Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Combustion and Flame: Lei Zhang, Xiaohua Ren, Zhigang Lan

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

Combustion and Flame 212 (2020) 377–387

Contents lists available at ScienceDirect

Combustion and Flame


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

A reduced reaction mechanism of biodiesel surrogates with low


temperature chemistry for multidimensional engine simulation
Lei Zhang a,∗, Xiaohua Ren a, Zhigang Lan b
a
Beijing Key Laboratory of Process Fluid Filtration and Separation, College of Mechanical and Transportation Engineering, China University of Petroleum,
Beijing 102249, China
b
CNOOC Research Institute Co. Ltd., Beijing 100028, China

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

Article history: A reduced biodiesel mechanism composed of 156 species and 589 reactions is reduced from an origi-
Received 10 August 2019 nal complex mechanism (3299 species and 10806 reactions) based on MD, MD9D, and n-heptane as the
Revised 10 September 2019
surrogates. The mechanism reduction is conducted using the path flux analysis method, which considers
Accepted 6 November 2019
multiple reaction path generations in the analysis of species interactions, and isomer lumping. Calcula-
tions of homogeneous auto-ignition and perfectly stirred reactor (PSR) combustion on a variety of reac-
Keywords: tion states, including pressures from 1 to 100 atm and equivalence ratios from 0.5 to 2, are the basis of
Biodiesel surrogate the reduction. The initial temperatures are from 700 to 1800 K for the auto-ignition, and the inlet temper-
Mechanism reduction ature is 300 K for the PSR. These reaction states cover the high-pressure and low-temperature operating
Low-temperature chemistry
conditions of future engines using advanced combustion technologies characterized by fuel–air premix-
Auto-ignition
ing and auto-ignition. The fidelity of the resulting reduced mechanism with low-temperature chemistry
Engine emission
is examined using a variety of applications. Close agreements between the reduced and original mecha-
nisms are obtained in the predictions of ignition delay, history of mixture temperature, and species mole
fraction during homogeneous auto-ignition and the temperature profile in PSR. The reduced mechanism,
further integrated with a nitrogen oxides chemistry and a two-step soot model, is implemented into the
KIVA/CHEMKIN program for the 3D simulation of biodiesel spray combustion. The predicted liquid and
vapor penetrations agree with the experimental data in a non-reactive biodiesel spray simulation, indi-
cating an accurate estimation of biodiesel physical properties. In the simulation of biodiesel spray com-
bustion, predicted spatial distributions of hydroxyl radical and soot also agree with the corresponding
experimental data.
© 2019 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

1. Introduction using renewable fuels, such small-scale turbulence-chemistry in-


teractions need to be accurately accounted for in multidimensional
To further enhance thermal efficiency and reduce emission engine simulation, in addition to the in-cylinder dynamics of en-
levels, future international combustion engines tend to rely on gine cylinder scales. Therefore, stringent requirement has been im-
advanced combustion technologies characterized by fuel–air pre- posed on the accuracy of the combustion reaction mechanism of
mixing and volumetric auto-ignition under low-temperature and renewable fuels for engine simulation.
high-pressure operating conditions [1,2]. The combustion processes Biodiesel has been well recognized as a promising alterna-
under such extreme conditions are strongly determined by fuel re- tive fuel for conventional compression ignition (CI) engines with-
activity, which is in turn a function of fuel molecular structure. out the need of major modification to engine design. The fuel
The adaptability to renewable alternative fuels will also be empha- is oxygenated and composed of numerous components with var-
sized for future engines to enhance energy sustainability. Thus, fu- ious carbon chain lengths and differs in composition determined
eling of renewable fuels in future engines can inevitably lead to by the raw material (e.g., vegetable oils and animal fats) for pro-
previously unexplored turbulence-chemistry interactions and new duction. Despite slightly increased emission of nitrogen oxides
turbulent flame regimes. To assist the design of future engines (NOx ) at certain operating conditions [3,4], fueling of biodiesel
(usually by mixing with diesel fuel) in CI engines can effectively
reduce the emissions of unburned hydrocarbon, carbon monox-

Corresponding author. ide, and particulate matters [5–7]. In addition, the fuel also con-
E-mail address: zlei@cup.edu.cn (L. Zhang). tains no sulfur elements, producing no sulfur oxides or sulfides.

https://doi.org/10.1016/j.combustflame.2019.11.002
0010-2180/© 2019 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
378 L. Zhang, X. Ren and Z. Lan / Combustion and Flame 212 (2020) 377–387

Fig. 1. Biodiesel combustion mechanism based on MD, MD9D, and n-heptane as the surrogates.

