute a major part of total friction are the pistons and
piston rings. Figure 5 shows the friction forces on a typical piston assembly as it goes
through one cycle. The forces are greatest near TDC and BDC, where the piston mo-
mentarily stops. When there is no relative motion between the piston and cylinder
walls, the oil film between these surfaces gets squeezed out by the high forces between
them. When the piston then starts a new stroke, there is very little lubricant between
surfaces, and some metal-to-metal contact occurs, with resulting high friction forces. As
the piston gains speed over the lubricated cylinder wall surface, it drags a film of oil
with it and hydraulic floating occurs. This is the most effective form of lubrication be-
tween moving surfaces, and friction forces are minimized.
It can be seen in Fig. 5 that there is even a small measurable friction force at TDC
and BDC where the piston velocity is considered zero. This shows that there are de-
flections in the connecting components and stretching or compression of the piston oc-
curring at these points due to mass inertia and high acceleration rates. This is the
reason the maximum allowable average piston speed is about 5 to 20 m/sec for all en-
gines regardless of size. With speeds higher than these, there would be a danger of
structural failure, with too small of a safety margin for the materials in the piston as-
semblies of most engines (i.e., iron and aluminum).
The magnitude of the friction forces is about the same for the intake, compres-
sion, and exhaust strokes. It is much higher during the expansion stroke, reflecting the
higher pressure and forces that occur at that time.
The piston assemblies of most engines contribute about half of the total friction
and can contribute as much as 75% at light loads. The piston rings alone contribute
about 20% of total friction. Most pistons have two compression rings and one or two
oil rings. The second compression ring reduces the pressure differential that occurs
437
Friction and Lubrication
ENGINE SETUP 45
A- COMPLETE ENGINE
B- COMPLETE ENGINE MINUS INTAKE 40 A
FIGURE 4
Friction losses for various engine components as measured by motoring of the engine. All losses, which
are given in terms of fmep (psia), increase with increasing engine speed. Reprinted with permission from
SAE Paper No. 730150 © 1973 SAE International, [20].
across the first compression ring during combustion and power stroke. To reduce fric-
tion, the trend has been to make compression rings thinner, some engines having rings
as thin as 1 mm. Oil rings distribute and remove an oil film on the cylinder walls and
sustain no pressure differential. All rings are spring loaded against the walls, which re-
sults in high friction forces.
Adding an additional compression ring can add about 10 kPa to fmep of an en-
gine. Increasing the compression ratio by one will increase fmep by about 10 kPa. In-
creasing the compression ratio also requires heavier bearings on the crankshaft and
connecting rods and may require an additional piston compression ring.
The valve train of an engine contributes about 25% of total friction, crankshaft
bearings contribute about 10% of total, and engine-driven accessories contribute about
15% of total.
In Figs. 6 and 7, friction mean effective pressure 1Lmmep2 is plotted as a func-
tion of engine speed. Engine speed of the ordinate (X axis) can be replaced with aver-
age piston speed without changing the shape of the curves. When data are generated to
make curves like these, a Reynolds number is defined in terms of an average piston
speed such that
Re = UpB/v (24)
438
Friction and Lubrication
40
20
10
10
Exhaust
TDC BDC
Piston Position
per Unit Piston Area (kPa)
30
Piston Friction Force
20
Intake
10
10
Compression
TDC BDC
Piston Position
FIGURE 5
Friction force on piston and piston rings through one engine cycle. Greater friction during
the expansion stroke reflects the higher pressure and forces during that part of the cycle.
Adapted from [80].
where
B = bore
v = kinematic viscosity of lubricating oil
Data from engines of different sizes can be compared at the same piston speed
and temperature if the kinematic viscosity of the lubricating oil is adjusted to be pro-
portional to the cylinder bore B (i.e., if B/v is kept constant). When this is done, the or-
dinate variable of piston speed can be replaced with the Reynolds number, resulting in
the same curves. This is limited by the amount the kinematic viscosity of an oil can be
adjusted without affecting the lubrication of the engine.
439
Friction and Lubrication
30
ENGINE IDLING
20 BMEP LOAD
40 BMEP LOAD
Fmep, psi
20 30 BMEP LOAD
10
0
800 1000 1200 1400 1600 1800 2000 2200 2400
RPM
FIGURE 6
Friction mean effective pressure as a function of engine speed and load for six-cylinder CI engine.
Pressure in psia. Reprinted with permission from SAE Paper No. 730150 © 1973 SAE
International, [158].
10
0
800 1000 1200 1400 1600 1800 2000 2200 2400
RPM
FIGURE 7
Comparison of friction mean effective pressure of a motored engine vs. fired engine. Engine is a six-
cylinder CI engine. Pressure in psia. Reprinted with permission from SAE Paper No. 730150 © 1973
SAE International, [158].
Example Problem 1
A five-cylinder, in-line engine has an 8.15-cm bore, a 7.82-cm stroke, and a connecting rod length
of 15.4 cm. Each piston has a skirt length of 6.5 cm and a mass of 0.32 kg. At a certain engine
speed and crank angle, the instantaneous piston speed is 8.25 m/sec, and clearance between the
piston and cylinder wall is 0.004 mm. SAE 10W-30 motor oil is used in the engine, and at the tem-
perature of the piston–cylinder interface, the dynamic viscosity of the oil is 0.006 N-sec/m2. Cal-
culate the friction force on one piston at this condition.
440
Friction and Lubrication
30
12:1 COMPRESSION RATIO
7:1 COMPRESSION RATIO
25
20 FRICTION
MEP
15
10
PUMPING
0
0 20 40 60 80 100 120 140
LOAD, BMEP
Friction and pumping work versus load at two compression ratios
FIGURE 8
Friction mean effective pressure (fmep) and pumping mean effective pressure (pmep) as a
function of brake mean effective pressure (bmep) for two compression ratios. Engine is a
four-cylinder, 3.25 liter, SI engine, with 9.53 cm bore and 11.40 cm stroke, operating at 1600
RPM. Pressures in psia. Reprinted with permission from SAE Transactions © 1958 SAE
International, [177].
ts = m1dU/dy2 = m1¢U/¢y2
= 10.006 N-sec/m22[18.25 m/sec2/10.000004 m2] = 12,375 N/m2
441
Friction and Lubrication
Example Problem 2
Dynamometer data of a five-cylinder, four-stroke cycle, 260-in.3 displacement, CI engine gives:
at N = 1000 RPM torque t = 230 lbf-ft mechanical efficiency hm = 88%
3000 RPM 255 lbf-ft 78%
5000 RPM 224 lbf-ft 62%
Calculate:
1. indicated power at 3000 RPM
2. fmep at 1500 RPM
3. friction power lost at 4000 RPM
(1) Brake power at 3000 RPM is
#
Wb = [13000212552]/152522 = 145.7 hp
Indicated power is
Use Eqs. (82), (47), (49), and (90) from the chapter, Operating Characteristics, to find
fmep at 5000 RPM:
#
Wb = [15000212242]/152522 = 213.3 hp
Wi = 1213.32/10.622 = 344.0 hp
#
Use Eq. (12) to find fmep at 1000 RPM, at 3000 RPM, and at 5000 RPM:
fmep = A + BN + CN 2
At 1000 RPM: 18.3 = A + B110002 + C1100022
At 3000 RPM 42.0 = A + B130002 + C1300022
At 5000 RPM 79.6 = A + B150002 + C1500022
442
Friction and Lubrication
Use Eq. (90) from the chapter, Operating Characteristics, to find the friction power
lost at 4000 RPM:
#
Wf = [159.1212602140002]/[1396,0002122] = 77.6 hp = 57.9 kW
Engine Accessories
There are many engine and automobile accessories powered off the crankshaft
which reduce the brake power output of the engine. Some of these are continuous
(fuel pump, oil pump, supercharger, engine fan), and some operate only part of the
time (power brake pump, air conditioner compressor, emission control air pump,
power steering pump). When an engine is motored to measure friction, it has been
found that three essential accessories (water pump, fuel pump, and alternator) can
account for as much as 20% of total friction power. The fuel pump and water pump
on many older engines were driven mechanically off the crankshaft. Most modern
engines have electric fuel pumps and some have electric water pumps. The power to
drive these comes from the alternator, which in turn is driven off the engine crank-
shaft. Most engines have a cooling fan that draws external air through the radiator
and blows it through the engine compartment. Some are powered by direct mechan-
ical linkage to the crankshaft. As engine speed goes up, fan speed also goes up.
Power needed to drive an air fan goes up as fan speed cubed, so power requirements
can get high at higher engine speeds. Higher engine speeds often mean higher auto-
mobile velocity, which is when fan cooling is not necessary. At high automobile ve-
locity, enough air is forced through the radiator and engine compartment to
adequately cool the engine just by the forward motion of the car. A fan is not need-
ed. To save power, some fans are driven only when their cooling effect is needed.
This can be done with mechanical or hydraulic linkage that disconnects at higher
speeds or at cooler temperatures (i.e., with a centrifugal or thermal clutch). Most
fans are electrically driven and can be turned on with a thermal switch only when
needed. Automobiles with air conditioners often require a larger fan due to the
added cooling load of the AC condenser.
3 FORCES ON PISTON
Figure 9 shows the forces that act on a piston. The centerline of the cylinder is used as
the X axis, with positive being down in the direction of piston motion during the power
443
Friction and Lubrication
X
Gas
Pressure P
Side Thrust
Force Ft
Friction
Force Ff
FIGURE 9
Force balance on a piston. Side thrust force is a
reaction to the connecting rod force and is in the Connecting
plane of the connecting rod. When the piston passes Rod Force Fr
BDC, the side thrust force switches to the other side
of the cylinder. The connecting rod force and the
resulting side thrust force are greatest during the Crank Angle
power stroke, and this is called the major thrust side.
Lesser forces during the exhaust stroke occur on the
minor thrust side. The friction force is in the
direction opposite to the piston motion and changes
direction after TDC and BDC. Crank Rotation
stroke. The Y axis is in the radial direction outward, with zero at the centerline. A force
balance in the X direction gives
2
a Fx = m1dUp /dt2 = -Fr cos f + P1p/42B ; Ff (25)
where
the sign on the friction force term depends on the crank angle u:
- when 0° 6 u 6 180°
+ when 180° 6 u 6 360°
444
Friction and Lubrication
a Fy = 0 = Fr sin f - Ft (26)
Combining Eqs. (25) and (26) gives the side thrust force on the piston as
This side thrust force is the Y-direction reaction to the force in the connecting
rod and lies in the plane of the connecting rod. From Eq. (27) it can be seen that Ft is
not a constant force but changes with piston position (angle f), acceleration 1dUp /dt2,
pressure (P), and friction force 1Ff2, all of which vary during the engine cycle. During
the power and intake strokes, the side thrust force will be on one side of the cylinder
(the left side for an engine rotating as shown in Fig. 9) in the plane of the connecting
rod. This is called the major thrust side of the cylinder because of the high pressure
during the power stroke. This high pressure causes a strong reaction force in the con-
necting rod, which in turn causes a large side thrust reaction force. During the exhaust
and compression strokes, the connecting rod is on the other side of the crankshaft and
the resulting side thrust reaction force is on the other side of the cylinder (the right side
in Fig. 9). This is called the minor thrust side due to the lower pressures and forces in-
volved and is again in the plane of the connecting rod. The side thrust forces on the pis-
ton are less in planes turned circumferentially away from the plane of the connecting
rod, reaching a minimum in the plane at a right angle to the connecting rod plane.
There will still be a small force reacting to the spring-loaded piston rings.
The side thrust force also varies with crank angular position as the piston moves
back and forth in the cylinder. Thus, there is continuous variation both in the circum-
ferential direction and along the length of the cylinder from TDC to BDC. One result
of this force variation is the variation in wear that occurs on the cylinder walls. The
greatest wear occurs in the plane of the connecting rod on the major thrust side of the
cylinder. Significant, but less, wear will occur on the minor thrust side. This wear will
also vary along the length of the cylinder on both sides. Additional wear to various de-
grees will occur in the other rotation planes and at various distances along the length of
the cylinder. As an engine ages, this wear can become significant in some spots. Even if
the cross section of an engine cylinder is perfectly round when the engine is new, wear
will erode this roundness with time [125].
To reduce friction, modern engines use pistons that have less mass and shorter
skirts. Less mass lowers the piston inertia and reduces the acceleration term in Eq. (27).
Shorter piston skirts reduce rubbing friction because of the smaller surface contact
area. With closer tolerances between piston and cylinder wall, skirts can be made as
short as about 60% of bore diameter. In addition, parts of the skirt can be removed to
lower piston mass. Fewer and smaller piston rings are common compared with earlier
engines, but these also require closer manufacturing tolerances. In some engines, the
wrist pin is offset from center by 1 or 2 mm towards the minor thrust side of the piston.
This reduces the side thrust force and resulting wear on the major thrust side.
The philosophy of some manufacturers is to reduce friction by having a shorter
stroke. However, for a given displacement this requires a larger bore, which results in
445
Friction and Lubrication
10
2.5
.25
FIGURE 10
Oil film thickness between piston
compression ring and cylinder wall as a .1
function of speed and load (bmep).
Film is minimum when the piston stops
at TDC and BDC and oil gets
Power Exhaust Inlet Compression
squeezed out. When the piston moves
in the opposite direction, the oil film is 0 180 360 540 720
again dragged between the surfaces, Crank Angle
reaches a maximum thickness at
maximum piston speed, and again bmep
decreases with decreasing piston
KEY(I) 0
speed. Data are for the major thrust (II) 40 1500 RPM
side of a cylinder in a Perkins AT6-354 (III) 150
diesel engine. Reprinted with (IV) 0
permission, © Institute of Mechanical (V) 40 2200 RPM
Engineers, London, UK, [3]. (VI) 136
greater heat losses due to the larger cylinder surface area. Greater flame travel dis-
tance also increases knock problems. This is why most medium-sized engines (automo-
bile engines) are close to square, with B L S.
Figure 10 shows how the oil film thickness between the piston and cylinder wall
varies with speed during an engine cycle for one circumferential position of the piston.
Piston friction force will be proportional to oil viscosity, engine speed, and imep.
On some large CI engines, piston side thrust force is eliminated by adding a sec-
ondary sliding mechanism called a crosshead (Fig. 11). The crosshead is contained in an
extension of the cylinder and is connected to the crankshaft by the connecting rod. The
piston is connected to the crosshead by a secondary connecting rod which stays paral-
lel with the cylinder walls. This eliminates side forces on the piston, transferring them
to the crosshead, and reducing wear on the main cylinder walls. This system adds mass,
height, and more mechanism to the engine, and is rarely found on smaller or vehicle
engines [150].
Example Problem 3
At the conditions given in Example Problem 1, the engine performs a power stroke in the cylin-
446
Friction and Lubrication
FIGURE 11
Cross-sectional view of large Sulzer RTA62 two-stroke cycle CI engine. The pistons of this engine
are connected to the crankshaft through a secondary crosshead mechanism which eliminates side
thrust forces on the pistons. The crosshead is connected to the crankshaft by connecting rods and
thus experiences the side thrust forces. Reprinted with permission from SAE Paper 851219 © 1985
SAE International, [198].
447
Friction and Lubrication
der described and the crank angle is as shown in Fig. 12. At this point, pressure in the cylinder is
3200 kPa and the compressive force in the connecting rod is 8.1 kN. Calculate the thrust force on
the cylinder wall at this time.
Crank offset equals half of stroke length = 3.91 cm.
Angle between the connecting rod and the centerline of the cylinder:
tan f = 3.91/15.4 = 0.2539
f = 14.25°
Use Eq. (25) to find the acceleration term:
m1dUp /dt2 = -Fr cos f + P1p/42B2 - Ff
= -18.1 kN2 cos114.25°2 + 13200 kPa21p/4210.0815 m22 - 1205N2
= 8638 N
Use Eq. (27) to find the thrust force:
Ft = [-m1dUp /dt2 + P1p/42B2 - Ff]tanf
= [-18638 N2 + 13200 kPa21p/4210.0815 m22 - 1205 N2]tan114.25°2
= 1994N
This force would be in the plane of the connecting rod on the major thrust side of the cylinder.
Ft
Side
Thrust
Force
Piston
Connecting Rod
Right Angle
448
Friction and Lubrication
FIGURE 13
Lubrication of a 1980s automobile engine consisting of a combination of a
pressurized system and splash. The oil pump distributes oil under pressure
through passages in the engine components. Systems usually include filtration and
sometimes an oil cooler. Reprinted with permission from SAE Paper 820111 ©
1982 SAE International, [135].
449
Friction and Lubrication
the piston crowns. Most automobiles actually use dual distribution systems, relying on
splash within the crankcase in addition to the pressurized flow from the oil pump. Most
large stationary engines also use this kind of dual system. Most aircraft engines and a
few automobile engines use a total pressurized system with the oil reservoir located sep-
arate from the crankcase. These are often called dry sump systems (i.e., the crankcase
sump is dry of excess oil). Aircraft do not always fly level, and uncontrolled oil in the
crankcase may not supply proper lubrication or oil pump input when the plane banks or
turns. A diaphragm controls the oil level in the reservoir of a dry sump system, assuring
a continuous flow into the oil pump and throughout the engine. Some automobile en-
gines with overhead camshafts have a secondary oil sump in the engine head to supply
the cam and valve mechanism.
