Please wait a minute...
Frontiers in Energy

ISSN 2095-1701

ISSN 2095-1698(Online)

CN 11-6017/TK

邮发代号 80-972

2019 Impact Factor: 2.657

Frontiers in Energy  2021, Vol. 15 Issue (2): 405-420   https://doi.org/10.1007/s11708-021-0718-3
  本期目录
Development of a simplified n-heptane/methane model for high-pressure direct-injection natural gas marine engines
Jingrui LI1, Haifeng LIU1(), Xinlei LIU2, Ying YE1, Hu WANG1, Xinyan WANG3, Hua ZHAO3, Mingfa YAO1
1. State Key Laboratory of Engines, Tianjin University, Tianjin 300072, China
2. Clean Combustion Research Center, King Abdullah University of Science and Technology, Thuwal 23955-6900, Saudi Arabia
3. Brunel University London, Kingston Lane, Uxbridge, Middlesex UB8 3PH, UK
 全文: PDF(5511 KB)   HTML
Abstract

High-pressure direct-injection (HPDI) of natu- ral gas is one of the most promising solutions for future ship engines, in which the combustion process is mainly controlled by the chemical kinetics. However, the employment of detailed chemical models for the multi-dimensional combustion simulation is significantly expensive due to the large scale of the marine engine. In the present paper, a reduced n-heptane/methane model consisting of 35-step reactions was constructed using multiple reduction approaches. Then this model was further reduced to include only 27 reactions by utilizing the HyChem (Hybrid Chemistry) method. An overall good agreement with the experimentally measured ignition delay data of both n-heptane and methane for these two reduced models was achieved and reasonable predictions for the measured laminar flame speeds were obtained for the 35-step model. But the 27-step model cannot predict the laminar flame speed very well. In addition, these two reduced models were both able to reproduce the experimentally measured in-cylinder pressure and heat release rate profiles for a HPDI natural gas marine engine, the highest error of predicted combustion phase being 6.5%. However, the engine-out CO emission was over-predicted and the highest error of predicted NOx emission was less than 12.9%. The predicted distributions of temperature and equivalence ratio by the 35-step and 27-step models are similar to those of the 334-step model. However, the predicted distributions of OH and CH2O are significantly different from those of the 334-step model. In short, the reduced chemical kinetic models developed provide a high-efficient and dependable method to simulate the characteristics of combustion and emissions in HPDI natural gas marine engines.

Key wordshigh-pressure direct-injection    natural gas    chemical kinetics    combustion modelling    marine engine
收稿日期: 2020-06-06      出版日期: 2021-06-21
Corresponding Author(s): Haifeng LIU   
 引用本文:   
. [J]. Frontiers in Energy, 2021, 15(2): 405-420.
Jingrui LI, Haifeng LIU, Xinlei LIU, Ying YE, Hu WANG, Xinyan WANG, Hua ZHAO, Mingfa YAO. Development of a simplified n-heptane/methane model for high-pressure direct-injection natural gas marine engines. Front. Energy, 2021, 15(2): 405-420.
 链接本文:  
https://academic.hep.com.cn/fie/CN/10.1007/s11708-021-0718-3
https://academic.hep.com.cn/fie/CN/Y2021/V15/I2/405
Fig.1  
Number Reaction (k = ATBexp(−E/(RT))) A B E/(kcal?mol1) Ref.
