Vous êtes sur la page 1sur 10

Combustion and Flame 207 (2019) 186–195

Contents lists available at ScienceDirect

Combustion and Flame


journal homepage: www.elsevier.com/locate/combustflame

The internal combustion engine as a CO2 reformer


Hendrik Gossler a, Simon Drost b, Sylvia Porras b, Robert Schießl b, Ulrich Maas b,
Olaf Deutschmann a,∗
a
Institute for Chemical Technology and Polymer Chemistry, Karlsruhe Institute of Technology, Karlsruhe, Germany
b
Institute for Technical Thermodynamics, Karlsruhe Institute of Technology, Karlsruhe, Germany

a r t i c l e i n f o a b s t r a c t

Article history: This paper investigates the concept of using piston engines as chemical reactors to convert CO2 to more
Received 17 January 2019 useful substances. Conditions are presented under which significant conversions of CO2 and CH4 to syn-
Revised 6 March 2019
thesis gas (CO and H2 ) can be achieved in internal combustion engines (ICE). Numerical optimization of
Accepted 20 May 2019
the process based on detailed chemistry is employed to identify initial values of temperature, pressure
and gas composition which theoretically yield maximal CO2 conversions. The optimization results are
Keywords: used to guide experiments in a rapid compression machine (RCM) setup that is used in place of a real
CO2 reforming ICE. The optimization predicts that the addition of molecular oxygen enables the endothermal dry re-
Internal combustion engine forming reaction (CO2 + CH4  2 H2 + 2 CO) to proceed by delivering the required energy from burning
Rapid compression machine
a part of the fuel. Despite combustion taking place, experiments show that over 50% of the CO2 con-
Numerical optimization
tained in mixtures of Ar, O2 , CH4 , DME (dimethylether) and CO2 can be converted to syngas with H2 O as
Detailed chemical kinetics
a by-product.
© 2019 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

1. Introduction reforming, these systems have other disadvantages [3–5]. For ex-
ample, nickel catalysts are susceptible to coking, while cobalt cat-
The environmental effects of CO2 emissions are a major prob- alysts are reported to suffer from deactivation by oxidation [6,7].
lem of industrialized societies. Beside its appearance as a waste- An exploration of alternative methods for CO2 reforming there-
product of hydrocarbon combustion [1], CO2 also occurs as one fore appears worthwhile. The physico-chemical properties of the
of the main byproducts in chemical industry. CO2 reforming, or overall reaction (Eq. 1) impose important constraints for such a
dry reforming, has gained increased attention in recent years as it method:
opens a route to utilize the CO2 as a carbon source for more use-
ful products [2], and therefore represents a possibility to reduce
1. Hypothetically, if all kinetic constraints are removed, the system
CO2 emissions. Dry reforming is strongly endothermic, and the net
will instantaneously approach chemical equilibrium. The chem-
reaction can be formulated as
ical composition at this equilibrium, and therefore also the ex-
CO2 + CH4  2 H2 + 2 CO H  = 247 kJ mol−1 . (1) tent of conversion, depends on the thermodynamic conditions.
For the endothermal reaction (Eq. (1)), chemical equilibrium
Besides offering the benefit of removing undesired CO2 , this re-
moves further to the right (i.e. to the side of CO and H2 )
action is also attractive for applications that require synthesis gas
with increasing temperature (at constant pressure). At constant
with low H2 /CO ratios.
temperature, equilibrium is shifted to the right when pressure
In the chemical industry, reforming processes are generally car-
is decreased. Therefore, based on equilibrium considerations,
ried out using catalysts that often contain noble metals such as
high-temperature and low-pressure conditions favor high CO2
rhodium and palladium. However, as in the case of dry reform-
conversions. Starting from unreacted “cold” low-pressure con-
ing, the use of precious metal catalysts is often cost prohibitive at
ditions, input of energy is hence required to shift equilibrium
commercial-scale. Although current research suggests that less ex-
to the desired state.
pensive catalysts based on nickel or cobalt could be used in dry
2. Some of the elementary reaction steps underlying Eq. (1) have
very high activation energies. Due to the Arrhenius-like de-

Corresponding author. pendence of elementary reaction rate coefficients, high tem-
E-mail address: deutschmann@kit.edu (O. Deutschmann). peratures are required to allow those reactions to proceed at
URL: http://www.itcp.kit.edu/deutschmann (O. Deutschmann) sufficient speeds and approach equilibrium.

https://doi.org/10.1016/j.combustflame.2019.05.031
0010-2180/© 2019 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
H. Gossler, S. Drost and S. Porras et al. / Combustion and Flame 207 (2019) 186–195 187

Compression engines have the potential to fulfill the above- outline a concept in which engines form the key component of
mentioned requirements, and also feature some more properties compact gas-to-liquids plants. The concept allows the plant to be
which make them appropriate candidates as reactors: operated at a significantly smaller scale in a profitable way, due
to the lower costs associated with engines compared to conven-
• The required input of energy (item (1) above) can be provided
tional chemical reactors. Furthermore, a unique feature of their ex-
in the form of mechanical compression work. This work can
perimental work was the ability to preheat the inlet gas to over
be supplied within relatively short times (down to the scale
480° C.
of milliseconds). Within these time scales, heat loss processes
There are also numerous publications that studied the in-
can interfere strongly with chemical kinetics. The ability for fast
cylinder production of syngas via partial oxidation numeri-
repetition also means that high mass-flows of CO2 can be con-
cally, for example [16–18]. In their modeling study, Hegner and
verted.
Atakan [18] focused on thermodynamic aspects of a polygenera-
• The high temperatures required to overcome kinetic barriers
tion process to produce hydrogen. The authors claim that exergetic
(point (2) above) can be provided in engines via compression.
efficiencies of up to 80% are achievable compared to the separated
• In contrast to conventional large-scale chemical reactors, com-
production of power, heat and hydrogen. Their study was later ex-
pression engines can be operated in a very flexible manner, tol-
tended by accompanying experiments in a rapid compression ma-
erating quite large variations of inputs (mass flow rates, input
chine [14]. Here, dimethyl ether (DME) was used as a fuel addi-
pressure and temperature).
tive to allow methane conversions at moderate, technically acces-
• Engines are also able to react promptly to changes in input con-
sible compression ratios and inlet temperatures, as was done in
ditions, and their operation can be adapted to optimal opera-
the present work. Apart from syngas, small amounts (2 mol %) of
tion.
ethylene were produced in the experiments.
Also, from an economic point of view, engines have the advan- In another modeling study, numerical optimization was applied
tage of being cheap as the technology is well developed and ro- to maximize the hydrogen yield at the end of the expansion stroke
bust. of an HCCI engine [17]. In the present study we expand on the ap-
Note that, the higher the conversion, i.e., the more the reaction proach in [17]. In contrast to previous studies that investigated the
(Eq. (1)) is moved to its right side, the more net input of energy use of engines as reformers to produce syngas, this work focuses
is required for the process. In energy conversion technology, situa- on converting an undesirable species (CO2 ). The production of syn-
tions where a temporal and/or local surplus of an otherwise “unus- gas, which goes along with the CO2 conversion, is regarded as a
able” form of energy (e.g., heat or work) is available for exactly this side effect (albeit a positive one) only.
purpose, occur quite frequently. The availability of energy from “re- The objective of this paper is to demonstrate the concept of us-
newable” energy sources such as solar and wind power are subject ing a piston engine process to convert CO2 to syngas by dry re-
to high fluctuations. This leads to gaps between energy demand forming. We employ a coupled simulation/experimental approach
and availability. When there is an abundance of (electric) energy, in which simulations, together with numerical optimization, are
the energy should ideally be stored in order to compensate times used to identify operation conditions that theoretically yield opti-
where the renewable energy sources are not able to meet the de- mal conversions. A rapid compression machine (RCM) is used as an
mand of energy. engine simulator to assess whether the predicted conversion levels
While the literature contains enormous amounts of studies con- can be realized in a process of semi-technical scale. Rapid com-
cerning ICEs operated under conventional (i.e. total-oxidation) con- pression machines are devices which are frequently employed to
ditions, substantially less research has been dedicated to their use investigate chemical reactions under engine-like conditions [19].
as chemical reactors. Recently, a review article has been published In the remainder of this paper we will first describe the mod-
that gives an overview of the past research on fuel reforming in elling approach, which involves simulation of an engine process
conjunction with internal combustion engines [8]. Most of the pre- with detailed chemistry. The procedure for using the simulation
vious work on in-cylinder fuel reforming has focused on burning to identify optimal engine operation points is then outlined. Next,
rich fuel-air mixtures to produce synthesis gas via partial oxida- the RCM experiment setup is described. The results presented af-
tion following the reaction terwards show that CO2 conversion levels of over 50 percent are
1 possible, both in theory and in the experiments, with suitably cho-
CH4 + O2  2 H2 + CO H  = −35.7 kJ mol−1 . (2) sen initial composition of the mixture and initial temperatures and
2
pressures.
The in-cylinder production of syngas has been shown to work
in engine experiments by numerous groups, e.g. [9–15]. In these
studies, various ways to achieve stable operating conditions (i.e. 2. Methodology
low cycle-to-cycle variations) were investigated. Spark ignition (SI),
compression ignition (CI) and homogeneous charge compression In an engine, there are many operating parameters that can be
ignition (HCCI) were successfully applied to convert methane- varied independently. Examples include the engine speed (rpm),
based fuels to syngas. the inlet gas temperature (T0 ), the inlet pressure (p0 ) and the over-
Karim and Wierzba [11] operated an engine on rich mix- all gas composition (mole fractions, x). Finding operating points
tures of methane and oxygenated air, where the mixture was ig- that will lead to high CO2 conversions is difficult due to the
nited by injecting diesel fuel (CI mode). Yang et al. [12] were vast number of combinations of possible operating parameters
able to demonstrate syngas production in an HCCI process with and the typically strongly nonlinear dependence of the chemical
added CO. In a combined modeling and experimental study, Wie- conversion on these parameters. For this reason, conducting em-
mann et al. [15] investigated a single-cylinder engine operated pirical experiments on a trial-and-error basis is not a practical
on methane/air using spark ignition and homogeneous charge approach.
compression-ignition. In their experiments, n-heptane was added Instead, the approach taken in this paper is to employ numer-
to reduce auto-ignition temperatures and to stabilize engine oper- ical optimization in conjunction with detailed chemical kinetics
ation. simulations to predict promising operating points. The optimal op-
In a study by a group of researchers at MIT [13], the use of erating points proposed by the model are then verified in a rapid
an engine as a reformer is motivated by economical factors. They compression machine setup.
188 H. Gossler, S. Drost and S. Porras et al. / Combustion and Flame 207 (2019) 186–195