Biodiesel derived from soybean and rapeseed is mainly composed MD (or MD9D), as described in Fig. 1. This complex mechanism
of two saturated fatty acid methyl esters, i.e., methyl palmitate can emulate the features of long carbon chain and ester group
(C17 H34 O2 ) and methyl stearate (C19 H38 O2 ), and three unsaturated of actual biodiesel methyl esters, and the unsaturation due to
fatty acid methyl esters, i.e., methyl oleate (C19 H36 O2 ), methyl double bond can be considered by MD9D. However, due to the
linoleate (C19 H34 O2 ), and methyl linolenate (C19 H32 O2 ), featuring large numbers of species and reactions, this mechanism is com-
long carbon chains with an ester group and different levels of un- putationally expensive and cannot be used in three-dimensional
saturation. simulation of engine in-cylinder dynamics. Thus, reducing this
Considering the large molecular size of biodiesel methyl esters, complex mechanism to an appropriate size without significantly
a more practical way to simulate biodiesel combustion is to find impairing the computational fidelity is desired.
appropriate surrogates and develop the chemistry based on the The reduction of reaction mechanism has been extensively
surrogates. Commonly used surrogates for biodiesel mechanisms studied with various categories of methods developed over the
are mainly from two categories of molecules, i.e., methyl esters past decades. Skeletal reduction is one of the categories, and it
and n-alkanes. Methyl butanoate (MB) has been used as the sur- finds and eliminates unimportant species and reactions to finally
rogate to simulate the biodiesel combustion, because of the exis- obtain a computationally efficient mechanism without losing the
tence of an ester group in its molecular structure [8–10]. However, key features of the original mechanism. Starting from a target
limited by its short carbon chain length, MB has been found to species, this method estimates the relation between the target
be insufficient for biodiesel combustion modeling, primarily due to species and each of other species and reactions by calculating a
its lower reactivity compared to soybean biodiesel. n-hexadecane correlation coefficient. If the calculated correlation coefficient is
is the long-chain alkane found to be a good surrogate for rapeseed less than a prespecified error tolerance, this species or reaction is
methyl ester (RME) [11,12]. The mechanism of n-hexadecane shows found to be unimportant and can be eliminated from the mech-
a reactivity in an agreement with that of actual RME measured anism. Either a species or a reaction can be used as the starting
by the experiment, but it failed to predict the early production of point of the reduction process. Commonly used methods include
CO2 due to the ester group, as observed in the same experiment. the principal component analysis [17,18], Jacobian analysis [19],
Consequently, according to the performance of the above two cat- sensitivity analysis [20,21], directed relation graph (DRG) [22,23],
egories of surrogates, methyl esters with long carbon chain are the and its various derived versions, such as the DRG with error
optimum surrogates for biodiesel. propagation [24,25]. Note that the DRG method is characterized
In the molecular structure of some soy and rapeseed methyl by a linear reduction time, making it highly efficient to reduce
esters, there can be multiple double bonds, leading to different extremely large mechanisms, and is usually used as the first step
levels of unsaturation. The existence of unsaturated fatty acid in the mechanism reduction. Lu et al. [26] extended the DRG
methyl esters in the fuel can have a strong negative impact on method with expert knowledge (DRGX) by specifying a specific
the overall reactivity of the fuel–air mixture, especially in the low error tolerance for each target species. The resulting DRGX method
temperature regime [13]. Double bonds can lead to the formation can produce a skeletal mechanism capable of highly accurate pre-
of unsaturated species which can also be the precursors of soot dictions of heat release and concentrations of the target species,
particles. Therefore, a biodiesel reaction mechanism needs to care- while moderate accuracies are also achieved for other species.
fully deal with the effects of unsaturation. Biodiesel mechanisms Note that the DRG based skeletal method only considers the
based on the mixture of various types of surrogates have been pro- connection between the target species and species directly related.
posed to consider the combined effects of long carbon chain, ester However, species not directly related to the target species can also
group, and unsaturation. Golovitchev and Yang [14] modeled RME have strong impacts on the target species via intermediate species,
as the mixture of methyl butanoate, n-heptane, and phenyl methyl and it is valuable to consider this factor in the reduction process
ether (C7 H8 O). Ng et al. [15] proposed a compact biodiesel-diesel to further improve the accuracy of the reduced mechanism. Based
combined reaction mechanism based on MB, methyl crotonate, on the second or even higher generations of reaction fluxes, Sun et
and n-heptane, which represent saturated fatty acid methyl ester, al. [27] developed the path flux analysis (PFA) method to identify
unsaturated fatty acid methyl ester, and straight-chain hydrocar- the importance of species indirectly related to the target species.
bon, respectively. A complex mechanism including 3299 species Compared to the DRG method, improved accuracy of the reduced
and 10806 reactions was developed by Lawrence Livermore Na- mechanism was achieved using the PFA method considering the
tional Laboratory (LLNL) using the mixture of methyl decanoate first two generations of fluxes, without increasing the mechanism
(MD) as a saturate surrogate, methyl-9-decenoate (MD9D) as an size. In addition, isomer lumping [28,29] is another category of
unsaturated surrogate, and n-heptane as an alkane surrogate for reduction method, and it lumps similar species and reactions
biodiesel fuels [16]. It is assumed that each of the five main to reduce the number of variables that need to be tracked in
components undergoes a rapid decomposition into n-heptane and chemical reaction computation.
L. Zhang, X. Ren and Z. Lan / Combustion and Flame 212 (2020) 377–387 379

The DRG-based methods have been used to develop reduced range of reaction states ensures that the resulting reduced mecha-
biodiesel reaction mechanisms for high-temperature conditions. nism is applicable for the operating conditions of traditional CI en-
Using DRG, isomer lumping, and optimized DRG-aided sensitiv- gines and advanced engine combustion technologies characterized
ity analysis, Luo et al. [30] developed a reduced biodiesel mech- by volumetric auto-ignition and low-temperature combustion.
anism including 118 species and 837 reactions from the original
LLNL mechanism. The reduced biodiesel mechanism is only valid 2.1.1. Path flux analysis reduction
for temperatures above 10 0 0 K. However, considering the dom- The PFA method is used first to identify and remove relatively
inance of low-temperature ignition in future engines using ad- unimportant species and reactions from the original biodiesel reac-
vanced combustion technologies, mechanisms only valid for high- tion mechanism. As an alternative to the DRG approach, which is
temperature conditions are not applicable for the corresponding based on forward and reverse reaction rates, the PFA method uses
numerical studies. Under low-temperature (below 10 0 0 K) and production and consumption fluxes to evaluate the interaction be-
high-pressure conditions, determined by the long-chain molecu- tween two species and identify important reaction pathways. The
lar structures of major biodiesel components, elementary reac- production flux, PA , and consumption flux, CA , for species A are de-
tions can lead to strong negative temperature coefficient (NTC) fined by
effect, which is manifested by the short ignition delay compara-

