A workbench for the reduction of detailed chemical kinetic

6 days ago - In skeletal reduction, DRG with DRGEP was found to be the optimal one and finally selected for skeletal reduction. Within 10% error toler...
0 downloads 0 Views 705KB Size
Subscriber access provided by Kaohsiung Medical University

Combustion

A workbench for the reduction of detailed chemical kinetic mechanisms based on DRG and its deduced methods: Methodology and n-cetane as an example Yue Qiu, Liang Yu, Leilei Xu, Yebing Mao, and Xingcai Lu Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.8b00981 • Publication Date (Web): 31 May 2018 Downloaded from http://pubs.acs.org on May 31, 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 27 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 workbench for the reduction of detailed chemical kinetic mechanisms based on DRG

2

and its deduced methods: Methodology and n-cetane as an example

3

Yue Qiu, Liang Yu, Leilei Xu, Yebing Mao, Xingcai Lu*

4

Key Lab. for Power Machinery and Engineering of M. O. E, Shanghai Jiao Tong University, 200240, Shanghai, PR China

5 6

Abstract:

Reduction of detailed mechanisms of large hydrocarbons is of significant importance to multi-CFD

7

simulations, while the scale and accuracy of the reduced mechanism are closely related to the reduction method adopted. In

8

this study, a workbench for systematic reduction of detailed mechanism was developed. It is operated on MATLAB

9

platform and integrated with CHEMKIN PRO software. In the scheme, a skeletal reduction module was firstly employed to

10

identify and eliminate unimportant species and associate reactions, in which four different algorithms based on DRG and its

11

deduced methods, including single DRG, single DRGEP, two-stage DRG and DRG with DRGEP were applied and

12

compared to find the optimal solution for further reduction. Then a subsequent reaction sensitivity analysis module was

13

implemented to eliminate less important reactions. The potential and feasibility of the proposed scheme were presented with

14

an example of reduction of a recently proposed detailed n-cetane mechanism. In skeletal reduction, DRG with DRGEP was

15

found to be the optimal one and finally selected for skeletal reduction. Within 10% error tolerance, a comprehensive

16

reduced mechanism consisting of 521 species and 1623 reactions was generated with ~75% reduction of species and 80%

17

reduction of reactions. The reduced mechanism was well validated against the detailed mechanism in the ignition delay time,

18

temperature profiles and important species concentrations in 0-D homogeneous batch reactor, species evolution in the

19

jet-stirred reactor and 1-D premixed laminar flame speeds over a wide range of pressures (10-40 bar), temperatures

20

(680-1600 K) and equivalence ratios (0.5-1.5).

21

Keywords: Mechanism reduction; Skeletal mechanism; Deduced DRGs; N-Cetane; Diesel Surrogate

*

Corresponding author: E-mail address: [email protected]. Tel.: +86-21-34206039; Fax: +86-21-34205949. 1

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

22

1. Introduction

23

Computational Fluid Dynamics (CFD) simulation is an effective technique in the design and development of modern

24

engines. The flow, spray and combustion phenomenon of fuel inside cylinder are investigated by numerically solving fluid

25

flow equations coupled with the chemical mechanism of fuel combustion in CFD simulation. However, on the one hand, the

26

realistic fuel is a complex mixture of hundreds of thousands of alkanes, alkenes and aromatics. Even if we can constitute a

27

surrogate model with limited representative components to characterize its major physiochemical properties, still the

28

mechanism is too large for simulation. For instance, a complete mechanism describing the pyrolysis, partial oxidation and

29

combustion of hydrocarbon and oxygenated fuels developed by E. Ranzi et al1 contains 451 species and 17848 reactions. A

30

detailed diesel fuel surrogate model with binary mixture of n-dodecane and m-xylene also consists of 2885 species and

31

11754 reactions2. While such a large scale may provide more details of the underlying chemical interactions, it is almost

32

impossible to apply them in multi-CFD simulations. On the other hand, the species and reactions vary a lot in the detailed

33

mechanism, which will induce severe chemical stiffness problems. The situation gets worse when a larger mechanism is

34

involved3. Therefore, the reduction of detailed mechanisms is essential for multi-CFD. In addition, mechanism reduction

35

can distinguish critical species and reactions which contributes to a clearer and more thorough understanding of the reaction

36

process.

37

Set against this background, much effort has been dedicated to the development of mechanism reduction in recent

38

years3. In general, there are two major categories of reduction methods: skeletal reduction and time-scale analysis reduction.

39

Time-scale analysis methods include typical quasi-steady state assumption (QSSA), partial-equilibrium approximation

40

(PEA), computational singular perturbation (CSP) analysis, intrinsic low dimensional manifold (ILDM) and so on. More

41

information can be found in Lu’s3 review. Skeletal reduction methods focus on identifying and eliminating unimportant

42

species and reactions. It is efficient and typically the first step of mechanism reduction. Among them the Directed Relation

43

Graph (DRG) first proposed in 2005 by Lu and Law4 gains much concern recently. It uses a directed graph to map the

44

coupled species, wherein the vertex represents species and a direct weighted edge quantifies the dependences of one species

45

on another. A skeletal mechanism can be generated by carefully setting the threshold to remove unimportant species. 2

ACS Paragon Plus Environment

Page 2 of 27

Page 3 of 27 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

Following DRG, many developments and improvements have been made: Luo et al.5 revised the quantification of species

47

interaction with ‘max’ operator in replace of original ‘sum’ operator for better reduction of mechanisms with large isomers;

48

Lu proposed a two-stage DRG reduction strategy6 which restarts the reduction after the first round of DRG implementation

49

and DRGASA7 which performs sensitivity analysis to further reduce unimportant species. While DRG is efficient, concise

50

and easy for implementation, it only focuses on the direct relations between two species within one reaction without

51

considering that the influences will be weakened if two species are far from each other. To compensate for this,

52

Pepiot-Desjardins suggested a geometric error propagation strategy and developed DRGEP method8 which takes into

53

consideration the length of chemical paths between two candidate species. A path dependent coefficient is adopted to

54

quantify species interactions. Niemeyer9 compared the reliability and effectiveness of different graph search algorithms used

55

in DRGEP and found that Dijkstra’s algorithm generated more compact skeletal mechanisms than depth first search (DFS),

56

breadth first search (BFS) and revised breadth first search (RBFS) algorithm. Subsequently, Dijkstra algorithm was applied

57

in DRGEP and later Niemeyer and Sung10 combined DRGEP with sensitivity analysis to obtain DRGEPSA. Besides

58

DRGEP, there are also many other methods based on DRG such as the path flux analysis (PFA)11, 12, the flux project tree

59

method13 and the Jacobi based DRGEP14. To sum up, DRG and its deduced skeletal reduction methods have been

60

researched heatedly in recent years due to its simplicity, effectiveness, and minimal requirement of knowledge with fuel

61

chemistry.

