Subscriber access provided by University at Buffalo Libraries
Combustion
A simplified mechanistic model of three-component surrogate fuels for RP-3 aviation kerosene Yunpeng LIU, Yuchen LIU, Dengbing CHEN, Wen FANG, Jinghua Li, and Yingwen Yan Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.8b02094 • Publication Date (Web): 01 Aug 2018 Downloaded from http://pubs.acs.org on August 6, 2018
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
1
A simplified mechanistic model of three-component surrogate fuels for
2
RP-3 aviation kerosene
3
Yunpeng Liu, Yuchen Liu, Dengbing Chen, Wen Fang, Jinghua Li, Yingwen Yan
4
College of Energy and Power Engineering, Nanjing University of Aeronautics and Astronautics,
5
Nanjing, 210016, China
6
ABSTRACT
7
In this study, a detailed chemical reaction kinetic model of a three-component surrogate fuel (73%
8
n-dodecane, 14.7% 1,3,5–3-trimethylcyclohexane, and 12.3% n-propylbenzene) for RP-3 aviation
9
kerosene was simplified, and a simplified mechanism for this fuel was obtained and validated. The
10
detailed chemical reaction kinetic model included 257 components and 874 elementary reactions.
11
In the first step, a mechanism consisting of 109 components and 423 elementary reactions was
12
constructed from the detailed model using a directed relation graph (DRG). The second step,
13
based on the first, was to construct an 84-component, 271-elementary-reaction mechanism using a
14
DRG based on error propagation (DRGEP) and computational singular perturbation (CSP). In the
15
third step, path analysis was applied to analyze the combustion paths under atmospheric pressure
16
and high-temperature conditions; the results were compared with those of the detailed mechanism
17
and the simplified mechanism of the second step to remove unimportant reaction paths or to
18
supplement important paths that were reduced in the first two steps. In the final step, a simplified
19
mechanism of the three-component surrogate fuel suitable for high temperature and atmospheric
20
pressure was obtained, which involved 59 components and 158 elementary reactions. Test data of
21
the ignition delay time and the laminar flame velocity of RP-3 kerosene were used to verify the
22
simplified mechanism for the three-component surrogate fuel. The results showed that the
23
numerical results for the proposed simplified mechanism were consistent with the test data. Finally,
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
24
to verify the engineering practicality of the simplified mechanism proposed in this study, taking a
25
Bunsen burner as the physical model, a premixed pre-evaporation combustion flame was
26
numerically simulated by using the simplified mechanism for the three-component surrogate fuel.
27
The numerical calculation results of the simplified mechanism were consistent with the test data,
28
and the computation time was within the acceptable range of the engineering application.
29
Therefore, the simplified mechanism for the three-component surrogate fuel can be used for
30
numerical simulation of engineering combustion with kerosene fuel.
31
Keywords: Three-component surrogate fuels; RP3-Aviation kerosene; Ignition delay; Laminar
32
flame speed; Simplified mechanism; Bunsen burner;
33
1. Introduction
34
Research on the combustion characteristics, combustion rate, flame structure, and stable
35
combustion range of aviation kerosene has been key to the development of aeroengine combustors
36
(1), so it is essential to study the combustion of aviation kerosene in depth. Aviation kerosene
37
combustion in aeroengine combustors is a complex and turbulent combustion process, which is
38
jointly determined by turbulent mixing and chemical kinetics. Therefore, to accurately simulate
39
flame combustion in an aeroengine combustor, besides studying the turbulent combustion model,
40
research on a chemical kinetic model of aviation kerosene is necessary. A combination of
41
computational combustion and a chemical kinetic model can accurately predict the turbulent
42
combustion characteristics of aeroengine combustor, and provide technical support for the design
43
and optimization of aeroengine combustors. Since RP-3 aviation kerosene is the main fuel for
44
combustion chambers of aeroengines in China, the multi-step chemical reaction kinetic model
45
directly establishes the accuracy and reliability of numerical simulations of turbulent flow in
ACS Paragon Plus Environment
Page 2 of 32
Page 3 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
46
combustion chambers. However, the combustion of macromolecular hydrocarbon in RP-3 aviation
47
kerosene with oxygen is extremely complex. This involves two aspects (2): 1) aviation fuels are
48
mixtures of hundreds of components and it is therefore impractical to provide a detailed chemical
49
mechanism model to study and test for each component. 2) Generally, a detailed chemical kinetic
50
mechanism for a surrogate fuel contains hundreds of components and thousands of elementary
51
reactions. For such a complex reaction system, a detailed chemical kinetic model is too complex
52
to use to calculate the three dimensional flow field at the current computer level. Therefore,
53
numerical simulation of turbulent combustion in aeroengine combustors requires the detailed
54
chemical kinetic mechanism to be simplified on the premise of ensuring the accuracy of the
55
numerical simulation of the chemical reaction mechanism. The aim is to ensure the accuracy of
56
the numerical simulation while reducing the computation time, which requires an increased
57
calculation efficiency. Simplifying the kinetic model of the aviation kerosene combustion reaction
58
is important for numerical simulation of combustion in aeroengine combustors at this stage. This
59
can improve the computational efficiency and reduce the convergent stiffness for the CFD
60
simulation of aviation kerosene combustion, to introduce the chemical reaction mechanism to the
61
practical engineering problems of three-dimensional turbulent combustion and perform more
62
accurate numerical simulations. While the detail reaction mechanisms could well reflect the
63
complex reaction process, the mechanisms are hard to be applied to the engineering applications,
64
for example, combustion numerical simulation of the aero engine combustor. Thus, appropriate
65
methods and verifications should be employed to simplify the detail mechanisms. As conducted in
66
the paper, the simplified mechanism could well be introduced to the engineering calculation.
67
This study is mainly concerned with the engineering application of the chemical reaction
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
68
mechanism of aviation kerosene. So it is different from the fundamental research about the
69
surrogate fuel of aviation kerosene or the detail combustion reaction mechanism. The detail
70
mechanism of three-component surrogate fuel of aviation kerosene is selected and simplified in
71
the paper. The present study is devoted to investigating the simplification of the detail mechanism
72
for engineering application, and the unique simplification scheme makes this paper especially
73
concern the ignition delay time in the simplification process. A set of verification ensures the
74
reasonability and accuracy of the simplified mechanism in this paper.