I
ble to the combustion time scale in engines. The low-temperature PA = max (0, νA,i ωi ) (1)
auto-ignition of biodiesel can have a strong impact on the en-
i=1
gine combustion process, and thus the reduced biodiesel mecha-
nism also needs to be valid for conditions below 10 0 0 K. Since and
the low-temperature chemistry involves numerous key radicals 
I

and intermediate species, as well as complex reaction pathways, CA = max (0, −νA,i ωi ). (2)
the reduced biodiesel reaction mechanism with low-temperature i=1

chemistry is significantly larger than those only valid for high- In the above two equations, I is the total number of reactions in
temperature conditions. Compact biodiesel schemes, including less which species A is involved. ν A, i and ωi are the stoichiometric co-
than 100 species and 200 reactions, have been developed using efficient and net production rate for species A by the ith reaction,
the DRG method for multidimensional engine simulation [31–33]. respectively. The net production rate (ωi ) is the difference between
The accuracy of these compact mechanisms was found to be in- forward (ωf, i ) and reverse (ωb, i ) reaction rates, i.e.,
sufficient, due to either the limited mechanism size or inappro-
priate biodiesel surrogates (e.g. MB). Using the DRG-based reduc- ωi = ωf,i − ωb,i . (3)
tion methods, Luo et al. [34,35] obtained larger reduced biodiesel To evaluate the interaction coefficient, rAB , which evaluates the
mechanisms (123 species and 394 reactions) with low-temperature error induced to the target species A due to the elimination of
chemistry. However, as mentioned above, the limitation of DRG- species B from the mechanism, the production and consumption
based methods is that only direct species interactions are analyzed. fluxes of species A related to species B are calculated using the net
Considering the complexity of low-temperature chemistry, it is im- production rate as
portant to analyze indirect species interactions via intermediate
species in the reduction of biodiesel mechanism. 
I

The purpose of the present study is to develop a reduced


PAB = max (0, νA,i ωi δB,i ) (4)
i=1
biodiesel mechanism, which can be used in the numerical study
of biodiesel applications in conventional CI engines and advanced and
engine combustion technologies characterized by volumetric auto- 
I
ignition and low-temperature combustion. The original biodiesel CAB = max (0, −νA,i ωi δB,i ). (5)
mechanism, including 3299 species and 10806 reactions, is the one i=1
developed by LLNL based on the three surrogates, i.e., MD, MD9D, δ B ,i is the delta function, which is equal to 1, if species A and B are
and n-heptane, and it is reduced in the present study using the simultaneously involved in the ith reaction, and 0, otherwise. Using
PFA method to consider multiple reaction path generations and the production and consumption fluxes, the interaction coefficients
isomer lumping. The resulting reduced mechanism is further com- for the production and consumption of species A in direct relation
bined with a NOx chemistry and a two-step soot model for use with species B are defined by
in the multidimensional simulation of engine operation and emis-
sions. The fidelity of the reduced mechanism is comprehensively p PAB
rAB = (6)
examined by comparing with the original mechanism in calcu- max (PA , CA )
lating the combustion characteristics of various homogeneous ap- and
plications. Experimental data of three-dimensional biodiesel spray
c CAB
flame under CI engine operating conditions are also used to vali- rAB = , (7)
date the reduced mechanism. max (PA , CA )
where superscripts “p” and “c” denote production and consump-
2. Model formulation tion, respectively.
The interaction coefficients determined by the above two equa-
2.1. Mechanism reduction tions only involve the first-generation reaction fluxes, which evalu-
ate the directly relation with the target species A. However, species
In the present study, the reduction of the biodiesel reaction B can also be indirectly related to the target species A via in-
mechanism is performed on a large range of reaction states, with termediate species, involving the second or higher generations of
pressures from 1 to 100 atm and equivalence ratios () from 0.5 reaction fluxes, and this indirect interaction can also be impor-
to 2.0, representing lean, stoichiometric, and rich combustion con- tant to the production or consumption of species A. In the present
ditions. The initial temperatures for homogeneous auto-ignition study, in addition to the first-generation reaction flux, the second-
are from 700 to 1800 K, and the inlet temperature for the com- generation reaction flux is also considered in the interaction anal-
bustion in the perfectly stirred reactor (PSR) is 300 K. The large ysis, and it has been proven that improved fidelity of the reduced
380 L. Zhang, X. Ren and Z. Lan / Combustion and Flame 212 (2020) 377–387

mechanism can be obtained by adding the second-generation re- Table 1


Lumped isomers for the reduction of biodiesel reaction mechanism.
action flux in the analysis [27]. Assume that species A and B are
also indirectly related via a third species Ni , and the interaction Lumped species Isomer species
coefficients for the second-generation are calculated by md3j md3j, md4j, md5j, md6j, md7j, md8j, md9j