R1 NC7H16+O2=C7H15+·HO2 7.00 × 1015 0 47380 modified from [42]
Reverse Arrhenius coefficients 4.62 × 1011 0.3 172.0/ [42]
R2 ·OH+NC7H16=C7H15+H2O 1.00 × 1014 0 3000 modified from [40]
R3 ·C7H15 + O2 = C7H15O2 1.50 × 1012 0 0 modified from [40]
Reverse Arrhenius coefficients 2.51 × 1013 0 27400.0/ [40]
R4 C7H15O2 = C7H14OOH 8.94 × 1011 0 19000 modified from [40]
Reverse Arrhenius coefficients 1.00 × 1011 0 11000.0/ [40]
R5 C7H14OOH+ O2 = OOC7H14OOH 2.20 × 1010 0 0 modified from [40]
Reverse Arrhenius coefficients 2.51 × 1013 0 27400.0/ [40]
R6 OOC7H14OOH=OC7H13OOH+·OH 1.20 × 1011 0 17000 [40]
R7 OC7H13OOH= C3H7CHO+ CH2O+ C2H3 + OH 9.00 × 1014 0 41100 modified from [42]
R8 C7H14OOH=C3H6+C3H7CHO+·OH 1.00 × 1013 0 25000 modified from [40]
R9 C3H7CHO+ ·OH= C3H7CO+ H2O 1.66 × 1012 0 0 [40]
R10 C3H7CO+O2=CO+C3H6+·HO2 6.03 × 1016 0 15000 [45]
R11 C7H15=C2H4+·C2H5+C3H6 7.04 × 1013 0 34600 modified from [42]
R12 ·C3H7=C2H4+·CH3 9.60 × 1013 0 30950 [45]
R13 ·C2H5+O2=C2H4+·HO2 2.00 × 1010 0 - 2200 [45]
R14 C2H4+·OH=C2H3+H2O 6.02 × 1013 0 5955 modified from [45]
R15 ·C2H3+O2=CH2O+HCO 4.00 × 1013 0 -251 [45]
R16 C3H6 + ·OH= ·C3H5 + H2O 3.12 × 106 2 -298 [45]
R17 ·C3H5+O2=·CH3+CH2O+CO 4.00 × 1012 0 0 [45]
R18 ·CH3+HO2=CH3O+·OH 5.00 × 1013 0 0 [45]
R19 ·CH3+·O=CH2O+·H 8.00 × 1013 0 0 [45]
R20 CH4 + O2 = ·CH3 + ·HO2 2.02 × 107 2.1 53210 modified from [47]
R21 CH4 + ·OH= ·CH3 + H2O 1.83 × 105 2.6 2190 modified from [47]
R22 CH4 + ·O= ·CH3 + ·OH 1.02 × 109 1.5 8604 modified from [47]
R23 CH4 + ·HO2 = ·CH3 + H2O2 1.13 × 101 3.7 21010 modified from [47]
R24 CH3O+ O2 = CH2O+ ·HO2 2.41 × 1014 0 5017 [46]
R25 CH2O+ ·OH= HCO+ H2O 3.47 × 109 1.2 -477 [46]
R26 HCO+ O2 = CO+ ·HO2 1.35 × 1013 0 400 [45]
R27 CO+ ·O+ M= CO2 + M 4.17 × 1014 0 3000 [46]
R28 CO+ ·OH= CO2 + ·H 3.51 × 107 1.2 69 [46]
Reverse Arrhenius coefficients: 9.45 × 1013 0 26089.0/ [46]
R29 H2O2 + M= 2·OH+ M 1.20 × 1017 0 45500 [46]
R30 ·HO2+·HO2=H2O2+O2 2.00 × 1012 0 0 [46]
R31 ·H+O2=·O+·OH 2.63 × 1016 -0.7 17040 [46]
Reverse Arrhenius coefficients: 7.87 × 1013 -0.3 -278.0/ [46]
R32 ·H+O2+M=·HO2+M 2.80 × 1018 - 0.9 0 [46]
Reverse Arrhenius coefficients: 4.66 × 1018 -0.9 48023.0/ [46]
R33 H2 + ·OH= H2O+ ·H 1.17 × 109 1.3 3626 [45]
Reverse Arrhenius coefficients: 1.24 × 1010 1.2 19092.0/ [45]
R34 ·O+H2=·OH+·H 5.06 × 104 2.7 6290 [45]
Reverse Arrhenius coefficients 2.37 × 104 2.7 4287.0/ [45]
R35 ·HO2+·OH=H2O+O2 7.50 × 1012 0 0 [45]
Tab.1  
Number Reaction (k = ATB exp( -E/lRT)) A B E/(kcal?mol-1) Ref.
