25.07.2013 Views

Investigation of Two-Photon Laser-Induced Fluorescence Detection ...

Investigation of Two-Photon Laser-Induced Fluorescence Detection ...

Investigation of Two-Photon Laser-Induced Fluorescence Detection ...

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>Investigation</strong> <strong>of</strong> <strong>Two</strong>-<strong>Photon</strong> <strong>Laser</strong>-<strong>Induced</strong> <strong>Fluorescence</strong> <strong>Detection</strong> <strong>of</strong><br />

Carbon Monoxide for Applied Combustion Diagnostics<br />

M. Richter, Z. S. Li * and M. Aldén<br />

Division <strong>of</strong> Combustion Physics, LTH, P.O. Box 118<br />

Lund University, S-221 00 Lund, Sweden<br />

Abstract<br />

The properties <strong>of</strong> two-photon laser-induced fluorescence (LIF) signals from carbon monoxide have been studied in<br />

order to investigate the applicability <strong>of</strong> this technique for measurement and visualization <strong>of</strong> CO molecules in applied<br />

combustion environments. Spectrally and spatially resolved measurements were carried out in a test cell, in a flame<br />

and in an engine. In the cell measurements, the two-photon fluorescence signal from CO was studied as a function<br />

<strong>of</strong> laser flux and ambient pressure. In addition, measurements with buffer gas at pressures from 0.1 up to 10 bars<br />

were carried out. In the flame measurements, problems from spectral interference due to undesired C2 molecules<br />

produced by photodecomposition <strong>of</strong> fuel and fuel fragments were investigated at different stoichiometries. <strong>Two</strong><br />

dimensional LIF imaging <strong>of</strong> CO was also performed in an engine. The interference from C2 and photolysis <strong>of</strong> hot<br />

CO2 was excluded from the spectral and spatial resolved measurements.<br />

.<br />

Introduction<br />

<strong>Detection</strong> <strong>of</strong> Carbon monoxide has always been a<br />

important task for laser combustion diagnostics.<br />

Various laser diagnostic techniques have been<br />

developed during the last decades and they have<br />

become <strong>of</strong> utmost importance for a deepened<br />

understanding <strong>of</strong> combustion processes. This<br />

statement is to a large degree due to the outstanding<br />

features like non-intrusiveness in combination with<br />

high temporal and spatial resolution for<br />

measurements <strong>of</strong> species concentrations,<br />

temperatures, velocities and particle characteristics,<br />

see e.g. [1-3]. The technique that has attracted most<br />

attention is <strong>Laser</strong>-<strong>Induced</strong> <strong>Fluorescence</strong> (LIF), see<br />

