1 s2.0 S0016236123028156 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Fuel 358 (2024) 130201

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full length article

Spectroscopic investigation of diesel-piloted ammonia spray combustion


Valentin Scharl ∗, Thomas Sattelmayer
Chair of Thermodynamics, Technical University of Munich, Boltzmannstraße 15, 85748 Garching, Germany

ARTICLE INFO ABSTRACT

MSC: This work investigates the flame emissions of diesel-piloted ammonia spray combustion under conditions
00-01 relevant to internal combustion engines. Due to ammonia’s carbon-free structure, its flame emissions differ
99-00 significantly from carbon-containing fuels. This work provides new insights into the combustion process by
Keywords: distinguishing pilot and main fuel combustion by analyzing spectrally resolved flame emission footprints
Ammonia between 275 nm and 750 nm in a rapid compression expansion machine (RCEM). Additional results based
Chemiluminescence on Shadowgraphy (SG), OH*-chemiluminescence (CL), visible natural flame luminosity (NL) imaging, and
Direct-injection
heat release rate (HRR) analysis describe the combustion process in detail. OH*, NH*, NH2 *, as well as
Combustion
broadband CL from NH2 *, NO2 *, and H2 O* contribute to ammonia flame luminosity. The results show that the
Spectroscopy
Dual-fuel combustion process encompasses a pilot ignition phase, a flame transition phase, and an ammonia combustion
phase, during which only ammonia-related flame emissions occur and the main share of heat is released.
Background corrected OH*, NH*, and NH2 * intensities reveal the evolution of individual contributions to the
flame emissions. NH2 *-CL is the main contributor to visible ammonia flame emissions under the conditions
investigated. Therefore, visible flame emissions are a suitable indicator for the flame front during the ammonia
combustion phase.

1. Introduction production by renewable power plants. Among the dual-fuel concepts,


the high-pressure dual-fuel (HPDF) combustion, where both fuels are
With the background of climate change, ammonia is of interest injected close to top dead center (TDC), offers the potential to reduce
as an alternative to fossil fuels. Due to its carbon-free structure, it unburned fuel emissions caused by wall-quenching effects and valve
offers the potential to reduce greenhouse gas emissions significantly. timing overlap. For this combustion process, the primary fuel burns
Easier storage and higher energy densities provide an advantage for
mainly diffusively, similar to diesel engines. Comprehensive reviews of
transportation or mobile applications compared to hydrogen. A possible
ammonia combustion in ICEs can be found in [5] and [6].
option to convert the energy stored in ammonia is using internal com-
Imaging of the natural flame luminosity (NL) is used in experimental
bustion engines, which offer a robust and cost-effective way of energy
combustion research to infer important features of combustion pro-
conversion. Therefore, internal combustion engines could be part of
future energy systems [1]. However, as the combustion characteristics, cesses, such as flame front location, ignition time and location, or heat
physical properties, and chemical reaction pathways differ immensely release rate (HRR) distribution [7,8]. Most investigations target specific
from carbon-based fuels, research is necessary to develop internal wavelength ranges using bandpass-filtered 2-D imaging techniques.
combustion engines suitable for using ammonia. As ammonia has a Therefore, interference between the chemiluminescence (CL) of several
high auto-ignition temperature, slow laminar burning velocities, high excited species emitting in the investigated band or with other sources
enthalpy of evaporation and narrow flammability limits, measures to of luminosity, such as soot, distorts the information obtained. For exam-
improve its combustion behavior are commonly used [2]. These include ple, under high-temperature conditions, OH*-CL is known to interfere
hydrogen addition [2], ammonia dissociation prior to combustion [3] with soot radiation [9,10]. Under low-temperature combustion condi-
and the use of pilot fuel [4]. Furthermore, pure, spark-ignited ammonia tions, OH*-CL interferes with CO2 * and HCO* broadband CL [11,12].
combustion is feasible if compression ratios and fuel preheating are By dispersing flame emissions based on their wavelengths, spectro-
sufficiently high [4]. Dual-fuel concepts that use conventional fuel
scopic investigations can help distinguish flame emissions’ source. A
as pilot may also run on the conventional fuel only. This option
detailed overview of the chemiluminescent species present in various
decreases the exposure of engine operators to the risk of fluctuating
flames is given in [13]
ammonia availability, e.g., caused by seasonal variability of power

∗ Corresponding author.
E-mail addresses: valentin.scharl@tum.de (V. Scharl), sattelmayer@td.mw.tum.de (T. Sattelmayer).

https://doi.org/10.1016/j.fuel.2023.130201
Received 24 July 2023; Received in revised form 30 September 2023; Accepted 26 October 2023
Available online 3 November 2023
0016-2361/© 2023 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
V. Scharl and T. Sattelmayer Fuel 358 (2024) 130201