Oil pumps can be electric or mechanically driven off the engine. Pressure at the
pump exit is typically about 300 to 400 kPa. If an oil pump is driven directly off the en-
gine, some means should be built into the system to keep the exit pressure and flow
rate from becoming excessive at high engine speeds.
A time of excess wear is at engine start-up before the oil pump can distribute
proper lubrication. It takes a few engine cycles before the flow of oil is fully estab-
lished, and during this time, many parts are not properly lubricated. Adding to the
problem is the fact that often the oil is cold at engine start-up. Cold oil has much high-
er viscosity, which further delays proper circulation. A few engines have oil preheaters
that electrically heat the oil before start-up. Some engines have preoilers that heat and
circulate the oil before engine start-up. An electric pump lubricates all components by
distributing oil throughout the engine.
It is recommended that turbocharged engines be allowed to idle for a few sec-
onds before they are turned off. This is because of the very high speeds at which the
turbocharger operates. When the engine is turned off, oil circulation stops and lubricat-
ed surfaces begin to lose oil. Stopping the oil supply to a turbocharger operating at
high speed invites poor lubrication and high wear. To minimize this problem, the en-
gine and turbocharger should be allowed to return to low speed (idle) before the lubri-
cation supply is stopped.
The oil systems on some large vehicle engines are designed to remove oil from
the circulation system very slowly and burn it in the combustion chambers. Makeup oil
is then automatically added on a one-for-one basis from a separate oil reservoir filled
with new oil. Replacing old oil with new unused oil keeps the oil circulation system
clean, and can delay the need for an oil change for up to 50,000 miles [166]. Instead of
discarding used oil from an oil change, some ship companies mix this oil with fuel and
burn it in the large CI ship engines. There is some concern for an increase of ash and
wear when this is done.
450
Friction and Lubrication
and an alternate method must be used to lubricate the crankshaft and other compo-
nents in the crankcase. In these engines, oil is carried into the engine with the inlet air
in much the same way as the fuel. When the fuel is added to the inlet air, usually with a
carburetor, oil particles as well as fuel particles are distributed into the flow. The air
flow then enters the crankcase, where it is compressed. Oil particles carried with the air
lubricate the surfaces they come in contact with, first in the crankcase and then in the
intake runner and cylinder.
In some systems (model airplane engines, marine outboard motors, etc.), the oil is
premixed with the fuel in the fuel tank. In other engines (automobiles, some golf carts,
etc.), there is a separate oil reservoir that feeds a metered flow of oil into the fuel sup-
ply line or directly into the inlet air flow. Fuel-to-oil ratio ranges from 30:1 to 400:1, de-
pending on the engine. Some modern high-performance engines have controls that
regulate the fuel–oil ratio, depending on engine speed and load. Under conditions of
high oil input, oil sometimes condenses in the crankcase. Up to 30% of the oil is recir-
culated from the crankcase in some experimental automobile engines. It is desirable to
get at least 3000 miles per liter of oil used. Most small, lower-cost engines have a single
average oil input setting. If too much oil is supplied, deposits form on the combustion
chamber walls and valves will stick (if there are valves). If too little oil is supplied, ex-
cess wear will occur and the piston can freeze in the cylinder. Engines that add oil to the
inlet fuel obviously are designed to use up oil during operation. This oil also con-
tributes to HC emissions in the exhaust due to valve overlap and poor combustion of
the oil vapor in the cylinders. New oils that also burn better as fuel are being developed
for two-stroke cycle engines.
Some experimental two-stroke cycle automobile engines and other medium- and
large-size engines use an external supercharger to compress inlet air. These engines use
pressurized/splash lubrication systems similar to those on four-stroke cycle engines
with the crankcase also serving as the oil sump.
Example Problem 4
A four-cylinder, two-stroke cycle engine, with a 2.65-liter displacement and crankcase compres-
sion, is running at 2400 RPM and an air–fuel ratio of 16.2:1. At this condition, the trapping effi-
ciency is 72%, relative charge is 87%, and the exhaust residual from the previous cycle in each
cylinder is 7%. Oil is added to the intake air flow such that the input fuel-to-oil ratio is 50:1.
Calculate:
1. rate of oil use
2. rate of unburned oil added to the exhaust flow
(1) The total charge trapped in the engine with time is
#
mtc = VdralrcN/n
= 10.00265 m3/cycle211.181 kg/m3210.87212400/60 rev/sec2/11 rev/cycle2
= 0.1089 kg/sec
451
Friction and Lubrication
Oil input
moil = 10.00818 kg/sec2/1502
#
(2) Of the oil input, 72% is trapped in the cylinders and burned, and 28% enters the ex-
haust during valve overlap.
Mass of unburned oil in the exhaust
6 LUBRICATING OIL
The oil used in an engine must serve as a lubricant, a coolant, and a vehicle for remov-
ing impurities. It must be able to withstand high temperatures without breaking down
and must have a long working life. The development trend in engines is toward higher
operating temperatures, higher speeds, closer tolerances, and smaller oil sump capacity.
All of these require improved oils compared with those used just a few years ago. Cer-
tainly, the technology of the oil industry has to continue to improve along with the
technology growth of engines and fuel.
Early engines and other mechanical systems were often designed to consume the
lubricating oil as it was used, requiring a continuous input of fresh oil. The used oil was
either burned up in the combustion chamber or allowed to fall to the ground. Just a
couple of decades back, the tolerances between pistons and cylinder walls was such
that engines burned some oil that seeped past the pistons from the crankcase. This re-
quired a periodic need to add oil and frequent oil changes due to blowby contamina-
tion of the remaining oil. HC levels in the exhaust were high because of the oil in the
combustion chamber. A rule in the 1950s and 1960s was to have an oil change in an au-
tomobile every 1000 miles.
Modern engines run hotter, have closer tolerances which keep oil consumption
down, and have smaller oil sumps due to space limitations. They generate more power
with smaller engines by running faster and with higher compression ratios. This means
higher forces and a greater need for good lubrication. At the same time, many manu-
facturers now suggest changing the oil every 6000 miles. Not only must the oil last
longer under much more severe conditions, but new oil is not added between oil
changes. Engines of the past that consumed some oil required periodic makeup oil to
be added. This makeup oil mixed with the remaining used oil and improved the overall
lubrication properties within the engine. Some modern high-performance automobiles
(Mercedes, Corvette) have sensors in the oil sump that monitor oil level, age, tempera-
ture, etc. [191]. These systems tell the operator when the oil has degraded to a point
452
Friction and Lubrication
that an oil change is required. These vehicles can sometimes go 40,000 km (25,000
miles) or two years between oil changes.
The oils in modern engines must operate over an extreme temperature range.
They must lubricate properly from the starting temperature of a cold engine to beyond
the extreme steady-state temperatures that occur within the engine cylinders. They
must not oxidize on the combustion chamber walls or at other hot spots such as the
center crown of the piston or at the top piston ring. Oil should adhere to surfaces so
that they always lubricate and provide a protective covering against corrosion. This is
often called oiliness. Oil should have high film strength to assure no metal-to-metal
contact even under extreme loads. Oils should be nontoxic and nonexplosive. Lubri-
cating oil must satisfy the following needs:
1. Lubrication
It must reduce friction and wear within the engine. It improves engine efficiency
by reducing friction forces between moving parts.
2. Coolant
3. Removal of contaminants
4. Enhancement of ring seal and reduction of blowby
5. Slow corrosion
6. Stability over a large temperature range
7. Long life span
8. Low cost
The base ingredients in most lubricating oils are hydrocarbon components made
from crude oil. These are species with larger molecular weight is obtained from the dis-
tillation process (Table 1). Various other components are added to create a lubricant
that will allow for maximum performance and life span of the engine. These additives
include the following:
1. Antifoam agents
These reduce the foaming that would result when the crankshaft and other com-
ponents rotate at high speed in the crankcase oil sump.
2. Oxidation inhibitors
Oxygen is trapped in the oil when foaming occurs, and this leads to possible oxi-
dation of engine components. One such additive is zinc dithiophosphate.
3. Pour-point depressant
SAE 10 25–35 28
SAE 20 30–80 38
SAE 30 40–100 41
Adapted from [56].
453
Friction and Lubrication
4. Antirust agents
5. Detergents
These are made from organic salts and metallic salts. They help keep deposits and
impurities in suspension and stop reactions that form varnish and other surface
deposits. They help neutralize acid formed from sulfur in the fuel.
6. Antiwear agents
7. Friction reducers
8. Viscosity index improvers
Viscosity
Lubricating oils are generally rated using a viscosity scale established by the Society of
Automotive Engineering (SAE). Dynamic viscosity is defined from the equation
ts = m1dU/dy2 (13)
where
The higher the viscosity value, the greater is the force needed to move adjacent
surfaces or to pump oil through a passage. Viscosity is highly dependent on tempera-
ture, increasing with decreasing temperature (Fig. 14). In the temperature range of en-
gine operation, the dynamic viscosity of the oil can change by several orders of
magnitude. Oil viscosity also changes with shear, du/dy, decreasing with increasing
shear. Shear rates within an engine range from very low values to extremely high val-
ues in the bearings and between piston and cylinder walls. The change of viscosity over
these extremes can be several orders of magnitude. The following viscosity grades are
commonly used in engines:
SAE 5
SAE 10
SAE 20
SAE 30
SAE 40
SAE 45
SAE 50
The oils with lower numbers are less viscous and are used in cold-weather op-
eration. Those with higher numbers are more viscous and are used in modern high-
temperature, high-speed, close-tolerance engines. Oils become more viscous with
age, because the components with lower molecular weights evaporate quicker.
454
Friction and Lubrication
0.5
0.2
Dynamic Viscosity, (kg/m-sec)
0.1
0.05
SA
SA 0
SA 0
SA 30
SA 0
E5
E4
E
E2
E1
0.02
0.01
0.005
0.002
FIGURE 14
0.001 Dynamic viscosity as a function of
10 20 30 40 50 60 70 80 90 100 110 120 130 140 temperature for common engine oils. Adapted
Temperature, T ( C) from [113].
455
Friction and Lubrication
If oil viscosity is too high, more work is required to pump it and to shear it be-
tween moving parts. This results in greater friction work and reduced brake work and
power output. Fuel consumption can be increased by as much as 15%. Starting a cold
engine lubricated with high-viscosity oil is very difficult (e.g., an automobile at -20°C
or a lawn mower at 10°C).
Multigrade oil was developed so that viscosity would be more constant over the
operating temperature range of an engine. When certain polymers are added to an oil,
the temperature dependency of the oil viscosity is reduced, as shown in Fig. 15. These
oils have low-number viscosity values when they are cold and higher numbers when
they are hot. A value such as SAE 10W-30 means that the oil has properties of 10 vis-
cosity when it is cold 1W = winter2 and 30 viscosity when it is hot. This gives a more
constant viscosity over the operating temperature range (Fig. 15). This is extremely im-
portant when starting a cold engine. When the engine and oil are cold, the viscosity
must be low enough that the engine can be started without too much difficulty. The oil
flows with less resistance and the engine gets proper lubrication. It would be very diffi-
cult to start a cold engine with high-viscosity oil, because the oil would resist engine ro-
tation and poor lubrication would result because of the difficulty in pumping the oil.
On the other hand, when the engine gets up to operating temperature, it is desirable to
have a higher viscosity oil. High temperature reduces the viscosity, and oil with a low
viscosity number would not give adequate lubrication.
Some studies show that polymers added to modify viscosity do not lubricate as
well as the base hydrocarbon oils. At cold temperatures SAE 5 oil lubricates better
than SAE 5W-30, and at high temperatures SAE 30 oil lubricates better. However, if
SAE 30 oil is used, starting a cold engine will be very difficult, and poor lubrication and
very high wear will result before the engine warms up.
The following multigrade oils are commonly available:
456
Friction and Lubrication
0.5
SAE
SAE
20W 0W-4
SA
50
E2
-50
0.2
Dynamic Viscosity, (kg/m-sec)
0.1
SA
E1
0W
-30
0.05
SA
E
5W
-30
0.02
0.01
SA
E
10
0.005
0.002
FIGURE 15
0.001 Dynamic viscosity as a function of temperature
10 20 30 40 50 60 70 80 90 100 110 120 130 140
for common multigrade engine oils. Adapted
Temperature, T ( C) from [113].
457
Friction and Lubrication
standards were up to SL. Standards for oils intended for use in diesel engines were given
a similar set of letter designations: CA, CB, etc. 1C = compression ignition2. In 2002,
the diesel standards were up to CH-4.
Synthetic Oils
A number of synthetically made oils are available that give better performance than
those made from crude oil. They are better at reducing friction and engine wear, have
good detergency properties which keep the engine cleaner, offer less resistance for
moving parts, and require less pumping power for distribution. The main reason for
this better performance is that the base synthetic material is a homogeneous fluid with
similar molecules of the same structure and molecular weight. Crude oil lubricant, on
the other hand, has a base fluid structure made up of dissimilar molecules with a range
of molecular weights. With good thermal properties, synthetic oils provide better en-
gine cooling and less variation in viscosity. Because of this, they contribute to better
cold-weather starting and can reduce fuel consumption by as much as 15%. These oils
cost several times as much as those made from crude oil. However, they can be used
longer in an engine, with 24,000 km (15,000 miles) being the oil change period suggest-
ed by most manufacturers.
Various oil additives and special oils that can be added in small quantities to stan-
dard oils in the engine are available on the market.These claim, with some justification, to
improve the viscous and wear resistance properties of normal oils. One major improve-
ment that some of them provide is that they stick to metal surfaces and do not drain off
when the engine is stopped, as most standard oils do. The surfaces are thus lubricated im-
mediately when the engine is next started. With standard oils, it takes several engine rota-
tions before proper lubrication occurs, a major source of wear.
Solid lubricants, such as powdered graphite, have been developed and tested in
some engines. These are attractive for adiabatic engines and engines using ceramic com-
ponents, which generally operate at much higher temperatures. Solid lubricants remain
functional at high temperatures that would break down and destroy more conventional
oils. Distribution is a major difficulty when using solid lubricants.
7 OIL FILTERS
A filtration system to remove impurities from the engine oil is included in most pressur-
ized oil systems. One of the duties of engine oil is to clean the engine by carrying contam-
inant impurities in suspension as it circulates.As the oil passes through filters that are part
of the flow passage system, these impurities are removed, cleaning the oil and allowing it
to be used for a greater length of time. Contaminants get into an engine in the incoming
air or fuel, or can be generated within the combustion chamber when other-than-ideal sto-
ichiometric combustion occurs. Dust and other impurities are carried by the incoming air.
Some, but not all, of these are removed by an air filter. Fuels have trace amounts of impu-
rities such as sulfur, which create contaminants during the combustion process. Even pure
fuel components form some contaminants, such as solid carbon in some engines under
some conditions. Many engine impurities are carried away with the engine exhaust, but
some get into the interior of the engine, mainly in the blowby process. During blowby, fuel,
air, and combustion products are forced past the pistons into the crankcase, where they
458
Friction and Lubrication
mix with the engine oil. Some of the water vapor in the exhaust products condenses in
the crankcase, and the resulting liquid water adds to the contaminants. The gases of
blowby pass through the crankcase and are routed back into the air intake. Ideally,
most of the contaminants are trapped in the oil, which then contains dust, carbon, fuel
particles, sulfur, water droplets, and many other impurities. If these were not filtered
out of the oil, they would be spread throughout the engine by the oil distribution sys-
tem. Also, the oil would quickly become dirty and lose its lubricating properties, result-
ing in greater engine wear. Instead of an oil filter, Cummins Diesel uses a centrifuge on
some of their large engines [166]. The oil impurities are forced to the outer edge of the
centrifuge where they are removed.
Flow passages in a filter are not all the same size but usually exist in a normal
bell-shaped size distribution (Fig. 16). This means that most larger particles will be fil-
tered out as the oil passes through the filter, but a few as large as the largest passages
will get through.
The choice of filter pore size is a compromise. Better filtration will be obtained
with smaller filter pores, but this requires a much greater flow pressure to push the oil
through the filter. This also results in the filter becoming clogged quicker and requiring
earlier filter cartridge change. Some filter materials or material of too small a pore size
can even remove some additives from the oil. Filters are made from cotton, paper, cel-
lulose, and a number of different synthetic materials. Filters are usually located just
downstream from the oil pump exit.
As a filter is used, it slowly becomes saturated with trapped impurities. As these
impurities fill the filter pores, a greater pressure differential is needed to keep the same
flow rate. When this needed pressure differential gets too high, the oil pump limit is
reached and oil flow through the engine is slowed. The filter cartridge should be re-
placed before this happens. Sometimes, when the pressure differential across a filter
gets high enough, the cartridge structure will collapse and a hole will develop through
Percent of Pores
FIGURE 16
Pore size distribution for common filters.
459
Friction and Lubrication
the cartridge wall. Most of the oil pumped through the filter will then follow the path
of least resistance and flow through the hole. This short circuit will reduce the pressure
drop across the filter, but the oil does not get filtered.
