23.11.2012 Views

KCFP_Annual_Report_2010.pdf

KCFP_Annual_Report_2010.pdf

KCFP_Annual_Report_2010.pdf

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>KCFP</strong><br />

Kompetenscentrum Förbränningsprocesser<br />

Centre of Competence Combustion Processes<br />

<strong>Annual</strong> <strong>Report</strong> 2010<br />

Faculty of Engineering, LTH<br />

Lund University


<strong>KCFP</strong><br />

Centre of Competence Combustion Processes<br />

The Centre of Competence Combustion Processes, <strong>KCFP</strong>, started July 1 1995.<br />

The main goal of this centre is to better understand the combustion process in internal<br />

combustion engines. Of particular interest are the combustion processes with low enough<br />

temperature to suppress formation of NOx and particulates, PM, often called Low Temperature<br />

Combustion, LTC or Homogeneous Charge Compression Ignition, HCCI.<br />

The Centre of Competence Combustion Processes has a budget of 22.25 MSEK per year.<br />

This is roughly one third each from the Swedish Energy Agency, STEM, Lund University and<br />

the Industry.<br />

CONTENTS<br />

1. The Partially Premixed Combustion Project 6<br />

1.1 PPC - Heavy Duty 6<br />

1.2 PPC - Light Duty 10<br />

1.3 Optical Diagnostics 12<br />

1.4 Combustion Modelling 16<br />

1.5 PPC Fuels 18<br />

1.6 Combustion Control 20<br />

2. The Generic Diesel Project 23<br />

2.1 Analysis of the Initial Soot Formation During an EGR Sweep 23<br />

2.2 Post-Injection Effects on Near-Injector Over-Lean Mixtures<br />

and UHC Emissions 24<br />

2.3 Optical Diagnostics Development 26<br />

3. The Gas Engine Project 30<br />

3<br />

INDUSTRY<br />

PARTNERS<br />

Volvo Cars<br />

Volvo Powertrain<br />

Volvo Penta<br />

Saab<br />

Scania<br />

Toyota<br />

Caterpillar<br />

Cargine<br />

Chevron<br />

Finnveden<br />

Hoerbiger<br />

Wärtsilä<br />

The Swedish Gas<br />

Centre, SGC


Associate<br />

Professor<br />

Per Tunestål<br />

Supervisor for:<br />

CC and SIGE<br />

<strong>KCFP</strong><br />

ORGANISATION<br />

BOARD<br />

Sören Udd, Volvo Powertrain<br />

Börje Grandin, Volvo Cars<br />

Annika Kristoffersson, Saab (Chair)<br />

Jonas Hoffstedt, Scania CV<br />

Professor<br />

Marcus Aldén<br />

Supervisor for:<br />

PPC and<br />

GenDies<br />

Associate<br />

Professor<br />

Mattias Richter<br />

Supervisor for:<br />

PPC and<br />

GenDies<br />

Professor<br />

Rolf Johansson<br />

Supervisor for:<br />

CC<br />

DIRECTOR<br />

Professor<br />

Bengt Johansson<br />

Supervisor for:<br />

PPC<br />

Assistant<br />

Professor<br />

Öivind Andersson<br />

Supervisor for:<br />

GenDies<br />

4<br />

Ulla Holst, LTH<br />

Ida Truedsson, LTH<br />

(student representative)<br />

Professor<br />

Xue-Song Bai<br />

Supervisor for:<br />

CM<br />

ADMINISTRATOR<br />

Maj-Lis Roos<br />

Assistant<br />

Professor<br />

Martin Tunér<br />

Supervisor for:<br />

Fuel<br />

ADMINISTRATOR<br />

Elna Andersson


Vittorio Manente<br />

PhD Student<br />

1. The Partially Premixed Combustion Project<br />

Partially Premixed Combustion, PPC, is a combustion process between Homogeneous<br />

Charge Compression Ignition, HCCI and the classical diffusion controlled diesel combustion.<br />

With PPC it is possible to moderate the charge stratification and thus control the<br />

burn rate better than with HCCI. In comparison to classical diesel combustion the NOx and<br />

particulates can be suppressed with orders of magnitude. <strong>KCFP</strong> has five different but linked<br />

subprojects on PPC.<br />

In the Heavy Duty PPC subproject, the first Ph.D. student, Vittorio Manente, has been running<br />

a Scania D13 truck size diesel engine with PPC using diesel fuel as well as gasoline in<br />

the 69-99 octane range. With gasoline it was possible to run from idle up to 30 bar IMEP<br />

even though only 26 bar was reported. Vittorio presented his thesis in September 2010 and<br />

was replaced by Peter Andersson. Peter has been focusing on ethanol as PPC fuel and initial<br />

tests has shown great potential with much less soot. In The Light Duty PPC project, PPC-LD,<br />

the Ph.D. student Patrick Borgqvist has been working with SI and HCCI reference tests in the<br />

low load regime. He has also designed a combustion chamber that should enable PPC combustion<br />

with the active valve train from Cargine. In the PPC-Fuels subproject the student,<br />

Hadeel Solaka, has been working with fuel and combustion interactions with the same fuels<br />

previously used in the PPC-HD project. Larger than expected differences between Heavy<br />

Duty and Light Duty engine geometries were found. The PPC optical diagnostics have been<br />

in a build-up phase with of two kind of activities 2010. A new Scania D13 optical engine design<br />

has been developed and machined but due to delays in the lab, it has not been up and<br />

running during 2010 but will do so 2011. Due to the delay, the optical diagnostics activities<br />

2010 have been performed in stationary test rigs instead. The PPC modelling subproject<br />

is based on the HCCI and SACI models developed in the previous phase of <strong>KCFP</strong>. The PPC<br />

model with LES started with using gaseous jet modelling and will eventually move to include<br />

also sprays. The fundamental DNS simulations of hydrogen spray and combustion can<br />

give additional insight on the PPC type of combustion. The former controls project has due<br />

to budget constraints been reduced in the current phase of <strong>KCFP</strong>, thus less results can also<br />

be expected. Even so, the PPC control project has done mode switch SI-HCCI-SI using model<br />

based control in the PPC-LD engine having the fully flexible valve system. The developed<br />

control strategies can also be useful for PPC.<br />

1.1. PPC - Heavy Duty<br />

Gasoline PPC Combustion in Scania D13 Engine<br />

During 2010 tests with a Scania D13 engine were performed using various types of gasolines<br />

with a range of octane numbers. Figure 1 shows the resulting operating range as a function<br />

of octane number with fixed low inlet temperature. For low ON it is possible to use PPC<br />

mode of combustion from idle to 26 bar IMEP but with a octane number above 70 low load<br />

becomes a problem. With commercial gasoline with RON in the 95-99 range the low limit s<br />

15 bar. How to enable also low load operation will be investigated in the future.<br />

6


Figure 1: Load range with inlet temperature of 25C.<br />

Figure 2: Emissions as a function of load for two low octane gasoline in comparison to future<br />

emission limits in US and Europe.<br />

7


Peter Andersson<br />

PhD Student<br />

Figure 2 shows that the emission levels with low octane gasoline can be below the legal limits<br />