The chemiluminescent signature of flames generally depends Table 1


strongly on the fuel used, combustion temperatures and pressures, RCEM specifications and operating parameters. The timings in CAD refer to OP4 at
1000 rpm.
equivalence ratios, and strain rates. Only a few studies investigated
Operating point OP3 OP4
the flame luminosity under engine-relevant temperatures and pressures
for ammonia. Wang et al. [14] used visible flame luminosity to study TDC temperature [K] 865 ± 10 920 ± 10
TDC pressure [bar] 98 ± 2 125 ± 2
pure ammonia spray combustion. They observed a transparent-like,
Effective compression ratio [−] 17.2 20.5
orange-red flame with less intense luminosity than sooting diesel flames 1
Equivalent engine rpm [ min ] ≈920 ≈1000
under similar conditions. Wüthrich et al. [15] used simultaneously Bore diameter [mm] 220
acquired Schlieren and OH*-CL imaging to analyze the combustion of Maximum Stroke [mm] 368.9
homogeneous ammonia mixtures ignited by diesel pilots (low-pressure Fuel injection temperature [K] 293 ± 3
dual-fuel combustion), which allowed a detailed description of pilot Diesel injection duration [μs∕CAD] 520∕1.56
fuel ignition, as well as flame transition to the premixed ammonia Start of diesel injection [μs∕CAD btdc] 2000 ± 50∕6 ± 0.15
charge and subsequent flame propagation. The authors of the study Diesel nozzle diameter [μm] 200
Diesel injection pressure [bar] 2000 ± 20
at hand carried out OH*-CL investigations of ammonia combustion
Injected diesel mass [mg] 5
to analyze the combustion of diesel-piloted ammonia sprays (HPDF
Ammonia injection duration [μs∕CAD] 2700∕8.1
combustion) in a rapid compression expansion machine (RCEM). We
Relative ammonia injection timing [μs∕CAD] +500∕1.5
found the OH*-radiation intensity of the burning ammonia spray to Diameter [μm] 940
be more than an order of magnitude lower than the OH* intensity Injection pressure [bar] 503 ± 10 530 ± 10
emitted by the diesel pilot combustion, although diesel only accounted Ammonia mass [mg] 210
for 5% of the LHV supplied [16,17]. Spectroscopic measurements
of premixed mixtures containing ammonia have been carried out at
atmospheric pressures for CH4 /NH3 /air mixtures by Zhu et al. [18]
and for CH4 /NH3 /H2 /air by Mashruk et al. [19]. The studies detected 2.1. Test rig
NO*, OH*, NH*, CN*, CH*, NH2 * peaks, as well as CO2 *, H2 O* and
NO2 * broadband radiation. A strong reduction in the intensity of all The investigations are conducted in a pneumatically driven RCEM,
peaks (including NH*) except NH2 * was found for increasing ammonia which is optically accessible via a quartz-glass piston. The machine
shares. While the behavior of CH* and CN* can be explained by the simulates single cycles of large-bore engines. Essential specifications
reduced availability of carbon atoms, the decline of NH*-CL intensity and operating parameters are listed in Table 1, and a scheme of the
was attributed to the reduced OH* concentration in ammonia flames, as RCEM driving mechanism, the cylinder head, and the spectroscopy
NH* is mainly produced via the reaction NH2 + OH → NH + H2 O [19]. setup is shown in Fig. 1. This section only outlines the main features of
Under engine-relevant conditions, Ichikawa et al. [20] reported the the RCEM, while a more detailed description can be found in [24].
flame emission profiles obtained from stratified spray injection of Before each experiment, the driving-air bottles are filled with pres-
ammonia and n-hexadecane through a single nozzle. By detecting the surized air. After starting the experiment, the outer piston, which is
NH2 * peaks, they corroborated that the supporting fuel successfully hydraulically coupled to the inner piston, accelerates. As a result, the
induced ammonia combustion. inner working piston accelerates in the opposite direction, which can-
For HPDF combustion, pilot/supporting fuel and ammonia are not cels net momentum and reduces unwanted vibrations. The compression
mixed prior to combustion, which differs from the spectroscopy inves- stroke ends at TDC when a dynamic equilibrium is reached, and the
tigations discussed above. Instead, the two fuel sprays interact within expansion stroke starts. The piston trajectory is determined by driving
the combustion chamber. This interaction is a major influencing factor pressure, flow resistance, and combustion chamber pressure. For the
on the performance of the combustion process [21]. HPDF combustion conditions investigated in this work, the piston motion resembles the
relies on the favorable auto-ignition properties of the pilot fuel, whose motion induced by a crankshaft at 920 or 1000 rpm, depending on the
ignition may fail or be delayed when the mixture formation process operating point. The combustion chamber pressure is measured during
deteriorates by interaction with the less reactive primary fuel [22,23]. the experiment using a Kistler 7061C piezo-electric pressure transducer.
When using ammonia as the main fuel, the unique CL features of the An inductive linear incremental encoder measures the piston position.
ammonia flame (e.g., NH*, NH2 * emissions) may allow us to distinguish Temperature is only measured before the experiment via a thermo-
between the combustion of pilot fuel and ammonia. couple. During the experiment, the temperature is calculated under
This work aims to characterize the natural flame emissions of the assumption of an undisturbed gas core that undergoes isentropic
diesel-piloted ammonia HPDF combustion using spectrally resolved compression. This assumption is valid when flat pistons and large
high-speed CL measurements. The data obtained may serve as a basis on crevice volumes prevent charge motion and roll-up of the boundary
which wavelength ranges to choose for investigating ammonia combus-
layer [25], as in the RCEM employed. Deviations in driving pressure,
tion optically and help interpret the observed luminosity. Furthermore,
initial chamber pressure, and oil temperature lead to deviations in the
we try to take advantage of unique flame emissions of ammonia (NH*,
chamber pressure obtained at TDC. This deviation also causes different
NH2 *) to characterize the flame transition process from pilot to primary
TDC temperatures according to the relations for isentropic compression.
fuel. The work is structured as follows: First, the test rig and the
The deviation in temperature due to different compression ratios is
optical measurement system employed are presented. Then, temporally
superimposed by changes in initial chamber temperature. Furthermore,
and spatially integrated spectra are discussed to reveal which species
the ammonia injector opening timing deviates slightly between indi-
contribute significantly to the observable natural flame emissions. Sub-
vidual experiments, as can be extracted from optical imaging data. In
sequently, time-resolved flame spectra are discussed in context with SG,
OH*-CL, and NL images to characterize the combustion process with a addition, the accuracy with which the injection pressures can be set is
focus on flame transition. Furthermore, NH*, OH*, NH2 *, and broad- limited. Table 1 provides typical ranges of the quantities discussed.
band luminosity intensity evolutions are extracted from the spectra and The thermodynamic model also assumes this isentropic core to
discussed in context to the HRR. The work concludes with a summary calculate HRRs based on chamber pressure and piston position. The
and recommendations regarding future optical investigations. multi-zone model assumes heat loss to occur only in representative
piston crevices. Heat loss model parameters are calibrated using unfired
2. Experimental setup and methods experiments. Therefore, the HRRs presented in this work include the
heat sink due to evaporation of the fuels and effects due to species
This section briefly describes the test rig used, the optical methods conversion. Detailed information on the thermodynamic model is given
employed, and the processing of the spectroscopy results. in [22,23,26].

2
V. Scharl and T. Sattelmayer Fuel 358 (2024) 130201

Fig. 1. RCEM driving system and cylinder head: (1) driving-air bottles, (2) driving piston, (3) hydraulic fluid, (4) flow orifice, (5) working piston, (6) combustion chamber,
(7) cylinder head, (8) diesel injector, (9) ammonia injector, (10) UV lens, (11) monochromator, (12) image intensifier, (13) high-speed camera.