e.g [4]. Most <strong>of</strong> the combustion diagnostic<br />

experiments using LIF have been directed towards<br />

single photon excitation. There are, however, several<br />

species <strong>of</strong> combustion interest that do not have easily<br />

accessible excitation wavelengths in the UV/visible<br />

part <strong>of</strong> the spectrum, e.g. the light atoms O, N, H, C<br />

and molecules like H2, H2O, NH3, CO. Many <strong>of</strong> these<br />

species have their main electronic resonances in the<br />

vacuum ultraviolet region <strong>of</strong> the electromagnetic<br />

spectrum. In order to overcome the problems <strong>of</strong><br />

generating and using VUV laser beam and the<br />

problems associated with excitation in this spectral<br />

region, multi-photon excitation processes have been<br />

utilized for the probing <strong>of</strong> these species.<br />

One <strong>of</strong> the species mentioned above, CO, as an<br />

indicator <strong>of</strong> the combustion efficiency and as a major<br />

pollutant, is one <strong>of</strong> the most important molecules in<br />

combustion processes. In combustion environments<br />

CO has been measured commonly by exciting the<br />

molecule from its ground state, X 1 Σ + , to the excited<br />

* Corresponding author: zhongshan.li@forbrf.lth.se<br />

Associated Web site: http://spartacus.forbrf.lth.se/<br />

Proceedings <strong>of</strong> the European Combustion Meeting 2005<br />

state, B 1 Σ + , via two-photon excitation at 230.1 nm<br />

and observing the fluorescence from the excited level<br />

to the lower lying A 1 Π level by detecting the<br />

fluorescence in the spectral range between 451 nm<br />

(v’=0, v’’=0) and 662 nm (v’=0, v’’=5), see Figure 1<br />

where the collision induced triplet transitions are also<br />

indicated. Using this excitation scheme several flame<br />

measurements have been performed [5-9]. Quenching<br />

measurements in flames using pico-second excitation<br />

[10, 11] have also been reported. In order to achieve<br />

a quantitative interpretation <strong>of</strong> the two-photon LIF<br />

signal, substantial efforts have been put in measuring<br />

important parameters such as the two-photon<br />

excitation cross-sections [12], photoionization crosssections<br />

and ac Stark shift [13], collisional<br />

broadening and shift [14], species-specific and<br />

temperature-dependent quenching cross-sections [15,<br />

16]. Photochemical effects from vibrational hot CO2<br />

[17] has been investigated as an interference for twophoton<br />

LIF CO detection in atmospheric pressure<br />

flames. Among other alternative spatial-resolved CO<br />

detection techniques, Linow et al. reported the<br />

comparison between the excitation scheme<br />

mentioned above and excitation from the ground<br />

state, X 1 Σ + , to the excited state, C 1 Σ for two photon<br />

LIF measurements [18]; infrared LIF detection <strong>of</strong> CO<br />

has been reported by Hanson’s group [19, 20], by<br />

probing infrared active ro-vibrational overtone bands;<br />

two-photon polarization spectroscopic was<br />

investigated by Nyholm et al. for spatial resolved CO<br />

detection [21].<br />

Despite the relatively large number <strong>of</strong><br />

publications including flame measurements <strong>of</strong> CO<br />

detection using two-photon LIF, so far to our


knowledge no investigation at high pressure and<br />

engine applications have been reported. The aim <strong>of</strong><br />

the present work was to investigate the applicability<br />

<strong>of</strong> CO detection with two-photon LIF in combustion<br />

engine. In order to understand the behaviour and<br />

achieve a quantitative interpretation <strong>of</strong> the CO LIF<br />

signal from the combustion engine, where both high<br />

pressure and high temperature presented, a pressure<br />

dependent measurement in a static cell and a flame<br />

measurement with different stoichiometries have<br />

been performed. Single shot, two dimensional CO<br />

LIF imaging in a test engine was finally<br />

demonstrated with exclusion <strong>of</strong> the potential<br />

interference from C2 and from photolysis <strong>of</strong> hot CO2.<br />

Figure 1. Schematic energy level diagram for<br />

carbon monoxide. Depicted also the two-photon<br />

excitation, the fluorescence decay and the collisioninduced<br />

triplet state transitions.<br />

Cell and flame measurements<br />

In order to investigate the potential for<br />

quantitative LIF measurements <strong>of</strong> CO in general and<br />

at elevated pressure in specific, it is necessary to<br />

control a range <strong>of</strong> properties affecting the relation<br />

between laser power and the generated signal level<br />

and the signal dependences on pressure and possible<br />

photolytic creation <strong>of</strong> C2 radicals. When probing CO,<br />

in addition to the more conventional quenching<br />

phenomenon (marked with a Q in Figure 1),<br />

properties like two-photon absorption, stimulated<br />

emission and ionization effects have to be taken into<br />

account, which are all dependent on the laser<br />

intensity. The laboratory experiments served to<br />

investigate these properties in simulated combustion<br />

environments, i.e. to investigate the dependencies on<br />

laser energy, pressure, temperature and<br />

stoichiometry, as well as, the influence from the<br />

presence <strong>of</strong> other species.<br />

A Nd:YAG (Continuum NY82-S)) pump dye<br />