75
2. State-of-the-art and development trends
76
Because the composition of aviation kerosene is very complex, many researchers investigate
77
surrogate fuel of aviation kerosene. For example, Yi W et al (3) investigated the laminar flame
78
speed measurement of kerosene and its surrogate fuel (n-decane, n-propyl benzene, and propyl
79
cyclohexane) and results confirm the reasonableness of surrogate fuel. To date, research on the
80
reaction mechanisms of chemical combustion of aviation kerosene includes the following aspects.
81
1) Surrogate fuel modes developed from monocomponent surrogate fuel modes to
82
double-component surrogate fuel modes and subsequently to multi-component surrogate fuel
83
modes. 2) Surrogate fuel modes developed from single chain hydrocarbon modes to
84
multi-component surrogate fuel modes, which contain aliphatic and aromatic hydrocarbons. 3)
85
The intermediate components increased from dozens to hundreds and elementary reactions
86
increased from hundreds to thousands.
87
Kundu et al. (4) studied the surrogate fuel C12H23, and a simplified mechanism involving 16
88
components and 23 reactions was proposed to simulate the combustion of Jet A fuel. By
89
comparing the temperature, the concentration of NOx, and other products calculated using the
ACS Paragon Plus Environment
Page 4 of 32
Page 5 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
90
simplified mechanism with the experimental data, it was shown that the simplified mechanism can
91
be applied to the numerical simulation of Jet A fuel combustion. Many other researchers (5-7) also
92
investigated the skeletal mechanisms and mechanism reduction for aviation kerosene surrogate
93
fuel. Moreover, using n-dodecane as a monocomponent surrogate fuel for RP-3 kerosene, a
94
reaction model involving 11 components and 17 elementary reactions was proposed for
95
n-dodecane by Wang et al.(8). The cracking of kerosene, the formation mechanism of carbon
96
smoke, and the principle of oxidation were also explored. Patterson et al. (9) developed a
97
surrogate for Jet A-1 kerosene comprising 89% n-decane and 11% methylbenzene to conduct
98
experimental and mechanistic research. Honnet et al. (10) developed a surrogate for Jet A-1
99
kerosene consisting of 80% n-decane and 20% 1,2,4–3-methylbenzene. The proposed mechanism
100
involved 118 components and 914 elementary reactions. Vovelle et al. (11) developed a surrogate
101
for Jet A-1 kerosene consisting of 90% n-decane and 10% methylbenzene. They proposed a
102
simplified kinetic model consisting of 39 components and 207 reactions. Subsequently,
103
researchers developed more complex multi-component dynamic models. Prior to this, the main
104
principle involved adding naphthenic hydrocarbons to surrogate fuels consisting of saturated
105
alkanes and aromatic hydrocarbons. For example, a detailed model was proposed involving a
106
three-component surrogate fuel for Jet A-1 comprising 74 mol% n-decane, 11 mol%
107
1,3,5-trimethylcyclohexane, and 15 mol% n-propylbenzene (12). This model consisted of 209
108
components and 1,673 reactions. Chitral kumar V (13) introduced a detailed chemical kinetic
109
mechanism for surrogate aviation fuel. First, using an available mechanism, the reaction
110
mechanism was constructed by coupling the mechanisms of components of surrogate fuel, then
111
adding the reaction mechanism for NOx. Second, a reaction mechanism for the three-component
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
112
surrogate fuel was proposed. Finally, ideal numerical results were obtained after refining the main
113
mechanism. Using 71% n-decane, 13% trimethylcyclohexane, and 16% ethylbenzene as a
114
three-component surrogate fuel for kerosene, a detailed three-component surrogate fuel
115
mechanism model including 109 components and 946 elementary reactions was proposed by Xiao
116
(14). By using a quasi-steady-state hypothesis simplification method, a simplified reaction model
117
of an overall reaction with 22 components and 18 elementary reactions was obtained. Additionally,
118
the ignition delay time and the laminar flame speed were calculated by using the simplified
119
mechanism. The calculation results were consistent with the experimental data for kerosene
120
ignition. A mixture composed of 73 mol% n-dodecane, 14.7 mol% 1,3,5-trimethylcyclohexane,
121
and 12.3 mol% n-propylbenzene was applied as a surrogate for RP-3 kerosene (15). The
122
simplified mechanism involved 138 components and 530 reactions. A mixture consisting of 40
123
mol% n-decane, 42 mol% n-dodecane, 13 mol% ethylcyclohexane, and 5 mol% p-xylene was
124
proposed by Yu (2) as a four-component surrogate fuel for RP-3 kerosene. This was developed
125
into a detailed mechanism containing 168 components and 1,089 elementary reactions. The results
126
showed that the flame propagation speed obtained by this detailed mechanism was consistent with
127
the measured RP-3 kerosene flame propagation speed. A four-component surrogate fuel for JP-8
128
aviation kerosene proposed by Montgomery (16) consisted of 32.6 mol% decane, 34.7 mol%
129
n-dodecane, and 16.7 mol% methylcyclohexane. The detailed chemical reaction mechanism of the
130
four-component surrogate fuel involved 164 components and 1,162 elementary reactions. The
131
mechanism of the detailed reaction was simplified to obtain two simplified reaction mechanisms
132
composed of 15 and 20 components. The ignition delay times of the two simplified reaction
133
mechanisms were consistent with the test data. From the above studies, it can be concluded that
ACS Paragon Plus Environment
Page 6 of 32
Page 7 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
134
surrogate fuels for aviation kerosene consisted of three major categories: alkanes (straight-chain
135
and branched paraffin), cycloalkanes, and aromatics. The main surrogate fuels for aviation
136
kerosene were n-decane, n-dodecane, n-propylbenzene, n-butyl benzene, 1,2,4–3-methylbenzene,
137
methylbenzene,
138
hydrocarbons and other trace elements were not generally considered by researchers.
139
3. Construction of simplified mechanism for three-component surrogate fuel
140
In
this
1,3,5-trimethylcyclohexane,
study,
a
mixture
composed
and
of
methylcyclohexane,
73%
n-dodecane
while
unsaturated
(n-C12H26),
14.7%
141
1,3,5–3-methylcyclohexane (C9H18), and 12.3% n-propylbenzene (PHC3H7) was selected as a
142
surrogate fuel for RP-3 aviation kerosene (15). The suitability of the proportion of each
143
component was verified and a semi-detailed mechanism involving 257 components and 874
144
elementary reactions was proposed (15). Additionally, by utilizing test data of aviation kerosene
145
ignition delay times published by Tang (17) to verify the ignition delay time of this detailed
146
mechanism, the results showed that the model can accurately describe the high-temperature
147
ignition characteristics of RP-3 kerosene.