(except for Pm at higher loads). With some fine tuning and optimization also PM could be<br />

within target.<br />

Tests with modified combustion chamber, spray umbrella angle and swirl levels were performed<br />

in order to repeat the 2008-2009 results in the former Scania D12 geometry. Swirl<br />

was increased to 3.7, spray angle set to 120 degrees and a piston with 14.3 nominal compression<br />

ratio was installed. An improvement in soot-NOx trade-off was found with increased<br />

swirl but the combustion chamber shape was found less suitable with the narrow spray angle<br />

of 120 degrees. A new combustion chamber must thus be designed and evaluated during<br />

2011.<br />

Ethanol PPC Results<br />

Figure 3: Combustion chambers and resulting soot-NOx trade-off<br />

with increased swirl level (HS) at loads 50-100% at A speed.<br />

As with the gasoline PPC the whole load range can be optimized for low emissions and relatively<br />

high efficiency with 40-45% of EGR and low λ using ethanol as fuel. The main diffrence<br />

to gasoline is that no soot is formed


Figure 4: Heat release, injection signal and cylinder pressure for low and medium load ethanol PPC.<br />

Figure 5: Combustion efficiency.<br />

Even though a low lambda can be used with ethanol without producing soot a high level of<br />

EGR is needed to slow down the combustion and prevent excessive the pressure rise rate. For<br />

low loads the EGR level has a minor influence while it is of great interest for middle and high<br />

loads.<br />

Planned activities<br />

For 2011, the current campaign with 95% ethanol is to be finished and investigations with<br />

ethanol and λ = 1 is in the pipeline. Further work is to modify the engine with a new combustion<br />

chamber, i.e. a new piston, to investigate how this influence the efficiency.<br />

9


Patrick Borgqvist<br />

PhD Student<br />

1.2 PPC - Light Duty<br />

Introduction<br />

Gasoline partially premixed combustion is a promising way to achieve simultaneous high efficiency<br />

and low soot and NOx emissions in a diesel engine. Running on gasoline, there is<br />

a potential to get adequate premixing of fuel and air to avoid soot formation. With proper<br />

fuel injection strategy, the combustion is controlled so that excessive heat-release rates are<br />

avoided, which is one of the problems with HCCI combustion at higher loads.<br />

The problem with gasoline partially premixed combustion is the limeted operating region<br />

running with higher octane number fuels. See Figure 1. The purpose of this project is to<br />

increase the attainable operating region and increase the understanding of the combustion<br />

process. The approach is to use a variable valve timing system and different fuel injection<br />

strategies. At a later stage, laser diagnostics will be applied to get increased knowledge and<br />

detailed information of the combustion.<br />

During the first year, an extensive pre-study has been carried out. The purpose is to investigate<br />

and compare spark ignited and HCCI combustion with high octane number fuel (RON95)<br />

using variable valve timings within the target area (circled in the diagram).<br />

Figure 6: Experimental engine and piston crown.<br />

Experimental Setup<br />

The research engine is based on a Volvo D5. It has been modified to run on only one of the five<br />

cylinders, and is equipped with a Cargine fully flexible pneumatic variable valve train system,<br />

a spark plug and both a direct injection and port-fuel injection system. At the first stage, only<br />

the port-fuel injection system has been used. The piston crown had to be modified in order to<br />

get valve clearance around top-dead centre. The piston crown design was inspired by a design<br />

in an article from Toyota [1]. The research engine and piston crown can be seen in Figure 6..<br />

10


Figure 7: Collected rate of heat release curves, operating point 1500 rpm, 3.5 bar IMEP net.<br />

Results<br />

The engine was run port-fuel injected with gasoline (RON95) at engine speeds 1500 rpm and<br />

2000 rpm. Spark ignited and HCCI combustion using different variable valve timings strategies<br />

were compared; spark ignited combustion using conventional valve timings and a throttle to<br />

control load (SI Throttle); spark ignited combustion using late intake valve closing timing (SI<br />

LIVC); spark ignited combustion using early intake valve closing (SI EIVC); HCCI combustion using<br />

negative valve overlap (HCCI NVO); HCCI combustion using retarded closing of the exhaust<br />

valve (HCCI Reb 1); HCCI combustion using a second opening of the exhaust valve late during<br />

the intake stroke (HCCI Reb 2); HCCI with conventional valve timings and pre-heated air. Combustion<br />

timings were optimized for maximum indicated efficiency for each strategy.<br />

The collected rate of heat-release curves from operating point 3.5 bar IMEPnet is seen in Figure<br />

7. The apparent difference between spark ignited combustion and HCCI combustion are<br />

clearly seen. With Partially premixed combustion the goal is to achieve a curve which would<br />

be somewhere in between the spark ignited and HCCI curves. This would be obtained with<br />

a more advanced injection strategy. With a controllable combustion, excessive heat-release<br />

rates can be avoided, and high efficiency with acceptable pressure rise rates can be obtained.<br />

11


Figure 8: Net indicated efficiency, operating point 1500 rpm.<br />

The net indicated efficiencies are summarized in Figure 8. The relative improvement in efficiency<br />

is up to approximately 25% comparing the HCCI Rebreathing strategies with the spark<br />

ignited strategies. The HCCI negative valve overlap strategy suffers mainly from lower gasexchange<br />

efficiency and the pre-heated air HCCI strategy suffers from low combustion efficiency.<br />

References<br />

[1] Improvement of Diesel Engine Performance by Variable Valve Train System, T. Tomoda,<br />

T. Ogawa, H. Ohki, T. Kogo, Dr. K. Nakatani, E. Hashimoto, Toyota Motor Corporation, Japan,<br />

30. Internationales Wiener Motorensymposium 2009<br />

1.3 Optical Diagnostics<br />

Further work on multiple scattering suppression for spray imaging<br />

Structured Laser Illumination Planar Imaging, SLIPI, is a planar imaging technique capable of<br />

suppressing multiple scattering. The principle of the technique is to use a laser sheet which<br />

is spatially modulated in intensity instead of a homogeneous laser sheet. This modulation,<br />

which has a sinusoidal pattern, acts as a ”finger print” allowing the singly scattered photons<br />

to be recognized, as they will scatter accordingly. In contrast, photons which are scattered<br />

several times within the sample before detection appear as a position-dependent intensity<br />

noise on the recorded image and will not follow the modulation pattern. To extract the single<br />

scattering from the multiple light scattering intensity, the method consists in extracting the<br />

amplitude of the modulated component. This does, however, require three modulated images<br />

to be recorded for each event.<br />

Imaging of optical depth<br />

The optical depth describes the turbidity of a sample and is usually measured by illuminating<br />

the medium with a laser beam and to measure the reduction of light in the far field. The<br />

undesired detection of multiple scattering makes imaging of optical depth extremely challenging<br />

and the resulting value is often underestimated using conventional imaging methods.<br />

By combining structured illumination with transmission imaging it is possible to visualize the<br />

optical depth in two dimensions while minimizing the influence from multiple scattered light.<br />

Figure 9 shows an example of such a measurement.<br />

12


Figure 9: Two-dimensional measurements of optical dept in a water spray. The image to the left shows<br />

the results obtained with conventional technique and the image to the right is obtained using SLIPI to<br />

suppress multiply scattered light. Note the underestimation that occurs when scattered light taken into<br />

account.<br />

3D imaging with correction for signal losses<br />

Imaging of turbid media, such as sprays, is challenging not only because of multiple scattering<br />

but also because of losses of both the light from the light source as well as the signal itself.<br />

Figure 10 shows a three-dimensional view of an aerated spray corrected for these losses of<br />

light. This was made possible by measuring the optical depth simultaneously imaging the<br />

sample at different positions.<br />

Figure 10: Three-dimensional view of an aerated water spray that has been corrected for the above<br />

mentioned intensity losses.<br />

13


First test of SLIPI in a diesel spray<br />

The first preliminary test of applying the SLIPI technique to a diesel spray has been conducted.<br />

The tests were performed in a spray chamber with an ambient pressure of 19 bar and the injection<br />

pressure was 600 bar. Commercial diesel fuel was used throughout the test. Although<br />

the conditions were not replicating true in-cylinder conditions, the results are promising for<br />

future applications in engines.<br />

Figure 11: Mie scattering at different times after start of injection (TSOI) recorded with conventional<br />

sheet imaging (above) and SLIPI (below). Injection pressure 600 bar, chamber ambient pressure 19 bar.<br />

Note the significant amount of multiple scattered light in the conventional images.<br />

Wall temperature measurements<br />

As outlined in the previous report there is an interest in measuring the in-cylinder wall temperature.<br />

This is especially the case when running combustion systems with high pressure<br />

rise rates such as HCCI and PPC. There is a possibility that the induced pressure oscillations<br />

can rupture the thermal boundary layer and thus cause a severe increase in heat transfer to<br />

the relatively cold wall. Such phenomena could be investigated using thermographic thermometry.<br />

The surface of interest is coated with a thin layer of a suitable thermo sensitive<br />

material often referred to as a thermographic phosphor. Then the coating is illuminated with<br />

laser radiation and shortly thereafter, when the excited phosphor relaxes, light is emitted. This<br />

resulting radiation from most thermographic phosphors has a duration and spectral characteristics<br />

that is dependent on the temperature. Measurements of the signal decay generally<br />

results in a temperature reading with higher precision compared to corresponding spectral<br />

measurements. Decay times are measured using photo multiplier tubes (PMT) which limits<br />

this approach to point temperature measurements. Spectral investigations are often based<br />

on taking the ratio between two recordings performed at two different wavelengths. This can<br />

be performed using high sensitivity cameras which enable two-dimensional measurements.<br />

These techniques have been applied in engines before both within kcfp and at other research<br />

centers. Thermographic phosphors have proven to be a robust technique capable of delivering<br />

qualitative and quantitative data also when applied in harsh environments such as engines.<br />

However, what has been lacking is detailed knowledge about the precision and accuracy of<br />

the measurement technique.<br />

14


The two main approaches (decay time and spectral) are both in need of thorough studies in<br />

order to reach a level where absolute temperatures can be presented with reliable error bars.<br />

The latter is highly important if small temperature variations on an in-cylinder wall should be<br />

possible to detect in a trustful manner. An extensive effort has been focused on refining the<br />

two-dimensional temperature measurement technique in order to bring it to the desired level.<br />

This, still ongoing, study includes detailed investigations on: reproducibility, spectral ratio<br />

dependence of phosphorescence intensity, laser intensity, detector gain, detector integration<br />

time and phosphor layer thickness. All these parameters must be taken into account if real<br />

quantitative data with known measurement errors should be possible. A fundamental, but<br />

far from trivial, requirement is that the spectral intensity ratio should depend on the surface<br />

temperature only. This means that even if the original measurements performed at two different<br />

wavelengths should show intensity variations due to laser intensity variations or variations<br />

in phosphor concentration, the resulting ration between these two images should only<br />

reflect temperature, se example in Figure 12. So far the investigations have revealed a much<br />

greater complexity than initially expected. However, if carried out carefully the thermographic<br />

phosphor technique shows great potential of reaching a measurement precision of only a few<br />

percent also for 2D temperature probing and that is truly outstanding for a probing technique<br />

with such high temporal resolution.<br />

Figure 12: For the ratio method two images (upper two) are detected through different optical filters and then<br />

the ratio is calculated. This resulting ratio image (above) holds the temperature information. In this particular<br />

case it can clearly be seen how variations in the raw data, caused by laser intensity variations, are canceled out<br />

when the ratio is calculated. This is unfortunately not always the case, and if reliable temperature data is to be<br />

retrieved all the above mentioned parameters must be considered.<br />

15


Rickard Solsjö<br />

PhD Student<br />

Tobias Joelsson<br />

PhD Student<br />

1.4 Combustion Modelling<br />

Introduction<br />

The goal of the PPC modelling project is to use high fidelity simulation methods to simulate<br />

to a rather detailed level the spray injection, the gas jet mixing with ambient EGR and air, and<br />

the ignition/reaction fronts propagation in PPC engines. Modeling tools developed in <strong>KCFP</strong><br />

phase 5 and CeCOST 3 are used for the PPC studies. Specially, large eddy simulation (LES) will<br />

be used, instead of Reynolds averaged Navier-Stokes (RANS) modeling, as LES can capture the<br />

most energetic turbulence structures down to the Taylor micro-scales, and only the sub-scale<br />

turbulence transport for mass, momentum and energy needs to be modeled. This way the<br />

most important phenomena in PPC are simulated properly, such as the vortex shedding and<br />

Kelvin-Helmholtz instability associated with the gas jet and the liquid spray injection, the gas<br />

jet flow interaction with piston wall, etc. For the high speed spray injection study an open<br />

source code OpenFoam is used to take into account the compressibility effect. In addition to<br />

the available tools developed previously, particular flow/chemistry coupling models for PPC<br />

ignition and combustion wave propagation needs to be revisited and validated. This involves<br />

study of PPC using direct numerical simulation (DNS). DNS is developed under generic conditions<br />

for simple fuels such as hydrogen. The modeling project is operated in close collaboration<br />

with the engine experiments and laser diagnostics activities within <strong>KCFP</strong>, to gain maximum<br />

synergy effect to improve the understanding of the fundamental physical and chemical<br />

processes involved in PPC, and to be able to explain phenomena found in the engine tests. The<br />

DNS data is used to validate different modeling concepts, e.g. progress variable based ignition<br />

tabulation approach and multi-zone chemistry mapping (MZCM) approach.<br />

Large eddy simulation of high speed gaseous fuel injection<br />

As a first step we investigate the effect of high speed vapor fuel injection on fuel stratification<br />

and fuel-air mixing, at different injection and ambient conditions. To focus the effort, only the<br />

mixing of fuel and air inside a vessel has been considered. The work is conducted using LES<br />

with an open source CFD-software OpenFoam. To achieve improved understanding on the behavior<br />

of the fuel/air mixing, more attention has been placed on the physics at the early stages<br />

of fuel injection. A constant volume enclosure similar to the combustion vessel developed at<br />

Sandia and Lund University has been studied.<br />

(a)(a) (a)(b) (a)(c)<br />

Figure 13: Instantaneous distribution of mass fraction of fuel in the combustion vessel; (a) 1100 bar<br />

injection pressure and ambient density of 22.8kg/m 3 , (b) 2000 bar injection pressure and ambient density<br />

of 22.8kg/m 3 , (c) 2000 bar injection pressure and ambient density of 40kg/m 3 .<br />

16


Figure 13 shows the effect of injection pressure and ambient density on the mixing of fuel<br />

and the ambient air. With lower injection pressure and lower density (case a) the instability of<br />

the jet develops later and mixing of the fuel with ambient is slower; increasing the injection<br />

pressure leads to an earlier onset of Kelvin-Helmholtz instability and enhanced mixing (case<br />

b); increasing density, however, leads to a delay of instability and slower mixing (case c).<br />

Figure 14. Probability density function (PDF) of fuel mass fraction in the vessel; injection pressure<br />

1100 bar, ambient density 22.8kg/m3 .<br />

The mixing field may be characterized using probability density function of fuel mass fraction<br />

in the vessel. For a homogeneous mixture the PDF shows a spike at the given fuel mass fraction<br />

of the mixture. For a very inhomogeneous mixture the PDF shows a wide distribution<br />

over the mass fraction of fuel. As seen in Figure 14, at later times, 2 ms to 5 ms, the fuel distribution<br />

evolves with time to more homogeneous inside the vessel. This can be seen by the<br />

growing peak at mass fraction value of 0.005 on the fuel mass fraction coordinate. The reason<br />

for the high value of the PDF at fuel mass fraction of zero is the large air rich region in the<br />

vessel. The highest value of mass fraction of fuel in the vessel is shown to decrease with time.<br />

At early times, e.g. 0.5 ms, the fuel is still in the early mixing process; there is a wide interval<br />

of different mass fraction values. A systematic parameter investigation for different injection<br />

conditions including multiple injections of fuel is currently undertaken.<br />

Direct numerical simulation of PPC<br />

To improve the understanding of the structures and mechanisms of PPC ignition and reaction<br />

front propagation, DNS is employed to simulate a H2 jet injected into hot air in a constant volume<br />

vessel of 10x10x20 mm3 . The hot air has a temperature of 1000K and pressure of 30 bar.<br />

The fuel is injected through a 1 mm diameter nozzle at 300K and 100 m/s. Figure 15 shows<br />

the auto-ignition process characterized with the H radicals and fuel distribution at 5 consecutive<br />

instants. The first frame from the left shows the start of ignition where the ignition site<br />

is found at the low left side, slightly downstream of the nozzle. At this instant, the fuel jet is<br />

not disturbed by the heat release; the upper half of the jet flow shows a wave shape due to<br />

flow instabilities. After the onset of the ignition, the ignited zone grows quickly outwards to<br />

both upstream and downstream directions. The fuel jet flow is disturbed by the gas expansion<br />

due to heat release. The intensity of H radicals is highest at the propagation fronts indicating<br />

more intense chemical reactions at the fronts; the ignition fronts propagates upwards at a<br />

speed approximately of 400 m/s along the stoichiometric mixture surface surrounding the<br />

fuel jet. Further DNS simulations are being undertaken for conditions with EGR and pressure/<br />

temperature as well as fuels similar to the PPC experiments to be carried out in KC-FP. The<br />

DNS data are currently being analyzed to find correlation between the onset of ignition and<br />

the local flow and mixture properties, and the nature of reaction front propagation. In turn<br />

the results are being used to validate LES models for simulation of flame and ignition under<br />

PPC conditions.<br />

17


Hadeel Solaka<br />

PhD Student<br />

Figure 15: Instantaneous H radicals (color contour, upper row), H2 (color contour, low row) and vorticity<br />

isolines (white lines) in the symmetric plane of the constant volume combustion vessel. Results<br />

are from DNS of H2/air mixture with injection velocity of 100 m/s at 300 K into hot air with 1000K.<br />

1.5 PPC Fuels<br />

Introduction<br />

The traditional gasoline and diesel engine types have totally dominated the market for internal<br />

combustion engines during the last century. The quantification and research on fuels<br />

has during this time revolved around these engine types, creating indexes and understanding<br />

that do not necessarily apply to other combustion principles in engines. Novel combustion<br />

principles such as HCCI and PPC, principles that show potential for high efficiency and low<br />

emissions operation, require new understanding on fuel properties.<br />

Since most studies show that the oil reserves have started to decrease and that CO2 emissions<br />

affect the climate in a negative way while at the same time the need for transportation is increasing,<br />

the development of combustion engines faces three major challenges: Replacement<br />

of traditional fuels- Low emissions operation – Low cost operation. Engine with high efficiency,<br />

such as the PPC engine, are needed to meet these challenges. To fully exploit the potential<br />

of the PPC engine, new understanding of the impact from fuel properties is a key issue.<br />

Research<br />

The Volvo research engine D5 has been converted to single cylinder operation, see Fig.16. It<br />

is equipped with the new version of fuel system, piezoinjector, which replaced by a solenoid<br />

injector. Replacing the new injector with an old version was due to mechanical problems that<br />

occurred while running with gasoline as a fuel. The injector failed after reaching 800bars in<br />

injection pressure. This failure was due to the fuel properties differences between diesel and<br />

gasoline.<br />

18


Figure 16 shows the research engine Volvo D5.<br />

However, some experiments have been made with the piezoinjector in PPC-mode. These experiments<br />

were run with gasoline (70 RON and 66 MON) and diesel MK1. By using fuels with<br />

high octane number in diesel engine it is possible to achieve high efficiency and low emissions.<br />

Previous research has shown that the heat release rate shape is improved by higher resistance<br />

to auto-ignition which leads to higher efficiency.<br />

With the new modification of the fuel system some experiments have been run with gasoline<br />

69 octane number and diesel.<br />

Results<br />

The engine-out specific emissions of NOx and Soot are shown in Figs. 17 and 18, respectively.<br />

The trends in these emissions from both experiments are quite similar: diesel gave higher<br />

engine-out emissions than gasoline. Although not presented here, engine-out UHC and CO<br />

emissions were also measured. The values for NOx and Soot were low but there is a penalty<br />

from this that both UHC and CO were high. The trade off between Soot and NOx can be seen<br />

in figs. 17 and 18.<br />

180<br />

160<br />

140<br />

120<br />

NOx [ppm]<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 5 7,5<br />

SOC [CAD]<br />

19<br />

Nox-Diesel @ 4IMEP[bar]<br />

Nox-Diesel @ 7.2 IMEP[bar]<br />

Nox-Diesel @10.5IMEP[bar]<br />

Figure 17: Engine out emissions of NOx as function of load and start of combustion.


Anders Widd<br />

PhD Student<br />

Patrick Borgqvist<br />

PhD Student<br />

FSN [-]<br />

2,1<br />

2<br />

1,9<br />

1,8<br />

1,7<br />

1,6<br />

1,5<br />

1,4<br />

1,3<br />

1,2<br />

1,1<br />

1<br />

0,9<br />

0,8<br />

0,7<br />

0,6<br />

0,5<br />

0,4<br />

0,3<br />

0,2<br />

0,1<br />

0<br />

0 5<br />

SOC [CAD]<br />

7,5<br />

20<br />

Soot-Diesel @4IMEP[bar]<br />

Soot-Diesel @7.2IMEP[bar]<br />

Soot-Diesel @10.5IMEP[bar]<br />

Soot-Diesel @13.2IMEP[bar]<br />

Soot-Gasoline @4IMEP[bar]<br />

Soot-Gasoline @7.2IMEP[bar]<br />

Soot-Gasoline @10.5IMEP[bar]<br />

Soot-Gasoline @13.2IMEP[bar]<br />

Figure 18: Engine out emissions of Soot as function of load and start of combustion.<br />

Future work<br />

The coming research is focused on completing the current campaign which is comparing gasoline<br />

fuel with different octane numbers and understanding how this is affecting the main<br />

parameters in combustion (e.g. efficiency, Soot, NOx and maximum pressure rise rate). The<br />

plan is to run 3 different engine speeds (1200, 1500 and 1800rpm) and make a load sweep (4,<br />

6 and 8 bar) for each rpm and make another sweep of start of combustion (SOC) for each load<br />

around (0, 5 and 10 CAD). Start of combustion (SOC) is defined as 1 % of heat release.<br />

1.6 Combustion Control<br />

Introduction<br />

To the purpose of fuel efficiency and low emissions, the focus of this project is model-based<br />

control of PPC. Physics-based models are favored due to the improved portability of the resulting<br />

controllers as well as the added insight into the physical process. During the year, an<br />

experimental study on mode switch between SI and HCCI has been completed. The attainable<br />

high octane number gasoline partially premixed operation region is limited. Even with<br />

advanced injection strategies and variable valve actuation, low load operating conditions may<br />

be unattainable. One option is to apply combustion mode switching to and from spark ignited<br />

combustion. Future and ongoing work involves modeling of PPC combustion to enable<br />

optimization-based methods to be used for steady-state and transient operation.<br />

Experimental Apparatus and Control System<br />

The research engine hardware is a modified Volvo D5 passenger car size engine. The engine<br />

is modified to run on only one of the five cylinders and is equipped with a fully flexible pneumatic<br />

variable valve train system.<br />

The valve timings are the most important parameters for mode switch control. The states<br />

of the combustion and engine actuators are continuously updated at different rates by the<br />

LabVIEW based engine control system. The engine control software is executed on a separate<br />

target-PC, which is a dedicated real-time system, PXI-8110, running the LabVIEW real-time<br />

operating system. The user interface is run on a separate Windows host computer. The host<br />

PC communicates with the dedicated real-time PC over TCP/IP. The front panel of the control<br />

system is shown in Figure 19.


Figure 19: LabVIEW engine control system front panel<br />

Mode Switch Between SI and HCCI<br />

During mode switches between SI and HCCI in early experiments, rather large variations in<br />

IMEP were observed. The goal of this investigation was to reduce these variations and obtain<br />

a smooth transition. It was discovered that a substantial amount of fuel was retained in<br />

the intake port during SI operation, which produced large spikes in IMEP when the effective<br />

compression ratio was changed to one suitable for HCCI operation. The switch was therefore<br />

carried out by disabling the fuel injection during one cycle, and performing the change<br />

in valve timings over 3 cycles. Closed-loop control was then activated to obtain the correct<br />

steady-state combustion phasing and load after the switch. Two different control strategies<br />

were investigated: PI and State Feedback control.<br />

Figure 20 shows a mode switch sequence. The fuel injection is deactivated at cycle 0 while the<br />

valve timing changes take effect. When closed-loop control was used, the PI-controllers were<br />

activated at cycle 4 and the state feedback controllers at cycle 3.<br />

Figure 20: Mode switch sequence showing the in-cylinder pressure, valve movements, and fuel injection<br />

21


Two PI-controllers were implemented, one that governed IMEP using the fuel injection duration<br />

as control signal, and one that governed the combustion phasing using the Negative Valve<br />

Overlap (NVO). The controllers were tuned to produce optimized steady-state performance<br />

around the nominal set points of combustion phasing of 6 degrees after top dead center and<br />

an IMEP of 3.5 bar. Figure 21 shows the response following a mode switch around cycle 80.<br />

There are noticeable oscillations in IMEP. These oscillations could likely be reduced by further<br />

tuning of the controllers to this particular mode switch, but the PI approach is inherently limited<br />

since there is no communication between the two controllers.<br />

Figure 21: PI control following a mode switch showing the outputs IMEP and<br />

CA50 (left axes) and the control signals fuel injection duration and NVO (right<br />

axes).<br />

To enable state feedback control, system identification was used to obtain a model with the<br />

fuel injection duration and NVO as inputs and combustion phasing and IMEP as outputs. The<br />

model was of fourth order and produced a VAF (Variance Accounted For) score of around<br />

70-80 for both outputs in cross-validation tests. The controller was combined with a state<br />

estimator and implemented in the LabVIEW-based engine control system. Figure 22 shows the<br />

response following a mode switch around cycle 60. The response is a lot smoother than that<br />

of the PI controllers.<br />

Figure 22: State feedback control following a mode switch showing the outputs<br />

IMEP and CA50 (left axes) and the control signals fuel injection duration<br />

and NVO (right axes).<br />

22


2. The Generic Diesel Project<br />

Due to its relatively high efficiency, the diesel engine is an important power source for road<br />

transport. Its importance can be expected to increase as the demands on fuel efficiency increase.<br />

However, its exhaust emissions of oxides of nitrogen and particulates remain challenging.<br />

One recent focus of the Generic Diesel (GenDies) project has been to clarify the mechanisms<br />

behind trends in soot emissions during a wide sweep in exhaust gas recirculation (EGR). When<br />

diluting the charge with EGR, soot emissions typically increase drastically. At a certain EGRlevel<br />

they decrease again. These trends have been investigated by for the first time applying<br />

time-resolved laser-induced incandescence (LII) in the cylinder of an optical heavy-duty<br />

diesel engine.<br />

Another focus has been sources of unburned hydrocarbons (UHC) in a low-temperature combustion<br />

mode. The closing flank of the injector has been shown to cause drastically increased<br />

air entrainment into the diesel jet. The resulting over-lean mixtures have been suspected to<br />

contribute significantly to the UHC emissions. An experiment was devised to see if this overlean<br />

mixture can be rendered combustible by a small post-injection. This experiment was<br />

performed in collaboration with Dr. Mark Musculus at Sandia National Labs and showed that,<br />

with the right injection strategy, this is indeed feasible.<br />

2.1 Analysis of the Initial Soot Formation During an EGR Sweep<br />

Introduction<br />

During an Exhaust Gas Recirculation (EGR)-sweep, the engine out smoke level is typically<br />

characterized by an initial increase with decreasing O mole fraction, followed by a drastic<br />

2<br />

decrease below a certain O level. In this investigation, a time-resolved analysis of the soot<br />

2<br />

distribution is made, in a reacting jet, during such a sweep in EGR.<br />

Methodology<br />

A Scania D12 heavy duty diesel engine, converted into an optically accessible single cylinder<br />

engine of Bowditch design, was used in this investigation. A Multi-YAG laser system was<br />

utilized to create time-resolved Laser Induced Incandescence (LII). The test matrix consisted<br />

of a sweep in inlet O mole fraction between 9 and 21 %. The engine was operated at 6 bar<br />

2<br />

indicated mean effective pressure (IMEP). The crank angle of 50% completion of the heat<br />

release and was kept constant at 9 CAD After Top Dead Center over the whole test matrix.<br />

Results<br />

As the O fraction was lowered from 21% to 9%, the engine out smoke level increased from<br />

2<br />

zero filtered smoke number (FSN) up to a peak of 2 FSN, occurring between 11 and 12% O , 2<br />

from where it decreased down to zero again at 9% O . 2<br />

Although it only represents a limited portion of the combustion chamber, the illuminated section<br />

of the jet should capture the important trends during the early phases of soot formation.<br />

As the portion of the laser sheet containing LII signal is related to the volume of the soot cloud<br />

in the jet, the growth rate of the area is thus related to the soot forming rate. Figure 23 shows<br />

the initial soot area growth rate as function of O fraction. The area growth rate is evaluated<br />

2<br />

as the slope of a linear curve-fit to the soot areas during the first 5 CAD with LII signal. The<br />

soot area growth rate increases rapidly as the O fraction is decreased from 21% and peaks<br />

2<br />

at 13% O . As the O fraction is decreased further the soot area growth rate decreases again.<br />

2 2<br />

23<br />

Clément Chartier<br />

PhD Student<br />

Ulf Aronsson<br />

PhD Student<br />

Johan Sjöholm<br />

PhD Student<br />

Joakim Rosell<br />

PhD Student


Figure 23: The left graph shows soot area growth rate during the first 5 images with LII-signal from 10%<br />

to 21% O 2 . The Images to the right show the raw signal and the derived soot area at -1 CAD ATDC for<br />

14% O 2 .<br />

Conclusions<br />

The evolution of the in-cylinder soot distribution during an EGR sweep is consistent with and<br />

helps explain the trends in engine-out smoke:<br />

The initial growth rate of the soot area, in the plane of the laser sheet, increases when the O2 mole fraction decreases from 21% to 13% which is an indication of a greater net soot production<br />

with decreasing O . It is reasonable to assume that the increased net soot production is<br />

2<br />

partially associated with reduced soot oxidation at low O fractions, although this cannot be<br />

2<br />

quantified from the present data.<br />

Below 13% O , past the peak in smoke emissions the area growth decreases rapidly which<br />

2<br />

is explained by a decreased soot formation. As before, the oxidation is probably aggravated<br />

under these conditions, but this is less important when a limited amount of soot is formed.<br />

2.2 Post-Injection Effects on Near-Injector Over-Lean Mixtures<br />

and UHC Emissions<br />

Background<br />

Low temperature combustion (LTC) is a concept aimed at reducing NOx and PM from diesel<br />

engines by dilution of the charge with large amounts of cooled EGR. However, drawbacks in<br />

terms of elevated UHC emissions are known for low load LTC conditions.<br />

Previous studies have shown that over leaning of the jet after end of injection, due to enhanced<br />

air entrainment, results in mixtures too lean to ignite in the near-injector region. This<br />

behavior is especially seen at low loads when the ignition dwell, i.e. time between end of<br />

injection and start of combustion, is large.<br />

The potential was investigated of post-injections to locally increase the equivalence ratio and<br />

render the mixtures ignitable, in order to reduce UHC emissions. Combined laser diagnostics<br />

were used for indirect measurements of mixing. Combustion intermediates were used as indicators<br />

of the local stoichiometry. Formaldehyde can be found in regions ranging from fuel<br />

lean to fuel rich but it persists late in the cycle only in fuel lean regions. Furthermore, OH<br />

concentrations are high in near-stoichiometric regions. Third, PAH is known as a soot precursor<br />

and is found in overall fuel rich regions. Using combined OH and 355nm PLIF, the different<br />

combustion intermediates can be imaged and give information about the local stoichiometry.<br />

Formaldehyde and PAH can be extracted from spectral analysis of the 355-PLIF fluorescence<br />

signal, since formaldehyde has a particular spectrum with seven distinct peaks between 385<br />

and 445 nm.<br />

24


Results<br />

It was found that for a single injection, OH radicals are formed in the downstream regions<br />

close to the bowl wall and stagnate in this location late in the cycle as shown in Figure 24.<br />

Formaldehyde, i.e. lean regions, are found after EOI around the injector in the center of the<br />

bowl as seen in Figure 25. No sign of fuel rich regions were observed for this case.<br />

Figure 24: Evolution of the OH-radicals downstream position relatively to the injector<br />

nozzle in the jet axis plane for the single injection compared to both investigated postinjection<br />

strategies.<br />

A post-injection strategy able to reduce measured exhaust UHC emissions was found to be<br />

achievable, though very sensitive to the settings of the post-injection. A case with 20% decrease<br />

in UHC emissions was found and repeatability tests showed good results. The postinjection<br />

for this case was closely spaced after the actual end of main injection, by one CAD.<br />

Rate of injection measurements showed that the momentum of the post injection is very low<br />

compared to the main injection and that the fuel mass delivered in the post only is small (1.8<br />

mg). Pressure wave dynamics in the injector body and common rail system are suspected to<br />

cause this behavior since a one CAD shift in the start of post-injection, without changing the<br />

energizing time, gave a similar momentum as the main injection.<br />

The small post-injection caused a sudden appearance of OH radicals in the upstream part of<br />

the jet by 10 CAD ATDC, see Figure 25. This observation indicates that second stage ignition<br />

was made possible close to the nozzle by the post-injected fuel, which must have raised the<br />

local equivalence ratio towards stoichiometry.<br />

A second case of post-injection was investigated to show the sensitivity of the local enrichment<br />

mechanism to changes in post injection strategy. The post-injection duration was kept<br />

constant but its start was delayed by one CAD. Exhaust measurements showed no influence<br />

of the post on UHC emissions for this case compared to the single injection. Furthermore, no<br />

OH radicals were seen closer to the injector. Instead the same trend as for the single injection<br />

was observed, see Figure 25.<br />

25


Figure 25: Simultaneous OH (green) and 355 PLIF (red) images. The left frame is extracted from a sequence<br />

with the single injection whereas the right one is from the small post injection strategy. The<br />

timings are given relative to TDC and to the main EOI. The formaldehyde correlation bar indicates the<br />

similitude of the 355 PLIF signal to a formaldehyde reference spectrum. The darkest colors in the correlation<br />

bar indicate a low correlation whereas the brightest colors indicate that the LIF signal has a<br />

high probability of being formaldehyde-LIF signal.<br />

The fuel lean regions in the center of the bowl remained lean, correlating with the unchanged<br />

UHC emissions. Rate of injection measurements showed that post-injected fuel mass was nine<br />

times higher than the small post case and the momentum was similar to the main injection.<br />

Conclusions<br />

A comparison of the two post-injection strategies suggests that the momentum is crucial.<br />

The post injection momentum should be low enough for the penetration to remain relatively<br />

short. The post-injected fuel should thereby remain in the near-injector region to enrich the<br />

over-mixed zones. Higher momentum in the post would carry the post-injected fuel too far<br />

downstream. The discovery of the small post-injection strategy leading to exhaust UHC reduction<br />

is the result of a trial and error process on the engine. This strategy can be seen as using<br />

pressure dynamics in the fuel system to achieve such a small post injection. Injector dynamics<br />

is not a controlled variable and the transfer of this particular injection strategy to another<br />

engine might not be straightforward. However, the working principle of the local enrichment<br />

by the post is valid and is the major novelty of this study.<br />

2.3 Optical Diagnostics Development<br />

New Ph.D. student in laser/optical diagnostics<br />

During the fall the Ph.D. student Johan Sjöholm, who has been involved with the Gendies<br />

program for almost five years, went on parental leave. At the same time a new Ph.D. student<br />

was recruited. In late August Joakim Rosell joined the group. He has studied physics at Lund<br />

University. After taking the master exam in Physics he studied pedagogics and thereafter he<br />

has been working as a teacher both in Sweden and abroad. Joakim will be focusing on highspeed<br />

diagnostics, mainly using the unique multi-YAG system, which is very well in line with<br />

the needs of kcfp. Initially Joakim worked along with Johan Sjöholm in order to learn as much<br />

as possible about the highly specialized equipment before Johan went on parental leave. During<br />

the winter Joakim has focused on a feasibility study where the multi-YAG and an OPO-laser<br />

are combined in order to target CO molecules. Joakim’s work includes both setting up the<br />

laser hardware and performing test on CO in a controlled environment. If proven successful,<br />

this approach will be used within in the GenDies program, for detection of partially oxidized<br />

zones, in collaboration with visiting professor Paul Miles from Sandia National Labs.<br />

26


Probing of CO molecules<br />

Carbon monoxide, CO, is not only a major pollutant but can also be used as an indicator of<br />

combustion efficiency, since incomplete oxidation of hydrocarbons result in increased concentration<br />

of CO in the burnt gas region. It would thus be useful if this important species could be<br />

monitored in combustion processes. In the ongoing feasibility study CO is probed by exciting<br />

the molecule from its ground state to the excited state and then observing the fluorescence<br />

that occurs when the molecule relaxes from the excited level to a lower. What is special about<br />

laser induced fluorescence of CO is that it requires two-photon excitation in the UV (230.1<br />

nm). This means that two photons must be absorbed for the excitation to take place. This is<br />

in contrast to conventional laser induced fluorescence where only one photon is needed. The<br />

reason for applying two-photon excitation in this case is that the single-photon transitions in<br />

CO occur at wavelengths so short that the light would suffer from severe absorption when<br />

transmitted in air. There are several issues with two-photon excitation. The most obvious are<br />

the reduced signal strength and quadratic dependence on laser power, which means that high<br />

power is required. A perhaps trivial but not less important task is to investigate how strong<br />

CO-signals the new laser system can generate. Another important issue that needs careful<br />

consideration is the interference from C -molecules. Spectrally CO and C can overlap, i.e. they<br />

2 2<br />

produce fluorescence of the same color. C can be formed naturally during combustion but it<br />

2<br />

can also be formed by photo decomposition of the fuel, i.e. introduced by the laser light. In<br />

order to tell these two species apart it must be possible to distinguish CO-fluorescence from<br />

C -fluorescence. There are strategies for this that are based on careful selection of the spec-<br />

2<br />

tral range where the detection is performed. The drawback of this approach is a significantly<br />

reduced detectability, remember that the CO-signal is generally weak from the beginning and<br />

further filtering worsens this. Quite recently another excitation scheme for CO, where laser<br />

light at 217.5 nm is being employed (instead of 230 nm) was proposed. The amount of presented<br />

work using this alternative scheme is very limited and Joakim Rosell is currently performing<br />

comparisons of the two approaches. The potential advantage with the less applied 217.5 nm<br />

excitation is a reduced interference from C . 2<br />

Figure 26: Spectra from CO molecules after two-photon excitation.<br />

27


Implications of applying LII in engine environments<br />

Laser-Induced Incandescence (LII) has traditionally been considered a straightforward and reliable<br />

optical diagnostic technique for in-cylinder soot measurements. As a result, it is nowadays<br />

even possible to buy turn-key LII measurement systems. During recent years, however,<br />

attention has been drawn to a number of unresolved challenges with LII. Many of these are<br />

especially evident in regions with high soot concentrations typically found in combustion engines.<br />

During the work of performing high-repetition rate measurements with LII in diesel<br />

engines several specific challenges were identified. As a consequence, a substantial part of the<br />

high-speed soot measurement campaign performed within the Gendies program was focused<br />

on identifying and understanding the various issues associated with the application of LII in<br />

engine environments in general and at high-speed in particular. This part of the work resulted<br />

in a dedicated technical paper, on the application of high-speed LII in engines. This paper is<br />

accepted for publication at SAE World Congress in 2011. All the identified issues might not<br />

be possible to overcome but they should nevertheless be known and their potential impact<br />

should be considered when interpreting the results.<br />

In short, the LII signal depends on laser fluence. Ideally it is desired to have a plateau on the<br />

fluence curve (LII-signal vs. laser fluence), i.e. a region where the LII signal is independent on<br />

the laser fluence. This means that variations in laser energy do not influence the measured<br />

soot amount. In practice the shape of the fluence curve is highly dependent on the laser beam<br />

profile. A Gaussian profile will produce a LII-signal that continuously increases with laser fluence<br />

and thus, does not show any plateau region. A perfect top-hat profile will introduce severe<br />

sublimation and hence a weaker signal with increasing fluence. In Figure 27, the fluence<br />

curves are shown for different beam profiles generated by altering the sheet forming optics.<br />

Figure 27: The LII-signal versus laser fluence for different laser beam profiles. The<br />

green curve represents a close to top-hat shaped profile that causes severe sublimation<br />

of the soot particles. The blue curve represents a profile with more spatial<br />

wings, i.e. closer to a Gaussian shape, which generates the desired plateau region<br />

where the LII-signal is independent of the laser fluence.<br />

28


In the initial feasibility study, high-speed LII measurements were made in a laminar ethylene/<br />

air flame to test the characteristic features of high-speed LII in a well defined environment<br />

before approaching the engine. Excessive influence of soot sublimation (damages of the soot<br />

caused by the laser) was observed on the LII signals detected for later pulses in the pulse train<br />

when running with pulse separations relevant for engine conditions, see Figure 28<br />

