ArticlePDF Available

Utilizing low airflow strategies, including cylinder deactivation, to improve fuel efficiency and aftertreatment thermal management

Authors:

Abstract and Figures

Approximately 30% of the fuel consumed during typical heavy-duty vehicle operation occurs at elevated speeds with low-to-moderate loads below 6.5 bar brake mean effective pressure. The fuel economy and aftertreatment thermal management of the engine at these conditions can be improved using conventional means as well as cylinder deactivation and intake valve closure modulation. Airflow reductions result in higher exhaust gas temperatures, which are beneficial for aftertreatment thermal management, and reduced pumping work, which improves fuel efficiency. Airflow reductions can be achieved through a reduction of displaced cylinder volume by using cylinder deactivation and through reduction of volumetric efficiency by using intake valve closure modulation. This paper shows that, depending on load, cylinder deactivation and intake valve closure modulation can be used to reduce the fuel consumption between 5% and 25%, increase the rate of warm-up of aftertreatment, maintain higher temperatures, or achieve active diesel particulate filter regeneration without requiring dosing of the diesel oxidation catalyst.
Content may be subject to copyright.
Standard Article
International J of Engine Research
2017, Vol. 18(10) 1005–1016
ÓIMechE 2017
Reprints and permissions:
sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/1468087417695897
journals.sagepub.com/home/jer
Utilizing low airflow strategies,
including cylinder deactivation, to
improve fuel efficiency and
aftertreatment thermal management
Aswin K Ramesh
1
,GregoryMShaver
1
, Cody M Allen
1
, Soumya Nayyar
2
,
Dheeraj B Gosala
1
,DinaCaicedoParra
2
, Edward Koeberlein
2
,James
McCarthy
3
and Doug Nielsen
3
Abstract
Approximately 30% of the fuel consumed during typical heavy-duty vehicle operation occurs at elevated speeds with
low-to-moderate loads below 6.5 bar brake mean effective pressure. The fuel economy and aftertreatment thermal man-
agement of the engine at these conditions can be improved using conventional means as well as cylinder deactivation
and intake valve closure modulation. Airflow reductions result in higher exhaust gas temperatures, which are beneficial
for aftertreatment thermal management, and reduced pumping work, which improves fuel efficiency. Airflow reductions
can be achieved through a reduction of displaced cylinder volume by using cylinder deactivation and through reduction
of volumetric efficiency by using intake valve closure modulation. This paper shows that, depending on load, cylinder
deactivation and intake valve closure modulation can be used to reduce the fuel consumption between 5% and 25%,
increase the rate of warm-up of aftertreatment, maintain higher temperatures, or achieve active diesel particulate filter
regeneration without requiring dosing of the diesel oxidation catalyst.
Keywords
Heavy-duty federal test procedure, fuel efficiency, cylinder deactivation, variable valve actuation, aftertreatment thermal
management
Date received: 11 September 2016; accepted: 24 January 2017
Introduction
Heavy-duty vehicle fuel consumption is expected to
double by 2050.
1
The majority of heavy-duty vehicles
incorporate diesel engines, given their efficiency bene-
fits compared with spark-ignited engines. However, die-
sel engines emit air pollutants, including particulate
matter, unburnt hydrocarbons, and oxides of nitrogen
(NOx). To reduce the production of these harmful
gases, several on-engine strategies, including multiple
fuel injection, high injection pressure, late fuel injec-
tion, and exhaust gas recirculation,
2
have been incorpo-
rated in modern diesel engines. These strategies are
unable to meet the regulations at all operating points.
As a result, complex aftertreatment systems are coupled
with the diesel engine to meet tailpipe emission regula-
tion limits.
Typical aftertreatment systems incorporate selective
catalytic reduction to reduce NOx emissions, a diesel
oxidation catalyst to reduce hydrocarbons and carbon
monoxide, and a diesel particulate filter to reduce parti-
culate matter emissions. Selective catalytic reduction is
effective in converting NOx into N
2
and H
2
O,
3
pro-
vided the catalyst temperature is between approxi-
mately 250 °C and 450 °C.
4–6
Urea is injected upstream
of the selective catalytic reduction system, decomposing
into ammonia and carbon dioxide. This process
requires selective catalytic reduction inlet temperatures
above approximately 200 °C to avoid build up of solid
deposits on the catalyst bed.
1
The selective catalytic
reduction efficiency is maximized for selective catalytic
1
Purdue University, USA
2
Cummins, Inc., USA
3
Eaton Valvetrain Engineering, USA
Corresponding author:
Gregory M Shaver, Ray W. Herrick Laboratories, Purdue University, 177
S Russel Street, West Lafayette, IN 47906, USA.
Email: gshaver@purdue.edu
reduction catalyst bed temperatures from 300 °Cto
450 °C. To keep the aftertreatment warm in a robust
manner, turbine outlet temperatures between 350 °C
and 500 °C are desirable. The diesel particulate filter is
regularly regenerated in a passive manner by oxidizing
the collected particulate matter, provided the turbine
outlet temperature is approximately 250 °Cto300°Cin
the presence of NO
2
.
7,8
Active regeneration of the diesel
particulate filter requires temperatures in excess of 450
°C. The diesel oxidation catalyst must reach approxi-
mately 200 °C to perform effective oxidation of carbon
monoxide and hydrocarbons. Once activated, the diesel
oxidation catalyst oxidizes hydrocarbons, resulting in
increased aftertreatment temperatures, while converting
NO to NO
2
, both of which enable passive reactions in
the diesel particulate filter at temperatures above
250 °C.
9,10
In short, the aforementioned aftertreatment
components work effectively in reducing emissions
when operated at the correct temperatures; however,
during cold start and low-load engine operation, the
exhaust gas temperature is too low to keep the after-
treatment catalysts at effective operating temperatures.
Thermal management strategies designed to increase
aftertreatment component temperatures are needed for
efficient aftertreatment operation over a wide range of
engine operating conditions.
11,12
Research continues on methods for improving after-
treatment thermal management by means of increased
diesel engine exhaust temperatures.
13,14
One technology
currently being studied to accomplish this goal is vari-
able valve actuation.
15–19
To assess the impact of advanced engine system stra-
tegies on aftertreatment thermal management, it is
helpful to compare strategies during operation of stan-
dardized test procedures. The heavy-duty federal test
procedure (HD-FTP) is used for regulatory emission
testing of heavy-duty, on-road engines in the USA.
This test was developed by taking actual operating data
over a variety of heavy-duty truck and bus driving pat-
terns on roads and expressways.
1
Figure 1 shows the
speed and normalized brake mean effective pressure
(BMEP) profile through time for the HD-FTP.
Figure 2 shows the fuel consumption over a HD-FTP
mapped around eight steady-state engine operating
regions. The number displayed next to a bubble sig-
nifies the percentage of fuel consumed at these operat-
ing conditions. In this study, the benefits of using
variable valve actuation for improving fuel efficiency
and aftertreatment thermal management are studied
for the cross-hatched bubbles at an engine speed of
2200 r/min, for two reasons.
1. More than 30% of the fuel used during the HD-
FTP is consumed in these operating regions.
2. As will be shown, there is an opportunity to
increase fuel efficiency and exhaust temperatures
at these conditions by reducing airflow via both
conventional engine actuators and variable valve
actuation.
During these high-speed low-load conditions, the
air-to-fuel ratio is elevated, given the reduced fueling
required at low loads and the engine ‘‘over breathing’’
at low loads, and to prepare for a sudden increase in
fueling resulting from a commanded increase in desired
torque and power. There are several possible strategies
for improving the engine torque response to acceptable
levels when air-to-fuel ratios are reduced at high-speed,
low-load operating conditions, including:
(a) early exhaust valve opening;
20
(b) internal exhaust gas recirculation via combustion
gas trapping or re-induction;
(c) turbocharger electrification;
21
(d) supercharging;
22,23
Figure 1. Speed and brake mean effective pressure versus time
for the heavy-duty federal test procedure. Highlighted sections
shows high-speed, low-load conditions, as cross-hatched in
Figure 2.
BMEP: brake mean effective pressure; LAFY: Los Angeles Freeway;
LANF: Los Angeles Non Freeway; NYNF: New York Non Freeway.
Figure 2. Fuel consumption distribution over heavy-duty
federal test procedure, mapped around eight operating
conditions.
BMEP: brake mean effective pressure; HDFTP: heavy-duty federal test
procedure.
1006 International J of Engine Research 18(10)
(e) powertrain hybridization;
24,25
(f) availability of look-ahead information,
26
through
vehicle data connectivity with other vehicles or
the Cloud, to allow anticipation of an upcoming
transient.
These methods are not the subject matter discussed in
detail in this paper. Instead, this paper outlines strate-
gies for achieving and the benefits of low air-to-fuel
ratio operation at high-speed, low-load conditions. Air-
to-fuel ratio reduction strategies considered in this
paper include:
(a) ‘‘opening up’’ the variable geometry turbine turbo-
charger (VGT);
(b) reducing the displaced volume through cylinder
deactivation;
(c) reducing volumetric efficiency via late intake valve
closure (IVC);
(d) some combination of these strategies.