p−2
   md3o2 md3o2, md5o2, md6o2
rAB = rApN rNp B (8) md3ooh5j md3ooh5j, md5ooh3j, md6ooh4j
i i
Ni = A ,B md3ooh5o2 md3ooh5o2, md5ooh3o2, md6ooh4o2
md9d3j md9d3j, md9d4j, md9d6j, md9d7j
md9d3o2 md9d3o2, md9d6o2, md9d8o2, md8dxo2
and
c7h15-2 c7h15-2, c7h15-3, c7h15-4
c−2
   c7h15o2-2 c7h15o2-2, c7h15o2-3
rAB = rAc Ni rNc i B . (9) c7h14ooh2-4o2 c7h14ooh2-4o2, c7h14ooh3-5o2
Ni = A ,B nc7ket24 nc7ket24, nc7ket35
md9doh9 md9doh9, md9doh10
The above interaction coefficients can be then used in the
mechanism reduction with an error tolerance prespecified. Species
B is found to be important to species A, if the interaction coeffi- L2 , participate in the following elementary reactions with the cor-
cient is greater than the error tolerance, and it is retained in the responding reaction rates, ωi , i.e.,
mechanism. Otherwise, the species is eliminated. The four interac-
k1
tion coefficients defined by the above equations are grouped into R 1 : A 1 → L 1 , ω 1 = k 1 [A 1 ]
an overall interaction coefficient for species B, and thus only one
k2
error tolerance is needed, i.e., R 2 : A 2 → L 2 , ω 2 = k 2 [A 2 ]
p c p−2 c−2
rAB = rAB + rAB + rAB + rAB . (10) k3
R 3 : L 1 → B 1 , ω 3 = k 3 [L 1 ]
Since the purpose of the present study is to obtain a reduced k4
biodiesel mechanism for use in the three-dimensional simulation R4 : L2 → B2 , ω4 = k4 [L2 ]. (11)
of engine in-cylinder dynamics, it is important for the reduced ki is the reaction rate coefficient of the ith reaction. If L1 and L2
mechanism to have as few species as possible to enhance the com- can be lumped into species L, namely,
putational efficiency. On the other hand, considering the complex-
ity of low-temperature chemistry, the reduced mechanism needs to [L] = [L1 ] + [L2 ], (12)
be as detailed as possible to adequately capture the auto-ignition the above four reactions can be rewritten as
behavior under low-temperature conditions. Thus, reduction of the ˜
biodiesel reaction mechanism in the present study is performed to  k1
R1 : A 1 → L, ω
˜ 1 = k˜ 1 [A1 ]
achieve a balance between computational accuracy and efficiency.
˜
The target species for the PFA reduction in this study include  k2
R2 : A 2 → L, ω
˜ 2 = k˜ 2 [A2 ]
the fuel, air (N2 and O2 ), CO, HO2 , and C2 H2 to accurately deal
with the processes of fuel decomposition, oxidation, CO emission, ˜
 k3
R3 : L → B1 , ω
˜ 3 = k˜ 3 [L]
H2 –O2 chemical reactions, and soot emission, respectively. Molar
fractions of the three biodiesel surrogates, i.e., MD, MD9D, and n-  k4˜
heptane, in the mixture are prespecified to be 25%, 25%, and 50%,
R4 : L → B2 , ω
˜ 4 = k˜ 4 [L]. (13)
respectively. Despite the fact that the fuel surrogate composition In the above equation, rate coefficients of the production reac-
for reduction does not significantly change the resulting mecha- tions of isomers are simply k˜ 1 = k1 and k˜ 2 = k2 , indicating that the
nism, which also usually features good extensibility in predicting lumping operation does not affect the production reaction of ini-
the ignition characteristics of fuels with different initial surrogate tial isomer species. Rate coefficients for the consumption reactions
compositions [30,34], this surrogate composition for the present of isomers are adjusted by multiplying a factor proportional to the
biodiesel mechanism reduction is selected to be intermediate, giv- contribution of the isomer species in the lumped species composi-
ing equal moles of saturated and unsaturated long-chain methyl tion, i.e., the ratio of the isomer species composition to that of the
esters according to the scheme in Fig. 1. A large threshold value of lumped species:
45% is selected as the worst-case error tolerance for the PFA reduc-
[L 1 ] ˜ [L 2 ]
tion to obtain a relatively compact reaction mechanism, and the k˜ 3 = k3 , k4 = k4 . (14)
resulting mechanism is composed of 185 species and 735 reactions [L ] [L ]
after this reduction stage. Table 1 shows the isomer groups to be lumped in the biodiesel
mechanism. The isomers are lumped only if the deviation of igni-
tion delay from the original mechanism is less than 5% at all the
2.1.2. Isomer lumping validation conditions. After this reduction stage of isomer lump-
A further reduction of the mechanism is to consider the fact ing, the resulting mechanism consists of 156 species and 589 reac-
that the reaction mechanism of large hydrocarbons usually has tions. To evaluate the production rate of each isomer species dur-
many isomers, which have the same molecular weight and the ing the reaction computation, its concentration is needed and can
same function groups. Since isomers also have similar trans- be constructed from the lumped species concentration. The recon-
port properties, they can be grouped to obey a single transport struction is also a complex and unknown function of temperature,
equation. By grouping a few isomer species into one lumped pressure, and concentrations of all other species. In this study, the
species and eliminating the corresponding unimportant reactions, reconstruction is approximated using a simple strategy according
the number of variables to be tracked in combustion simulation to Eq. (14), which assumes each isomer species concentration to
can be effectively reduced. In this study, a simplified isomer lump- be the product of a constant factor and the lumped species con-
ing strategy is used to identify isomer groups, in which species centration [37].
have similar transport properties and participate in similar reac- Furthermore, to simulate the emission of NOx during biodiesel
tions [36]. Assume that two isomer species, for example, L1 and combustion, a chemistry describing NOx formation needs to be
L. Zhang, X. Ren and Z. Lan / Combustion and Flame 212 (2020) 377–387 381

present in the reaction mechanism. Since nitrogen is not contained


in the fuel, a simple chemistry including 4 species and 12 reactions
related to NOx is integrated with the mechanism to model the
thermal formation of NOx [38]. Finally, the final reduced mecha-
nism for biodiesel combustion is composed of 160 species and 601
reactions.

