Sunteți pe pagina 1din 16

TRCP 3709

Combustion Chamber Deposits in Spark-Ignition Engines:


A Literature review
by
Gautam T.Kalghatgi

Shell Research Ltd., Thornton Research Centre, P.O.Box 1, Chester CH1 3SH, U.K.

Paper proposed for presentation at the SAE Fuels and Lubricants Meeting,
October 16-19, Toronto, Canada

April 1995

1
Kalghatgi

952443

Combustion Chamber Deposits in Spark-Ignition Engines: A


Literature Review
Gautam T. Kalghatgi
Shell Global Solutions U.K.

viz carburetters, port fuel injectors and inlet valves has


matured and the use of such additives is being increasingly
mandated. There has been concern that some of this
technology might cause an increase in CCD formation and
associated engine performance problems. The spurt in
interest in CCD is reflected in industry initiatives like the
1993 CRC Workshop on Combustion Chamber Deposits ;
the Proceedings of this workshop (2) contain a lot of
valuable information. An earlier source of information on
CCD is (3). This paper is a review of the literature more
focused on CCD than in (1). It contains material that was not
considered in (1) and is meant to be read in conjunction with
(1). The fuel of interest is unleaded gasoline unless
otherwise mentioned.

ABSTRACT
Deposits, derived primarily from the fuel but with some
contribution from the oil are formed inside the combustion
chamber of a spark-ignition engine with use. The growth of
combustion chamber deposits (CCD) is a dynamic and, to an
extent, reversible process which at any given time reflects
the balance between the formation and removal processes.
Engine surface temperature is the most important parameter
that affects their formation and changes in engine operation
which tend to increase surface temperature, reduce deposit
growth. At a fixed temperature, less volatile fuels tend to
form more deposits than more volatile fuels. Some detergent
additive packages tend to increase CCD levels. CCD reduce
the heat lost to the coolant and increase charge temperature
thereby increasing flame propagation rates but reducing
volumetric efficiency; they might also affect the final phase
of combustion by as yet undefined chemical means. This is
reflected as an increase in octane requirement and NOx and
a reduction in maximum power but an improvement in fuel
economy and a reduction in CO2 emissions. They might also
lead to higher HC emissions but not necessarily always
since there might be competing mechanisms which come
into play. CCD effect on CO emissions is not clearly
established. They can also cause other interference
problems like carbon rap. It is not known to what extent
engine performance is affected by small changes in CCD
levels. There is a large variation in deposit growth and its
response to changes in fuel, additives and engine operating
conditions across the combustion chamber and between
different engines. Similarly, the performance of different
engines will be affected to different extents by the deposits.
While assessing the effects of different fuels or additives on
engine performance and emissions through their effects on
CCD, the simultaneous effects on other aspects such as inlet
valve deposits which might have their own effects on engine
operation, should also be considered. The paper reviews the
literature on these topics.

2. GROWTH OF COMBUSTION CHAMBER


DEPOSITS
2.1 DEPOSIT FORMATION AND REMOVAL Combustion chamber deposits are derived primarily from the
fuel and to some extent from the engine lubricating oil in
typical modern engines with low oil consumption running on
typical, full-boiling-range fuels (e.g. 4,8,9,10). The primary
deposit formation mechanism appears to start with radical
initiated addition/substitution reactions on fuel and lubricant
components to produce coordinated oxidation products;
these products then condense on the hot surfaces and
undergo polymerisation (11,13). The polymerisation
mechanism is not simple pyrolysis (12). For example
deposits formed by heating oils in shallow dishes or from
fuels burnt in diffusion flames contain less than 2% oxygen
as opposed to 16-25% in CCD. However when films of oil
were subjected to periodic ignition and extinguishment,
deposits similar to CCD in terms of oxygen concentration
and pyridine solubility were obtained (12). Lauer and Friel
(12) suggest an "oxidative pyrolysis" of deposit precursors
which condense on the surfaces. Price et al suggest that this
process starts with the reactions caused by radicals that
diffuse into the quench layer adjacent to the wall from the
flame (13). Partial oxidation reactions occur ; further
surface reactions prevent the evaporation of such material
and the deposit formation process continues (13). The
condensation of the deposit precursors on the surface is the
critical step in this process and deposit formation is strongly
dependent on the surface temperature (4,5,6,7,14,22) higher surface temperatures reduce deposit levels.

1. INTRODUCTION
The literature on the deposits that form with use on the
internal surfaces of a spark ignition engine was reviewed in
(1). In the period following the publication of (1), new field
problems like carbon rap (see below) attributed to
combustion chamber deposits (CCD) have been recognised.
Moreover, fuel and additives technology to ensure
satisfactory control of deposits in other parts of the engine

As the deposits grow, the surface temperature increases


because of the insulating properties of the deposits
2

Kalghatgi

mechanisms like oxidation and gasification (29), physical


mechanisms like desorption and evaporation of volatile and
gaseous components and mechanical removal like abrasion,
flaking brought about by thermal stresses and mechanical
wash-off. In many of these mechanisms e.g., oxidation, as
the surface temperature increases, the rate of deposit
removal would increase (29). Price et al have studied the
deposit removal process to an extent in a laboratory rig (13).
The details of such removal mechanisms and the extent of
their individual importance are difficult to assess in the
engine context though some useful attempts have been made
recently (29). Deposit weight and any other property
measured at a given time reflect the balance of the deposit
formation and removal processes upto that time.

(1,5,15,16). This can be inferred from the composition


variation across the thickness of the deposit (12,44,51,60)
with low-melting-point materials at the wall to
high-melting-point materials at the deposit surface. For
instance, Nakamura et al (44) demonstrated sharp
concentration gradients from unleaded gasoline with high
concentrations of carbon near the wall and high
concentrations of inorganics such as Ca and Zn (presumably
from the oil) at the deposit surface which Chapman and
Williamson (60) have shown to be in the form of calcium
sulphate and zinc phosphate. Lauer and Friel (12) speculate
that acid and anhydride groups in the carbon-rich layers
adjacent to the wall are responsible for their adhesion to the
wall. The measurement and modelling of heat transfer
through CCD (see section 5 below) also confirm the increase
in surface temperature as deposits grow. Hence as CCD
grow, only the less volatile deposit precursors will condense
and contribute to deposit growth; the results of
thermogravimetric analysis (TGA) of deposits collected at
different stages of an engine test by Daly et.al (18) confirm
this. Daly et al (18) also report that as an engine test
progressed, the aliphatics to aromatics ratio as determined
by solid state NMR decreased i.e, the newer deposits were
relatively more aromatic. It has to be said that Daly et.al (18)
interpret their results in terms of a more complicated model
for deposit formation. However their observations are fully
in line with the simpler physical picture of deposit growth
described above and in (5). Ebert et al (42) have also shown
the ability of the deposits to adsorb aromatics from the fuel.
They suggest that deposit growth occurs through a
chain-growth mechanism with adsorbed species being
oxidatively bound to the existing deposit.

2.2 DYNAMIC NATURE OF DEPOSIT GROWTH Deposit growth is a dynamic and to some extent reversible
process (e.g.30). For instance deposit levels in an engine can
be reduced to a new equilibrium level if some deposit
formation is switched off and deposit removal rate is
increased say by increasing the surface temperature for a
given fuel (4). Similarly for the same engine condition,
replacing the fuel with a low boiling point fuel like isooctane
(1) or even replacing the highest boiling fraction of the fuel
with another of lower deposit forming tendency (30) will
switch the deposit formation off temporarily because the
surface temperature will be above the critical temperature
for the new fuel and deposit weights and thicknesses will
move to a new, lower equilibrium.
2.3 VARIATION IN DEPOSIT GROWTH - The
average metal surface temperature varies widely across the
combustion chamber, from around 120 C (31,32) in the
cooler areas to about 320 C at the exhaust valve seats to over
800 C at the exhaust valves (33). During an engine cycle,
surface temperature can increase by 400 C from its base line
value (29) in the cooler areas. Piston surface temperatures
vary between 200 C and 300 C depending on location and
engine operating condition (32,33). Piston temperatures are
also 10 C to 25 C higher than corresponding combustion
chamber surface temperatures (32). The exposure of the
surface to liquid fuel or additive droplets or to combustion
products which might be important in deposit removal and
formation mechanisms will also not be uniform across the
cylinder. Hence there are very large differences in the
weight, thickness and properties of deposits from different
parts of the combustion chamber (1,21,22,34-36).