Previous research by several of the authors has
demonstrated that cylinder deactivation is an effective
method for increasing exhaust temperatures and fuel
efficiency when the diesel engine is idle,
19
and for
enabling active diesel particulate filter regeneration
without requiring diesel oxidation catalyst fuel dosing
during highway cruise conditions.
17
The effect of early
and late IVC, including a notable increase in the
exhaust gas temperature is outlined by Ojeda
27
and
Lombard and le Forestier.
28
Prior efforts have also
shown that late IVC can be used to increase exhaust
temperature.
18
The first results section of this paper focuses on eval-
uating the fuel economy benefits of the specified strate-
gies at elevated speed (2200 r/min) and low-load
operating conditions (1.3–6.3 bar BMEP). The second
results section focuses on which strategy works best
when the main goal is to keep the aftertreatment tem-
perature elevated. The third results section focuses on
which strategies allow the fastest warm-up of the after-
treatment system. Finally, the fourth results section
focuses on which strategy is preferable for enabling
active diesel particulate filter regeneration.
Experimental setup
The system under observation in this study is a camless
six-cylinder Cummins diesel engine outfitted with an
electro-hydraulic variable valve actuation system, high-
pressure cooled exhaust gas recirculation, a sliding
nozzle-type VGT, an air-to-water charge air cooler,
and a common rail fuel injection system. A schematic
of the engine architecture is presented in Figure 3. The
fresh intake air flows through the laminar flow element
into the compressor and is then cooled in the charge air
cooler. The exhaust flows either through the turbine of
the VGT to the exhaust pipe or into the exhaust gas
recirculation system. The exhaust temperature is mea-
sured at the outlet of the turbine and is referred to in
this paper as the turbine outlet temperature. Two
Kistler 6067 and four AVL QC34C in-cylinder pressure
transducers are used in tandem with an AVL 365C
crankshaft position encoder together with an AVL 621
Indicom module for high-speed in-cylinder pressure
data acquisition. Laboratory-grade fuel flow measure-
ment is used to measure the fuel consumption.
Both the intake and exhaust valve pairs for each of
the six cylinders are actuated by the variable valve
actuation system, such that it has a total of twelve
actuators. The actuators use position feedback for
closed-loop control, enabling the desired cylinder-inde-
pendent, cycle-to-cycle operation of the system.
Figure 4 presents a schematic of the variable valve
actuation system. The valve profiles are generated in
Simulink and the dSPACE hardware is used to transmit
Figure 3. Camless Cummins multicylinder engine testbed at
Purdue University.
CAC: charge air cooler; EGR: exhaust gas circulation; LFE: laminar flow
element; VGT: variable geometry turbine turbocharger.
Figure 4. Purdue variable valve actuation system.
LVDT: linear variable differential transformer.
Ramesh et al. 1007
voltage feedback to the servo valves via the controller
and amplifier. The servo valves shuttle high-pressure
hydraulic oil to one side of the valve actuators. These
actuators push on the valve pairs through a valve
bridge to open them. The return forces from the valve
springs close the valves as the actuators retract.
The valve profiles have six key features that describe
the valve lift shape:
(a) Intake valve opening (IVO);
(b) Peak intake valve lift (IVL);
(c) Intake valve closure (IVC);
(d) Exhaust valve opening (EVO);
(e) Peak exhaust valve lift (EVL);
(f) Exhaust valve closure (EVC).
Using these inputs, valve strategies including IVC
modulation (early IVC or late IVC) can be realized on
the engine. Figure 5 illustrates the late IVC valve strat-
egy, wherein the IVC timing is delayed from the nom-
inal position.
Cylinder deactivation is achieved by deactivating the
valve motions and fuel injections for either two or three
cylinders, resulting in four-cylinder or three-cylinder
modes, respectively.
A butterfly valve is used in the exhaust pipe to simu-
late the back pressure that would be caused by a typical
aftertreatment system. The aforementioned equipment,
full access to adjustment of parameters in the engine
control module, and additional temperature and pres-
sure sensors are integrated using a dSPACE system.
The dSPACE system simultaneously controls the vari-
able valve actuation system, sends and receives data
with the engine control module, and samples all of the
external measurement channels.
Methodology and nomenclature
A total of five loads (1.3, 2.6, 3.8, 5.1, and 6.4 bar
BMEP) were selected to study the fuel economy and
aftertreatment thermal management benefits of reduced
air flow conditions at 2200 r/min. All experimental data
shown and discussed in the following sections were sub-
ject to strict emissions and mechanical constraints. The
mechanical constraints are shown in Table 1. The parti-
culate matter, NOx, and unburnt hydrocarbon limits
for each operating load are the same as the engine
achieves with stock calibration and conventional valve
motions. For the ‘‘open VGT’’ and ‘‘open VGT +
IVC modulation’’ strategies, the start of injection, and
rail pressure were modulated to screen for operating
strategies to achieve improved efficiency or thermal
management responses. During cylinder deactivation,
the VGT position, start of injection, and rail pressure
were modulated to achieve improved efficiency or ther-
mal management.
The brake thermal efficiency of the engine includes
contributions from the closed-cycle efficiency, open-
cycle efficiency, and mechanical efficiency, as
BTE = hclosed cycle 3hopen cycle 3hmechanical ð1Þ
Closed-cycle efficiency is affected by combustion com-
pleteness, piston expansion work, and in-cylinder heat
transfer. The open-cycle efficiency quantifies the effec-
tiveness of the gas exchange and is affected by the tur-
bine and compressor efficiency, and pressure
differences between the intake and exhaust manifold.
The mechanical efficiency captures losses from friction
and parasitic loads. For a more detailed explanation of
the cycle efficiency analysis strategy refer to Stanton.
1
Experimental results
The benefits of the low air flow strategies described in
the following four sections include:
(a) improving fuel efficiency, enabled through
increased open-cycle efficiency via reduced pump-
ing work;
(b) maintaining elevated aftertreatment component
temperatures through elevated engine exhaust tem-
perature as a result of reduced airflow operation;
Figure 5. Late intake valve closure modulation valve strategy
compared with nominal valve positions. The black line
represents the nominal intake valve position.
IVC: intake valve closure; LIVC: late intake valve closure; TDC: top dead
center.
Table 1. Mechanical constraints.
Mechanical parameter Unit Limit
Turbine inlet temperature °C 760
Compressor outlet temperature °C 230
Turbo speed kr/min 193
Peak cylinder pressure bar 172
Exhaust manifold pressure kPa 500
In-cylinder pressure rise rate bar/ms 100
1008 International J of Engine Research 18(10)
(c) increasing the rate of warm-up of the aftertreat-
ment components through elevated engine exhaust
gas temperatures;
(d) enabling active diesel particulate filter regenera-
tion through elevated engine exhaust gas
temperatures.
Fuel efficiency
Figure 6 summarizes the fuel efficiency benefits, over
the BMEP range, of using cylinder deactivation, IVC
modulation, or opening up the VGT. The results are
normalized with respect to the stock calibration, and
show that fuel savings between 5% and 30% are possi-
ble, depending on load (e.g., BMEP). Cylinder deacti-
vation leads to a significant (25%) fuel consumption
reduction at loads less than 2.5 bar. This is primarily a
result of a 35% increase in the open-cycle efficiency,
shown in Figure 7, achieved through a reduction in
pumping work.
Pumping work is lower because airflow is lower, as
shown in Figure 8. Airflow is lower as a result of a
reduction in displaced volume via cylinder deactivation.
Airflow is lower for the ‘‘open VGT’’ strategy, owing
to the reduction is boost pressure via VGT opening. In
general, the three-cylinder mode is more fuel-efficient
than the four-cylinder mode, when it is feasible to deac-
tivate three cylinders. The fuel efficiency benefit relative
to the stock calibration decreases as the load increases,
owing to reductions in the open-cycle efficiency
benefits.
At 2.5 bar BMEP, the increase in open-cycle effi-
ciency for cylinder deactivation compared with ‘‘open
VGT’’ operation is negated by a decrease in closed-
cycle efficiency during cylinder deactivation, as shown
in Figure 9. As a result, the brake-specific fuel con-
sumption is similar for the cylinder deactivation and
‘‘open VGT’’ strategies.
At loads greater than, and equal to, 2.5 bar, cylinder
deactivation has a slightly higher fuel consumption than
‘‘open VGT’’ operation as a result of decrease in closed-
cycle efficiency via:
(a) higher in-cylinder heat transfer, as shown in
Figure 10, and longer heat release caused by
higher per-cylinder fueling in active cylinders, as
shown in Figure 11;
(b) later injection timings required to maintain
engine-out NOx levels, as shown in Figure 11.
Cylinder deactivation was not implemented at loads
above 5.1 bar BMEP, as there is not enough oxygen
Figure 7. Comparison of open-cycle efficiency versus brake
mean effective pressure at 2200 r/min. The open-cycle efficiency
is highest for the three-cylinder mode at these low-load
conditions. Open-cycle efficiency quantifies the efficiency of the
gas exchange, which is impacted by turbine and compressor
efficiency and pressure differences between the intake and
exhaust manifolds. The benefit seen in open-cycle efficiency
decreases as the load increases, as there is a lower reduction in
airflow with higher loads.