3. Results and discussion

3.1. Mechanism validation based on homogeneous applications

The present reduced biodiesel mechanism including 156 species


and 589 reactions is first validated by comparing with the origi-
nal complex mechanism in the calculations of homogeneous auto-
ignition and PSR combustion at various combustion conditions.
Figure 2 shows the predicted ignition delay as a function of ini-
tial temperature in a closed homogeneous reactor. The ignition de-
lay is defined as the time when the temperature of the mixture
increases by 400 K. The data are obtained at three pressures, rep-
resenting low-, moderate-, and high-pressure conditions, and three
equivalence ratios, representing lean, stochiometric, and rich com-
bustion conditions. Ignition delay curves predicted using the orig-
inal mechanism at the same operating conditions are also shown
in the same figure for the validation. The comparison indicates that
the ignition delay data predicted by the reduced mechanism agrees
closely with those by the original mechanism at all the selected
conditions, especially in the low-temperature (below 10 0 0 K) re-
gion which behaves strong NTC effect at high pressures. This close
agreement proves that the present reduced mechanism can ade-
quately reproduce the behavior of the original mechanism in pre-
dicting the low-temperature auto-ignition of biodiesel. Noted that
no adjustment, which is frequently used to improve the prediction
capability of reduced mechanisms, is made to the constants of the
Arrhenius reaction rate coefficients. To further examine the fidelity
of the reduced mechanism, predicted instantaneous mixture tem-
peratures during the auto-ignition process from four different ini-
tial temperatures are also compared as shown in Fig. 3. The pre-
dicted results are for a pressure of 1 atm and an equivalent ratio
of 1. It is seen that temperature histories predicted by the reduced
mechanism also agree closely with those by the original mecha-
nism for all the four initial temperatures.
Figure 4 shows the evolution of instantaneous molar fraction
of a few important species during the auto-ignition of stoichio-
metric biodiesel-air mixture in the closed homogeneous reactor at
constant pressure. Results are predicted using both the reduced
and original mechanisms with equal molar fractions of MD, MD9D,
and n-heptane as the fuel surrogate composition, different from
the original composition used for mechanism reduction. The ini-
tial temperature of the reactor is 1600 K, and the pressure is fixed
to be 1 atm. Comparison of the results by both mechanisms indi-
cates that the reduced mechanism agrees with the original mecha-
nism in predicting the mole fractions of all the selected important
species under the current operating condition. This agreement also
proves the extensibility of the present reduced mechanism to dif-
ferent compositions of fuel surrogate mixtures.
In addition to ignition, extinction is also a critical transient
phenomenon for premixed combustion, and prediction of ex-
tinction is an important quality of a reaction mechanism. In
the present study, the combustion in PSR at various operating
conditions is simulated for a further validation of the present
reduced mechanism. Figure 5 shows the comparison of calculated
temperature profiles as a function of residence time in PSR at Fig. 2. Predicted ignition delay for biodiesel-air mixture using both the reduced
and original mechanisms at pressures from 1 to 100 atm and equivalence ratios
different equivalence ratios and pressures. For each of the selected from 0.5 to 2.
conditions, the reduced mechanism shows a close agreement with
the original complex mechanism in the calculation of the tempera-
ture profile, especially in the branch above the curve turning point,
382 L. Zhang, X. Ren and Z. Lan / Combustion and Flame 212 (2020) 377–387

Fig. 3. Comparison of the predicted temperature histories from four different initial temperatures during auto-ignition by the reduced and original mechanisms.

Table 2
Operational parameters of the biodiesel spray combustion.

Fuel SME
Injection orifice diameter 90.8 μm
Injection pressure 150 MPa
Volumetric compositions of ambient gas Inert: O2 =0%, N2 =0.8971%, CO2 =0.0652%, H2 O=0.0377%
Reactive: O2 =0.15%, N2 =0.7515%, CO2 =0.0622%, H2 O=0.0363%
Gas density 22.8 kg m− 3
Gas temperature 900 K, 1000 K
Pressure 6.0 MPa, 6.7 MPa
Liquid fuel temperature 363 K
Injection fuel mass 22.7 mg
Discharge coefficient 0.94

which indicates the extinction of flame. This branch of the curve is Table 3
Composition of SME biodiesel used in the experiment [39].
a physical stable branch related to strong flames in realistic stable
combustion systems operating in steady state. Noticeable deviation Component Mole fraction
exists in the branch below the turning point for some operating Methyl Palmitate (C17 H34 O2 ) 11.0%
conditions. Since this lower branch behaves simultaneous decreas- Methyl Stearate (C19 H38 O2 ) 4.0%
ing temperature and increasing residence time (or Damköhler Methyl Oleate (C19 H36 O2 ) 25.0%
number) from the extinction state, it is composed of physically Methyl Linoleate (C19 H34 O2 ) 53.0%
Methyl Linolenate (C19 H32 O2 ) 7.0%
unrealistic states and cannot be obtained experimentally. Thus,
deviation in this branch does not impair the validity of the present
reduced mechanism in the simulation of PSR combustion.

The simulation is performed using an improved version of the


KIVA-3V R2 code, with a CHEMKIN solver embedded in the pro-
3.2. Mechanism validation based on three-dimensional biodiesel gram [40]. The present reduced biodiesel mechanism is then put in
spray combustion the KIVA/CHEMKIN program for the simulation of fuel spray com-
bustion. The computational domain is simplified to be a cylindri-
The present reduced biodiesel mechanism is also used to simu- cal geometry with a diameter of 10 cm and a height of 10 cm,
late a three-dimensional biodiesel spray combustion. The operating and the mesh is generated to have an average grid size of 1 mm.
conditions of the simulation are consistent with those of the op- The gas-phase flow field is resolved using the Reynolds-Averaged
tically accessible experiment of fuel spray combustion conducted Navier–Stokes equation, with the RNG k–ε model for turbulence
by Nerva et al. [39] in a constant-volume combustion chamber. modeling. The breakup of biodiesel spray is simulated using the
The chamber has a cubical geometry, with an identical length of KH-RT model, which describes the breakup of liquid fuel spray into
108 mm in each dimension. Soy methyl ester (SME) biodiesel is a two-stage process [41]. The soot emission is predicted using a
injected into the chamber through an injector mounted on the top two-step soot model [40], considering the generation and oxida-
center of the chamber. The injector has a single hole with a diame- tion of soot particles. According to the scheme shown in Fig. 1,
ter of 90 μm, and it is connected to a common rail, which fixes the mole fractions of the three biodiesel surrogates in the combus-
injection pressure to be 150 MPa. The main operational parameters tion simulation are determined by the composition of the actual
of the experiment are listed in Table 2. SME biodiesel used in the experiment as shown in Table 3. The
L. Zhang, X. Ren and Z. Lan / Combustion and Flame 212 (2020) 377–387 383

Fig. 4. Comparison of the molar fractions of a few important species during the constant-pressure auto-ignition from an initial temperature of 1600 K.

determination of surrogate mole fractions is to obey a simple rule fuel spray is defined as the axial distance from the injector to the
that the carbon content and the ester group content are in balance position behind which 95% of the total injected fuel exists, and
with those of the actual biodiesel components [16]. The physical the vapor penetration of fuel spray is the maximum axial distance
properties of biodiesel used in the spray modeling are calculated from the injector to the position where the mass fraction of fuel
using the methods proposed by Zhang and Kong [42]. vapor is greater than 0.001%. Spray penetration is a critical param-
To verify first the calculation of biodiesel physical properties, eter evaluating the combustion performance: Excessive penetration
the present models for biodiesel spray combustion are used to sim- may cause impingement of spray on cylinder wall, whereas defi-
ulate a non-reactive biodiesel spray in an inert (no oxygen) ambi- cient penetration can lead to insufficient atomization and mixing
ence as shown in Table 2. Figure 6 shows the development of liq- with ambient air. It is shown in Fig. 6 that liquid and vapor pen-
uid and vapor penetrations of the spray at two different ambient etrations coincide at a very early stage of the biodiesel spray and
temperatures. Also shown in the figure are the corresponding ex- start to separate shortly due to the vaporization of biodiesel drops.
perimental data [39] for comparisons. The liquid penetration of The liquid penetration fluctuates near a mean value, which is
384 L. Zhang, X. Ren and Z. Lan / Combustion and Flame 212 (2020) 377–387

Fig. 5. Temperature profiles of biodiesel-air mixture combustion in PSR as a function of residence time for a fixed inlet temperature of 300 K and various pressures and
equivalent ratios.

usually referred to as the liquid length. Beyond this axial distance, Table 4
Comparison of the predicted ignition delay and flame lift-off length with experi-
no liquid fuel drops exist in the cylinder, whereas the gas-phase
mental data at two initial temperatures.
fuel vapor continues to penetrate and mix with ambient air. The
calculated liquid and vapor penetrations of the biodiesel spray Temp. (K) Parameter Exp. Sim. RE
at the two ambient temperatures are in good agreements with 900 Ignition delay (ms) 0.709 0.656 −7.48
the experimental data, indicating accurate predictions of biodiesel Flame lift-off length (mm) 26.18 28.58 9.17
physical properties. 1000 Ignition delay (ms) 0.377 0.384 1.86
Flame lift-off length (mm) 17.27 21.91 26.87
Furthermore, the combustion of biodiesel spray in the re-
active ambience (including 15% oxygen) is simulated by solv-
ing the present reduced mechanism using the CHEMKIN solver. shows both experimental and simulated images of hydroxyl radical
Comparisons of predicted ignition delay and flame lift-off length (OH) distributions on the central cut plane of the cylindrical cham-
with the corresponding experimental data at two ambient gas tem- ber at 3 ms after the start of injection (ASI). Note that quantities
peratures are shown in Table 4. Both predicted and experimental used to indicate the OH distributions in the experimental and sim-
data show that higher ambient gas temperature leads to shorter ulated images are not identical. The experimental images are the
ignition delay and flame lift-off length, indicating enhanced fuel ensemble average of OH chemiluminescence images proportional
spray atomization and ignition due to the higher ambient temper- to the OH mass concentration, which is the quantity plotted in
ature. The predicted ignition delay at both ambient temperatures simulated images. Thus, quantities in the experimental and sim-
agrees with the experimental data, with the maximum relative er- ulated images are equivalent in indicating the instantaneous spa-
ror (RE) less than 10%. For both ambient temperatures, the present tial OH distribution. As shown by both experimental and simulated
combustion model slightly overpredicts the flame lift-off length. images, the OH concentrates in thin regions along the periphery of
The flame lift-off lengths at both 900 and 1000 K ambient tem- the flame, meaning high reaction rates in these regions. At the se-
peratures are also marked by the red dotted lines in Fig. 7, which lected instant, the OH distributions predicted using the biodiesel
L. Zhang, X. Ren and Z. Lan / Combustion and Flame 212 (2020) 377–387 385

Fig. 6. Histories of liquid and vapor penetrations of biodiesel spray in comparison with experimental data at two ambient temperatures.

combustion models with the present reduced mechanism approxi- biodiesel spray flame. The simulated soot distributions agree with
mately agree with those shown by the experimental images. those captured by the experiment in the overall geometry and ax-
In addition to the OH radical, experimental and simulated dis- ial position of the soot region and the relative intensity of concen-
tributions of soot particles on the central cut plane of the cylinder tration within the soot region.
at 3 ms ASI are also plotted for comparisons as shown in Fig. 8.
The experimental images plot the contours of soot volume fraction 4. Summary
measured using the planer laser-induced incandescence, whereas
in the simulated images, the distribution of soot particles is indi- In this study, a reduced biodiesel mechanism composed of
cated by the mass concentration. The soot emission is simulated 156 species and 589 reactions based on the mixture of three
using the two-step model with acetylene (C2 H2 ) selected as the surrogates (MD, MD9D, and n-heptane) is developed by reducing
precursor species for the generation of soot particles. Soot distri- the LLNL complex mechanism using the PFA reduction and isomer
butions shown in both experimental and simulated images indicate lumping. The selected three surrogates can adequately emulate
that soot particles concentrate in the central fuel-rich region of the the molecular structures of the five major biodiesel components
386 L. Zhang, X. Ren and Z. Lan / Combustion and Flame 212 (2020) 377–387

Fig. 7. Experimental and simulated distributions of OH at 3 ms ASI. (For interpretation of the references to color in this figure, the reader is referred to the web version of
this article.)

Fig. 8. Experimental and simulated distributions of soot particles at 3 ms ASI.

with different levels of unsaturation. Reduction of the mechanism cant financial support for this work that could have influenced its
is performed based on homogeneous auto-ignition and PSR com- outcome.
bustion on a variety of combustion states covering a large range of
operation conditions, especially in the region of low-temperature Acknowledgments
auto-ignition. The fidelity of the present reduced mechanism is
first examined by comparing with the original complex mechanism This work is supported by the National Natural Science Foun-
in the simulations of homogeneous auto-ignition and PSR combus- dation of China (Grant No. 51306209) and the China National Off-
tion under a variety of reaction conditions. Good agreements are shore Oil Corporation (No. YXKY-2018-ZY-09).
obtained in the predicted ignition delay, history of mixture tem-
perature, and history of major species concentration during auto- Supplementary materials
ignition. In the simulation of PSR combustion, close agreement
is also obtained in the upper stable branch of the temperature Supplementary material associated with this article can be
profile. The present reduced mechanism, further integrated with a found, in the online version, at doi:10.1016/j.combustflame.2019.11.
simple NOx chemistry and a two-step soot model, is implemented 002.
into the KIVA/CHEMKIN program to simulate three-dimensional
biodiesel spray combustion and emission under CI engine operat- References
ing conditions. Agreements with experimental data are obtained
in the liquid and vapor penetration lengths, flame lift-off lengths, [1] J. Chen, Petascale direct numerical simulation of turbulent combustion—fun-
and the spatial distributions of OH and soot. Consequently, the damental insights towards predictive models, Proc. Combust. Inst. 33 (2011)
99–123.
applicability of the present reduced biodiesel mechanism in the [2] R.D. Reitz, Directions in internal combustion engine research, Combust. Flame
numerical study of biodiesel combustion and emission in conven- 160 (2013) 1–8.
tional CI engine is demonstrated. Considering its accuracy in the [3] W. Yuan, A.C. Hansen, Z. Tan, Modeling of NOx emissions of biodiesel fuels,
2005 ASAE Annual International Meeting (2005) paper 056116.
prediction of low-temperature auto-ignition, the present reduced
[4] J. Sun, J.A. Caton, T.J. Jacobs, Oxides of nitrogen emissions from biodiesel-fu-
mechanism also has a great potential to be used in the future elled diesel engines, Prog. Energy Combust. 36 (2010) 677–695.
study of engines operating on advanced combustion technologies. [5] E. Rajasekar, A. Murugesan, R. Subramanian, N. Nedunchezhian, Review of NOx
reduction technologies in CI engines fuelled with oxygenated biomass fuels,
Renew. Sust. Energy Rev. 14 (2010) 2113–2121.
Declaration of Competing Interest [6] A.K. Agarwal, Biofuels (alcohols and biodiesel) applications as fuels for internal
combustion engines, Prog. Energy Combust. 33 (2007) 233–271.
[7] C. He, Y. Ge, J. Tan, K. You, X. Han, J. Wang, Characteristics of polycyclic
We wish to confirm that there are no known conflicts of inter- aromatic hydrocarbons emissions of diesel engine fueled with biodiesel and
est associated with this publication and there has been no signifi- diesel, Fuel 89 (2010) 2040–2046.
L. Zhang, X. Ren and Z. Lan / Combustion and Flame 212 (2020) 377–387 387