laser (continuum ND60) system are utilized in the<br />

cell and the flame measurements. The dye laser (with<br />

a mixture <strong>of</strong> Rhodamine 590 and Rhodamine 610<br />

dissolved in methanol) was pumped by the second<br />

2<br />

harmonic from the Nd:YAG laser, producing<br />

radiation at 587.2 nm. Frequency doubling in a KDPcrystal<br />

resulted in 293.6 nm. By mixing this with the<br />

fundamental 1064 nm from the Nd:YAG laser,<br />

approximately 4.5 mJ <strong>of</strong> the required radiation at<br />

230.1 nm was generated.<br />

The investigated gas mixtures were kept in a<br />

stainless steel high pressure vessel equipped with<br />

quartz windows. The laser beam was focused with a<br />

single spherical lens through the cell with the focus<br />

point centred in the middle. The interaction region<br />

between the CO molecules and the laser beam was<br />

imaged with an image intensified CCD (ICCD)<br />

camera (Princeton Instruments, ICCD-5765S)<br />

perpendicular to the excitation laser beam. A long<br />

pass filter was placed in front <strong>of</strong> the ICCD to block<br />

the scattered light from the 230 nm excitation laser.<br />

Figure 2. Plot showing the CO LIF signal measured<br />

at twelve different pressures ranging from 0.1 to 3.0<br />

bar.<br />

Shown in Figure 2 is the resulting CO LIF signal<br />

versus pressure when the cell was filled with pure<br />

carbon monoxide and the pressure varied from 0.1<br />

bar to 3 bar. The laser was propagating from left to<br />

right and the focus point located at 0 mm. The<br />

absorption effect <strong>of</strong> the excitation laser as it<br />

transverse the cell was obvious as indicated by the<br />

decrease <strong>of</strong> the LIF signal along the laser beam and<br />

the absorption increase with the increased CO<br />

pressure in the cell. Note that even for a fix laser<br />

energy output the laser flux density will vary along<br />

the focal line, i.e. it will decrease with the distance to<br />

focus. As can be seen this results in a variation <strong>of</strong> the<br />

fluorescence intensity along the beam. The ionization<br />

potential <strong>of</strong> the CO molecule is 14.0139 eV. This<br />

gives that when the B 1 Σ + level is excited, absorption<br />

<strong>of</strong> one additional photon would be sufficient to ionize<br />

the molecule. The dip at the focus location is<br />

probably due to the increased influence <strong>of</strong> photo<br />

ionization. To understand this better, spatial resolved<br />

CO LIF measurement with 400 mbar pure CO in the<br />

cell was performed with varied excitation laser pulse<br />

energy. Shown in Figure 3 is an example <strong>of</strong> such a<br />

measurement with the laser pulse energy varied from<br />

0.2 to 2.4 mJ. The CO LIF emission is highly


dependent on several parameters, like laser intensity,<br />

ionization, non-linear absorption which all contribute<br />

to the question <strong>of</strong> correct signal interpretation in an<br />

applied combustion situation. In order to understand<br />

the phenomena taking place and the interaction<br />

between these phenomena a simulation based on a<br />

four level rate equation system including photoionisation,<br />

quenching, two-photon absorption, and<br />

spontaneous emission was applied to calculate the<br />

spatial distribution <strong>of</strong> the CO LIF signal [22]. The<br />

thick line shown in Figure 3 is a simulation <strong>of</strong> the<br />

measured curves with 2.4 mJ per pulse the laser<br />

intensity. A general similarity between the simulated<br />

pr<strong>of</strong>ile and the measured one was achieved, which<br />

indicates that the model are proper in prescribe the<br />

observed the phenomenon.<br />

Figure 3. The variation <strong>of</strong> the CO LIF intensity along<br />

the laser beam. The gas pressure <strong>of</strong> CO in the cell was<br />

held constant at 400 mbar, while the energy <strong>of</strong> the laser<br />

pulse was varied from 0.2 to 2.4 mJ. The thick line<br />

represents a simulated fluorescence signal along the<br />

laser beam with 2.4 mJ per pulse.<br />

Signal intensity (counts)<br />

10 4<br />

10 3<br />

k = 2<br />

k = 1<br />

0.1 bar<br />

0.2 bar<br />

0.5 bar<br />

1.0 bar<br />

2.0 bar<br />

3.0 bar<br />

106 107 108 109 <strong>Laser</strong> fluence (W/cm2) Figure 4. Plot showing the measured LIF<br />

signal vs excitation laser fluxes at different CO<br />

pressures.<br />

Shown in Figure 4 are the LIF signal intensities<br />

as the laser energy was increased for different<br />

pressures <strong>of</strong> CO in the cell. Normally, the<br />

3<br />

fluorescence signal, S, in a two-photon excitation<br />

processes depends quadratically on the incident laser<br />

intensity, I, i.e. S ∼ I k , where k = 2. However, at<br />

higher laser energies photo-ionization <strong>of</strong> the CO<br />

molecule becomes significant. Moreover, the<br />

transition starts to saturate and also at higher laser<br />

energy two-photon absorption decreases the signal<br />

intensity. Hence, the fluorescence signal scales with k<br />

= 2 in the low laser intensity limit, and turns into k is<br />

close to one when the laser intensity is increased. If<br />

the laser intensity is further increased the signal goes<br />

through a maximum until the signal actually<br />

decreases as the laser intensity is increased further.<br />

Figure 5. LIF signal dependence on the buffer N2 pressures (100 mbar CO, N2 pressure varied).<br />