BMEP: brake mean effective pressure; IVC: intake valve closure; VGT:
variable geometry turbine turbocharger.
Figure 8. Comparison of normalized air flow versus load at
2200 r/min. There is a significant decrease in air flow when
cylinders are deactivated or when the variable geometry turbine
turbocharger is opened. There is a minor decrease in air flow
when intake valve closure modulation is used, as the volumetric
efficiency is reduced.
BMEP: brake mean effective pressure; IVC: intake valve closure; VGT:
variable geometry turbine turbocharger.
Figure 6. Comparison of brake-specific fuel consumption
versus brake mean effective pressure at 2200 r/min. The lower
the brake-specific fuel consumption, the higher the brake
thermal efficiency.
BMEP: brake mean effective pressure; BSFC: brake-specific fuel
consumption; IVC: intake valve closure; VGT: variable geometry turbine
turbocharger.
Ramesh et al. 1009
available in cylinder deactivation mode to keep particu-
late matter within the constraints.
The ‘‘open VGT + IVC modulation’’ strategy has
a slightly lower fuel consumption than the ‘‘open
VGT’’ strategy, given better open-cycle efficiency and
closed-cycle efficiency at 2.5 bar BMEP. The open-
cycle efficiency increases for the ‘‘open VGT + IVC
modulation’’ strategy, as the air flow is reduced and
pumping work is reduced, resulting in an increase in
open-cycle efficiency (as shown in Figures 8 and 7).
The IVC modulation also allows for a reduction in
effective compression ratios, which decrease NOx
through reductions in in-cylinder charge temperatures
prior to, during, and following combustion. This
enables earlier injection timing for the ‘‘open VGT +
IVC’’ strategy at 2.5 bar, as shown in Figure 11, thereby
providing a small closed-cycle efficiency improvement,
as shown in Figure 9.
The ‘‘open VGT’’ strategy and ‘‘open VGT + IVC
modulation’’ strategies have similar fuel consumption
benefits (Figure 6), and lower airflow (Figure 8), when
compared with the stock calibration for loads greater
than 2.5 bar BMEP. However, a shortcoming of these
low air flow strategies is that their turbo speeds are
lower when compared with the stock calibration
(Figure 12), as the exhaust flow through the turbine is
lower when cylinders are deactivated, IVC modulation
is used, or VGT is opened.
Lower pre-acceleration airflows and turbo speeds
increase turbo-lag in turbocharged diesel engines
Figure 11. Comparison of apparent heat release rate of an
active cylinder in cylinder deactivation and six-cylinder open
variable geometry turbine turbocharger strategy at 2200 r/min,
2.5 bar brake mean effective pressure. Cylinder deactivation has
higher per-cylinder fueling in an active cylinder to maintain the
demanded torque. This leads to a larger and more spread-out
heat release rate when compared with six-cylinder strategies.
IVC: intake valve closure; VGT: variable geometry turbine turbocharger.
Figure 10. Comparison of in-cylinder heat rejection of an
active cylinder in cylinder deactivation and six-cylinder open
VGT strategy at 2200 r/min, 2.5 bar brake mean effective
pressure. The in-cylinder heat rejection is calculated by
performing a first-law analysis with the boundary around the
cylinder. The fuel energy is distributed as brake work, exhaust
stream energy, and in-cylinder heat rejection or loss.
BMEP: brake mean effective pressure; IVC: intake valve closure; VGT:
variable geometry turbine turbocharger.
Figure 9. Comparison of closed-cycle efficiency versus brake
mean effective pressure at 2200 r/min. Closed-cycle efficiency
quantifies combustion completeness, piston compression and
expansion work, in-cylinder energy release during combustion,
and in-cylinder heat transfer. The closed-cycle efficiency
decreases for cylinder deactivation, owing to increased in-
cylinder heat loss during combustion.
BMEP: brake mean effective pressure; IVC: intake valve closure; VGT:
variable geometry turbine turbocharger.
Figure 12. Turbocharger speed versus brake mean effective
pressure at 2200 r/min for different strategies. All the strategies
have a lower turbocharger speed when compared with stock
calibration. All these strategies have similar turbocharger shaft
speeds despite the variation in air flow.
BMEP: brake mean effective pressure; IVC: intake valve closure; VGT:
variable geometry turbine turbocharger.
1010 International J of Engine Research 18(10)
during transients.
20
In addition, as shown in Figure 13,
the pressure differential across the exhaust gas recircu-
lation loop is decreased for these strategies, as com-
pared with the stock calibration, owing to the
reduction in airflow (as shown in Figure 8). Lower
pre-acceleration exhaust gas recirculation pressure dif-
ferentials decrease the potential to drive exhaust gas
recirculation flow during transients, increasing transi-
ent NOx emissions.
The implication of these phenomena for transient
performance at these conditions requires further study.
As mentioned in the introduction, early exhaust valve
opening, internal exhaust gas recirculation, turbochar-
ger electrification, supercharging, and powertrain
hybridization are all potential solutions, but are beyond
the scope of this paper. Availability of look-ahead
information, through vehicle data connectivity with
other vehicles or the cloud, to allow anticipation of
upcoming transients would also allow the use of one of
the strategies outlined in this paper when an upcoming
transient is not immediately pending.
Maintaining elevated aftertreatment component
temperatures
The selective catalytic reduction system operates most
efficiently when temperatures are between approxi-
mately 300 °C and 450 °C. Once the aftertreatment has
reached these temperatures, it is preferable to maintain
turbine outlet temperature at the upper end of this
range so that diesel oxidation catalyst fuel dosing
(which still requires temperatures above 250 °C) is not
required to keep the selective catalytic reduction system
temperatures elevated.
Reducing the air-to-fuel ratio is the most direct way
to increase turbine outlet temperature. Figure 14 specif-
ically shows a decreasing relationship between turbine
outlet temperature and air-to-fuel ratio, regardless of
operating strategy or load. As shown, and expected, a
reduction in air-to-fuel ratio results in an increase in
turbine outlet temperature. The most fuel-efficient way
to reduce the air-to-fuel ratio is by reducing the airflow
(as opposed to increasing the amount of fuel required).
As discussed in the previous section, cylinder deactiva-
tion and IVC modulation are fuel-efficient ways to
reduce engine airflow, and as such, are also effective
strategies for maintaining exhaust aftertreatment com-
ponent temperatures. More specifically, at a given load,
lower air-to-fuel ratios are possible via cylinder deacti-
vation and IVC modulation, as shown in Figure 15,
resulting in higher turbine outlet temperature, as shown
in Figure 16.
Figure 16 and Table 2 show that:
1. three-cylinder operation is preferred for loads
below 2.5 bar;
Figure 13. Comparison of delta pressure across the exhaust
gas recirculation versus brake mean effective pressure at 2200 r/
min for different strategies. All the strategies have a lower
exhaust gas recirculation delta pressure, compared with stock
calibration. The main reasons for the lower delta pressures are
the reduction in boost pressures for ‘‘open VGT’’ cases and
displaced volume for the cylinder deactivation strategy.
BMEP: brake mean effective pressure; IVC: intake valve closure; VGT:
variable geometry turbine turbocharger.
Figure 14. Direct correlation between air fuel ratio and
turbine outlet temperature at an engine speed of 2200 r/min for
different operating strategies. The shaded region indicates the
turbine outlet temperature range consistent with maintaining
selective catalytic reduction temperature between 300 °C and
450 °C. This temperature range is typically the aftertreatment
sweet spot wherein the conversion efficiency is maximum.
IVC: intake valve closure; VGT: variable geometry turbine turbocharger.
Figure 15. Air-to-fuel ratio versus brake mean effective
pressure at 2200 r/min.
BMEP: brake mean effective pressure; IVC: intake valve closure; VGT:
variable geometry turbine turbocharger.
Ramesh et al. 1011
2. four-cylinder operation is preferred for loads between
2.5 and 4 bar, and six-cylinder operation with an
open VGT is preferred for loads above 4 bar.
Also, note from Figure 16 that below ;6.5 bar BMEP,
using stock engine calibration and valve profiles, it is
not possible to meet desirable turbine outlet tempera-
tures for maintaining already elevated exhaust after-
treatment component temperatures. Conversely, the
combination of opening the VGT and valvetrain flexi-
bility (IVC modulation or cylinder deactivation) allows
desirable temperatures to be reached for loads greater
than ;1.5 bar.
Reaching desired aftertreatment component
temperatures quickly
During cold-start conditions, the aftertreatment system
would ideally warm-up as quickly as possible. This
section compares the catalyst warm-up characteristics
of various strategies at different loads. As a first
approximation, the heat transfer rate between the
exhaust gas and an aftertreatment catalyst depends on
the turbine outlet temperature, exhaust flow rate, and
instantaneous catalyst bed temperature. As an approxi-
mation, consider the heat transfer rate between an
incoming gas and the wall within a round pipe,
19
as
given by
q=C3_
m4=53(TTurbine Outlet TCatalyst)ð2Þ
where _
mis the experimental mass flow rate of the
exhaust gas going through the catalyst, TCatalyst is the
temperature of the catalyst, and Cis a constant that
depends on the geometry and material of the catalyst.