R1 NC7H16 + O2 = ·C7H15 + ·HO2 1.00 × 1016 0 47380 [42], modified
Reverse Arrhenius coefficients: 4.62 × 1011 0.3 172.0/ [42]
R2 ·OH+ NC7H16 = ·C7H15 + H2O 5.00 × 1013 0 3000 [40], modified
R3 ·HO2 + NC7H16 = ·C7H15 + H2O2 9.01 × 1013 0 16000 [40], modified
Reverse Arrhenius coefficients: 6.31 × 101 0 8000.0/ [40]
R4 ·C7H15 = C2H4 + ·C2H5 + C3H6 7.04 × 1013 0 34600 [42], modified
R5 ·C2H5 + O2 = C2H4 + ·HO2 2.00 × 1010 - 0.9 - 2200 [45], modified
R6 C2H4 + ·OH= ·C2H3 + H2O 6.02 × 1014 0 5955 [45], modified
R7 ·C2H3 + O2 = CH2O+ HCO 4.00 × 1013 0 - 251 [45]
R8 C3H6 + ·OH= ·C3H5 + H2O 3.12 × 106 2 - 298 [45]
R9 C3H5 + O2 = ·CH3 + CH2O+ CO 4.00 × 1012 0 0 [45]
R10 ·CH3 + ·HO2 = CH3O+ ·OH 5.00 × 1013 0 0 [45]
R11 ·CH3 + ·O= CH2O+ ·H 8.00 × 1013 0 0 [45]
R12 CH4 + O2 = ·CH3 + ·HO2 2.02 × 107 2.1 53210 [47], modified
R13 CH4 + ·OH= ·CH3 + H2O 1.83 × 105 2.6 2190 [47], modified
R14 CH4 + ·O= ·CH3 + ·OH 1.02 × 109 1.5 8604 [47], modified
R15 CH4 + ·HO2 = ·CH3 + H2O2 1.13 × 101 3.7 21010 [47], modified
R16 CH3O+ O2 = CH2O+ HO2 2.41 × 1014 0 5017 [46]
R17 CH2O+ ·OH= HCO+ H2O 3.47 × 109 1.2 - 477 [46]
R18 HCO+ O2 = CO+ HO2 1.35 × 1013 0 400 [45]
R19 CO+ ·O+ M= CO2 + M 4.17 × 1014E+ 14 0 3000 [46]
R20 CO+ ·OH= CO2 + ·H 3.51 × 107 1.2 69 [46]
Reverse Arrhenius coefficients: 9.45 × 1013 0 26089.0/ [46]
R21 H2O2 + M= 2·OH+ M 1.20 × 1017 0 45500 [46]
R22 ·HO2 + ·HO2 = H2O2 + O2 2.00 × 1012 0 0 [46]
R23 ·H+ O2 = ·O+ ·OH 2.63 × 1016 - 0.7 17040 [46]
Reverse Arrhenius coefficients: 7.87 × 1013 - 0.3 - 278.0/ [46]
R24 ·H+ O2 + M= ·HO2 + M 2.80 × 1018 - 0.9 0 [46]
Reverse Arrhenius coefficients: 4.66 × 1018 - 0.9 48023.0/ [46]
R25 H2 + ·OH= H2O+ ·H 1.17 × 109 1.3 3626 [45]
Reverse Arrhenius coefficients: 1.24 × 1010 1.2 19092.0/ [45]
R26 ·O+ H2 = ·OH+ ·H 5.06 × 104 2.7 6290 [45]
Reverse Arrhenius coefficients: 2.37 × 104 2.7 4287.0/ [45]
R27 ·HO2 + ·OH= H2O+ O2 7.50 × 1012 0 0 [45]
Tab.2  
Fig.2  
Fig.3  
Fig.4  
Fig.5  
Fig.6  
Fig.7  
Fig.8  
Parameters Value/type
Type 4T50ME-GI
Number of cylinders 4
Bore/stroke 500 mm/2200 mm
Connecting rod 2885 mm
Geometric compression ratio 18.14
Engine speed at MCR(maximum continuous rating) 123 r/min
Power at MCR 7050 kW
IMEP(indicated mean effective pressure) at MCR 2.0 MPa
Turbocharger MAN TCA55-VTA
Tab.3  
334-step 35-step 27-step
CPU-hour/h 96 49 47
Tab.4  
Fig.9  
27-step (error) 35-step (error) 334-step (error) Experiment
CA10 (°CA ATDC) 3.7 (1.08%) 3.8 (3.68%) 4.0 (8.5%) 3.66
CA50 (°CA ATDC) 10.8 (5.56%) 10.7 (6.54%) 11.1 (2.7%) 11.4
pmax/MPa 16.76 (0.36%) 16.89 (0.42%) 16.73 (0.53%) 16.82
pmax phase (°CA ATDC) 11.2 (2.1%) 11.2 (2.1%) 10.9 (4.72%) 11.44
ISFC/(g·kW-1·h-1) 164.0 (0.49%) 163.5 (0.18%) 164.6 (0.86%) 163.2
Power output /kW 5331.8 (0.01%) 5346.7 (0.29%) 5311.6 (0.36%) 5331
Tab.5  
27-step (error) 35-step (error) 334-step (error) Experiment
NOx/(g·kW-1·h-1) 10.68 (11.4%) 10.54 (12.9%) 11.37 (4.93%) 11.9
HC/(g·kW-1·h-1) 0.0045 0.0039 0.0057
CO/(g·kW-1·h-1) 0.0237 (700%) 0.0092 (1900%) 0.0351 (441%) 0.19
CO2/(g·kW-1·h-1) 415.7 415.0 416.9
Soot/(g·kW-1·h-1) 0.243 0.241 0.0238
Tab.6  
Fig.10  
Fig.11  
Fig.12  
Fig.13  
Fig.14  
Fig.15  