In general, at any given engine condition, it is the highest


boiling fractions of the fuel that contribute most to CCD
formation (20). For a given fuel, there will be a maximum
surface temperature above which little deposit will form.
This critical temperature was found to be 310 C for indolene
by Cheng and Kim (4) and 320 C for another unleaded
gasoline by Nakic et al.(14). For lubricating oil, which has a
higher boiling point compared to the fuel, Cheng (5) reports
that the critical surface temperature for deposit formation is
about 60 C higher. For fuels of low boiling point like
isooctane this critical temperature will be lower than any
normally found in the combustion chamber and very little
deposit from the fuel will be formed (e.g. 6,7,12,19). In such
cases the deposits will primarily be derived from engine oil;
Tsukasaki et al report that in cars running on methanol,
which has a low boiling point, CCD consisted mainly of
calcium originating from the oil (17). Engine oil appears to
affect the formation of piston-top deposits more than the
head deposits (see Section 5 below). For a given engine
operating condition and hence a surface temperature regime,
if higher boiling point materials, such as some fuel additives,
are added to a given fuel, more deposits will be formed (e.g.
10,14,15, 18, 21-27) and this difference is likely to be more
marked, the cleaner the original fuel is.

In general, surface temperatures increase with engine


speed, throttle opening, coolant temperature and to a smaller
extent with ignition advance and compression ratio except
for exhaust valve temperatures which decrease slightly with
an increase in compression ratio (32,33). They also increase
as the mixture is made richer, because of the increase in the
temperature of the burnt gas, peaking at an air fuel ratio
about 15% fuel rich compared to stoichiometric conditions
(33).
Most of the experimental observations discussed below
can be understood within the context discussed above. These
studies range from single cylinder engine studies using

As the deposits are being formed, they are also being


removed through various mechanisms as listed by
Lepperhoff and Houben (28). These include chemical

3
Kalghatgi

are high, are composed almost entirely of inorganic


compounds (6,19,36,44). Deposits in the end-gas region i.e
the region where the flame front arrives last, have lower
oxygen content and higher carbon content (19,36),
indicating a less oxidised state, compared to deposits from
other parts. Gebhardt et al (19) also found lower quantities
of volatile material compared to non-volatiles in deposits in
the end-gas region which they interpret as indicating
condensation of heavier aromatics in this area. There can
also be substantial differences, e.g. in H/C values, in
deposits from different engines (19,36) and the same engine
operating at different conditions (36) .

removable probes (e.g. 4,8,9) to tests on production engines


on test stands and road tests.
3. PROPERTIES OF COMBUSTION CHAMBER
DEPOSITS
In comparison with fuel and oil, on average, CCD from
unleaded gasolines have much higher weight concentrations
of oxygen (16%-25%) and nitrogen (1%-2.5%) and much
lower H/C ratios (0.6 to 1 compared to 1.7 to 1.9 molar)
(22,36,42,48,60); the carbon content is between 60% to 70%
by weight. Thus a substantial change of the fuel and oil
involving a large loss of alkyl components and a significant
degree of oxidation takes place before their inclusion in
CCD. The deposits also have a substantial volatile
component - around 35 weight percent of the initial deposit
is lost if it is heated to 500 C (19,42). The presence of lead
increases deposit weights considerably - by a factor of up to
8 over unleaded gasoline ; the weight of the deposit is not
related to lead concentration in the fuel above a limit of
about 0.15 g/l (1). In deposits from leaded gasolines, lead
constitutes a substantial (upto 70% wt) proportion of
elemental composition (60). Other metallic elements,
predominantly from the oil also show up in the deposits
though, naturally, they are more prominent - between 2%
and 6% - in deposits from unleaded gasolines (42,60). The
concentrations of Zn, Ca and P, all coming from lubricant
additives, are upto four times higher in piston-top deposits
compared to head deposits (22,41,61); thus engine oil
appears to affect the formation of piston deposits more
compared to head deposits (22,61).

The most commonly measured deposit properties are


thickness and weight. Thickness measurements can now be
routinely made in different parts of the combustion chamber
using commercial probes (e.g. 10,21,34,35,37-39) though
other techniques like an optical microscope have also been
used (15). Such measurements show the differences in
deposit thickness in different parts of the combustion
chamber, e.g lower thicknesses in hotter regions (5,15,37),
and different ways of estimating an average thickness from
such measurements have been recorded (e.g.34,35). Pistontop deposit weights are usually around 45% of the total CCD
weight (21,40). In a given engine test with fixed
fuel/additive combination, there is usually a reasonable
correlation between any of the thicknesses measured and
deposit weight although there is a lot of scatter (34,38,39).
However, there can be significant differences in the response
of different engines to the same additive - for instance, an
additive which more than doubled deposit thickness
compared to the base fuel in one engine test can produce a
statistically insignificant effect in another engine test ( e.g.
Engine C vs Engine F in 35).

Densities of around 1530 kg/m have been reported for


deposits removed from the combustion chamber for
commercial unleaded fuels (41). In other studies where
deposit volume was estimated through detailed thickness
3
measurements, average densities of between 1100 kg/m and
3
2000 kg/m were found with unleaded gasolines (43).
Measured heat capacities of deposits removed from the
combustion chamber vary from 0.84 kJ/kgK to 1.84 kJ/kgK
(41,44); there appears to be a good positive correlation
between the carbon content of the deposit and its heat
capacity (44). Thermal conductivity is more difficult to
measure and for deposits taken out of the engine depends
very much on the way they are prepared into pellets prior to
measurement (44). These values have ranged from 0.17
W/mK to 0.8 W/mK (see Table I in Ref.46). Thermal
characteristics are likely to depend on the morphology of the
deposit and need to be measured in-situ. Such methods have
been developed and depend on making in-situ temperature
measurements and inferring the thermal properties by
modelling the heat transfer through the deposit (e.g.15,16).
Thermal diffusivity of CCD from an unleaded gasoline has
-7 2
been measured to be around 3.5 x 10 m /s (15). Hayes et
al. report a wide variation in the thermal diffusivity from
-7
-7 2
0.59 x 10 to 3.4 x 10 m /s in deposits from different
locations in the engine (47).

4. DEPOSIT STRUCTURE
Different descriptions of the CCD structure have been
given from different view-points. The carbon is amorphous
(non-graphitic), extremely porous and characterised by a
heterogeneous granular structure representative of a carbon
produced by a pyrolitic mechanism (41). The exhaustive
analytical studies of deposits from unleaded fuel by Ebert
and co-workers (42) show that CCD are volatile
carbonaceous materials more like bituminous coal than
graphite. Between 85% and 90% of the carbon is aromatic this aromaticity is higher than that found for coals of
comparable chemical composition. Their structure is a
composite one consisting of a refractory "skeleton" which is
rich in oxygen and poor in hydrogen compared to the
volatile, smaller molecules adsorbed on this backbone (42).
Similarly, Chapman and Williamson (60) report columnar
and dendritic microstructures made up of inorganic material
with amorphous carbon filling the gaps between these
structures. The volatile material which is not bound so
tightly to the deposit closely resembles the heavy ends of the
fuel from which the deposit is probably formed. Unburned
fuel components and oxidation products derived from fuel
and oil can be detected in the deposits; these include
carboxylic acids, metal carboxylates, esters, ketones,
aldehydes and lactones (8).The primary molecular building

There are substantial differences in deposits from


different parts of the combustion chamber. For instance,
deposits on exhaust valves, where the surface temperatures

4
Kalghatgi

blocks appear to be 2-4 ring PNAs originating in the fuel


(11). These building blocks appeared to be independent of
the fuel composition or the severity of the test cycle in a
given production engine (11).

starts decreasing again if the mixture is made leaner beyond


a limit ( also see 1.3 above).
Compression Ratio - Changing the compression ratio
from 7 to 8.75 had little effect on deposition rate (4).
However Shore and Ockert (7) state that deposition rate
went down as wall temperature was increased by raising the
compression ratio though they do not show any direct
evidence to support this claim. The change in surface
temperature brought about by an increase in compression
ratio is not uniform across the combustion chamber (33).

High-aromatic fuels have been shown to yield


condensed, high-aromatic deposits and low-aromatic fuels
produce "fluffier" deposits (11,49). The structures of the
deposits on the head and on the piston were not found to be
substantially different but there were cylinder to cylinder
variations probably caused by temperature gradients (11).
The molecular structure of deposits from two production
engines running on two different cycles was similar (11,49)
but differed significantly from the deposits from a Honda
generator running on low load (11). This was attributed to
differences in surface temperature and the consequent
differences in the degree of condensation of deposit
precursors (11). On increasing the load on this engine,
deposit molecular structure changed, to resemble that in the
production engines (11).

Ignition Advance -Increasing the spark advance


increases metal temperatures (33) and Shore and Ockert (7)
claim, again without showing direct evidence, that this
reduces CCD formation.
Surface Material - Cheng and Kim (4) found that
deposition rate was independent of surface material as long
as the thermal conductivity was high enough; they tested
aluminium, brass, cast iron, sapphire, stainless steel and
macor ceramic. Deposition rate was very low, however, on
the ceramic coupon. Aluminium heads and pistons run
roughly 40 C to 80 C cooler compared to cast iron ones (31)
and consequently could be expected to be more prone to
deposit formation.

5. FACTORS THAT INFLUENCE DEPOSIT


GROWTH
5.1 EFFECT OF ENGINE PARAMETERS -In
general, operating conditions that tend to increase surface
temperatures lead to a reduction in CCD formation.