[8] E.M. Fisher, W.J. Pitz, H.J. Curran, C.K. Westbrook, Detailed chemical kinetic [26] T. Lu, M. Plomer, Z. Luo, S.M. Sarathy, W.J. Pitz, Directed relation graph with
mechanisms for combustion of oxygenated fuels, Proc. Combust. Inst. 28 expert knowledge for skeletal mechanism reduction, 7th U.S. National Com-
(20 0 0) 1579–1586. bustion Meeting (2011) paper 1F01.
[9] S. Gaïl, M.J. Thomson, S.M. Sarathy, S.A. Syed, P. Dagaut, P. Diévart, A.J. March- [27] W. Sun, Z. Chen, X. Gou, Y. Ju, A path flux analysis method for the reduction of
ese, F.L. Dryer, A wide-ranging kinetic modeling study of methyl butanoate detailed chemical kinetic mechanisms, Combust. Flame 157 (2010) 1298–1307.
combustion, Proc. Combust. Inst. 31 (2007) 305–311. [28] G. Li, H. Rabitz, J. Tóth, A general analysis of exact nonlinear lumping in chem-
[10] A. Farooq, D.F. Davidson, R.K. Hanson, L.K. Huynh, A. Violi, An experimental ical kinetics, Chem. Eng. Sci. 49 (1994) 343–361.
and computational study of methyl ester decomposition pathways using shock [29] R. Fournet, V. Warth, P.A. Glaude, F. Battin-Leclerc, G. Scacchi, G.M. Côme, Au-
tubes, Proc. Combust. Inst. 32 (2009) 247–253. tomatic reduction of detailed mechanisms of combustion of alkanes by chem-
[11] P. Dagaut, S. Gaïl, Chemical kinetic study of the effect of a biofuel additive on ical lumping, Int. J. Chem. Kinet. 32 (2015) 36–51.
Jet-A1 combustion, J. Phys. Chem. A 111 (2007) 3992–4000. [30] Z. Luo, T. Lu, M.J. Maciaszek, S. Som, D.E. Longman, A reduced mechanism
[12] P. Dagaut, S. Gaïl, M. Sahasrabudhe, Rapeseed oil methyl ester oxidation over for high-temperature oxidation of biodiesel surrogates, Energy Fuel 24 (2010)
extended ranges of pressure, temperature, and equivalence ratio: experimental 6283–6293.
and modeling kinetic study, Proc. Combust. Inst. 31 (2007) 2955–2961. [31] J.L. Brakora, Y. Ra, R.D. Reitz, J. McFarlane, C.S. Daw, Development and valida-
[13] Y. Zhang, Y. Yang, A.L. Boehman, Premixed ignition behavior of C9 fatty acid tion of a reduced reaction mechanism for biodiesel-fueled engine simulations,
esters: a motored engine study, Combust. Flame 156 (2009) 1202–1213. SAE Int. J. Fuel Lubr. 1 (2008) 675–702.
[14] V.I. Golovitchev, J. Yang, Construction of combustion models for rapeseed [32] J.L. Brakora, R.D. Reitz, Investigation of NOx predictions from biodiesel fueled
methyl ester bio-diesel fuel for internal combustion engine applications, HCCI engine simulations using a reduced kinetic mechanism, SAE Technical
Biotechnol. Adv. 27 (2009) 641–655. Paper, 2010-01-0577, 2010.
[15] H.K. Ng, S. Gan, J.-H. Ng, K.M. Pang, Development and validation of a reduced [33] J.L. Brakora, Y. Ra, R.D. Reitz, Combustion model for biodiesel-fueled engine
combined biodiesel-diesel reaction mechanism, Fuel 104 (2013) 620–634. simulations using realistic chemistry and physical properties, SAE Int. J. Eng. 4
[16] O. Herbinet, W.J. Pitz, C.K. Westbrook, Detailed chemical kinetic mechanism for (2011) 931–947.
the oxidation of biodiesel fuels blend surrogate, Combust. Flame 157 (2010) [34] Z. Luo, M. Plomer, T. Lu, S. Som, D.E. Longman, A reduced mechanism for
893–908. biodiesel surrogates with low temperature chemistry for compression ignition
[17] P. Gokulakrishnan, A.D. Lawrence, P.J. McLellan, E.W. Grandmaison, A function- engine applications, Combust. Theory Model. 16 (2012) 369–385.
al-PCA approach for analyzing and reducing complex chemical mechanisms, [35] Z. Luo, M. Plomer, T. Lu, S. Som, D.E. Longman, S.M. Sarathy, W.J. Pitz, A re-
Comput. Chem. Eng. 30 (2006) 1093–1101. duced mechanism for biodiesel surrogates for compression ignition engine ap-
[18] S. Vajda, P. Valko, T. Turányi, Principal component analysis of kinetic models, plications, Fuel 99 (2012) 143–153.
Int. J. Chem. Kinet. 17 (2010) 55–81. [36] P. Pepiot-Desjardins, H. Pitsch, An automatic chemical lumping method for the
[19] Y. Liu, Y. Wu, Y. Gao, T. Lu, A linearized error propagation model based on Jaco- reduction of large chemical kinetic mechanisms, Combust. Theory Model. 12
bian analysis for skeletal mechanism reduction, 9th U.S. National Combustion (2008) 1809-1118.
Meeting (2015) paper RK38. [37] T. Lu, C.K. Law, Strategies for mechanism reduction for large hydrocarbons:
[20] Y. Zhang, M.A. Dubé, D.D. McLean, M. Kates, Biodiesel production from waste n-heptane, Combust. Flame 154 (2008) 153–163.
cooking oil: 2. Economic assessment and sensitivity analysis, Bioresour. Tech- [38] S.-C. Kong, Y. Sun, R.D. Reitz, Modeling diesel spray flame liftoff, sooting ten-
nol. 90 (2003) 229–240. dency, and NOx emissions using detailed chemistry with phenomenological
[21] A.S. Tomlin, M.J. Pilling, T. Turányi, J.H. Merkin, J. Brindley, Mechanism reduc- soot model, J. Eng. Gas Turb. Power 129 (2007) 245–251.
tion for the oscillatory oxidation of hydrogen: sensitivity and quasi-steady-s- [39] J.-G. Nerva, C.L. Genzale, S. Kook, J.M. García-Oliver, L.M. Pickett, Fundamental
tate analyses, Combust. Flame 91 (1992) 107–130. spray and combustion measurements of soy methyl-ester biodiesel, Int. J. Eng.
[22] T. Lu, C.K. Law, A directed relation graph method for mechanism reduction, Res. 14 (2013) 373–390.
Proc. Combust. Inst. 30 (2005) 1333–1341. [40] S.-C. Kong, Y. Sun, R.D. Reitz, Modeling diesel spray flame liftoff, sooting ten-
[23] T. Lu, C.K. Law, On the applicability of directed relation graphs to the reduction dency, and NOx emissions using detailed chemistry with phenomenological
of reaction mechanisms, Combust. Flame 146 (2006) 472–483. soot model, J. Eng. Gas Turb. Power 128 (2006) 245–251.
[24] J. An, Y. Jiang, Differences between direct relation graph and error-propaga- [41] M.A. Patterson, R.D. Reitz, Modeling the effects of fuel spray characteristics on
tion-based reduction methods for large hydrocarbons, Proc. Eng. 62 (2013) diesel engine combustion and emissions, SAE Technical Paper 980131, 1998.
342–349. [42] L. Zhang, S.-C. Kong, Vaporization modeling of petroleum–biofuel drops using
[25] P. Pepiot-Desjardins, H. Pitsch, An efficient error-propagation-based reduction a hybrid multi-component approach, Combust. Flame 157 (2010) 2165–2174.
method for large chemical kinetic mechanisms, Combust. Flame 154 (2008)
67–81.

You might also like