There are several ways in which the oil circulation system can be filtered:
1. Full-flow oil filtration. All oil flows through the filter. The filter pore size must be
fairly large to avoid extreme pressures in the resulting large flow rate. This results
in some larger impurities in the oil.
2. Bypass oil filtration. Only part of the oil leaving the pump flows through the filter,
the rest bypassing it without being filtered. This system allows the use of a much
finer filter, but only a percentage of the oil gets filtered during each circulation loop.
3. Combination. Some systems use a combination of full-flow and bypass. All the
oil first flows through a filter with large pores and then some of it flows through a
second filter with small pores.
4. Shunt filtration. This is a system using a full-flow filter and a bypass valve. All oil
at first flows through the filter. As the filter cartridge dirties with age, the pres-
sure differential across it needed to keep the oil flowing increases. When this
pressure differential gets above a predetermined value, the bypass valve opens
and the oil flows around the filter. The filter cartridge must then be replaced be-
fore filtering will again occur.
HISTORIC—FILTER CARTRIDGES
During the 1950s, one manufacturer sold an accessory oil filtration system that could be
added onto an automobile engine. The system consisted of a filter cartridge canister that
was bolted onto the oil distribution system. The filter cartridge that was to be used with
this canister was a standard roll of toilet paper.
8 CRANKCASE EXPLOSIONS
There is a remote chance of having an explosion in the crankcase of a reciprocating en-
gine [152]. The crankcase contains oxygen (air) and oil vapor and/or fuel vapor from
blowby. There is, therefore, a remote chance of an explosion if an ignition source devel-
ops. This could be a hot spot (bad bearing, worn component, etc.), flame getting past a
piston (broken piston ring), or a spark from a broken component. Even when oxygen,
fuel, and ignition are all present, the chance of an explosion is very remote because the
very rich fuel–air mixture will fail to ignite. If an explosion does happen in a large en-
gine, the oil sump can be ruptured with dangerous and destructive results. An explosion
in a small engine is not considered dangerous. Because of the very small volume of the
crankcase, the likelihood of a combustible mixture is extremely remote, and if an explo-
sion did occur, it probably would not damage the engine. The mechanical structure and
strength of a small engine relative to the size of the crankcase is such that an explosion
probably would not even be noticed, and the engine would continue to run. Interna-
tional safety codes consider safe any engine with less than 6.1 m3 of crankcase volume
or bores less than 0.2 m. Engines larger than this require explosion relief valves.
460
Friction and Lubrication
Explosion relief valves are designed to release the pressure buildup caused by a
crankcase explosion without damaging the engine or injuring anyone near the engine
[164]. A well-designed valve will open, direct the flow of hot gases away from opera-
tors, arrest the flame, and immediately close without allowing air to flow back into the
crankcase. Most valves will open when a pressure pulse in the range of 5 to 20 kPa (0.7
to 3 psia) is detected. The engine will usually continue to run, but should be stopped to
determine the cause of explosion. Engines should be allowed to cool before they are
opened; the onrush of air into the hot crankcase could be very dangerous.
Brake power output from an engine is less than the power generated in the combustion
chambers, due to engine friction. Two types of friction occur that result in dissipation
and loss of useful power. Mechanical friction between moving parts is a major source
of engine power loss, with piston motion in the cylinders causing a large percentage of
this loss. Fluid friction occurs in the intake and exhaust systems, in flow through valves,
and because of motion within the cylinders. Operation of engine accessories, although
not friction in the normal sense, is often included as part of the engine friction load.
That is because the accessories are powered, directly or indirectly, off the engine crank-
shaft and reduce final crankshaft output power.
To minimize friction and to reduce engine wear, lubrication systems are a major
required facet of any engine. Oil distribution can be accomplished by using a pressur-
ized system supplied by a pump, as with automobile engines, or by splash distribution,
as on many small engines. In addition to lubricating, engine oil helps to cool the engine
and is a vehicle for removing engine contaminants.
PROBLEMS
1 The connecting rod in Fig. 12 experiences a force of 1000 N in the position shown during
the power stroke of a four-cylinder, four-stroke cycle SI engine operating at 2000 RPM.
Crankshaft offset equals 3.0 cm and connecting rod length equals 9.10 cm.
Calculate:
(a) Side thrust force felt in the cylinder wall at this moment. [N]
(b) Distance the piston has traveled from TDC. [cm]
(c) Engine displacement if S = 0.94 B. [L]
(d) Side thrust force felt in the cylinder wall when the piston is at TDC. [N]
2 Why do cylinders in IC engines get out-of-round as the engine is operated for a long peri-
od of time? Why is wear on the cylinder walls not the same along the length of the cylin-
der? Theoretically, why is piston frictional force equal to zero at TDC and BDC? In
actuality, why is piston frictional force not equal to zero at TDC and BDC?
3 A six-cylinder IC engine has a 6.00-cm bore, a 5.78-cm stroke, and a connecting rod length
of 11.56 cm. In the power stroke of the cycle for one cylinder at a crank position of 90°
aTDC, the pressure in the cylinder is 4500 kPa and the sliding friction force on the piston
is 0.85 kN. Piston acceleration at this point can be considered zero.
461
Friction and Lubrication
Calculate:
(a) Force in the connecting rod at this point. [kN] Is it compressive or tensile?
(b) Side thrust force on the piston at this point. [kN] Is it on the major thrust side or the
minor thrust side?
(c) Side thrust force on the piston at this point if the wrist pin is offset 2 mm to reduce the
side thrust force. (Assume rod force and friction force are the same as above.) [kN]
4 A V6, two-stroke cycle SI automobile engine has a 3.1203-inch bore and 3.45-inch stroke.
The pistons have a height of 2.95 inches and diameter of 3.12 inches. At a certain point
during the compression stroke piston speed in one cylinder is 30.78 ft/sec. The lubricating
oil on the cylinder walls has a dynamic viscosity of 0.000042 lbf-sec/ft2. Calculate the fric-
tion force on the piston under this condition. [lbf]
5 A four-cylinder, four-stroke cycle, 2.8-liter, opposed-cylinder SI engine has brake mean ef-
fective pressure and mechanical efficiency as follows:
at 1000 RPM bmep = 828 kPa hm = 90%
2000 RPM bmep = 828 kPa hm = 88%
3000 RPM bmep = 646 kPa hm = 82%
Calculate:
(a) Brake power at 2000 RPM. [kW]
(b) Friction mean effective pressure at 2500 RPM. [kPa]
(c) Friction power lost at 2500 RPM. [kW]
6 A 110-in.3-displacement, six-cylinder SI automobile engine operates on a two-stroke cycle
with crankcase compression and throttle body fuel injection. With AF = 17.8 and the en-
gine running at 1850 RPM, the automobile cruises at 65 mph and gets 21 miles per gallon
of gasoline. Oil is added to the inlet air at a rate such that input fuel-to-oil ratio is 40:1. Rel-
ative charge is 64% and the exhaust residual from the previous cycle is 6%. Combustion
efficiency hc = 100% and the density of gasoline rg = 46.8 lbm/ft3.
Calculate:
(a) Rate of oil use. [gal/hr]
(b) Trapping efficiency of the engine. [%]
(c) Rate of unburned oil added to the exhaust flow. [gal/hr]
7 When a supercharger is installed on the four-stroke cycle SI engine with a compression
ratio rc = 9.2:1, the indicated thermal efficiency at WOT is decreased by 6%. Mass of air
in the cylinders is increased by 22% when operating at the same speed of 2400 RPM. En-
gine mechanical efficiency stays the same, except that 4% of the brake crankshaft output
is needed to run the supercharger.
Calculate:
(a) Indicated thermal efficiency without a supercharger. [%]
(b) Indicated thermal efficiency with a supercharger. [%]
(c) Percent increase of indicated power when a supercharger is installed. [%]
(d) Percent increase of brake power when a supercharger is installed. [%]
DESIGN PROBLEMS
1D Design a two-stroke cycle SI engine with crankcase compression that uses a conventional
oil distribution system (i.e., a pressurized system with an oil pump and an oil reservoir in
the crankcase).
462
Friction and Lubrication
463
This page intentionally left blank
References
From References of Engineering Fundamentals of the Internal Combustion Engine, Second Edition.
Willard W. Pulkrabek. Copyright © 2004 by Pearson Education, Inc. All rights reserved.
465
References
[1] Abthoff, J., H. Schuster, H. Langer, and G. Loose, “The Regenerable Trap Oxidiz-
er—An Emission Control Technique for Diesel Engines,” SAE paper 850015,
1985.
[2] Alkidas, A. C. and J. P. Myers, “Transient Heat–Flux Measurements in the Com-
bustion Chamber of a Spark Ignition Engine,” Journal of Heat Transfer, ASME
Trans., vol. 104, pp. 62–67, 1982.
[3] Allen, D. G., B. R. Dudley, J. Middletown, and D. A. Panka, “Prediction of Piston
Ring–Cylinder Bore Oil Film Thickness in Two Particular Engines and Correla-
tion with Experimental Evidences,” Piston Ring Scuffing, p. 107, London: Me-
chanical Engineering Pub. Ltd., 1976.
[4] Amann, C. A., “Control of the Homogeneous-Charge Passenger Car Engine—
Defining the Problem,” SAE paper 801440, 1980.
[5] Amann, C. A., “Power to Burn,” Mechanical Engineering, ASME, vol. 112, no. 4,
pp. 46–54, 1990.
[6] Amsden, A. A., T. D. Butler, P. J. O’Rourke, and J. D. Ramshaw, “KIVA—A Com-
prehensive Model for 2-D and 3-D Engine Simulations,” SAE paper 850554,
1985.
[7] Amsden, A. A., J. D. Ramshaw, P. J. O’Rourke, and J. K. Dukowicz, “KIVA—A
Computer Program for Two- and Three-Dimensional Fluid Flows with Chemical
Reactions and Fuel Sprays,” report LA-10245-MS, Los Alamos National Labora-
tory, 1985.
[8] “A Stirling Briefing,” NASA, Cleveland: Lewis Research Center, March 1987.
[9] “A Survey of Variable Valve Actuation,” Automotive Engineering, vol. 98, no. 1,
pp. 29–33, 1990, SAE International.
[10] Austen, A. E. W. and W. T. Lyn, “Relation Between Fuel Injection and Heat Re-
lease in a Direct-Injection Engine and the Nature of the Combustion Processes,”
Proc. Institute of Mechanical Engineers, pp. 47–62, 1960.
[11] Automotive Engineering, a monthly publication by SAE International.
[12] Birch, S., “NGVs in the U.K.,” Automotive Engineering, vol. 102, no. 3, pp. 26–27,
1994, SAE International.
[13] Birch, S., “Thermo Accumulator,” Automotive Engineering, vol. 100, no. 2, pp.
85–86, 1992, SAE International.
466
References
[14] Birch, S., “Two-Stroke Power,” Automotive Engineering, vol. 100, no. 8, pp. 45–47,
1992, SAE International.
[15] Birch, S., J. Yamaguchi, A. Demmler, and K. Jost, “Honda’s Oval-Piston Mega-
Bike,” Automotive Engineering, vol. 100, no. 6, pp. 46–47, 1992, SAE International.
[16] Borgnake, C., V. S. Arpaci, and R. J. Tabaczynski, “A Model for the Instanta-
neous Heat Transfer and Turbulence in a Spark Ignition Engine,” SAE paper
800287, 1980.
[17] Bracco, F. V., “Modeling of Engine Sprays,” SAE paper 850394, 1985.
[18] Brandstatter, W., R. J. R. Johns, and G. Wigley, “The Effect of Inlet Port Geome-
try on In-Cylinder Flow Structure,” SAE paper 850499, 1985.
[19] Brooks, D., “Development of Reference Fuel Scales for Knock Rating,” SAE
Journal, vol. 54, no. 8, August 1946.
[20] Brown, W. L., “The Caterpillar imep Meter and Engine Friction,” SAE paper
730150, 1973.
[21] Butler, T. D., L. D. Cloutman, J. K. Dukowicz, and J. D. Ramshaw, “Multidimen-
sional Numerical Simulation of Reactive Flow in Internal Combustion Engines,”
Proc. Energy Combustion Science, vol. 7, pp. 293–315, 1981.
[22] Cameron, K., “NR750,” Cycle World, pp. 30–35, Jan. 1992.
[23] Catania, A. E., C. Dongiovanni, and A. Mittica, “Further Investigation into the
Statistical Properties of Reciprocating Engine Turbulence,” JSME International
Journal, vol. 35, pp. 255–265, 1992.
[24] Catania, A. E. and A. Mittica, “Autocorrelation and Autospectra Estimation of
Reciprocating Engine Turbulence,” Trans. ASME, J. Eng. Gas Turbines Power,
vol. 112, 1990.
[25] Catania, A. E. and A. Mittica, “Extraction Techniques and Analysis of Turbulence
Quantities from In-Cylinder Velocity Data,” Trans. ASME, J. Eng. Gas Turbines
Power, vol. 111, 1989.
[26] Catania, A. E. and A. Mittica, “Induction System Effects on Small-Scale Turbu-
lence in a High-Speed Diesel Engine,” Trans. ASME, J. Eng. Gas Turbines Power,
vol. 109, 1987.
[27] Caton, J., “A Brief Review of Coal-Fueled Engines,” Internal Combustion Engine
Division Newsletter, ASME, Summer 1995.
[28] Chapman, M., J. M. Novak, and R. A. Stein, “Numerical Modeling of Inlet and
Exhaust Flows in Multi-Cylinder Internal Combustion Engines,” Flows in Inter-
nal Combustion Engines, ASME, 1982.
[29] Cummins, C. L. Jr, Internal Fire, SAE International Inc., 1989.
[30] Demmler, A., “Smog-Treating Catalyst,” Automotive Engineering, vol. 103, no. 8,
p. 32, 1995, SAE International.
[31] Diesel and Gas Turbine Worldwide, a monthly publication by Diesel and Gas Tur-
bine Publications.
[32] Dinsdale, S., A. Roughton, and N. Collings, “Length Scale and Turbulence Inten-
sity Measurements in a Motored Internal Combustion Engine,” SAE paper
880380, 1988.
467
References
468
References
469
References
470
References
471
References
[118] Tabaczynski, R. J., “Turbulence and Turbulent Combustion in Spark Ignition En-
gines,” Proc. Energy Combustion Science, vol. 2, pp. 143–165, 1976.
[119] Tabaczynski, R. J., F. H. Trinker, and B. A. S. Shannon, “Further Refinement and
Validation of a Turbulent Flame Propagation Model for Spark Ignition Engines,”
Combustion and Flame, vol. 39, pp. 111–122, 1980.
[120] Taylor, C. F., The Internal Combustion Engine in Theory and Practice. Cam-
bridge, MA: M.I.T. Press, 1977.
[121] “The Changing Nature of Gasoline,” Automotive Engineering, vol. 102, no. 1, pp.
99–102, 1994, SAE International.
[122] “The Road to Clean Air: Powered with Alternate Fuels,” pamphlet by Wisconsin
Alternate Fuels Task Force, 1994.
[123] Thomas, F. J., J. S. Ahluwalia, E. Shamah, and G. W. Van der Horst, “Medium-
Speed Diesel Engines Part I: Design Trends and the Use of Residual/Blended
Fuels,” ASME paper 84-DGP-15, 1984.
[124] Thompson, K. D., “Applying Parametrics and AutoCad in Engine Design,”
Cadence, pp. 63–68, March 1988.
[125] Ting, L. L., and J. E. Mayer Jr., “Piston Ring Lubrication and Cylinder Bore Wear
Analyses, Part II—Theory Verification,” Trans. ASME, J. Lub. Tech., pp. 258–266,
1974.
[126] Tizard and Pye, Philosophical Magazine, July 1922.
[127] Uzkan, T., Flows in Internal Combustion Engines—II, FED-vol. 20, pp. 39–46,
ASME, New York, 1984.
[128] Uzkan, T., C. Borgnakke, and T. Morel, “Characterization of Flow Produced by a
High-Swirl Inlet Port,” SAE paper 830266, 1983.
[129] Uzkan, T., W. G. Tiederman, and J. M. Novak, International Symposium on Flows
in Internal Combustion Engines—III, FED-vol. 28, pp. 125–134, ASME, New
York, 1985.
[130] Valenti, M., “Alternate Fuels: Paving the Way to Energy Independence,”
Mechanical Engineering, vol. 113, no. 12, pp. 42–46, 1991, ASME.
[131] Valenti, M., “Insulating Catalytic Converters,” Mechanical Engineering, vol. 117,
no. 5, pp. 14–16, 1995, ASME.
[132] Valenti, M., “Pollution-Reducing Cars,” Mechanical Engineering, vol. 117, no. 7, p.
12, 1995, ASME.
[133] “Variable Valve Actuation,” Automotive Engineering, vol. 99, no. 10, pp. 12–16,
1991, SAE International.
[134] Wakisaka, T., Y. Shimamoto, and Y. Isshiki, “Three Dimensional Numerical
Analysis of In-Cylinder Flows in Reciprocating Engines,” SAE paper 860464,
1986.
[135] Walker, J. W., “The GM 1.8 Liter Gasoline Engine Designed by Chevrolet,” SAE
paper 820111, 1982.
[136] Wallace, T. F., “Buick’s Turbocharged V-6 Powertrain for 1978,” SAE paper
780413, 1978.