Fuel injection is triggered based on the piston position via a Com- Previous studies illustrate the high repeatability of experiments in
pactRIO field-programmable gate array. Single-hole injectors inject the the RCEM by showing low variations in pilot fuel ignition delay [22,23]
fuels on a plane parallel to the flat piston surface to study the interac- and repeatable OH* CL intensities [16]. Furthermore, the results pre-
tion between the two sprays in detail. A custom Woodward L’Orange sented in this work rely on clear qualitative trends robust to the small
injector with a 0.94 mm spray hole injects ammonia, while a Bosch variations between individual experiments.
CRI 2.20 injector with a nozzle diameter of 0.2 mm supplies diesel. The
injection pressure of ammonia is 530 bar for the 125 bar TDC pressure 2.2. Optical setup
operating point, and the same relative pressure between the ammonia
injector and the chamber is maintained for the lower compression ratio The optical investigations presented in this work rely on three differ-
investigated. In contrast, the diesel injection pressure is kept constant ent optical setups: Spectrally resolved CL, simultaneous SG and OH*-CL
at 2000 bar, as the change in relative pressure is low. Air-driven high- imaging, and NL imaging. First, the setup for spectrally resolved CL
pressure pumps (Maximator G400-2 for diesel and Maximator G150-2 measurements is presented.
with EPDM gaskets for ammonia) supply the two fuels. The ammonia Fig. 1 includes a scheme of the spectroscopy setup. A Cerco UV lens
fuel line features an additional counter-pressure valve that creates a with a focal length of 100 mm collects light emitted by the combus-
back-pressure of 15 bar in the leakage stream after the fuel injector. tion event and projects it onto the 10 μm slit of the monochromator.
The increased pressure inhibits fuel evaporation within the injector, The SpectraPro 275 spectroscope from Acton Research Corporation is
which would deteriorate injector behavior. The injectors were analyzed a Czerny-Turner type monochromator. It allows to change between
in an IAV injection analyzer using calibration fluid. The correspond- different gratings, which feature different densities of grooves per mm
ing injection mass was then derived by assuming that the discharge and result in different spectral dispersions. In this work, a grating with
coefficient is independent of the fuel used. Coppo et al. [27] showed 150 grooves per mm is used to resolve the CL in the UV–VIS range
that opening and closing behavior of the injector are not affected by from 275–750 nm with a spectral resolution of 1 nm, while 600 grooves
changing the fuel. However, Gaucherand et al. [28] pointed out that per mm are used to resolve the UV range from 275 nm to 400 nm with
ammonia may show significantly lower injection masses due to intense a higher spectral resolution of 0.25 nm. A Hamamatsu C10880 image
cavitation under engine-relevant conditions, particularly when the fuel intensifier coupled with Photron Fastcam SAX2 records the spectra at
is preheated. Therefore, ammonia was injected into the RCEM combus- the outlet of the monochromator. For the less reactive case (OP3), flame
tion chamber under non-reacting conditions (i.e., no pilot) to confirm emissions in the UV–VIS range are recorded with 20 000 f ps, while in
the determined ammonia injection mass. The resulting concentration the UV range, longer exposure times were necessary, which resulted in
in the chamber was measured using an ABB Advanced Optima Limas a frame rate of 10 000 f ps. For the highly reactive case (OP4), the soot
21 HW gas analyzer. The difference in injected masses was small due luminosity of the pilot fuel prohibits any investigations in the visible
to the cold initial fuel temperature and the high combustion chamber wavelengths. The UV emissions were significantly brighter than in the
pressure, which reduce cavitation. The angle between the two fuel less reactive case and were recorded at 20 000 f ps. The resolution of
sprays and the timing of the two injections can be freely adjusted. the spectrally resolved images is 1024 × 632 pixels. Different image
However, in this work, timing and spray angle are fixed at values found intensifier gains prohibit a comparison of the flame emission intensities
suitable in a previous publication, where a more detailed description of obtained for different combustion cases. Therefore, the results do not
the injection setup is given [23]. Based on these findings, the spray provide absolute intensity values for most cases.
interaction angle is set to −7.5◦ , which leads to slightly converging Simultaneous SG and OH*-CL images were obtained using the
and strongly interacting sprays. The start of diesel pilot injection is double-pass setup shown in Fig. 2. A point source of light is created by
set to 2 ms BTDC, corresponding to 12 CAD BTDC for 1000 rpm, while focusing a xenon-arc lamp on a pinhole. The pinhole is placed in the
ammonia is injected 0.5 ms after diesel. The work presented in the focus of a large concave mirror redirecting the beams onto a flat mirror
paper at hand investigates two different TDC conditions. The less and then into the combustion chamber. The light passes the combustion
reactive condition (OP3) features a TDC pressure of 98 bar and a TDC event before it is reflected onto the same optical trajectory as before by
temperature of 865 K. Soot luminosity is weak for these conditions, a polished stainless steel plate. Therefore, the density gradients in the
and the CL emissions in the UV and visible range can be investigated combustion chamber deflect the light beams twice. Two beam splitters
undisturbed. The highly reactive case (OP4) features a TDC pressure redirect the reflected light. For SG imaging, the light is detected directly
of 125 bar and a temperature of 920 K, which are the most reactive via a Photron Fastcam SAX with 20 000 f ps. The flame emissions, which
conditions possible in the RCEM employed. Under these conditions, the follow the same optical trajectory, are redirected by a second beam
strongly sooting diesel combustion leads to overexposure of the image splitter and filtered with a bandpass filter centered at 307 nm with an
intensifier in the visible wavelength range. Therefore, the investigation FWHM of 10 nm. The same combination of Hamamatsu C10880 and
considers only spectrally resolved UV emissions for this operating point. Photron Fastcam SAX2 as above is used for detecting the CL. For SG

3
V. Scharl and T. Sattelmayer Fuel 358 (2024) 130201

enable reproducible interpolation of the background radiation. The


filter considered 21 interpolation points to approximate fifth-degree
polynomials. In addition, broadband luminosity intensity is consid-
ered, which is obtained by integrating the intensities over the entire
range of investigated wavelengths. While OH* and NH* intensities
are obtained by considering the area of individual peaks, the NH2 *
intensity is obtained by adding the intensity of several of its major
peaks. NH2 * flame emissions consist of a multiple-line system appearing
as a continuous band with superimposed heads [13,29]. Therefore, the
approach followed in this work only detects a fraction of the actual
NH2 * intensity. However, the intensity inferred from the peak areas
can give information on when NH2 * radiation occurs and its relative
intensity. OH* intensities were derived from images obtained using
both gratings. NH* intensities are derived using the higher spectral
dispersion, as the NH* peak was not sufficiently prominent for the
Fig. 2. SG and CL setup: (1) xenon-arc lamp, (2) UV-filter, (3) pin hole, (4) concave post-processing routine in the images obtained using grating 3. The
mirror, (5) planar mirror, (6) reflective polished steel plate, (7) beam splitter, broadband radiation can mainly be assigned to CO2 * at shorter wave-
(8) focusing lens, (9) bandpass filter, (10) image intensifier, (11) and (12) camera. lengths and at longer wavelengths to NO2 *, H2 O*, and NH2 * [13]. For
diesel combustion, soot might occur additionally.