The dry reforming reaction (Eq. (1)) has a few characteristics


that are important to the scope of the optimization approach. Pre-
liminary simulation studies showed that temperatures exceeding
10 0 0 K are required for any reactions to proceed at sufficiently high
rates. Thus, we can identify the temperature as a key limiting fac-
tor for CO2 conversion. We can assume that the higher the peak
temperature during an engine revolution is, the higher the CO2
conversion becomes. However, it is difficult to attain these high
temperatures in typical experimental setups due to their limited
preheating capabilities and/or insufficiently high compression ra-
tios. Because the initial gas composition itself can affect the re-
sulting temperature profile in a number of ways, the composition
gains an essential role as an engine tuning parameter. The objec-
tive is therefore to find optimal initial gas compositions that im-
plicitly lead to maximum temperatures, that in turn cause a maxi-
mum amount of CO2 to be converted. However, it is not clear what Fig. 1. Schematic drawing of the RCM used in this work. When the bumper is in
the exact gas composition should be because there are opposing place as shown, the piston remains at TDC after compression (“RCM mode”). If the
effects and constraints. bumper is removed altogether, the piston immediately retracts after reaching TDC
Argon can be added to reduce the heat capacity of the gas mix- (“RCEM mode”, with zero hold time, tH = 0).

ture and hence increase the temperature during the compression


phase. But more argon reduces the amount of the reactants (CO2
The equation system is closed by using the ideal gas equation
and CH4 ) that can be added to the mixture. Another way to reach
higher temperatures is to burn a part of the CH4 by adding O2 . pV = nRT . (6)
However, the mixture may fail to ignite if there is not enough O2
present; too much O2 , on the other hand, will lead to the produc- The first differential equation (Eq. (3)) describes the temporal
tion of CO2 . Therefore, under the conditions we study, a trade-off change of moles of each species, and the second (Eq. (4)) describes
has to be taken: No oxygen addition offers potentially high CO2 the change in temperature. In these equations, t is the time, n is
conversion, but the inferred low reactivity of the oxygen-free mix- the amount of substance, and ω˙ is the effective reaction rate (in
ture may prevent the engine from realizing this potentially high mol/m3 /s) for the i-th species. Furthermore, H denotes the molar
conversion. Higher amounts of oxygen tend to increase the mix- enthalpy, R the universal gas constant, T the temperature, p the
ture’s reactivity, but also favor formation of CO2 , which is coun- pressure, and cV the mixture’s heat capacity (in J/K) at constant
terproductive in a CO2 reforming process. Furthermore, reactivity volume. The temporal rates of change of moles and volume are de-
could be increased by adding reaction enhancers to the mixture, noted by n˙ and V˙ , respectively. Lastly, q˙ represents the heat flux (in
such as DME. Adding these enhancers can increase CO2 conversion; W) from the combustion chamber through the cylinder walls.
on the other hand, overly large amounts of enhancer increase the The general formula for calculating the effective reaction rates
cost and complexity of the process. when using a detailed reaction scheme is given by
These considerations show that the problem of finding condi-
 

NR
 
NS
νs j
tions for which high CO2 conversion can be achieved is a non- ω˙ i = f (c, T ) = kj νij − νi j
 cs , (7)
trivial task. Therefore, an approach based on mathematical opti- j s
mization is advantageous.
which is a function of both species concentrations c and tempera-
2.1. Modeling approach ture. Here, ν  denotes the stoichiometric coefficients of the prod-
ucts, while ν  denotes those of the reactants. The total number of
2.1.1. Detailed chemistry simulations reactions and species are given by NR and NS , respectively. The rate
The combustion chamber of the RCM is modeled as a batch coefficient k is described either by an extended Arrhenius law, or,
reactor with a user-definable variable volume profile. The model in case the reaction is pressure-dependent, by the “F-Center” for-
accounts for the temporal change of volume, temperature and mulation by Troe et al. [21]. The unit of k depends on the reaction
species concentrations using a detailed gas-phase mechanism. Sur- order.
face reactions and spatial variations within the chamber are not The heat flux q˙ through the chamber walls is described by the
considered in this model. Heat loss through the cylinder walls is basic relationship for convective heat transfer

accounted for using the global heat transfer model according to q˙ (t ) = h(t ) · Apiston + Ahead + Aliner (t ) T (t ) , (8)
Woschni [20].
The modeling approach gives rise to the following system of where h is the heat transfer coefficient that is obtained by the
coupled differential equations that must be solved simultane- Woschni relation (in W/m2 /K, see below). Furthermore, A is the
ously: area of the particular engine part as denoted by the index. The
area of the liner is the cylindrical part of the combustion cham-
dni
= V · ω˙ i , (3) ber, and is therefore dependent on time (cf. Fig. 1). The tempera-
dt ture difference in the above equation applies to the gas and wall
   temperatures, i.e.
dT n˙ i H i (T ) − n˙ i RT − pV˙ − q˙
=
i
. (4) T (t ) = Tgas (t ) − Twall . (9)
dt cV
The temperature Twall was taken to be equal to the inlet tempera-
The volume V enters the simulation as a temporal constraint; it
ture of the gas (462 K) for all simulations; in practice, typical wall
is computed from tabulated time/volume pairs that are obtained
temperature variations in engine-like devices [22] are small rela-
from the experimental setup:
tive to the gas temperature Tgas , so that they do not impede the
V = Vcyl (t ) . (5) realism of the model.
H. Gossler, S. Drost and S. Porras et al. / Combustion and Flame 207 (2019) 186–195 189