5.2 EFFECT OF FUEL PROPERTIES -The


conclusions from the pioneering study of Shore and Ockert
(7) have been confirmed in many other studies.

Engine Speed and Load - As engine speed or load is


increased, deposit formation is reduced (7,40,50). In a fixed
duration test, total fuel consumption increases as speed/load
increases and it was found that there was a good negative
correlation between total fuel consumption and CCD weight
(40) as speed/load increased. In fact deposit levels can be
substantially reduced in an engine already with CCD by
running it at a high speed or load for a relatively short time
(51-53).

The most important fuel property is the boiling point and


the most critical fuel components in deposit formation are
those with the highest boiling points (20). At a given boiling
point, taking the major gasoline components, aromatics are
the most prone, paraffins are the least prone with the olefins
in between for CCD formation and boiling point is the only
factor determining the deposit forming tendencies of
aromatics (7). Price et al have confirmed these findings of
Shore and Ockert (13). They also report that cyclic alkanes
and olefins show greater deposit forming tendencies
compared to their straight chain counterparts (13).
Polynuclear aromatics (PNAs) are strong deposit formers
depending on their structure; angular arrangements of
aromatic rings produce more deposits than structures with
linear configuration of rings (13). Also the deposit-forming
tendencies of PNAs increases with the number and length of
the alkyl side chains attached to the backbone of the
molecule (13).

Coolant Temperature- Cheng and Kim found that as


coolant temperature was increased from 40 C to 95 C,
deposition rate decreased (4). At the higher coolant
temperature, dark brown, smooth deposits which were hard
to remove were formed as compared to black soot-like
material at lower coolant temperature which could be easily
removed with acetone. Nagao et al also report that
increasing the coolant temperature from 50 C to 90 C
reduced CCD levels in an engine test and deposits produced
at the higher coolant temperature were less soluble in hexane
and benzene (22).
Inlet Temperature -When inlet temperature was
changed from 26 C to 100 C in (4) there was no change in
deposition rate. However, Myers et al. have reported that by
reducing the liquid phase in the intake system by increasing
the mixture temperature, CCD could be reduced (20); this
might have been because they conducted their tests in
carburetted engines.

Price et al have been able to unify the treatment of the


CCD-forming tendencies of different generic forms of
hydrocarbons by considering their molecular structure. They
have defined a parameter derived from both the boiling point
and the degree of unsaturation present in a precursor
hydrocarbon - termed the nominal boiling point - which
correlates well with deposit-formation rates in a laboratory
rig spanning over four orders of magnitude with different
hydrocarbon molecules (13).

Air/Fuel Ratio - Mixtures leaner than stoichiometric led


to more deposits than rich mixtures (4) ; deposition rate

5
Kalghatgi

Maxey (48) also found that even when some additives


increased CCD growth, the additive chemical structure had
little effect on the deposit chemical structure. They found no
intact additive backbones or head groups in the deposits.
Thus additives seem to undergo substantial transformation
before being incorporated into the deposits. However Nagao
et al were able to identify structures associated with the
additives in the CCD (22).

There are now many published empirical studies which


show that combustion chamber deposit formation tendency
of a fuel increases as the aromatic content of a gasoline
increases (8,11,24,30,49,55-57). Gibbs (57) states that CCD
formation tendency increases, in terms of hydrocarbon
groups, with saturates, C7+C8 aromatics, olefins and C9 and
higher aromatics . However in all these studies the effects of
changing composition are not decoupled from the effects of
changes in the volatility characteristics. It is possible that, at
the mechanistic level, the observed changes in CCD
formation in many of these reported tests result primarily
from changes in the boiling characteristics of the fuel .

Nagao et al (22) have been able to correlate the CCD


forming tendencies of a polyether and four polyolefin
detergents to the amount of residue left after three hours in a
thermogravimetric analysis (TGA) rig at 300 C. Thus
additives which are more thermally stable tended to cause
more CCD in these tests. However, such detergents are also
more likely to be effective for inlet valve deposit (IVD)
control in engines which have high inlet-valve temperatures usually small engines runnning generally at higher speeds
which are more common in Europe. It is also possible that
some additives decompose at higher temperatures into
components which promote CCD formation. Such additives
might produce less residue in a TGA test such as described
in (22) but might still lead to more CCD formation.

Jackson and Pocinki (27) have demonstrated a rough


positive correlation between CCD weight and washed gum
level in the fuel in two different engine tests. Takei et al (50)
have shown reasonable positive correlations between the
unwashed gum level in the fuel on the one hand and
piston-top and cylinder-head deposit weights and thicknesses
on the other. Nagao et al (22) have also shown a weak
correlation between unwashed gum levels of a gasoline
containing different additives and CCD levels. Gum levels
will reflect the concentrations of the most involatile
components in the fuel and hence the observed correlations
are not surprising.

5.4 EFFECT OF ENGINE OIL - Deposits usually


contain small quantities of Ca, Zn and Mg which come from
lubricant additives. The concentration of these elements,
usually in the form of sulphates or phosphates, is highest on
the surface of the deposits (60). Compared to deposits in the
head, deposits on the piston top have higher (by upto four
times) concentrations of these elements (22,41,61)
suggesting that oil plays a bigger role in forming piston
deposits.

5.3 EFFECT OF FUEL ADDITIVES -Modern


detergent packages, many of which contain carrier oils and
other co-additives in addition to the detergents, can lead to
an increase in CCD formation (e.g. 10,21-28) in unleaded
gasolines depending on their concentration and composition.
Some detergent additives (10,23-25,30,35) and some
carriers (27) are more likely to increase CCD than others in
particular engine tests. In several engine and road tests, a
polyether amine detergent has been shown to produce less
CCD compared to polyolefin derived detergents (22,25,30).
In other engine tests a PIB amine detergent package at the
right dose rate was shown to be at least as good as a
polyether amine for CCD (24,50,59). Of course, the
detergent additives have to be used at sufficient
concentrations to do their primary job, i.e. to keep the fuel
and inlet systems adequately clean. Hence their propensity to
form CCD has to be compared at concentrations that deliver
comparable inlet system detergency performance. The
ranking of additive packages, in terms of their ability to
control inlet system deposits, can depend on the engine or
vehicle test used for such ranking (1). This is quite likely to
be the case also for CCD formation from additive packages.

CCD weights increase if oil consumption increases


beyond a limit (9,36,63). However modern engines are
designed to use little oil and if they are running normally, no
significant correlation between CCD weight and total oil
consumption is found (40).
Thus engine oil contributes to CCD formation but the
extent of this contribution is difficult to assess (41). When
engines are run on fuels which themselves are unlikely to
form deposits such as hydrogen (62), methanol (17) or
alkylate (39), substantial amounts of deposits, predominantly
derived from oil (17), can be formed. For example Moore
(39) reported that in a car running on alkylate, 472 mg of
deposits were formed in the squish area of a car in a road
test compared to 2286 mg in the same car running on an
unleaded gasoline. However, one would not be justified in
concluding from these figures that in the latter case, 20% of
the CCDs were derived from oil because the surface
conditions which determine the dynamics of the CCD
growth would have been quite different in the two cases.
With normal full boiling range fuels running in modern
engines which burn little oil, CCDs are likely to be
predominantly derived from the fuel. If the route to deposit
formation through the fuel is switched off e.g. by using a
low-boiling-point fuel, the lubricant components come into
play and CCD grow towards a different equilibrium.

As discussed in section 3, there can be significant


differences in the response of different engines in terms of
CCD formation to the same additive (e.g. 35). Also the
effect of an additive in a given engine test would probably
be more marked in a clean fuel than in a fuel which would
produce more deposits by itself (24).
Compared to CCDs from base fuel, deposits produced
with fuels containing detergent additives tend to have lower
aromaticity (18,48), higher hydrogen content (48), higher
fraction of organic solubles (22) but similar quantity and
kinds of organic oxygen functionalities (48). Keleman and

6
Kalghatgi

ORI has been reviewed in (1) and only the salient points will
be mentioned below.

The dominant oil component for CCD formation is the


base oil - increasing the high-molecular-weight and
low-volatility content of the oil increases CCD formation
(9,64,65). Cheng (9) found that changes in lubricant
additives had negligible effect on CCD growth. However,
work by Takei et al (50) suggests that lubricant additive
dosage levels might be important. They report that an SG
grade oil, compared to an SE grade oil made from the same
base oil, resulted in higher CCD weight but lower thickness
(50).

