Sei sulla pagina 1di 12

DOI: 10.1002/chem.

201602676

Concept

& High-Temperature Chemistry

Combustion Chemistry Diagnostics for Cleaner Processes


Katharina Kohse-Hinghaus*[a]

Chem. Eur. J. 2016, 22, 13390 13401

13390

T 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Concept
Abstract: Climate change, environmental problems, urban
pollution, and the dependence on fossil fuels demand
cleaner, renewable energy strategies. However, they also
ask for urgent advances in combustion science to reduce
emissions. For alternative fuels and new combustion regimes, crucial information about the chemical reactions
from fuel to exhaust remains lacking. Understanding such
relations between combustion process, fuel, and emissions needs reliable experimental data from a wide range
of conditions to provide a firm basis for predictive modeling of practical combustion processes.

1. New Combustion Approaches


Combustion still dominates global primary energy consumption, with a fraction of more than 80 % from fossil fuels.[1] Combustion heat is also used for large-scale industrial processes
such as the production of glass and ceramics. Combustion processes are known to emit carbon dioxide and undesired pollutants, including carbon monoxide, nitrogen oxides (NOx), polycyclic aromatic hydrocarbons (PAHs) and soot as well as further
species including, for example, alkenes, aromatics, and carbonyl compounds. They contribute to particulate pollution[2] and
secondary aerosol formation[3] in many regions of the world,
especially in megacities with high vehicle traffic. Although air
quality, environmental, climate, and health issues demand to
replace combustion by sustainable and renewable energy
sources, this transformative process will need substantially
more than a decade because of the scale of energy needed;
the growth of population, industrial production, and transportation; and the considerable lifetimes of existing technology.
Advanced combustion strategies for the mid-term future must
thus strive for improved efficiency, CO2 savings, and reduction
of harmful emissions. Two representative examples for such
new combustion technology developments from the area of
automotive engines are given in the following.
Low-temperature combustion (LTC) processes are expected
to significantly reduce NOx and particulate matter (soot).[4, 5]
This strategy is illustrated in Figure 1,[6] where the local mixture
of fuel and oxygen, represented by the equivalence ratio f, is
given as a function of the local temperature. Combustion regimes are indicated where the formation of NOx and soot
emissions must be expected: Whereas NOx is formed particularly at high temperatures and near-stoichiometric (f& 1) fueloxygen mixtures, soot develops mainly in fuel-rich zones (f+
2). The operating conditions of a conventional diesel engine
favor the formation of both types of undesired emissions.
Advanced low-emission combustion processes in the LTC
regime, including homogeneous charge compression ignition
[a] Prof. Dr. K. Kohse-Hinghaus
Department of Chemistry, Bielefeld University, Universit-tsstrae 25
33615 Bielefeld (Germany)
E-mail: kkh@uni-bielefeld.de
Chem. Eur. J. 2016, 22, 13390 13401

www.chemeurj.org

Figure 1. Comparison of diesel fuel combustion modes in a diagram of local


equivalence ratio vs. local temperature, from Lu et al.;[6] LTC: low-temperature combustion, HCCI: homogeneous charge compression ignition, PCCI:
premixed charge compression ignition. Reprinted with permission by Elsevier from Ref. [6].

(HCCI) and premixed charge compression ignition (PCCI), are


being designed to provide almost homogeneous mixing
before ignition and to operate at considerably lower than conventional temperatures, practically achieved by significant exhaust gas recirculation (EGR). These combustion processes, in
principle adaptable to a wide spectrum of fuels, need considerable engine management and control. Regarding the fuel-specific auto-ignition reactions, ignition timing is crucial and demands intricate knowledge of the low-temperature fuel-specific oxidation chemistry. A cartoon that represents a low-temperature diesel concept, adapted by Bergthorson and Thomson[7]
from a number of recent studies, is shown in Figure 2. With
higher premixing and reduced temperature, low-temperature
strategies can lead to the simultaneous reduction of NOx and
soot; compare the shaded areas for LTC, PCCI, and HCCI combustion with that for the conventional diesel process in
Figure 1. A partially premixed flame (PPF in Figure 2) evolves
that is broader and longer than in a usual diesel combustion.
Since no soot is formed, the typical yellow flame color from
the black-body radiation of hot soot particles will not be observed; rather, the flame will appear blue due to chemiluminescence of excited species (denoted with an asterisk*) such
as CH2O* and the CH* radical.
Biofuels can contribute to CO2 reductions, but their combustion may lead to new or different emissions, including oxygenated pollutants such as aldehydes and ketones.[8, 9] Prediction
of the combustion behavior and emissions of practical/industrial systems as well as their optimization and control towards
more efficient and cleaner performance thus need detailed information on the underlying physico-chemical reaction processes including the ignition and low-temperature oxidation
behavior of conventional and alternative fuels and their
blends, and pathways leading to undesired emissions.
With regard to transportation, an area where replacement of
the present high-density fuels by more sustainable alternatives
is particularly urgent, combustion reactions of conventional petroleum-based fuels have been studied since a long time,

13391

T 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Concept

Figure 2. Cartoon of a low-temperature diesel combustion process; PPF: partially-premixed flame, from Bergthorson and Thomson.[7] Reprinted with permission by Elsevier from Ref. [7].

whereas the knowledge on alternative fuels is considerably


less established. While replacement fuels are being suggested,
including, e.g., biodiesel[10] and dimethyl ether (DME)[11] as
diesel substitutes, detailed combustion kinetics of biodiesel
esters[12, 13] and even for the rather simple fuel molecule
DME[14, 15] are still under study, especially in the LTC regime. The
physico-chemical combustion mechanisms must be understood, however, as an important prerequisite to optimize the
process. Such knowledge will assist explaining how auto-ignition in the conventional and LTC regimes depends on the
fuels chemical nature. Understanding the reaction details will
also help to assess how fuel improvers, additives, and
blendsdue to their chemical structurewill influence the
combustion reactions. Such information will potentially prove
suitable to reduce or block chemical pathways that form hazardous emissions. To obtain and use such information, the
combustion process must be inspected directly while it happens, with in-situ diagnostics.