and OH* imaging, both cameras acquire images at 20 000 f ps. NL images 3. Results and discussion
are captured separately in additional experiments. Due to the light
emitted by the xenon-arc lamp, simultaneous performance of NL and In the first part of this section, temporally and spatially inte-
SG imaging is impossible. As an image intensifier is not necessary for grated CL signatures of the lower reactivity combustion case with low
detecting the bright visible flame luminosity, the setup consists only of soot luminosity are presented to enable a detailed assignment of the
the Photron Fastcam SAX2 and a lens. observed flame emissions to individual species. Subsequently, time-
resolved spectra are interpreted with SG, OH*-CL, and NL imaging
2.3. Spectroscopy image processing to describe the HPDF combustion process in detail. In addition, the
temporal evolution of background corrected OH*, NH*, NH2 *-CL and
The monochromator preserves the spatial coordinate along the broadband luminosity intensity are discussed in context to the HRRs
spray axis (vertical), while the incident light disperses on the coor- for two combustion chamber conditions. Three repetitions were carried
dinate perpendicular to the spray axis (horizontal) depending on its out for each case and measurement technique. The HRRs are averaged
wavelength. Therefore, the horizontal positions of the pixels of the over all experiments conducted for the combustion case, while the
images taken reflect specific wavelengths, while the vertical positions spectroscopic data is averaged over the three experiments conducted
of the pixels correspond to a certain height in the combustion chamber. for each grating. The 2D imaging data is taken from the cases with the
Before the combustion experiments, light emitted by a mercury-arc HRRs most similar to the averaged HRR.
lamp was recorded. This light consists of several distinct spectral lines
specific to mercury-arc lamps. Therefore, specific pixels of the images 3.1. Integral flame spectra
taken correspond to certain wavelengths, and a conversion rule from
the wavelength to the pixel scale is derived based on interpolation. Fur- The integral flame emissions shown in Fig. 4 are obtained by
thermore, the coordinate along the spray axis was calibrated by varying temporally and spatially integrating the flame emissions. This results
the position of the mercury-arc lamp within the combustion cham- in well-defined spectra suitable for identifying individual peaks and
ber. A third calibration procedure corrects the wavelength-dependent the corresponding species. While the spatial integration was conducted
sensitivity of the setup employed. This calibration requires a light over the entire combustion chamber for diesel, the combustion chamber
source with a continuous emission spectrum. Therefore, a halogen region starting from 20 mm downstream of the diesel nozzle was
lamp, whose spectrum is known from the manufacturer, was located considered for the ammonia spectra shown to exclude the influence of
in the combustion chamber, and the light emitted was recorded. A diesel combustion due to injector dripping. The diesel spectrum was
wavelength-dependent calibration function was derived by comparing obtained by temporally integrating up to 900 μs ASOI. In comparison,
the intensities obtained by the spectroscopy setup to the reference the ammonia spectrum was obtained by integrating from 1600 μs to
spectrum. This function corrects the spectra obtained during the exper- 6000 μs ASOI (of diesel). As diesel is injected before ammonia, diesel
iments. The correction considers the wavelength-dependent sensitivity combustion progresses undisturbed at first. Subsequently, the sprays
of the entire setup, encompassing the quartz-glass piston and lens interact, and both fuels burn simultaneously from approximately 900 μs
transmissivities, as well as the monochromator’s sensitivities and the to 1600 μs ASOI. This flame transition period is not included in the
image intensifier’s sensitivities. analysis conducted in this section and will be discussed in more detail
While significant parts of this work use raw spectra to discuss in the next section. The corresponding HRR and injection mass fluxes
the evolution of the dual-fuel combustion process, the results also are shown in Fig. 5.
discuss the background corrected intensities of OH*, NH*, NH2 *, and In the following, A-X or B-X refers to the electron state transition,
broadband emissions. The background corrected intensities result from while (𝑣′ , 𝑣′′ ) refers to the vibrational states of the electrons in their
integrating peak areas corresponding to the respective species’ CL upper and lower states. The UV range of the ammonia flame spectrum
emissions. These peak areas superimpose on a continuous background shown in Fig. 4 a (grating 3, left part) and (c) (grating 2) consists
of broadband flame emissions. Fig. 3 illustrates the areas considered of the distinct NH* (A-X (0,0)) peak system at 336 nm and the OH*
for the background corrected OH*, NH*, and NH2 * intensities. The (A-X (0,0)/(1,1)) peak system at around 308 nm superimposed on
background radiation is interpolated linearly within the range, in which weak broadband radiation. A weak peak visible at 282 nm results from
each considered species causes a peak in CL emissions to distinguish OH* (A-X(1,0)). Continuous broadband CL is observed for wavelengths
between species-specific and background radiation. The spectra were longer than 400 nm (see Fig. 4 a). A possible source for this radiation
filtered using a Savitzky-Golay filter in 𝑀𝑎𝑡𝑙𝑎𝑏 to reduce noise and is the CL of NO2 *. The approach by Sheehe and Jackson [30] is

4
V. Scharl and T. Sattelmayer Fuel 358 (2024) 130201

Fig. 3. Typical CL spectrum of ammonia combustion. Left: UV-range obtained using grating 2, right: UV–VIS spectrum obtained using grating 3. The colored peak areas represent
the background corrected CL intensity of the corresponding species. The peaks are extracted by interpolating the background radiation linearly in the respective wavelength ranges.

Fig. 4. Spatially and temporally integrated flame spectra (OP3). (a) and (b) were obtained using grating 3, while (c) was obtained using the higher spectral dispersion offered by
grating 2. (a) and (c) are obtained by temporal integration ranging from 1600 μs to 6000 μs ASOI, while (b) is obtained by integrating until 900 μs ASOI.

adopted to test whether NO2 * can explain the broadband emissions 432 nm. The highly bright sodium D-line at 589 nm results from sodium
observed for the ammonia flame. Therefore, the NO2 * intensity dis- impurities in the diesel fuel. An additional peak at approximately
tribution presented by [31] is compared to the spectrum obtained in 460 nm could not be identified and is thought to result from impurities
this work (see Fig. 4 a). The NO2 * intensity distribution is fitted to the in the diesel fuel. Similar to the ammonia case, broadband emissions
ammonia flame emission spectrum so that the NO2 * emissions do not at longer wavelengths result from NO2 * and H2 O*. In addition, weak
exceed the ammonia flame emission intensity at any wavelength. As a soot luminosity occurred. The absence of distinct features of diesel
result, the NO2 * intensity distribution fits well in a range from 400 nm combustion in the ammonia emission spectra clearly shows that the
to approximately 450 nm. However, the broadband emissions of the ammonia combustion phase after 1600 μs ASOI, in which the major part
ammonia flame are more intense for wavelengths longer than 450 nm of heat release occurs (see Fig. 5), is not directly supported by diesel
than to be expected if only NO2 * was contributing to its flame emis- fuel combustion.
sions. Therefore, the maximum possible contribution of NO2 * to the
flame luminosity during ammonia combustion is small. The broadband 3.2. Temporal and spatial evolution of flame emissions
emissions unaccounted for by NO2 * may be attributed to H2 O* and to
NH2 * [20]. In addition, the NH2 * peaks of the ammonia-alpha band are The temporal evolution of the combustion using SG, OH*-CL and
superimposed on the broadband spectrum. In addition to the OH* (A- NL imaging is shown in Fig. 6 for 𝑡 = 700–1600 μs and in Fig. 7 for 𝑡 =
X (0,0)/(1,1)) peak system, the diesel spectrum recorded during earlier 2000–5000 μs. The corresponding temporal evolution of flame spectra
combustion phases and shown in Fig. 4b shows more intense broadband for the same combustion case is shown in Fig. 8. While SG and OH*-CL
radiation in the UV range. This radiation is attributed to HCO (Vaidya’s images were taken simultaneously, NL and spectroscopy imaging could
hydrocarbon flame bands), HCHO (Emeleus’s cool flame bands), as only be performed one at a time, which required repeated experiments.
well as CO2 * [13,32,33]. A weak CH* peak (A-X (0,0)) is visible at SG contours are overlayed onto the OH*-CL images to enable a detailed