In general, the larger the initial octane requirement


(IOR), the lower the ORI (1). New cars, even of the same
model, can have widely varying octane requirements but
with use, the octane requirements are distributed in a
narrower band. In very general terms, ORI is increased by
factors that tend to increase combustion chamber deposit
quantity such as fuel properties like high boiling points (77),
aromatics (30,55,56) or the presence of some additives
(64,77-79). However, some fuel additive packages have
been shown to be either neutral or beneficial i.e. reduce ORI,
compared to base fuel (74,80,81) in some engine and road
tests. Similarly, factors that cause removal of CCDs such as
hard driving modes (51,52,62,64,78) or running a dirty
engine on isooctane (66), can reduce the octane requirement.
In well controlled engine tests a reasonable overall
correlation between CCD weight and ORI can be established
(44,80). However this is the exception rather than the rule
and usually, CCD weight does not correlate with ORI
(44,79,82-85). Deposits in different parts of the combustion
chamber have varying degrees of effect on ORI. For
instance, deposits in the end-gas region, the region where
combustion takes place last, appear to be the most
significant (64) whereas piston-top deposits contribute little
to ORI compared to head deposits (37,80). In some engines
inlet valve deposits contribute significantly to ORI
(63,74,87,89).

6. EFFECTS OF COMBUSTION CHAMBER


DEPOSITS ON ENGINE OPERATION
The effect of CCD on heat transfer at the combustion
chamber walls has been demonstrated experimentally in
many independent sudies (e.g.15,16,32,37,41,44,46,47,53,
54). They act as insulators and also as heat resesrvoirs,
storing heat in one cycle and releasing it to the fresh charge
in the next cycle. They also occupy volume, increasing the
compression ratio. Finally they might also absorb and
release unburnt fuel, pro-knock species and promote
chemical reactions through catalytic effects.
As deposits build up in the combustion chamber, the
maximum (32) as well as the average (37) heat flux away
from the combustion chamber decreases, heat lost to the
coolant decreases (51,66,67), air consumption rate is
reduced (51,66,68) volumetric efficency decreases (53,6870) and flame propagation rates increase (37,67,68). Thus
the incoming charge is heated by the deposits; Harrow and
Orman estimated that deposits increased the temperature of
the trapped charge by about 11 C in their experiments (68).
Modelling of the heat transfer through the deposit layer also
shows that the deposit surface temperature, the temperature
of the bulk gas and of the end-gas all rise as deposits build
up. (69, 72-76). Studzinsky et al. estimate that the bulk gas
temperature increases by about 10 C because of volumetric
effects and by about 50 C from heat transfer changes owing
to the deposits.

The primary reason why CCD cause ORI appears to be


because they increase the end-gas temperature by storing up
heat from one cycle and giving it up to the fresh charge in
the succeeding cycle - the increase in surface temperature
caused by CCD also plays a role in this. The volumetric
effect of CCD on ORI accounts for about 10% of the total
effect in unleaded gasolines (64). In addition, there might be
a chemical mechanism - one possibility is that CCD absorb
stable pro-knock species like hydrogen peroxide and release
them in the end-gas to promote knock (75). The extent of
such chemical mechanisms and the balance between
chemical and physical mechanisms will almost certainly
depend on the nature of the deposits and the fuel being used
and at this stage is not understood. How inlet system
deposits affect ORI, when they do, is also not understood.
Thus ORI is a complex phenomenon which is not well
understood even after decades of extensive study.

The observed effects of CCD on engine operation


discussed below can be understood to an extent in the
context of the discussion above.
6.1 OCTANE REQUIREMENT INCREASE ( ORI ) The tendency of a car to knock and hence its octane
requirement (OR), increases with use as deposits build up. In
bench engine tests or vehicle dynamometer tests where
operating conditions are either constant or consist of
repetitive cycles, OR increases rapidly in the early stages
and reaches an equilibrium value as deposits stabilise.
However in field tests, an equilibrium octane number as well
as the distance when equilibrium is reached are some times
difficult, if not impossible to define (61,63). It has been
argued that concepts such as ORI can be applied to car
populations only in statistical terms (63). Typical values of
ORI ("stabilised" OR minus initial OR) for modern cars
running on unleaded gasoline are between 4 and 5 numbers
over 12000 miles; there can be large variations between
individual cars - from 0 to 15 numbers. The literature on

Testing for ORI, like in other deposit related studies, is


characterised by poor repeatability and reproducibility; it is
also labour-intensive and expensive. The difficulties in ORI
testing and in drawing sound and useful conclusions from
such testing have been highlighted by many studies (e.g.
30,61,63,64,79,83,90). It is important to be aware of such
difficulties while assessing the differences between say, the
effect of two additives on ORI.
6.2 EMISSIONS - In general, as cars grow older, there is
an increase, typically of the order of 25% though there could
be a wide variation between individual cars, in engine-out
hydrocarbon (HC),carbon monoxide (CO) and nitrogen

7
Kalghatgi

into an active component of the deposits or oil rather than


adsorption, a physical effect, of the fuel (100).

oxides (NOx) emissions (91-96). Some of the increase is


caused by inlet valve deposits (1,93) and carburetter/port
fuel injector deposits (1) and some presumably by changes
in engine characteristics like valve seat and ignition system
condition. The effects of CCD on emission have been
separately assessed by measuring them before and after
cleaning the combustion chambers of used engines.

Thus there are opposing mechanisms which determine


the effect of CCD on HC emissions and this might explain
why this effect is not clearly established.
CO Emissions - There is one paper reporting significant
effects of CCD on CO emissions (94). In this paper by
Bitting et al, CO emissions were reduced, on average by
around 30%, when CCD were removed . In another paper
Bitting et al (101) show that in a road test, average CO
levels either decreased or increased over the first 9000 miles
depending on which of the two IVD control additives was
used. In another road test where two IVD control additive
packages were compared with the base fuel in 15-car fleets
(45), average CO levels increased for the base-fuel fleet
while they decreased for the cars running on the additised
fuels during a 10,000 mile test. It might be that in this test
the CCD effects on CO were counterbalanced by the effects
of the detergents on IVD and the consequent effects on CO.

NOx Emissions -are reduced when CCD are completely


removed (76,94,97,98,100). In fleet tests, NOx generally
increases as the test progresses and deposits build up
(45,101). Houser (95) found that in three engines, CCD
weights were reduced by using a particular additive rather
than another and this was accompanied by a reduction in
NOx. Bitting et.al have demonstrated this effect in four
vehicles of different models; their results were statistically
significant at the 95% confidence level. The results of
Studzinski et al. (76) suggest that piston-top deposits have a
smaller effect compared to head deposits on NOx; their
modelling study relates the increase in NOx to the increase
in bulk gas temperature brought about by the CCD.

There is no convincing mechanistic explanation as to


why CCD should increase CO. It is possible that because of
the reduction in volumetric efficiency, as deposits build up,
if there is no adequate closed-loop control, the mixture gets
richer as air consumption is reduced. This might lead to an
increase in CO and if this is the mechanism, the effect of
CCD on CO emissions would be negligible if the mixture
strength is maintained by proper closed-loop control.

Hydrocarbon (HC) Emissions- were reduced when


CCD were removed in some studies (92,94,96,97,100) but
Bower et al reported no change (99). Only Bitting et al. (94)
have tried to assess the statistical significance of their
results; their results were not significant at the 90%
confidence level. In another paper, Bitting et al report that as
deposits built up in another road test (on fuels containing
IVD detergents and hence giving low levels of IVD) there
was a trend that HC emissions actually decreased (101).
Spink et al (45) also found that in three 15-car fleets each
running on an unleaded base fuel and two different IVD
control additives, average HC emissions decreased over a
10,000 mile test. Bitting et al speculate that some amount of
CCD may be of benefit for HC emissions (101). Thus
though there are indications that CCD may contribute to an
increase in HC emissions, this effect has not been clearly
established.

CO2 Emissions - Bitting et al report that CO2 emissions


decreased during a road test running on fuels containing IVD
detergents (101) i.e. as primarily CCD levels increased.
Spink et al (45) also report that in three 15-car fleets, CO2
levels decreased on average by 3-4% over 10,000 miles in a
road test. This is to be expected since CCD lead to improved
fuel economy (see below).
6.3 FUEL ECONOMY - Nakamura et.al found, in
eleven different cars, that as deposits grew and octane
requirement increased, fuel economy improved, by 13% in
the most extreme case; it was confirmed that most of this
improvement was due to CCD by removing CCD and
remeasuring fuel consumption (44,54). These cars were
running on unleaded gasoline on the road and fuel economy
was measured in a "10 mode" cycle. They also show (44)
that in another car that ran on a chassis dynamometer at a
constant speed of 60 kmph, fuel economy improved by about
9% in 10000 km, with most of the change occuring in the
first 5000 km. In another experiment (44), they coated the
cylinder head surface of a research engine with teflon and
found that octane requirement increased and fuel economy
improved depending on the thickness of the coating; 15.4%
improvement in fuel economy at a fixed speed and load with
a coating 0.08 mm thick was measured.