2. Diagnostics in Model Combustors and Engines


Combustion processes in engines and gas turbines used, for
example, for road and air transportation, present themselves
as difficult and challenging to diagnostics. The phases before,
during, and after combustion in a technical combustion
system are, in principle, accessible to diagnostics with a multitude of different analysis techniques. Ex situ or off-line standard chemical analytics, however, will not permit to fully understand or optimize the actual combustion process. In a real
engine, complex fluid dynamics, mixing, fuel evaporation,
changing pressure and temperature occur in combination with
complex chemistry of thousands of reactions among hundreds
of species[6] and provide an environment with extreme conditions for in-situ diagnostics. One of the main challenges is the
multi-scale nature of the combustion process: It spans about
ten orders of magnitude in dimension from the molecular
Chem. Eur. J. 2016, 22, 13390 13401

www.chemeurj.org

scale of chemical reactions to the macroscopic size of the


engine or combustor, and it encompasses a similar range in
time from sub-nanosecond molecular rearrangements up to
seconds needed for mass transport. Direct inspection of practically relevant combustion systems thus must ideally provide
appropriately resolved information in space-time for all relevant quantities, requirements that should be met with multispecies or multi-quantity four-dimensional (4D) measurements.
More realistically, the in-situ characterization of a system must
rely on techniques that yield the most relevant information for
a given purpose. Such combinative measurements can jointly
address, for example, flow field structure and flame front, or
temporal development of individual vortices and their interaction with the flame, or spatial images of a few important parameters that define the reaction progress.
Optical, and especially, laser techniques belong to the most
suitable methods to permit in-situ combustion analyses, and
they are often applied in combinations.[16, 17] Mostly they are
non-invasive and offer respectable temporal and spatial resolution; based on a few selected quantities they can provide
flow-stopping momentary images in two and three dimensions or time sequences as movies of the process. Characteristics of the flow field can be obtained as well as temperature
and concentration fields for many major or minor species.
Recent reviews have focused on high-speed imaging of combustion phenomena;[18] visualization of combustion properties;[19] on velocity, temperature, and species measurements;[20, 21] on optical investigations of flame-wall interactions;[22] or on soot formation studies with optical methods.[23]
Significant progress has been made recently, and milestones
include hybrid femtosecond/picosecond coherent anti-Stokes
Raman scattering (CARS) techniques to study flame-wall interactions,[24] high-speed combinations of 2D particle imaging velocimetry (PIV) with simultaneous planar laser-induced fluorescence (PLIF) imaging to study misfiring cycles in a dedicated
optical engine,[25] and high-speed fiber-optic diode laser techniques to analyze the dynamic conditions in a near-production
engine for high EGR.[26] Similar progress to that of in-situ
engine diagnostics is being made for further combustion applications, including the characterization of the turbulent swirl
flames representative of gas turbine combustors. Advanced
high-speed imaging measurements combining chemiluminescence and PIV with PLIF of fuel tracers and species present in
the flame front permit deep insights into the flame-flow field
interaction, local extinction, and oscillations due to acoustic instabilities.[27] Besides velocity, temperature, selected concentrations and their spatio-temporal variations during the combustion process, laser diagnostics can provide important characteristics of combustion beyond direct observables, including, for
example, heat release from simultaneous PLIF images of the
OH radical and of formaldehyde, CH2O, that were multiplied
pixel-by-pixel to provide an indirect estimate of this quantity.[28]
The example of combinative laser diagnostics in Figure 3 is
taken from a recent investigation that has been instrumental
in developing new concepts of low-temperature engine combustion.[5] Multiple quantities have been measured, and only

13392

T 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Concept
a small part of the information is presented for an LTC condition of a heavy-duty engine running on a fuel blend of n-heptane, isooctane, and toluene. Color-coded PLIF images of CH2O
on the left are combined with stoichiometry contours from the
same conditions on the right, obtained from toluene PLIF. In
all images, the injector is on the left side, and the piston bowl
on the right; a cross section through the jet from fuel injection
is thus analyzed. Under these conditions, first-stage ignition
was determined from further laser measurements to happen at
a crank angle of 68 after start of injection (ASI), and formaldehyde as a reaction intermediate appears suddenly at 78 ASI.
Note that the color code in the two CH2O images does not
represent the relative intensity but the spectral content of
CH2O, calibrated from ensemble-averaged PLIF images, since
other fluorescence signatures such as from PAHs could interfere in this spectral region. The ensemble-averaged contour
maps on the right show a broad range of equivalence ratios to
be present in the jet, from fuel-lean at f< 1 to very rich at f>
3.5. Ignition in the LTC regime can occur for this whole range
of fuel-air mixtures. However, it was not possible to simulate
that behavior consistently with available combustion models.[5]

Figure 3. Left: Single-shot images of formaldehyde PLIF for a late-injection


condition in a heavy-duty engine. The color bar indicates the relative contribution of formaldehyde to the total fluorescence. Right: Ensemble-averaged
equivalence ratio contours measured with toluene fuel-tracer PLIF under
oxygen-free (non-combustion) conditions at the same thermodynamic state
as the formaldehyde PLIF images, from Musculus et al.[5] Reprinted with permission by Elsevier from Ref. [5].

While a wealth of information can be obtained today directly from a running near-production engine or a model gas turbine combustor, the analysis requires coupling of several advanced laser-optical techniques, including multi-dimensional,
highly spatially resolved, high-speed imaging of the flow field
and of some selected chemical constituents. Such combined
instrumentation is available for this purpose in few specialized
laboratories around the world, and more details become detectable, for example, with improved equipment including
faster lasers, new wavelength regimes, higher repetition rates,
more sensitive detectors, and high-quality optics and sensors.
Only few selected species, such as some small molecules and
radicals, can typically be probed, and many important details
Chem. Eur. J. 2016, 22, 13390 13401

www.chemeurj.org

of the ongoing chemical reactions are not revealed from these


species, although such information would be highly desirable.
This is true especially for LTC conditions with the need to address fuel-specific auto-ignition as well as the cycle-resolved
composition of unburnt hydrocarbons (HC), since due to EGR,
chemical species from one cycle will influence the next.

3. Diagnostics To Validate Chemical Mechanisms and Combustion Models


Rather than in engines, the fundamental physico-chemical
combustion reactions must be studied in dedicated, specifically designed laboratory-scale set-ups under well-characterized
and idealized pressure, temperature, flow, and stoichiometry
conditions, often quite different from those of practical systems. Not just phenomenological reaction models, but detailed
reaction mechanisms built from information on these fundamental processes are needed to transfer the physico-chemical
knowledge obtained from laboratory investigations to technical combustion processes.[29, 30] Such mechanisms can only be
predictive for a large range of conditions if they are critically
validated against highest-quality experimental data. Validation
today often relies on rather global information, obtained from,
for example, ignition delay times[31] or flame speeds.[32] However, more detailed species analysis is necessary for a comprehensive validation and can be provided for a limited range of conditions, for example, from jet-stirred reactor experiments,[33, 34]
typically coupled with GC analysis, or from laminar premixed
low-pressure flame investigations, often using molecular-beam
mass spectrometric (MBMS) detection.[35, 36] The structure-selective identification of numerous stable, labile, and reactive intermediate species is often the first step to improve mechanisms
that may not have considered reactions of a given species or
family of species. Examples include the consideration and detection of NCN[37, 38] for NOx formation and of enols[39] in flames
that have had a significant influence on mechanism development.
The chemical reaction sequences, and consequently intermediates, depend on the structure of the fuel molecule. With
alternative (bio)fuels, different species can occur in addition to
the species pool from hydrocarbon combustion.[9] Further diagnostic challenges arise in the low-temperature ignition regime
where numerous fuel-specific labile oxygen-rich intermediates
occurstructures only recently detected experimentally.[40, 41]
Exhaust gas recirculation will add potentially reactive CO2, H2O,
and HC species to a fuel mixture. Fuel blends, additives, and
fuel improvers may lead to interaction of the reaction sequences through intermediates from different fuel components. The
highly complex chemistry of PAH and soot formation proceeds
along higher-molecular intermediates and poses extreme difficulties to experimentally corroborate structures between the
molecular phase with 34 aromatic rings and the nanometersized nucleating particles.[42, 43]
Single-photon ionization molecular-beam mass spectrometry
(PI-MBMS) using tunable vacuum ultraviolet radiation from synchrotrons has become one of the most powerful, universally
applicable strategies for versatile, multiplexed in-situ chemical