CA10 The crank angle that releases 10% of the heat
CA50 The crank angle that releases 50% of the heat
CI Compression ignition
CFD Computational fluid dynamic
HPDI High-pressure direct-injection
HRR Heat release rate
HTC High-temperature combustion
HyChem Hybrid Chemistry
ICEs Internal combustion engines
ISFC Indicated specific fuel consumption
LTC Low-temperature combustion
pmax Maximum pressure
pmax phase Phase of maximum pressure
RCM Rapid compression engine
  
1 M Y Luo, D Liu. Kinetic analysis of ethanol and dimethyl ether flames with hydrogen addition. International Journal of Hydrogen Energy, 2017, 42(6): 3813–3823
https://doi.org/10.1016/j.ijhydene.2016.08.208
2 Y Y Ying, D Liu. Detailed influences of chemical effects of hydrogen as fuel additive on methane flame. International Journal of Hydrogen Energy, 2015, 40(9): 3777–3788
https://doi.org/10.1016/j.ijhydene.2015.01.076
3 M M Abdelaal, A H Hegab. Combustion and emission characteristics of a natural gas-fueled diesel engine with EGR. Energy Conversion and Management, 2012, 64: 301–312
https://doi.org/10.1016/j.enconman.2012.05.021
4 R Florea, G D Neely, Z Abidin, et al.. Efficiency and emissions characteristics of partially premixed dual-fuel combustion by co-direct injection of NG and diesel fuel (DI2). In: SAE 2016 World Congress and Exhibition, Detroit, USA, 2016, 121606
5 H Q Wei, J Y Qi, L Zhou, et al.. Ignition characteristics of methane/n-heptane fuel blends under engine-like conditions. Energy & Fuels, 2018, 32(5): 6264–6277
https://doi.org/10.1021/acs.energyfuels.7b04128
6 Y Z He, Y D Wang, C Grégoire, et al.. Ignition characteristics of dual-fuel methane-n-hexane-oxygen-diluent mixtures in a rapid compression machine and a shock tube. Fuel, 2019, 249: 379–391
https://doi.org/10.1016/j.fuel.2019.03.105
7 G P McTaggart-Cowan, J Huang, S Munshi. Impacts and mitigation of varying fuel composition in a natural gas heavy-duty engine. SAE International Journal of Engines, 2017, 10(4): 1506–1517
https://doi.org/10.4271/2017-01-0777
8 G P McTaggart-Cowan, H L Jones, S N Rogak, et al.. The Effects of high-pressure injection on a compression-ignition, direct injection of natural gas engine. Journal of Engineering for Gas Turbines and Power, 2007, 129(2): 579–588
https://doi.org/10.1115/1.2432894
9 A M L M Wagemakers, C A J Leermakers. Review on the effects of dual-fuel operation, using diesel and gaseous fuels, on emissions and performance. In: SAE 2012 World Congress and Exhibition, Detroit, USA, 2012, 92224
10 K Zeng, Z H Huang, B Liu, et al.. Combustion characteristics of a direct-injection natural gas engine under various fuel injection timings. Applied Thermal Engineering, 2006, 26(8–9): 806–813
https://doi.org/10.1016/j.applthermaleng.2005.10.011
11 H F Liu, J R Li, J T Wang, et al.. Effects of injection strategies on low-speed marine engines using the dual fuel of high-pressure direct-injection natural gas and diesel. Energy Science & Engineering, 2019, 7(5): 1994–2010
https://doi.org/10.1002/ese3.406
12 H M Cho, B Q He. Spark ignition natural gas engines—a review. Energy Conversion and Management, 2007, 48(2): 608–618