In fact, an increase in combustion chamber surface


temperatures, usually brought about by increasing the
coolant temperature, reduces HC emissions (100,103-105)
and since CCD increase surface temperatures they might be
expected to reduce HC emissions. Also, it is possible that
deposits in the crevices could reduce crevice volumes which
are major sources of HC emissions; CCD could reduce HC
emissions through this mechanism (100). On the other hand
CCD can absorb fuel hydrocarbons and release them, largely
unchanged, during the exhaust stroke; this increase might be
directly related to the relative solubility of the fuel in the oil
which might be in the deposit (106). Tsukasaki et al have
attributed the observed increase in formaldehyde emissions
from a car running on methanol to the deposits absorbing the
unvaporised methanol (17). Harpster et al (100) found that
the observed increase in HC emissions with the growth of
deposits was dependent on the fuel structure; of the fuels
tested, MTBE and ethylbenzene showed the largest effects.
They interpret this to mean that the primary mechanism
responsible for the HC increase from CCD is fuel absorption

Kalghatgi et al (37) have measured a reduction in specific


fuel consumption of about 7% at part throttle, with the
reduction being most pronounced at the start of the test, as
deposits built up in an engine test. By removing the deposits

8
Kalghatgi

Most modern engines have flat areas around the rim of


the piston which cause radial gas motion known as "squish"
as they approach similar flat areas on the head during the
compression stroke. At top dead centre, the clearance
between the head and the piston will be minimum in this
squish area. There has been a design trend to reduce squish
clearances, to as low a value as 0.7 mm by some
manufacturers, since it helps to reduce emissions (107). The
squish clearance will be inevitably less than the design value
in some cars because of production tolerances. In addition,
when the engine is cold, the piston is tilted slightly as it rises
(107) particularly at low engine speeds (22). This rocking
motion further reduces the minimum clearance between the
piston and head. This minimum clearance can be eliminated
by combustion chamber deposits and the piston top can
actually hit the head surface under some circumstances. The
clattering sound that results in the frequency range of 1kHz
to 10kHz (22) has been variously described as "carbon rap",
"carbon knock", "deposit induced noise" and "combustion
chamber deposit interference". The problem usually occurs
when a car which suffers from it is started at low ambient
temperatures e.g. after it had been parked outside on a cold
night, and disappears in around five minutes as the car
warms up. However during this time the noise can be quite
unpleasant and during 1992, several car manufacturers
received hundreds of complaints about this problem (107109). No permanent engine damage was observed in cars
which experienced CCDI (110) and the problem disappeared
after prolonged shut-downs (110). Engines with small squish
clearances did not suffer this problem before the summer of
1991 in the U.S. (107). It also appeared that the problem was
more prevalent in the eastern part of the U.S.(108,109).

in stages, they established that almost all this effect was


attributable to head deposits and piston-top deposits had
little effect.
Spink et al (45) report that, in three test fleets, each of
fifteen cars and each fleet running on a different unleaded
fuel and additive combination, on-the-road fuel economy
improved by around 17% over 10000 miles. Almost all of
this improvement, part of which can be attributed to the
reduction in friction losses during "running in", was in the
first 3000 miles. However, in an ECE test cycle, for the same
fleets, the fuel economy improvement was between 2.4%
and 4.5% (45). The initial measurement in this case was
1000 miles after the start of the test and the final
measurement, after 10000 miles. Fuel economy appears to
change rapidly at the start of deposit build-up. If the "clean"
engine measurement is not made in a truly clean engine, the
first measurement of fuel consumption and the apparent
change brought about by the deposits will be smaller. The
fuel economy improvements reported by Spink et al (45)
cannot be clearly attributed to CCD because they did not
make any check-back measurements.
Graiff measured fuel consumption at different speeds and
road load conditions using two bench engines and a vehicle
(87). He reports that removing combustion chamber deposits
caused an increase of about 2% in fuel consumption at all
conditions. A similar figure is quoted by Woodyard (38) for
two bench engines.
Tsutsumi et al (88) measured fuel consumption at wide
open throttle at different speeds in an engine test as part of
their investigations on the effects of surface finish of the
combustion chamber. Their Figure 15 shows that when
combustion chamber deposits built up, brake specific fuel
consumption decreased by about 2% at all speeds upto 6000
RPM.

The problem was exclusively associated with engines with


design squish clearances of 1 mm or less (39,107,108,110).
There is a large variation in squish clearances associated
with production tolerances even in engines of the same
model; engines with smaller clearances ran into CCDI
earlier- some as early as after 1500 km - compared to
engines with larger squish clearances (39). CCDI could
occur even with unadditised base fuel (39,85). Moore shows
that there are significant differences between the effect of
different deposit control additives on CCDI (39). However,
in a fuel with a high aromatic content and high end-point,
additives had little effect on CCDI (84) presumably because
they do not affect CCD formation significantly (see 1.1
above) compared to such a base fuel. Even in cars
susceptible to the problem, fuels like motor alkylate which
are likely to produce low levels of CCD alleviate the CCDI
problem (39,110).

Yonekawa et al have assessed the changes caused by


CCD on the energy balance in a single cylinder engine by
analysing the pressure curves and attribute the improvement
in fuel economy at a fixed speed and load primarily to a
reduction in heat lost to the coolant (67); faster flame
development also plays a role. The heat lost to coolant
constitutes a much larger fraction of fuel energy at low
speeds and loads than at full throttle operation (Ref.112
Ch.12). Hence the fuel economy benefits of CCD are likely
to be less marked at high loads.
Thus CCD generally have a beneficial effect on fuel
economy for some of the same reasons that they increase
NOx and octane requirement, though the extent of this
improvement might be different in different engines and
operating conditions. The problem requires further study.

Moore has described a test protocol using a vibration


signal from the engine to characterise CCDI intensity (39).
He found that there was no correlation between CCDI
intensity and CCD weight. As Megnin (110) suggests, the
maximum deposit thickness is likely to be more relevant
than the average deposit thickness or weight for CCDI. In
fact Iwamoto et al (111) suggest that CCD thickness on the
anti-thrust side in the squish region is important for CCDI.

6.4 COMBUSTION CHAMBER DEPOSIT


INTERFERENCE (CCDI) OR CARBON RAP -The
excellent paper by Moore (39) describes the phenomenon in
detail and what follows borrows heavily from that work.

9
Kalghatgi

The problem could be eliminated quickly - in around 160


km - by using a very high dose, greater than 0.7% in (39), of
a polyether amine detergent (39,110). However, this relief is
temporary and the problem returns quickly. An effective and
permanent cure was to increase the squish clearance (107).

These assessments need to be made on modern engines


with unleaded fuels. However, during the process of
establishing maximum power, maximum power will increase
initially as CCD are removed because of the high
temperatures reached in the engine (53).

The CCDI problem has drawn attention to combustion


chamber deposits and helped generate new interest in
controlling them. It is unlikely that it will continue to be a
problem associated with CCD since there are simple design
options available to engine manufacturers to eliminate it.
This is in contrast to the other effects of CCD like ORI
where there are no such simple engine design solutions.

6.6 OTHER EFFECTS -Pre-ignition and post-ignition


are abnormal combustion phenomena in which the charge is
ignited independent of the spark by hot spots in the
combustion chamber and can cause serious engine damage
(112,113). Build-up of CCD especially those containing lead
or calcium and barium salts from the oil can promote these
phenomena (113). It is unlikely that this problem is of much
relevance in modern engines using unleaded gasolines.
Spark plug fouling can be caused by deposits on the body
of the plug. Greenshields published an authoritative study on
plug fouling caused by deposits from leaded gasolines (114)
on the ceramic core of the spark plug. An occasional modern
problem is plug fouling caused by carbon deposits during the
numerous very short trips taken by a car during its
production and initial delivery (115-117). Moisture
deposition on the carbon layer causes a conductive layer
between the central electrode and the shell causing charge
leakage and failure of the plug-gap to fire (116,117). In
contrast, low electron work function deposits produced by
spark-aider fuel additives on plug electrodes, particularly the
cathode, can improve the energy transfer efficiency of the
spark (118,119). This can lead to faster early flame
development (120) and a dramatic improvement in the
ignition ability of the spark near the ignition limits
(118,119). This results in reduction in misfires,extension of
the lean limit and improvements in driveability (121).

6.5 MAXIMUM ENGINE POWER -There does not


appear to be any publication in this area after 1971 and
many of the studies used leaded fuels.
Cornetti et al.(53) report results from tests using 5
engines of different types of combustion chamber design
after they had built up deposits using leaded fuels. The dirty
engines were then rated at full throttle at 3000 RPM and the
power increased smoothly by about 14% in 80 minutes ; they
attribute this effect to the removal of some CCD. The final,
stabilised power was 6% to 10% lower than for the clean
engine. This is in line with earlier results (70,71). At a given
engine speed, the torque is maximum at a particular ignition
timing - Maximum Brake-Torque (MBT) timing. When
deposits build up, at a given engine speed, MBT timing
occurs earlier, reflecting the faster flame growth and the
torque decreases (70); the change in the power/ignition
timing curve has to be considered while assessing the extent
of the power/torque loss. This loss in peak power is largely
attributable to reductions in volumetric efficiency (53,70).
The power loss depends on the combustion chamber design,
being least (around 4%) for the most compact combustion
chamber (70) ; the loss in maximum power reduced from
5.8% to 3.2% when the compression ratio was increased
from 8 to 12 (71). The deposits in the area "most thoroughly
scrubbed" by the incoming charge have the largest effects on
the power loss; piston top deposits have a relatively smaller
effect (70).