62

There have been several successful applications of DRG and its deduced methods in the reduction of single

63

components such as n-alkanes13, 15-17, iso-alkanes10, 18 and esters19-21, as well as multi-component surrogate models like

64

gasoline22-25, diesel26-29, jet fuel30 and biodiesel5, 31, 32. In constructing surrogate models, hydrocarbons with large molecules

65

are deemed as important components to characterize the physicochemical properties of target fuel. Among them, n-cetane is

66

a favorable candidate in diesel surrogate models33, 34. First, it has comparable average carbon number with that of real diesel

67

(C10-C22). In addition, it is a primary reference fuel for cetane number (CN) rating which permits fuel blending to adapt to

68

the wide cetane-number range of real diesel. However, the chemical kinetic mechanism of n-cetane is not comprehensively

69

studied owing to its complex long-chain structure. Westbrook et al.35 built a detailed mechanism with 2116 species and 8130 3

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

70

reactions for combustion of n-alkanes (C8-C16)35 based on the classification of 25 reaction classes. Later, Sarathy et al.36

71

made some modifications to the above n-alkane mechanism35 in their combined mechanism of C7-C20 n-alkanes and singly

72

methylated alkanes. Those updates were demonstrated to increase low temperature reactivity, as validated by the authors.

73

Researches on the reduction of detailed n-cetane mechanism have also been rare. Liang et al.37 generated a reduced

74

381-species n-cetane baseline mechanism from Westbrook’s35 detailed mechanism using CHEMKIN-PRO Workbench to

75

test advanced solution strategies in HCCI combustion, whilst the reduction details were not elaborately explained. Poon et

76

al.34 proposed a five-stage reduction scheme and implemented it with the example of n-cetane. a reduced mechanism with

77

79 species was derived from detail mechanism of Westbrook et al35 and validated in the 0-D closed homogeneous batch

78

reactor and 2-D spray combustion simulation in a constant volume bomb. Nonetheless, the maximum error adopted is too

79

large, reaching 40%. On the whole, large hydrocarbons such as n-cetane feature high boiling temperature, low saturate

80

pressure and homogenously mixing with oxidizer. This leads to difficulties in experimental studies of ignition, oxidation,

81

pyrolysis as well as laminar flame, and limits the development of detailed mechanism and subsequent reduction research.

82

This paper attempts to further explore the feasibility of systematic reduction of detailed mechanism of large

83

hydrocarbons. A reduction workbench was constructed based on MATLAB 2017b platform and integrated with

84

CHEMKIN-PRO 1513138. It is composed of a skeletal reduction module followed by a reaction analysis module. In the

85

skeletal reduction module, four different DRG-based algorithms, including single DRG, single DRGEP, two-stage DRG and

86

DRG with DRGEP were discussed and compared to select an optimal method for further reduction. As an illustration, a

87

reduced mechanism of n-cetane was generated from the newly constructed detailed mechanism by Sarathy et al.36 following

88

the scheme and validated in the 0-D homogeneous batch reactor, jet stirred reactors (JSR) and 1-D premixed laminar flame

89

speed simulator.

90

2. Mechanism reduction workbench

91

The reduction scheme proposed in this work focuses on skeletal reduction of detail mechanisms. An overall flow chart

92

of the reduction process is illustrated in Fig.1. The original detailed mechanism, desired range of conditions (temperature,

93

pressure and equivalence ratio) and the desired accuracy of the reduced mechanism are required inputs. 4

ACS Paragon Plus Environment

Page 4 of 27

Page 5 of 27 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

94

First, CHEM is called to interpret the original detailed mechanism on MATLAB platform. Species and reactions are

95

numbered and the thermal data and kinetic reactions are sorted in corresponding order. A matrix containing stoichiometric

96

coefficients of species in the reactions is also generated. In this stage, the species and reactions are indexed by number so

97

that any desired mechanism can be generated with given numbers.

98

After that, constant volume simulations of autoignition within desired range of conditions are performed with

99

CKReactorGenericClosed solver to sample ignition delay times and net reaction rates data. As current reduction focuses on

100

the autoignition characteristic of fuels, the accuracy of the generated reduced mechanism refers to the maximum absolute

101

relative error in ignition delay time prediction between the reduced mechanism and detailed mechanism. Herein, the ignition

102

delay time is defined as the timepoint of the maximum temperature gradient along the temperature profile.

103

Next, skeletal reduction module is implemented to identify and eliminate unimportant species. Sampled net reaction

104

rates are used to calculate coupling coefficients between species. Thus, a weighted digraph is constructed where vertices

105

denote species and edges denote the coupling of one species on another. Given a threshold, those edges with coefficients

106

smaller than this value are removed from the graph while the remained edges connect the strongly coupled species. With a

107

specific graph search method, a set of coupled important species can be identified and a skeletal mechanism containing only

108

those selected species and associated reactions is generated. The desired skeletal mechanism which has the smallest scale

109

under given accuracy is achieved by iteratively varying the thresholds and calculating the accuracy of generated skeletal

110

mechanism. Four different skeletal reduction methods are implemented and compared, including single DRG, single

111

DRGEP, two-stage DRG and DRG with DRGEP. DRG4 and DRGEP8 methods differ in calculation of species coupling

112

coefficients and graph search algorithm. Herein, DRG method considers the net production of a species in calculating

113

species coupling coefficients and adopts a depth first search algorithm to find a set of species coupled with initial target

114

species. While in DRGEP reduction, the production and consumption of a species are considered separately and Dijkstra

115

algorithm is applied to calculate the coupling coefficient between two species. One concern of implementation of DRG and

116

DRGEP methods is that error can be induced after the elimination of some species due to the change of the mechanism

117

coefficient matrix. Hence, multi-stage reduction is also considered while studies found that a two-stage reduction is 5

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

118

adequate32. Therefore, single DRG, singe DRGEP, two-stage DRG and DRG with DRGEP are adopted and discussed at this

119

stage to select the optimal one for further reduction.