148
Therefore, the detailed mechanism of the three-component surrogate fuel used in this study
149
was the semi-detailed mechanism (257 components and 874 elementary reactions) of RP-3
150
aviation kerosene that was previously verified (15, 18). However, as the number of components
151
and elementary reactions remained high, this could not be applied to the numerical simulation for
152
combustion of RP-3 aviation kerosene in an aeroengine combustor. Further simplification was
153
therefore necessary. A three-step simplification scheme was adopted in this study. In the first step,
154
a directed relation graph model (DRG) (19) was used to simplify the semi-detailed chemical
155
kinetic model, and skeleton simplification mechanism Ⅰ was obtained. In the second step, a
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
156
directed relation graph based on error propagation model (DRGEP) (20) was used for further
157
simplification on the basis of simplification mechanism Ⅰ. Additionally, computational singular
158
perturbation model (CSP) (21) was used to remove redundant reactions to obtain simplification
159
mechanism. In the third step, a path analysis method in CHEMKIN-PRO software (22) was used
160
to analyze the combustion paths at atmospheric pressure and high temperature. By comparing the
161
detailed mechanism and simplification mechanism Ⅰ, the unimportant reaction paths (elementary
162
reactions with low chemical reaction rates) were removed, and the important paths removed in the
163
first and second steps were reinstated. Finally, a simplified mechanism for a three-component
164
surrogate fuel suitable for combustion at atmospheric pressure and high temperature was obtained,
165
which involved 59 components and 158 elementary reactions. Experimental data of RP-3 aviation
166
kerosene ignition delay times published by Changhua Zhang (23) and test data of RP-3 kerosene
167
laminar flame speeds published by Yu (2) were used to verify the simplified mechanism proposed
168
in this study.
169
The aim of the mechanistic study was to obtain a detailed understanding of each elementary
170
reaction, to clarify the paths and the order of the reactions, and finally to clarify the rate and trend
171
of the actual combustion reaction. Clarifying the reaction paths of the mechanism is essential for
172
controlling the reaction and predicting the external macroscopic phenomena. To study the reaction
173
mechanism of RP-3 kerosene combustion, it is necessary to understand how specific combustion
174
paths of aviation kerosene develop. Some studies have been published on the analysis of paths.
175
Generally, the reaction rate analysis method in CHEMKIN-PRO is used to analyze the main
176
reaction paths of fuel molecules at high and low temperatures during combustion. At low
177
temperatures, fuel is consumed by dehydrogenation, oxygenation, and isomerization of large
ACS Paragon Plus Environment
Page 8 of 32
Page 9 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
178
molecules and cracking reactions of ketones. At high temperatures, intermediate products
179
generated at low temperatures are cracked into smaller molecules such as CH4. This may occur via
180
cleavage of C-O, C-C, and β-bonds. The first and second steps of the simplification process have
181
already been described (24). Based on the first two steps, the third step involving simplification
182
and optimization using path analysis was performed in this study.
183
In the third step, path analysis was adopted for quantitative and qualitative analysis of
184
three-component surrogate fuel cracking using the simplified mechanism obtained in the second
185
step and the detailed mechanism separately under the working conditions, including equivalence
186
ratio Ø = 1, P = 1 atm, T = 1,150 K, 1,600 K, and 2,000 K. Additionally, paths that are not
187
applicable at atmospheric pressure, those with low reaction rates, and unimportant paths were
188
removed, while some important paths that were removed by the simplification program were
189
reinstated. The simplified mechanism obtained involved 59 components and 158 elementary
190
reactions (See Appendix A The simplified mechanism of three-component surrogate fuels for RP-3
191
aviation kerosene for details).
192
Figure 1 shows the probability of path analysis diagrams of fuel cracking during ignition
193
delay times at three different initial temperatures for the simplified mechanism proposed in this
194
study. As the ignition process of fuel plays an important role in mechanistic studies, and more than
195
half of the ignition delay time is occupied by macromolecular cracking, this study focuses on the
196
path analysis of macromolecules of three-component surrogate fuel during the ignition delay time.
197
In this study, the mechanism was simplified, while its cracking process was explored. In the first
198
step of Figure 1(a), s0C12H26 was cracked to nC3H7, s10C9H19, s12C5H11, and s13C7H15
199
through 109th and 110th reactions. The three intermediate components finally formed nC3H7
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
200
through three paths via a series of reactions. Finally, nC3H7 was dehydrogenated to form C3H6
201
through the 97th reaction. Meanwhile, small molecules C2H4 and CH3 were formed through the
202
78th elementary reaction. As shown in Figure 1(b), 1,3,5–3-methylcyclohexane (s1C9H18) was
203
gradually cracked and oxidized via two paths into small elementary molecular groups. First,
204
s1C9H18 formed s58C9H18 and s59C9H18 via the 118th and 119th isomerization reactions,
205
respectively. The two components were then dehydrogenated with CH3 to form s164C9H17 and
206
s167C9H17 via the 121st and 122nd reactions. Next, nC3H7, iC3H7, and s482C6H10 were formed
207
via the 123rd and 124th cracking reactions. Then, s482C6H10 reacted with H, CH3, and OH to
208
crack into C4H6 and C2H3 via the 125th, 126th, and 127th reactions. Finally, nC3H7, iC3H7, and
209
C4H6 cracked into small molecules. Figure 1(c) shows that paths of n-propylbenzene PHC3H7
210
were significantly different from those of alkanes. The unique step of the high-temperature
211
combustion of aromatic hydrocarbons is a ring-opening reaction; the first mode of ring opening
212
was cleavage of a carbon-carbon bond on a single ring to generate a diradical, which cracked at
213
high temperature to form isoolefin or other small molecular groups. The second mode was the
214
reaction of alkyl on a single ring to form a mono-ring alkyl by hydrogen extraction, followed by
215
ring opening by cleavage of a β-bond. Reactions of the species after ring opening were consistent
216
with alkane cracking. As regards PHC3H7, in the first step, BPHC3H6 and CPHC3H6 were
217
formed by dehydrogenation via the 128th, 129th, 130th, 131st, 132nd, 133rd, 134th, 135th, and 136th
218
elementary reactions. Simultaneously, cracking was performed to form nC3H7, C6H6, PHCH2,
219
and APHC2H4 via the 137th, 138th, and 139th elementary reactions. C6H5 was then generated
220
through a series of cracking and dehydrogenation reactions, after which a portion of C6H5
221
cracked into C5H5 via the 105th reaction, while the remainder was oxidized to C6H5O via the
ACS Paragon Plus Environment
Page 10 of 32
Page 11 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
222
104th and 106th reactions. Finally, small molecules such as C3H3 and C2H2 were formed by
223
cleavage of C-O, C-C, and β-bonds, cracking reactions, partial dehydrogenation, and
224
isomerization of C5H5.