https://doi.org/10.1016/j.enconman.2006.05.023
13 R P Roethlisberger, D Favrat. Investigation of the prechamber geometrical configuration of a natural gas spark ignition engine for cogeneration: part I. numerical simulation. International Journal of Thermal Sciences, 2003, 42(3): 223–237
https://doi.org/10.1016/S1290-0729(02)00023-6
14 J B Zheng, J H Wang, Z B Zhao, et al.. Effect of equivalence ratio on combustion and emissions of a dual-fuel natural gas engine ignited with diesel. Applied Thermal Engineering, 2019, 146: 738–751
https://doi.org/10.1016/j.applthermaleng.2018.10.045
15 Z H Huang, S Shiga, T Ueda, et al.. Effect of fuel injection timing relative to ignition timing on the natural-gas direct-injection combustion. Journal of Engineering for Gas Turbines and Power, 2003, 125(3): 783–790
https://doi.org/10.1115/1.1563243
16 E Demosthenous, G Borghesi, E Mastorakos, et al.. Direct numerical simulations of premixed methane flame initiation by pilot n-heptane spray autoignition. Combustion and Flame, 2016, 163: 122–137
https://doi.org/10.1016/j.combustflame.2015.09.013
17 E Demosthenous, E Mastorakos, R Stewart Cant. Direct numerical simulations of dual-fuel non-premixed autoignition. Combustion Science and Technology, 2016, 188(4–5): 542–555
https://doi.org/10.1080/00102202.2016.1139391
18 H J Curran, P Gaffuri, W J Pitz, et al.. A comprehensive modeling study of n-heptane oxidation. Combustion and Flame, 1998, 114(1–2): 149–177
https://doi.org/10.1016/S0010-2180(97)00282-4
19 K W Zhang, C Banyon, J Bugler, et al.. An updated experimental and kinetic modeling study of n-heptane oxidation. Combustion and Flame, 2016, 172: 116–135
https://doi.org/10.1016/j.combustflame.2016.06.028
20 H K Ciezki, G Adomeit. Shock-tube investigation of self-ignition of n-heptane-air mixtures under engine relevant conditions. Combustion and Flame, 1993, 93(4): 421–433
https://doi.org/10.1016/0010-2180(93)90142-P
21 D F Davidson, R K Hanson. Interpreting shock tube ignition data. International Journal of Chemical Kinetics, 2004, 36(9): 510–523
https://doi.org/10.1002/kin.20024
22 K Fieweger, R Blumenthal, G Adomeit. Self-ignition of S.I. engine model fuels: a shock tube investigation at high pressure. Combustion and Flame, 1997, 109(4): 599–619
https://doi.org/10.1016/S0010-2180(97)00049-7
23 B M Gauthier, D F Davidson, R K Hanson. Shock tube determination of ignition delay times in full-blend and surrogate fuel mixtures. Combustion and Flame, 2004, 139(4): 300–311
https://doi.org/10.1016/j.combustflame.2004.08.015
24 E J Hu, X T Li, X Meng, et al.. Laminar flame speeds and ignition delay times of methane–air mixtures at elevated temperatures and pressures. Fuel, 2015, 158: 1–10
https://doi.org/10.1016/j.fuel.2015.05.010
25 J Huang, P G Hill, W K Bushe, S R Munshi. Shock-tube study of methane ignition under engine-relevant conditions: experiments and modeling. Combustion and Flame, 2004, 136(1–2): 25–42
https://doi.org/10.1016/j.combustflame.2003.09.002
26 J J Liang, Z H Zhang, G S Li, et al.. Experimental and kinetic studies of ignition processes of the methane–n-heptane mixtures. Fuel, 2019, 235: 522–529