120

Once DRG-based skeletal reduction efficiently identifies and removes a large number of unimportant species, a

121

reaction sensitivity analysis module is implemented to further eliminate less important reactions in the skeletal mechanism.

122

Herein, a simple and reliable method is proposed in which reactions are ranked in ascending order of maximum reaction

123

rates along evolution among all the sampling points. They are removed one by one until the accuracy of generated

124

mechanism meets error limit. Note that a reversible reaction should be treated as a single reaction in case of fast reversible

125

reactions that quickly reach partial equilibrium7.

126

Following this scheme, a reduced mechanism with considerably fewer species and reactions is finally generated.

127

Further external validations in the Jet Stirred Reactor(JSR), Laminar Flame Speed Calculator and so on are also conducted

128

to test the accuracy and applicability of the generated reduced mechanism.

129 130

3. Reduction of a detailed n-cetane mechanism

131

3.1 Data sampling

132

Recently, Sarathy et al.36 have made some updates to the n-alkane mechanisms of Westbrook et al.35 in construction of

133

a detailed mechanism of C7-C20 2-methylalkanes (7200 species and 31400 reactions). Specifically, a new n-heptane

134

mechanism by Mehl39,

135

elementary reactions. Moreover, low-temperature reactivity of the mechanism was increased by a decrease in the activation

136

energy of the alkyl peroxy radical isomerization class and the addition of a new concerted elimination reaction class. In this

137

example, a sub-mechanism of n-cetane is generated from the above updated mechanism with 2136 species and 8076

138

reactions (methyl-alkanes excluded) to serve as the initial detailed mechanism for reduction.

40

was adopted as the core-mechanism with some adjustments of kinetic parameters of two

139

Mechanism reduction is usually targeted at selected conditions. In this example, wide engine-relevant conditions were

140

considered in order to ensure good applicability of the reduced mechanism. Pressures of 10, 20 and 40 bar, equivalence

141

ratios of 0.5, 1.0 and 1.5, and a temperature range of 680-1600 K were selected. Within the temperature range, about 10 6

ACS Paragon Plus Environment

Page 6 of 27

Page 7 of 27

Energy & Fuels

142

temperature points were selected to cover low, NTC and high temperature regions, as is illustrated in Fig.2. Kinetic data

143

were sampled at ~100 timepoints along temperature or pressure evolution profile (Fig.3). Denser sampling was made where

144

temperature or pressure profile undergoes sharp rise to capture more information of the reaction.

145

3.2 Comparison of different skeletal reduction methods

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

146

Four skeletal reduction methods, including single DRG, single DRGEP, two-stage DRG and DRG with DRGEP were

147

compared in the proposed workbench. They are different combinations of DRG and DRGEP algorithm, single stage and

148

two-stage methods. In DRG method, fuel, O2, H2O and CO2 were selected as initial target species. In DRGEP method,

149

careful selection should be made because the result is very sensitive to initial target species such as HO2, CO and CH2O

150

near NTC regions for fuels with two-stage ignition22. In this example, fuel, O2, H, CO, HO2 and CH2O were selected. For

151

the sake of simplicity, a preliminary skeletal reduction of n-cetane was conducted at a pressure of 20 bar, an equivalence

152

ratio of 1, and a temperature range of 680-1600 K to illustrate the differences among the four methods. The desired accuracy

153

of the generated skeletal mechanism is set at around 10%.

154

Fig.4 compares the maximum error and species number at increasing thresholds of the investigated four methods.

155

Fig.4(a) presents the result with single DRG method. As threshold increases, the species number of generated skeletal

156

mechanism gradually decreases, while the maximum error of ignition delay time changes nonlinearly. It remains almost

157

unchanged until the threshold increases to 0.1, after which the error steeply increases and then meets a trough as the

158

threshold reaches 0.16. This indicates it is possible to generate higher accuracy of skeletal mechanism at an even smaller

159

scale. Tosatto et al.41 found similar nonlinearity change of error curve in the reduction of JP-8 jet fuels. One reasonable

160

explanation is that the calculated coefficient is an absolute value, not with sigh. It does not distinguish between the

161

promotion and prohibition effects of one species on another. For example, the elimination of some species can reduce the

162

error of predicted ignition delay time, while subsequent elimination of other species may increase the error. Thus, it is

163

possible that the overall accuracy becomes higher. The nonlinear change of maximum error with increasing thresholds

164

indicates that there probably exists a global optimal threshold for the reduction. In this case, considering the given 10%

165

error tolerance, 0.11 was selected as the optimal threshold and 921 species were remained in skeletal mechanism with a 7

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

166

maximum error of 12.89 %.

167

Fig.4(b) demonstrates the case of single DRGEP method. As chemical path is considered, DRGEP can generate a much

168

smaller scale of skeletal mechanism than DRG at a given threshold. The species number generally declines as the threshold

169

increases, which undergoes a rapid decrease at the beginning and then slows down at a threshold of around 0.015. In

170

contrast, the maximum error of predicted ignition delay undergoes fluctuation as threshold increases. The final optimal

171

threshold was selected as 0.044 and a skeletal mechanism containing 511 species and 2061 reactions was generated. The

172

maximum error was 10.86%.