Deposits that form in the ring belt area could cause


piston-ring sticking or ring plugging and ring breakage, all
of which could lead to engine malfunction and short engine
life (122). Such deposits are likely to be more associated
with the lubrication system than the fuel/combustion system.
Thus CCDs affect many aspects of engine performance
and emissions. Some effects, like the improvement in fuel
economy, are beneficial , for more or less the same reasons

TABLE I
Major effects of CCD on engine performance and emissions
Note: These effects are usually observed when CCD are completely removed. The effect of partial changes of say about 20%
in CCD levels are not known.
Parameter
Possible Mechanism (s)
Effect of CCD compared to a
clean engine
Octane Requirement
Volumetric effect (small around 10% of total)
Octane requirement increase (ORI)
Thermal - heat reservoir/insulation
(No clear correlation between CCD
Chemical- absorption of knock precursors ?
weights and ORI)

10
Kalghatgi

HC emissions
(engine out)

Absorption of partly burned fuel by CCD


Reduced crevice volume ( would reduce HC
emissions)
Increased surface temperature (would reduce
HC emissions)

Increase in HC in some cases


but not in others

CO emissions
(engine out)

Not known. Possible that mixture is richened


by reduction in volumetric efficiency. If so, the
effect will not be seen in engines with good
closed-loop control.

Not clear

NOx emissions
(engine out)

Increase in charge temperature because of


heating of charge (see ORI above), reduction in
heat lost to coolant.

Increase in NOx

Fuel economy
and CO2 emissions

Reduction in heat lost to coolant, faster flame


development because of increased charge
temperature leading to higher combustion
efficiency. (same as for NOx).

Improvement in fuel economy.


Reduction in CO2 emissions

Maximum Power

Reduction in volumetric efficiency because of


heating of charge.

Reduction in maximum power


(No publications since 1971)

Carbon Rap

Mechanical interference between piston and


head surfaces in the squish area

deposit growth vary across the combustion chamber. Hence


there is a large variation in thickness, properties, structure
and the effect on engine operation of deposits across the
cylinder.

that CCD might increase octane requirement or NOx


emissions. Table I summarises these effects.
These effects have generally been observed by
completely removing CCD. The effect of smaller changes in
CCD levels of say about 20% are not known. While
assessing the effect of fuels and IVD control additives, some
of which might lead to higher levels of CCD, on engine
performance and emissions, it is important not to concentrate
exclusively on CCD. There are simultaneous effects of fuels
and additives on other factors such as IVD levels which have
effects on engine performance and emissions which might
counteract some of the effects of CCD. What should matter
is the overall effect of a fuel or additive. Fleet tests often
demonstrate an overall benefit in emissions and fuel
economy brought about by good IVD control additives
compared to the base fuel alone (e.g. 45).

Changes in engine operating parameters that tend to


increase surface temperatures reduce deposit formation.
Thus increasing the speed,load and coolant temperature or
making the mixture strength slightly richer than
stoichiometric all reduce deposit growth. Materials of low
conductivity like ceramics can attain high surface
temperatures and on such surfaces, deposit growth is
reduced. Boiling point is the most important fuel or lubricant
property in CCD formation - the higher the boiling point, the
greater the deposit forming tendency. Of the common fuel
components aromatics have been reported to be more prone
to form deposits than olefins and saturates but this might be
primarily because they have higher boiling points. Many fuel
detergent additive packages cause an increase in CCD
formation but there can be a large variation between
different additives and even for the same additive, between
different engines and in different fuels in this regard. Some
additive packages have been shown to be neutral or
beneficial (i.e lower CCD) compared to the base fuel in
some engine tests. Changes in engine oil additives have little
effect on CCD formation.

7. CONCLUSION
In a modern engine with low oil consumption running on
a typical full boiling range fuel, combustion chamber
deposits are primarily formed by the fuel though the
lubricant also contributes to their formation. Deposit growth
is a dynamic and reversible process which at a given time,
reflects the balance between the formation and removal
processes. The single most important parameter that controls
deposit formation is the surface temperature which itself
changes as deposits grow because of their insulating
properties. This and other conditions which might affect

Combustion chamber deposits increase the charge


temperature and flame propagation rate, and reduce
volumetric effeciency and the heat lost to the coolant and
11

Kalghatgi

Increased likelihood. Can be


designed out by increasing squish
clearances

5. Cheng, S.S., " A Physical Mechanism for Deposit


Formation in a Combustion Chamber", SAE Paper No.
941892, also in SAE SP-1054, 1994.

may affect the final phases of combustion through, as yet


undefined, chemical means. This is reflected as an increase
in octane requirement and NOx and a reduction in maximum
power but an improvement in fuel economy and so a
reduction in CO2 emissions. Their effect on HC and CO
emissions is not clear though there is evidence that they
might cause HC emissions to increase in some cases. There
are different opposing mechanisms which determine the
CCD effect on HC emissions and this might explain why this
effect is not unequivocally established. How CCD effect CO
emissions has not been established but if it is through their
effect on mixture strength, engines with effective closedloop control should not suffer any increase in CO because of
CCD. They can lead to carbon rap or combustion chamber
deposit interference which might be avoided by simple
design changes to the combustion chamber and hence is
unlikely to remain a problem in the future. Deposits in
different parts of the combustion chamber affect engine
performance to varying degrees e.g. piston-top deposits have
relatively small effects on ORI, NOx emissions, fuel
economy and power. It is very likely that different engines
will respond to different degrees to the build up of CCD.
The mechanisms by which CCD affect engine performance
are not fully understood. Further work is needed in
establishing and understanding the effects of CCD on
emissions and engine performance.

6. Shore, L.B. and Ockert, K.F. ," Combustion Chamber


Deposits - A Radio Tracer Study", SAE Transactions,
vol.66, pp 285-294, 1958.
7. Newby, W., " Emphasises the effect of boiling point on
deposit formation", SAE Transactions, vol 66, p 294, 1958.
8. Kim,C., Cheng, S.S. and Majorski, S.A. " Engine
Combustion Chamber Deposits : Fuel Effects and
Mechanisms of Formation", SAE Paper No. 912379, 1991.
9. Cheng, S.S., " The Effects of Engine Oils on Intake Valve
Deposits and Combustion Chamber Deposits", SAE Paper
No.932810, 1993.
10. Shipinsky, J.H., p 4-19 in (2) above.
11. Edwards,J.C. and Choate,P.J.," Average Molecular
Structure of Gasoline Engine Combustion Chamber Deposits
Obtained by 13C,31P and 1H Nuclear Magnetic Resonance
Spectroscopy" , SAE paper No.932811, 1993
12. Lauer, J.L. and Friel, P.J., " Some properties of
carbonaceous deposits accumulated in internal combustion
engines", Combust. Flame, vol 4, p 107, 1960

Engine and vehicle tests used in CCD studies are


difficult, long, expensive and are characterised by poor
repeatability and reproducibilty reflecting the complexity of
the problem. This makes it difficult to arrive at definitive
conclusions or to generalise such conclusions to embrace
other fuels and engines and test procedures unless such
experiments are conducted with care. The effects of CCD
on engine performance and emissions have been assessed by
removing the deposits completely. It is not clear to what
degree performance and emissions are affected by smaller
changes, say of 20% , in CCD levels. While assessing the
effects of different fuels or additives on engine performance
through their effect on CCD, it is important to remember that
there would be simultaneous effects on other aspects such as
inlet valve deposits which might have their own,
counterbalancing effects on engine performance.

13. Price, R.J., Wilkinson,J.P.T., Jones,D.A.J. and


Morley,C.," A laboratory simulation and mechanism for the
fuel dependence of SI combustion chamber deposit
formation", SAE Paper No.95
, 1995
14. Nakic,D.J., Assanis,D.N. and White,R.A.," Effect of
elevated piston temperature on Combustion Chamber
Deposit Growth", SAE Paper No.940948, 1994.
.
15. Hayes,T.K., White,R.A. and Peters,J.E.," The In-situ
Measurement of Thermal diffusivity of Combustion
Chamber Deposits in a Spark Ignition Engine", SAE paper
No.920513, 1992

8 . REFERENCES

16. Hafnan,M. and Nishiwaki,K.," Determination of


Thermal Conductivity and Diffusivity of Engine Combustion
Chamber Deposits", JSAE Technical Paper No.9305931,
1993.

1. Kalghatgi, G.T., " Deposits in Gasoline Engines - A


Literature Review", SAE Paper No.902105, 1990
2. "Proceedings of the CRC Workshop on Combustion
Chamber Deposits, Orlando, Florida, November 15-17,
1993.", CRC Inc., 219 Perimeter Center Parkway, Suite 400,
Atlanta, GA 30346, U.S.A.

17 Tsukasaki,M., Yoshida,A., Ito,S. and Nohira,H.," Study


of mileage related formaldehyde emission from methanol
fueled vehicles" SAE Paper No.900705, 1990.