[137] Watkins, A. P., A. D. Gosman, and B. S. Tabrizi, “Calculation of Three Dimension-
al Spray Motion in Engines,” SAE paper 860468, 1986.
472
References
[138] Wentworth, J. T., “Effects of Top Compression Ring Profile on Oil Consumption
and Blowby with the Sealed Ring-Orifice Design,” SAE paper 820089, 1982.
[139] Wise, D. B., The Illustrated Encyclopedia of the World’s Automobiles. New York:
A and W Publishers, 1979.
[140] Witze, P. O., “Measurements of the Spatial Distribution and Engine Speed De-
pendence of Turbulent Air Motion in an I. C. Engine,” SAE paper 770220, 1977.
[141] Woehrle, W. J., “A History of the Passenger Car Tire: Part I,” Automotive Engi-
neering, vol. 103, no. 9, pp. 71–75, 1995.
[142] Yamada, T., T. Inoue, A. Yoshimatsu, T. Hiramatsu, and M. Konishi, “In-Cylinder
Gas Motion of Multivalve Engine—Three Dimensional Numerical Simulation,”
SAE paper 860465, 1986.
[143] Yamaguchi, J., “Honda’s Oval-Piston Mega-Bike,” Automotive Engineering, vol.
100, no. 6, pp. 46–47, 1992.
[144] Arai, M., M. Tabata, and H. Hiroyasu, “Disintegrating Process and Spray Charac-
teristics of Fuel Jet Injected by a Diesel Nozzle,” SAE paper 840275, 1984.
[145] Ashley, A., “Fuel Cells Start to Look Real,” Automotive Engineering Internation-
al, vol. 109, no. 3, pp. 64–80, 2001, SAE International.
[146] “A Survey of Variable Valve Actuation,” Automotive Engineering International,
vol. 98, no. 1, pp. 29–33, 1990, SAE International.
[147] Auburn–Cord–Duesenberg Museum, Auburn, IN, private correspondence in 2002.
[148] Birch, S., “Combustion and Expansion at Saab,” Automotive Engineering Interna-
tional, vol. 109, no. 3, pp. 46–48, 2001, SAE International.
[149] Birch, S., “New Saab and Citroen Technology at Geneva,” Automotive Engineer-
ing International, vol. 108, no. 5, pp. 96–101, 2000, SAE International.
[150] Birch, S., “Two-Seaters from MG and Fiat,” Automotive Engineering Internation-
al, vol. 103, no. 6, pp. 51–53, 1995, SAE International.
[151] Birch, S., “Variations on a Theme by Saab,” Automotive Engineering Internation-
al, vol. 109, no. 4, pp. 54–57, 2001, SAE International.
[152] Blizzard, D. T., and W. N. Shade, “Crankcase Explosion Safety Revisited—Good
Maintenance Practices Necessary for Prevention,” Diesel & Gas Turbine World-
wide, pp. 10–14, March 1996, Diesel and Gas Turbine Publications.
[153] “BMW Returns to Formula One with V10,” Automotive Engineering Internation-
al, vol. 108, no. 3, pp. 42, 2000, SAE International,
[154] “BMW’s Hydrogen Message,” Automotive Engineering International, vol. 110,
no. 5, 2002, SAE International.
[155] Bower, G. R., and D. E. Foster, “A Comparison of the Bosch and Zuech Rate of
Injection Meters,” SAE paper 910724, 1991, SAE International.
[156] Broge, J. L., “Revving up for Diesel,” Automotive Engineering International, vol.
110, no. 2, pp. 40–49, 2002, SAE International.
[157] Broge, J. L., “Smog-Eating Technology,” Automotive Engineering International,
vol. 108, no. 2, pp. 141–142, 2000, SAE International.
[158] Brown, W. L., “The Caterpillar imep Meter and Engine Friction,” SAE paper
730150, 1973.
473
References
[159] Buchholz, K., “Chevy Revs for 2002 IRL Season,” Automotive Engineering Inter-
national, vol. 110, no. 3, pp. 33–34, 2002, SAE International.
[160] “Butane/Propane Mixtures as Fleet Fuel,” Automotive Engineering Internation-
al, vol. 107, no. 12, pp. 41–44, 1999, SAE International.
[161] Carney, D., “Denso Heats Up Diesels,” Automotive Engineering International,
vol. 109, no. 7, pp. 47–48, 2001, SAE International.
[162] Carney, D., “Developments in Fuel Cells,” Automotive Engineering International,
vol. 110, no. 3, pp. 47–52, 2002, SAE International.
[163] Catania, A. E., C. Dongiovanni, and A. Mittica, “Further Investigation into the
Statistical Properties of Reciprocating Engine Turbulence,” JSME International
Journal, series II, vol. 35, no. 2, pp. 255–265, 1992, JSME.
[164] Chellini, R., “Improved Design Explosion Relief Valves,” Diesel & Gas Turbine
Worldwide, p.46, March 2002, Diesel and Gas Turbine Publications.
[165] Clymer, F., and L. R. Henry, Ford Model A Album, Polyprints, 1960.
[166] Cummins Diesel, private communication, 2002.
[167] Czadzeck, G. H., “Ford’s 1980 Central Fuel Injection System,” SAE paper
790742, 1979.
[168] Date, T., S. Yagi, A. Ishizuya, and I. Fuji, “Research and Development of the
Honda CVCC Engine,” SAE paper 740605, 1974.
[169] DeAngelis, G., E. P. Francis, and L. R. Henry, The Ford Model A, Motor Cities
Publishing Company, 1983.
[170] Dent, J. C., and P. S. Mehta, “Phenomenological Combustion Model for a Quies-
cent Chamber Diesel Engine,” SAE paper 811235, 1981.
[171] “Emission Standards,” DieselNet Website, 2002.
[172] “Emissions with Butane/Propane Blends,” Automotive Engineering Internation-
al, vol. 104, no. 11, pp. 49–54, 1996, SAE International.
[173] “Engine Tech 2001,” Diesel & Gas Turbine Worldwide, pp. 45–48, Dec 2001,
Diesel and Gas Turbine Publications.
[174] “EPA Approves Clean Diesel Measure,” Automotive Engineering International,
vol. 109, no. 2, p. 240, 2001, SAE International.
[175] “Federal and California Exhaust and Evaporative Emission Standards for Light-
Duty Vehicles and Light-Duty Trucks,” EPA Website, 2002.
[176] “Ford Invention Will Promote Natural Gas Vehicles,” Automotive Engineering
International, vol. 106, no. 6, p. 120, 1998, SAE International.
[177] Gish, R. E., J. D. McCullough, J. B. Retzloff, and H. T. Mueller, “Determination of
True Engine Friction,” SAE Trans., vol. 66, pp. 649–661, 1958, SAE International.
[178] Glockler, O., H. Knapp, and H. Manger, “Present Status and Future Develop-
ment of Gasoline Fuel Injection Systems for Passenger Cars,” SAE paper
800467, 1980.
[179] Greiner, M., P. Romann, and U. Steinbrenner, “BOSCH Fuel Injectors—New
Developments,” SAE paper 870124, 1987.
[180] Hamburg, D. R., and J. E. Hyland, “A Vaporized Gasoline Metering System for
Internal Combustion Engines,” SAE paper 760288, 1976.
474
References
[181] Hames, R. J., R. D. Straub, and R. W. Amann, “DDEC Detroit Diesel Electronic
Control,” SAE paper 850542, 1985.
[182] Hara, S., A. Hidaka, N. Tomisawa, M. Nakamura, and T. Todo, “Application of a
Variable Valve Event and Timing System to Automotive Engines,” SAE paper
2000-10-1224, 2000.
[183] Hardenberg, H. O., and F. W. Hase, “An Empirical Formula for Computing the
Pressure Rise Delay of a Fuel from its Cetane Number and from the Relevant
Parameters of Direct-Injection Diesel Engines,” SAE paper 790493, 1979.
[184] “Harm Free Use of Diesel Additives,” Automotive Engineering International, vol.
107, no. 7, pp. 84–88, 1999, SAE International.
[185] “Holley’s Nitrous Oxide Systems (NOS) Brand,” Holley Website, 2002.
[186] “Hydrogen as an Alternative Automobile Fuel,” Automotive Engineering Inter-
national, vol. 102, no. 10, 1994, SAE International.
[187] Jost, K., K. Buchholz, and R. Gehm, “Advances in Fuel-Cell Development,”
Automotive Engineering International, vol. 110, no. 6, pp. 67–71, 2002, SAE
International.
[188] Jost, K., “Fuel Cell Autonomy,” Automotive Engineering International, vol. 110,
no. 2, pp. 35–37, 2002, SAE International.
[189] Jost, K., “Fuel-Stratified Injection from VW,” Automotive Engineering Interna-
tional, vol. 109, no. 1, pp.63–65, 2001, SAE International.
[190] Jost, K., “Mercedes-Benz Launches Cylinder Cutout,” Automotive Engineering
International, vol. 107, no. 1, pp. 38–39, 1999, SAE International.
[191] Jost, K., “New Diesel V8 for S-Class,” Automotive Engineering International, vol.
109, no. 1, pp. 78–80, 2001, SAE International.
[192] Kates, E. J., Diesel and High Compression Gas Engines, American Technical So-
ciety, 1954.
[193] Khan, I. M., G. Greeves, and C. H. T. Wang, “Factors Affecting Smoke and
Gaseous Emissions from Direct Injection Engines and a Method of Calculation,”
SAE paper 730169, 1973.
[194] Li, C. H., “Piston Thermal Deformation and Friction Considerations,” SAE paper
820086, 1982.
[195] Liou, T. M., M. Hall, D. A. Santavicca, and F. N. Bracco, “Laser Doppler Velocime-
try Measurements in Valved and Ported Engines,” SAE paper 840375, 1984.
[196] “Lotus Active-Valve Control,” Automotive Engineering International, vol. 101,
no. 1, pp. 97–98, 1993, SAE International.
[197] Lumley, John L., Engines, An Introduction, Cambridge University Press, 1999.
[198] Lustgarten, G. A., and S. Winterthur, “The Latest Sulzer Marine Diesel Engine
Technology,” SAE Technical Paper 851219, 1985.
[199] Malchow, G. L., S. C. Sorenson, and R. O. Buckius, “Heat Transfer in the Straight
Section of an Exhaust Port of a Spark Ignition Engine,” SAE paper 790309, 1979.
[200] “MAN B & W Announces IS Diesels,” Diesel & Gas Turbine Worldwide, p.54,
April 1999, Diesel and Gas Turbine Publications.
[201] Matekunas, F. A., “Modes and Measures of Cyclic Combustion Variability,” SAE
paper 830337, 1983.
475
References
[202] Matsui, Y., T. Kamimoto, and S. Matsuoka, “Formation and Oxidation Processes
of Soot Particles in a D.I. Diesel Engine—An Experimental Study Via the Two-
Color Method,” SAE paper 820464, 1982.
[203] Matsumura, R., K. Higashiyama, and K. Kojima, “The Turbocharged 2.8L Engine
for the Datsun 280ZX,” SAE paper 820442, 1982.
[204] Matsuo, I., S. Nakazawa, H. Maeda, and E. Inada, “Development of a High-Per-
formance Hybrid Propulsion System Incorporating a CVT,” SAE paper 2000-01-
0992, 2000.
[205] “Mayflower’s Variable Engine Technology,” Automotive Engineering Interna-
tional, vol. 110, no. 1, pp. 43–44, 2002, SAE International.
[206] “Mitsubishi Variable Displacement and Valve Timing/Lift,” Automotive Engi-
neering International, vol. 101, no. 1, pp. 99–100, 1993, SAE International.
[207] Moklegaard, L.,A. G. Stefanopolulou, and J. Schmidt,“Transition from Combustion
to Variable Compression Braking,” SAE paper 011228, 2001, SAE International.
[208] “More than One Way to Stop a Truck,” Mechanical Engineering, vol. 121, no. 11,
pp. 34–36, 1999, ASME International.
[209] Mullins, P., “Future Emissions Control at Wartsila,” Diesel & Gas Turbine World-
wide, pp. 32–35, May 2001, Diesel and Gas Turbine Publications.
[210] Mullins, P., “Lubricating Landfill Gas Engines,” Diesel & Gas Turbine Worldwide,
pp. 40–42, March 2000, Diesel and Gas Turbine Publications.
[211] Mullins, P., “Ro-Ro Vessels Will have Water Injection,” Diesel & Gas Turbine
Worldwide, pp. 28–32, Sept 1999, Diesel and Gas Turbine Publications.
[212] Mullins, P., “Shipbuilding in Asia,” Diesel & Gas Turbine Worldwide, pp. 22–26,
March 2001, Diesel and Gas Turbine Publications.
[213] “Natural Gas Fueling of Diesel Engines,” Automotive Engineering International,
vol. 104, no. 11, pp. 87–90, 1996, SAE International.
[214] “Next-Generation Power Sources,” Automotive Engineering International, vol.
107, no. 9, p. 57, 1999, SAE International.
[215] “NOx Reduction at a Vermont Ski Resort,” Diesel & Gas Turbine Worldwide,
pp.14–16, July 2002, Diesel and Gas Turbine Publications.
[216] Pierik, R. J., and J. F. Burkhard, “Design and Development of a Mechanical Vari-
able Valve Actuation System,” SAE paper 2000-01-1221, 2000.
[217] Pischinger, M., W. Salber, F. van der Staay, H. Baumgarten, and H. Kemper, “Ben-
efits of the Electromechanical Valve Train in Vehicle Operation,” SAE paper
2000-01-1223, 2000.
[218] Ponticel, P., “High Time for Hybrids,” Automotive Engineering International, vol.
110, no. 2, pp. 77–80, 2002, SAE International.
[219] “Saab Variable,” www.saabnet.com, 2000.
[220] Seinosuke, H., A. Hidaka, N. Tomisawa, M. Nakamura, T. Todo, S. Takemura, and
T. Nohara, “Application of a Variable Valve Event and Timing System to Auto-
motive Engines,” SAE paper 011224, 2001, SAE International.
[221] Sharke, P., “Power of 42,” Mechanical Engineering, vol. 124, no. 4, pp. 40–42, 2002,
ASME International.
476
References
[222] Shimizu, R., T. Tadokoro, T. Nakanishi, and J. Funamoto, “Mazda 4-Rotor Rotary
Engine for the LeMans 24-Hour Endurance Race,” SAE paper 920309, 1992.
[223] Silverstein, C. C., “Staying Cool,” Mechanical Engineering, vol. 121, no. 7, pp.
64–65, 1999, ASME International.
[224] Smith, J., “Nitrous vs. Blowers,” Hot Rod, pp. 60–62, Nov 1996.
[225] Stone, Richard, Introduction to Internal Combustion Engines, 2nd ed., 1992, SAE
International.
[226] “The Stretch for Better Passenger Car Fuel Economy: A Critical Look, Part 1,”
Automotive Engineering International, vol. 106, no. 2, pp. 305–313, 1998, SAE In-
ternational.
[227] Suzuki, T., The Romance of Engines, 1997, SAE International.
[228] “Toyota Prius: Best Engineered Car of 2001,” Automotive Engineering Interna-
tional, vol. 109, no. 3, pp. 27–28, 2001, SAE International.
[229] Tsuchiya, K, and S. Hirano, “Characteristics of 2-Stroke Motorcycle Exhaust HC
Emission and Effects of Air–Fuel Ratio and Ignition Timing,” SAE paper
750908, 1975.
[230] “Two-Stroke Engine Technology,” Automotive Engineering International, vol. 99,
no. 7, pp. 11–14, 1991, SAE International.
[231] Uchiyama, H., T. Chiku, and S. Sayo, “Emission Control of Two-Stroke Automo-
bile Engine,” SAE paper 770766, 1977.
[232] Urciuoli, V., B. R. Mason, and M. Marcacci, “Simulation Techniques Applied to
the Development of a 125cc 4-Stroke Scooter Engine,” SAE paper 951819, 1995.
[233] “Urea Selective Catalytic Reduction,” Automotive Engineering International,
vol. 108, no. 11, pp.125–128, 2000, SAE International.
[234] “Valeo and Ricardo Team for 42-V Diesel Engine,” Automotive Engineering In-
ternational, vol. 109, no. 11, pp. 56–59, 2001, SAE International.
[235] “Variable Valve Actuation,” Automotive Engineering International, vol. 89, no.
10, pp. 12–16, 1991, SAE International.
[236] Wilson, K., “Reducing Emissions: The Effects on Shipowners,” Diesel & Gas Tur-
bine Worldwide, pp.28–30, Sept 1996, Diesel and Gas Turbine Publications.
[237] Wong, C. L., and D. E. Steere, “The Effects of Diesel Fuel Properties and Engine
Operating Conditions on Ignition Delay,” SAE paper 821231, 1982.
[238] Wong, V. W., M. Stewart, G. Lundholm, and A. Hoglund, “Increased Power Den-
sity via Variable Compression/Displacement and Turbocharging Using the Alvar-
Cycle Engine,” SAE technical paper 981027, 1998, SAE International.
[239] Yamaguchi, J., “Opa - Toyota’s New-Age Vehicle,” Automotive Engineering Inter-
national, vol. 108, no. 9, pp. 23–34, 2000, SAE International.