For each TCatalyst, this simple model yields a pre-
dicted heat transfer rate from the exhaust gas to the
catalyst by using the experimentally measured exhaust
mass flow rate and the turbine outlet temperature for
each load for different strategies. A positive heat trans-
fer rate corresponds to catalyst warm-up, as heat is
transferred from the exhaust gas to the catalyst.
Negative heat transfer rate corresponds to catalyst
cooling down, as the heat is transferred from the cata-
lyst to the exhaust gas.
For each load, predicted heat transfer rates are nor-
malized by the ‘‘six-cylinder stock calibration’’ case at a
catalyst bed temperature of 0 °C for that particular
load. The result allows assessment of the relative warm-
up characteristic of each strategy, at various loads. A
higher heat transfer rate is preferred during the catalyst
warm-up phase and is achieved using the optimal com-
bination of exhaust flow and turbine outlet temperature
for a particular catalyst bed temperature.
Figure 17 illustrates the catalyst warm-up character-
istics of the stock calibration, ‘‘open VGT,’’ ‘‘open
VGT + IVC modulation,’’ and cylinder deactivation
strategies at 1.3 bar BMEP. The six-cylinder stock cali-
bration mode has the highest heat transfer rate when
Table 2. Summary of experimental results illustrating ideal operating modes for fuel efficiency, aftertreatment warm-up,
aftertreatment temperature maintenance (e.g, ‘‘stay warm’’), and active diesel particulate filter regeneration. The top row shows
percentage fuel savings when compared with stock calibration. The bottom two rows show the turbine outlet temperature for the
strategy. Results show significant fuel consumption and aftertreatment thermal management benefits for cylinder deactivation and
intake valve closing modulation within the load range 1.3–6.3 bar BMEP.
Operating load, bar 1.3 2.5 3.8 5.1 6.3
Fuel efficiency 3 cylinder IVC modulation IVC modulation Open VGT Open VGT
(25.1%) (28.7%) (17.7%) (12%) (5.6%)
Aftertreatment warm-up Stock calibration Stock calibration Stock calibration 4 cylinder IVC modulation
(T\1008C)
Aftertreatment warm-up 3 cylinder 3 cylinder 4 cylinder 4 cylinder IVC modulation
(T.1008C)
Aftertreatment 3 cylinder 4 cylinder IVC modulation Open VGT Open VGT
maintain temperature (314 °C) (371 °C) (377 °C) (397 °C) (417 °C)
Active diesel particulate 3 cylinder 3 cylinder 4 cylinder 4 cylinder IVC modulation
filter regeneration (314 °C) (473 °C) (484 °C) (537 °C) (440 °C)
BMEP: brake mean effective pressure; IVC: intake valve closing; VGT: variable geometry turbine.
Figure 16. Turbine outlet temperature versus brake mean
effective pressure at 2200 r/min. The shaded region indicates
the turbine outlet temperature range consistent with
maintaining selective catalytic reduction temperature between
300 °C and 450 °C. This temperature range is typically the
aftertreatment sweet spot wherein the conversion efficiency is
maximum. BMEP: brake mean effective pressure; IVC: intake
valve closure; VGT: variable geometry turbine turbocharger.
1012 International J of Engine Research 18(10)
the catalyst bed temperature is lower than 100 °C . This
is because the positive impact of elevated exhaust mass
flow is more important than elevated turbine outlet
temperature at lower catalyst bed temperatures.
As the catalyst bed temperature increases, higher
exhaust gas temperatures are required to maintain heat
flow from gas to bed (equation (2)). As a result (see
Figure 17), the three-cylinder mode outperforms all
others above bed temperatures of 100 °C. The catalyst
can only reach a temperature of 200 °C if the engine
only operates in the six-cylinder stock calibration mode,
whereas the three-cylinder mode enables catalyst tem-
peratures up to 300 °C. Figure 17 illustrates that the
fastest way to warm up the catalyst bed at 1.3 bar is to
operate in the stock six-cylinder mode until the bed
reaches 100 °C, at which point a mode switch to three-
cylinder operation should be made. Combination of
Figures 6 and 17 suggests that the six-cylinder ‘‘open
VGT’’ mode could be used instead of the stock six-
cylinder mode, if fuel economy is paramount during the
warm-up.
Figure 18 illustrates the catalyst warm-up character-
istics of the stock calibration, ‘‘open VGT,’’ cylinder
deactivation, and ‘‘open VGT + IVC modulation’’ at
2.5 bar BMEP. Like the 1.3 bar load condition (Figure
17), operation in the stock six-cylinder stock mode will
result in the highest gas-to-bed heat transfer rates when
the bed temperature is below 100 °C, while the three-
cylinder mode is preferred above this temperature.
Significantly, the three-cylinder strategy can heat the
catalyst to temperatures in excess of 470 °C.
Figure 19 shows the catalyst warm-up characteristics
of the stock calibration, ‘‘open VGT,’’ ‘‘open VGT +
IVC modulation,’’ and four-cylinder strategies at 3.8
bar BMEP. Again, until the catalyst bed temperature
reaches 100 °C, the six-cylinder stock calibration has
the highest heat transfer rate, while at this load four-
cylinder mode is preferred above 100 °C. This is a
result of the four-cylinder mode having a turbine outlet
temperature about 200 °C higher than the other two
modes, as shown in Figure 16.
Figure 20 illustrates the catalyst warm-up character-
istics of stock calibration, ‘‘open VGT,’’ ‘‘open VGT
+ IVC modulation,’’ and the four-cylinder strategy at
5.1 bar BMEP torque. The four-cylinder strategy has a
higher heat transfer rate than the other three cases at
all catalyst bed temperatures. The four-cylinder mode
Figure 19. Catalyst warm-up characteristics of different
strategies at 2200 r/min, 3.8 bar brake mean effective pressure.
Predicted heat transfer rates are normalized using the heat
transfer rate of the ‘‘six-cylinder stock calibration’’ case at a
catalyst bed temperature of 0 °C.
IVC: intake valve closure; VGT: variable geometry turbine turbocharger.
Figure 18. Catalyst warm-up characteristics of different
strategies at 2200 r/min, 2.5 bar brake mean effective pressure.
Predicted heat transfer rates are normalized using the heat
transfer rate of the ‘‘six-cylinder stock calibration’’ case at a
catalyst bed temperature of 0 °C.
IVC: intake valve closure; VGT: variable geometry turbine turbocharger.
Figure 17. Catalyst warm-up characteristics of different
strategies at 2200 r/min, 1.3 bar brake mean effective pressure.
Predicted heat transfer rates are normalized using the heat
transfer rate of the ‘‘six-cylinder stock calibration’’ case at a
catalyst bed temperature of 0 °C.
IVC: intake valve closure; VGT: variable geometry turbine turbocharger.
Ramesh et al. 1013
at this load can also heat a catalyst up to temperatures
in excess of 520 °C.
Figure 21 illustrates the catalyst warm-up character-
istics of the stock calibration, ‘‘open VGT,’’ and ‘‘open
VGT + IVC modulation’’ strategy at 6.31 bar BMEP.
The open VGT + IVC strategy has the highest heat
transfer rate, and can heat a catalyst to temperatures in
excess of 450 °C.
Active diesel particulate filter regeneration
The engine needs to operate in such a way that the tur-
bine outlet temperature is above 450 °C in order to
enable an active diesel particulate filter regeneration
without dosing the diesel oxidation catalyst. The three-
cylinder mode yields the highest turbine outlet tempera-
ture when compared with the stock calibration and
open VGT strategy at 1.3 bar BMEP. Cylinder deacti-
vation (three-cylinder and four-cylinder) enables tur-
bine outlet temperature to reach 450 °C for loads
between 2.5 and 5.1 bar BMEP, which is enough to
realize active regeneration in the diesel particulate filter
without dosing the diesel oxidation catalyst. However,
cylinder deactivation is not possible at a load of 6.31
bar BMEP or higher, as the air flow is too low to sus-
tain smoke-free combustion. As such, at 6.31 bar,
BMEP the best way to operate would be to use the
‘‘open VGT + IVC’’ strategy, as it has a higher tur-
bine outlet temperature as well as a lower fuel con-
sumption when compared with the stock calibration.
Conclusions
This paper demonstrates the benefits of cylinder deacti-
vation and other low airflow strategies under elevated
speed, low-load operation when the goal is to:
(a) reduce fuel consumption;
(b) warm up the aftertreatment;
(c) maintain elevated aftertreatment temperatures; or
(d) enable active diesel particulate filter regeneration
without dosing the diesel oxidation catalyst.
A summary of experimental results is shown in
Table 2. The key conclusions are as follows.
1. The fuel savings of cylinder deactivation, ‘‘open
VGT,’’ and IVC modulation are between 5% and
25%, depending on the load. The fuel savings are
primarily due to a reduction in airflow leading to
an increased open-cycle efficiency. For loads less
than 2.54 bar BMEP, cylinder deactivation has
higher fuel efficiency than ‘‘open VGT’’ operation.