3. Chemistry of Engine Combustion Chamber Deposits


,(Editor: Ebert, L.B.), Plenum Press, 1985.

18. Daly, D.T., Bannon, S.A., Fog, D.A. and Harold, S.M., "
Mechanism of combustion chamber deposit formation", SAE
Paper No.941889, 1994

4. Cheng, S.S. and Kim,C., "Effect of Engine Operating


Parameters on Engine Combustion Chamber Deposits", SAE
Paper No. 902108, 1990

19. Gebhardt, L.A., Lunt, R.S. and Silbernagel, B.G., "


Electron spin resonance studies of internal combustion
engine deposits", pp 145-175 in 3 above.

12
Kalghatgi

36. Adams, K.M. and Baker, R.E., " Effects of combustion


chamber deposits location and composition", pp 19-37 in 3
above.

20. Myers, P.S., Uyehara,O.A. and DeYoung,R., " Fuel


composition and vaporization effects on combustion
chamber deposits", U.S.Dept. of Energy, Report No.
DOE/CS/50020-1, December 1981.

37. Kalghatgi, G.T., McDonald,C.R. and Hopwood,A.B., "


An experimental study of combustion chamber deposits and
their effects in a spark-ignition engine", SAE Paper
No.950680, 1995.

21. Keller,H. and Shimkoski,D.A. ,p 6-45 in (2) above


22. Nagao,M, Kaneko,T., Omata,T., Iwamoto,S.,
Ohmori,H. and Matsuno, S.," Mechanism of combustion
chamber deposit interference and effects of gasoline
additives on CCD formation", SAE Paper No.950741 also
in SAE SP-1095, 1995

38. Woodyard, M.E., p 9-61 in (2) above

23. Owens,J., p 7-55 in (2) above

39. Moore, S.M., " Combustion Chamber Deposit


Interference Effects in Late Model Vehicles", SAE Paper
No. 940385, 1994.

24. Wagner,R, p 7-85 in (2) above

40. Shoppe,D. and Friedland,G., p 9-1 in (2) above.

25. Peyla,R.J, p 7-109 in (2) above


26. Koehler,D, p 7-141 in (2) above

41. Bachman, H.E. and Prestridge, E.B.," The use of


Combustion Deposit Analysis for Studying Lubricant
Induced ORI", SAE Paper No. 750938, 1975.

27. Jackson, M.M. and Pocinki, S.B., " Effects of fuel and
additives on combustion chamber deposits", SAE Paper
No.941890, also in SAE SP-1054, 1994.

42. Ebert, L.B., Davis, W.H., Mills, D.R.; Dannerlein, D.I.


and Rose, D.L.; Melchior, M.T., " The chemistry of internal
combustion engine deposits - Parts I,II and III", in 3 above.

28. Lepperhoff,G. and Houben,M.," Mechanisms of Deposit


Formation in Internal Combustion Engines and Heat
Exchangers", SAE Paper No.931032, 1993

43. McDonald, C.R. and Hopwood, B., Personal


Communication.

29. Anderson,C.L. and Wood,B.S.," Gasification of Porous


Combustion Chamber Deposits in a Spark Ignition Engine",
SAE Paper No.930773, 1993.

44. Nakamura,Y.,Yonekawa,Y. and Okamoto,N.," The


Effect of Combustion Chamber Deposits on Octane
Requirement Increase and Fuel Economy" pp 199-211 in 3
above.

30. Peyla, R.J., " Motor Gasoline and Deposit Control


Additives - A Challenge for the 90s", Paper No.FL-91-118,
presented at the 1991 NPRA National Fuels and Lubricants
Meeting, Houston, 1991.

45. Spink, C.D., Barraud, P.G. and Morris,G.E.L., " A


critical road test evaluation of two high-performance
gasoline additive packages in a fleet of modern European
and Japanese vehicles", SAE Paper No. 912393, 1991.

31. Finlay,I.C., Harris,D.,Boam,D.J. and Parks,B.I., "Factors


Influencing Combustion Chamber Wall Temperature in a
Liquid Cooled Automobile Spark Ignition Engine.", Proc.
Instn. Mech. Engrs., vol 199, No D3, p 207, 1985.

46. Harder, R.F. and Anderson, C.L., " Investigation of


combustion chamber deposit thermal behaviour utilising
optical radiation measurements in a fired engine",
Combust.Sci.Tech., vol 60, pp 423-439, 1988

32.Hayes, T.K., White, R.A. and Peters, J.E., " Combustion


Chamber Temperature and Instantaneous Local Heat Flux
Measurements in a Spark Ignition Engine", SAE Paper No.
930217, 1993.

47. Hayes,T.K., Peters,J.E. and White,R.A., " The thermal


properties of engine deposits and their effect on combustion
chamber heat transfer", (C382/011), Proceedings of the 2nd
Int.Conf.on New Developments in Powertrain and Chassis
Engineering, Strassbourg, June 1989, I.Mech.E, 1989.

33. French, C.C.J. and Atkins,K.A., " Thermal Loading of a


Petrol Engine", Proc. Instn. Mech. Engrs., vol 187, 49/73,
pp 561-573, 1973.

48. Kelemann, S.R. and Maxey, C.T., p 7-125 in (2) above.

34. Megnin,M.K. and Choate, P.J., " Combustion Chamber


Deposit Measurement Technique", SAE Paper No. 940346,
1994.

49. Choate, P.J. and Edwards, J.C., " Relationships between


combustion chamber deposits, fuel composition and
combustion chamber deposit structure", SAE Paper No.
932812, 1993.

35. Keller, C.T. and Corkwell, K.C., " Honda Generators


Used to Evaluate Fuels and Additive Effects on Combustion
Chamber Deposits", SAE Paper No. 940347, 1994.

50. Takei, Y., Uehara, T., Hoshi, H. and Okada, M.," Effects
of gasoline and gasoline detergents on combustion chamber
deposit formation", SAE Paper No.941893, 1994.

13
Kalghatgi

67. Yonekawa, Y., Kokubo, K., Nakamura, Y. and


Okamoto, N., " The study of combustion chamber deposit
(part 5): The role of combustion chamber deposits in fuel
economy", J.Japan Petroleum Institute, vol 25, pp 177-182,
1982.

51. Dumont, L.F., " Possible mechanisms by which


combustion chamber deposits accumulate and influence
knock", SAE Quart.Trans., vol 5, pp 565-576, 1951.
52. Duckworth J.B., " Effects of Combustion Chamber
Deposits on Octane Requirement and Engine Power
Output", SAE Quart. Trans., vol 5, pp 577 -583, 1951.
53. Cornetti,G., Liguori, V. and Amendola, L., " Power loss
due to combustion chamber deposits", Journal of
Automotive Engineering, pp8-14, June 1971.

68. Harrow, G.A. and Orman, P.L., " Study of flame


propagation and cyclic dispersion in a spark ignition
engine", ASAE Symposium on Combustion in Engines, July
1965. Advances in Automobile Engineering, pp 3-27,
Pergamon Press, 1966.

54. Yonekawa,Y., Nakamura,Y. and Okamoto,N.," The


study of combustion chamber deposit (Part 4) - Octane
number requirement and fuel economy with deposit build
up", J.Japan Petrol. Inst., vol 25 (3), pp 173-176, 1982.

69. Bradish, J.P., Myers, P.S. and Uyehara, O.A., " Effects
of deposit properties on volumetric efficiency, heat transfer
and preignition in internal combustion engines", SAE Paper
No. 660130, 1966.

55. Megnin, M.K. and Furman, J.B., " Gasoline effects on


octane requirement increase and combustion chamber
deposits", SAE Paper No. 92258, 1992.

70. Gibson H.J., Hall, C.A. and Huffman, A.E., "


Combustion chamber deposition and power loss", SAE
Quart. Trans., vol 6, no.4, p 595, 1952.

56. Nishizaki, T., Maeda, Y., Date, K and Maeda, T., " The
effect of fuel composition and fuel additives on intake
system detergency of Japanese automobile engines", SAE
Paper No. 790203, 1979.

71. McDuffie, A. and Mitzelfeld, T.H., Discussion of Ref.50


above, SAE Quart. Trans., vol 6, no.4, p 604, 1952.
72. De Gregoria, A.J., " A theoretical study of engine
deposit and its effect on octane requirement using an engine
simulation", SAE Paper No. 820072, 1982.

57. Gibbs, L.M., p 7-1 in (2) above.

73. Nishiwaki, K., " Unsteady Thermal Behaviour of Engine


Combustion Deposits", SAE Paper No.881225, 1988.

58. Cunningham, L., p 7-15 in (2) above.


59. Swaynos, D. and More, I., p 8-21 in (2) above.

74. Valtadoros, T.H., Wong, V.W. and Heywood, J.B., "


Fuel additive effects on deposit build-up and engine
operating characteristics", Symposium on fuel
composition/deposit formation tendencies, Division of
Petroleum Chemistry, ACS, Atlanta, p 66, April 1991.