173

As for two-stage reduction methods, by setting a threshold of 1st DRG as 0.1(Fig.4(a)), a preliminary skeletal

174

mechanism was generated and subsequent 2nd reduction was carried out with the method of DRG and DRGEP, respectively.

175

Fig.4(c) and Fig.4(d) are plots of species number and maximum error of predicted ignition delay time against different

176

thresholds in 2nd DRG and 2nd DRGEP methods. As can be seen from Fig.4, the curves of species number and maximum

177

error against thresholds in 2nd reduction exhibit similar tendency to those of single stage reduction. Besides, the global

178

optimum thresholds for maximum error are almost the same in the two stages, although the mechanisms are not as it was.

179

This demonstrates the consistency of reduction algorithm on different reduction stages of mechanisms. In 2nd DRG

180

reduction, the maximum error goes through a sharp rise at a threshold of ~0.14, at which point the maximum error is ~3%.

181

However, further reduction of species will make the maximum error rapidly jump over 10% setting value. This possibly

182

implies the strongly coupling of many species at that point. Finally, 0.138 was selected to generate a reduced mechanism

183

with 777 species and 3.23% maximum error. As for 2nd DRGEP, 0.04 was selected as the threshold and the generated

184

skeletal mechanism contains 518 species with 9.19% maximum error. Table 1 lists the detailed information of the skeletal

185

mechanisms generated by four methods.

186

Fig.5 depicts the maximum error of n-cetane skeletal mechanism versus species number so as to compare more

187

intuitively the performances of four skeletal reduction methods. Overall, the DRG and its deduced methods can effectively

188

reduce the scale of detailed mechanism at a small maximum error. In the example of n-cetane skeletal reduction, the

189

maximum error is even less than 1.5% with ~50% reduction of species. As the species continue to be removed, the 8

ACS Paragon Plus Environment

Page 8 of 27

Page 9 of 27

Energy & Fuels

190

maximum error of generated skeletal mechanism raises rapidly and then goes through a trough. Similar tendency is found in

191

all the four methods, indicating nonlinear change of maximum error during reduction process. In comparison with DRG,

192

DRGEP algorithm can reach a larger reduction scale and generate a globally smaller mechanism within setting accuracy,

193

although it shows big deviations at initial reduction stage. Besides, the implementation of a second stage reduction does

194

have improved the reduction performance in this example. A larger number of species can be eliminated in the beginning

195

and the maximum error curves are less fluctuating. Taken together, the two-stage reduction, DRG with DRGEP, was finally

196

selected as the optimal method for skeletal reduction.

197

3.3 Comprehensive reduction of n-cetane

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

198

With the implementation of DRG with DRGEP method, skeletal reduction was carried out at all the sampled conditions

199

covering a wide range of temperatures, pressures and equivalence ratios. The ultimate skeletal mechanism was generated by

200

union of the retained species set at each case. It’s worth mentioning that mechanism reduction in this case exhibits little

201

sensitivity to pressure and equivalence ratio. The reduced mechanism generated at P=20 bar and ER=1 also showed good

202

performance at all other sampled conditions, with an overall maximum error being 11.11%. This is quite interesting as it

203

indicates that the oxidation of n-cetane under different conditions can be characterized by simply some critical reactions,

204

which in turn confirms the feasibility of mechanism reduction.

205

After skeletal reduction, reaction sensitivity analysis was performed to further identify and eliminate unimportant

206

reactions. Ranked reactions were removed iteratively, as explained in Section 2. Fig.6 displays the maximum error curve (all

207

sampled conditions included) versus reaction number. With the removing of the reactions, the maximum error of the

208

reduced mechanism undergoes a gradual decline in the beginning and then rises steeply as the reaction number approaches

209

1620. This is interesting because the elimination of some unimportant species can actually increase the overall reduction

210

accuracy. Finally, 411 reactions were removed from 2034 reactions under given 10% accuracy and the optimal mechanism

211

contains 521 species and 1623 reactions, as is attached in Supplementary material. The overall maximum error decreases to

212

8% after the unimportant reaction elimination.

213

Fig.7 plots in detail the relative error of ignition delay times at each sampled condition. Similar tendency of relative 9

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 10 of 27

214

error evolution with increasing temperature is found at different pressures and equivalence ratios. Overall, the reduced

215

mechanism has agreeable accuracy at low and relatively high temperature (~1200 K), while large deviation occurs in NTC

216

(~850-950 K) and higher temperature (>1400 K) region. Meanwhile, it can be concluded that the error is more sensitive to

217

temperature than pressure and equivalence ratio.

218

3.4 Validation of the reduced n-cetane mechanism

219

Validation was performed via comparing the predicted ignition delay times of both the reduced mechanism and

220

detailed mechanism in 0-D closed homogenous batch reactor within reduction range, as is shown in Fig.8. From the figure,

221

the reduced mechanism performs well within a pressure of 10-40 bar, a temperature of 600-1600 K and an equivalence ratio

222

of 0.5-1.5. The results of the reduced mechanism almost overlap with those of the detailed mechanism. Even the deviation

223

in NTC region where the largest error occurs is narrow from the figure. Within 10% error tolerance, the generated reduced

224

mechanism features ~75% reduction of species and ~80% reduction of reactions.

225

Existing studies23, 24 found that mechanism reduction targeted at ignition delay time prediction cannot guarantee the

226

prediction of time and spatial evolution of some parameters. Hence, temperature and major species evolution were also

227

validated in this work, as is shown in Fig.9-10. Three initial temperature points (T=730 K, 910 K and 1200 K) at P=20 bar

228

and ER=1 were selected to represent low, NTC and high temperature regions, respectively. It is revealed from Fig.9 that the

229

reduced mechanism overall predicts well the temperature and OH evolution, with a bit shift towards left at NTC point and

230

right at low temperature point. Similar tendency is also found in Niemeyer’s10 reduction of n-heptane and n-decane.

231

Meanwhile, two-stage ignition of n-cetane at low and medium temperature is well captured by the reduced mechanism.