For loads greater than 2.54 bar BMEP, the ‘‘open
VGT’’ and IVC modulation strategy have better
fuel economy than cylinder deactivation, as there
is higher in-cylinder heat loss during cylinder deac-
tivation, which causes overall efficiency to decrease
for the cylinder deactivation strategy.
2. The VGT is squeezed under these low-load condi-
tions in order to prepare for a transient event.
Opening up the VGT decreases the delta pressure
across the engine, thereby increasing the fuel effi-
ciency. However, this might have detrimental
effects on transients; this needs to be understood.
3. At catalyst bed temperatures below 100 °C, the
stock calibration warms up the aftertreatment
faster for loads between 1.3 and 3.8 bar BMEP.
For loads between 3.8 and 5.1 bar BMEP, the
four-cylinder mode will warm up the aftertreat-
ment system more quickly.
Figure 20. Catalyst warm-up characteristics of different
strategies at 2200 r/min, 5.1 bar brake mean effective pressure.
The predicted heat transfer rates are normalized using the heat
transfer rate of the ‘‘six-cylinder stock calibration’’ case at a
catalyst bed temperature of 0°C.
IVC: intake valve closure; VGT: variable geometry turbine turbocharger.
Figure 21. Catalyst warm-up characteristics of different
strategies at 2200 r/min, 6.31 bar brake mean effective pressure.
The predicted heat transfer rates are normalized using the heat
transfer rate of ‘‘six-cylinder stock calibration’’ case at a catalyst
bed temperature of 0 °C.
IVC: intake valve closure; VGT: variable geometry turbine turbocharger.
1014 International J of Engine Research 18(10)
4. At 6.31 bar BMEP, there is not enough oxygen
available in the cylinder to sustain smoke-free com-
bustion during cylinder deactivation. However, the
‘‘open VGT + IVC modulation’’ strategy can
help realize faster warm-up at all catalyst bed
temperatures.
5. When the aftertreatment is warm, cylinder deacti-
vation helps keep the engine and aftertreatment
system above 250 °C for loads below 2.54 bar
BMEP. For loads above 2.54 bar BMEP, the open
VGT and IVC modulation strategies help keep the
system warm in a fuel-efficient manner. Cylinder
deactivation on average gives rise to a turbine out-
let temperature increment of 150 °C–250 °C in the
1.3–5.1 bar BMEP load range.
6. Cylinder deactivation not only helps to warm up
the diesel particulate filter faster under certain con-
ditions but also enables diesel particulate filter gas
inlet temperatures required for active diesel parti-
culate filter regeneration without using a diesel oxi-
dation catalyst doser.
7. Strategies including turbocharger electrification,
supercharging, powertrain hybridization, and
availability of look-ahead information can be used
to improve the transient behavior of the engine.
Several variable valve actuation strategies, such as
zero valve overlap or early exhaust valve opening,
could also prove viable strategies to improve tran-
sients but are beyond the scope of this paper.
Acknowledgements
The heavy-duty engine was generously provided by
Cummins Inc. Technical assistance was provided by
both Cummins Inc. and Eaton for this work. The
authors would also like to thank the engines lab per-
sonnel at Ray W Herrick labs, particularly David
Meyer and Ron Evans, for their immense support
toward this work.
Declaration of conflicting interests
The author(s) declared no potential conflicts of interest
with respect to the research, authorship, and/or publi-
cation of this article.
Funding
The author(s) disclosed receipt of the following finan-
cial support for the research, authorship, and/or publi-
cation of this article: Funding for this project was
provided by Cummins Inc. and Eaton.
References
1. Stanton DW. Systematic development of highly efficient
and clean engines to meet future commercial vehicle
greenhouse gas regulations. SAE Int J Engines 2013; 6:
1395–1480.
2. Johnson T. Vehicular emissions in review. SAE Int J
Engines 2016; 9: 1258–1275.
3. Koebel M, Elsener M and Kleemann M. Urea-SCR: A
promising technique to reduce NOx emissions from auto-
motive diesel engines. Catal Today 2000; 59(3): 335–345.
4. Girard J, Cavataio G, Snow R, et al. Combined Fe-Cu
SCR systems with optimized ammonia to NOx ratio for
diesel NOx control. SAE Int J Fuels Lubr 2009; 1(1):
603–610.
5. Lambert C, Hammerle R, McGill R, et al. Technical
advantages of urea SCR for light-duty and heavy-duty die-
sel vehicle applications. SAE Trans 2004; 113(4): 580–589.
6. Charlton S, Dollmeyer T and Grana T. Meeting the US
heavy-duty EPA 2010 standards and providing increased
value for the customer. SAE Int J Commer Veh 2010;
3(1): 101–110.
7. Walker A. Controlling particulate emissions from diesel
vehicles. Top Catal 2004; 28(1–4): 165–170.
8. Allansson R, Blakeman PG, Cooper BJ, et al. Optimising
the low temperature performance and regeneration effi-
ciency of the continuously regenerating diesel particulate
filter (CR-DPF) system. SAE paper 2002-01-0428, 2002.
9. Stadlbauer S, Waschl H, Schilling A, et al. DOC tem-
perature control for low temperature operating ranges
with post and main injection actuation. SAE paper 2013-
01-1580, 2013.
10. Song X, Surenahalli H, Naber J, et al. Experimental and
modeling study of a diesel oxidation catalyst (DOC)
under transient and CPF active regeneration conditions.
SAE paper 2013-01-1046, 2013.
11. Naseri M, Aydin C, Mulla S, et al. Development of emis-
sion control systems to enable high NOx conversion on
heavy duty diesel engines. SAE Int J Engines 2015; 8:
1144–1151.
12. Stanton D, Charlton S and Vajapeyazula P. Diesel engine
technologies enabling powertrain optimization to meet
US greenhouse gas emissions. SAE Int J Engines 2013;
6(3): 1757–1770.
13. Parks J, Huff S, Kass M, et al. Characterization of in-
cylinder techniques for thermal management of diesel
aftertreatment. SAE paper 2007-01-3997, 2007.
14. Singh P, Thalagavara AM, Naber JD, et al. An experi-
mental study of active regeneration of an advanced cata-
lyzed particulate filter by diesel fuel injection upstream of
an oxidation catalyst. SAE paper 2006-01-0879, 2006.
15. Schwoerer JA, Kumar K, Ruggiero B, et al. Lost-motion
VVA systems for enabling next generation diesel engine
efficiency and after-treatment optimization. SAE paper
2010-01-1189, 2010.
16. Magee M. Exhaust thermal management using cylinder
deactivation. MSME Thesis, Purdue University, USA,
2013.
17. Lu X, Ding C, Ramesh A, et al. Impact of cylinder deac-
tivation on active diesel particulate filter regeneration at
highway cruise conditions. Front Mech Eng 2015; 1: 9.
18. Garg A, Magee M, Ding C, et al. Fuel-efficient exhaust
thermal management using cylinder throttling via intake
valve closing timing modulation. Proc IMechE, Part D: J
Automobile Engineering 2016; 230(4): 470–478.
19. Ding C, Roberts L, Fain DJ, et al. Fuel efficient exhaust
thermal management for compression ignition engines
during idle via cylinder deactivation and flexible valve
actuation. Int J Engine Res 2016; 17(6): 619–630.
Ramesh et al. 1015
20. Rakopoulos CD and Giakoumis EG. Diesel engine transi-
ent operation: principles of operation and simulation analy-
sis. London: Springer, 2009.
21. Lee W, Schubert E, Li Y, et al. Electrification of turbo-
charger and supercharger for downsized internal combus-
tion engines and hybrid electric vehicles-benefits and
challenges. In: 2016 IEEE transportation electrification
conference and expo (ITEC), Dearborn, MI, 27–29 June
2016. Piscataway, NJ: IEEE.
22. Grondin O, Thibault L and Querel C. 3rd IFAX work-
shop on engine and powertrain control, simulation and
modeling transient torque control of a diesel hybrid
powertrain for NOx limitation. IFAC Proc Vol 2012;
45(30): 286–295.
23. Ishikawa N. A study on emissions improvement of a die-
sel engine equipped with a mechanical supercharger. Int J
Engine Res 2012; 13(2): 99–107.
24. Wang Y, Zhang H and Sun Z. Optimal control of the
transient emissions and the fuel efficiency of a diesel
hybrid electric vehicle. Proc IMechE, Part D: J Automo-
bile Engineering 2013; 227(11): 1546–1561.
25. Sivertsson M and Eriksson L. An optimal control bench-
mark: Transient optimization of a diesel-electric power-
train. In: Proceedings of the 55th conference on simulation
and modelling (SIMS 55), modelling, simulation and opti-
mization, Aalborg, Denmark, 21–22 October 2014, paper
no. 006, vol. 108, pp.59–63, Linko
¨ping: Linko
¨ping Uni-
versity Electronic Press.
26. Florell C. Utilizing look-ahead information to minimize
fuel consumption and NOx emissions in heavy duty vehicles.
Master’s Thesis, Linko
¨ping University, Sweden, 2015.
27. Ojeda WD. Effect of variable valve timing on diesel com-
bustion characteristics. SAE paper 2010-01-1124, 2010.