60. Chapman, J.L., Williamson, J. and Preston, W.H., "


Deposits in internal combustion engines" in Deposition from
Combustion Gases (Ed. Jones, A.R.), pp 113-127, IOP
Publishing Ltd., Bristol, U.K., 1989.

75. Bussovansky, S., Heywood, J.B. and Keck, J.C., "


Predicting the effects of air and coolant temperature,
deposits, spark timing and speed on knock in spark ignition
engines", SAE Paper No. 922324, 1992.

61. Keller, B.D., Meguerin, G.H., Tracy, C.B. and Smith,


J.B., " ORI of today's vehicles", SAE Paper No.760195,
1976.
62. Kunc, J.F., " Effect of lubricating oil on octane
requirement increase", SAE Quart. Trans., vol 5, p 582,
1951.

76. Studzinsky, W.M., Liiva, P.M., Choate, P.J., Acker,


W.P., Smooke, M., Brezinsky, K., Litzinger, T. and Bower,
S., " A Computational and Experimental Study of
Combustion Chamber Deposit Effects on NOx Emissions",
SAE Paper No. 932815, 1993.

63. Alquist, H.E., Holman, G.E. and Wimmer,D.B., " Some


observations of factors affecting ORI", SAE Paper No.
750932, 1975.

77. Niles, H.T., McConnell, R.J., Roberts, M.A. and


Saillant, R., " Establishment of ORI characteristics as a
function of selected fuels and engine families", SAE Paper
No. 750451, 1975.

64. Benson, J.D.," Some factors which affect octane


requirement increase", SAE Paper No. 750933, 1975.
65. McNab, J.B., Moody, L.E. and Hakala, N.V., " Effect of
lubricant composition on combustion chamber deposits",
SAE Transactions, vol. 62, pp 228-242, 1954.

78. Saillant, R.B., Pedrys, F.J. and Kidder, H.E., " More
data on ORI variables", SAE Paper No. 760196, 1976.
79. Fuentes-Afflick, P., p 9-29 in (2) above.

66. Warren, J.," Combustion Chamber Deposits and Octane


Number Requirement" SAE Transactions, vol 62,
pp583-594, 1954.

80. Peyla, R.J., p 7-109 in (2) above.

14
Kalghatgi

81. Schreyer, P., Starke, K., Thomas, J. and Crema, S., "
Effect of multifunctional fuel additives on octane number
requirement of internal combustion engines", SAE Paper
No.932813, 1993.

98. Huls, T.A. and Nickol, H.A., " Influence of engine


variables on exhaust oxides of nitrogen concentrations from
a multicylinder engine", SAE Paper No. 670482, 1967.
99. Bower, S.L., Litzinger, T.A. and Frottier, V., " The
effects of fuel composition and engine deposits on emissions
from a spark ignition engine", SAE Paper No.932707, 1993.

82. Mitchell, K., p 6-77 in (2) above


83. Wittenbrock, D.G., p 6-89 in (2) above.

100. Harpster, M.O., Matas, S.E., Fry, J.H. and


Litzinger,T.A., " An experimental study of fuel composition
and combustion chamber deposit effects on emissions from a
spark ignition engine", SAE Paper No. 950740, Also in SAE
SP-1095, 1995

84. Cunningham, L., p 7-15 in (2) above.


85. Avery, T., Axelrod,J and Carey, J., p 8-43 in (2) above.
86. Haury, E.J., p 9-41 in (2) above.

101. Bitting,W.H., Firmstone,G.P. and Keller,C.T., " A fleet


test of two additive technologies comparing their effects on
tailpipe emissions", SAE Paper No.950745, Also in SAE
SP-1095, 1995.

87. Graiff, L.B., " Some new aspects of deposit effects on


engine octane requirement increase and fuel economy", SAE
Paper No. 790938, 1979.
88. Tsutsumi, Y., Nomura, K. and Nakamura, N., " Effect of
mirror-finished combustion chamber on heat loss", SAE
Paper No. 902141, 1990.

102. Wentworth, J.T., " More on origins of exhaust


hydrocarbons - effect of zero oil consumption, deposit
location and surface roughness", SAE Paper No.720939,
1972.

89. Eng, K.D., Carlson, C.A., Haydn, T.E. and Sung, R.L., "
Engine test procedures to evaluate octane requirement
increase and intake system cleanliness", SAE Paper No.
892122, also in SP-794, Fuel and Induction System
Deposits, SAE, 1989.

103. Wentworth, J.T., " Effect of combustion chamber


surface temperature on exhaust hydrocarbon concentration",
SAE Paper No.710587, 1971
104. Myers, J.P. and Alkidas, A.C., " Effects of
combustion-chamber surface temperature on the exhaust
emissions of a single -cylinder spark-ignition engine", SAE
Paper No. 780642, 1978.

90. Sengers, H.P.M. and ter Rele, R.R.J., " Improved


precision in ORI field and engine tests", Paper No.
CEC/93/EF18, Fourth international symposium on the
performance evaluation of automotive fuels and lubricants,
Birmingham, 1993.

105. Russ,S.G., Kaiser,E.W., Siegl,W.O., Podsiadlik,D.H.


and Barrett, K.M., " Compression ratio and coolant
temperature effects on HC emissions from a spark-ignition
engine", SAE Paper No.950163, also in SAE SP-1094,
1995.

91. Gagliardi, J.C. and Ghanam, F.E., " Effect of tetraethyl


lead concentration on exhaust emissions in customer type
vehicle operation", SAE Paper No. 690015,1969.
92. Leikkanen, H.E. and Beckman, E.W.," The effect of
leaded and unleaded gasolines on exhaust emissions as
influenced by combustion chamber deposits", SAE Paper
No.710843, 1971.

106 Adamczyck, A.A. and Kach, R.A., " The effect of


engine deposit layers on hydrocarbon emissions from closed
vessel combustion", Combust.Sci.Tech., vol 47, pp 193-212,
1986.

93. Houser, K.R. and Crosby, T.A., " The impact of intake
valve deposits on exhaust emissions", SAE Paper No.
92259, 1992.

107. Russ, M.J., p 4-31 in (2) above.


108. Bannon, S.A., p 4-1 in (2) above.

94. Bitting, W.H., Firmstone, G.P. and Keller, C.T., "


Effects of combustion chamber deposits on tailipipe
emissions", SAE Paper No. 940345, 1994.

109. Carlson, C.A., p 4-7 in (2) above.


110. Megnin, M.K., p 8-57 in (2) above.

95. Houser, K., p 9-85 in (2) above.


111. Iwamoto, S. et al, " Carbon knock analysis concerning
combustion chamber deposit", JSAE 9433290, 1994.

96. Pahnke, A.J. and Conte, J.F., " Effect of combustion


chamber deposits and driving conditions on vehicle exhaust
emissions", SAE Paper No.690017, 1969.

112. Heywood, J.B., " Internal combustion engine


fundamentals", McGraw Hill Book Co., 1988.

97. Gagliardi, J.C., " The effects of fuel anti-knock


compounds and deposits on exhaust emissions", SAE Paper
No.670128, 1967.

113. Wheeler, R.W., "Abnormal combustion effects on


economy" in " Fuel economy in road vehicles powered by

15
Kalghatgi

spark ignition engines", (Ed.s, Hilliard, J.C. and Springer,


G.S.), Ch.6, Plenum Press, New York, 1984.

119. Kalghatgi, G.T., " Flame initiation and development


from glow discharges - Effect of electrode deposits",
Combustion and Flame, vol 77, p 321, 1989.

114. Greenshields, R.J., "Spark plug fouling tendencies",


SAE Trans., vol 61, p 3, 1953.

120. Kalghatgi, G.T., " Improvement in early flame


development in a spark ignition engine brought about by a
spark-aider fuel additive", Combust. Sci. Tech., vol 52, p
427, 1987.

115. Kimbara, Y., Noguchi, Y. and Ishiguro, T., " Study on


spark plug carbon fouling", SAE Paper No. 800832, 1980.
116. Quader, A.," Spark plug fouling: A quick engine test",
SAE Paper No. 920006, 1992.

121. Kalghatgi, G.T., " Effect of a spark-aider fuel additive


on the misfire characteristics of a spark ignition engine",
Comust. Sci. Tech., vol 62, p 1, 1989.

117. Collings, N., Dinsdale, S. and Hands, T., " Plug fouling
investigations on a running engine - an application of a novel
multi-purpose diagnostic system based on the spark plug",
SAE Paper No.912318, 1991.

122. Kipp, K.L., Ingamells, J.C., Richardson, W.L. and


Davis, C.E., " Ability of gasoline additives to clean engines
and reduce exhaust emissions", SAE Paper No. 700456,
1970.

118. Kalghatgi, G.T., " Improvements in the ignition ability


of glow discharges brought about by deposits of potassium
sulphate on the electrodes", Paper No. C50/80, International
Conference on combustion in engines- Technology and
applications, I. Mech.E., London, 1988.

16
Kalghatgi

S-ar putea să vă placă și