Table 1 2.1.3. Chemical kinetic model


Nomenclature.
Development of chemical kinetic models has been a main sub-
Notation Unit ject of research in the last decades, mainly because of their major
α Heat transfer scaling factor – role in the simulation of internal combustion systems as well as in
ν Stochiometric coefficient – the description of species time-profiles in fuel combustion and ox-
φ Fuel equivalence ratio – idation [23]. Several reaction mechanisms have been developed in
ω˙ Effective reaction rate mol/m3 /s the past in order to describe hydrocarbon fuel combustion [24–32].
A Area m2
Using this well-validated and valuable information, Por-
B Bore m
c Species concentration mol/m3 ras et. al. [33] compiled a reaction mechanism based on the exist-
cV Mixture heat capacity at J/mol/K ing methane combustion and pyrolysis model of Hidaka et al. [30],
constant volume the C1 -C4 hydrocarbon oxidation mechanism of Heghes [31] devel-
C1 Constant (2.892) –
oped for the description of conventional combustion and the DME
C2 Constant (3.24×10−3 ) –
h Heat transfer coefficient W/m2 /K mechanism of Zhao et al. [32]. The resulting assembled mechanism
Hi Molar enthalpy of species i J/mol was adapted specifically for the description of polygeneration pro-
k Rate coefficient mol, m, s cesses of CH4 /DME mixtures at large equivalence ratios φ up to
n Amount of substance mol 20. It consists of 578 elementary steps and 90 species and was
NR Total number of reactions –
validated against RCM and shock tube ignition delay times mea-
NS Total number of species –
p Pressure Pa surements as well as against RCEM and flow reactor species pro-
rc Compression ratio – files experimental data between a temperature range of 600 K to
q˙ Heat flux W 1500 K and pressures of 10 bar to 30 bar [33,34].
R Universal gas constant J/mol/K
Detailed chemistry simulations presented in this work were
t Time s
T Temperature K
performed using this kinetic model. The mechanism will be pub-
U Internal energy J lished in a separate article, however, an electronic version of it is
Ui Molar internal energy of species i J/mol attached to this work as supplemental material.
V Volume m3
Vd Displacement volume m3
w Instantaneous gas velocity m/s
2.2. Optimization procedure
x Mole fraction –
X Conversion m/s The purpose of the optimization is to predict engine operating
conditions (the optimization variables) under which as much CO2
as possible (the objective function) is converted. There are many
The formula by Woschni provides a spatially-averaged instanta- operating parameters that can be varied, for example the engine
neous heat transfer coefficient and is given by speed (N, commonly in rpm, represented in the RCM by the com-
pression duration), the compression ratio (rc ), the inlet gas temper-
h(t ) = α · B−0.2 p(t )0.8 T (t )−0.53 w(t )0.8 . (10) ature (T0 ), the inlet pressure (p0 ) or the overall initial gas compo-
sition (mole fractions, x0 ). However, the problem can be simplified
Here, B is the bore (equal to the diameter of the cylindrical com- given the preliminary considerations outlined in the beginning of
bustion chamber), p is the instantaneous pressure (in bar) and T Section 2. The number of optimization variables can be reduced
the instantaneous temperature. A scaling factor (α ) exists to match because one of the key limiting factors is the maximum attainable
a particular experimental setup (see Section 2.1.2). The instanta- temperature. Hence, in the optimization, the initial gas tempera-
neous gas velocity w is given by ture is not included as an optimization variable and is set to al-
ways be T0 = 462 K. This value corresponds to the maximum tem-
Vd Tref
w(t ) = C1 c̄piston + C2 [ p(t ) − pmot (t )] , (11) perature admissible in our RCM setup. For better transferability of
prefVref
the results to the technical system, the initial pressure was taken
where C1 and C2 are constants (see Table 1), c̄piston is the mean to be 1.013 bar for all simulations and experiments. Furthermore,
piston speed, and Vd is the displaced volume. The variables Tref , the compression ratio is set to rc = 10, as this is representative for
pref and Vref denote “reference values” of temperature, pressure and a typical spark-ignition engine.
volume, for which the corresponding initial values are used. The In a real engine, the engine speed represents a suitable tun-
last term in the above equation in square brackets is a measure for ing parameter, as it influences ignition timing and residence time.
the combustion taking place and is equal to the difference between However, in an RCM, the engine speed (corresponding to compres-
the instantaneous pressures when the engine is fired and when it sion rate) is constrained to a small range, which considerably limits
is motored. In our simulations, the motoring pressure is obtained the usefulness of this parameter as an optimization variable. In all
by simulating the cycle while neglecting chemical reactions. For optimization simulations, we therefore use the same volume pro-
more details, the reader is referred to the original publication by file. The only remaining optimization variables are hence the mole
Woschni [20]. fractions x0 of the initial gas composition. In this study, the gas is
composed of the five species Ar, CH4 , DME, O2 and CO2 . The opti-
mization is carried out by keeping the initial CO2 amount constant
2.1.2. Scaling the heat transfer coefficient at a certain value and allowing the optimizer to vary the remain-
The scaling factor (α ) in the Woschni heat transfer correlation ing four species. This procedure is repeated for initial CO2 values
(Eq. 10) was varied in order to best match a pressure curve from from 10% to 50%.
the experimental setup prior to the optimization. The pressure pro- The objective function to be maximized is the CO2 conversion
file from the experiment was recorded using a gas feed of pure that we define as
nitrogen to rule out any misleading effects due to chemical reac- nCO2 (t0 ) − nCO2 (t ) nCO2 (t )
tion. Other experimental parameters were set to match those of XCO2 (t ) = =1− , (12)
nCO2 (t0 ) nCO2 (t0 )
the later optimization simulations: T0 = 462 K, p0 = 1.0 bar, rc =
10. The fit to the pressure profile resulted in a value of α = 99.6, where nCO2 is the time-dependent number of moles of CO2 . In the
which was then used in all subsequent simulations. optimization, the value is calculated for tf = 300 ms, which is the
190 H. Gossler, S. Drost and S. Porras et al. / Combustion and Flame 207 (2019) 186–195

compression time plus a certain hold time. We formulate the opti- Table 2
Parameters of the RCM setup.
mization problem as
Bore B 82 mm
nCO2 (tf ) Stroke L 74 mm
max XCO2 = 1 − , (13)
x0 nCO2 (t0 ) Compression ratio rc 8 to 12
Displaced volume Vd 0.4 L
subject to the linear constraints Clearance volume Vc 6.6 cm3
 Initial temperature T0 300 K to 462 K
x0,i = 1 , (14) Initial pressure p0 0.5 to 1.5 bar
i

The equilibria represent the states that would result in the RCM
9x0,DME − 1x0,CH4 = 0 , (15)
if the system would instantaneously react to the maximum possi-
ble extent, always adapting to transient thermodynamical condi-
tions (U(t) and V(t)), which constantly change during the process
x0,i ≥ 0 , (16) due to compression and heat losses.
The equilibrium for given values of Ufix and Vfix is given by the
state of maximum entropy at those conditions. While a direct so-
x0,Ar ≥ 0.02 . (17) lution of U, V equilibrium via entropy maximization is possible, it
is sometimes plagued by numerical problems, especially for rich
The linear constraints formulated above are to ensure that the
mixtures. On the other hand, for the related problem of finding the
sum of all mole fractions equals unity (Eq. (14)), that the fuel
equilibrium for given values of T, p (by minimizing Gibbs enthalpy),
is composed of 90% CH4 and 10% DME by volume (Eq. 15), and
a highly robust and efficient algorithm is known [39]. Therefore,
that each component can only take on positive values (Eq. (16)).
we used this as a numerical tool to obtain U, V-equilibria by an
Furthermore, Eq. (17) constrains the amount of argon to above
iterative procedure. The first step involves expressing the equilib-
2%. This last constraint is in place to reflect requirements of the
rium moles (ni ) for isothermal-isobaric conditions as a function of
accompanying experiments, where a certain amount of helium
T and p:
is required as a reference for the gas analysis instrument (see
Section 3 for more details). For simplicity, the optimization does ni = n˜ i,eq (T , p) (i = 1, . . . , NS ) (19)
not additionally take helium into account. Instead, a part of the ar-
To obtain the equilibrium for given U and V rather than for given T
gon is substituted by helium in the experiments.
and p, the nonlinear system of two simultaneous equations in the
The optimization problem is solved using the derivative-based
two unknowns T, p
SQP algorithm from the MATLAB Optimization Toolbox (see Chap-
ter 18 in [35] for a description of the algorithm). The actual simula- 
NS

tion code is part of the DETCHEM package [36] and is written in C, U i (T ) · n˜ i,eq (T , p) − Ufix = 0 , (20)
C++ and Fortran. The code computes the objective function value i=1

along with the derivatives with respect to the optimization vari-


p
ables. MATLAB calls the simulation code via a dynamic library that NS − Vfix = 0 (21)
passes the objective function value and the derivatives back. The i
n˜ i,eq (T , p) · RT
derivatives are calculated by algorithmic differentiation using the was solved numerically using a Newton method [40]. The solution
Adept [37] software in conjunction with the CVODES solver that values T, p are the desired equilibrium values Teq , peq , respectively,
is part of the sundials package [38]. The CVODES solver is used to for the given Ufix and Vfix . The species moles ni,eq at this equilib-
compute sensitivities in forward-mode. rium are ni,eq = n˜ i,eq (Teq , peq ), thus fully determining the equilib-
rium at Ufix , Vfix .
2.3. Equilibrium chemistry simulations
3. Experimental methods
In addition to the detailed simulations, states along a RCM
trajectory corresponding to chemical equilibrium were simulated. A rapid compression machine (RCM) was used to perform the
This was done to bring the results of the detailed chemistry sim- experiments in place of a real engine to reduce apparatus complex-
ulations in context with a (hypothetical) process where there are ity and to allow more well-defined initial and boundary conditions.
no kinetic obstacles to the reactions. This theoretical limit aids in The RCM setup has been described in detail by Werler et al. [41],
the interpretation of the modeling and experimental results from therefore only a brief overview of the machine is given, with a de-
detailed chemistry, as will be described below. scription of some modifications.
For this, the temporal histories of internal energy U(t) and vol- The main components of the RCM are shown in Fig. 1; Table 2
ume V(t) of the RCM simulation were used as an input; internal gives a technical overview of geometrical and operating parame-
energy U(t) was computed as the mole-weighted sum of the molar ters. The central part of the setup is a piston-cylinder device that
internal energies of all species in the mixture: is used to compress a test gas within the combustion chamber.
The combustion chamber is surrounded by an oil bath to adjust

NS
the initial temperature. The piston head is equipped with a quartz
U (t ) = U i (T (t ) ) · ni (t ) , (18)
pressure transducer (Kistler 6061 B) for measuring the time re-
i=1
solved in-cylinder pressure. The position of the piston is measured
with the temperature-dependent molar internal energy of the by a potentiometric position sensor (Burster type 7812) which is
i−th species U i and its molar amount ni . Instantaneous values screwed to the piston rod. Post-processing the piston position data
T(t) and ni (t) were taken from the corresponding detailed chem- yields the volume profile of the combustion chamber over time.
istry simulation trajectory. The equilibrium states have the same A mixing vessel is used for preparation of the test gas in order
(time-dependent) internal energy and volume as the correspond- to achieve homogeneous conditions. The gases are added sequen-
ing states from detailed simulations. tially and an absolute pressure transducer (Baratron type 121A)
H. Gossler, S. Drost and S. Porras et al. / Combustion and Flame 207 (2019) 186–195 191

0.45 45 40

CO2 conversion / %
0.4 40 35
0.35 35
30
0.3 30

Pressure / bar
25
Volume / L

0.25 25
20 RCM
0.2 RCM 20
RCEM
0.15 15 15
RCEM 0.6 0.8 1 1.2 1.4 1.6
0.1 10 Initial pressure / bar
RCEM
0.05 RCM 5 Fig. 3. Comparison of CO2 conversions for RCM and RCEM operation modes at
T0 = 462 K. With increasing initial pressures, ignition occurs and the CO2 conversion
0 0
0 50 100 150 200 tends towards the same value (32%) for both modes of operation. Mole fractions of
Time / ms the gas before compression: 8% CO2 , 14.8% CH4 , 1.6% DME, 15.1% O2 , 55.5% Ar, 5%
He.
Fig. 2. Exemplary volume and pressure profiles of the experimental setup using an
ignitable CH4 /O2 /Ar mixture, highlighting the differences between RCM and RCEM
modes of operation. At t = 50 ms top-dead center is reached, which corresponds to sion. Therefore, no pressure decrease after TDC (which would be
an engine running at a speed of 600 rpm. common in a typical RCM experiment when ignition delay times
are measured) is visible.
After compression (from t = 50 ms), the difference between
measures the accumulated partial pressures. The volume of the RCM and RCEM modes becomes evident. In RCM mode, isochoric
mixing vessel and its total interior pressure are both one order of conditions prevail, with the volume remaining at the constant min-
magnitude larger than in the combustion chamber, resulting in the imum value of 0.05 L. The pressure decay in this region is solely
gas mass inside the vessel being two orders of magnitude larger. due to heat loss through the cylinder walls. In RCEM mode, the
This allows performing multiple RCM shots with the same initial expansion phase immediately follows the compression phase (hold
mixture. time tH = 0), like in a piston engine. The volume increases again af-
The machine is equipped with a creviced piston that swallows ter reaching its minimum value at TDC, resulting in a much faster
the roll-up vortex, helping to grant homogeneous thermal condi- pressure decay than in the RCM case. However, the piston only par-
tions in the compressed gas [42–44]. Furthermore, the combustion tially retracts due to geometrical constraints of the setup. Conse-
chamber is surrounded by a thermostatically controlled oil bath to quently, the volume does not reach its initial value of 0.45 L. Fur-
keep the temperature at the desired level. thermore, it can be seen that the piston slightly oscillates during
A knee lever system pushes the piston towards the right into the expansion phase.
the cylinder, thus compressing the gas in the combustion cham- After an experiment, the piston is retracted and it is waited un-
ber. Maximum compression is reached when the knee lever is fully til the temperature has fallen to near its pre-compression value.
stretched and aligned with the piston rod and the piston itself (top Then, samples from the in-cylinder gas mixture are extracted and
dead center, TDC). A bumper prevents the knee lever from moving analyzed in a gas chromatographic (GC) analyzer (Agilent 490 Mi-
past the stretched position further downwards in order to keep the cro GC). Using helium as an inert reference and argon as a carrier
piston from retracting back. In this way, isochoric conditions are gas, this GC allows to measure multiple species, including CO2 and
met after the compression phase. We refer to this mode of opera- lower hydrocarbons.
tion as “RCM mode”. Experiments were carried out using the gas mixtures of CH4 ,
Alternatively, the bumper can be removed from the setup after CO2 , O2 , DME and Ar predicted by the optimization. For simplic-
some adjustable hold time (tH = 0, . . . , ∞), and the knee lever can ity, helium was not included in the optimization. Therefore, 2 to
then move downwards past the stretched position immediately af- 5% of the argon was substituted by helium in the experiments to
ter the compression phase, which in effect retracts the piston after be able to quantify the mixture’s exact composition from the chro-
the hold time. We refer to this operation mode as “RCEM mode”, matograms. The CO2 conversion was then calculated according to
short for rapid compression expansion machine mode. This mode, Eq. (12).
with compression and subsequent expansion, is similar to a recip-
rocating internal combustion engine; various realizations of such 4. Results
rapid compression-expansion machines have been presented, and
have been used as a simplified, well-controlled replacement of pis- 4.1. RCM and RCEM operation modes
ton engines [45].
Figure 2 shows typical volume and pressure profiles of the If operated with short hold times, RCEM mode more closely
setup using an ignitable CH4 /O2 /Ar mixture, highlighting the dif- resembles conditions in a real engine than RCM mode. However,
ferences between RCM and RCEM operation modes. The solid lines running experiments in RCEM mode over prolonged time periods
denote the volume (left axis) and the dashed lines denote the pres- causes excessive wear to the setup. Therefore, RCM mode is pre-
sure (right axis). The blue lines are the profiles recorded in RCM ferred in order to avoid damage to the setup. This leads to the
mode (i.e. with the bumper in place), while the red lines were question whether results obtained in RCM mode can be meaning-
recorded in RCEM mode. The traces start at t = 0, where the cylin- fully extrapolated to real engine conditions.
der is filled with the unreacted mixture. In the compression phase Figure 3 shows a comparison between both operation modes. In
(until t = 50 ms), the volume and pressure curves of both modes these measurements, the gas mixture (specified in the figure cap-
coincide. As the volume decreases from its initial value of 0.45 L to tion) was compressed at constant initial temperature (T0 = 462 K)
0.05 L, the pressure rises from 1.0 bar to about 20 bar as per near- at varying initial pressures. Afterwards, the gas was analyzed and
adiabatic compression. In these examples, ignition occurs in both the CO2 conversions were obtained. The measurements were car-
operation modes near TDC (at 50 ms), detectable by a sharp pres- ried out in RCM mode (red crosses) and in RCEM mode (blue
sure rise up to 40 bar. Here, the ignition directly follows compres- circles).
192 H. Gossler, S. Drost and S. Porras et al. / Combustion and Flame 207 (2019) 186–195

10 15 20 25 30 35 40 45 50 CO2 conversion of 51.0% is shown in the top graph (blue line). Only
55
Experiment CH4 , DME and O2 enter the formula to calculate φ ; The oxygen
50
CO2 conversion / %

Simulation contained in the CO2 is not included in the calculation.


45
The dots are the experimental values as determined by gas-
40 chromatographic analysis of gas samples from the RCM combus-
35 tion chamber. Using the gas composition predicted by the numer-
30 ical optimization, the CO2 conversions in the experiment closely
25 follow the values of the simulations.
2.9 Note that the optimization was carried out with a constraint
Fuel
40 O2 2.8
in place that restricts the argon mole fractions to values greater
Optimal inlet mole fractions / %

Ar than or equal to 2%, as the GC unit requires a certain amount of


35 2.7 reference gas to be present. With increasing CO2 mole fractions,
30 2.6 it can be seen that the mole fraction of argon tends towards this
minimum value. While it is conceivable that the inert gas is not
25 2.5

Optimal
required here to attain similarly high CO2 conversions, we would
20 2.4 not have been able to quantify the results.
15 2.3
4.3. Evolution of states during the RCM process
10 2.2

5 2.1 Figures 5 and 6 both show the simulated temporal evolution of


0 2 selected species (top) and temperature (bottom). The conditions in
10 15 20 25 30 35 40 45 50 the simulation are the optimal values for CO2 initial mole fraction
CO2 inlet mole fractions / %
of 20%, as shown in Fig. 4. Figure 5 shows the data on a conven-
Fig. 4. Top graph: predicted (line) and experimental (dots) CO2 conversions as func-
tional time axis; this gives a realistic impression of the dynamics
tion of CO2 inlet mole fractions at T0 = 462 K and p0 = 1.0 bar. Each dot corresponds of the process. Figure 6 shows the same data, but with a differently
to the mean value of three samples analyzed in the gas chromatograph. Experimen- scaled time axis, helping to better discern the different phases of
tal and simulation results obtained in RCM mode. Bottom graph: gas composition the RCM process.
required to achieve the CO2 conversion shown on the top. The fuel is composed of
For the detailed kinetics curves (solid lines, labeled with “kin.”),
90% CH4 and 10% DME.
starting at t = 0 with unreacted mixture at the initial temperature
of 462 K and pressure of 1.0 bar, compression leads to a temper-
In Fig. 3, the initial pressure serves as a parameter to control ature rise that is visible at approx. t = 50 ms. This triggers auto-
ignition. At low inlet pressures, the mixture fails to ignite and ignition near 75 ms which then evolves into combustion, visible
hence none of the CO2 is converted. When the initial pressure is by the very steep rise in temperature and almost instantaneous
increased, ignition occurs and the CO2 conversion approaches a species conversion. At this point in time, CH4 and CO2 get con-
constant value, in this case of around 32%. This is true for both sumed very quickly, while CO rises from 0 to about 35% mole frac-
operation modes. Both modes show a minimum initial pressure tion. The temperature reaches a maximum at about t = 80.6 ms,
that is required for ignition. From the data in Fig. 3, one can rec- and then decreases rapidly due to heat losses (“cool-down” phase).
ognize that for RCEM mode, a pressure of 1.2 bar is required for It asymptotically approaches the pre-compression temperature of
stable ignition, while for RCM mode, it is 0.7 bar (although this 462 K. The species, in contrast, do not notably change in this cool-
might not be easily discernible in the figure). The reason for this is down phase. Instead, their mole fraction values remain close to
that ignition delay is longer at lower pressures. If the ignition de- the values at t = 80 ms. After t = 300 ms temperature has nearly
lay is shorter than the RCEM’s expansion time (for pressures above reached its initial value of 462 K.
1.2 bar in Fig. 3), there is no difference in CO2 conversion between The curves for equilibrium show a different behavior (dashed
RCM mode and RCEM mode. At lower pressures, the expansion in lines, labeled with “equil”.). At t = 0, the temperature at the U, V-
RCEM mode does not allow ignition to happen. Therefore, the CO2 equilibrium is already high (close to 20 0 0 K). The species mole
conversions differ between the two modes at low pressures. fractions differ from the detailed chemistry solution as well. There
Hence, RCEM mode and RCM mode deliver similar results, pro- is practically no methane at equilibrium, while more CO2 and CO
vided that ignition occurs. If ignition does occur, then the CO2 con- is present than in the kinetics simulations. During compression,
versions will be the same for both RCM and RCEM modes of oper- the equilibrium temperature changes only slightly. Its maximum
ation. We therefore conclude that data obtained in RCM mode can value of about 2100 K is reached near t = 70 ms, which is more
be used to infer information on the feasibility of the process under than 300 K below the maximum temperature in the kinetics simu-
engine conditions. A real engine may even be advantageous over lations.
the RCM, as there are more options to trigger ignition in an ICE, Between t = 80 ms and 82 ms, the chemical reactions proceed
for example by a spark. quickly due to the high temperatures, allowing the system to reach
equilibrium for all state variables (species, temperature). Conse-
4.2. CO2 conversion quently, the curves for detailed chemistry and equilibrium coin-
cide in this region (near t = 82 ms). This can be seen well in the
Figure 4 shows the optimal gas composition (bottom graph) center section of the time axis in Fig. 6. Afterwards, when the
along with the related CO2 conversions (top graph), both as func- temperature has dropped to about 1500 K, some reactions are no
tions of CO2 mole fractions at the inlet. The bottom graph should longer fast enough and the species states become “frozen”. From
be interpreted as follows. If the engine is to be operated with 20% this point on, the system can no longer adapt its chemical state to
(per volume) of CO2 in the inlet, then the optimal composition of the changing conditions during the cool-down phase. This is visi-
the remaining 80% is: 3.6% argon (red line, left axis), 33.0% oxy- ble in Fig. 6 where the detailed states then deviate from the equi-
gen (blue line, left axis), and 43.5% fuel (9:1 mixture of CH4 and librium curve at t = 88 ms. The deviation between detailed solu-
DME, black line, left axis). This mixture corresponds to a φ -value tion and equilibrium is considerable for all species. Notably, dur-
(gray line, right axis) of 2.77. The resulting maximum attainable ing cool-down, equilibrium methane rises from its previous value
H. Gossler, S. Drost and S. Porras et al. / Combustion and Flame 207 (2019) 186–195 193

0 50 100 150 200 250 300 librium. In contrast, the detailed kinetics simulations predict that
45
during cool-down CO2 stays near its low value of the “hot” phase,
40 CH (kin.)
4
where both equilibrium and detailed simulations coincide and pre-
CO (equil.) H (kin.) CO (kin.) dict xCO2 ≈ 5%.
35 2
The observed high CO2 conversion levels exceeding 50% in both
H (equil.)
Mole fractions / %

30 2
the detailed chemistry simulations and the experiment therefore
25 CO (equil.)
reflect the tendency to preserve the chemical composition of equi-
H O (kin.)
libria (that are reached in the “hot” phase of the process) through-
20 CO2 (kin.) 2
H O (equil.)
2
out the expansion. The lower temperatures that result during
CO (equil.)
15 2 expansion quench CO2 producing reactions, so that the low CO2
levels from earlier “hot” equilibria are preserved.
10
Interestingly, heat loss alone leads to a sufficiently fast tem-
CO2 (kin.)
5 CO2 (equil.)
perature decay, allowing this effect to be observed in the RCM,
CH (equil.)
4
even in absence of the expansion stroke. The expansion stroke lets
the mixture attain low temperatures much faster than heat loss
through the cylinder walls alone and therefore may present an op-
Temperature / K

2000 T (equil.)
tion to further improve CO2 conversions. However, when the RCM
1500 is operated in this way (RCEM mode), ignition timing becomes a
key issue (see also Fig. 3).
T (equil.)
1000 Figure 6 shows that CO2 conversion occurs on a timescale of
T (kin.) a few milliseconds. Within this timescale, a mixture with initially
T (kin.)
500 uniform temperature distribution will develop a non-uniform dis-
0 50 100 150 200 250 300 tribution due to heat transfer from the gas to the cylinder walls.
Time / ms Despite the fact that the simulation model does not take into ac-
Fig. 5. Temporal evolution of selected species (top) and temperature (bottom) dur- count effects of thermal stratification, the experiment and simula-
ing a RCM process, showing detailed kinetics (solid lines, “kin.”) and equilibrium tion show a good agreement in achieved CO2 conversion.
(dashed lines, “equil.”) simulations. Initial conditions correspond to the optimal val- At a first glance, this agreement would be surprising. There is
ues for CO2 initial mole fraction of 20%: T0 = 462 K, p0 = 1.0 bar, 20% CO2 , 39.15% a plausible explanation, however: We propose that for the studied
CH4 , 4.35% DME, 33.0% O2 , 3.6% Ar.
conditions, the detailed dynamics of combustion and heat release
during the in-cylinder process are not crucial for the final amount
0 20 40 60 80 80.5 81 81.5 82 100 150 200 of CO2 conversion. The important requirements for the conversion
45
are that
40 CH 4 (kin.)
CO (equil.) • the gas reaches a sufficiently high maximum temperature to be
35 H
2 able to overcome kinetic barriers and approach chemical equi-
Mole fractions / %

30 librium
CO (equil.) • the cool-down phase is fast enough to let the chemical state
25
freeze at a point of some high-temperature equilibrium
20 CO (kin.)
2
H O
2
These requirements can be achieved in different regions of the
15 CO2 (equil.) in-cylinder mixture (e.g., central or closer to walls), even if they
10
are affected by heat transfer to a different extent, and with differ-
CO2 (kin.) ent timing during the cycle. For instance, even near-wall regions,
5 CO (equil.)
2
CH (equil.)
which during an engine cycle first would remain relatively cold
4
by heat transfer, can later attain the high temperatures required
for efficient CO2 conversion by thermal contact to burned regions.
Temperature / K

2000 T (equil.)
They can auto-ignite by mixing with hot burned gas, and then fur-
ther increase their temperature by chemical reaction.
1500
5. Conclusions
1000

T (kin.) Our work shows that CO2 conversion by dry reforming yielding
500 CO and H2 can be carried out in piston engines. As a by-product,
0 20 40 60 80 80.5 81 81.5 82 100 150 200 water is produced. From this viewpoint, the engine can be seen
Time / ms
as a multi-purpose chemical converter, which is both capable of
Fig. 6. Temporal evolution of selected species (top) and temperature (bottom) dur- producing power (in conventional combustion mode), but also able
ing a RCM process, showing detailed kinetics (solid lines, “kin.”) and equilibrium to deliver useful products (syngas) from CO2 .
(dashed lines, “equil.”) simulations. Same data as in Fig. 5, but with differently A combination of numerical simulation and optimization was
scaled time axis, helping to better discern the different phases of the RCM process.
used to identify operating conditions that yield high conversions,
Center section: “hot” phase. Right section: “cool-down” phase.
exceeding 50%; experiments in a rapid compression machine show
that these predictions are realistic, both with respect to the opti-
of zero up to xCO2 ,eq = 7%. This is in accordance with the princi- mal conditions and with respect to the achieved conversions. The
ple of Le Chatelier and Braun for the endothermal CO2 reforming optimization approach was not only helpful to reduce the number
reaction CO2 + CH4  2 CO + 2 H2 . Also, the equilibrium mole of experiments, but also to determine whether CO2 conversion in
fraction of CO2 rises during the cool-down phase to approx. 16%, an engine is possible in the first place.
which is close to its pre-compression value. Hence, no notable One result of the optimization is that addition of oxygen to the
CO2 conversion would result if the system always followed equi- initial mixture of CH4 and CO2 helps improve the conversion by
194 H. Gossler, S. Drost and S. Porras et al. / Combustion and Flame 207 (2019) 186–195

providing the high temperatures (via exothermal chemical reac- References


tions during oxidation of CH4 ) that overcome kinetic barriers of
the underlying elementary reactions, and that also shift the equi- [1] E. Schwab, A. Milanov, S.A. Schunk, A. Behrens, N. Schödel, Dry reforming and
reverse water gas shift: Alternatives for syngas production? Chemie Ingenieur
librium of the overall chemical reaction to the desired right side. Technik 87 (4) (2015) 347–353, doi:10.10 02/cite.20140 0111.
Despite the combustion of CH4 that then takes place, there is still [2] A. Giehr, L. Maier, S.A. Schunk, O. Deutschmann, Thermodynamic considera-
a net conversion of CO2 . tions on the oxidation state of Co/γ -Al2 O3 and No/γ -Al2 O3 catalysts under
dry and steam reforming conditions, ChemCatChem 10 (4) (2018) 751–757,
In fact, for the typical engine conditions studied here, ignition doi:10.1002/cctc.201701376.
is crucial for any significant CO2 conversions to occur. The methane [3] A.W. Budiman, S.-H. Song, T.-S. Chang, C.-H. Shin, M.-J. Choi, Dry reforming
oxidation was a required aid that delivered the required tempera- of methane over cobalt catalysts: a literature review of catalyst development,
Catal. Surv. Asia 16 (4) (2012) 183–197, doi:10.1007/s10563- 012- 9143- 2.
tures which were not attainable by compression alone in the par-
[4] T. Roussière, K.M. Schelkle, S. Titlbach, G. Wasserschaff, A. Milanov,
ticular RCM setup used here. How these temperatures are attained G. Cox, E. Schwab, O. Deutschmann, L. Schulz, A. Jentys, J. Lercher,
does not play a role. If alternative methods for providing the high S.A. Schunk, Structure–activity relationships of Nickel–Hexaaluminates in
reforming reactions Part I: Controlling Nickel nanoparticle growth and
temperatures (e.g., higher compression rations, higher initial tem-
phase formation, ChemCatChem 6 (5) (2014a) 1438–1446, doi:10.1002/cctc.
peratures, etc.) would exist, then these would allow dry reforming 201300960.
to proceed without the oxidation. In our study, we relied on auto- [5] T. Roussière, L. Schulz, K.M. Schelkle, G. Wasserschaff, A. Milanov, E. Schwab,
ignition to achieve conversion. In a real engine, there are more O. Deutschmann, A. Jentys, J. Lercher, S.A. Schunk, Structure–activity relation-
ships of Nickel–Hexaaluminates in reforming reactions Part II: Activity and
options to trigger ignition, for example, spark ignition. Although stability of nanostructured nickel–hexaaluminate-based catalysts in the dry re-
RCEM mode more closely resembles real engine conditions with forming of methane, ChemCatChem 6 (5) (2014b) 1447–1452, doi:10.1002/cctc.
its compression and expansion strokes, we found experimentally 201300958.
[6] K. Takanabe, K. Nagaoka, K. Nariai, K.-i. Aika, Titania-supported cobalt and
that considering the simpler RCM mode (compression stroke only) nickel bimetallic catalysts for carbon dioxide reforming of methane, J. Catal.
is sufficient to predict the CO2 conversion under engine conditions. 232 (2) (2005) 268–275, doi:10.1016/j.jcat.2005.03.011.
The dynamics of the conversion process can be regarded as a [7] L.A. Schulz, L.C.S. Kahle, K.H. Delgado, S.A. Schunk, A. Jentys, O. Deutschmann,
J.A. Lercher, On the coke deposition in dry reforming of methane at elevated
two-step sequence: After compression and combustion, high tem- pressures, Appl. Catal. A: Gen. 504 (2015) 599–607, doi:10.1016/j.apcata.2015.
peratures exceeding 2400 K result. This allows reactions to quickly 03.002.
proceed to chemical equilibrium, which at these high temperatures [8] L. Tartakovsky, M. Sheintuch, Fuel reforming in internal combustion engines,
Prog. Energy Combust. Sci. 67 (2018) 88–114, doi:10.1016/j.pecs.2018.02.003.
is on the CO/H2 side of the overall reaction equation. In the fol-
[9] L. Von Szeszich, Herstellung von Synthesegas im Otto-Motor bei gleichzeitiger
lowing cool down/expansion phase, internal energy and pressure Arbeitsgewinnung, Chemie Ingenieur Technik 28 (3) (1956) 190–195, doi:10.
decrease rapidly. The chemistry in the combustion chamber first 1002/cite.330280310.
[10] M.H. McMillian, S.A. Lawson, Experimental and modeling study of hydro-
follows the equilibrium for these transient thermodynamical con-
gen/syngas production and particulate emissions from a natural gas-fueled
ditions, with a tendency to proceed back to the CH4 /CO2 side. partial oxidation engine, Int. J. Hydrogen Energy 31 (7) (2006) 847–860, doi:10.
However, these reverse reactions are quenched at lower tempera- 1016/j.ijhydene.2005.08.013.
tures (approx. 1500 K at the studied conditions), largely preserving [11] G.A. Karim, I. Wierzba, The production of hydrogen through the uncatalyzed
partial oxidation of methane in an internal combustion engine, Int. J. Hydrogen
the favorable equilibrium from the previous “hot” phase. This is a Energy 33 (8) (2008) 2105–2110, doi:10.1016/j.ijhydene.2008.01.051.
key phenomenon for the success of the conversion process. [12] Y.C. Yang, M.S. Lim, Y.N. Chun, The syngas production by partial oxidation us-
The experimental setup used in this work had a compression ing a homogeneous charge compression ignition engine, Fuel Process. Technol.
90 (4) (2009) 553–557, doi:10.1016/j.fuproc.2009.01.002.
ratio of approximately 10, and the maximum initial temperature [13] E.G. Lim, E.E. Dames, K.D. Cedrone, A.J. Acocella, T.R. Needham, A. Arce,
was 462 K. These two points are the limiting factors to the maxi- D.R. Cohn, L. Bromberg, W.K. Cheng, W.H. Green, The engine reformer: Syn-
mum attainable temperature in the combustion chamber by work gas production in an engine for compact gas-to-liquids synthesis, Can. J. Chem.
Eng. 94 (4) (2016) 623–635, doi:10.1002/cjce.22443.
alone. Future work will focus on whether CO2 can be converted [14] R. Hegner, M. Werler, R. Schießl, U. Maas, B. Atakan, Fuel-rich HCCI engines
without combustion. Furthermore, the H2 /CO ratio of the syngas as chemical reactors for polygeneration: A modeling and experimental study
mixture that exits the engine is quite low. For downstream conver- on product species and thermodynamics, Energy Fuels 31 (12) (2017) 14079–
14088, doi:10.1021/acs.energyfuels.7b02150.
sion processes the mixture would need to be enriched with H2 .
[15] S. Wiemann, R. Hegner, B. Atakan, C. Schulz, S.A. Kaiser, Combined production
The conversion levels we found are encouraging for use of dry of power and syngas in an internal combustion engine – Experiments and sim-
reformation in technical applications. A real engine offers addi- ulations in SI and HCCI mode, Fuel 215 (2018) 40–45, doi:10.1016/j.fuel.2017.11.
002.
tional parameters to fine-tune the process, so that the achievable
[16] M.H. Morsy, Modeling study on the production of hydrogen/syngas via par-
conversions can likely be increased even more. tial oxidation using a homogeneous charge compression ignition engine fueled
with natural gas, Int. J. Hydrogen Energy 39 (2) (2014) 1096–1104, doi:10.1016/
Acknowledgments j.ijhydene.2013.10.160.
[17] H. Gossler, O. Deutschmann, Numerical optimization and reaction flow analysis
of syngas production via partial oxidation of natural gas in internal combus-
This research was supported by the Deutsche Forschungs- tion engines, Int. J. Hydrogen Energy 40 (34) (2015) 11046–11058, doi:10.1016/
gemeinschaft within the framework of the DFG research unit j.ijhydene.2015.06.125.
[18] R. Hegner, B. Atakan, A polygeneration process concept for HCCI-engines –
FOR 1993 “Multi-functional conversion of chemical species and Modeling product gas purification and exergy losses, Int. J. Hydrogen Energy
energy” (projects SCHI647/3-2 and DE659/10-2). Hendrik Gossler 42 (2) (2017) 1287–1297, doi:10.1016/j.ijhydene.2016.09.050.
gratefully acknowledges financial support by the Karlsruhe House [19] C.-J. Sung, H.J. Curran, Using rapid compression machines for chemical kinetics
studies, Prog. Energy Combust. Sci. 44 (2014) 1–18, doi:10.1016/j.pecs.2014.04.
of Young Scientists (KHYS) at KIT for a research visit at the Col-
001.
orado School of Mines, USA. A cost-free license of DETCHEM TM [20] G. Woschni, A Universally Applicable Equation for the Instantaneous Heat
and financial support by Steinbeis GmbH & Co. KG für Technolo- Transfer Coefficient in the Internal Combustion Engine, SAE International, War-
rendale, PA, 1967.
gietransfer (STZ 240 Reaktive Strömung) are gratefully acknowl-
[21] R.G. Gilbert, K. Luther, J. Troe, Theory of thermal unimolecular reactions in
edged. the fall-off range. II. Weak collision rate constants, Berichte der Bunsenge-
sellschaft für physikalische Chemie 87 (2) (1983) 169–177, doi:10.1002/bbpc.
Supplementary material 19830870218.
[22] N. Fuhrmann, M. Schneider, C.-P. Ding, J. Brübach, A. Dreizler, Two-dimensional
surface temperature diagnostics in a full-metal engine using thermographic
Supplementary material associated with this article can be phosphors, Meas. Sci. Technol. 24 (2013) 9pp.
found, in the online version, at doi:10.1016/j.combustflame.2019.05. [23] K. Kohse-Höinghaus, Clean combustion: Chemistry and diagnostics for a sys-
tems approach in transportation and energy conversion, Prog. Energy Combust.
031. Sci. 65 (2018) 1–5, doi:10.1016/j.pecs.2017.10.001.
H. Gossler, S. Drost and S. Porras et al. / Combustion and Flame 207 (2019) 186–195 195

[24] H. Wang, X. You, A.V. Joshi, S.G. Davis, A. Laskin, F. Egolfopoulos, C.K. Law, USC [35] J. Nocedal, S. Wright, Numerical optimization, Springer Series in Operations
Mech Version II. High-temperature combustion reaction model of H2 /CO/C1-C4 Research and Financial Engineering, 2, Springer-Verlag, New York, 2006.
compounds, 2007, http://ignis.usc.edu/USC_Mech_II.htm. [36] O. Deutschmann, S. Tischer, C. Correa, D. Chatterjee, S. Kleditzsch, V.M. Janard-
[25] H.J. Curran, S.L. Fischer, F.L. Dryer, The reaction kinetics of dimethyl ether. II: hanan, N. Mladenov, H.D. Minh, H. Karadeniz, M. Hettel, H. Gossler, DETCHEM
Low-temperature oxidation in flow reactors, Int. J. Chem. Kinet. 32 (12) (20 0 0) Software Package v2.7 (www.detchem.com), 2018,
741–759, doi:10.1002/1097- 4601(2000)32:12<741::AID- KIN2>3.0.CO;2- 9. [37] R.J. Hogan, Fast reverse-mode automatic differentiation using expression tem-
[26] C.K. Westbrook, M. Mehl, W.J. Pitz, G. Kukkadapu, S. Wagnon, K. Zhang, Multi- plates in C++, ACM Trans. Math. Softw. 40 (4) (2014) 26:1–26:16, doi:10.1145/
fuel surrogate chemical kinetic mechanisms for real world applications, Phys. 2560359.
Chem. Chem. Phys. 20 (16) (2018) 10588–10606, doi:10.1039/C7CP07901J. [38] A.C. Hindmarsh, P.N. Brown, K.E. Grant, S.L. Lee, R. Serban, D.E. Shumaker,
[27] B. Chan, J.M. Simmie, Barriometry – an enhanced database of accurate bar- C.S. Woodward, SUNDIALS: Suite of nonlinear and differential/algebraic equa-
rier heights for gas-phase reactions, Phys. Chem. Chem. Phys. 20 (16) (2018) tion solvers, ACM Trans. Math. Softw. 31 (3) (2005) 363–396, doi:10.1145/
10732–10740, doi:10.1039/C7CP08045J. 1089014.1089020.
[28] D.L. Baulch, C.T. Bowman, C.J. Cobos, R.A. Cox, T. Just, J.A. Kerr, M.J. Pilling, [39] S.B. Pope, Gibbs function continuation for the stable computation of chemical
D. Stocker, J. Troe, W. Tsang, R.W. Walker, J. Warnatz, Evaluated kinetic data equilibrium, Combust. Flame 139 (3) (2004) 222–226.
for combustion modeling: Supplement II, J. Phys. Chem. Ref. Data 34 (3) (2005) [40] P. Deuflhard, Newton methods for nonlinear problems: Affine invariance and
757–1397, doi:10.1063/1.1748524. adaptive algorithms, Springer, 2011, doi:10.1007/978- 3- 642- 23899- 4.
[29] B.S. Haynes, H.G. Wagner, Soot formation, Prog. Energy Combust. Sci. 7 (4) [41] M. Werler, L. Cancino, R. Schießl, U. Maas, C. Schulz, M. Fikri, Ignition delay
(1981) 229–273, doi:10.1016/0360-1285(81)90 0 01-0. times of diethyl ether measured in a high-pressure shock tube and a rapid
[30] Y. Hidaka, K. Sato, Y. Henmi, H. Tanaka, K. Inami, Shock-tube and modeling compression machine, Proc. Combust. Inst. 35 (1) (2015) 259–266, doi:10.1016/
study of methane pyrolysis and oxidation, Combust. Flame 118 (3) (1999) 340– J.PROCI.2014.06.143.
358, doi:10.1016/S0010-2180(99)00010-3. [42] C.-j. Sung, H.J. Curran, Using rapid compression machines for chemical kinetics
[31] C. Heghes, C1 -C4 hydrocarbon oxidation mechanism, University of Heidelberg, studies, Prog. Energy Combust. Sci. 44 (2014) 1–18, doi:10.1016/j.pecs.2014.04.
Germany, 2006 (Ph.D. thesis). 001.
[32] Z. Zhao, M. Chaos, A. Kazakov, F.L. Dryer, Thermal decomposition reaction and [43] D. Lee, S. Hochgreb, Rapid compression machines: Heat transfer and suppres-
a comprehensive kinetic model of dimethyl ether, Int. J. Chem. Kinet. 40 (1) sion of corner vortex, Combust. Flame 114 (3–4) (1998) 531–545, doi:10.1016/
(2008) 1–18, doi:10.1002/kin.20285. S0 010-2180(97)0 0327-1.
[33] S. Porras, R. Schießl, U. Maas, A chemical kinetic modeling study of ignition in [44] G. Mittal, M.P. Raju, C.-J. Sung, Vortex formation in a rapid compression ma-
fuel-rich methane/dimethyl ether mixtures for polygeneration processes, 8th chine: Influence of physical and operating parameters, Fuel 94 (2012) 409–417,
European Combustion Meeting 2017 (2017). Dubrovnik, Croatia doi:10.1016/J.FUEL.2011.08.034.
[34] D. Kaczmarek, S. Porras, F. Sen, T. Kasper, R. Schießl, U. Maas, B. Atakan, [45] S.S. Goldsborough, C.J. Potokar, The Influence of Crevice Flows and Blow-By
Kinetische Untersuchung der partiellen Oxidation von Methan/DME- und on the Charge Motion and Temperature Profiles Within a Rapid Compression
Methan/Ethanol-Gemischen, 28. Deutscher Flammentag (2017). Darmstadt, Expansion Machine Used for Chemical Kinetic (HCCI) Studies, SAE Technical
Germany Paper Series (2007-01-0169) (2007) 776–790, doi:10.4271/2007- 01- 0169.

Vous aimerez peut-être aussi