28. Lombard B and le Forestier R. Advanced combustion
and engine integration of a hydraulic valve actuation sys-
tem (camless). Inge
´nieurs de l’automobile 2007; 787:
61–67.
1016 International J of Engine Research 18(10)
... To perform calculations, we made the following assumptions: we consider the transmission efficiency to be constant at each of the actual gears; a car (tractor) is moving at a constant speed (Vm=const), uniformly (j=0), on a horizontal surface (α=0º), without longitudinal vibrations affecting the changes in the tractive effort and torque of the engine, without slipping (δ= 0); we neglect aerodynamic drag force Рw because of the low movement speed in the case of a tractor [33][34][35][36][37]. ...
Article
Full-text available
In fuel economy, a rising level of interest in heavy duty diesel engines that industry has witnessed over the last few years continues to go up and this is not likely to change. Lowering the fuel consumption of all internal combustion engines remains a priority for years to come, driven by economic, legislative, and environmental reasons. According to statistics, the share of operating expenses to ensure transport operations in industrial production is 15-20%, wherein 16-30% of the total volume of transport operations concerns a car, tractor, and trailer. During transport operations, the engine load by the torque, in most cases, does not exceed 40-50%. The paper investigates the increase in fuel efficiency of cars and tractors by disconnecting some of the engine cylinders operated in low-load and idling modes. The research has led to the establishment of the theoretical dependencies between the effective power, engine efficiency, mass of the transported cargo, speed of the car (tractor) and the number of disconnected engine cylinders. Results of experiments suggest the interdependencies of the performance parameters of the car (tractor) when disconnecting some of the engine cylinders. It has also been established that the maximum reduction in the hourly fuel consumption occurs in the idling mode while it decreases along with an increase in the load.
... One engine technology at a time shows that cylinder deactivation technology can save 4.5 ~ 6 % on petrol (Knight, 2010). Cylinder deactivation in an IC engine under part load improves engine economy and exhaust condition by increasing F/A ratio and reducing pumping loss Ramesh et al., 2017). When a 4-cylinder SI engine is run on two cylinders at 2000 and 4000 RPM, there is a considerable improvement in fuel consumption and pumping work (Zhao et al., 2018). ...
Article
Full-text available
This article compares engine performance, combustion, and emissions at part-load in cylinder deactivation (CD) mode versus the traditional spark ignition mode. This comparison demonstrates the applicability of cylinder deactivation for a compact 3-cylinder engine. Experiments have been performed on a multi-point fuel injection engine having a 1000 cc displacement and equipped with an open engine control unit. A response surface approach has been applied to design experiments and parametric engine optimization. The analysis of variance test indicates that the load is the input factor that impacts the deactivation mode most. At 3500 RPM and 45 N•m load, the engine performs best in CD mode, with BSFC, BTE, CP, HRR, and UHC values of 332.805 g/kW-hr, 25.26 %, 52.51 bar, 43.24 J, and 10.38 ppm, respectively. The validation test results for CD mode show that the percentage error for the BSFC, BTE, CP, HRR, and UHC responses was found to be 2.86, 2.91, 3.01, 3.47, and 3.97, respectively, within the acceptable range. As compared to SI mode, the current analysis has discovered a considerable decrease in BSFC (11.47 %), an improvement in BTE (12.25 %), a higher CP (80.65 %), a greater HRR (91.74 %), and a decrease in UHC (95.48 %).
... As the number of cylinders increased, it would cost more in the range of $200-$250 to reduce NVH in medium-to heavy-duty engines to implement cylinder deactivation [8]. In IC engines at part-load conditions, cylinder deactivation promotes a higher F/A ratio and reduces pumping loss; consequently, engine efficiency and exhaust condition were improved [9,10]. Zhao J et al. use a similar method in a four-cylinder gasoline engine where two intake manifold was attached with a variable valve timing strategy. ...
Article
Full-text available
This research paper aims to compare an engine's probable and contemporary condition concerning performance, combustion, and emission characteristics to identify obstructions and assess the suitability of deactivating the cylinder. A 998 cc three-cylinder spark-ignition, multi-point fuel-injection engine with an open engine control unit that features normal and cylinder deactivation modes has been selected. Experimentation was carried out at a constant speed of 3000 rpm under three load conditions: 15, 30, and 45 N m. The result shows that 45 N m acts as full engine load when two cylinders are active. Consequently , a reduction in pumping loss during deactivation mode facilitates brake-specific fuel consumption drops by 11.55% and brake thermal efficiency rises by 3.18%. In deactivation mode, peak pressure has been observed to be 1.75 times that of the conventional method, where the maximum heat release rate was 1.82 times that of the normal mode. The fraction of mass burned during cylinder deactivation is higher for all crank angle levels. It indicates an enhanced combustion rate and a higher degree of combustion efficiency than those in normal mode. An increase in average gas temperature in deactivation mode facilitates the catalytic converter to work effectively and efficiently. In deactivation mode, unburned hydrocarbons are reduced significantly, while carbon dioxide reduction is 13.86% at 30 N m. This research paves the path for adopting cylinder deactivation mode in smaller engines and offers prospective benefits such as better fuel usage, enhanced engine performance, more efficient combustion, and reduced pollution levels.
... However, in most of those aforementioned techniques, there is usually a need for extreme delay or advance of valve timings-such as in EEVO and EIVC modes-or the need to control both valve timing and fuel injection-such as in CDA mode-for high exhaust temperature improvement. Moreover, the EAT warmup period is degraded in EIVC or CDA modes due to highly reduced exhaust flow rates [40]. Despite those negative Energies 2023, 16, 4542 3 of 25 effects, VVT is a proven technique to increase engine-out temperature. ...
Article
Full-text available
The exhaust after-treatment (EAT) threshold temperature is a significant concern for highway vehicles to meet the strict emission norms. Particularly at cold engine start and low loads, EAT needs to be improved above 250 °C to reduce the tailpipe emission rates. Conventional strategies such as electrical heating, exhaust throttling, or late fuel injection mostly need a high fuel penalty for fast EAT warmup. The objective of this work is to demonstrate using a numerical model that a combination of the Miller cycle and delayed exhaust valve opening (DEVO) can improve the tradeoff between EAT warmup and fuel consumption penalty. A relatively low-load working condition (1200 RPM speed and 2.5 bar BMEP) is maintained in the diesel engine model. The Miller cycle via retarded intake valve closure (RIVC) is noticeably effective in increasing exhaust temperature (as high as 55 °C). However, it also dramatically reduces the exhaust flow rate (over 30%) and, thus, is ineffective for rapid EAT warmup. DEVO has the potential to enhance EAT warmup via increased exhaust temperature and increased exhaust flow rate. However, it considerably decreases the brake thermal efficiency (BTE)—by up to 5%—due to high pumping loss in the system. The RIVC + DEVO combined technique can elevate the exhaust temperature above 250 °C with improved fuel consumption—up to 10%—compared to DEVO alone as it requires a relatively lower rise in pumping loss. The combined method is also superior to RIVC alone. Unlike RIVC alone, the RIVC + DEVO combined mode does not need the extreme use of RIVC to increase engine-out temperature above 250 °C and, thus, provides relatively higher heat transfer rates (up to 103%) to the EAT system through a higher exhaust flow rate. The RIVC + DEVO combined method can be technically more difficult to implement compared to other methods. However, it has the potential to maintain accelerated EAT warmup with improved BTE and, thus, can keep emission rates at low levels during cold start and low loads.
... An accompanying phenomenon of piston deactivation is a change in the vibration spectrum of entire drive. From the point of dynamics of the mechanical system, the deactivation of the cylinders is accompanied by the emergence of resonances even from the so-called secondary harmonic components that have a significant character of dominant load amplitudes, and the torsional oscillation in the mechanical system will increase significantly [46][47][48][49]. ...
Article
Full-text available
The industries of shipping, shipbuilding and port operations are among those in which mechanical drives with piston machines are widely used. The wide use of piston machines is the result of many years of experience and many years of development and modernization of piston machines. Usually, they operate as mechanical drives with constant operating speeds, with the exception of drives with combustion engines, which operate in a wider range of operating speeds. The limiting condition of innovation of mechanical drives with piston machines, resulting from the nature of the piston machine operation, is the torsional oscillation. The effort to decrease an energy demand of mechanical drives requires the application of non-traditional working modes, which can be considered as a deactivation of the cylinders of piston machine or an expansion of the working speed range. One of the possibilities of eliminating these limiting factors is an application of a pneumatic tuner in mechanical drives, which, in contrast to traditional solutions, has a wide range of torsional stiffness that can be smoothly changed. During experimental measurements in the resonance area, at the operating speed of 700 rpm after torsional stiffness change, a torsional vibration value of 15 Nm decreased to 5 Nm.