5
V. Scharl and T. Sattelmayer Fuel 358 (2024) 130201

Fig. 5. HRR and mass fluxes of the investigated combustion cases featuring a slightly advanced diesel pilot injection (+500 μs).

spatial allocation of the flame zone. To render the weak OH*-CL phases. The low relative intensity of the NH* peak compared to the
during late ammonia combustion visible, the intensities in Fig. 7 are diesel background radiation intensity in the respective wavelength
brightness-corrected by potentiating the initial intensity values ranging range may prohibit its use as a flame marker for optical combustion
from 0 to 1 with 0.4. The flame emission spectra corresponding to the diagnostics.
timings of the images shown in Fig. 6 and Fig. 7 are shown in Fig. 8. During later combustion phases, the flame spectra do not show
During diesel ignition, first weak NL is observed at 𝑡 = 700 μs (see significant qualitative changes apart from different intensities (see
Fig. 6). The flame spectra (see Fig. 8 a) show first broadband CL with Fig. 8 c). The UV emissions, in particular OH*-CL, become weaker,
only weak OH* luminosity at 𝑡 = 750 μs, which might hint at a strong while the peak intensity of visible broadband CL is reached between
contribution of cool flame ignition chemistry to the flame emissions 𝑡 = 2000 μs and 𝑡 = 3000 μs. Both OH*-CL and NL drift off from the
observed. However, as experiments are averaged over three repetitions injector until and after the end of injection at 𝑡 = 3200 μs (see Fig. 7).
and the camera’s exposure time was set to 𝑡 = 40 μs, no isolated Therefore, combustion progresses close to the cylinder wall, and more
spectrum containing only cool flame chemistry was obtained. Subse- air entrains up to the flame front. As a result, average equivalence
quently, flame emission intensity increases significantly, and peaks, ratios become leaner, and combustion temperatures drop. The lack
such as OH*, CH*, and the sodium D-line, become more prominent in of flame stabilization of ammonia flames observed can be attributed
the spectra. At 𝑡 = 850 μs, both, OH*-CL and NL, encompass the entire to an inhibition of the stabilization mechanism predominant for con-
diesel spray. Although the ammonia spray starts interacting with the ventional diesel fuels, as mixing requirements of the fuel/air mixture
diesel spray at approximately 𝑡 = 750 μs, no sign of ammonia-specific with combustion products to undergo auto-ignition are exceptionally
flame emissions, such as NH* and NH2 *, can be observed in the spectra high for ammonia [16,36]. As OH*-CL gradually weakens, it cannot
until 𝑡 = 1000 μs. be detected reliably for later combustion phases, while NL still permits
At 𝑡 = 1000 μs, NH* and NH2 * peaks first emerge on the spectra proper observation of the combustion event.
(see Fig. 8 b). Simultaneously, OH*-CL decreases, reflected both in the
spectra and the OH*-CL images. The decrease starts at 𝑡 = 1000 μs on 3.3. Temporal evolution of background corrected intensities of OH*, NH*
the side, where ammonia interacts with the diesel cloud first (right) and NH2 *
before an intense reduction of the CL in almost the entire spray occurs.
At 𝑡 = 1600 μs, intense OH*-CL is restricted to a small pocket of gases The temporal evolution of background corrected intensities of OH*,
on the sides of the spray, which most likely contain diesel combustion NH2 *, and broadband luminosity in the lower reactivity case OP3 are
products. The emission spectra from 𝑡 = 1000 μs to 𝑡 = 1600 μs show shown in Fig. 9 a. The data is derived from the spectra obtained using
a decrease in overall intensity. The UV emission intensity drops to grating 3 according to the abovementioned procedure. Furthermore,
particularly low levels. The emergence of the NH2 * peaks and the each intensity is normalized with its highest value. OH* and broadband
decline of the sodium D-line mark the shift from diesel to ammonia luminosity rise sharply after diesel ignition at 𝑡 ≈ 750 μs, with broad-
combustion. The location of both OH*-CL and NL indicates that the band luminosity reaching its first peak slightly earlier. Both intensities
flame does not encompass the ammonia spray tip in this combustion start declining almost immediately, as pilot injection ends at 𝑡 = 520 μs,
phase. Instead, the reaction zone location changes only slightly during and no new diesel fuel is supplied to the reaction zone. At 𝑡 ≈ 900 μs,
the flame transition process, while the ammonia spray penetrates into NH2 * intensity increases as ammonia starts reacting. Starting from 𝑡 ≈
the combustion chamber. Therefore, unburned ammonia emissions may 1600 μs, broadband luminosity and NH2 * intensity correlate strongly,
originate in the spray tip region. This behavior is attributed to the indicating that the flame luminosity results from ammonia combustion
low ignitability of ammonia. In addition, the tip region features the only. Therefore, the background corrected intensities corroborate the
typical fuel-rich head vortex, which shows particularly poor mixture findings inferred from the raw spectra above: After a short pure pilot
formation, resulting in slower evaporation and lower temperatures [34, combustion phase, the flame transition occurs in which both fuels con-
35]. Furthermore, the HRR does not increase significantly between 𝑡 = tribute to combustion. However, the heat release during this transition
1000 μs to 𝑡 = 1600 μs, which also indicates slow and limited combustion phase is low (see Fig. 5). The main part of combustion and heat release
of ammonia in this phase. In contrast to recent spectroscopic studies is entirely attributed to the combustion of the ammonia spray.
burning ammonia and a second, carbon-containing fuel [18–20], the Furthermore, the strong correlation between NH2 * peak intensity
recorded spectra lack CN* peaks (e.g., B-X(0,0) and B-X(1,1) around and broadband luminosity suggests that a major part of the ammonia
388 nm) during the entire combustion event. As CN originates from flame emissions result from NH2 * under the conditions investigated.
interactions between carbon- and nitrogen-containing fuels, the lack of The correlation is particularly strong for the time interval between
its flame emissions hints at low rates of inter-species reactions between 𝑡 ≈ 1600 μs and 𝑡 ≈ 3300 μs. Afterwards, broadband luminosity intensity
ammonia and diesel for the combustion process investigated in this increases relative to the NH2 * intensity, which indicates the increasing
work. The NH* peak is weak during flame transition and subsequent importance of other flame emissions, such as NO2 *-CL. The conversion