232

Fig.10 further compares other mole fraction evolution of major species, wherein NC16H34, O2 and CO2 are major

233

reactants and products, HO2 is an important species in chain branching reactions and CH2O is critical in the formation of

234

CO. In correspondence with temperature and OH profile, the reduced mechanism performs well at high temperature, while

235

there exist some deviations at NTC point. The mole fraction evolution curves are found to be shifted left in the reduced

236

mechanism, which means NC16H34 was consumed earlier to initiate low-temperature reactions. Temperature rose earlier

237

and the overall reaction process was advanced. This is possibly due to the elimination of some chain-inhibiting reactions at 10

ACS Paragon Plus Environment

Page 11 of 27 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

238

Energy & Fuels

low temperature.

239

Extended validation of the reduced mechanism was also performed in jet stirred reactor(JSR) at P=10 atm, ER=1 with

240

residence time being 0.1s. Mole fractions of major species are shown in Fig.11. It is demonstrated that the reduced

241

mechanism reproduces well the species concentration at different temperatures, even in low temperature region. The

242

nonmonotonic change of the species mole fraction as temperature rises is also observed, which can be explained by the

243

NTC behavior of the fuel. Meanwhile, 1-D simulation of premixed laminar flame speeds was also conducted using both the

244

detailed and reduced mechanism at 10-40 atm under an unburnt gas temperature of 400 K. The effects of thermal diffusion

245

(the Soret effect) was taken into consideration and the mixture-averaged diffusion model was applied. There were around

246

100 grid points in the solving domain and the final results of calculated laminar flame speeds were plotted in Fig.12. As can

247

be seen from the figure, the reduced mechanism predicts perfectly well the laminar flame speeds as the equivalence ratio

248

varies over 0.5-1.5, with nearly no deviations. The high temperature chemistry of the detailed mechanism is well captured

249

by the reduced mechanism. This further confirms the feasibility of characterizing the oxidation of n-cetane with a

250

small-scale reduced mechanism.

251

4. Concluding Remarks

252

A workbench for the systematic reduction of the detailed hydrocarbon mechanisms was proposed in this study. It was

253

operated on MATLAB platform and integrated with CHEMKIN PRO software. In the scheme, a skeletal reduction module

254

was firstly employed to identify and eliminate unimportant species and associate reactions, in which four different

255

algorithms based on DRG and its deduced methods, single DRG, single DRGEP, two-stage DRG and DRG with DRGEP

256

were applied and compared to find the optimal one for further reduction. After that, a reaction sensitivity analysis module

257

was implemented to eliminate less important reactions.

258

The potential and flexibility of the proposed scheme was presented with an example of the reduction of n-cetane,

259

which is a favorable diesel surrogate component. In the skeletal reduction module, all four methods can effectively reduce

260

the scale. The maximum error of predicted ignition delay time changes nonlinearly as threshold increases and there exists a

261

global optimum point. In comparison with DRG, DRGEP algorithm can reach a larger reduction scale and generate a 11

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 12 of 27

262

globally smaller mechanism within setting accuracy, although it shows bigger deviations at initial reduction stage.

263

Two-stage methods can generate a reduced mechanism with a higher accuracy at the same species number than single-stage

264

methods. Overall, DRG with DRGEP was found to be the optimal one and finally selected as the skeletal reduction method.

265

Within 10% error tolerance, a comprehensive reduced mechanism consisting of 521 species and 1623 reactions was

266

generated with around 75% reduction of species and 80% reduction of reactions. The reduced mechanism was validated

267

against the ignition delay times, temperature profiles and important species concentrations in the 0-D homogeneous batch

268

reactor, species evolution in the jet-stirred reactor and laminar flame speeds in 1-D flame simulator over a wide range of

269

pressures (10-40 bar), temperatures (680-1600 K) and equivalence ratios (0.5-1.5). The results overall can well reproduce

270

the oxidation of n-cetane with narrow deviations. The performances in high temperature region are slightly better than those

271

of low temperature region and the maximum error usually lies in NTC region. Besides, it is found in the reduction of

272

n-cetane that mechanism reduction is less sensitive to pressure and equivalence ratio as the generated mechanism at one

273

condition also showed agreeable performances at full reduction ranges. In summary, the systematic reduction workbench

274

exhibits great feasibility and potential for large-scale reduction of detailed mechanisms.

275

Acknowledgements

276

This work was supported by the Natural Science Foundation of China (Grant No. 51425602, 91641202) and High

277

Technology Ship Research Program-Marine Low Speed Project (Phase I).

278

References

279 280 281 282 283 284 285 286 287 288 289 290 291

1.

2. 3. 4. 5. 6.

Ranzi, E.; Frassoldati, A.; Grana, R.; Cuoci, A.; Faravelli, T.; Kelley, A. P.; Law, C. K., Hierarchical and comparative kinetic modeling of laminar flame speeds of hydrocarbon and oxygenated fuels. Progress in Energy and Combustion Science 2012, 38, (4), 468-501. Pei, Y.; Mehl, M.; Liu, W.; Lu, T.; Pitz, W. J.; Som, S., A multicomponent blend as a diesel fuel surrogate for compression ignition engine applications. Journal of Engineering for Gas Turbines and Power 2015, 137, (11), 111502. Lu, T. F.; Law, C. K., Toward accommodating realistic fuel chemistry in large-scale computations. Progress in Energy and Combustion Science 2009, 35, (2), 192-215. Lu, T. F.; Law, C. K., A directed relation graph method for mechanism reduction. Proceedings of the Combustion Institute 2005, 30, 1333-1341. Luo, Z. Y.; Lu, T. F.; Maciaszek, M. J.; Som, S.; Longman, D. E., A Reduced Mechanism for High-Temperature Oxidation of Biodiesel Surrogates. Energy & Fuels 2010, 24, (12), 6283-6293. Lu, T. F.; Law, C. K., Linear time reduction of large kinetic mechanisms with directed relation graph: n-Heptane and iso-octane. Combustion and Flame 2006, 144, (1-2), 24-36. 12

ACS Paragon Plus Environment

Page 13 of 27 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

292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338

7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19.

20.

21. 22.

23. 24. 25. 26.

27.