... The main difficulties remain in the complicated balance when matching the Radial Inflow Turbine (RIT) to maintain the engine operates far from surge and choke lines of the CC's map. To reach these objectives, several works tried to redesign or to propose a new design of TC compressor to enhance the engine output power, 9,10 improve mechanical sustainability, 11,12 secure working of the CC (surge margin), [13][14][15][16] or to optimize the design procedure. 17,18 Some researches were focused on the components of the TC or ICE including inlet and exhaust manifolds, 19 intercooler, 20,21 the CC and RIT casings, 22,23 variable turbine geometry, 24 exhaust gas recirculation, 25,26 or twin-entry radial turbine. ...
Article
Full-text available
The present paper proposes a novel methodology of turbocharging automotive engines to reach targeted performance. The actual method is tested and validated against simulation test results of two turbocharged diesel engines; engine I, three cylinders, 1.5 L, and engine II, six cylinders, 5.9 L. The present procedure is subdivided into four key parts; namely, database construction, selection procedure, turbocharger preliminary design, and engine modeling. Based on geometric dimensions and aerodynamic parameters provided by the preliminary design procedure, 3D geometries of the turbine and compressor are generated for each studied engine. After integrating previous data into a constructed turbocharger database, two turbochargers are selected for the engine I, while only one turbocharger for the engine II. The findings show that, at the engine speed of 4000 rpm, engine I matched with the adequate turbocharger reached a target power about 2.7%, compared to the original turbocharger equipping engine I. Furthermore, engine II reached a rated power of 299.3 kW at 2500 rpm which is slightly under the original one by 2.64 kW. The superimposition of the engine operating area on compressor and turbine maps provided satisfactory results in terms of turbocharger-engine output performance, fuel consumption, secure functioning and engine thermal strength. Finally, the main advantage of the developed methodology consists of its ability to be applied at both earlier and last stages of the engine turbocharging process or to find new adequate turbochargers to replace the original one for economic, mechanical or for safety reasons.
Article
Today’s CI engines are subject to strict regulations of pollutant emissions and ambitious fuel consumption targets. Therefore, the interaction between the engine and the exhaust aftertreatment system (ATS) has become increasingly important. Numerous studies have shown that a variable valve train (VVT) improves the interaction between engine and ATS. However, most of these studies either quantify the advantage on a specific engine or only present complex CFD models, such that the results are not easily transferable to different engines. Thus, engine manufacturers cannot directly use these results to assess the advantage of various VVT strategies for their engines. In this paper, we propose a cycle-discrete cylinder model based on first principles which allows to simulate various VVT strategies. In contrast to present methods based on CFD, the proposed cylinder model can be realized with the equations presented. Furthermore, the model is identified with measurement data of an engine without a VVT. A separate engine, which is retrofitted with a fully VVT, is used to validate the proposed modeling approach. Using the identified model in combination with a mean-value model of the air path, we are able to simulate the effects of early intake valve closing, early exhaust valve opening, and cylinder deactivation for a complete CI engine that has no VVT installed. The model is then used to highlight the advantage of a VVT for two scenarios at part-load operation. At cold start, where the temperature of the ATS must be increased quickly, variable valve timing achieves higher enthalpy flows to the ATS while also lowering engine-out NOx emissions when compared to a standard engine strategy. If the ATS is at the operating temperature, cylinder deactivation achieves significantly higher enthalpy flows which prevents the ATS from cooling down. In addition, cylinder deactivation also lowers fuel consumption and engine-out NOx emissions.
Article
Exhaust thermal management (ETM) plays a prime role in reducing pollutant emissions from internal combustion engines (ICEs), especially during cold-start and warm-up conditions. Under ever-stringent emissions and fuel-efficiency regulations, it is challenging to achieve a better trade-off between energy efficiency and emissions. This review paper is focused on the various ETM technologies and their applications, both engine-based and device-added. Furthermore, a comprehensive explanation and analysis of the various ETM technology principles, working characteristics, advantages, and limitations of their applications are provided to aid pertinent researchers in the choice and design of thermal management solutions. Engine-based ETM technologies cannot be used to their maximum potential because of additional fuel consumption or limited application range. Solutions integrating multiple technologies are becoming more prevalent to mitigate the fuel consumption and emissions penalties associated with engine-based ETM technologies. From an energy efficiency standpoint, closely coupled the ETM device, such as an electrically heated catalyst (EHC), with the controlled device is more efficient, due to the reduction of heat losses. In addition, ETM techniques based on waste heat recovery (WHR), especially thermodynamic cycles and thermal energy storage (TES) systems, are progressively gaining attention because of the demand to minimize CO2 emissions in the field of marine, locomotive and stationary power generators. However, the difficulty of miniaturization limits their development in the vehicle sector. Finally, the review concludes with suggested technological solutions and future work directions to address the challenges that ETM technologies confront in applications.
Article
Full-text available
Intake throttling has been verified as an effective approach to increase the exhaust temperature of diesel engines, which could benefit the catalytic efficiency aftertreatment. To better understand the influence of intake throttling on the combustion characteristics and exhaust emissions of light-duty diesel engines operating under idle mode, a light-duty diesel engine was experimentally investigated. This study is a follow-on to previous studies on the effect of throttling on light-duty diesel engine exhaust temperatures and emissions. Tests were conducted at a fixed idle speed of 1100 rpm, and the throttle position and intake manifold air pressure (MAP) were varied. The in-cylinder pressure, pressure rise rate, heat release rate (HRR), in-cylinder temperature, exhaust temperature, and regular gaseous emissions were analyzed. The results indicated that under the influence of intake throttling, the MAP decreased from 101 kPa under wide-open-throttle (WOT) conditions to 52.5 kPa under the heaviest throttling conditions, and the exhaust temperature increased from 100 °C to 200 °C, with a fuel penalty associated with the increase in the pumping indicated mean effective pressure (IMEP). The in-cylinder pressure continuously declined with decreasing MAP, while the HRR generally increased with increasing MAP. Under WOT conditions, the ignition delay decreased, while the combustion duration decreased under heavier throttling conditions. The in-cylinder temperature with throttling was higher than that under WOT conditions, and after post-injection treatment, the in-cylinder temperature exhibited an increasing trend with decreasing MAP. The CO2, CO, NOx, and HC emissions increased with increasing throttling amounts.
Article
Worldwide automotive emission regulations for heavy-duty diesel engines are increasingly stringent, especially for nitrogen oxide (NOx) and particulate matter (PM). The development of high-efficiency aftertreatment technologies and the optimization of the urea injection strategies are key factors for the diesel engines to meet the future regulations. In this paper, the effects of temperature and space velocity on close coupled selective catalyst reduction (ccSCR) and SCR are investigated firstly. The highest NOx conversion efficiency of ccSCR and main SCR is observed at 350 ℃, and an increase of space velocity leads to a more significant decrease on NOx conversion efficiency of ccSCR than that of SCR. Then three phases of urea injection state for both nozzles are determined according to the catalyst inlet temperature in the Federal Test Procedure (FTP) cycle. Urea injection strategy is optimized by urea injection ratio calibration and urea injection strategy updating based on the characteristic of three phases, as well as NH3 sensor feedback. Finally, the composite tailpipe NOx emission under FTP cycle is reduced below 0.027 g/kW·h, satisfying California air resources board (CARB) ultra-low NOx emission regulation with a penalty of slight increase in CO2 emissions. However, adopting a high reactivity gasoline can improve the efficiency while maintaining the low level of NOx emissions.
Conference Paper
Full-text available
Forced induction technology (turbocharging and supercharging) can enhance the performance of an internal combustion engine by compressing inlet air charge, allowing full engine power to be produced efficiently. As the fuel economy and greenhouse gas emission standards are projected to be much more stringent globally, the use of a forced induction engine in passenger cars and light duty trucks has become a new inevitable trend in the automotive industry. However, the conventional forced induction system suffers from the slow transient response, especially when the engine speed is low, a phenomenon typically known as turbo lag. The electrification of forced induction system, called electric forced induction system (EFIS), has emerged as a feasible solution and it also possesses numerous benefits depending on its topologies. This paper provides a comprehensive study on EFIS by investigating system level topologies, performance, various types of high-speed machines, power electronics, and control techniques. The advantages and disadvantages of existing electric forced induction system are summarized and the new challenges and opportunities are also introduced.
Article
This review paper summarizes major and representative developments in vehicular emissions regulations and technologies from 2015. The paper starts with the key regulatory advancements in the field, including newly proposed Euro 6 type regulations for Beijing, China, and India in the 2017-20 timeframe. Europe is continuing developments towards real driving emissions (RDE) standards with the conformity factors for light-duty diesel NOx ramping down to 1.5X by 2021. The California heavy duty (HD) low-NOx regulation is advancing and may be proposed in 2017/18 for implementation in 2023+. LD (light duty) and HD engine technology continues showing marked improvements in engine efficiency. Key developments are summarized for gasoline and diesel engines to meet both the emerging criteria and greenhouse gas regulations. LD gasoline concepts are achieving 45% BTE (brake thermal efficiency or net amount of fuel energy gong to the crankshaft) and closing the gap with diesel. Projections indicate tight CO2 regulations will require some degree of hybridization and/or high-performing diesel engines. HD engines are demonstrating more than 50% (BTE) using methods that can reasonably be commercialized; and proposals are developed for reaching 55% BTE. Lean NOx control technologies are summarized, including SCR (selective catalytic reduction), SCR filters, and combination systems. Emphasis is on durability, N2O, and greatly reduced emissions. Diesel PM (particulate matter) reductions are evolving around the nature of soot and the distribution in the filters. Gasoline direct injection (GDI) particulates carry PAHs (polycyclic aromatic hydrocarbons) through the three way catalyst, but filters can remove most of them. Gasoline particulate filter regeneration is now better understood. Improved understanding of oxidation catalyst formulations are reported with further quantification of the impact of precious metal formulations. Finally, the paper discusses some key developments in three-way catalysts, with improved understanding of the catalyst-support interactions, and the introduction of a low-mass cellular substrate that improves TWC cold start performance.