https://doi.org/10.1016/j.fuel.2018.08.041
27 Y T Wu, Y Liu, C L Tang, et al.. Ignition delay times measurement and kinetic modeling studies of 1-heptene, 2-heptene and n-heptane at low to intermediate temperatures by using a rapid compression machine. Combustion and Flame, 2018, 197: 30–40
https://doi.org/10.1016/j.combustflame.2018.07.007
28 S G Davis, C K Law, H Wang. Propene pyrolysis and oxidation kinetics in a flow reactor and laminar flames. Combustion and Flame, 1999, 119(4): 375–399
https://doi.org/10.1016/S0010-2180(99)00070-X
29 T J Held, A J Marchese, F L Dryer. A semi-empirical reaction mechanism for n-heptane oxidation and pyrolysis. Combustion Science and Technology, 1997, 123(1–6): 107–146
https://doi.org/10.1080/00102209708935624
30 Z H Zhang, H Zhao, L Cao, et al.. Kinetic effects of n-heptane addition on low and high temperature oxidation of methane in a jet-stirred reactor. Energy & Fuels, 2018, 32(11): 11970–11978
https://doi.org/10.1021/acs.energyfuels.8b03124
31 J F Griffiths, P A Halford-Maw, D J Rose. Fundamental features of hydrocarbon autoignition in a rapid compression machine. Combustion and Flame, 1993, 95(3): 291–306
https://doi.org/10.1016/0010-2180(93)90133-N
32 R Minetti, M Carlier, M Ribaucour, et al.. A rapid compression machine investigation of oxidation and auto-ignition of n-heptane: measurements and modeling. Combustion and Flame, 1995, 102(3): 298–309
https://doi.org/10.1016/0010-2180(94)00236-L
33 S Tanaka, F Ayala, J C Keck. A reduced chemical kinetic model for HCCI combustion of primary reference fuels in a rapid compression machine. Combustion and Flame, 2003, 133(4): 467–481
https://doi.org/10.1016/S0010-2180(03)00057-9
34 K P Grogan, S Scott Goldsborough, M Ihme. Ignition regimes in rapid compression machines. Combustion and Flame, 2015, 162(8): 3071–3080
https://doi.org/10.1016/j.combustflame.2015.03.020
35 T Aroonsrisopon, V Sohm, P Werner, et al.. An investigation into the effect of fuel composition on HCCI combustion characteristics. In: Powertrain and Fluid Systems Conference and Exhibition, San Diego, USA, 2002, 90919
36 T Aroonsrisopon, P Werner, J O Waldman, et al.. Expanding the HCCI operation with the charge stratification. In: 2004 SAE World Congress, Detroit, USA, 2004, 90264
37 J Andrae, D Johansson, P Björnbom, et al.. Co-oxidation in the auto-ignition of primary reference fuels and n-heptane/toluene blends. Combustion and Flame, 2005, 140(4): 267–286
https://doi.org/10.1016/j.combustflame.2004.11.009
38 K Seshadri. Numerical and asymptotic studies of the structure of stoichiometric and lean premixed heptane flames. Combustion and Flame, 1997, 108(4): 518–536
https://doi.org/10.1016/S0010-2180(96)00118-6
39 N Peters, G Paczko, R Seiser, et al.. Temperature cross-over and non-thermal runaway at two-stage ignition of n-heptane. Combustion and Flame, 2002, 128(1–2): 38–59
https://doi.org/10.1016/S0010-2180(01)00331-5
40 H L Li, D L Miller, N P Cernansky. Development of a reduced chemical kinetic model for prediction of preignition reactivity and autoignition of primary reference fuels. In: International Congress and Exposition, Detroit, USA, 1996, 90323