225
In this study, The Bunsen burner employed in this paper works at atmosphere environment.
226
The Bunsen burner is feed with gaseous aero kerosene, and the fuel is premixed with the air. So
227
the factors affecting the combustion are relatively simply, for example, there is no atomization,
228
evaporation and mixing in the combustion process. For the simplification process, the premixed
229
model is also utilized in CHEMKIN to explore the important chemical reactions. Then the
230
numerical simulation is conducted under the premixed and pre-vaporized circumstance. Based on
231
this, the simplified mechanism is prone to be compared and verified by the experimental data. In
232
addition to the qualitative analysis of the simplified mechanism for three-component surrogate
233
fuel during the ignition delay time under conditions including equivalence ratio Ø = 1,
234
atmospheric pressure P = 1 atm, and temperature T = 1,150 K, 1,600 K, and 2,000 K, quantitative
235
analysis was also performed. The rates of all reactions at all micro-moments during the ignition
236
delay time at three different ignition temperatures were statistically analyzed to compare the
237
effects of different initial temperatures on the chemical reaction paths. The main principle was to
238
quantitatively calculate the proportion of each small path (each arrow) in the current consumption
239
process (the tail of the arrow represents the consumed material) through chemical reaction rates
240
during the ignition delay time. Combined with Figure 1(a, b, and c), it can be seen that the
241
proportion of dehydrogenation and isomerization paths of macromolecules and most intermediate
242
components decreased as the temperature increased, and the proportion of cracking paths
243
increased. The proportion of cracking paths of some intermediate components was reduced, but
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
244
oxidation rather than dehydrogenation increased in proportion. Therefore, it can be concluded that
245
at relatively low temperatures, due to the energy being relatively low, C-H bonds were easily
246
broken, while C-C, C-O, and β-bonds were not, which increased the proportion of
247
dehydrogenation and isomerization paths at relatively low temperatures. At relatively high
248
temperatures, C-H, C-C, C-O, and β-bonds break, but C-C, C-O, and β-bonds react faster than
249
C-H. Thus, at high temperature, the proportion of cracking paths was larger than that of
250
dehydrogenation paths.
251 252
1(a) n-dodecane
ACS Paragon Plus Environment
Page 12 of 32
Page 13 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
253 254
1(b) 1,3,5-trimethylcyclohexane
255 256
1(c) n-propylbenzene
257
Figure 1. Path analysis of simplified mechanism of three-component surrogate fuel at three initial
258
temperatures during the ignition delay time
259
4. Sensitivity analysis
260
Sensitivity analysis is an ideal method available in CHEMKIN for judging and screening
261
important reactions in chemical reaction kinetic models. By studying the effects of small
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
262
disturbances on the whole system, the elementary reactions in the model that play important roles
263
in the combustion process can be found. Temperature sensitivity analyses mainly study the effects
264
of temperature disturbance on reaction rate of the simplified mechanism. That is, sensitivity
265
analyses investigate the internal chemical reaction mechanisms of all species. From temperature
266
sensitivity analysis, primary elementary reaction (controlled step) could be explored. Currently,
267
the Jacobi matrix method is used to analyze the sensitivity of mechanistic models. In this study,
268
sensitivity analysis of the steady adiabatic flame temperature of the simplified mechanism was
269
performed at atmospheric pressure and different initial temperatures.
270
Figure 2 shows the sensitivity analysis of the simplified mechanism proposed in this study at
271
atmospheric pressure and different initial temperatures. Each part of Figure 2 lists the 10 reactions
272
with the highest temperature sensitivity coefficients. The simulation conditions included P = 1 atm,
273
Ø = 0.1, T = 1,400 K, 1,600 K, and 2,000 K. In Figure 2(a)~(d), it can be seen that as the initial
274
temperature increased, the overall temperature sensitivity coefficient decreased. This is because
275
most elementary reactions can reach the required activation energy at high temperatures, so they
276
become increasingly insensitive to temperature as the temperature increases. Furthermore, the
277
elementary reaction H + O2 OH + O can be seen as the main OH formation reaction, wherein
278
the OH group is considered to symbolize ignition and combustion; this elementary reaction was
279
sensitive to temperature at different initial temperatures. As the initial temperature increased,
280
compared to other elementary reactions with positive sensitivity coefficients, the temperature
281
sensitivity of the reaction increased (although not significantly), indicating that this reaction
282
promoted combustion and temperature increase. As regards reactions with negative sensitivity
283
coefficients, it can be seen that the sensitivity of reaction CHO + O2 = CO + HO2 was relatively
ACS Paragon Plus Environment
Page 14 of 32
Page 15 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
284
larger, which indicated that this reaction inhibited ignition, combustion, and temperature increase.
285
Additionally, the cracking and dehydrogenation reactions of macromolecules such as s0C12H26
286
and PHC3H7 were among the first ten reactions at 1,200 K, but as the initial temperature
287
increased, the ten reactions with the highest temperature sensitivities were all violent reactions of
288
small molecular groups.
289 290
2(a) 1,200 K
2(b) 1,400 K
2(c) 1,600 K
2(d) 2,000 K
Figure 2. Sensitivity analysis of simplified mechanism of three-component surrogate fuel 5. Verification of multi-component simplification mechanistic model
291
A study on fuel combustion requires reliable and detailed kinetic analysis of the fuel ignition
292
process, flame propagation characteristics, flame stability, and concentration changes of
293
intermediate components. The ignition and flame propagation characteristics, adiabatic flame
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 16 of 32
294
temperature, and concentrations of some intermediate groups were selected to verify the proposed
295
simplification mechanism for three-component surrogate fuel for RP-3 aviation kerosene. The
296
most important parameter of the ignition characteristics was the ignition delay time, and that of the
297
flame propagation characteristics was the laminar flame speed. In this paper, different mechanisms
298
schemes are listed in Table.1.