6
V. Scharl and T. Sattelmayer Fuel 358 (2024) 130201

Fig. 6. SG (top), OH*-CL with overlayed SG contours (middle) and NL (bottom) imaging obtained at OP3 (865 k, 98 bar). The left section contains images taken during pilot ignition
(𝑡 = 700–850 μs ASOI), while the right section contains images taken during the flame transition from diesel to ammonia (𝑡 = 1000–1600 μs ASOI). The combustion phases are also
indicated via the HRRs at the top.

of ammonia to NH2 via hydrogen abstraction is generally the first step equilibrium. Owing to the low fuel mass supplied by the single-hole
of most ammonia oxidation mechanisms [37], NH2 will mainly occur injectors, this pressure rise due to combustion is low in the RCEM.
close to the flame front. Therefore, the NL of the ammonia spray flame Therefore, practical engines may show more intense afterglow in the
observed is a suitable indicator for the flame front during a significant burned zone.
part of the ammonia combustion phase. The findings align with a The background corrected OH* and NH* intensities of both oper-
recent study on visible ammonia flame emissions from atmospheric ating points investigated are shown in Fig. 9 b. The intensities are
burner flames: Weng et al. [29] found that NH2 * radiation is the main obtained by analyzing the UV spectra obtained using the higher spectral
source of CL at the flame front for fuel-rich and fuel-lean flames. In the dispersion offered by grating 2, as sooting diesel pilot combustion at
post-flame zone, NH2 * is the main contributor to flame emissions for OP4 permitted the investigation of longer wavelengths. As different im-
fuel-rich flames, while NO2 * radiation is the most important contribu- age intensifier gains prohibit the direct comparability of the intensities
tor for fuel-lean conditions. Equivalence ratios are non-uniform in the of the two operating points compared, the intensities (OH* and NH*)
spray combustion case investigated and subject to temporal and spatial are normalized with the highest value of OH*-intensity occurring for
variations. However, the general trend as combustion progresses and the respective operating point. Due to more reactive conditions, the
the lift-off length increases is a shift from conditions that show more pilot ignition delay is shorter for OP4, which results in an earlier onset
NH2 * radiation (high equivalence ratios, smaller amount of gas in post- of OH*-CL and heat release (see Fig. 5). The HRR of OP4 shows a more
flame zone) to conditions that show more NO2 *-CL (lower equivalence prolonged delay between pilot combustion and ammonia combustion
ratios, more gas in post-flame zone). In full engines, the pressure and due to the earlier pilot ignition. While the shape of OH*-CL intensity is
temperature rise due to combustion usually leads to ongoing reactions similar for both cases, the absolute OH* intensity was much higher for
in the burned charge, for example, caused by shifts in the chemical OP4 due to higher temperatures. Higher combustion temperatures in

7
V. Scharl and T. Sattelmayer Fuel 358 (2024) 130201

Fig. 7. SG (top), OH*-CL with overlayed SG contours (middle) and NL (bottom) imaging obtained at OP3 (865 k, 98 bar). OH* intensities are 𝛾-corrected (𝛾 = 0.4). The images are
taken during the main fuel combustion phase, in which ammonia is burning (𝑡 = 2000–5000 μs ASOI). The combustion phases are also indicated via the HRRs at the top.

OP4 result from higher air temperature and from higher equivalence 4. Summary and conclusions
ratios due to less premixing, which results from the shorter ignition
delay. A slightly earlier onset of NH* emissions for OP3 can be at-
We investigated the flame emissions of diesel-piloted ammonia
tributed to earlier interaction between ammonia and diesel due to
sprays in an RCEM. For this purpose, two single-hole injectors were
ammonia’s faster penetration at lower charge air densities, as is ap-
employed. The pilot injector provided a small diesel fuel injection (5%
parent from well-known spray penetration correlations (e.g., [38,39]).
of total LHV supplied) followed by the ammonia main injection from
As a result of the earlier interaction, ammonia also ignites earlier.
the second injector. Spectrally resolved CL, SG, OH*-CL, NL imaging,
NH*-CL intensity is lower for OP4 relative to OH* intensity, which
and HRR analysis were employed to obtain profound insights into the
results from the more intense OH*-CL mentioned above. Both NH*
combustion process.
intensities peak twice, directly after ammonia ignition and before the
Based on spatially and temporally integrated spectra, as well as
peak HRR. While the second peak is expected based on ammonia
temporally resolved spectra, three phases of the HPDF combustion
combustion intensity, the first peak is thought to result from increased
process were identified:
NH production as entrained diesel combustion products lead to higher
combustion temperatures [40]. Furthermore, the high OH concentra- • First, diesel pilot ignition and combustion lead to a peak in HRRs.
tion produced during diesel combustion increases NH concentration via The flame emissions were typical for diesel flames featuring
the reaction NH2 + OH → NH + H2 O [19]. As expected, temperature, OH*, CH*, the sodium D-line, UV-broadband luminosity (HCO*,
pressure, equivalence ratio, and strain rate variations prohibit any HCOH*, CO2 *), as well as visible broadband luminosity (H2 O*,
direct correlation between HRR and CL intensity. NO2 * and soot).

8
V. Scharl and T. Sattelmayer Fuel 358 (2024) 130201

Fig. 8. Spatially integrated flame emission spectra at different times during diesel pilot ignition (a), flame transition from diesel to ammonia (b) and ammonia combustion (c) at
OP3.