41 W Su, H Huang. Development and calibration of a reduced chemical kinetic model of n-heptane for HCCI engine combustion. Fuel, 2005, 84(9): 1029–1040
https://doi.org/10.1016/j.fuel.2005.01.015
42 A Patel, S C Kong, R D Reitz. Development and validation of a reduced reaction mechanism for HCCI Engine Simulations. In: 2004 SAE World Congress, Detroit, USA, 2004, 90264
43 F Maroteaux, L Noel. Development of a reduced n-heptane oxidation mechanism for HCCI combustion modeling. Combustion and Flame, 2006, 146(1–2): 246–267
https://doi.org/10.1016/j.combustflame.2006.03.006
44 S Lapointe, K Zhang, M J McNenly. Reduced chemical model for low and high-temperature oxidation of fuel blends relevant to internal combustion engines. Proceedings of the Combustion Institute, 2019, 37(1): 789–796
https://doi.org/10.1016/j.proci.2018.06.139
45 J Gustavsson, V I Golovitchev. Spray combustion simulation based on detailed chemistry approach for diesel fuel surrogate model. In: 2003 JSAE/SAE International Spring Fuels and Lubricants Meeting, Yokohama, Japan, 2003, 90284
46 J C Zheng, W Y Yang, D L Miller, et al.. A skeletal chemical kinetic model for the HCCI combustion process. In: SAE 2002 World Congress, Detroit, USA, 2002, 90920
47 Y Ra, R D Reitz. A reduced chemical kinetic model for IC engine combustion simulations with primary reference fuels. Combustion and Flame, 2008, 155(4): 713–738
https://doi.org/10.1016/j.combustflame.2008.05.002
48 K A Heufer, H Olivier. Determination of ignition delay times of different hydrocarbons in a new high pressure shock tube. Shock Waves, 2010, 20(4): 307–316
https://doi.org/10.1007/s00193-010-0262-2
49 H Wang, R Xu, K Wang, et al.. A physics-based approach to modeling real-fuel combustion chemistry- I. evidence from experiments, and thermodynamic, chemical kinetic and statistical considerations. Combustion and Flame, 2018, 193: 502–519
https://doi.org/10.1016/j.combustflame.2018.03.019
50 K Fieweger, R Blumenthal, G Adomeit. Shock-tube investigations on the self-ignition of hydrocarbon-air mixtures at high pressures. Symposium (International) on Combustion, 1994, 25(1): 1579–1585
51 K Kumar, J E Freeh, C J Sung, et al.. Laminar flame speeds of preheated iso-octane/O2/N2 and n-heptane/O2/N2 mixtures. Journal of Propulsion and Power, 2007, 23(2): 428–436
https://doi.org/10.2514/1.24391
52 A J Smallbone, W Liu, C K Law, et al.. Experimental and modeling study of laminar flame speed and non-premixed counterflow ignition of n-heptane. Proceedings of the Combustion Institute, 2009, 32(1): 1245–1252
https://doi.org/10.1016/j.proci.2008.06.213
53 C S Ji, E Dames, Y L Wang, et al.. Propagation and extinction of premixed C5–C12 n-alkane flames. Combustion and Flame, 2010, 157(2): 277–287
https://doi.org/10.1016/j.combustflame.2009.06.011
54 A P Kelley, A J Smallbone, D L Zhu, et al.. Laminar flame speeds of C5 to C8 n-alkanes at elevated pressures: experimental determination, fuel similarity, and stretch sensitivity. Proceedings of the Combustion Institute, 2011, 33(1): 963–970
https://doi.org/10.1016/j.proci.2010.06.074