299
Table.1 Mechanisms schemes of different references
Case A B C D E 300
Author XU Jia-Qi XU Jia-Qi Kundu P K M. I. Strelkova Dai Chao
reference (15) (15) (4) (25) (26)
Species and reactions 257 species and 874 reactions 138 species and 530 reactions 16 species and 23 reactions 25 species and 38 reactions 36 species and 62 reactions
5.1 Ignition delay time
301
The ignition delay time occurs because, although the initial pressure, initial temperature, and
302
other parameters of the combustion system meet the critical ignition conditions at t = 0,
303
combustion does not begin until the temperature reaches Tc at t = t0, where t0 is the ignition delay
304
time. The ignition delay time is a crucial parameter for fuel ignition and combustion. It is also an
305
essential criterion for judging the rationality of a chemical kinetic model.
306
Firstly, the ignition delay time under different working conditions was calculated using the
307
closed homogeneous reaction module of CHEMKIN. It was then compared with the results of
308
other researchers. Figure 3 compares the simulated ignition delay times for the mechanism
309
proposed in this study, the results of other researchers, and the test data for aviation kerosene
310
published by Changhua Zhang et al. (23) under the following conditions. The mechanisms from
311
the literature is used for the model predictions of detail and simplified mechanisms. The pressure
312
(P) was 1 atm, the equivalence ratio (Φ) was 1, and the temperature was in the range 1,150–1,550
ACS Paragon Plus Environment
Page 17 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
313
K. The graph shows that the ignition delay time of the simplified mechanism proposed in this
314
study was consistent with the test data. The difference between them is less than 6.5%, which is
315
acceptable for engineering. While the ignition delay times of case A, B and C were all close to the
316
test data, the trends were different. The difference of ignition delay time between case D and the
317
test data is larger under these working conditions. In summary, Figure 3 shows that the proposed
318
simplified mechanism is closer to the test data when P = 1 atm and Φ = 1.
Figure 3. Ignition delay time for P = 1 atm and Φ = 1 319
5.2 Laminar flame speed
320
The laminar flame speed is an important parameter of the propagation characteristics of fuel
321
combustion. To verify the accuracy of the proposed simplified mechanism, its laminar flame speed
322
had to be verified. The laminar flame speed (S) is a function of the temperature coefficient (A) and
323
the reaction rate (W). It is a physical and chemical parameter of a combustible gas (and the most
324
affected by temperature).
325
The laminar flame speed without stretch was calculated with the premixed laminar
326
flame-speed calculation module in CHEMKIN-PRO. Figure 4 shows a comparison of the laminar
327
flame speeds of case A, B, C, E and the simplified mechanism proposed in this study, and the
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
328
experimental data of the laminar flame speed is from the literature (2). The test data for RP-3
329
aviation kerosene published by Yu (2) worked under the same conditions where P = 1 atm, Φ =
330
0.7–1.4, and the initial unburned mixture temperature is 403 K. Figure 4 shows that while the
331
difference between the test data and the numerical simulation results of case E was 35%, they
332
follow the same trend. The difference between the results of case C and the test data was relatively
333
larger; furthermore, they followed different trends. As regards the simplified mechanism proposed
334
in this study, case A and case B, their results were consistent with test data. At Φ = 1.1, the laminar
335
flame speed calculated using the proposed mechanism was 67.2 cm/s, and the difference between
336
this and the test data (63.1 cm/s) was 6.5%. Additionally, they followed similar trends. Overall,
337
while the model predictions of the ignition delay time has a good agreement with the experimental
338
data, it has some different for the laminar flame speed, as shown in Figure 4. This is because the
339
numerical simulation is conducted under ideal conditions, in contrast, in the experimental
340
conditions, the heat transfer, measurement error, etc. will also have some effects on the results.
341
Even so, the simplified and detail mechanisms have good coincide characteristics with the
342
experimental data.
Figure 4. Laminar flame speed at P = 1 atm and T = 403 K
ACS Paragon Plus Environment
Page 18 of 32
Page 19 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
343
6. Application of simplified mechanism for three-component surrogate fuel for RP-3 aviation
344
kerosene in numerical simulation of premixed pre-evaporation combustion
345
6.1 CFD numerical simulation and calculation conditions
346
To verify the engineering applicability of the proposed simplified mechanism for RP-3
347
aviation kerosene, in this study, the CFD calculation software FLUENT was used to simulate the
348
premixed pre-evaporation combustion flame of a Bunsen burner. The proposed simplified
349
mechanism was applied to the calculation to simulate the chemical reactions of RP-3 aviation
350
kerosene (24). The simulation results were compared with experimental data (24,27) to verify the
351
feasibility of the simplified mechanism for practical combustion problems. Figure 5 shows a
352
physical model of a Bunsen burner adopted in the numerical simulation. The outlet diameter of the
353
Bunsen burner was 12 mm and its height was 20 mm. It was surrounded by a lampshade that was
354
600 mm high and 350 mm wide. As regards the calculation model, the boundary conditions were
355
as follows. The exit of the Bunsen burner was set as a mass flow inlet, the air entrance of the
356
lampshade was set as pressure inlet boundary condition, the exit of the lampshade was set as
357
pressure outlet boundary condition, and the wall of the lampshade was set as an adiabatic wall.
358
Figure 6 shows the computational grids for the physical model of the Bunsen lamp and Figure 7
359
shows local enlarged detail. Because of the obvious velocity shear layer at the outlet of the nozzle,
360
the grid of fuel nozzle and a part of region close to nozzle outlet are refined. These were
361
unstructured and the mesh number was 2,430,000. Additionally, the grid independence was
362
performed before simulation. The velocity magnitude distributions along y axis on z=0.3m for
363
different grid number are presented in Figure 8. The number of grid is 1.46, 2.07, 2.43 and 3.18
364
million, respectively. The figure shows that the velocity distributions of 2.43 and 3.18 million
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 20 of 32
365
grids are in good agreement, indicating that 2.43 million grids employed in this paper are
366
reasonable. The mathematical models used in the numerical calculation were the standard
367
K-epsilon turbulence model, the component transport model, and the eddy-dissipation-concept
368
model. The local mixture of fuel and air will be accelerated due to heat release of combustion. The
369
fluctuation in velocity will also be enhanced. In such, the flow under the combustion environment
370
is turbulent and the turbulent model should be employed, and the standard k-epsilon turbulence
371
model is employed in this paper. Discrete equations of pressure, momentum, energy, and all the
372
components were second order upwind scheme. Discrete equations of the turbulent kinetic energy
373
and turbulence dissipation were first order upwind scheme. In numerical simulation, the influence
374
of radiation is neglected. First of all, the radiation boundary conditions are not easy to determine.