13393

T 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Concept
analysis in laboratory reactors and flames.[35, 36, 4446] A recent example for isomer identification with PI-MBMS is given in
Figure 4.[15] Here, DME oxidation was analyzed in a jet-stirred
reactor at 540 K and atmospheric pressure, conditions representative of low-temperature ignition chemistry. Numerous
species in the range 2100 u were identified from their exact
masses and ionization thresholds, obtained by scanning the
energy of the ionizing photons. For an observed signal at
60.021 u, corresponding to species with the formula C2H4O2,
two ionization thresholds at 10.05 (: 0.05) eV and 10.80 (:
0.05) eV were seen that match the known or newly calculated
values for methyl formate, HC(O)OCH3 (10.835 eV) and for 1,3dioxetane, cyclic-CH2-O-CH2-O- (10.08 eV). Methyl formate had
been previously detected under such conditions, and its reactions had been included into detailed mechanisms for DME
before, whereas 1,3-dioxetane was newly identified by comparison with theoretically calculated adiabatic ionization energies
of C2H4O2 isomers.[15] The ability of synchrotron PI-MBMS to
detect, in principle, all species with high sensitivity down to
the sub-ppm level, including labile intermediates, is highly attractive, but it needs precious beam time at the synchrotron
multi-user facilities.

Established versatile techniques for chemical analysis of


stable intermediates and products are valuable complements
in combustion studies.[5155] For example, Harper et al.[52] have
analyzed the pyrolysis of the attractive biofuel n-butanol[30] in
a jet-stirred reactor with GC V GC-MS and have derived a fuel
decomposition mechanism from the main identified species
that include formaldehyde, ethanol, propanol, butanal, benzene, n-butanol, toluene, ethylbenzene, butyl acetate, and
naphthalene. A resulting scheme supported by this data is
shown in Figure 5. All pathways start from H-abstraction reactions by H and CH3 to form the different C4H9O radicals that
decompose to some of the stable products seen in the experiment. The reaction flux analysis in Figure 5 showing percentages for specific reaction sequences for a given set of conditions
was obtained from a comprehensive reaction mechanism developed in this study.[52]
Quantitative species measurements especially for radicals
are considerably more ambitious than species identification.
Calibration strategies as well as reliable absorption and ionization cross sections are needed, fragmentation of labile species,
probe effects, interferences, and other experimental influences
must be considered. Combinative approaches are significantly
more powerful than single-technique measurements, since
they rely on different physical effects and calibration procedures, and can thus complement each other and mutually support the information obtained. Extensive species measurements under a broad condition range of conditions would contribute significantly to understanding the chemistry of alternative fuels and fuel blends in unconventional but promising
combustion regimes and reveal the roles of many labile species and isomeric structures involved in pollutant formation.

4. Opportunities for Diagnostics


4.1. Fuel-Specific Auto-Ignition

Figure 4. Photoionization efficiency curve for 60.021 u, from Moshammer


et al.,[15] showing two ionization thresholds at 10.05(: 0.05) and
10.8(: 0.05) eV that match the ionization energies of 1,3-dioxetane and
methyl formate at 10.08 and 10.835 eV, respectively. Also shown is the photoionization efficiency curve of methyl formate, which also matches the experimentally observed PIE curve accurately. Reprinted with permission by
the American Chemical Society from Ref. [15].

Further recent diagnostics advances have been demonstrated that may hold promise in laboratory-scale combustion
chemistry analysis. These include, for example, laser techniques
for some individual targets, such as quantum-cascade laser absorption spectroscopy to detect labile species in a shock
tube[47] and infrared cavity ring-down spectroscopy to study
oxidation reactions in a jet-stirred reactor,[48] as well as universal analytical methods such as photoelectron-photoion coincidence spectroscopy that has only been recently applied to
flames.[49, 50] While such techniques are, in principle, capable to
detect reactive intermediates, they are still under development
for broader application.
Chem. Eur. J. 2016, 22, 13390 13401

www.chemeurj.org

One of the challenging yet essential topics in combustion


chemistry is the auto-ignition of fuels in the low-temperature
domain, since a large number of fuel-specific reactions are involved, and small variations of the temperature and pressure
conditions can have enormous effects. To predict the ignition
process with reasonable accuracy, comprehensive chemical reaction mechanisms for fuel oxidation and combustion must be
built up systematically from fundamental principles, including
a multitude of reaction rate expressions as a function of pressure and temperature.[12, 29, 30, 56] Instead of developing a mechanism for each individual fuel, such systematically built mechanisms for larger fuels include sub-mechanisms for reaction
classes or families. Reaction rate coefficients are obtained from
theoretical calculations and dedicated experiments[5759] that
also may use advanced diagnostics, especially for radical reactions and when very labile species are involved. Formalisms
able to reliably predict pressure-dependent kinetics are essential[60] to bridge between the lower pressures used in speciesresolved laboratory combustion experiments and those in engines and gas turbines. Comprehensive mechanisms, carefully
validated for a number of conditions, will then be reduced to

13394

T 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Concept

Figure 5. Reaction pathway analysis for n-butanol pyrolysis at inlet and maximum temperatures of 473 8C and 807 8C, respectively, and a pressure of
1.72 V 105 Pa, from Harper et al.[52] The arrow thickness and percentages represent the reaction rate of decomposition for that species. Reprinted with permission by Elsevier/The Combustion Institute from Ref. [52].