In order to estimate the delectability <strong>of</strong> using<br />

two-photon LIF at elevated pressure, e.g. in engines,<br />

measurements in the cell with and without buffer<br />

gases were carried out at pressures from 0.1 bar up to<br />

8 bar. It is more realistic to investigate the CO LIF<br />

signal intensity while adding a buffer gas to a<br />

constant pressure <strong>of</strong> CO in the cell. Shown in Figure<br />

5 is the signal distribution from CO at a pressure <strong>of</strong><br />

100 mbar while the pressure <strong>of</strong> the buffer gas, N2,<br />

was varied from 0 to 10 bar. As expected the signal<br />

decreases as the pressure is increased. Also the shape<br />

<strong>of</strong> the signal distribution changes as the pressure<br />

increases.<br />

In the flame investigations, the behaviour <strong>of</strong><br />

spectral interference from non-resonantly excited C2<br />

molecules, produced by photo-decomposition <strong>of</strong> fuel<br />

and fuel fragments, were studied at various<br />

stoichiometries and laser intensities. The first flame<br />

measurement was performed in a laminar, premixed,<br />

methane/air flame. During the flame measurements<br />

the cell was replaced by a premixed Bunsen-type<br />

conical burner, which has a prolonged intake duct in<br />

order to assure a laminar flow. Through a<br />

spectrometer (Acton, SpectraPro-150) equipped the<br />

ICCD camera, a spectral resolved CO LIF spectrum,<br />

as shown in Figure 6, was recorded for a Φ = 1.5<br />

methane/air flame with 50 laser shots averaged. As<br />

shown in the Figure, some C2 interference lines are<br />

clearly shown.