375
Moreover, radiation models in FLUENT are not accurate enough. Additional equations will
376
increase the calculating time of numerical simulation at the current calculation level. According to
377
the numerical simulation convergence criteria standard of FLUENT, the numerical simulation was
378
judged to be convergent under conditions for which the relative errors of the inlet and outlet flows
379
were less than 1%, the residual of the continuity equations was below 1.0 × 10−6, and other
380
residuals were below 1.0 × 10−3.
381 382
Figure 5. Physical computation model of Bunsen burner
Figure 6. computational grid
ACS Paragon Plus Environment
Page 21 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
383 Figure 7. Local enlarged detail of the grid
Figure 8. Grid sensitive analysis (velocity magnitude distribution along y axis on z=0.3m)
384
Table 2 shows the conditions of the numerical simulation for premixed pre-evaporation
385
combustion, which were consistent with those in references (24) and (27). The oil gas ratio for
386
condition 1 was slightly higher than that for ideal conditions. Condition 2 represented oil-rich
387
conditions and condition 3 represented oil-lean conditions. The atmospheric temperature of the
388
test environment was 300 K, and its atmospheric pressure was 101,325 Pa.
389
Table 2. Numerical simulation conditions Pipe
Case
Fuel flow
Air flow
Inlet
Equivalence
ml/h
L/h
temperature/K
ratio
1
80
700
430
1.07
2
90
900
430
0.94
3
100
800
430
1.19
diameter/mm
12
390
6.2 Analysis of numerical results
391
Figure 9 shows the temperature contour maps of the center section of the premixed
392
pre-evaporation combustion flame of a Bunsen burner. As shown in the figure, the temperature
393
distribution of the Bunsen burner flame was the same under three working conditions, and the
394
premixed pre-evaporation gas of aviation kerosene and air was ejected from the exit of the Bunsen
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 22 of 32
395
burner. This burned rapidly after ignition and formed a flame front, and the temperature rose to
396
form a red high-temperature zone. In front of the high-temperature zone, the inner flame exits
397
from a triangular thin layer region. At the back of the high-temperature zone, due to the decrease
398
in combustion intensity, the post-combustion airflow moved downstream, entraining surrounding
399
cold air, and the temperature gradually decreased and finally tended to ambient temperature.
400
However, there are some significant differences between the three conditions. Under condition 2
401
(oil-lean), the length of the high-temperature area (over 1,800 K) along the axis was 0.065 m,
402
which was shorter than those of condition 1 (0.069 m) and condition 3 (0.085 m). This was
403
because under oil-rich conditions, in addition to the premixed flame, there was diffusion flame,
404
which increased the length of the high-temperature area.
(a) Condition 1
(b) Condition 2
(c) Condition 3
405
Figure 9. Temperature contour maps of premixed pre-evaporation combustion flame in center
406
section
407
Figure 10 compares the temperatures on the central axis for the simulation results and test
408
data under the three conditions. The simplified mechanism is proposed in this paper and the
409
experimental data is resulted from the current work (from the Bunsen burner). In the experiment,
410
B-type thermocouple is used to measure the temperature of Bunsen burner, and its uncertainty is
411
±5K. In this figure, coordinate 0 represented the nozzle exit position. In the region of temperature
412
increase, the experimental data for each sampling point almost agreed with the calculated values,
ACS Paragon Plus Environment
Page 23 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
413
and they showed the same declining trend.
414
There were two explanations for the higher values of the calculation results in comparison
415
with the experimental data at the maximum value: 1) Errors originated from the measuring
416
equipment. 2) Numerical calculations were performed in an ideal environment, while the
417
experiment may have been subject to atmospheric interference in the environment, and the cooling
418
conditions for the experiment were different from those for the simulation.
(a) Condition 1
(b) Condition 2
(c) Condition 3
Figure 10. Temperature of premixed flame on central axis 419
Figure 11 shows the isoline nephogram of the OH mass fraction on the central section.
420
Hydroxyl, -OH, is regarded as a marker of the flame front and heat release rate. In this study, the
421
flame surface is judged using the OH diagram. In Figure 11(b), condition 2 corresponded to the
422
flame structure in oil-lean working conditions. It can be clearly seen that a high OH concentration
423
was present only at the outlet of the nozzle. This was because under this condition, premixed
424
flame existed without diffusion flame. Additionally, the cone of the premixed flame shrank to a
425
minimum under this condition. Figure 11(a) (condition 1) and Figure 11(c) (condition 3) show the
426
flame structures for oil-rich conditions. Condition 1 represents slightly oil-rich conditions, with
427
two layers having high OH concentrations. The diffusion flame front formed at the back of the
428
premixed flame. As condition 3 was more oil-rich than condition 1, the two layers can be seen
429
more clearly. Additionally, a crescent hollow area can be clearly seen between the two layers,
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 24 of 32
430
which indicates separation between the premixed and diffusion flame. From the above simulation
431
results, it can be concluded that under oil-rich conditions, both premixed and diffusion flame
432
existed. However, only premixed flame existed under oil-lean conditions. By comparing the three
433
conditions, it was found that under condition 2 (oil-lean condition), the maximum OH
434
concentration was the highest, that under condition 1 (slightly oil-rich) was second, while the
435
maximum OH concentration under condition 3 (oil-rich condition) was the lowest.
(a) Condition 1
(b) Condition 2
(c) Condition 3
Figure 11. Isoline nephogram of OH mass fraction of the central section 436
Figure 12 shows the contours of the CO2 mass fraction of the central section under the three
437
conditions. In the experiment, the sampling gas of the Bunsen burner is transmitted to an infrared
438
continuous gas analyzer (SIEMENS U23, Germany) to measure the volume concentrations of CO,
439
NO, O2, and CO2 via a sampling pipe. The sampling pipe is electrically heated to assure the
440
measurement accuracy. The measuring range of CO2 volume concentration is 0 to 10%, and the
441
measuring lever is 1%. The large molecules cracked into small intermediate molecules at the exit
442
area of the Bunsen burner and eventually converted from CO to CO2 (19th reaction step CO + OH
443
↔ CO2 + H). The OH concentration directly determined the generation rate and concentration
444
distribution of CO2, and the OH concentration in the flame front was generally the highest.
445
Therefore, the highest concentration of CO2 appeared behind the flame front. Under condition 2
446
(oil-lean conditions, Figure 12(b)) the high concentration region of CO2 was further forward on
ACS Paragon Plus Environment
Page 25 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
447
the central axis. This was because the premixed flame was further forward under these conditions.