Article
In a typical Diesel engine exhaust after treatment system consisting of a Diesel oxidation catalyst (DOC), a Diesel particulate filter (DPF) and a selective catalytic reduction (SCR) system the main purpose of the DOC, besides the oxidation of CO to CO2, is the oxidation of NO to NO2. The NO to NO2 conversion is an essential contribution for the downstream SCR system because the fast SCR reaction which provides the highest conversion rates of NOx to H2O and N2 works well only under roughly equal concentrations of NO and NO2. The typical amount of NO to NOx ratio produced by the engine is about 0.95, hence the DOC is necessary to decrease this coefficient close to 0.50. Due to the temperature dependency of the DOC reaction mechanism the oxidation of NO to NO2 takes only place sufficiently if the temperature of the DOC is higher than 200°C, which, however, cannot be reached during low engine speed and low load situations. As a consequence, under these circumstances the whole reduction system is ineffective and, moreover, the potentially high raw NOx emissions in this operating range may cause significant tailpipe emission values. Against this background this paper presents a strategy to raise the DOC temperature during lower temperature operating ranges in its most effective way by acting on the post or the main injection. The final temperature control approach is designed with respect to the additionally required fuel amount and the effect of the different injections on the total NOx emissions. Finally the control strategy was implemented on a test bench to evaluate the benefit of the DOC temperature increase and consequently the gain in SCR conversion efficiency against the fuel penalty. Of course, the additional consumption must be traded off against other system modifications, especially in the after treatment case.
Article
In this study, a DOC catalyst was experimentally studied in an engine test cell with a 2010 Cummins 6.7L ISB diesel and a production aftertreatment system. The test matrix consisted of steady state, active regeneration with in-cylinder fuel dosing and transient conditions. Conversion efficiencies of total hydrocarbon (THC), CO, and NO were quantified under each condition. A previously developed high-fidelity DOC model [1] capable of predicting both steady state and transient active regeneration gaseous emissions was calibrated to the experimental data. The model consists of a single 1D channel where mass and energy balance equations were solved for both surface and bulk gas regions. The steady-state data were used to identify the activation energies and pre-exponential factors for CO, NO and HC oxidation, while the steady-state active regeneration data were used to identify the inhibition factors. The transient data were used to simulate the thermal response of the DOC. A calibration procedure was developed and the identified model parameters are presented. The performance of the model was compared to the experimental results in terms of the outlet NO2/NOx ratios and CO, HC conversion efficiencies. The results show that the model is able to simulate the DOC performance under steady-state and transient conditions and during active regeneration conditions. The model is able to predict the outlet temperatures and HC concentrations within the measurement uncertainties. An inhibition factor that accounts for the inhibition of reactions in the presence of high hydrocarbon concentrations was determined so that the outlet HC and NO 2 concentrations during active regeneration were simulated accurately.
Article
Traditionally, the study of internal combustion engines has focused on the steady-state performance. However, the daily driving schedule of automotive and truck engines is inherently related to unsteady operation, whereas the most critical conditions encountered by industrial or marine engines are met during transients. Unfortunately, the transient operation of turbocharged diesel engines has been associated with poor driveability, as well as overshoot in particulate and gaseous emissions, making the study and modeling of transient engine operation an important scientific objective. Diesel Engine Transient Operation provides an in-depth discussion of all the complex thermodynamic and dynamic phenomena that are experienced by a diesel engine during load increase, acceleration, cold starting or Transient Cycle. Beginning with the fundamental and most influential turbocharger lag problem, the analysis covers a range of topics, including heat transfer, combustion, air-supply and friction. Diesel Engine Transient Operation presents the most important findings in the field, with special attention paid to the discussion of exhaust emission mechanisms and to the various methods of improving transient response. Moreover, the discussion of the main experimental techniques covers the measurement of exhaust emissions and particle size distribution, which has gained increasing interest in recent years due to stringent regulations imposed by the EU, USA, and Japan. Researchers and students in the field will find this book's comprehensive coverage of the latest research particularly informative, and will also appreciate the authors' analysis of available modeling techniques.
Article
This paper investigates the effects of variable valve actuation on combustion in a Diesel engine. Early inlet valve closing (EIVC) lowered the pressure and temperature during the compression stroke, resulting in a longer ignition delay as the fuel mixed more homogenously with the charge air ahead of combustion. Combustion was characterized by prominent cool flame chemistry and a faster, more energetic, premixed combustion. Tests were performed on a 6.4L V8 engine at loads up to 5 bar BMEP. The use of EIVC showed significant reductions of soot (above 90%) and fuel efficiency improvements (of 5%) with NOx levels below the US 2010 standard of 0.2g/bhp-hr. The improvements in emissions and fuel economy came from controlling in-cylinder temperatures and optimizing combustion phasing. For a constant engine-out NOx emission, EIVC improved fuel economy as the amount of EGR and the engine back pressure requirement were reduced.
Conference Paper
Passive regeneration (oxidation of particulate matter without using an external energy source) of particulate filters in combination with active regeneration is necessary for low load engine operating conditions. For low load conditions, the exhaust gas temperatures are less than 250°C and the PM oxidation rate due to passive regeneration is less than the PM accumulation rate. The objective of this research was to experimentally investigate active regeneration of a catalyzed particulate filter (CPF) using diesel fuel injection in the exhaust gas after the turbocharger and before a diesel oxidation catalyst (DOC) and to collect data for extending the MTU 1-D 2-layer model to include the simulation of active regeneration. The engine used in this study was a 2002 Cummins ISM turbo charged 10.8 L heavy duty diesel engine with cooled EGR. The exhaust after-treatment system consisted of a Johnson Matthey DOC and CPF (a CCRT®). Steady-state loading experiments at 20% load at rated speed were performed for different times in order to achieve three particulate matter loadings of 1.1, 2.2 and 4.1 grams of particulate/liter of filter. Active regeneration was carried out at three CPF-inlet temperatures of 500, 550 and 600°C to cover a range of temperatures and filter loadings for thermal regeneration. The dependent data of fuel usage, time of regeneration, mass of PM oxidized and maximum substrate temperature are presented as a function of mass loading and inlet CPF temperature. The results show that higher CPF-inlet temperature and particulate matter mass loading are more effective for regeneration of the CPF and lower fuel usage in grams of PM oxidized per gallon of fuel used whereas low temperatures and lower mass loadings were not as effective due to lower reaction rates. 90% of the HC from the diesel fuel injection were oxidized across the DOC while the other 10% were oxidized across the CPF under the test conditions.
Article
Fuel efficient thermal management of diesel engine aftertreatment is a significant challenge, particularly during cold start, extended idle, urban driving, and vehicle operation in cold ambient conditions. Aftertreatment systems incorporating NOx-mitigating selective catalytic reduction and diesel oxidation catalysts must reach ∼250 °C to be effective. The primary engine-out condition that affects the ability to keep the aftertreatment components hot is the turbine outlet temperature; however, it is a combination of exhaust flow rate and turbine outlet temperature that impact the warm-up of the aftertreatment components via convective heat transfer. This article demonstrates that cylinder deactivation improves exhaust thermal management during both loaded and lightly loaded idle conditions. Coupling cylinder deactivation with flexible valve motions results in additional benefits during lightly loaded idle operation. Specifically, this article illustrates that at loaded idle, valve motion and fuel injection deactivation in three of the six cylinders enables the following: (1) a turbine outlet temperature increases from ∼190 °C to 310 °C with only a 2% fuel economy penalty compared to the most efficient six-cylinder operation and (2) a 39% reduction in fuel consumption compared to six-cylinder operation achieving the same ∼310 °C turbine out temperature. Similarly, at lightly loaded idle, the combination of valve motion and fuel injection deactivation in three of the six cylinders, intake/exhaust valve throttling, and intake valve closure modulation enables the following: (1) a turbine outlet temperature increases from ∼120 °C to 200 °C with no fuel consumption penalty compared to the most efficient six-cylinder operation and (2) turbine outlet temperatures in excess of 250 °C when internal exhaust gas recirculation is also implemented. These variable valve actuation-based strategies also outperform six-cylinder operation for aftertreatment warm-up at all catalyst bed temperatures. These benefits are primarily realized by reducing the air flow through the engine, directly resulting in higher exhaust temperatures and lower pumping penalties compared to conventional six-cylinder operation. The elevated exhaust temperatures offset exhaust flow reductions, increasing exhaust gas-to-catalyst heat transfer rates, resulting in superior aftertreatment thermal management performance.