OH*-CL and NL showed a strong spatial correlation. However, OH*-


CL dropped during ammonia combustion until it could not be detected
reliably anymore. We also found that the ammonia spray first pene-
trates through the diesel products without being ignited at its tip during
the flame transition phase, which constitutes a possible mechanism for
unburned ammonia emissions.
The peak areas of NH*, OH*, and NH2 * were extracted from the
emissions spectra to discuss their evolution. A strong correlation be-
tween NH2 * peak intensities and broadband CL intensity during the
ammonia combustion phase suggested that most ammonia flame emis-
sions originated from NH2 *-CL. Additional background radiation during
late ammonia combustion originated from NO2 *-CL in lean post-flame
zones. Increasing combustion chamber temperature and pressure leads
to an earlier onset of diesel combustion. However, the first occur-
rence of ammonia-specific flame emissions did not change significantly,
indicating the high importance of spray interaction for the onset of
ammonia combustion.
Apart from improving the understanding of the HPDF combustion
process, the insights obtained in this study will help choose suitable
methods for future optical investigations of diesel-piloted ammonia
combustion. For cases with low soot formation, we suggest focusing on
wavelengths that capture NH2 *-CL, as they offer a strong signal, which
mainly originates from the flame front during ammonia combustion.
Although UV flame emissions are much stronger during pilot com-
Fig. 9. Temporal evolution of normalized background corrected CL intensities. (a): bustion, cases with strongly sooting diesel combustion might require
HRRs and mass fluxes to facilitate the allocation of flame emissions to the combustion observing wavelengths in the UV range. The best option available is
phases (see Fig. 5), (b): OH* and NH2 *, as well as broadband luminosity for the less the observation of established OH*-CL wavelengths (around 308 nm),
reactive operating point (OP3). (c): OH* and NH* for the less reactive operating point
(OP3) and the highly reactive operating point (OP4).
as they offer the strongest absolute and relative signal during ammonia
combustion. Observing the wavelengths relevant for NH*-CL (around
336 nm) is not recommended, as diesel UV broadband emissions are
strong relative to the intensity of the narrow NH* peak in this range.
• After interaction with the ammonia spray, the flame transition
started, during which flame emissions from NH3 and diesel com-
bustion were observed. HRRs only started to increase slowly. Nomenclature
Although both fuels burned simultaneously, CN*-CL was not de-
tected. Abbreviations
• The major share of heat release was found to occur during the ASOI after start of injection
ammonia-only combustion phase, in which OH*, NH*, NH2 * and btdc before top dead center
broadband luminosity at visible wavelengths (NO2 *, H2 O*, NH2 *) CAD crank angle degree
were observed. NO2 * provided a minor contribution to the overall CL chemiluminescence
broadband flame emission intensity.

9
V. Scharl and T. Sattelmayer Fuel 358 (2024) 130201

FWHM full width at half maximum [10] Dec JE. Dec 1997 - a conceptual model of DI diesel combustion based on laser
HPDF high-pressure dual-fuel sheet imaging. 1997.
HRR heat release rate [11] Pastor JV, García-Oliver JM, García A, Micó C, Durrett R. A spectroscopy study of
LHV lower heating value gasoline partially premixed compression ignition spark assisted combustion. Appl
Energy 2013;104:568–75. http://dx.doi.org/10.1016/j.apenergy.2012.11.030.
NL natural luminosity
[12] Srna A, Bombach R, Herrmann K, Bruneaux G. Characterization of the spectral
OP operating point signature of dual-fuel combustion luminosity: Implications for evaluation of
rpm revolutions per minute natural luminosity imaging. Appl Phys B 2019;125(7). http://dx.doi.org/10.
RCEM rapid compression expansion machine 1007/s00340-019-7222-z.
SG shadowgraphy [13] Gaydon A. The spectroscopy of flames. Dordrecht: Springer Netherlands; 1974,
URL https://ebookcentral.proquest.com/lib/kxp/detail.action?docID=6560931.
TDC top dead center
[14] Wang N, Li T, Zhou X, Zhang Z, Chen R, Li S. High-pressure spray and
UV ultra-violet combustion characteristics of ammonia jets under diesel-like conditions. In:
VIS visible The proceedings of the international symposium on diagnostics and modeling
of combustion in internal combustion engines. Vol. 2022.10. 2022, p. B7–1.
CRediT authorship contribution statement http://dx.doi.org/10.1299/jmsesdm.2022.10.B7-1.
[15] Wüthrich S, Cartier P, Süess P, Schneider B, Obrecht P, Herrmann K. Opti-
cal investigation and thermodynamic analysis of premixed ammonia dual-fuel
Valentin Scharl: Conceptualization, Methodology, Investigation, combustion initiated by dodecane pilot fuel. Fuel Commun 2022;12:100074.
Writing – original draft, Visualization, Project administration. http://dx.doi.org/10.1016/j.jfueco.2022.100074.
Thomas Sattelmayer: Conceptualization, Supervision, Funding [16] Scharl V, Lackovic T, Sattelmayer T. Characterization of ammonia spray com-
acquisition, Project administration, Writing – review & editing. bustion and mixture formation under high-pressure, direct injection conditions.
Fuel 2023;333:126454. http://dx.doi.org/10.1016/j.fuel.2022.126454.
[17] Scharl V, Sattelmayer T. Investigation of post-injections for emission reduction
Declaration of competing interest of diesel-piloted ammonia spray combustion. In: 30th CIMAC congress. 2023.
[18] Zhu X, Khateeb AA, Roberts WL, Guiberti TF. Chemiluminescence signature
The authors declare that they have no known competing financial of premixed ammonia-methane-air flames. Combust Flame 2021;231:111508.
interests or personal relationships that could have appeared to http://dx.doi.org/10.1016/j.combustflame.2021.111508.
[19] Mashruk S, Vigueras-Zuniga MO, Tejeda-del Cueto ME, Xiao H, Yu C, Maas U, et
influence the work reported in this paper.
al. Combustion features of CH4/NH3/H2 ternary blends. Int J Hydrogen Energy
2022;47(70):30315–27. http://dx.doi.org/10.1016/j.ijhydene.2022.03.254.
Data availability [20] Ichikawa Y, Niki Y, Takasaki K, Kobayashi H, Miyanagi A. Experimental study of
combustion process of NH3 stratified spray using imaging methods for NH3 fu-
Data will be made available on request. eled large two-stroke marine engine. Appl Energy Combust Sci 2023;13:100119.
http://dx.doi.org/10.1016/j.jaecs.2023.100119.
[21] Dumitrescu S, Hill PG, Li G, Ouellette P. Effects of injection changes on efficiency
Acknowledgments and emissions of a diesel engine fueled by direct injection of natural gas. In:
SAE technical paper series. SAE technical paper series, SAE International400
This research is funded by the German Federal Ministry for Eco- Commonwealth Drive, Warrendale, PA, United States; 2000, http://dx.doi.org/
10.4271/2000-01-1805.
nomic Affairs and Climate Action on the basis of a decision by the
[22] Fink G, Jud M, Sattelmayer T. Influence of the spatial and temporal interaction
German Bundestag (project no. 03SX534D) which is gratefully ac- between diesel pilot and directly injected natural gas jet on ignition and
knowledged. The authors would like to acknowledge the support of combustion characteristics. J Eng Gas Turbines Power 2018;140(10). http://dx.
the AmmoniaMot consortium (Woodward L’Orange, MAN Energy Solu- doi.org/10.1115/1.4039934.
tions, WTZ Roßlau and Neptun Ship Design). Furthermore, the authors [23] Scharl V, Sattelmayer T. Ignition and combustion characteristics of diesel piloted
ammonia injections. Fuel Commun 2022;11:100068. http://dx.doi.org/10.1016/
would like to express their sincere gratitude to Shane Martin for his
j.jfueco.2022.100068.
support. [24] Dorer F, Prechtl P, Mayinger F. Investigation of mixture formation and combus-
tion prozesses in a hydrogen fueled diesel engine. In: Saetre TO, editor. Hydrogen
References power: theoretical and engineering solutions. Dordrecht: Springer Netherlands;
1998, p. 49–54. http://dx.doi.org/10.1007/978-94-015-9054-9_6.
[1] International Energy Agency - IEA. The future of hydrogen: Technology report. [25] Lee D, Hochgreb S. Rapid compression machines: Heat transfer and suppression
2019, URL https://www.iea.org/reports/the-future-of-hydrogen. of corner vortex. Combust Flame 1998;114(3–4):531–45. http://dx.doi.org/10.
[2] Gray JT, Dimitroff E, Meckel NT, Quillian RD. Ammonia fuel - engine compatibil- 1016/S0010-2180(97)00327-1.
ity and combustion: technical paper 660156. SAE technical paper series, vol.75, [26] Fink G, Jud M, Sattelmayer T. Fundamental study of diesel-piloted natural gas
SAE International; 1966, p. 660002–164. http://dx.doi.org/10.4271/660156. direct injection under different operating conditions. J Eng Gas Turbines Power
[3] Bro K, Pedersen PS. Alternative diesel engine fuels: an experimental investigation 2019;141(9). http://dx.doi.org/10.1115/1.4043643.
of methanol, ethanol, methane and ammonia in a d.i. diesel engine with pilot [27] Coppo M, Negri C, Wermuth N, Garcia-Oliver J. Powering a greener future: The
injection: technical paper 770794. SAE technical paper series, SAE International; OMT injector enables high-pressure injection of ammonia and methanol. In: 30th
1977, http://dx.doi.org/10.4271/770794. CIMAC congress busan. 2023.
[4] Starkman ES, James GE, Newhall HK. Ammonia as a diesel engine fuel: Theory [28] Gaucherand J, Netzer C, Lewandowski MT, Løvås T. Modelling of liquid
and application: Technical paper 670946. SAE Trans 1968;3193–212, , http: injection of ammonia in a direct injector using Reynolds-averaged Navier–
//www.jstor.org/stable/44562852. Stokes simulation. In: Proceedings of the 63rd international conference of
[5] Tornatore C, Marchitto L, Sabia P, de Joannon M. Ammonia as green fuel in scandinavian simulation society. Linköping electronic conference proceedings,
internal combustion engines: State-of-the-art and future perspectives. Front Mech Linköping University Electronic Press; 2022, p. 405–12. http://dx.doi.org/10.
Eng 2022;8. http://dx.doi.org/10.3389/fmech.2022.944201. 3384/ecp192058.
[6] Chiong M-C, Chong CT, Ng J-H, Mashruk S, Chong WWF, Samiran NA, et al. [29] Weng W, Aldén M, Li Z. Visible chemiluminescence of ammonia premixed flames
Advancements of combustion technologies in the ammonia-fuelled engines. En- and its application for flame diagnostics. Proc Combust Inst 2023;39(4):4327–34.
ergy Convers Manage 2021;244:114460. http://dx.doi.org/10.1016/j.enconman. http://dx.doi.org/10.1016/j.proci.2022.08.012.
2021.114460. [30] Sheehe SL, Jackson SI. Identification of species from visible and near-infrared
[7] Lauer M, Sattelmayer T. On the adequacy of chemiluminescence as a measure spectral emission of a nitromethane-air diffusion flame. J Mol Spectrosc
for heat release in turbulent flames with mixture gradients. J Eng Gas Turbines 2019;364:111185. http://dx.doi.org/10.1016/j.jms.2019.111185.
Power 2010;132(6). http://dx.doi.org/10.1115/1.4000126. [31] Fontijn A, Meyer CB, Schiff HI. Absolute quantum yield measurements of the
[8] Ballester J, García-Armingol T. Diagnostic techniques for the monitoring and NO–O reaction and its use as a standard for chemiluminescent reactions. J Chem
control of practical flames. Prog Energy Combust Sci 2010;36(4):375–411. http: Phys 1964;40(1):64–70. http://dx.doi.org/10.1063/1.1724895.
//dx.doi.org/10.1016/j.pecs.2009.11.005. [32] Mancaruso E, Vaglieco BM. Spectroscopic analysis of the phases of premixed
[9] Fujimoto H, Kurata K, Asai G, Senda J. OH radical generation and soot forma- combustion in a compression ignition engine fuelled with diesel and ethanol.
tion/oxidation in DI diesel engine. SAE technical paper series, SAE International; Appl Energy 2015;143:164–75. http://dx.doi.org/10.1016/j.apenergy.2015.01.
1998, http://dx.doi.org/10.4271/982630. 031.