448
A high concentration region of CO2 (the region where the mass fraction of CO2 is higher than 0.14)
449
appeared at 0.018–0.056 m on the central axis. It can be clearly seen that high concentration
450
regions were relatively far from the nozzle exit under conditions 1 and 3 (oil-rich conditions). This
451
is because under oil-rich conditions, both premixed and diffusion flame existed; CO is generated
452
due to the oxygen deficiency of the premixed flame, and CO2 is formed by CO combustion in the
453
high OH concentration region of the diffusion flame. A high concentration of CO2 appeared at
454
0.03–0.061 m on the central axis under condition 1, and under condition 3, it appeared at
455
0.045–0.075 m. Figure 13 compares the CO2 along the central axis for the simulation results and
456
experimental data under the three conditions. Figure 13 shows that the CO2 concentration trend
457
was the same for the calculation results and experimental data. The highest concentration of CO2
458
was situated at Z = 0.032 m, which was closer to the nozzle exit under condition 2 than under
459
condition 1 (Z = 0.051 m) and condition 3 (Z = 0.66 m). This phenomenon was due to their
460
different flame structures; under oil-rich conditions, most of the CO2 was generated at the back of
461
the diffusion flame, which was further behind than for the premixed flame under oil-lean
462
conditions.
(a) Condition 1 (b) Condition 2 (c) Condition 3 Figure 12. Contours of CO2 mass fraction of the central section
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
(a) condition 1
(b) Condition 2
Page 26 of 32
(c) Condition 3
Figure 13. Mass fraction of CO2 along the central axis 463
Figure 14 shows the temperature contours of premixed flame under condition 2 at three
464
positions, Z = 9 mm, Z = 25 mm, and Z = 47 mm. The three figures show that the calculation
465
results for Z = 25 mm and Z = 47 mm are similar, with high-temperature regions. The
466
high-temperature areas are situated at the center of the sections and diffuse continuously along the
467
radial direction, forming concentric circles of different temperatures. The temperature along the
468
radial direction decreases gradually. Because the plane of Z = 25 mm is located in the center of the
469
high-temperature region of Figure 9(b), while that of Z = 47 mm is located in the tail of the
470
high-temperature region, the radius of the high-temperature region at Z = 25 mm is larger than that
471
at Z = 47 mm. As regards Z = 9 mm, there is a low-temperature region within the
472
high-temperature region due to Z = 9 mm being located at the core of the premixed flame, in
473
which there is unburned kerosene. This proves that the flame of the Bunsen burner is hollow, and
474
that its surface is conical. Figure 15 compares the premixed flame temperatures of the
475
experimental data and simulation results under condition 2 at Z = 9 mm, Z = 25 mm, and Z = 47
476
mm. As shown in Figure 14, the temperature initially increased and then decreased along the
477
radius. This was because the inner flame was located at height Z = 9 mm and there was unburned
478
low-temperature mixed gas in the inner flame. The radial direction passed through the mixed
479
low-temperature region to the flame surface, which caused the gradual increase in temperature. As
ACS Paragon Plus Environment
Page 27 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
480
the radius increased from the flame surface to the ambient atmosphere, the temperature decreased.
481
As regards heights Z = 25 mm and Z = 47 mm, these planes did not pass109th
482
Through the Bunsen inner flame region, so the temperature decreased as the radius
483
increased. By comparing experimental data and simulation results, it was found that, despite
484
minor differences between them, the trends of the experimental and calculated values were almost
485
identical.
(a) Z = 9 mm
(b) Z = 25 mm
(c) Z = 47 mm
Figure 14. Temperature contours of premixed flame for different heights under condition 2
(a) Z = 9 mm
(b) Z = 25 mm
(c) Z = 47 mm
Figure 15. Comparison of premixed flame temperatures at different heights 486
7. Conclusions
487
In this study, DRG, DRGEP, CSP, path analysis and sensibility analysis methods were
488
adopted to simplify the detailed mechanism for a three-component surrogate fuel, and a simplified
489
mechanism involving 59 components and 158 elementary reactions was obtained. The ignition
490
characteristics and flame propagation characteristics were then selected to verify the proposed
491
simplified mechanism using CHEMKIN. Finally, the simplified mechanism for three-component
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
492
surrogate fuel for RP-3 aviation kerosene was applied to numerical simulation of a premixed
493
pre-evaporation combustion flame, and a detailed comparison was made between the simulation
494
results and the experimental results. The conclusions are as follows.
495
(1) The path analysis showed that for fuel macromolecules and most intermediate
496
components, as the temperature increases, the proportion of dehydrogenation and isomerization
497
paths will decrease, while the proportion of decomposition paths will increase.
498
(2) The simplified mechanism for three-component surrogate fuel set up using DRG, DRGEP,
499
CSP, path analysis, and sensibility analysis was consistent with experimental data. The results
500
showed that the proposed simplified mechanism was consistent with the detailed mechanism.
501
(3) The simulation results for a premixed pre-evaporation combustion flame using the
502
proposed simplified mechanism showed that parameters such as combustion temperature
503
distribution and concentration of CO2 were consistent with the experimental data. The
504
computational load is acceptable.
505
In conclusion, by comparing the simulation calculated using the simplified mechanism for
506
the three-component surrogate fuel proposed in this study with experimental data, it was found
507
that the proposed mechanism can successfully simulate the combustion of kerosene and predict
508
the kinetic characteristics of combustion such as ignition delay time and flame propagation speed
509
within a large range of equivalence ratios. Additionally, the simplified mechanism was applied to
510
simulate a premixed pre-evaporation combustion flame using RP-3 aviation kerosene as fuel. The
511
simulation results were consistent with experimental data, and the computation time is acceptable
512
for engineering, which proved that the simplified mechanism for the three-component surrogate
513
fuel (59 components and 158 elementary reactions) proposed in this study can be used for
ACS Paragon Plus Environment
Page 28 of 32
Page 29 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
514
numerical simulation of RP-3 aviation kerosene combustion engineering.
515 516
Acknowledgments This work received funding from National Natural Science Foundation of China
517 518
(No.51676097, No. 91741118).
519 520
Corresponding author:
521
Dr. Yingwen YAN, E-mail:
[email protected], Tel: +86-13770507240.
522 523
References
524
1. Yan C.; Fan W. Combustion. Xi'an: Northwestern Polytechnic University Press. 2005.
525
2. Yu W. Study on flame speed and chemical reaction mechanism for alternative fuels of aviation
526
kerosene. Beijing, Tsinghua University 2014, 5–9.