Energy & Fuels Lu, T. F.; Law, C. K., Strategies for mechanism reduction for large hydrocarbons: n-heptane. Combustion and Flame 2008, 154, (1-2), 153-163. Pepiot-Desjardins, P.; Pitsch, H., An efficient error-propagation-based reduction method for large chemical kinetic mechanisms. Combustion and Flame 2008, 154, (1-2), 67-81. Niemeyer, K. E.; Sung, C.-J., On the importance of graph search algorithms for DRGEP-based mechanism reduction methods. Combustion and Flame 2011, 158, (8), 1439-1443. Niemeyer, K. E.; Sung, C. J.; Raju, M. P., Skeletal mechanism generation for surrogate fuels using directed relation graph with error propagation and sensitivity analysis. Combustion and Flame 2010, 157, (9), 1760-1770. Sun, W. T.; Chen, Z.; Gou, X. L.; Ju, Y. G., A path flux analysis method for the reduction of detailed chemical kinetic mechanisms. Combustion and Flame 2010, 157, (7), 1298-1307. Wang, W.; Gou, X., An improved path flux analysis with multi generations method for mechanism reduction. Combustion Theory and Modelling 2016, 20, (2), 203-220. Liu, A.-K.; Jiao, Y.; Li, S.; Wang, F.; Li, X.-Y., Flux Projection Tree Method for Mechanism Reduction. Energy & Fuels 2014, 28, (8), 5426-5433. Chen, Y.; Chen, J.-Y., Application of Jacobian defined direct interaction coefficient in DRGEP-based chemical mechanism reduction methods using different graph search algorithms. Combustion and Flame 2016, 174, 77-84. Sung, C. J.; Niemeyer, K. E., Skeletal Mechanism Generation of Surrogate Jet Fuels for Aeropropulsion Modeling. Iscm Ii and Epmesc Xii, Pts 1 and 2 2010, 1233, 1412-1417. Bahlouli, K.; Saray, R. K.; Atikol, U., Development of a Reduced Mechanism for n-Heptane Fuel in HCCI Combustion Engines by Applying Combined Reduction Methods. Energy & Fuels 2012, 26, (6), 3244-3256. Yao, T.; Pei, Y. J.; Zhong, B. J.; Som, S.; Lu, T. F.; Luo, K. H., A compact skeletal mechanism for n-dodecane with optimized semi-global low-temperature chemistry for diesel engine simulations. Fuel 2017, 191, 339-349. Li, R.; He, G.; Zhang, D.; Qin, F., Skeletal Kinetic Mechanism Generation and Uncertainty Analysis for Combustion of Iso-octane at High Temperatures. Energy & Fuels 2018, 32, (3), 3842-3850. Gerasimov, I. E.; Bolshova, T. A.; Zaev, I. A.; Lebedev, A. V.; Potapkin, B. V.; Shmakov, A. G.; Korobeinichev, O. P., Reduced Chemical Kinetic Mechanism for Methyl Pentanoate Combustion. Energy & Fuels 2017, 31, (12), 14129-14137. Lin, K. C.; Tao, H.; Kao, F.-H.; Chiu, C.-T., Minimized Skeletal Mechanism for Methyl Butanoate Oxidation and Its Application to the Prediction of C3–C4Products in Nonpremixed Flames: A Base Model of Biodiesel Fuels. Energy & Fuels 2016. Liu, C.; Zuo, Z.; Feng, H., Skeletal and Reduced Chemical Kinetic Mechanisms for Methyl Butanoate Autoignition. Energy & Fuels 2016, 31, (1), 891-895. Chen, Y.; Wolk, B.; Mehl, M.; Cheng, W. K.; Chen, J.-Y.; Dibble, R. W., Development of a reduced chemical mechanism targeted for a 5-component gasoline surrogate: A case study on the heat release nature in a GCI engine. Combustion and Flame 2017, 178, 268-276. Niemeyer, K. E.; Sung, C. J., Mechanism reduction for multicomponent surrogates: A case study using toluene reference fuels. Combustion and Flame 2014, 161, (11), 2752-2764. Niemeyer, K. E.; Sung, C. J., Reduced Chemistry for a Gasoline Surrogate Valid at Engine-Relevant Conditions. Energy & Fuels 2015, 29, (2), 1172-1185. Wang, H.; Yao, M.; Reitz, R. D., Development of a Reduced Primary Reference Fuel Mechanism for Internal Combustion Engine Combustion Simulations. Energy & Fuels 2013, 27, (12), 7843-7853. Poon, H. M.; Pang, K. M.; Ng, H. K.; Gan, S.; Schramm, J., Development of multi-component diesel surrogate fuel models – Part I: Validation of reduced mechanisms of diesel fuel constituents in 0-D kinetic simulations. Fuel 2016, 180, 433-441. Poon, H. M.; Pang, K. M.; Ng, H. K.; Gan, S.; Schramm, J., Development of multi-component diesel surrogate fuel models – Part II: Validation of the integrated mechanisms in 0-D kinetic and 2-D CFD spray combustion simulations. Fuel 2016, 181, 120-130. 13

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

339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371

Page 14 of 27

28. Qiu, L.; Cheng, X.; Wang, X.; Li, Z.; Li, Y.; Wang, Z.; Wu, H., Development of a Reduced n-Decane/α-Methylnaphthalene/Polycyclic Aromatic Hydrocarbon Mechanism and Its Application for Combustion and Soot Prediction. Energy & Fuels 2016, 30, (12), 10875-10885. 29. Sun, X.; Liang, X.; Shu, G.; Wang, Y.; Wang, Y.; Yu, H., Development of a Reduced n-Tetradecane–Polycyclic Aromatic Hydrocarbon Mechanism for Application to Two-Stroke Marine Diesel Engines. Energy & Fuels 2016, 31, (1), 941-952. 30. Wang, Q.-D.; Fang, Y.-M.; Wang, F.; Li, X.-Y., Systematic analysis and reduction of combustion mechanisms for ignition of multi-component kerosene surrogate. Proceedings of the Combustion Institute 2013, 34, (1), 187-195. 31. Luo, Z. Y.; Plomer, M.; Lu, T. F.; Som, S.; Longman, D. E., A reduced mechanism for biodiesel surrogates with low temperature chemistry for compression ignition engine applications. Combustion Theory and Modelling 2012, 16, (2), 369-385. 32. Poon, H. M.; Ng, H. K.; Gan, S.; Pang, K. M.; Schramm, J., Evaluation and Development of Chemical Kinetic Mechanism Reduction Scheme for Biodiesel and Diesel Fuel Surrogates. SAE International Journal of Fuels and Lubricants 2013, 6, (3), 729-744. 33. Pitz, W. J.; Mueller, C. J., Recent progress in the development of diesel surrogate fuels. Progress in Energy and Combustion Science 2011, 37, (3), 330-350. 34. Poon, H. M.; Ng, H. K.; Gan, S.; Pang, K. M.; Schramm, J., Development and Validation of Chemical Kinetic Mechanism Reduction Scheme for Large-Scale Mechanisms. SAE International Journal of Fuels and Lubricants 2014, 7, (3), 653-662. 35. Westbrook, C. K.; Pitz, W. J.; Herbinet, O.; Curran, H. J.; Silke, E. J., A comprehensive detailed chemical kinetic reaction mechanism for combustion of n-alkane hydrocarbons from n-octane to n-hexadecane. Combustion and Flame 2009, 156, (1), 181-199. 36. Sarathy, S. M.; Westbrook, C. K.; Mehl, M.; Pitz, W. J.; Togbe, C.; Dagaut, P.; Wang, H.; Oehlschlaeger, M. A.; Niemann, U.; Seshadri, K.; Veloo, P. S.; Ji, C.; Egolfopoulos, F. N.; Lu, T., Comprehensive chemical kinetic modeling of the oxidation of 2-methylalkanes from C-7 to C-20. Combustion and Flame 2011, 158, (12), 2338-2357. 37. Liang, L.; Puduppakkam, K.; Meeks, E., Towards Using Realistic Chemical Kinetics in Multidimensional CFD. 2009. 38. Reaction Design: San Diego, CHEMKIN-PRO. 15131; 2013. 39. Mehl, M.; Pitz, W. J.; Sjöberg, M.; Dec, J. E., Detailed Kinetic Modeling of Low-Temperature Heat Release for PRF Fuels in an HCCI Engine. In SAE International: 2009. 40. Mehl, M.; Pitz, W. J.; Westbrook, C. K.; Curran, H. J., Kinetic modeling of gasoline surrogate components and mixtures under engine conditions. Proceedings of the Combustion Institute 2011, 33, 193-200. 41. Tosatto, L.; Bennett, B. A. V.; Smooke, M. D., Comparison of different DRG-based methods for the skeletal reduction of JP-8 surrogate mechanisms. Combustion and Flame 2013, 160, (9), 1572-1582.

14

ACS Paragon Plus Environment

Page 15 of 27 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

Table 1 Comparison of skeletal mechanisms of n-cetane generated by different methods

Method

Thresholds

#Species

#Reactions

Single DRG Single DRGEP Two-stage DRG DRG+DRGEP

0.11 0.044 0.1, 0.138 0.1, 0.04

901 511 777 521

3736 2061 3079 2034

15

ACS Paragon Plus Environment

Max. Error /% 12.89 10.86 3.23 9.19

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

Fig.1. Flow chart of the reduction scheme

16

ACS Paragon Plus Environment

Page 16 of 27

Page 17 of 27

10

ignition delay (ms)

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

TL=680K

1 0.1 0.01 1E-3

TH=1600K

0.6

0.8

P=20 bar, ER=1 T=1000K

1.0

1.2

1.4

1.6

1000/T (1/K) Fig.2. An illustration of data sampling of temperature points

17

ACS Paragon Plus Environment

Page 18 of 27

5000

60

4000

40

3000

20

0

2000

P=20 bar, ER=1 T=1000K

1000 0.0

0.5

1.0

1.5

2.0

2.5

3.0

-20 3.5

time (ms)

Fig.3. An illustration of data sampling of timepoints

18

ACS Paragon Plus Environment

Pressure (bar)

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

Temperature (K)

Energy & Fuels

(b)

(a) 0.6

0.3

1000 0.2

optimal point

0.1

0.5 1500

0.4 0.3

1000 0.2 0.1

500

nd

2 reduction

0.00

0.05

0.10

0.15

optimal point

0.0

0.00

0.20

0.03

0.04

0.05

0.0 0.06

1000 (c)

(d) 0.6

0.6

Number of Species

1000

0.4 0.3 600 0.2 0.1

optimal point 0.05

0.10

Maximum Error

0.5 800

0.00

0.02

1-DRGEP Threshold

1-DRG Threshold

400

0.01

0.5 800 0.4 0.3 600 0.2 0.1 400

0.15

0.20

Maximum Error

500

Number of Species

0.4

Maximum Error

Number of Species

0.5

1500

0.6

2000

2000

Number of Species

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

Maximum Error

Page 19 of 27

optimal point

0.0 0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07

0.0 0.25

nd

nd 2 -DRGEP Threshold 2 -DRG Threshold Fig.4 Species number and maximum error of ignition delay time prediction against different thresholds for n-cetane skeletal mechanism generated by different methods (a) DRG, (b) DRGEP, (c)2nd DRG with 0.1 threshold at 1st DRG, (d)2nd DRGEP with 0.1 threshold at 1st DRG

19

ACS Paragon Plus Environment

Energy & Fuels

0.8 single DRG 2-stage DRG DRG+DRGEP single DRGEP

0.7 0.6

Maximum Error

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 27

0.5 0.4 0.3 0.2 0.1 optimal point

0.0 250

500

750

1000 1250 1500 1750 2000

Number of Species Fig.5 Comparison of the maximum error against species number of n-cetane skeletal mechanism generated by four methods.

20

ACS Paragon Plus Environment

Page 21 of 27

0.4

Maximum Error

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

0.3

411 reactions are removed from the skeletal mechanism; The final reduced mechansim contains 521 species and 1623 reactions

0.2

0.1

optimal point 0.0 1500

1600

1700

1800

1900

2000

Number of Reactions

Fig.6 Maximum error curve against reaction number

21

ACS Paragon Plus Environment

P=10, ER=0.5 P=10, ER=1.0 P=10, ER=1.5

600 800 1000 1200 1400 1600

T (K)

0.08 0.06 0.04 0.02 0.00 -0.02 -0.04 -0.06 -0.08 600

P=20, ER=0.5 P=20, ER=1.0 P=20, ER=1.5

800

1000 1200 1400 1600

T (K)

Page 22 of 27

Relative Error

0.08 0.06 0.04 0.02 0.00 -0.02 -0.04 -0.06 -0.08

Relative Error

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

Relative Error

Energy & Fuels

0.08 0.06 0.04 0.02 0.00 -0.02 -0.04 -0.06 -0.08 600

P=40, ER=0.5 P=40, ER=1.0 P=40, ER=1.5

800

1000 1200 1400 1600

T (K)

Fig.7 Comparison of maximum error of reduced mechanism at all sampled conditions: (a)10 bar, (b)20 bar, (c)40 bar

22

ACS Paragon Plus Environment

Page 23 of 27

100

100 (a)

10

10 bar

1

20 bar 40 bar

0.1 0.01

10 10 bar

1

0.8

1.0

1.2

20 bar 40 bar

0.1 0.01

line:detailed symbol:reduced

1E-3 0.6

1.4

line:detailed symbol:reduced 0.8

1.0

1000/T (1/K)

1.2

1.4

1000/T (1/K)

100 (c)

ER:1.5

ignition delay (ms)

1E-3 0.6

(a)

ER:1.0

ignition delay (ms)

ER:0.5

ignition delay (ms)

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

10 10 bar

1 20 bar 40 bar

0.1 0.01 1E-3 0.6

line:detailed symbol:reduced 0.8

1.0

1.2

1.4

1000/T (1/K) Fig.8 Comparison of ignition delay times of detailed and reduced mechanism within reduction range. (line: detailed, symbol: reduced) (a) ER=0.5, (b) ER=1.0, (c) ER=1.5

23

ACS Paragon Plus Environment

Energy & Fuels

3500

0.020

(a)

2500

Mole Fraction

3000

T (K)

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

1200K 910K

2000 730K 1500

P=20bar, ER=1 line:detailed symbol:reduced

1000 500 0.0

0.5

1.0

1.5

2.0

2.5

Page 24 of 27

(b)

0.015

OH 1200K

910K

0.010

730K 0.005

0.000 0.0

P=20bar, ER=1 line:detailed symbol:reduced 0.5

1.0

1.5

2.0

2.5

time (ms) time (ms) Fig.9 Temperature and OH profiles in ignition delay time prediction with both the detailed and reduced mechanism at P=20 bar, ER=1 (line: detailed, symbol: reduced) (a)Temperature, (b) OH mole fraction

24

ACS Paragon Plus Environment

Page 25 of 27

0.010

1200K 730k

0.004 910K

P=20bar, ER=1 line: detailed symbol:reduced

0.002

Mole Fraction

Mole Fraction

0.006

O2

0.20

0.000 0.0 0.5 1.0 1.5 2.0 2.5 3.0 time (ms)

0.15

1200K

0.10 910K 730k

0.05 0.00 0.0

0.5

1.0 1.5 2.0 time (ms)

0.10 Mole Fraction

NC16H34 0.008

1200K

0.05 910K 730k

0.00 0.0

2.5

CO2

0.5

1.0 1.5 2.0 time (ms)

2.5

0.0010 0.020

HO2

C2H4

1200K

0.010 910K

0.005 0.000 0.0

730k

0.5

1.0 1.5 2.0 time (ms)

2.5

0.0006

1200K

0.0004 910K 730k

0.0002 0.0000 0.0

0.5

1.0 1.5 time (ms)

2.0

Mole Fraction

Mole Fraction

0.015

CH2O

0.006

0.0008 Mole Fraction

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

0.004 1200K 910K

0.002 0.000 0.0

730k

0.5

1.0 1.5 2.0 time (ms)

2.5

Fig.10 Mole fractions of major species in ignition delay time simulation with both the detailed and reduced mechanism at P=20 bar, ER=1 (line: detailed, symbol: reduced)

25

ACS Paragon Plus Environment

Energy & Fuels

-3

-5

10

-7

10

-8

10

τ =0.1s P=20bar, ER=1 line: detailed symbol: reduced

Mole Fraction

-4

-6

-1

10

-1

10

10

10

O2

NC16H34 Mole Fraction

Mole Fraction

10

-2

10

-3

10

800 1000 1200 1400 1600 T (K)

800

1000 1200 1400 1600 T (K)

Mole Fraction

-3

10

-4

10

10

10

C2H4

-5

-3

10

800

1000 1200 1400 1600 T (K)

-2

10 Mole Fraction

-2

10

-2

10

10

1000 1200 1400 1600 T (K)

-2

CO

CO2

-3

800

-1

10 Mole Fraction

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 26 of 27

CH2O

-3

10

-4

10

-5

10

-6

800

1000 1200 1400 1600 T (K)

10

800

1000 1200 1400 1600 T (K)

Fig.11 Mole fractions of major species in JSR model simulation with both the detailed and reduced mechanism at P=10 atm, ER=1 and τ =0.1s (line: detailed, symbol: reduced)

26

ACS Paragon Plus Environment

Page 27 of 27

40

Laminar Flame Speed (cm/s)

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

35

Tu=400K

30

10 atm

25

20 atm

20 15

40 atm

10 line: detailed symbol: reduced

5 0 0.4

0.6

0.8

1.0

1.2

1.4

1.6

ER Fig. 12 Premixed laminar flame speeds versus equivalence ratios predicted by both the detailed and reduced mechanism At Tu=400 K (line: detailed, symbol: reduced)

27

ACS Paragon Plus Environment