C 2 interferences<br />

(c 1 Π g – b 1 Π u)<br />

(d 3 Π g – a 3 Π u)<br />

-7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7<br />

Distance (mm)<br />

400<br />

450<br />

500<br />

550<br />

600<br />

650<br />

Wavelength (nm)<br />

Figure 6. A CO fluorescence spectrum from a<br />

methane/air (φ=1.5) burner is shown above. The C 2<br />

lines around 474 nm and 520 nm are visible.<br />

Engine measurements<br />

The measurements in the cell and the premixed<br />

flame were followed by measurements in a small<br />

single-cylinder four-stroke Briggs & Stratton engine<br />

[23]. The engine featured a side-valve design with the<br />

spark plug located above the exhaust valve. The<br />

original crankcase, cylinder liner and piston remained<br />

unaltered, while the cylinder head was modified to<br />

allow for optical access to the combustion chamber.<br />

Three windows were mounted in the cylinder head<br />

for this purpose. <strong>Two</strong> vertical windows, one on the<br />

side <strong>of</strong> the combustion chamber, made it possible to<br />

have a horizontal laser sheet passing through it. The<br />

induced fluorescence was imaged through a third<br />

window mounted horizontally above the two valves.<br />

A picture taken from above, through the top window,<br />

is shown in Figure 7. In this image the intake valve<br />

can be seen to the left and the exhaust valve to the<br />

right. The spark plug electrodes are also visible in the<br />

lower right corner. The field <strong>of</strong> view for the ICCD<br />

camera is also shown in this figure, where the outer<br />

white frame illustrates the area covered during the<br />

flame chemiluminescence imaging and the inner<br />

frame illustrates the area covered during the CO LIF<br />

imaging.<br />

Isooctane was used as the fuel for the engine<br />

measurements. Shown in Figure 8 are single-shot<br />

images <strong>of</strong> the flame chemiluminescence. These<br />

images are recorded at subsequent crank angle<br />

positions to cover the different parts <strong>of</strong> the engine<br />

cycle. The exposure time for the ICCD camera was<br />

set to 5 µs, which corresponded to a crank shaft<br />

rotation <strong>of</strong> less than 0.04 crank angle degrees (CAD).<br />

At Top Dead Center (TDC), a developing flame<br />

kernel originating from the vicinity <strong>of</strong> the spark plug<br />

can be seen; 5 CAD after TDC, the flame is more<br />

spread out from the spark plug; the flame then<br />

continues to expand as expected and as shown by the<br />

images recorded at 15 and 30 CAD after TDC.<br />

4<br />

Figure 7. Vision <strong>of</strong> the engine chamber from the top<br />

window. Outer write frame indicate the chemiluminescence<br />

collection area; inner write frame<br />

indicates the LIF imaging area.<br />

300 600 900 400 1000 1800 2400<br />

TDC<br />

15 TDC<br />

5 TDC 30 TDC<br />

Figure 8. Flame chemiluminescence emission taken<br />

with 5 microsecond gate time.<br />

During the engine LIF experiments a singlemode<br />

Nd:YAG laser (Spectra Physics, PRO 290-10)<br />

pumping a OPO laser system (Spectra Physics,<br />

MOPO 730-10) was utilized to produce the required<br />

230 nm laser radiation. When pumped with 550 mJ<br />

per pulse at 355 nm, the OPO delivered 75 mJ per<br />

pulse at 460 nm. After frequency doubling in a BBO<br />

crystal, laser radiation at 230 nm with a pulse energy<br />

<strong>of</strong> 15 mJ was obtained. For this setup the 230 nm<br />

laser has an estimated linewidth <strong>of</strong> ~ 0.3 cm -1 . The<br />

230 nm beam was formed into a 12 mm wide laser<br />

sheet with a thickness <strong>of</strong> approximately 300 µm. The<br />

laser sheet was sent through the combustion chamber<br />

right above the valves as indicated in Figure 9. The<br />

ICCD camera was used for detection <strong>of</strong> the LIF<br />

through the top window perpendicular to the laser<br />

sheet. A long-pass filter was used for suppressing the<br />

scattered laser radiation while transmitting the redshifted<br />

fluorescence. The engine was run at 1200 rpm<br />

corresponding to a firing frequency <strong>of</strong> 10 Hz<br />

matching the repetition rate <strong>of</strong> the laser system.<br />

Setting the engine as master, a trigger signal for each<br />

engine cycle, locked at a selected crank angle<br />

position, was sent to a pulse-delay generator<br />

(Stanford, DG535), which was used to synchronize<br />

the laser system and the ICCD camera. This enabled


LIF images to be recorded at any desired crank<br />

angles <strong>of</strong> the engine cycle. The integration time for<br />

the ICCD camera was set to 50 ns in order to<br />

minimize the flame chemiluminescence background.<br />

200 300 400 500 600 700 800 900 1000 1100 1200<br />

TDC<br />

5 TDC<br />

10 TDC<br />

20 TDC<br />

30 TDC<br />

Figure 9. CO TPLIF images at different crank angles.<br />

Single-shot CO PLIF images from the engine were<br />

recorded at subsequent crank angles, examples are<br />

shown in Figure 9. At TDC just before the ignition, a<br />

faint CO LIF image was recorded. This small amount<br />

<strong>of</strong> CO originates from the exhaust gas <strong>of</strong> the previous<br />

cycle, remembering this is a side valve engine with<br />

high amounts <strong>of</strong> residuals. At 5 and 10 CAD after<br />

TDC, the increased intensity and spread out CO LIF<br />

distribution reveal the flame location. At 20 and 30<br />

CAD after TDC, the CO PLIF images show a more<br />

intense signal in the left part <strong>of</strong> the combustion<br />

chamber (furthest away from the spark plug) and<br />

only a weak signal in the right part indicating the<br />

fading away <strong>of</strong> the reaction zone <strong>of</strong> the flame.<br />