[240] Yu, R. C., and S. M. Shahed, “Effects of Injection Timing and Exhaust Gas Recir-
culation on Emissions from a D.I. Diesel Engine,” SAE paper 811234, 1981.
[241] Zhao, F. Q., M. C. Lai, and D. L. Harrington, “A Review of Mixture Preparation
and Combustion Control Strategies for Spark-Ignition Direct-Injection Gasoline
Engines,” SAE paper 970627, 1997.
477
This page intentionally left blank
Index
Page references followed by "f" indicate illustrated Angle, 41-43, 47, 82, 132, 207, 209, 216-217, 224, Bottle, 371
figures or photographs; followed by "t" indicates a 237, 248-249, 255, 260, 271-272, 275-277, Bottom, 101, 232, 270-271, 403, 415-416
table. 286, 291, 293-295, 298-299, 303, 311, Boundary, 274-276, 293, 354, 382-383, 434
313-314, 320, 322-324, 329, 332, 342, 360, lubrication, 434
399-400, 407, 430, 440, 444-446, 448 Brake horsepower, 45
#, 57-59, 62-67, 70-71, 74-76, 80-81, 90, 99, 133, 205, Angles, 276-277 Brakes, 78, 381
224-225, 228-229, 233, 235-240, 247, 250, Angular motion, 262 Breakdown, 310, 400, 420, 456
321, 334, 337, 339-340, 377, 389-390, 394, Arc, 289 Breaking down, 452
396-398, 402, 404, 406-407, 410, 423, area, 44, 48, 50-51, 53, 57, 59, 68, 117, 119, 123-124, Bubbles, 417
432-436, 442-443, 451-452 202, 206-210, 228-231, 233, 248, 254, 269, Bulk, 222-223, 260-261, 280, 286, 404, 417
276-277, 308, 311-312, 317, 330, 333-334, Burn, 66, 207, 260, 265, 270, 274, 285, 288, 291, 293,
A 346-347, 352, 367, 391, 394, 396-397, 298-301, 306-307, 310-311, 313, 322-324,
Abrasive, 378 401-402, 405-407, 412, 425-426, 434, 441, 349, 351, 356-357, 362, 373-374, 378, 407,
wear, 378 445-446, 454 450-451, 466, 470
Absolute pressure, 130 units of, 51, 347 Bus, 316
Accelerated, 41, 216, 225, 230, 321, 337 Arguments, 93 Butane, 474
Acceleration, 219, 228, 348, 409, 434, 437, 444-445, ARM, 46, 319
448, 461 Ash, 450 C
Accelerometers, 380 Assembly, 220-222, 437 Calculations, 91, 275-276, 405
Accumulator, 466 efficiency, 222 Calibration, 220, 415
Accuracy, 91, 274 systems for, 221-222 capacitor, 288
Acids, 363 Assumptions, 85, 256-257, 275, 324 Carbon, 75, 83, 100, 247, 278, 310, 315-316, 336,
Actuator, 212, 220, 289-291, 307, 309, 315, 318-319, Atmosphere, 89, 92-93, 104-105, 107-108, 130, 141, 346-349, 353-355, 358-359, 361-362,
319 225, 232, 309, 327, 331, 333, 348, 355, 374-375, 378-379, 381-382, 385, 398, 408,
Addition, 61, 89-90, 101-102, 107, 124, 133, 142, 199, 362-363, 373-374, 376, 379, 381, 385 453, 458-459
201, 204, 210, 232, 246, 251, 259, 271, 282, Atmospheric pressure, 103, 122, 135, 225-226, 241, black, 315
328, 331, 333, 337-338, 342, 352-353, 356, 299, 330, 341, 414, 417 Carbon atom, 348
363-365, 370, 372-373, 376, 382, 394-395, Atomization, 215, 217, 309 Carbon dioxide (CO2), 361
399, 408-409, 415, 445, 449-450, 461 Atoms, 453 Carbon monoxide (CO), 75, 346, 354
Additives, 232, 295, 314, 351, 359, 363, 369, 374, Autocorrelation, 467 Carboxylic acids, 363
409, 413, 415, 453, 458-459, 475 Automatic, 227, 319, 374, 417 Carnot, Sadi, 88
lubricants, 458 Automobiles, 43, 47, 58-59, 67, 69, 77, 79, 125, carry, 100, 232
Adjustment, 203, 225-226, 291, 293, 294-295, 299, 128-129, 137, 210, 212, 221, 225, 232, 236, Cartridge, 459-460
304, 319 251, 288, 302, 317, 327, 337-338, 346-347, Cartridges, 460
Advanced, 211-212, 356, 375 353, 359, 363, 365, 367, 369-370, 373, Cast, 318, 335, 398, 400, 411, 426
Air, 40-41, 46, 51, 58, 61-62, 64, 66-67, 71, 74-79, 375-376, 379-381, 411, 416-418, 420, 422, Cast iron, 335
83-85, 88-94, 96-97, 99-104, 106-120, 443, 450-452, 473 Catalysts, 78, 365, 369-370, 372-374, 470
122-126, 128-131, 133-142, 197-257, Average, 40-44, 47-48, 54-55, 57, 68-69, 81-85, 91, cell, 79, 475
259-264, 267-270, 272, 275, 277, 280-282, 138, 141-142, 207, 210, 216, 247, 255, Cells, 473-474
285-296, 299-303, 305-319, 321-323, 260-262, 266, 287, 292-293, 295-296, Center, 46, 101, 248, 255, 265, 298, 303, 307, 313,
327-329, 331, 333, 336-339, 341-343, 298-299, 303, 311, 313, 322, 333-335, 342, 318, 322-323, 393, 399, 430, 432, 445, 453,
345-386, 388-389, 391, 394-397, 401-405, 347-348, 353, 362, 382, 384-385, 397, 466
407, 409-412, 416-422, 424-426, 436-437, 399-400, 402-404, 408, 424-425, 430, 434, Center of gravity, 430
443, 450-452, 458-462, 468, 470, 472-473, 437-438, 451, 453 Centrifugal, 237, 293, 443
477 Average value, 91, 138, 397, 400 Centrifugal forces, 293
bending, 46 Avoidance, 202 Ceramic, 78, 206, 212, 237, 337, 365, 371-372,
compressors, 103, 234-235, 237, 338 Axis, 42, 211-212, 259, 267, 296, 438, 443-444 374-375, 384, 398, 400, 411, 420, 458
contaminants, 365, 458-459, 461 Ceramics, 367, 420
gases, 89-90, 93, 204, 237, 252, 256, 260-262, B structure, 367
267-270, 272, 275, 281, 288, 290-291, Back, 41, 89, 93, 108, 123-124, 128, 201-204, 219, Chain, 58, 128, 212, 353, 380, 411-412
293-294, 296, 301-302, 327-328, 331, 236, 239, 269-272, 276, 300, 302, 317-318, saws, 58, 128, 411-412
333, 336-339, 346-349, 352, 355-356, 330-332, 336-339, 341-342, 350-351, changing, 46, 55, 61, 88, 100, 102, 117, 137, 198, 204,
359, 362-363, 365, 367, 372, 374, 371-372, 374, 376-377, 379, 395, 399-402, 212, 274, 277, 285, 304, 319-320, 381, 411,
376-377, 391, 394, 402, 407, 459, 461 412, 416-417, 419, 421, 445, 449, 452, 459, 438, 452, 472
particulates, 75, 336, 346, 358, 360-361, 374, 382 461 Channel, 241, 417
vapors, 385 Backflow, 209, 219, 276, 331 Channels, 274
Air pollutants, 347 Bar, 312 Charge density, 237
Air pollution, 67, 317, 345-386, 470 Barrel, 230-233, 254, 385 Chatter, 206
Aircraft, 201, 232, 250, 307, 373, 409, 411-412, 450 base, 365, 415, 421, 453, 456, 458 Chemical, 64, 80, 85, 251, 274-275, 281, 290, 300,
Alcohol, 134, 201, 254, 314, 362, 382, 421-422 Basic, 71, 88, 133, 137, 210, 212-213, 221, 225-226, 336, 355-357, 362, 365, 372, 375-376,
Alcohols, 408, 418 449 382-383, 388-389, 466, 468, 471
Aldehydes, 357, 362 size, 212 Chemically, 365, 415
Alloy, 206 Bearings, 46, 401, 431-432, 434, 438, 454 Choke, 225-227
Alloy steel, 206 Bending, 46 Chromium, 359
Alloying, 420 force, 46 Circular, 319, 382-383
mechanical, 420 Biological, 356 Circulation, 416-417, 419, 449-450, 460
Alumina, 365, 384 Black, 315 distribution, 449-450, 460
Aluminum, 59, 272, 365, 398, 400, 411, 437 Blow, 102, 337 Cleaning, 458
porous, 365 Blowby, 100, 259, 270-272, 282, 316, 350, 352, 379, Clearance, 44, 46-48, 82-83, 91, 97, 107-108, 132,
Aluminum oxide, 365 452-453, 458-460, 473 135, 204, 207, 260, 264-265, 267-270,
American Petroleum Institute (API), 456 Blowdown, 89, 93, 95, 101, 105-107, 112, 114, 130, 272-273, 280-282, 293, 300, 304, 307-308,
Ammonia, 376, 383 132, 135-136, 210, 241, 262, 272, 327-333, 318, 320, 323, 330, 332, 351-352, 379, 383,
and, 40-44, 46-59, 61-80, 82-86, 88-89, 91-103, 105, 335-336, 338, 340-342, 351, 359, 364, 381, 401, 440
107-142, 197-257, 259-262, 264-278, 407, 435 fit, 207, 270, 308
280-282, 285-319, 321-324, 327-343, Boiling point, 414 Closed system, 89-90, 100, 107
345-386, 388-392, 394-427, 429-463, Bolts, 308, 319 Closing, 74, 78, 89, 102, 123-126, 141, 203, 206,
466-477 Boltzmann constant, 398 208-209, 275-276, 305, 330, 417
479
Closure, 101, 203 Convection heat transfer, 261, 394, 396-398, 401-408, 76, 79, 82-86, 97, 99, 104, 125, 132, 138,
Clusters, 359, 382 411, 426 141, 219, 234, 240, 243-245, 254, 257, 264,
Coal, 467 Convective heat transfer, 396, 408 303-304, 308-309, 312, 316, 318-319, 324,
Coated, 359 Conversion, 71, 78, 80, 309, 316, 368-369 339, 365, 382-384, 405, 433, 442, 445, 451,
Coefficient, 58, 206-207, 228-229, 233, 247, 250, Coolants, 412, 415, 418 461-462, 470, 476-477
254-255, 276, 394, 396-398, 402, 404, Cooling, 78, 93, 102-103, 105, 108, 110, 127, 201, Distances, 370, 445
406-408, 411, 420, 426 219, 232, 236, 248, 251-252, 295, 309, 328, Distillation, 453
drag, 58, 420 335, 338, 340, 343, 359, 365, 370, 391, Distribution, 201-202, 222, 246, 249, 265-266, 285,
Cold, 65, 78, 85, 222, 227, 229, 254, 287-288, 394-396, 398-399, 401, 408-417, 419-421, 295, 310, 314, 316, 389-391, 420-421,
305-306, 312, 315-317, 338, 350, 368, 424, 427, 443, 458 449-450, 458-462, 473
370-372, 376, 384-385, 391-393, 417-418, Coordinates, 50-51, 53-54, 92-94, 111, 116-117, Double, 133-134, 141, 371
421-422, 450, 453-454, 456, 458, 468 134-135, 235, 238, 329 Draw, 142, 257, 385, 427
shut, 305 Core, 275, 309-310, 313 Drawing, 275, 367, 427
College, 69 Corners, 199, 202, 206-207, 259, 270, 293, 321 driver, 337, 422
Combustion, 39-40, 44, 46-51, 54, 64-68, 70, 73-74, Corrections, 74 Drop, 93, 102, 202-203, 219, 227, 231-233, 237, 309,
76, 79, 82, 84-85, 87-91, 93, 95-97, 100-103, Correlation, 41, 314, 466 314, 328, 330, 337, 343, 359, 374-375, 404,
105, 107-108, 110-112, 114-116, 119, cost, 103, 134, 207, 225, 227, 229, 234, 239-241, 243, 407, 460
126-128, 130-131, 133-135, 137-142, 197, 273, 300, 308, 330, 353, 365, 372, 375, 413, Dry, 372, 431, 450
200-201, 204-208, 211-213, 215, 218, 415, 421-422, 424, 451, 453, 458 dual, 89, 114-119, 130, 137, 139, 204, 210-211, 221,
221-223, 227, 230, 234, 236, 239-242, justification, 458 239-241, 252, 255, 257, 301-302, 307, 312,
246-249, 252-253, 255-256, 259-283, Cotton, 459 318, 336, 372, 382, 405, 420, 422, 435, 450
285-325, 327, 330, 334-335, 338-340, 342, dust, 459 Ductwork, 412
345-364, 372-378, 382-385, 387-389, 391, Cracks, 414 Dummy, 212
394-402, 405-409, 411, 419, 421, 424, 426, Cross, 48, 115, 207, 229, 242-243, 445, 447 Dust, 458-459
429, 433, 435-438, 444, 450-453, 458, Crown, 264-265, 393, 399, 419, 453 Dynamometers, 59, 61, 435
461-462, 465-477 Crystal, 421 electric, 61, 435
heat of, 138, 201, 252-253, 384, 409 Curves, 52, 57, 61, 65, 73, 211, 234, 438-439 Prony brake, 59
of fossil fuels, 67 Cutoff, 111-113, 116-117, 122, 138-139, 282, 312, 341
Communication, 474 Cycles, 50, 87-143, 210, 227, 231, 239, 260, 273-274, E
component, 77, 261, 273-274, 295, 348, 361, 364, 277, 296-299, 305, 317, 319, 337, 341, 348, Ear, 380
391, 415, 453, 460 378, 399-400, 406, 421, 450 sensitivity, 380
Composite materials, 380 Cylinders, 41, 43, 46, 58-59, 61-62, 70, 79, 83, 85, 89, Earth, 362
Composites, 420 100-102, 113, 119, 138-141, 198-199, Economy, 64, 67, 69, 83, 212, 229-230, 242, 260, 269,
Compound, 223, 305, 351, 366 201-202, 205-206, 214-215, 225, 227, 299-301, 304, 308, 339, 358, 378-379, 382,
Compressed, 88, 103, 203, 235-237, 241, 280-281, 231-232, 238, 241, 251, 254-256, 259, 392, 409, 470, 477
293, 307, 339, 350, 384, 395, 451, 471 273-275, 282, 295, 298-300, 304-305, Efficiency, 44, 54-56, 58, 61, 63, 67, 70-71, 73-76,
Compression, 44, 46-48, 51, 53-55, 67, 72, 77, 79, 308-309, 312, 316-319, 335, 341-342, 352, 78-85, 90, 93, 95-99, 101-102, 108, 112-115,
82-83, 89-94, 96-98, 100-104, 107-115, 377, 382-383, 394-396, 398, 401, 410, 412, 117, 120-122, 124, 126-128, 131, 133-134,
117-120, 122-126, 130-133, 135, 137-142, 420-421, 424, 426, 430, 449, 451-453, 137-142, 198-208, 212, 219, 222-224, 230,
201-202, 204-205, 212-213, 221, 223-224, 461-462 232, 235-238, 240, 243-246, 249-257,
227, 234-236, 241-242, 246, 252-253, 259-261, 265, 267-268, 270, 277, 282, 291,
255-256, 259-260, 262, 264, 267, 270-273, D 295, 301, 304, 306, 308-309, 311, 316, 319,
275-276, 280-282, 288, 290-291, 298-301, data, 54, 211, 261, 274-277, 373, 397, 438-439, 442, 323-324, 339, 343, 353, 362, 367-369, 373,
303, 305-306, 308-309, 312, 314-320, 446, 467-468 375-376, 378, 382-384, 388-389, 394-395,
322-324, 330-331, 335, 341-343, 350-352, Decibels, 380 401, 403, 405-406, 408, 411, 416-417,
356, 359-360, 363, 374, 378, 381-383, 394, Deflections, 437 419-420, 422, 424-426, 432, 442, 451, 453,
396, 399, 401, 403, 405, 408, 411, 420, 433, Deformation, 475 462
435-439, 441, 445-446, 451-452, 458, 462, Defrosting, 417 Carnot, 134
469, 473, 475-477 Degree, 217, 250, 277, 291, 355 Electric, 40, 61, 78-79, 213, 222, 225, 289, 317,
axial, 212, 262 delay, 110, 247-248, 260, 271-272, 294, 303, 306, 371-372, 374, 416-417, 421-422, 435-437,
test, 420, 437 309-314, 316, 322, 371-372, 450, 469, 475, 443, 450, 469
Compressors, 103, 234-235, 237, 338 477 Electrochemical, 79
positive displacement compressors, 234 minimum, 272, 306, 309, 372 Electrode, 288-289, 301
Computations, 471 Delay time, 294, 306, 310-311, 314, 316, 322 Electrodes, 287-289, 295, 307, 321
Computer, 77, 207, 273-275, 300, 318, 466, 468, 471 Delays, 450 Electromechanical, 78, 212, 221, 476
simulation, 273-275, 468 density, 71, 74-75, 83, 90, 99, 108, 201, 204, 228, 233, Elements, 107, 328, 335, 350, 371, 471
Computer-aided, 207 237, 243, 248, 250, 252, 254-255, 270, 290, Elevation, 228
computer-aided design, 207 295, 330, 346-347, 358, 363, 382, 384, 394, Elevations, 362
Computers, 78, 212, 273-274, 381 404, 407-408, 410, 462, 477 Emergency, 315
Concentration, 309, 351, 356, 361-362, 385, 413-415, of water, 201, 410 Emissivity, 398
417 Depletion, 67, 422 Emulsion, 409
Conditioning, 77 Depth, 260, 383, 396, 398 energy, 55, 59, 61, 64, 71, 78, 80, 84-85, 89-90, 93,
Conduction, 288-290, 371, 395-396, 398, 400-402 Design, 44, 55-56, 58, 78, 80, 85-86, 117, 142, 202, 101-102, 107, 127, 235-237, 268-270, 281,
Conductivity, 371, 396-397, 400, 402, 404, 411, 414, 207-208, 232, 243, 252-257, 266, 273, 280, 287-289, 291, 293, 313, 317, 328, 335,
426 282, 302, 306-308, 318, 322-324, 329-330, 337-338, 340, 342, 355-356, 362, 365,
electrical, 371 336-337, 342-343, 351, 354, 357, 360, 375, 369-372, 377, 380-385, 388-390, 396, 400,
Connectors, 77 380, 385, 394, 411, 427, 462, 471-474, 476 402, 405, 407-408, 410, 415-417, 419,
Conservation of, 264, 335 Desorption, 351-352 421-424, 426-427, 467-468, 470, 472
angular momentum, 264 Deviation, 101-102, 295 BTUs, 426
energy, 335 standard, 101-102, 295 coefficient, 396, 402, 407-408, 426
mass, 264 Diagrams, 50 head, 270, 405
Constants, 91, 397, 433-434 Die, 299, 307 ideal, 71, 89, 101-102, 235-236, 291, 328, 337, 355
Construction, 83, 273, 365, 374, 380 Dies, 270 kinetic, 78, 80, 85, 107, 235, 281, 328, 335, 337,
Consultation, 456 Diesel, 40, 45, 64, 76, 82-83, 85, 110-115, 117-120, 340, 342, 381
Container, 365, 421, 425 122, 134, 137-139, 210, 239-240, 249-250, latent, 408, 421, 423, 470
Contaminants, 365, 453, 458-459, 461 255-257, 282, 312, 316-318, 328, 341-342, limited, 337, 355, 365, 372, 381, 389, 417, 422
Continuous, 134, 219, 254, 289, 443, 445, 450, 452 347-348, 353, 359-360, 362-363, 374-375, pipe flow, 424
Control, 74, 78-79, 125, 137, 142, 199, 203, 208, 381-385, 417, 426, 446, 458-459, 466-467, potential, 235, 287-289
210-213, 219, 223, 225, 227, 230-231, 469, 472-477 power and, 61, 71, 405, 408
247-248, 252, 272, 274, 276, 289, 295, 300, Differential equations, 471 specific, 55, 59, 64, 80, 84, 89-90, 93, 107,
308, 329-330, 336, 353, 360, 367, 372, 376, Diffusion, 318 235-236, 328, 335, 337, 342, 377, 380,
413, 443, 466, 470, 475-477 volume, 318 382, 384-385, 415-416, 423
segment, 330 Dimension, 206, 405 work, 55, 90, 93, 101-102, 127, 289, 291, 317, 377,
Control systems, 225, 372 Direct, 128, 198, 215, 218, 221, 224, 264-265, 274, 381, 385, 388, 419
Controlling, 202, 204, 212-213, 274, 316, 360, 380, 303, 311, 322, 339, 360, 409, 412, 415, 422, Energy loss, 408
420 443, 461, 466, 468, 475, 477 Engineering, 39, 87, 197, 259, 285, 327, 345, 387,
process of, 213 emulsion, 409 415, 429, 454, 465-477
Controls, 77-79, 211-212, 215, 222, 226, 230, 291, Directionality, 260 computer, 466, 468, 471
299, 305, 336, 375, 417, 421, 425, 450-451 Displacement, 41-44, 46, 51, 55-57, 59, 62-64, 68-71, environment, 415
480
value, 454 437-438, 446, 453, 466 476
Engineering design, 471 Filtering, 460 Generator, 61, 78, 80, 84-85, 343, 435, 468
Engineers, 446, 466 Filters, 78, 248, 316, 458-459 Geometric, 91, 206, 242
Engines, 40-42, 44, 46-47, 50, 52-59, 61, 66-68, 71, Filtration, 449, 458-460 Geometry, 41, 102, 242, 265, 276, 280, 287, 291, 295,
77-79, 86, 88-89, 91-92, 96, 100-101, 103, Fine, 221, 285, 291, 338, 459 302, 306, 314, 348, 375, 389, 434, 467
107, 110, 112, 115, 117-119, 122, 124-125, Fins, 396, 412, 420-421 Glass, 78
128-130, 133-134, 137, 198-204, 206-208, Firing, 79, 93, 256, 277, 287, 294-295, 421 Goals, 300
210-211, 213, 215, 219, 221-223, 225, 227, Fit, 59, 207, 270, 308 Graphite, 458
230-232, 234-237, 239, 241-243, 246, Fitting, 202, 330, 365 lubricant, 458
248-252, 260-261, 264-265, 267-269, Flame, 101, 202, 222, 260-261, 264-265, 267-270, Gravity, 199, 225, 228, 232, 414-415, 430-431
272-275, 285-286, 288-289, 291, 293, 295, 274-275, 277, 279, 282, 285-295, 297, 299, Greater than, 53-54, 65, 71, 100, 105, 112, 123, 128,
299-302, 305-309, 311-312, 314-319, 321, 301, 303-304, 306-310, 315, 318, 321-324, 137, 201, 208, 214, 235, 241, 243, 271, 287,
323, 329-331, 334-339, 341, 346-347, 349-350, 353-354, 356, 371, 374, 377, 382, 327, 330, 334, 390, 398, 402
351-360, 362-366, 368-370, 372-381, 385, 411, 446, 460-461, 468, 470, 472 Green, 471
387-427, 432, 437-439, 443, 445-446, Flat, 48, 308 Greenhouse gases, 362
449-454, 458-461, 466-468, 470-472, Flexibility, 78, 125, 134, 289, 308, 330, 420, 422 Group, 438
474-477 Flexible, 59, 225, 230, 324, 420
diesel, 40, 110, 112, 115, 117-119, 122, 134, 137, Flights, 373 H
210, 239, 249-250, 312, 316-318, 341, Float, 225-226 Hand, 77, 124, 198, 201, 203, 336, 388, 405, 419,
347, 353, 359-360, 362-363, 374-375, Flow rate, 41, 64-66, 71, 75-76, 83-85, 90, 100, 199, 424, 456, 458
381, 385, 417, 426, 446, 458-459, 201-205, 215, 217, 219, 224-227, 229-231, trucks, 336
466-467, 472, 474-477 233, 235, 237, 239-240, 247, 249-250, 252, Handling, 327
heat, 44, 55, 59, 67-68, 78, 88-89, 91, 96, 100-101, 254-256, 260, 281, 305, 321, 331, 333-334, Hard, 61, 206, 270
103, 110, 112, 115, 117, 119, 130, 337, 339-340, 342-343, 365, 370, 375, 377, Hardening, 363
133-134, 137, 201, 222, 227, 235-236, 389, 397, 402-403, 407, 412, 417, 423-426, Head, 48, 69, 198, 206-207, 228, 241-243, 246, 255,
251-252, 261, 269, 274-275, 288-289, 450, 459-460 262, 264-265, 267, 270, 276-277, 306, 308,
293, 299, 305, 308, 311-312, 315-317, volumetric, 71, 75-76, 83-85, 199, 201-205, 219, 318-319, 405, 420, 450
321, 335-338, 354-355, 365, 370, 372, 224, 230, 240, 249-250, 252, 254-256, loss, 308, 405, 420
385, 387-427, 446, 450, 466-468, 470, 260, 339, 343, 365, 403, 424, 426 pump, 228, 420, 450
475 Fluid, 55, 59, 89, 100, 103, 133, 202, 212-213, 236, Health, 346, 380, 385
internal combustion, 40, 54, 88, 91, 100, 133, 137, 259-283, 285, 354, 380, 388, 395, 400, 403, heat, 44, 49, 55, 59, 63, 67-68, 70, 78, 85, 88-91, 93,
200, 260, 275, 285, 317, 347, 385, 387, 409, 413, 415, 417-418, 423, 425, 431-432, 95-96, 98, 100-103, 110-112, 114-117, 119,
467-468, 470-472, 474, 477 458, 461, 466, 468, 471 130-131, 133-140, 142, 201, 222, 227,
jet, 134, 250, 300, 309, 311-312, 314, 321 Fluid flow, 100, 274, 417, 471 235-236, 251-253, 259, 261, 269-270,
Otto cycle, 92, 96, 101, 103, 110, 112, 115, smoothness of, 274 274-276, 280, 282, 287-290, 293, 298-299,
117-119, 124, 128, 137, 208, 210, 291, Fluid mechanics, 260 304-305, 308, 311-312, 315-317, 321, 328,
316, 319, 329, 341 Fluids, 306, 380, 427, 430, 436 335-338, 354-355, 365, 370-372, 384-385,
two-stroke, 42, 46, 56-59, 88, 128-130, 134, 137, viscosity of, 380, 436 387-427, 446, 450, 466-470, 475
198, 235, 239, 241-243, 252, 261, 265, Flux, 61, 398-399, 426, 466 Heat exchangers, 103, 395
319, 338-339, 341, 352-353, 373, Fluxes, 398-399, 405 Heat flow, 390, 398-399
450-451, 467-468, 477 Fog, 347 Heat transfer, 44, 89-90, 96, 100-101, 117, 201, 235,
Enthalpy, 71, 89-90, 93, 107-108, 235, 251, 328, Foil, 367, 371 259, 261, 274-276, 280, 288, 290, 304-305,
334-335, 337, 341-342, 378, 383, 388-390, Force, 46, 49, 56, 101, 203, 241, 269, 286, 291, 308, 321, 328, 387-427, 466-470, 475
414, 416, 419 314, 376, 431, 434, 437, 439-441, 444-446, conduction, 288, 290, 395-396, 398, 400-402
Environmental, 347 448, 454, 461-462, 472 convection, 261, 290, 394-398, 401-408, 411-412,
Environmental protection, 347 body, 462 416-417, 419, 426
Environmental Protection Agency (EPA), 347 Forging, 247 fin, 391
Equalization, 203 Formaldehyde, 347, 372 hot spots, 394-395, 412, 417
Equations, 56, 62, 80, 113-114, 120-121, 207, 239, Forming, 369 modes of, 395, 470
243, 262, 274, 276, 404, 412, 424, 434, 471 FORTRAN, 471 unsteady, 469
Equilibrium, 274, 355-357, 468, 471 Fossil fuels, 67 Height, 228, 232-233, 254, 260, 324, 415, 446, 462
Equilibrium constant, 355-357 Fouling, 288 Helium, 377
Equipment, 78, 274, 288, 338, 353, 374, 381, 433 Frames, 430 Help, 247, 273, 317, 331, 346, 359, 380, 401, 419,
Error, 100-101, 273 Freeze, 232, 412, 414, 451 454
Errors, 89, 102 Frequency, 203, 296, 299, 336, 375, 380-381 High-speed, 44, 46, 115, 202-203, 208, 211, 229, 248,
Estimation, 467 Friction, 44, 51, 53, 55, 57-59, 61-63, 66-68, 82-83, 261, 285, 306, 324, 338, 388, 409, 454, 467
Ethanol, 69, 85, 324, 383, 424 89, 99, 102, 138, 202, 212, 236, 279, 311, History, 128, 468, 473
Evaluation, 469 389-390, 418, 420, 425, 429-463, 467, 470, Hood, 232, 430
Evaporative, 201, 232, 251-252, 295, 309, 394, 396, 473-475 Horsepower, 45, 52, 58, 61, 69, 83, 85
408-410, 424, 474 Fuel cells, 473-474 Hot, 92, 102, 108-109, 131, 199, 204, 219, 237, 248,
event, 210, 475-476 Fuel NOx, 350 250, 289, 299, 304, 308-309, 314, 317, 331,
Events, 309, 427 Full, 58, 79, 107, 125, 370, 389, 392-393, 460 338, 369-370, 372, 375-376, 388, 391,
Exhaust, 40-41, 51, 53-55, 59, 70, 75, 77, 79, 82-83, Functions, 58, 69, 91, 142, 365 394-395, 399, 401-402, 412, 415-418,
85, 88-91, 93-95, 97-98, 100-102, 105-110, 420-421, 424, 430, 437, 453, 456, 460-461,
112-114, 119-120, 122-123, 126-128, G 477
130-132, 135-142, 199, 202, 204, 207-208, Gage, 215 spots, 309, 394-395, 412, 417, 420, 430, 453
210-214, 219, 227, 230, 236-238, 241-243, Gain, 201, 251, 253, 300, 419 Hot spots, 394-395, 412, 417, 453
245-247, 251, 253-256, 259, 261, 265, Gains, 437 HP, 44, 51-52, 58, 61-63, 65, 67, 69, 73, 75, 80-82, 84,
270-278, 281, 291, 299-302, 304-308, Gas constant, 74, 90, 209, 228, 328 99, 133-134, 142, 230, 239, 279, 324, 382,
314-316, 327-343, 346-352, 354-356, for air, 74, 328 442-443
358-365, 367, 369-385, 388-391, 393-395, Gases, 49, 89-90, 93, 105, 204, 237, 252, 256, Human, 376, 380
398-399, 402-408, 411-412, 418-420, 422, 260-262, 267-272, 275-276, 281, 288, Humidity, 74, 409
425-426, 433-435, 437-439, 444-446, 290-291, 293-294, 296, 301-302, 304, Hybrid, 69, 79-80, 85, 476
451-452, 458-459, 461-462, 467, 470, 327-328, 330-331, 333, 335-340, 346-349, Hydraulic, 46, 59, 78, 84, 211-212, 228, 318, 380,
474-475, 477 352, 355-356, 359, 362-363, 365, 367, 372, 431-432, 434, 437, 443
Exothermic reaction, 372 374, 376-377, 391, 394, 398-399, 402, 407, Hydraulic head, 228
Experiments, 261 435, 459, 461 Hydrocarbons, 75-76, 346-348, 364
Exposed, 391, 399, 401, 412, 419 Gasoline, 46-47, 65-66, 69, 80, 83, 85, 134, 137, 142, Hydrogen, 201, 349, 359, 362, 382, 469-470, 473, 475
199, 201, 205, 215, 218, 221, 224, 233,
F 253-254, 256, 281-282, 292, 295, 303, 306, I
Factors, 47, 77, 102, 202, 274, 305-306, 314, 365, 475 318-319, 323-324, 339, 343, 347-348, 351, Id, 312, 314, 322-323, 403
combined, 305 353, 362-363, 369, 371-372, 379, 382-385, Ignition, 44, 46-47, 51, 53, 65-66, 70, 72, 74, 79, 83,
Failure, 41, 122, 388, 419, 437 401, 409, 424, 462, 470, 472, 474, 477 101-103, 110, 114, 117, 119, 127, 130, 139,
mode, 388, 437 Gears, 430 141-142, 213, 220, 222-224, 236, 239-240,
Failures, 363, 412, 424 General, 44, 51-52, 80, 241, 274-275, 278-279, 306, 246-248, 253, 256-257, 260, 263, 269,
Feed, 198, 242, 380 319, 338, 353, 388, 405, 420, 468, 470 274-275, 277, 286-289, 291, 293-295,
Field, 61, 315 General Motors Research Laboratories, 470 297-298, 300-304, 306, 308-319, 321-324,
Figures, 277 Generation, 128, 222, 243, 274, 277, 288, 338, 341, 336, 356, 358, 360, 371-372, 374, 407-408,
Film, 199, 228, 271, 351-352, 354, 392, 400, 431-432, 349, 357-358, 360, 362, 375, 377, 383, 408,
481
420, 422, 435, 460, 466-467, 469-472, 475, 308, 311-312, 314, 317-318, 328-329, 461, 468, 472-473
477 335-336, 339, 357, 363-365, 370, 373-374, equations of, 274
Impact, 346 377-378, 383, 389, 398, 402, 408, 411, 413, relative, 319, 430-431, 437
Impurities, 346, 364, 369, 374, 401, 452, 454, 458-460 417, 420, 430, 434, 445, 451, 454 Motor vehicles, 361
Incomplete, 295, 348-349, 355, 362, 372 Lubricants, 420, 424, 458, 471 Motorcycle, 59, 77, 244-245, 477
Index, 75-76, 82-83, 382-383, 454 Lubrication, 78, 211, 227, 242, 271, 316, 337, 351, Mounting, 206, 422
Induction, 100-101, 197-257, 467, 469 353, 400, 420, 424, 429-463, 471-472 Mounts, 380, 468
heating, 101, 199, 232, 236, 246, 248 multiplier, 276
Industrial, 83, 203, 208, 227, 330 M
Information, 214, 274, 299, 336, 380, 471 Magnesium, 369 N
Inhibitor, 413 Magnetic, 61, 213 Natural, 134, 239-240, 302, 318, 346, 363, 416, 419,
Input, 40, 64-66, 79, 84, 93, 95, 101-102, 110-111, Maintenance, 274, 300, 473 470-471, 474, 476
115-116, 118-119, 127, 130-131, 133-135, Management, 64, 72, 223, 336, 360, 380 rubber, 318
200, 210, 219, 223-225, 230, 236, 239, Manual, 417, 468, 471 Neutral, 337, 365, 377
241-243, 250-252, 256, 275-277, 291, 299, Manufacturing, 207, 295, 308, 445 Nitrogen, 75, 204, 278, 338, 341, 346-347, 349,
305, 317-318, 336, 353, 377, 381, 385, 420, variations, 295 355-357, 361, 378, 383, 471
436, 450-452, 462 MASON, 477 Noble metals, 366, 373
Inputs, 224, 274, 433 Mass, 50, 62-66, 70-71, 74, 76-77, 80, 83-86, 90, 93, Noise, 77-78, 206, 234, 334, 338, 380-381, 412
Integrated, 337 97, 100, 102, 107-108, 114, 120, 122, 126, Normal, 66, 142, 212, 227, 231, 239, 241, 250, 259,
Integration, 275, 277 132, 199, 207-208, 217, 222, 224-225, 228, 274, 285, 288, 295, 299, 301-302, 305, 309,
Interest, 40, 88, 134, 277 235, 237, 239-240, 243, 245-247, 249-251, 314, 316-318, 321, 368, 370-374, 380-381,
Intermediate, 438 256, 259, 261, 263-264, 267-268, 270, 273, 391-392, 398, 401-402, 417, 421, 426,
shape, 438 276-277, 281, 285-286, 288-290, 293, 295, 436-437, 458-459, 461
Internal, 39-40, 54, 87-88, 90-91, 100, 133, 135, 137, 308, 319, 321, 333-334, 337-340, 342-343, Normal distribution, 314
197, 200, 259-260, 275, 285, 317, 327, 345, 359, 369, 374, 377, 379, 382, 384, 389, 394, Normal mode, 305
347, 385, 387-388, 401-402, 419, 429, 433, 397, 402-404, 406-408, 410-411, 417, Not equal to, 461
465, 467-472, 474, 477 422-424, 434, 437, 440, 444-446, 452, 462 Numbers, 47, 112, 314, 347, 353, 454, 456
Internal energy, 90 Mass moment of inertia, 263, 268
International standards, 380 Material, 41, 288, 308, 337, 351, 365, 367, 369,
Intervals, 469 O
374-375, 391, 419-421, 424, 458-459 Offset, 40-43, 47, 49, 131, 140-141, 265, 276, 294,
Iridium, 366, 373 cost, 308, 365, 375, 421, 424, 458
Iron, 272, 335, 359, 374, 398, 400, 411, 426, 430, 437 303, 312, 322, 407, 432, 445, 448, 461-462
movement, 419 Oils, 400, 420, 424, 451-458, 469
Materials, 44, 237, 288, 350, 356, 363-364, 366, 372, One, 43, 46, 48, 55-56, 58-59, 61-62, 67-68, 70-71,
K 374-375, 380, 388, 400, 405, 411, 413, 415, 74-75, 77-84, 89-90, 92-93, 97-99, 104-105,
kernel, 296, 298 420-421, 424, 427, 437, 459 107-108, 112, 114, 120-122, 126, 130,
Kerosene, 418 Mathematical models, 273-274 132-134, 141, 199, 202-204, 206-209, 211,
Kinematic viscosity, 439 Maximum density, 358 214, 219, 223-225, 230-234, 241, 245, 247,
Kinetic energy, 78, 80, 85, 107, 235, 281, 328, 335, Maximum thickness, 446 249, 251-256, 259-263, 265-272, 274-275,
337, 340, 342, 381 Mean, 55-56, 62-63, 81-84, 98-99, 105, 125-128, 281-282, 285, 287-289, 295, 301-304,
rotational, 281, 337 132-133, 138-139, 141, 244-245, 260, 296, 306-307, 309, 313, 318-319, 322, 327,
Kinetics, 274, 470 298, 405, 433-434, 436, 438, 440-441, 443, 329-340, 342, 346-348, 352-353, 355-356,
462 366, 370, 372-373, 377, 379-380, 385, 388,
L Measurement, 219, 470 391, 397, 399, 401-402, 405, 409-410, 412,
Lag, 237, 337 Measurements, 40, 438, 466-467, 473, 475 416, 418-422, 425-426, 430-431, 434-435,
Land, 376, 393 Mechanical, 40, 50-51, 54-56, 58-59, 61, 63, 67, 71, 437-440, 445-446, 450, 453, 458, 460-462,
Large errors, 89 76-78, 81-83, 85, 97, 123-124, 131, 134, 473, 476
Laser, 475 138-139, 141-142, 204-207, 211-214, 236, Open, 53, 74, 79, 88-89, 93-95, 100-101, 105, 107,
Latent heat of, 423 238, 240, 251, 274, 285, 293, 299, 308, 323, 110, 112, 130, 136, 142, 198, 202-203, 206,
Laws, 47, 67, 128, 225, 241, 316, 347, 353, 362, 376, 327, 336-337, 382, 419-420, 430, 432, 438, 208, 210, 212, 219, 223, 227, 241, 248,
380, 422 442-443, 446, 452, 460-462, 466, 470-472, 254-255, 276, 302, 311, 317-318, 322-323,
Per, 67, 128, 316, 347 476-477 327, 329-331, 333, 339-341, 351-352, 405,
layers, 431 advantage, 236, 420 449, 461
layout, 218 alloying, 420 Optical, 274, 299
Lead, 314, 347, 351, 363-364, 369, 382, 385, 415 Melting point, 403, 421 Optimization, 208, 210-211
Lead emissions, 363 Metal, 59, 227, 270, 293, 316, 351, 363, 366-367, 369, Optimum, 213, 248, 291, 299-300, 329, 372, 392, 394
Leading, 295, 403 371, 374, 400, 430, 434, 437, 453, 458 Order, 41, 44, 54, 92, 100, 122, 130, 206, 216, 241,
Leakage, 271, 338 Metallic, 454 256, 259-260, 262, 270, 287, 290, 293, 295,
Less than, 42, 51, 65, 67, 71, 92, 101-102, 125, 200, Metals, 363, 366, 373, 412, 420 314, 349, 367, 380, 388, 390, 394, 396, 398,
202-203, 206-207, 215, 225-226, 232, Meter, 214, 438, 467, 473 405, 407-408, 417, 420, 424, 426, 432, 434
235-236, 238, 285, 306, 333, 335, 347, 353, Methane, 201, 240 Ordinate, 438-439
370, 397, 431, 460-461 Methanol, 141, 255-256, 318, 324, 372, 470 Out of, 41, 52, 79, 89, 105, 135, 139, 203-204, 208,
Light, 65, 69, 76, 79, 107, 119-120, 138-139, 212, 223, Method of, 69, 210, 229, 242, 256, 319, 337, 377, 381, 219, 225, 241, 243, 259, 269-270, 273-274,
239-240, 249, 255-256, 282, 298, 318, 337, 385, 432, 475 280, 320, 330-332, 335-336, 347-348,
347-348, 359, 368, 370-372, 381-382, 384, Methods, 59, 75, 125, 204, 211, 230, 241, 262, 352-353, 363-364, 381, 405, 459
408, 437, 474 274-275, 288-289, 299, 338, 370, 375, 388, Output, 40, 44, 49, 51, 55-56, 58-59, 61-63, 65, 71,
speed of, 79, 255-256, 382 394, 397, 409, 418, 420-421 73, 79, 82-85, 93, 99-102, 125, 127-128,
Lighting, 77 direction, 262, 397, 418 133-134, 138, 205, 212, 234, 236-237, 240,
Limits, 66, 299, 310, 316, 391 indirect, 274, 375 251, 255, 277-279, 298-299, 302, 304, 307,
Linear, 260, 405 Minutes, 84, 305, 370, 373, 392, 422 309, 316, 318, 338, 343, 371, 389-390, 405,
Lines, 51, 61, 117, 317 model, 41-44, 46, 58, 92, 105, 107, 110, 141, 225, 408, 410, 419-420, 430, 434-435, 443, 456,
Liquid, 71-72, 77, 83, 100, 129, 199, 201, 219, 228, 227, 234, 253, 260, 262, 266, 268, 274, 276, 461-462
236, 248-251, 295, 309-310, 312-315, 322, 281, 289, 328, 411, 418, 425, 451, 466-469, Outputs, 100, 433
354, 395-396, 402, 412-413, 416-418, 472, 474 Overhead, 206-208, 308, 405, 450
420-421, 423-426, 431, 459 components, 41, 411, 451 Overlap, 101, 107, 117, 202, 204, 208-210, 213, 227,
Liquids, 418 elements, 107, 328 230, 242-243, 255, 305, 330-332, 339-340,
List, 276, 323, 382 mathematical, 274 351, 384, 451-452
Loads, 51, 54, 107, 210, 223, 298, 309, 318, 359, 389, Modeling, 262, 275, 467-471 Oxygen, 64, 70, 219, 250, 252, 304, 315, 336,
408, 437, 453 Models, 58, 77, 129, 221, 260, 273-275, 285, 338, 402 348-349, 353-356, 359-360, 362, 364-365,
Logic, 223 Module, 220 374, 376, 384, 453, 460
Loop, 51, 53-54, 102, 123-124, 242-243, 309, 318, Mole, 251, 355, 377, 382, 410 cylinders, 70, 304, 453
381, 416-417, 435-436, 460 Molecular weight, 352, 409, 453, 458
closed, 102, 124, 242-243, 318, 381, 416-417 Moles, 378-379, 409-410
Moment of inertia, 263, 268 P
Loops, 417 Parameters, 40, 42, 55, 59, 64, 66, 72, 80, 109, 127,
lot, 77 Moments, 337
Moments of inertia, 337 219, 223, 229, 254, 260, 267-268, 276, 285,
Lower, 44, 47, 51, 53, 58, 68, 74, 77, 100-104, 124, 318, 336, 348, 436, 475
127-128, 199-201, 204, 208, 211-212, Momentum, 262, 264, 268, 280-281
Motion, 89, 222-223, 248, 259-283, 293, 295, 305, determining, 219
214-215, 225, 227, 229, 232, 234, 236-237, Particles, 100, 199, 246, 260, 295, 310-311, 315-316,
242, 248, 272, 291, 295, 299, 301, 304, 306, 308, 319, 330, 354, 359, 383, 397-398, 406,
411-412, 416, 430-431, 434, 437, 443-445, 349-351, 353-354, 358-360, 365, 374, 382,
482
398, 451, 459, 476 321-322, 327-328, 330-331, 333, 338-342, rise, 47, 93, 101-102, 107, 116, 286, 288, 290-291,
Particulates, 75, 336, 346, 358, 360-361, 374, 382 346, 349-350, 352-353, 355, 360, 364, 375, 294, 299, 307, 310-311, 314, 321, 340, 371,
Parts, 138, 202, 219, 225, 227, 269, 307, 315, 317, 381, 394, 399, 407-408, 453, 458, 473 374, 401, 405-406, 411, 436, 475
353, 385, 389, 392, 419, 430, 435, 445, information, 274, 299 Roll, 460
449-450, 453, 456, 458, 461 Processing, 468 Root-mean-square average, 260
Parts per million (ppm), 385 Product, 59, 362 Rotary, 72-73, 206, 470-471, 477
Passengers, 430 Production, 47, 125, 219, 418 Rotation, 211, 216, 224, 226, 231, 248, 254, 256, 265,
Patterns, 207 rate, 219 267-268, 281-282, 287, 291, 294, 303, 309,
Pedal, 78, 218, 223, 227, 337 Products, 77, 88, 100, 270, 315, 348, 363, 398, 415, 311-312, 316, 318-319, 322-323, 330, 407,
Period, 58, 101, 210, 215, 243, 246, 265, 282, 286, 437, 458-459, 471 435, 444-445, 456
293-294, 305, 330, 348, 363, 370, 421, 437, Profile, 212, 324, 365, 424, 473 Rotational kinetic energy, 281
458, 461 Programming, 213 Rotational motion, 259, 261, 267, 280
Phase angle, 320 Programs, 128, 275 Rotations, 458
Phase shift, 211 Projects, 471 Roughness, 430-431
Photochemical smog, 348, 355-356 Propagation, 274, 286-289, 291, 293-295, 321-324, Roundness, 445
formation of, 356 470, 472 routing, 417
Photographs, 297, 373, 468 Propane, 297, 351, 371, 474 Rubber, 318, 356
Physical properties, 273, 413 Property, 89, 91, 102, 274, 404, 423, 431 Rules, 59
pin, 42, 59, 276, 392, 400, 445, 462 Prototype, 77, 307, 418
Pipe, 237, 327, 336, 365, 370, 381, 401-404, 418, Proximity, 259, 270 S
422, 424-425 Pulley, 212, 319 Saddle, 469
Pipes, 198, 335, 401-402, 424 Pump, 51-55, 78, 102, 105, 123-124, 141, 214-215, Safety, 77, 307, 417, 420, 437, 460, 473
Planes, 250, 445 218, 221, 223, 225, 228, 254, 300, 309, 317, factors, 77
Plastics, 356 369, 372, 412-414, 416, 418-421, 433, Sampling, 274
Platinum, 288, 359, 366, 373-374 436-438, 443, 449-450, 454, 456, 459-462 Sampling rate, 274
Plots, 50, 277 housing, 372 Satellites, 373
Plug, 72, 81, 84, 93, 130, 140-141, 207, 253, 265, Pumping, 57, 61, 66, 79, 105, 137, 140, 212, 279, Savings, 79, 240, 273
269-270, 282, 287-289, 291, 294-298, 317-318, 329, 381, 421, 433, 435-436, 438, Saws, 58, 128, 411-412
300-301, 303, 307, 316-317, 321-324, 350, 441, 456, 458 table, 128
356, 366, 391, 393, 399 Pumps, 78, 216, 363, 379, 427, 443, 450 Scale, 234, 260, 380, 417, 454, 467, 469
Point, 56, 74, 92-93, 104-105, 107, 110, 123-126, 130, Scope, 275
132-133, 202-204, 208, 236, 265, 275, 289, Q Scuffing, 466
293, 311, 328, 330-331, 333, 336, 341, 359, Quality, 44, 91, 219, 242, 252, 272, 288, 295, 299, 400 Seconds, 78, 254-256, 287, 303, 305, 309, 312, 316,
398-399, 401, 403, 407, 413-414, 421, 425, Quality control, 272, 295 323-324, 372-373, 380, 393, 421-422, 450
430, 448, 452-453, 461-462 Quantity, 93, 100, 228, 295 Selective, 477
Points, 54, 107, 126, 310, 336, 350, 391, 399, 430, Quenching, 101, 349-350 Sensor, 50, 218, 220, 222, 334, 399
437 Sensors, 78, 214, 219, 248, 274, 299-300, 319, 336,
Poison, 369 435, 452, 471
Polymers, 456 R
Radiation, 290-291, 362, 373, 395, 398, 406, 408 Service stations, 415
Population, 346-347, 363 Shape, 200, 207, 260, 280, 416, 426, 438, 471
population density, 346-347 Radiation shield, 362, 373
Radio frequency, 375 Shear, 431, 434, 454, 456, 469
Positioning, 265, 269, 420 Shearing, 431
Potential energy, 235 random variations, 298
Range, 40-44, 46, 56, 58-59, 65-66, 70-71, 74, 77, 91, Shell, 366
Power, 40, 42, 44, 51-53, 55-59, 61-63, 65-66, 69-71, Shield, 362, 373
73-79, 81-86, 88, 90, 93, 95, 98-102, 109, 100, 102, 107, 138, 199-200, 203-204, 212,
229, 238, 247, 274, 276, 308-311, 314-315, Short circuit, 460
112, 114, 122, 128-129, 132-139, 141-142, Short-circuiting, 243
201, 205, 212, 220, 229-230, 234-241, 243, 322, 336, 353, 356, 359, 363, 369-370, 374,
376, 380, 389, 395, 398, 402, 422, 453-454, Shutoff, 225
246, 250-256, 259-260, 262, 269-271, 273, SI units, 67, 80-81, 83
277, 279, 291, 294, 299-302, 304, 306-307, 456, 458, 461
argument, 336 Side, 73, 131, 225, 232, 242-243, 289, 302, 318, 338,
309, 311, 314, 316, 318-319, 321, 323-324, 359, 372, 396-397, 401, 412, 417, 426, 434,
327, 329, 335-337, 339-340, 342-343, 346, Rapid, 47, 260-261, 327, 348, 363
Rapid flow, 327 444-448, 461-462
349, 358-360, 371, 378, 382-385, 389-390, Silicon, 359, 418, 420
392, 398-399, 405-408, 410-411, 419-420, Ratios, 44, 47, 66, 91, 107, 112, 115, 117, 119, 124,
236, 272, 275-276, 300-302, 305-306, nitride, 420
422, 425, 430, 432-438, 442-446, 452, 456, Silicon nitride, 420
458, 461-462, 466-467, 470, 476-477 308-309, 311, 315-316, 335, 351, 356, 359,
363, 373, 411, 441, 452 SIMPLE, 115, 142, 225, 241, 250, 260, 262, 274, 285,
Power plants, 346 316, 353, 375, 379, 385, 405
Press, 200, 470-472, 475 Reaction, 64, 66, 70, 251, 287, 290-291, 293-294,
298, 306, 315, 336, 348, 350, 354, 356, Simple process, 250
Pressure, 46, 49-51, 53-56, 62-63, 74, 79, 81-84, Simulation, 273-275, 278-279, 467-468, 470, 473, 477
89-90, 92-95, 97-107, 109-142, 199, 358-359, 364-366, 372, 376, 383, 409,
444-445, 471 termination, 274
201-204, 210, 213-215, 218-219, 221-223, Single, 46, 69, 71, 78, 84, 89, 134, 141, 203-204,
225-228, 230-238, 241, 244-248, 250-255, Reactive, 355-357, 467
Reading, 61 214-215, 230-231, 256, 275, 288-289,
260, 269-272, 277-282, 285-286, 288, 295-297, 300, 302, 317-318, 342, 359, 365,
290-291, 294, 296-299, 302, 304, 307-312, Reasonable, 85, 109
Record, 52 405, 451
314-315, 318, 321, 327-333, 336-337, Sizing, 207, 308
339-342, 349-351, 354, 356-357, 360, 363, Recycling, 79, 341, 378, 383
Reduction, 77, 93, 102, 109, 205, 218, 228, 232, 253, Slope, 228, 291
371, 374-375, 379-380, 383, 405, 407-408, Smart, 299, 336
411, 414, 417, 420, 433-441, 444-445, 336-338, 347, 357-358, 365, 367, 369-371,
375-377, 380, 400, 430-431, 453, 476-477 Smog, 346-348, 355-356, 467, 473
448-450, 459-462, 468, 475 Soft, 78, 212
absolute, 130, 215, 252, 405, 417, 434 Refrigeration, 134
Refueling, 379 Software, 274
atmospheric, 74, 92, 103, 122, 135, 202, 225-226, Solar energy, 385
237, 241, 299, 330, 337, 341, 356-357, Regulators, 289
Rejection, 89, 93, 95, 112, 133, 136 Solid, 75, 77, 100, 247, 264, 310, 315-316, 346, 353,
414, 417 358-359, 365, 371, 374-375, 378, 398, 402,
center of, 248, 255, 298, 307 Relationships, 89, 96, 266, 274
Relative motion, 430-431, 437 408, 420-421, 423, 425-426, 430-431, 458
vacuum, 203-204, 222, 225, 227, 230, 371 Solid lubricants, 420, 458
vapor, 89, 100, 199, 201, 219, 227-228, 232, 251, reliability, 237, 388
Relief valves, 460-461, 474 Sources, 133, 286, 362, 364, 371, 379, 476
260, 272, 309-311, 351, 356, 363, 379, Sparks, 289
383, 417, 459-460 requirements, 40, 77, 128, 205, 306, 318, 348, 352,
412, 415, 424, 443 Specific gravity, 414-415
Pressure (p), 445 Specific heat, 91, 138, 235, 335, 337, 384, 409, 423,
pressure, 445 Reservoirs, 379
Resistance, 58, 61, 198-199, 207, 231, 271, 288, 308, 425
primary, 74, 79, 204, 230-231, 263, 269, 301, 319-322, Specific heats, 89-91, 100, 138, 142, 328, 377, 412
336, 356, 395 316, 371, 400-401, 417, 431, 456, 458, 460
Resistors, 219, 371 Specific volume, 50, 53-54, 59, 62-64, 82, 84, 90, 94,
Principal, 275 103-104, 106-108, 111, 113, 116, 118-119,
Problem statement, 97, 113, 120, 126 end, 371
Resolution, 274 123-125, 134, 235, 238, 329
Process, 51, 70, 88-90, 92-95, 100-103, 105-107, Specific weight, 59, 128
110-112, 115-117, 121, 123, 125, 129-131, Response, 77, 234, 274, 337, 380
Rhodium, 366, 373 Spheres, 359, 365-366
133, 135-136, 142, 201, 213, 227-228, 235, Spraying, 399, 419
238, 241-243, 247, 250, 260-261, 265, 267, Ribs, 380
Ring, 270-272, 282, 350, 391, 393, 434, 437-438, 446, Spread, 246, 267, 285-286, 289, 296, 301, 309, 321,
274-276, 285-286, 288-289, 291, 293-294, 459
296-299, 302, 307, 309-311, 314-315, 317, 453, 460, 466, 468, 470-473
Spreading, 261, 275, 310
483
Springs, 207, 212, 430 417, 419-426, 434, 436-437, 439-440, 460
Square, 42, 44, 48, 82, 84, 141, 202, 209, 229, 260, 452-457 Values, 41-43, 46, 55-56, 65, 70, 74, 85, 89, 91, 93,
434, 446 and friction, 437 96, 102, 105, 107-109, 128, 138, 228, 238,
Stability, 296, 308, 397, 453 Temperature sensors, 248 243, 251, 257, 262, 268, 274-275, 288, 290,
stable, 355-356, 415 Temporary, 100 330, 335, 371, 378, 388, 391, 396, 404-405,
Standard, 74-75, 77-78, 83, 88-89, 91-92, 94, 96-97, Termination, 274, 286, 293-294, 321-322 423-424, 443, 454, 456
99-104, 106, 108, 110-112, 114-119, Test, 40, 58, 274, 278, 297, 415, 420, 437 Valves, 41, 59, 70, 78-79, 94-95, 101, 107, 111-112,
122-126, 129-131, 133-142, 203, 212, 229, Testing, 58, 273-274, 346, 359, 415 115-116, 125, 128-129, 135-136, 140, 142,
242, 252-253, 274, 281, 289, 291, 295, 306, Tetraethyl lead, 363 198, 202, 204, 206-213, 241-243, 252,
312, 328-329, 331, 336-337, 341-343, 348, thermal, 44, 67, 70-71, 74-75, 85, 93, 95-96, 98, 254-255, 267, 269, 282, 288, 299, 301-302,
380, 382-383, 404, 412, 424, 458, 460 101-102, 112-114, 117, 120-122, 124, 304, 307-308, 328, 330-331, 333-335, 342,
atmospheric pressure, 103, 122, 135, 341 126-128, 133-134, 137-140, 142, 199, 205, 351, 363, 395, 399, 402, 407-408, 411,
deviation, 101-102, 295 216, 222-223, 227, 240, 251, 253, 255-256, 437-438, 440, 451, 460-461, 474
of mass, 100 261, 267, 270, 272, 274, 291, 299, 304, 306, Vapor pressure, 201
of temperature, 91, 102, 142 308-309, 311, 316, 319, 321, 323-324, Vaporization, 201, 215, 221-222, 227, 252-253, 307,
Standard deviation, 295 335-336, 346, 348, 350, 355, 362, 364-365, 309-310, 312, 335, 352, 409-410, 414
standards, 77, 305, 318, 347-348, 353, 375, 380, 456, 369-373, 376, 378, 382, 388-389, 391, 394, gasoline, 201, 215, 221, 253, 409
458, 469, 471, 474 396-398, 400-403, 405-406, 411-412, Variability, 212, 475
Stationary, 42, 58-59, 134, 219, 289, 318, 374-376, 414-417, 419-427, 436, 443, 458, 462, 475 Variables, 41, 90, 102, 200, 277, 300, 324, 376, 405
380, 383, 402, 408, 426, 450 Thermal conductivity, 371, 396-397, 400, 402, 411, Variations, 101, 212, 219, 242, 260, 295, 297-299,
Statistical, 252, 260, 295, 467, 474 426 302, 334, 367, 398-399, 473
quality control, 295 Thermal energy, 102, 355, 405 Vehicle, 58, 65, 67, 69, 77-78, 80, 83, 85, 227-228,
Status, 474 temperature and, 102 232, 330, 337-338, 347-348, 359, 373, 375,
Steady state, 70, 198, 391-392, 398, 417 Thermal expansion, 216, 272, 350, 365, 391, 400, 380-381, 385, 411-412, 422, 446, 450, 452,
Steady-state flow, 71, 231-232, 337, 403 412, 417, 436 461, 476-477
Steel, 206, 338, 365 Thermal radiation, 362, 373, 398 Velocity, 107, 139, 198-199, 201-203, 211-212, 219,
Stefan-Boltzmann constant, 398 Thermal resistance, 400-401 225, 229-230, 232, 248, 255, 260, 263-268,
Step size, 277 Thermally, 365, 405, 417 281-282, 288, 309-311, 313-314, 327-329,
Stiffness, 380 Thermocouple, 334 331, 333-335, 337, 340-342, 376, 397,
Stoichiometric air-fuel ratio, 343, 356 Thermodynamics, 91, 235, 275, 376, 378, 469-471 404-405, 407, 424-426, 434, 437, 443, 454,
Stops, 78, 85, 286, 312, 317, 323, 337, 354, 364, 379, first law of, 235, 275 467
421, 437, 446, 450 Thickness, 349, 396, 432, 446, 466 actual, 107, 310, 334-335, 425
Stored energy, 421-422 Three-dimensional, 466, 469, 471 Ventilation, 272
Strain gage, 215 Titanium, 338 Vertical, 232
Strategies, 285, 477 Tolerances, 272, 432, 445, 452 Vibration, 380, 430
Streamlines, 199, 207 Tool, 273 acceptable levels, 380
Strength, 41, 61, 207, 308, 397, 453, 460, 470 Tools, 80, 411 Vibrations, 380
bearing, 460 Top, 46, 79, 101, 232, 270-272, 282, 289, 350, 354, Viscosity, 212, 248, 316-317, 380, 397, 402, 404,
Stretching, 437 375, 393, 399, 403, 416, 438, 453, 473 420-421, 431-432, 434, 436-437, 439-440,
Strides, 69 Total, 44, 63, 76-77, 83, 85, 97, 100, 108, 141, 199, 446, 450, 454-458, 462
String, 353 201, 204, 207, 210, 212, 219, 224, 232, 240, kinematic, 439
Structure, 251, 365-367, 371-372, 375-376, 458-460, 243, 246-247, 269-270, 272-273, 280, 287, temperature of, 402, 404, 440
467 290, 300, 303, 308, 312, 330, 333-334, 336, Viscosity index, 454
of ceramics, 367 338-340, 348, 350, 353-354, 359, 370, 374, Viscous dissipation, 260
Sulfuric acid, 362, 369 377-379, 383-385, 388-390, 398, 407-408, VOL, 297, 361, 466-477
Surface, 44, 48, 68, 73, 199, 202, 206-207, 219, 260, 415, 419, 422-423, 425, 430, 434-435, Voltage, 77-79, 287-289, 301, 380
269-270, 276-277, 288-289, 308, 311-312, 437-438, 443, 450-451 Volume, 41-44, 46-50, 53-56, 59, 62-64, 68, 71-72,
316-317, 353-354, 359, 363, 365, 367, Total energy, 287, 388-390, 422 82-84, 89-90, 92-95, 97-98, 101, 103-104,
372-373, 375, 379, 382, 388, 391, 394-399, Total pressure, 199, 201, 232, 336 106-108, 110-116, 118-120, 123-125,
402, 405-407, 411-412, 415, 419, 424, 426, problem, 199, 232 128-136, 139, 141, 200, 203-204, 235, 238,
430-432, 434, 437, 445-446, 454 Toxicity, 415 243, 245, 255, 260, 262, 264-265, 267-273,
roughness, 430-431 Trace, 211, 355, 359, 458 276-278, 280-282, 290-291, 293, 300,
treatments, 363 Tractors, 470 303-304, 306-308, 311-312, 316, 318-320,
Surface damage, 411 Trade, 361 323-324, 328-330, 332-333, 339, 349-352,
Surface hardening, 363 trade-off, 361 354, 365, 370, 372, 375, 382-384, 401, 405,
Surface roughness, 431 Training, 415 414, 417, 433-435, 460
and lubrication, 431 Transducer, 50, 215 Volumes, 44, 91, 107, 272, 350, 409
Surface temperature, 277, 308, 373, 426 Transfer, 44, 89-90, 96, 100-101, 117, 201, 235, 259, Voluntary, 456
Surface treatments, 363 261, 274-276, 280, 288, 290-291, 304-305, Vortices, 295
Surfaces, 89, 207, 242, 261, 270, 351-352, 359, 363, 321, 327-328, 380, 387-427, 466-470, 475 Vt, 268
373, 383, 391, 394, 400-401, 411-412, mechanisms, 394
415-416, 419-420, 430-432, 434, 437, 446, Transportation, 59, 67 W
450-451, 453-454, 458 Tube, 225-230, 232-233, 254, 379 Walls, 89, 199, 201-202, 227-228, 241-243, 248, 259,
Switches, 444 Tunnels, 58 261-262, 265, 267, 270-271, 275, 280, 293,
Synthetic, 458-459 Turbulent flow, 404 305, 308, 313-314, 316-317, 349, 351-352,
System, 46, 55, 64-65, 72, 77-79, 82, 88-90, 93, 100, Turning, 40, 237, 300, 432 363, 371, 382-383, 392, 394, 396-401, 404,
102, 105, 107, 123-124, 128, 130, 142, Two-stroke engines, 468 411, 414-415, 420-421, 424, 434, 437-438,
198-199, 201-204, 206, 210-219, 221-223, Types, 61, 91, 125, 199, 221, 239, 242, 286, 289, 302, 444-446, 449, 451-454, 461-462
225, 227, 230, 232, 234, 236-237, 241, 243, 365, 449, 461 assemblies, 437
246-247, 251, 254, 256-257, 261, 263, Warning, 300
268-269, 272, 274-276, 282, 288-289, 293, U Waste, 236, 417, 421
295, 305, 315, 319-320, 323-324, 327-331, Uniform, 291, 412, 420 Water, 59, 78, 84, 201, 232, 253, 356, 362, 367,
334-339, 341, 343, 359, 363, 365, 367, Units, 51, 56, 67, 75, 78, 80-81, 83, 217, 347-348, 372-373, 376, 382, 384, 388, 391, 394, 397,
371-372, 374-377, 379-381, 383, 385, 394, 380, 470 399-401, 408-410, 412-421, 423-427, 430,
402, 407, 409, 412-415, 417-423, 425-427, unknown, 378 438, 443, 459, 476
435, 446, 449-450, 458-462, 467-468, Urea, 376, 477 cooling, 78, 201, 232, 391, 394, 399, 401, 408-410,
474-476 Us, 41, 88 412-417, 419-421, 424, 427, 443
physical properties of, 413
T V specific gravity, 414-415
Team, 69, 477 Vacuum, 108, 203-204, 222, 225, 227, 230, 305, 371, specific heat of, 384, 423
temperature, 70, 74, 84, 90-94, 96-98, 100-111, 381, 421 Water vapor, 232, 356, 362, 367, 372-373, 384, 410,
113-114, 116, 118-122, 126-128, 132-135, evaporation, 227, 421 459
137-142, 201, 209, 212, 219, 222-223, Value, 66, 71, 74, 89-91, 97-98, 101, 105, 108-109, Wave, 204, 336, 381
227-229, 235-239, 246, 248, 251-255, 260, 112, 120, 125, 127, 138, 141, 207, 209, 225, speed, 204, 336, 381
270, 272-273, 277, 279, 281-282, 285, 229, 236, 252, 267, 275, 291, 296, 309, 334, Wear, 78, 89, 206, 288, 317, 330, 351-352, 363, 378,
287-288, 290-291, 294-295, 297, 299, 362, 374, 389, 397, 400-401, 410-411, 417, 400, 421, 445-446, 450-451, 453, 456,
306-310, 312, 314, 316, 321-322, 327-328, 454, 456, 460 458-459, 461, 472
330-332, 334-343, 349, 355-360, 362, added, 91, 98, 138, 207, 236, 252, 296, 410, 456, parts, 317, 445, 450, 453, 456, 458, 461
364-365, 367-378, 383-385, 388, 391-415, severe, 351
484
Weather, 85, 316-317, 338, 346, 373, 392, 414, 418,
421, 454, 458
Weight, 59, 69, 128, 201, 234, 348, 351-352, 362, 375,
409, 411-413, 420, 424, 430, 453, 458
Weld, 431
Well, 110, 134, 200, 236, 246, 251, 274, 289, 291,
299, 304, 311, 316, 327, 346, 355, 366, 372,
381, 389, 397, 419, 422, 451, 456, 461
Wire, 77-78, 219, 374
Work, 40, 49-57, 62-63, 70, 82, 90, 93, 98-105, 114,
121-124, 127-128, 132-133, 137, 139-141,
202, 212, 221, 223, 274-275, 286, 289, 291,
294, 302, 309, 317-318, 321, 327, 329-331,
346, 373, 375-377, 381, 385, 388, 419-420,
432-436, 441, 456, 469
Y
Yielding, 315
Z
Zinc, 359, 369, 453
485