Figure 28: Normalized LII signal from the second of two laser pulses sent through an atmospheric<br />

flame as a function of pulse separation. The effect of soot sublimation is seen for<br />

separations shorter than 20 ms.<br />

Fortunately, corresponding data from the following in-cylinder measurements does not show<br />

signs of the same behavior, which is believed to be due to high mixing and/or very fast soot<br />

formation/oxidation in the measurement volume. This implies that time-resolved LII is a technique<br />

applicable for qualitatively following the evolution of the soot distribution inside a reacting<br />

diesel jet during a single combustion event.<br />

29


Mehrzad Kaiadi<br />

PhD Student<br />

3. The Gas Engine Project<br />

Introduction<br />

Most heavy-duty engines are diesel operated. Severe emission regulations, high fuel prices,<br />

high technology costs (e.g. catalysts, fuel injection systems) and unsustainably in supplying<br />

fuel are enough reasons to convince engine developers to explore alternative<br />

technologies or fuels. Using natural gas/biogas can be a very good alternative due to the<br />

attractive fuel properties regarding emission reduction and engine operation. Heavy-duty<br />

diesel engines can be easily converted for natural gas operation which is a very cost effective<br />

process for producing gas engines. However, due to the high throttle losses and low<br />

expansion ratio the overall engine efficiency is lower than for the corresponding diesel engines.<br />

Moreover the lower density of natural gas results in lower maximum power level.<br />

The existing project is the fourth phase of the natural gas engine project at Lund University.<br />

The previous results show that the stoichiometric operation is a better choice than lean operation<br />

since by using a 3-way catalyst the emissions level can be kept at very low levels and<br />

differences in efficiency are not significant. Based on the previous results the engine operates<br />

stoichiometric and the project focuses mostly on the problems associated with this type of<br />

operation.<br />

Extending the dilution limit of the engine and developing closed-loop control to operate the<br />

engine at its dilution limit has been the main method to reduce throttle losses. New methods<br />

to determine combustion stability during transients have been developed and used together<br />

with model based control methods. New modifications are also performed on EGR and turbocharging<br />

system of the engine to increase the overall efficiency and extend the maximum<br />

load limit.<br />

Objectives<br />

The main objectives of the project during 2010 have been to identify and apply new hardware/strategies<br />

to improve the overall engine efficiency and also extend the maximum load<br />

level of the engine. A summary of the results achieved during 2010 are presented in this<br />

report.<br />

Experimental Results<br />

Some modifications have been performed on the engine in order to improve the efficiency<br />

and extend the maximum load limit. The main modifications were performed on the pistons,<br />

the turbocharging system and the EGR system.<br />

As already mentioned in last year’s report the major modifications on the pistons are<br />

the shape and compression volume. The new pistons generate very high turbulence<br />

and increase the compression ratio from 10.5 to 12. The results show roughly 40% reduction<br />

in combustion duration and at least 2 percent units improvement in gross indicated<br />

efficiency. The new combustion chambers are called “Quartette” in the figures.<br />

The new piston modification results in some changes of the exhaust gas characteristics<br />

and since the engine is equipped with a turbocharger with wastegate, the changes in exhaust<br />

gas characteristics may have direct influence on the boost pressure. The absolute<br />

boost pressure with Quartette pistons is lower which results in lower maximum BMEP.<br />

30


Higher expansion ratio means that more heat is converted into mechanical work in the cylinder<br />

and less energy in the form of exhaust gas enthalpy is available. According to the manufacturer<br />

the exhaust gas temperature should be lower than 760oC. The lower exhaust gas<br />

temperature obtained with higher compression ratio is beneficial in two ways. First, there is a<br />

higher margin to the highest allowable exhaust gas temperature and second, there is potential<br />

to extend the maximum load limit by replacing the turbocharger with a better matched unit.<br />

A Variable Geometry Turbocharger (VGT) has the ability to change the turbine geometry in<br />

order to obtain the desired boost pressure. VGT offers attractive properties such as high flexibility,<br />

minimal amount of lag and wide operating range. VGT technology is extensively used in<br />

diesel engines; however the use of VGT is very much ignored in gasoline engines due to their<br />

relatively high exhaust temperatures. Ordinary VGT materials and designs cannot withstand<br />

temperatures over 890°C. The exhaust gas temperature of gasoline engines could, however,<br />

reach up to 1000°C, versus 650°C in diesel engines.<br />

Normally natural gas engines use the same combustion technology as gasoline engines due<br />

to similar fuel properties. In heavy-duty applications, since the engine speed does not exceed<br />

2000 RPM, the exhaust gas temperature is not very high. After consulting with the R&D group<br />

at Volvo a VGT was designed and mounted on the engine. The engine is operated at the same<br />

engine speeds with WOT and by altering the geometry of the turbine housing the boost was<br />

increased until either knock occurred or the pumping losses started to increase. To suppress<br />

knock, EGR was added and the ignition timing was adjusted accordingly. Figure 29 shows the<br />

boost pressure achieved by the VGT in comparison with the by-pass turbocharger with original<br />

and Quartette pistons. The boost pressure was increased up to much higher levels with<br />

VGT without sacrificing gas-exchange efficiency. The increase in boost pressure resulted in<br />

increased maximum load as presented in Figure 30 compared to previous configurations. The<br />

peak load increased by 18 percent from 16 to 19 bar BMEP.<br />

Figure 29 Boost pressure achieved after replacing the turbocharger with a VGT<br />

31


Figure 30 Maximum load achieved after replacing the turbocharger with a<br />

VGT<br />

Using a throttle always leads to pumping losses. Normally a throttle is used to control the desired<br />

torque in the engines, but it is also possible to use a VGT instead of the throttle in a large<br />

operating range to control the desired torque. This is possible due to the flexibility of the VGT<br />

to adjust the inlet pressure by altering the geometry of the turbine housing. In this operation<br />

region the throttle is kept fully open and the VGT is used to adjust the inlet pressure and finally<br />

control the demanded torque. This means that the throttle will not be used in a large part of<br />

the operating region (roughly 60% of the entire operating area of the engine) and no throttle<br />

use means no throttle losses. The results have been validated in a large operating region and<br />

showed that at least 2 percent units improvement in gas-exchange efficiency is gained. The<br />

dynamics of VGT was also comparable to the throttle dynamics. An innovative driving strategy<br />

with lowest possible throttle losses has been suggested. In summary the key features to<br />

improve the natural gas engine performance are identified as: Right amount of EGR at different<br />

operating regions, Right compression ratio, Variable Geometry Turbocharger (VGT), High<br />

turbulence pistons, Long route EGR system and model-based control.<br />

Planned Activities<br />

Mehrzad Kaiadi will defend his doctoral thesis 2011-02-11. After that the project will focus on<br />

dual fuel operation using direct injection of Diesel fuel for ignition.<br />

32


<strong>KCFP</strong><br />

Kompetenscentrum Förbränningsprocesser<br />

Centre of Competence Combustion Processes<br />

Faculty of Engineering, LTH<br />

P.O. Box 118<br />

SE-221 00 Lund<br />

Sweden

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!