10
V. Scharl and T. Sattelmayer Fuel 358 (2024) 130201

[33] Augusta R, Foster DE, Ghandhi JB, Eng J, Najt PM. Chemiluminescence mea- [38] Naber J, Siebers DL. Effects of gas density and vaporization on penetration and
surements of homogeneous charge compression ignition (HCCI) combustion. In: dispersion of diesel sprays: technical paper 960034. SAE technical paper series,
SAE technical paper series. SAE technical paper series, SAE International400 SAE International; 1996, http://dx.doi.org/10.4271/960034.
Commonwealth Drive, Warrendale, PA, United States; 2006, http://dx.doi.org/ [39] Hiroyasu H, Arai M. Structures of fuel sprays in diesel engines. SAE Trans
10.4271/2006-01-1520. 1990;(99):1050–61. http://dx.doi.org/10.4271/900475.
[34] Rizk W. Experimental studies of the mixing processes and flow configurations [40] Klippenstein SJ, Harding LB, Glarborg P, Miller JA. The role of NNH in NO
in two-cycle engine scavenging. Proc Inst Mech Eng 1958;172(1):417–37. http: formation and control. Combust Flame 2011;158(4):774–89. http://dx.doi.org/
//dx.doi.org/10.1243/PIME_PROC_1958_172_037_02. 10.1016/j.combustflame.2010.12.013.
[35] Musculus MPB, Kattke K. Entrainment waves in diesel jets. SAE Int J Eng
2009;2(1):1170–93. http://dx.doi.org/10.4271/2009-01-1355.
[36] Pickett LM, Siebers DL, Idicheria CA. Relationship between ignition processes and
the lift-off length of diesel fuel jets. SAE technical paper series, SAE International;
2005, http://dx.doi.org/10.4271/2005-01-3843.
[37] Miller JA, Bowman CT. Mechanism and modeling of nitrogen chemistry in
combustion. Prog Energy Combust Sci 1989;15(4):287–338. http://dx.doi.org/
10.1016/0360-1285(89)90017-8.

11

You might also like