Compared with the flame chemiluminescence images<br />

shown in Figure 8, one can find that although a<br />

strong flame chemiluminescence was observed at 30<br />

CAD after TDC the CO LIF had already decreased. It<br />

should be kept in mind that the recording <strong>of</strong><br />

chemiluminescence is a line-<strong>of</strong>-sight technique<br />

integrating across the entire thickness <strong>of</strong> the<br />

combustion chamber whereas the PLIF technique<br />

monitors a thin slice in the middle <strong>of</strong> the combusting<br />

volume. The results indicate that the CO<br />

concentration reaches the highest value in the flame<br />

reaction zone in this high pressure engine combustion<br />

environment. This is in agreement with the expected<br />

scenario where CO is formed as an intermediate<br />

species when the fuel is decomposed and then<br />

consumed as the temperature increases. At high<br />

5<br />

temperature and pressure, the interference from<br />

photolysis <strong>of</strong> hot CO2 molecule [24] can be a<br />

problem in the CO LIF detection. From the fact that<br />

only weak CO LIF was detected at 30 TDC, where<br />

high temperature are revealed from the<br />

chemiluminescence measurement, one can judge that<br />

the hot CO2 photo-fragmentation are <strong>of</strong> little concern<br />

in this experiment. This might be due to the high<br />

nascent CO concentration in the engine flame.<br />

Intensity (arb. Units)<br />

Intensity (arb. units)<br />

Intensity (arb. units)<br />

Intentisty (arb. units)<br />

6 x104<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

7 x104<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

6 x104<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

14 x104<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

scattered 460 nm laser light<br />

emission from C2<br />

5 degree ATDC<br />

15 degree ATDC<br />

scattered 460 nm laser light<br />

25 degree ATDC<br />

scattered 460 nm laser light<br />

a<br />

b<br />

c<br />

20 degree BTDC<br />

exhaust cycle<br />

scattered 460 nm laser light<br />

d<br />

0<br />

450 500 550 600 650 700<br />

Wavelength (nm)<br />

Figure 10. Spectra <strong>of</strong> TPLIF <strong>of</strong> CO from engine<br />

chamber.<br />

In order to clarify any other possible<br />

interference in the recorded CO LIF images, spectral


investigations <strong>of</strong> the fluorescence emission from the<br />

engine were performed. The CO LIF was sent<br />

through a spectrometer (Acton, SpectraPro-150) and<br />

detected with the ICCD camera. Shown in Figure 10<br />

are CO LIF spectra from the engine recorded at<br />

different crank angles. A spectral line at 460 nm,<br />

which is residuals from the OPO signal beam, is<br />

evident in every spectrum. The relative intensity <strong>of</strong><br />

this line can be used as a measure <strong>of</strong> the CO LIF<br />

intensity. The emission from C2 can introduce strong<br />

interference in CO LIF measurements especially in<br />

fuel rich flames. In Figure 10 (a), the strongest C2<br />

line showed only a very small peak, which indicates<br />

that the C2 interference were negligible in the<br />

presented measurement. Shown in Figure 10 (d) is a<br />

CO spectrum recorded at 20 CAD before TDC in the<br />

exhaust stroke, hence, the detected CO LIF signal<br />

originates from the CO in the exhaust gases. The low<br />

temperature and low pressure at this crank angle may<br />

explain the relatively intense CO LIF spectrum<br />

observed there, as indicated in Figure 5 there is a<br />

strong pressure dependence <strong>of</strong> the LIF signal. One<br />

might also be encouraged to use CO LIF technique to<br />

study engine exhaust gas recycling (EGR)<br />

distributions by performing single-shot 2D EGR<br />

measurements.<br />

Summary<br />

<strong>Two</strong>-photon LIF detection <strong>of</strong> carbon monoxide<br />

has been performed in a cell at elevated pressure, in<br />

flames with different stoichiometric and in an engine.<br />

The dependence <strong>of</strong> the LIF signal on pressure and<br />

excitation laser power density was investigated.<br />

Despite the complex involving photo-ionization,<br />

quenching, absorption etc, a general understand <strong>of</strong><br />

the behaviour <strong>of</strong> the CO LIF signal was achieved. In<br />

flame measurements, C2 interference was clearly<br />

observed especially in rich flames. Finally in the<br />

engine measurements, single-shot, two dimensional<br />

CO imaging was achieved. With the spatial- and<br />

spectral- resolved measurements, interference both<br />

from C2 and from hot CO2 photo fragmentation are<br />

excluded. To the best <strong>of</strong> our knowledge, this<br />

represents the first single shot CO imaging in<br />

combustion engines.<br />

Acknowledgments<br />

This work is supported by the Swedish Science<br />

Council (VR) and The Swedish Energy<br />

Administration (STEM).<br />

References<br />

1. A. C Eckbreth, “<strong>Laser</strong> Diagnostics for<br />

Combustion Temperature and Species”, 2 nd<br />

edition, Gordon and Breach, UK, 1996.<br />

2. K. Kohse-Höinghaus, and J. B. Jeffries, (eds.),<br />

“Applied Combustion Diagnostics”, Taylor and<br />

Francis, New York, 2002.<br />

6<br />

3. J. Wolfrum, Twenty-Seventh Symposium<br />

(International) on Combustion, The Combustion<br />

Institute, Pittsburgh, 1-41 (1998).<br />

4. K. Kohse-Höinghaus, Prog. Energy Combust.<br />

Sci. 20, 203-279 (1994).<br />

5. M. Aldén, S. Wallin and W. Wendt, Appl. Phys.<br />

B 69, 505 (1984).<br />

6. J. Haumann, J. M. Seitzman and R. K. Hanson,<br />

Opt. Letts. 11, 776-779 (1986).<br />

7. J. M. Sertzman, J. Haumann and R. K. Hansson,<br />

Appl .Opt. 26, 2892-2899 (1987).<br />

8. D. A. Everest, C. R. Shaddix and K. C. Smyth,<br />

Twenty-Sixth Symposium (International) on<br />

Combustion, The Combustion Institute, 1161-<br />

1169 (1996).<br />

9. N. Georgiev and M. Aldén, Appl. Spec. 51,<br />

1229-1237 (1997).<br />

10. S. Agrup and M. Aldén, Chem. Phys. Letts. 189,<br />

211-216 (1992).<br />

11. S. Agrup and M. Aldén, Applied Spectroscopy<br />

48, 1118 (1994).<br />

12. M. D. Di Rosa and R. L. Farrow, J. Opt. Soc.<br />

Am. B 16, 1988-1994 (1999).<br />

13. M. D. Di Rosa and R. L. Farrow, J. Opt. Soc.<br />

Am. B 16, 861-870 (1999).<br />

14. M. D. Di Rosa and R. L. Farrow, J. Quant. Spec.<br />

Rad. Trans. 68, 363-375 (2001).<br />

15. F. D. Teodoro, J. E. Rehm, R. L. Farrow and P.<br />

H. Paul, J. Chem. Phys. 113, 3046-3054 (2000).<br />

16. T. B. Settersten, a. Dreizler and R. L. Farrow,<br />

“Temperature- and species-dependent quenching<br />

<strong>of</strong> CO B probed by tow-photon laer-induced<br />

fluorescence using a picosecond laser”, J. Chem.<br />

Phys. 117, 3173-3179 (2002).<br />

17. A. P. Nefedov, V. A. Sinel’shchikov, A. D.<br />

Usachev and A. V. Zobnin, Appl. Opt. 37, 7729<br />

(1998).<br />

18. S. Linow, A. Dreizler, J. Janicka and F. P.<br />

Hassel, Appl. Phys. B 71, 689-696 (2000).<br />

19. B.J. Kirby and R.K. Hanson, Appl. Phys. B 69,<br />

505-507 (1999).<br />

20. B. J. Kirby and R. K. Hanson, Appl. Opt. 41,<br />

1190-1201 (2002).<br />

21. K. Nyholm, R. Fritzon, N. Georgiev and M.<br />

Aldén, Opt. Commu. 97, 2246-2250 (1995).<br />

22. M. Karlsson, “Experimental and theoretical<br />

studies <strong>of</strong> two-pnoton laer-induced fluorescence<br />

<strong>of</strong> carbon monoxide”, Master’s thesis, Lund<br />

Instituted <strong>of</strong> Technology, LRCP-44 (1998).<br />

23. C. F. Kaminski, X. S. Bai, J. Hult, M. Richter, J.<br />

Nygren, A. Franke, M. Aldén, S .Lindenmaier,<br />

A. Dreizler, U. Maas, R. B. Williams, SAE<br />

Paper 2000-01-2833, Journal <strong>of</strong> Fuels &<br />

Lubricants (2000).<br />

24. W. G. Bessler, C. Shculz, t. Lee, J. B. Jeffries<br />

and R. K. Hanson, Chem. Phys. Letts. 375, 344-<br />

349 (2003).

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

Saved successfully!

Ooh no, something went wrong!