a much smaller number of representative species and reactions


that can be handled in full engine models.
Validation for the LTC regime needs temperature- and pressure-dependent data relevant for auto-ignition. Typically, ignition delay times are measured that can be compared to comprehensive combustion models. An example for the two-stage
ignition of n-heptane, a prototypical fuel for diesel combustion, is given in Figure 6, using the mechanism of Seidel
et al.[61] in comparison with shock tube results[62] for 13.5 bar
(left) and 40 bar (right) and different stoichiometries. Generally,
the two-stage ignition behavior, characteristic for the longerchain alkanes in realistic petroleum-based fuels, is well reproduced by the model simulation. Early fuel consumption at temperatures below 600 K is followed by a regime with negative
temperature dependence where the overall reaction is slowed
down until high-temperature reactions consume the fuel rapidly near 1000 K.
Oxygen addition to hydrocarbon radicals and rapid rearrangements are involved in the low-temperature range, following a sequence of reactions from the fuel to oxygenated inter-

mediates schematically depicted in Figure 7.[63] The initial alkyl


radical, R, can react with molecular oxygen preferably at low
temperature and high pressure to form an RO2 radical that can
isomerize to a hydroperoxyalkyl radical, denoted as QOOH,
and undergo a series of further reactions. The competition between the formation of the R + O2 !RO2 reaction and the destruction of the fuel radical to smaller species contributes sensitively to the noted negative temperature dependence. For
successful ignition, sufficient radicals must be generated to
allow for chain branching. An important reaction is the lowtemperature H-abstraction from the fuel by HO2 that leads to
an alkylhydroperoxide ROOH which can decompose towards
reactive radicals. Further reactions start from QOOH, adding
a second oxygen molecule to form O2QOOH that in turn decomposes to a ketohydroperoxide, OQOOH, and an OH radical, with further reactions to form additional OH. In total, the
sequence in Figure 7 thus can generate three radicals for each
starting fuel radical in reactions that are highly sensitive to
temperature.[63] There are a number of complications not evident from the scheme, including intramolecular hydrogen

Figure 6. Ignition delay times t for n-heptane/air in a shock tube, from Seidel et al.[61] Left: 13.5 bar, right: 40 : 2 bar; solid lines correspond to model predictions, symbols to experiments reported in Ref. [62ac]. Reprinted with permission by Elsevier/The Combustion Institute from Ref. [61].
Chem. Eur. J. 2016, 22, 13390 13401

www.chemeurj.org

13395

T 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Concept
transfer reactions and cyclic transition states. Although the
core scheme is, in principle, quite well established,[64] further
reactions are being included, for example, by adding the
Korcek reactions leading to carbonyl species and carboxylic
acids,[59] for which experimental evidence from low-temperature oxidation experiments is becoming available.[15]
The crucial interplay between reactions involving labile oxygenated species that produce reactive radicals to support ignition or lead to their consumption slowing down the overall reaction, is shown in Figure 8 for a prototypical long-chain
alkane (such as n-heptane whose overall ignition behavior is
evident from Figure 6). The initial H-abstraction reactions now
form a number of radicals R with different structures that can
undergo the subsequent oxygen addition, isomerization, and
scission steps, and to verify their presence in the reaction
mechanism, even highly discriminative techniques such as synchrotron PI-MBMS come to their limits. However, general acceptance and corroboration of the overall scheme involving
such reaction classes may guide experiments to look for certain structures or decomposition products. It may also be attempted from such fundamental knowledge and experimental
observations to derive reduced models that focus on the prototypical behavior and controlling reaction in the different
stages of ignition and thus facilitate the description and prediction of the ignition behavior of families of fuels.[63]
Observation of the species including radicals and labile oxygenated intermediates in low-temperature oxidation and combustion is still in its early stages.[14, 15, 40] The overview mass

Figure 7. Schematic mechanism for low-temperature alkane oxidation and


auto-ignition chemistry, from Merchant et al.[63] who considered the work of
Z#dor et al.[64] and included the Korcek reaction.[59] Reprinted with permission by Elsevier/The Combustion Institute from Ref. [63].

spectra in DME oxidation at 540 K of Moshammer et al.[15] exhibit in the range of 25100 u, besides a signal from H2O2,
those of numerous oxygenated species of composition CxHyOz
with x = 1,2; y = 16; and z = 14. Assignment of those signals
to the contributing structures was possible in part by analyzing
fragmentation pathways and from extensive theoretical calculations. For example, species contributing to C2H4O4 at 92.011 u
could be identified through fragmentation and from calculated

Figure 8. Schematic of a long n-alkane fuel undergoing low-temperature peroxy chemistry, from Merchant et al.[63] As the carbon chain increases every site
can be treated as an equivalent site. The blue arrows are OH formation reactions while the red arrows show loss of OH radicals. The dotted arrows are reactions that compete with low-temperature branching reactions and delay the first stage ignition of the fuel. Reprinted with permission by Elsevier/The Combustion Institute from Ref. [53].
Chem. Eur. J. 2016, 22, 13390 13401

www.chemeurj.org

13396

T 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Concept
photoionization efficiency (PIE) curves. As an important support for the postulated low-temperature oxidation mechanism,
hydroperoxymethyl formate (HPMF), HOOCH2OCHO, could
thus be assigned and detected for the first time. This identification was possible by considering the observed fragments
with composition C2H3O4, CH4O3, and C2H3O2 arising from loss
of H, CO, and HO2 that appear at ionization thresholds theoretically calculated in that work. In the theoretical analysis of the
measured PIE curves, fifteen conformeric structures for HPMF
and their populations were considered, showing how cuttingedge kinetics must rely on a firm basis from experiment and
theory.
Wang et al.[14] very recently studied the oxidation of 2,5-dimethylhexane, a branched alkane with a size typically present
in realistic fuels, at 510 K and reported species with up to five
oxygen atoms, originating from addition of up to three oxygen
molecules. A wealth of structures was detected, and a few of
C8H16O5 composition at 192.10 u and their likely formation reactions are depicted in Figure 9.[14] Their unambiguous identification and quantification must be regarded as work in progress with significant unknown territory. Interestingly, similar
and even more highly oxygenated structures are also involved
in the atmospheric chemistry of compounds forming secondary organic aerosol,[65] and such reactions are thus attractive
targets for future work.

views summarize the status of regulated vehicle emissions, including NOx, and their dependence on biofuel addition.[6670]
Present knowledge primarily addresses physico-chemical parameters and technical boundary conditions relevant for NOx
emission. Overall reaction routes involved in NOx formation are
quite well known and include thermal NO from high-temperature zones, prompt NO from fuel-rich zones and NO from
fuel-bound nitrogen. Relevant influence parameters for (bio-)
diesel combustion can be linked with these major formation
mechanisms as has been pointed out recently by Sun et al.[68]
For example, double-bond structures in the fuel will influence
the ignition delay and the adiabatic flame temperature, the
latter with a direct relevance for thermal NOx. Similarly, fuelbound oxygen as found in biodiesel can influence the choice
of feasible stoichiometries, again with consequences for the
adiabatic flame temperature and NOx formation.
The basis for in-situ diagnostics of NO in engines that captures the formation under realistic pressure, temperature, and
mixing conditions, has been developed some time ago[7173]
and in-cylinder laser measurements in heavy-duty diesel[74] and
realistic gasoline engines have been successfully demonstrated.[75] Such diagnostic developments have enabled singlepulse instantaneous two-dimensional imaging of the NO mole
fraction during combustion and have revealed highly inhomogeneous NO distributions and large cycle-to-cycle variations.[75]
From phase-averaging such images for different crankshaft positions, a good correlation of exhaust values with in-cylinder
4.2. Pollutant Formation
NO concentrations was established, and parameters affecting
the engine performance could be assessed.[75] An example is
Similarly important as for fuel-specific auto-ignition is diagnostics to understand emissions formed from novel fuels or fuel
given in Figure 10: an enormous reduction (up to about tenmixtures or with advanced combustion concepts. Recent refold) of the in-cylinder NO concentration was achieved by increasing the level of EGR, thus lowering the combustion temperature (compare Figure 1). With technical
solutions for catalytic aftertreatment of NOx under
near-stoichiometric conditions, the urgency of in-cylinder measurements has seemingly decreased, and
newer diagnostics developments for NO measurements are scarce. However, new engine concepts
and conditions will contribute to renewed interest,
especially in view of problems to ensure control of
performance and compliance with regulations on the
road.
The formation of polycyclic aromatic hydrocarbons
and soot belongs to the most complex chemical
issues in combustion and has been investigated extensively as discussed in recent reviews.[42, 43] Important open questions remain, however, as to the
impact of the molecular fuel structure on the physico-chemical nature of the formed PAHs and emerging particles[7678] that should be understood and potentially predicted with the aid of reaction mechanisms.[7981] To characterize vehicle exhaust, a large
Figure 9. The reaction network of alternative isomerization of OOQOOH from the primary
[14]
fuel radical, from Wang et al., showing the most feasible pathways. The reaction senumber of techniques and instruments is typically
quence following the solid arrows, i.e., O2 addition, intramolecular H-abstraction, and deemployed[82] that can assess important parameters incomposition produce C8H16O5 species at 192.10 u; the decomposition of P(OOH)2, dashed
cluding size, number density, structure, and chemical
arrows, leads to hydroperoxy cyclic ethers and olefinic hydroperoxides. KDHP: keto-dihyeffects. Diagnostics to validate PAH and soot formadroperoxide, DHPCE: dihydroperoxy cyclic ether. Reprinted with permission by Elsevier/
The Combustion Institute from Ref. [14].
tion models is, however, more challenging, because it
Chem. Eur. J. 2016, 22, 13390 13401

www.chemeurj.org

13397

T 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Concept

Figure 10. Average NO concentrations for different EGR ratio from Bessler
et al.[75 j Reprinted with permission by The Combustion Institute from
Ref. [75].

must couple the molecular composition of the reacting system


with the nucleation of nanoparticles and their properties. It encounters a dark phase that extends from small molecular radicals to species with 34 aromatic rings and further via carbonrich clusters to structures of about a nanometer size. Such
structures from PAH and soot growth models are shown in Figures 11 and 12.
While pathways leading to the first and second aromatic
rings as starting points for higher-molecular growth have been
extensively investigated, the reactions leading further from
those species to structures with about four aromatic rings are
not probed and understood in much detail. With theoretical
methods, transition states for some PAH conversion pathways
are being analyzed,[79] albeit without experimental corrobora-

Figure 12. Low-energy molecular clusters of (a) 50 pyrene molecules and


(b) 50 coronene molecules, where E/N is the total interaction energy per
molecule, from Totton et al.[80] Reprinted with permission by Elsevier/The
Combustion Institute from Ref. [80].

tion for the existence of these particular structures (Figure 11).


When combining smaller four-ring aromatics such as pyrene,
or larger ones such as coronene, by stacking and minimizing
energies in the resulting clusters, particle-like structures arise
with sizes of about 1 nm in the case where 50 coronenes were
assembled (Figure 12).[80] While such clusters appear highly
suggestive and may resemble emerging soot particles, no diagnostics today will support that they exist and play a role in
the soot formation process.
Nevertheless, diagnostics methods should be combined that
would allow access to all relevant structures along this growth

Figure 11. Potential energy diagram showing energies of the chemical species (CS) and transition states (TS) involved in the conversion of benzo[c]phenanthrene to pyrene at 0 K, from Raj et al.[79] Reprinted with permission by Elsevier/The American Carbon Society from Ref. [79].
Chem. Eur. J. 2016, 22, 13390 13401

www.chemeurj.org

13398

T 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Concept
sequence, from species with a few carbon atoms to particles
containing thousands. Recent advances include mass-spectrometric analysis of high-molecular-weight soot precursors,[83]
laser applications such as jet-cooled laser fluorescence[84] and
laser-induced incandescence[85] as well as high-resolution electron and helium ion microscopy.[86, 87] Nevertheless, in spite of
such advances and further diagnostics developments, structures like those in Figures 11 and 12 cannot be detected unambiguously. To understand relationships between fuel structure
and soot particle properties that will have climate, environmental, and health effects, theoretical prediction and sound
physico-chemical models must be combined and critically examined with decisive experimentsa situation with great opportunities for advanced diagnostics.

5. Conclusion
Combinative diagnostics is significantly more powerful than
single-technique, single-condition measurements and therefore
holds high promise for combustion chemistry analysis and
mechanism validation. Structure-selective physico-chemical information on the combustion chemistry of prototypical fossil
and alternative fuels should be made available to significantly
enhance the fundamental understanding of such processes as
ignition, low-temperature oxidation, and formation of harmful
species such as nitrogen oxides, polycyclic aromatic hydrocarbons, and soot. Non-regulated toxics such as certain carbonyls
or other harmful species can be emitted as a function of fuel
choice and engine conditions, and their specific formation
mechanisms must be understood and included in engine
models to reliably assess and predict the pollutant potential.
Specifically, the identity and provenance of fuel- and condition-specific unburnt exhaust species will need to be addressed, particularly for high exhaust gas recirculation rates.
Looking deeply into the combustion process, interesting
chemistry is revealed that is of fundamental value as well as of
practical relevance. Some quite similar reactions to those in
combustion are being considered under different conditions,
for example, to understand atmospheric or interstellar chemistry and combustion-assisted material synthesis including
carbon nanotubes and functional oxides. Molecules that are
presently considered only as fuels in combustion processes
can be important building blocks in energy conversion and
storage, and their reactions under different boundary conditions will matter. Alternative fuels produced in future from different feedstocks in a more sustainable and carbon-reduced
fashion may or may not resemble those in structure that we
are using today. To explore the underlying chemistry that is
useful for these and other purposes, high emphasis must be
placed on the quality of experimental data. For combustion diagnostics and beyond, desired features of diagnostics methods
include in-situ capability, spatial and temporal resolution, structure selectivity, high sensitivity, and transferability of the results
from laboratory environments to near-production engines.
Combinative diagnostics using a suite of appropriate instruments and techniques to address these needs in fundamental
and applied studies are advisable to ensure completeness and
Chem. Eur. J. 2016, 22, 13390 13401

www.chemeurj.org

reliability. Beyond the laboratory, however, real-time sensors


for portable on-board diagnostics to assess the combustion
performance with a chance for in-situ control are highly desirable next developments, needing advances in techniques, instruments, theory, mechanism reduction, and modeling speed,
and to be useful, they should become available, of course, at
competitive costs.

Acknowledgements
The author would like to acknowledge the fruitful collaboration on low-temperature combustion chemistry with Prof.
Heinz Pitsch, RWTH Aachen University; helpful discussions and
careful reading of this manuscript are gratefully acknowledged.
Thanks go also to the PC1 group, especially to Julia Pieper and
Christian Hemken. The partial support by the Deutsche Forschungsgemeinschaft under contract KO1363/31-1 is gratefully
acknowledged.
Keywords: analytical methods gas-phase reactions hightemperature chemistry
[1] Key World Energy Statistics 2013, http://www.iea.org/publications/freepublications/publication/KeyWorld2013.pdf.
[2] R.-J. Huang, Y. Zhang, C. Bozzetti, K.-F. Ho, J.-J. Cao, Y. Han, K. R. Daellenbach, J. G. Slowik, S. M. Platt, F. Canonaco, P. Zotter, R. Wolf, S. M.
Pieber, E. A. Bruns, M. Crippa, G. Ciarelli, A. Piazzalunga, M. Schwikowski,
G. Abbaszade, J. Schnelle-Kreis, R. Zimmermann, Z. An, S. Szidat, U. Baltensperger, I. El Haddad, A. S. H. Pr8vit, Nature 2014, 514, 218 222.
[3] T. D. Gordon, A. A. Presto, N. T. Nguyen, W. H. Robertson, K. Na, K. N.
Sahay, M. Zhang, C. Maddox, P. Rieger, S. Chattopadhyay, H. Maldonado,
M. M. Maricq, A. L. Robinson, Atmos. Chem. Phys. 2014, 14, 4643 4659.
[4] T. J. Jacobs, D. N. Assanis, Proc. Combust. Inst. 2007, 31, 2913 2920.
[5] M. P. B. Musculus, P. C. Miles, L. M. Pickett, Prog. Energy Combust. Sci.
2013, 39, 246 283.
[6] X. Lu, D. Han, Z. Huang, Prog. Energy Combust. Sci. 2011, 37, 741 783.
[7] J. M. Bergthorson, M. J. Thomson, Ren. Sust. Energy Rev. 2015, 42, 1393
1417.
[8] E. G. Giakoumis, C. D. Rakopoulos, A. M. Dimaratos, D. C. Rakopoulos,
Prog. Energy Combust. Sci. 2012, 38, 691 715.
[9] K. Kohse-Hinghaus, P. Owald, T. A. Cool, T. Kasper, N. Hansen, F. Qi,
C. K. Westbrook, P. R. Westmoreland, Angew. Chem. Int. Ed. 2010, 49,
3572 3597; Angew. Chem. 2010, 122, 3652 3679.
[10] J. Y. W. Lai, K. C. Lin, A. Violi, Prog. Energy Combust. Sci. 2011, 37, 1 14.
[11] S. H. Park, C. S. Lee, Energy Convers. Manage. 2014, 86, 848 863.
[12] C. K. Westbrook, C. V. Naik, O. Herbinet, W. J. Pitz, M. Mehl, S. M. Sarathy,
H. J. Curran, Combust. Flame 2011, 158, 742 755.
[13] L. Coniglio, H. Bennadji, P. A. Glaude, O. Herbinet, F. Billaud, Prog. Energy
Combust. Sci. 2013, 39, 340 382.
[14] Z. Wang, L. Zhang, K. Moshammer, D. M. Popolan-Vaida, V. S. B. Shankar,
A. Lucassen, C. Hemken, C. A. Taatjes, S. R. Leone, K. Kohse-Hinghaus,
N. Hansenc, P. Dagauth, S. Mani Sarathy, Combust. Flame 2016, 164,
386 396.
[15] K. Moshammer, A. W. Jasper, D. M. Popolan-Vaida, A. Lucassen, P. Di8vart, H. Selim, A. J. Eskola, C. A. Taatjes, S. R. Leone, S. M. Sarathy, Y. Ju,
P. Dagaut, K. Kohse-Hinghaus, N. Hansen, J. Phys. Chem. A 2015, 119,
7361 7374.
[16] a) Applied Combustion Diagnostics (Eds.: K. Kohse-Hinghaus, J. B. Jeffries), Taylor & Francis, New York, 2002; b) M. A. Linne, Spectroscopic
Measurement, Academic Press, London, 2002; c) A. C. Eckbreth, Laser Diagnostics for Combustion Temperature and Species, Vol. 3, 2nd. Ed.,
Gordon and Breach, The Netherlands, 1995.
[17] K. Kohse-Hinghaus, R. S. Barlow, M. Ald8n, J. Wolfrum, Proc. Combust.
Inst. 2005, 30, 89 123.

13399

T 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Concept
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]

[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]

[40]

[41]

[42]

[43]
[44]

[45]
[46]
[47]
[48]

[49]
[50]

[51]

V. Sick, Proc. Combust. Inst. 2013, 34, 3509 3530.


M. Ald8n, J. Bood, Z. Li, M. Richter, Proc. Combust. Inst. 2011, 33, 69 97.
R. K. Hanson, Proc. Combust. Inst. 2011, 33, 1 40.
S. Roy, J. R. Gord, A. K. Patnaik, Prog. Energy Combust. Sci. 2010, 36,
280 306.
A. Dreizler, B. Bhm, Proc. Combust. Inst. 2015, 35, 37 64.
P. Desgroux, X. Mercier, K. A. Thomson, Proc. Combust. Inst. 2013, 34,
1713 1738.
A. Bohlin, M. Mann, B. D. Patterson, A. Dreizler, C. J. Kliewer, Proc. Combust. Inst. 2015, 35, 3723 3730.
B. Peterson, D. L. Reuss, V. Sick, Proc. Combust. Inst. 2011, 33, 3089
3096.
O. Witzel, A. Klein, S. Wagner, C. Meffert, C. Schulz, V. Ebert, Appl. Phys.
B 2012, 109, 521 532.
a) M. Sthr, C. M. Arndt, W. Meier, Proc. Combust. Inst. 2015, 35, 3327
3335; b) A. M. Steinberg, I. Boxx, M. Sthr, C. D. Carter, W. Meier, Combust. Flame 2010, 157, 2250 2266; c) I. Boxx, M. Sthr, C. Carter, W.
Meier, Combust. Flame 2010, 157, 1510 1525.
J. Kariuki, A. Dowlut, R. Yuan, R. Balachandran, E. Mastorakos, Proc. Combust. Inst. 2015, 35, 1443 1450.
E. Ranzi, A. Frassoldati, A. Stagni, M. Pelucchi, A. Cuoci, T. Faravelli, Int. J.
Chem. Kinet. 2014, 46, 512 542.
S. M. Sarathy, P. Owald, N. Hansen, K. Kohse-Hinghaus, Prog. Energy
Combust. Sci. 2014, 44, 40 102.
Y. Zhu, S. Li, D. F. Davidson, R. K. Hanson, Proc. Combust. Inst. 2015, 35,
241 248.
C. Ji, S. M. Sarathy, P. S. Veloo, C. K. Westbrook, F. N. Egolfopoulos, Combust. Flame 2012, 159, 1426 1436.
P. Dagaut, S. M. Sarathy, M. J. Thomson, Proc. Combust. Inst. 2009, 32,
229 237.
M. H. Hakka, P.-A. Glaude, O. Herbinet, F. Battin-Leclerc, Combust. Flame
2009, 156, 2129 2144.
N. Hansen, T. A. Cool, P. R. Westmoreland, K. Kohse-Hinghaus, Prog.
Energy Combust. Sci. 2009, 35, 168 191.
Y. Li, F. Qi, Acc. Chem. Res. 2010, 43, 68 78.
L. V. Moskaleva, M. C. Lin, Proc. Combust. Inst. 2000, 28, 2393 2401.
N. Lamoureux, X. Mercier, C. Western, J. F. Pauwels, P. Desgroux, Proc.
Combust. Inst. 2009, 32, 937 944.
C. A. Taatjes, N. Hansen, A. McIlroy, J. A. Miller, J. P. Senosiain, S. J. Klippenstein, F. Qi, L. Sheng, Y. Zhang, T. A. Cool, J. Wang, P. R. Westmoreland, M. E. Law, T. Kasper, K. Kohse-Hinghaus, Science 2005, 308, 1887
1889.
F. Battin-Leclerc, O. Herbinet, P.-A. Glaude, R. Fournet, Z. Zhou, L. Deng,
H. Guo, M. Xie, F. Qi, Angew. Chem. Int. Ed. 2010, 49, 3169 3172;
Angew. Chem. 2010, 122, 3237 3240.
O. Welz, J. Z#dor, J. D. Savee, M. Y. Ng, G. Meloni, R. X. Fernandes, L.
Sheps, B. A. Simmons, T. S. Lee, D. L. Osborn, C. A. Taatjes, Phys. Chem.
Chem. Phys. 2012, 14, 3112 3127.
Combustion Generated Fine Carbonaceous Particles (Eds.: H. Bockhorn, A.
DAnna, A. F. Sarofim, H. Wang), KIT Scientific Publishing, Karlsruhe,
2009.
H. Wang, Proc. Combust. Inst. 2011, 33, 41 67.
T. A. Cool, K. Nakajima, T. A. Mostefaoui, F. Qi, A. McIlroy, P. R. Westmoreland, M. E. Law, L. Poisson, D. S. Peterka, M. Ahmed, J. Chem. Phys.
2003, 119, 8356 8365.
C. A. Taatjes, N. Hansen, D. L. Osborn, K. Kohse-Hinghaus, T. A. Cool,
P. R. Westmoreland, Phys. Chem. Chem. Phys. 2008, 10, 20 34.
F. N. Egolfopoulos, N. Hansen, Y. Ju, K. Kohse-Hinghaus, C. K. Law, F. Qi,
Prog. Energy Combust. Sci. 2014, 43, 36 67.
M. B. Sajid, E. Es-sebbar, T. Javed, C. Fittschen, A. Farooq, Int. J. Chem.
Kinet. 2014, 46, 275 284.
C. Bahrini, P. Morajkar, C. Schoemaecker, O. Frottier, O. Herbinet, P.-A.
Glaude, F. Battin-Leclerc, C. Fittschen, Phys. Chem. Chem. Phys. 2013, 15,
19686 19698.
P. Owald, P. Hemberger, T. Bierkandt, E. Akyildiz, M. Khler, A. Bodi, T.
Gerber, T. Kasper, Rev. Sci. Instrum. 2014, 85, 025101.
J. Kreger, G. A. Garcia, D. Felsmann, K. Moshammer, A. Lackner, A. Brockhinke, L. Nahon, K. Kohse-Hinghaus, Phys. Chem. Chem. Phys. 2014, 16,
22791 22804.
M. Djokic, H.-H. Carstensen, K. M. Van Geem, G. B. Marin, Proc. Combust.
Inst. 2013, 34, 251 258.

Chem. Eur. J. 2016, 22, 13390 13401

www.chemeurj.org

[52] M. R. Harper, K. M. Van Geem, S. P. Pyl, G. B. Marin, W. H. Green, Combust.


Flame 2011, 158, 16 41.
[53] L. Hanley, R. Zimmermann, Anal. Chem. 2009, 81, 4174 4182.
[54] K. M. Van Geem, S. P. Pyl, M.-F. Reyniers, J. Vercammen, J. Beens, G. B.
Marin, J. Chromatogr. A 2010, 1217, 6623 6633.
[55] L. Vogt, T. Grger, R. Zimmermann, J. Chromatogr. A 2007, 1150, 2 12.
[56] C. W. Gao, J. W. Allen, W. H. Green, R. H. West, Comp. Phys. Commun.
2016, 203, 212 225.
[57] J. D. Savee, S. Borkar, O. Welz, B. Szt#ray, C. A. Taatjes, D. L. Osborn, J.
Phys. Chem. A 2015, 119, 7388 7403.
[58] C. A. Taatjes, O. Welz, A. J. Eskola, J. D. Savee, A. M. Scheer, D. E. Shallcross, B. Rotavera, E. P. F. Lee, J. M. Dyke, D. K. W. Mok, D. L. Osborn, C. J.
Percival, Science 2013, 340, 177 180.
[59] A. Jalan, I. M. Alecu, R. Meana-Paeda, J. Aguilera-Iparraguirre, K. R.
Yang, S. S. Merchant, D. G. Truhlar, W. H. Green, J. Am. Chem. Soc. 2013,
135, 11100 11114.
[60] A. W. Jasper, K. M. Pelzer, J. A. Miller, E. Kamarchik, L. B. Harding, S. J.
Klippenstein, Science 2014, 346, 1212 1215.
[61] L. Seidel, K. Moshammer, X. Wang, T. Zeuch, K. Kohse-Hinghaus, F.
Mauss, Combust. Flame 2015, 162, 2045 2058.
[62] a) K. Fieweger, R. Blumenthal, G. Adomeit, Combust. Flame 1997, 109,
559 619; b) H. K. Ciezki, G. Adomeit, Combust. Flame 1993, 93, 421
433; c) K. A. Heufer, H. Olivier, Shock Waves 2010, 20, 307 316.
[63] S. S. Merchant, C. F. Goldsmith, A. G. Vandeputte, M. P. Burke, S. J. Klippenstein, W. H. Green, Combust. Flame 2015, 162, 3658 3673.
[64] J. Z#dor, C. A. Taatjes, R. X. Fernandes, Prog. Energy Combust. Sci. 2011,
37, 371 421.
[65] M. Ehn, J. A. Thornton, E. Kleist, M. Sipil-, H. Junninen, I. Pullinen, M.
Springer, F. Rubach, R. Tillmann, B. Lee, F. Lopez-Hilfiker, S. Andres, I.-H.
Acir, M. Rissanen, T. Jokinen, S. Schobesberger, J. Kangasluoma, J. Kontkanen, T. Nieminen, T. Kurt8n, L. B. Nielsen, S. Jrgensen, H. G. Kjaergaard, M. Canagaratna, M. Dal Maso, Nature 2014, 506, 476 479.
[66] a) V. Franco, M. Kousoulidou, M. Muntean, L. Ntziachristos, S. Hausberger, P. Dilara, Atmos. Environ. 2013, 70, 84 97; b) K. Skalska, J. S. Miller, S.
Ledakowicz, Sci. Total Environ. 2010, 408, 3976 3989.
[67] a) S. M. Palash, H. H. Masjuki, M. A. Kalam, B. M. Masum, A. Sanjid, M. J.
Abedin, Energy Convers. Manage. 2013, 76, 400 420; b) J. Xue, T. E.
Grift, A. C. Hansen, Ren. Sust. Energy Rev. 2011, 15, 1098 1116.
[68] J. Sun, J. A. Caton, T. J. Jacobs, Prog. Energy Combust. Sci. 2010, 36, 677
695.
[69] S. K. Hoekman, C. Robbins, Fuel Process. Technol. 2012, 96, 237 249.
[70] B. M. Masum, H. H. Masjuki, M. A. Kalam, I. M. Rizwanul Fattah, S. M.
Palash, M. J. Abedin, Ren. Sust. Energy Rev. 2013, 24, 209 222.
[71] a) C. Schulz, V. Sick, U. E. Meier, J. Heinze, W. Stricker, Appl. Opt. 1999,
38, 1434 1443; b) W. G. Bessler, C. Schulz, T. Lee, D.-I. Shin, M. Hofmann,
J. B. Jeffries, J. Wolfrum, R. K. Hansen, Appl. Phys. B 2002, 75, 97 102.
[72] F. Hildenbrand, C. Schulz, J. Wolfrum, F. Keller, E. Wagner, Proc. Combust.
Inst. 2000, 28, 1137 1143.
[73] R. Stevens, P. Ewart, H. Ma, C. R. Stone, Combust. Flame 2007, 148, 223
233.
[74] a) G. C. Martin, C. J. Mueller, C.-F. F. Lee, Appl. Opt. 2006, 45, 2089 2100;
b) K. Verbiezen, R. J. H. Klein-Douwel, A. P. van Vliet, A. J. Donkerbroek,
W. L. Meerts, N. J. Dam, J. J. ter Meulen, Proc. Combust. Inst. 2007, 31,
765 773; c) K. Verbiezen, A. J. Donkerbroek, R. J. H. Klein-Douwel, A. P.
van Vliet, P. J. M. Frijters, X. L. J. Seykens, R. S. G. Baert, W. L. Meerts, N. J.
Dam, J. J. ter Meulen, Combust. Flame 2007, 151, 333 346.
[75] W. G. Bessler, M. Hofmann, F. Zimmermann, G. Suck, J. Jakobs, S. Nicklitzsch, T. Lee, J. Wolfrum, C. Schulz, Proc. Combust. Inst. 2005, 30, 2667
2674.
[76] K. Yehliu, R. L. Van der Wal, O. Armas, A. L. Boehman, Combust. Flame
2012, 159, 3597 3606.
[77] a) C. K. Gaddam, R. L. Vander Wal, Combust. Flame 2013, 160, 2517
2528; b) R. L. Vander Wal, V. M. Bryg, M. D. Hays, Aerosol Sci. Technol.
2010, 41, 108 117.
[78] J. Braga Dallarosa, J. Garcia Minego, E. Calesso Teixeira, J. Luis Stefens,
F. Wiegand, Atmos. Environ. 2005, 39, 1609 1625.
[79] A. Raj, P. L. W. Man, T. S. Totton, M. Sander, R. A. Shirley, M. Kraft, Carbon
2010, 48, 319 332.
[80] T. S. Totton, S. Chakrabarti, A. J. Misquitta, M. Sander, D. J. Wales, M.
Kraft, Combust. Flame 2010, 157, 909 914.

13400

T 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Concept
[81] a) N. A. Slavinskaya, U. Riedel, S. B. Dworkin, M. J. Thomson, Combust.
Flame 2012, 159, 979 995; b) S. B. Dworkin, Q. Zhang, M. J. Thomson,
N. A. Slavinskaya, U. Riedel, Combust. Flame 2011, 158, 1682 1695.
[82] B. Giechaskiel, M. Maricq, L. Ntziachristos, C. Dardiotis, X. Wang, H.
Axmann, A. Bergmann, W. Schindler, J. Aerosol Sci. 2014, 67, 48 86.
[83] a) S. A. Skeen, H. A. Michelsen, K. R. Wilson, D. M. Popolan, A. Violi, N.
Hansen, J. Aerosol Sci. 2013, 58, 86 102; b) B. Apicella, A. Carpentieri,
M. AlfH, R. Barbella, A. Tregrossi, P. Pucci, A. Ciajolo, Proc. Combust. Inst.
2007, 31, 547 553.
[84] T. Mouton, X. Mercier, P. Desgroux, Appl. Phys. B 2016, 122:123.
[85] a) H. Bladh, N.-E. Olofsson, T. Mouton, J. Simonsson, X. Mercier, A. Faccinetto, P.-E. Bengtsson, P. Desgroux, Proc. Combust. Inst. 2015, 35, 1843

Chem. Eur. J. 2016, 22, 13390 13401

www.chemeurj.org

1850; b) H. A. Michelsen, C. Schulz, G. J. Smallwood, S. Will, Prog. Energy


Combust. Sci. 2015, 51, 2 48.
[86] a) B. Apicella, P. Pr8, M. AlfH, A. Ciajolo, V. Gargiulo, C. Russo, A. Tregrossi, D. Deldique, J. N. Rouzaud, Proc. Combust. Inst. 2015, 35, 1895 1902;
b) K. Yehliu, R. L. Van der Wal, A. L. Boehman, Combust. Flame 2011, 158,
1837 1851.
[87] M. Schenk, S. Lieb, H. Vieker, A. Beyer, A. Glzh-user, H. Wang, K. KohseHinghaus, ChemPhysChem 2013, 14, 3248 3254.

Received: June 6, 2016


Published online on July 21, 2016

13401

T 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Potrebbero piacerti anche