55 P Dirrenberger, P A Glaude, R Bounaceur, et al.. Laminar burning velocity of gasolines with addition of ethanol. Fuel, 2014, 115: 162–169
https://doi.org/10.1016/j.fuel.2013.07.015
56 G S Li, J J Liang, Z H Zhang, et al.. Experimental investigation on laminar burning velocities and markstein lengths of premixed methane-n-heptane-air mixtures. Energy & Fuels, 2015, 29(7): 4549–4556
https://doi.org/10.1021/acs.energyfuels.5b00355
57 D Bradley. Burning velocities, markstein lengths, and flame quenching for spherical methane-air flames: a computational study. Combustion and Flame, 1996, 104(1–2): 176–198
https://doi.org/10.1016/0010-2180(95)00115-8
58 M Hassan. Measured and predicted properties of laminar premixed methane/air flames at various pressures. Combustion and Flame, 1998, 115(4): 539–550
https://doi.org/10.1016/S0010-2180(98)00025-X
59 C M Vagelopoulos, F N Egolfopoulos. Direct experimental determination of laminar flame speeds. Symposium (International) on Combustion, 1998, 27(1): 513–519
60 I V Dyakov, A A Konnov, J D Ruyck, et al.. Measurement of adiabatic burning velocity in methane-oxygen-nitrogen mixtures. Combustion Science and Technology, 2001, 172(1): 81–96
https://doi.org/10.1080/00102200108935839
61 T Tahtouh, F Halter, C Mounaïm-Rousselle. Measurement of laminar burning speeds and Markstein lengths using a novel methodology. Combustion and Flame, 2009, 156(9): 1735–1743
https://doi.org/10.1016/j.combustflame.2009.03.013
62 P Dirrenberger, H Le Gall, R Bounaceur, et al.. Measurements of laminar flame velocity for components of natural gas. Energy & Fuels, 2011, 25(9): 3875–3884
https://doi.org/10.1021/ef200707h
63 K J Richards, P K Senecal, E. PomraningCONVERGE Manual (Version 2.3), 2016
64 L R Juliussen, M J Kryger, A Andreasen. MAN B&W ME‐GI engines recent research & results. In: Proceedings of the International Symposium on Marine Engineering (ISME), 2011
65 G A Lavoie, J B Heywood, J C Keck. Experimental and theoretical study of nitric oxide formation in internal combustion engines. Combustion Science and Technology, 1970, 1(4): 313–326
https://doi.org/10.1080/00102206908952211
66 H Hiroyasu, T Kadota, M Arai. Development and use of a spray combustion modeling to predict diesel engine efficiency and pollutant emissions (part 1 combustion modeling). Bulletin of the JSME, 1983, 26(214): 569–575
https://doi.org/10.1299/jsme1958.26.569
67 J R Li, J T Wang, T Liu, et al.. An investigation of the influence of gas injection rate shape on high-pressure direct-injection natural gas marine engines. Energies, 2019, 12(13): 2571
https://doi.org/10.3390/en12132571
68 C R Larson. Injection study of a diesel engine fueled with pilot-ignited, directly-injected natural gas. Dissertation for the Doctoral Degree. Vancouver: University of British Columbia, 2003
69 M Jud, G Fink, T Sattelmayer. Predicting ignition and combustion of a pilot ignited natural gas jet using numerical simulation based on detailed chemistry. In: ASME 2017 Internal Combustion Engine Division Fall Technical Conference (ICEF 2017), Seattle, USA, 132256
Viewed
Full text


Abstract

Cited

  Shared   
  Discussed