527
3. Wu Y.; Modica V.; Yu X.; Grisch F. Experimental investigation of laminar flame speed
528
measurement for kerosene fuels: Jet A-1 surrogate fuel, and its pure components. Energy
529
Fuels 2018, 32(2), 2332-2343, DOI: 10.1021/acs.energyfuels.7b02731
530
4.
Kundu P.; Deur J. A simplified reaction mechanism for calculation of emissions in
531
hydrocarbon (Jet-A) combustion. Joint Propulsion Conference and Exhibit, Monterey, CA,
532
United States, AIAA paper 1993, 1993–2341, DOI: 10.2514/6.1993-2341
533
5. Zhong F.; Ma S.; Zhang X.; Sung C.; Niemeyer K. Development of efficient andaccurate
534
skeletal mechanisms for hydrocarbon fuels and kerosene surrogate. Acta Mechanica Sinica
535
2015, 31(5), 732-740, DOI: 10.1007/s10409-015-0434-5
536
6. Zettervall N.; Fereby C.; Nilson E. J. K. Small skeletal kinetic mechanism for kerosene
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 30 of 32
537
combustion. Energy Fuels 2016, 30(11), 9801-9813, DOI: 10.1021/acs.energyfuels.6b01664
538
7. Wu K.; Yao W.; Fan X. Development and fidelity evaluation of a skeletal ethylene mechanism
539
under scramjet-relevant conditions. Energy Fuels 2017, 31(12), 14296-14305, DOI:
540
10.1021/acs.energyfuels.7b03033
541 542 543 544 545 546 547 548 549
8. Wang T. Thermophysics characterization of kerosene combustion. J Thermophys Heat Transf. 2001, 15(2), 76-80, DOI: 10.2514/2.6602 9. Patterson P.; Kyne A.; Pourkhashanian M.; Williams A.; Wilson C. Combustion of kerosene in counter-flow diffusion flames. J Propuls Power 2000, 17(2), 453–460, DOI: 10.2514/2.5764 10. Honnet S.; Seshadri K.; Niemann U.; Peter N. A surrogate fuel for kerosene. Proc Combust Inst 2009, 32(1), 485–492, DOI: 10.1016/j.proci.2008.06.218 11. Vovelle C.; Delfau JL.; Reuillon M. Formation of aromatic hydrocarbons in decane and kerosene flames at reduced pressure. Soot Form Combust 1994, 59(59), 50–65 12. Dagaut P.; Ristori A.; Bakali A.; Cathonnet M. Experimental and kinetic modeling study of the
550
oxidation
of
n-propylbenzene.
551
10.1016/S0016-2361(01)00139-9
Fuel
2002,
81(2),
173–184,
DOI:
552
13. Chitral K.; Puduppakkam K.; Modak A.; Meeks E.; Wang Y. Detailed chemical kinetic
553
mechanism for surrogates of surrogate jet fuel. Combust Flame 2011, 158(3), 434–445, DOI:
554
10.1016/j.combustflame.2010.09.016
555 556
14. Xiao B.; Yang S.; Zhao H. Detailed and reduced chemical kinetic mechanisms for RP-3 aviation kerosene combustion. J Aerosp Power 2010, 25(9), 1949-1955
557
15. Xu J.; Guo J.; Liu A.; Wang J.; Tan N.; Li X. Construction of autoignition mechanisms for the
558
combustion of RP-3 surrogate fuel and kinetics simulation. Acta Phys.-Chim Sin 2015, 31(4),
ACS Paragon Plus Environment
Page 31 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
559 560 561 562
643–52, DOI: 10.3866/PKU.WHXB201503022 16. Montgomery C.; Cannon S.; Mawid M. Reduced chemical kinetic mechanisms for JP-8 combustion. AIAA Paper-2002-0336, 2002, DOI: 10.2514/6.2002-336 17. Tang H.; Zhang C.; Li P.; Wang L.; Ye B.; Li X. Experimental study of autoignition
563
characteristics
of
kerosene.
564
10.3866/PKU.WHXB201202161
Acta
Phys.-Chim
Sin
2012,
28(4),
1-6,
DOI:
565
18. Wang F.; Li X. A Fortran Program for Mechanism Reduction, China, 0288831[P]. 2014.
566
19. Lu T.; Law C. K. A directed relation graph method for mechanism reduction. Proc Combust
567 568
Inst 2005, 30(1), 1333–1341. DOI: 10.1016/j.proci.2004.08.145 20. Pepiot-Desjardins P.; Pitsch H. An efficient error propagation-based reduction method for large
569
chemical
570
10.1016/j.combustflame.2007.10.020
571
kinetic
mechanisms.
Combust.
Flame
2008,
154(1-2),
67–81,
DOI:
21. Lam S. Using CSP to understand complex chemical kinetics. Combust Sci Technol 1993,
572
89(5-6), 375–404. DOI: 10.1080/00102209308924120
573
22. CHEMKIN-PRO, Reaction Design, San Diego, CA, 2008.
574
23. Zhang C.; Li B.; Rao F.; Li P.; Li X. A shock tube study of the autoignition characteristics of
575
RP-3 jet fuel. Proc Combust Inst 2014, 35(3), 3151–3158, 10.1016/j.proci.2014.05.017
576
24. Yan Y.; Liu Y.; Di D.; Dai C.; Li J. Simplified chemica reaction mechanism for surrogate fuel
577
of aviation kerosene and its verification. Energy & fuels 2016, 30(11): 10847 -10857. DOI:
578
10.1021/acs.energyfuels.6b01852
579
25. Strelkova M.; Kirillov I.; Potapkin B.; YanUmanskiy S.; Bagaturyants A.; Safonov A.;
580
Sukhanov L.; Liventsov V.; Deminsky M.; Dean A.; Varatharajan B.; Tentner A. Detailed and
ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 32 of 32
581
reduced mechanisms of Jet A combustion at high temperatures. Combust Sci Tech 2008,
582
180(10-11), 1788–1802. DOI: 10.1080/00102200802258379
583
26. Dai C.; Wang Y.; Yan Y.; Li J.; Dang X. Reduced mechanism of surrogate fuel for RP-3
584
kerosene based on sensitivity analysis. J Nanjing Univ Aeronaut Astronaut 2014, 47(1),
585
579–587, DOI: 10.16356/j.1005-2615.2015.04.016
586
27. Liu Y.; Yan Y.; Dai C.; Li J. A simplified chemical reaction mechanism for surrogate fuel of
587
aviation
kerosene.
Chem
Res
588
10.1021/acs.energyfuels.6b01852
Chinese
Univ.
2017,
ACS Paragon Plus Environment
33(2),
274–281.
DOI: