bura.brunel.ac.uk · web viewthe equations were solved by the pressure-implicit with splitting of...

72
Title: Effect of Piston Shapes and Fuel Injection Strategies on Stoichiometric Stratified Flame Ignition (SFI) Hybrid Combustion in a PFI/DI Gasoline Engine by Numerical Simulations Author names and affiliations: Xinyan Wang a , Hua Zhao a, b , Hui Xie a a State Key Laboratory of Engines, Tianjin University, China. b Centre for Advanced Powertrain and Fuels, Brunel University London, UK Corresponding author: Hui Xie, State Key Laboratory of Engines, Tianjin University, P.R. China, Weijin Road 92, Nankai District, Tianjin, China, 300072. Tel: +86-22-27406842-8009, Mobil: +86-13001344531, Fax: +86-22+27406842-8009. Email: [email protected] . 1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 1 2

Upload: dangthu

Post on 25-Jun-2018

213 views

Category:

Documents


0 download

TRANSCRIPT

Title:

Effect of Piston Shapes and Fuel Injection Strategies on Stoichiometric Stratified Flame

Ignition (SFI) Hybrid Combustion in a PFI/DI Gasoline Engine by Numerical Simulations

Author names and affiliations:

Xinyan Wang a, Hua Zhao a, b, Hui Xie a

a State Key Laboratory of Engines, Tianjin University, China.

b Centre for Advanced Powertrain and Fuels, Brunel University London, UK

Corresponding author:

Hui Xie, State Key Laboratory of Engines, Tianjin University, P.R. China, Weijin Road 92,

Nankai District, Tianjin, China, 300072. Tel: +86-22-27406842-8009, Mobil: +86-

13001344531, Fax: +86-22+27406842-8009. Email: [email protected].

1

1

2

3

4

5

6

7

8

9

10

11

12

12

Abstract:

In this research, the stratified flame ignition (SFI) hybrid combustion process was proposed

to enhance the control of SI-CAI hybrid combustion and moderate the maximum pressure

rise rate (PRRmax) by the combination of port fuel injection (PFI) and direct injection (DI).

The effect of the stratified flame formed by different piston shapes, start of direct injection

(SOI) timings and direct injection ratios (rDI) on the stoichiometric SFI hybrid combustion

and heat release process was studied using the three-dimensional computational fluid

dynamics (3-D CFD) simulations. The spark ignited flame propagation near the spark plug

and the auto-ignition heat release process of the diluted mixture were modelled in the

framework of 3-Zones Extended Coherent Flame Model (ECFM3Z) by the extended coherent

flame model and tabulated auto-ignition chemistry of a 4-component gasoline surrogate,

respectively. The operating load of indicated mean effective pressure (IMEP) 3.6 bar was

selected to represent a typical part-load operation. The sweep of the spark timing (ST) was

performed for different pistons, SOI timings and direct injection ratios. The SFI hybrid

combustion process with the same combustion phasing was investigated in details. The

optimal stratified mixture pattern, characterized with the central rich mixture around spark

plug and stratified lean mixture at the peripheral region, formed by the newly designed Piston

A and B effectively lowers the PRRmax with a slight deterioration of IMEP. The later SOI

timing advances the crank angle of 50% total heat release (CA50) and significantly reduces

the PRRmax with a little deterioration of IMEP. As the direct injection ratio is increased, both

the PRRmax and IMEP decrease. During the SFI hybrid heat release process, spark timing is

effective to control CA50, IMEP and PRRmax regardless the piston shapes, SOI timings and

direct injection ratios. However, the sensitivity of SFI hybrid combustion to the stratified

mixture varies with the spark timing. The reduction of the PRRmax caused by the stratified

flame enables the advance of spark timing to achieve maximum IMEP.

2

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

34

Keywords: computational fluid dynamics, hybrid combustion, stratified mixture, controlled

auto-ignition, gasoline engine

1. Introduction

Controlled auto-ignition (CAI) combustion has been subject to extensive studies in the past

decade because of its significant advantage to reduce the NOx emissions and increase fuel

efficiency [1, 2]. The CAI combustion is characterized with multi-site auto-ignition at the top

dead center (TDC), as shown in Fig. 1 (a). However, the CAI combustion process is difficult

to control and the operation range of pure CAI combustion is relatively narrow, which

inhibits the practical application of this efficient combustion mode [3]. In order to enhance

the control of CAI combustion, the flame propagation induced by spark ignition (SI) has been

used to control the subsequent auto-ignition [4, 5], as shown in Fig. 1 (b). Compared to the

traditional SI combustion, the SI-CAI hybrid combustion, also known as spark assisted

compression ignition (SACI), shows a higher thermal efficiency and lower NOx emission,

while it obtains a lower maximum rate of pressure rise and higher load operation range

compared to the pure CAI combustion [6]. In addition, this hybrid combustion concept can be

used to bridge pure SI mode and CAI mode [3, 7-10].

Fig.1. Schematic diagram of (a) CAI combustion and (b) SI-CAI hybrid combustion.

3

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

56

In order to control the SI-CAI hybrid combustion, different control strategies, including spark

timing [11], intake temperature [11-13], wall temperature [13], in-cylinder flow [14] and

dilution composition [11, 15], have been studied to understand their effect on hybrid

combustion process. However, it was found that the optimal high load operation is limited by

the severe knock or unacceptable pressure rise rate (PPR) for the stoichiometric hybrid

combustion with the homogeneous charge [9, 15-17]. In order to moderate the maximum

pressure rise rate of hybrid combustion, Szybist el al. [16] used a combination of intake valve

closing (IVC) control and spark timing control to achieve IMEP up to 7.5 bar. In this

strategy, the spark timing allows the start of combustion to be controlled, while retarding the

intake valve closing angle reduces the effective compression ratio. Li et al. [15] proposed an

optimized positive valve overlapping (PVO) strategy in which the PVO was formed mainly

by advancing intake valve timing and by retarding exhaust valve timing to achieve the stable

SACI combustion in the range of 5–9 bar IMEP, with significant improvements in fuel

economy, pumping loss and NOx emission.

The fuel stratification accompanied with the enrichment of the central region would form the

diluted mixture in the peripheral region, which provides a natural resistance to severe auto-

ignitions leading to unacceptable PRR. The fuel stratification has been widely studied to

overcome the overly rapid combustion at high-load conditions in homogenous charge

compression ignition (HCCI) engines [18-23]. In terms of the SI-CAI hybrid combustion, the

fuel stratification was also proposed to evaluate its potential to extend the high load limit.

Yoshizawa et al. [24] developed a two-step combustion concept to expand the high load

operation in a gasoline compression ignition engine. In this two-step combustion concept, the

spark plug would be used only for igniting a small quantity of a fuel rich mixture at the bore

center and the remaining fuel would be ignited by auto-ignition. The core of this concept is to

create the separate heat release periods to reduce the PRR. Yun et al. [17] obtained the

4

58

59

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

81

82

78

identical combustion phasing at high load spark-assisted stoichiometric HCCI operation with

the lower ringing by retarding injection timing, advancing spark timing and employing higher

exhaust gas recirculation (EGR) in a direct injection gasoline engine. They also found the

double injection strategy could be used to reduce ringing and the split ratio has a significant

effect on the trade-off between ringing and combustion stability. It was reported by Li et al.

[25] that the stable stoichiometric assisted spark stratified compression ignition (ASSCI)

combustion was achieved at medium-high load from 3.5 bar to 6.5 bar IMEP using the fuel

direct injection before spark timing. Persson et al. [26] investigated the effect of fuel

stratification on SACI with overall diluted ethanol/air mixture using high speed fuel planar

laser-induced fluorescence (PLIF). They found an increased stratification would augment the

combustion duration although the combustion phasing remained constant. Meanwhile, the

optical results indicated that ignition occurred in the mixing zone between the rich and the

leaner regions instead of the most fuel-rich regions. In addition to the above research at high

load operation, the spark ignition with the stratified charge was used to control the

combustion phasing of HCCI combustion and successfully expand the operational range

towards lower loads, from 2.7 bar IMEP to 1.5 bar [27]. However, these previous works had

been carried out without considering the conditions to promote the formation of a stable

flame and its interactions with the subsequent auto-ignition reactions of the diluted mixture.

The dual-injection strategy utilizes both the port fuel injection (PFI) and direct injection (DI)

and has been successfully applied to different combustion concept to enhance the control and

fuel efficiency of the combustion [21, 22, 26, 28]. In this study, the stratified flame ignition

(SFI) hybrid combustion was proposed to generate a stable flame kernel and hence control

the heat release process of SI-CAI hybrid combustion by dual-injection strategy at stratified

charge operations. The stratified flame kernel is formed around the spark plug by fuel

stratification in the same way as a stratified charge DI gasoline engine through the

5

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

104

105

106

107

910

appropriate direct injection strategy in combination with a suitable bowl piston [29-33]. The

piston shape should be well designed to induce the direct injected fuel to the vicinity of spark

plug. In addition to the piston shapes, both the start of injection timing [20, 34-36] and the

direct injection ratio [20, 26, 31] play important roles on controlling in-cylinder fuel

stratification patterns and combustion process. Our previous study revealed the effect of

piston shapes, start of injection (SOI) timings and direct injection ratios on the in-cylinder

conditions, including in-cylinder flow conditions, fuel stratification patterns and thermal

conditions [37].

In the proposed SFI hybrid combustion concept, the central rich mixture offered by direct

injection can be used to stabilize the early flame propagation process, while the peripheral

diluted lean mixture controlled by the PFI can be consumed by the auto-ignition combustion

mode with a relatively lower heat release rate compared to the stoichiometric auto-ignition to

moderate the pressure rise rate. In addition, the gradually stratified mixture from the central

rich region to the peripheral lean region could be used to manage the whole combustion

process. The direct injection can provide fast response on controlling the patterns and degree

of the in-cylinder fuel stratification, which has the potential to enable the real-time control on

the SFI hybrid combustion. In this study, the effect of the fuel stratification formed by

different piston shapes, SOI timings and direct injection ratios on the SFI hybrid combustion

was investigated in detail to understand the fundamental performance of SFI hybrid

combustion. The stoichiometric fuel/air mixture was applied to enable the emission reduction

with the conventional three-way catalysts. The current study would present the fundamental

knowledge on the SFI hybrid combustion characteristics with different patterns and degrees

of fuel stratification. The results would provide effective guidance on selecting the optimal

bowl piston shape and fuel injection strategy to control the SI-CAI hybrid combustion in the

future experimental investigations.

6

108

109

110

111

112

113

114

115

116

117

118

119

120

121

122

123

124

125

126

127

128

129

130

131

132

1112

2. Methodology

2.1 SFI hybrid Combustion Modelling

The modelling and simulation of SFI hybrid combustion was implemented in the STAR-CD

TM software. Reynolds-averaged Navier Stokes (RANS) simulations were applied in this

study. RNG k-ε model was adopted as the turbulence model. The enthalpy conservation

equation for the fluid mixture was used to calculate the heat transfer [38]. The simulation of

wall heat transfer was implemented by Angelberger wall function [39]. The equations were

solved by the Pressure-implicit with splitting of operators (PISO) algorithm. A set of models

depicting spray and hybrid combustion process are adopted to simulate the stratified flame

ignition (SFI) hybrid combustion process.

2.1.1 Spray modeling

A multi-hole injector was selected in this study because of its superior performance and low

cost [40]. In order to calculate the velocity of the liquid fuel as it exits the nozzle and enters

the combustion chamber, the MPI2 nozzle model [41] was employed in this study. Reitz-

Diwakar model [42] was adopted to simulate the atomization and break-up of the liquid

droplets. O’ Rourke model [43] was used to predict the collisions between fuel droplets. In

order to consider the wall impingement, Bai model [44] was applied. Fig. 2 shows the

schematic diagram of the above spray modelling. The model parameters were well tuned and

the simulation results of the spray process showed good agreement with the corresponding

optical visualizations. The detailed validation results can be found in Ref. [37].

7

133

134

135

136

137

138

139

140

141

142

143

144

145

146

147

148

149

150

151

152

1314

Fig.2. Schematic diagram of spray modelling.

2.1.2 Combustion modeling

In order to cover both the turbulent mixing effects and chemical kinetics in the stratified

flame ignition (SFI) hybrid combustion, models for the premixed flame propagation and

auto-ignition combustion were constructed to model the hybrid combustion process. The

ECFM3Z combustion model [45] was adopted as the framework of the SFI hybrid

combustion model in this study. ECFM3Z uses a flame surface density equation to describe

the flame propagation process, and the gas state in ECFM3Z model is represented by a pure

fuel zone, a pure air plus possible residual gas zone and a mixed zone. The auto-ignition of

the unburned gas was predicted by the tabulated chemistry approach [46]. In order to

construct the tabulated database used for the tabulated chemistry approach, a reduced

gasoline surrogate mechanism [47] with 312 species was adopted for the chemical kinetics

calculations under various thermodynamic and dilution conditions.

With these models, the reaction regime of each cell is determined by two parameters: the

flame surface density and the auto-ignition tendency. The available fuel in the cell will be

consumed by the flame according to the ECFM3Z model when the flame surface density of a

certain cell is greater than 0. By contrast, the available fuel in the cell will be consumed by

8

153

154

155

156

157

158

159

160

161

162

163

164

165

166

167

168

169

170

1516

auto-ignition combustion according to the tabulated chemistry approach if the auto-ignition

tendency of a certain cell achieves 1. Therefore, this hybrid combustion model is capable to

predict the stratified flame ignition (SFI) hybrid combustion. The detailed modelling process

can be found in Ref. [48].

2.2 Experimental engine

In order to validate the CFD modelling, the engine experiments were carried out on a single

cylinder gasoline engine. Table 1 shows the basic engine specifications. The engine was

developed basing on a Ricardo Hydra single cylinder block. A specially designed cylinder

head equipped with a 4-variable valve actuation system (4VVAS) was mounted on both

intake and exhaust camshafts to enable the continuous adjustment of intake/exhaust valve lift

and the valve timing. The engine was coupled to an electric dynamometer to maintain the

constant engine speed. In order to realize the precise control of the air/fuel ratio, a linear

oxygen sensor with the ±1.5% accuracy was mounted in the exhaust pipe. The Kistler 6125B

piezoelectric transducer coupled with 5011B charge amplifier was used to monitor the in-

cylinder pressure. A laminar flow meter with the accuracy of ±1% was used to measure the

amount of airflow. The coolant and lubricant oil temperatures were controlled at 80 ± 1 ºC

and 55 ± 1 ºC respectively.

An operating load of IMEP 3.6 bar was selected as a typical part-load operation. Table 2

shows the detail operation conditions. The negative valve overlapping (NVO) strategy was

adopted in this study to realize the internal exhaust gas recirculation (iEGR). The external

exhaust gas recirculation (eEGR) was also used to achieve the stable SFI hybrid combustion.

The other experimental details could be found in Ref. [3] and [9].

9

171

172

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

191

192

1718

Table 1 Engine specifications.

Bore 86 mm

Stroke 86 mm

Displacement 0.5 L

Compression ratio 10.66

Combustion chamber Pent roof / 4 valves

Fuel injection PFI/DI

Fuel Gasoline 93 RON

Intake pressure Naturally aspirated

Throttle WOT

Table 2 Operation conditions.

IMEP 3.6 bar

Engine speed 1500 r/min

Piston shape Flat piston

Exhaust valve open (EVO) 177 °CA ATDC a

Exhaust valve close (EVC) 254 °CA ATDC a

Exhaust valve lift (EL) 1.9 mm

Intake valve open (IVO) 494 °CA ATDC a

Intake valve close (IVC) 603 °CA ATDC a

Intake valve lift (IL) 5.0 mm

Spark Timing 685 ºCA ATDC a

Fuel injection PFI

Fueling rate 13.4 mg/cycle

Fuel/air equivalence ratio 1

eEGR 0.08

iEGR 0.36a 720 ºCA is defined as the combustion top dead centre.

2.3 Simulations setup and validation

In this study, the effect of the fuel stratification formed by different bowl piston shapes, start

of direct injection (SOI) timings and direct injection ratios (rDI) on the stoichiometric SFI

hybrid combustion was modelled and analysed. Three piston bowl designs, as shown in Fig.

10

193

194

195

196

197

198

199

1920

3, were designed from the origin Flat Piston (FP) in order to achieve the stratified mixture

[37]. Compared to Piston A (PA), Piston B (PB) has a larger bowl diameter. Piston A and B

have been manufactured and ready for engine tests. Based on designs of Piston A and Piston

B, a deeper piston bowl was designed in Piston C (PC). The moving meshes for Flat Piston,

Pistons A, B and C were generated in ES-ICE TM using the mapping method. The engine

mesh with Piston A is shown in Fig. 4 as an example. All the four meshes have similar grid

size (around 0.8 mm).

Piston A Piston B Piston C

Fig. 3. Newly designed bowl pistons.

Fig. 4. Engine mesh with Piston A.

The detailed computational conditions of the simulations in this study are shown in Table 3.

The sweep of the spark timing (ST) was performed for all pistons, SOI timings and direct

injection ratios. The simulations were carried out from IVO to the end of combustion. In all

simulations, the inlet temperature and pressure were fixed at 355 K and 0.99 bar. The initial

in-cylinder temperature and pressure were 571 K and 0.49 bar, respectively. These conditions

were based on the one-dimensional engine simulations using GT-power. The GT-Power

engine model was built according to the real engine parameters and well validated [49]. The

11

200

201

202

203

204

205

206

207

208209

210

211

212

213

214

215

216

2122

wall temperature for cylinder head, piston head and cylinder liner were 400K, 442K and

371K respectively. The intake mixture prepared by the port fuel injection was set as a

homogeneous fuel/air mixture. The overall in-cylinder equivalence ratio after direct injection

for all cases was kept stoichiometric. The compression ratio was kept at 10.66 and the total

in-cylinder mass after the direct injection was kept constant for all the cases studied.

Table 3 Computational conditions.

Piston type SOI [ºCA] rDI [%] ST [ºCA]

Baseline case Flat Piston - 0 685

Group 1:

Effect of

Piston

Flat Piston

Piston A

Piston B

Piston C

660 28 670-700

Group 2:

Effect of SOI

Piston B 600

640

660

28 670-700

Group 3:

Effect of rDI

Piston B 660 16

20

28

35

50

670-700

In order to illustrate the applicability of the CFD simulation models, the predicted and

measured in-cylinder pressure and heat release rate profiles of the baseline case are plotted in

Fig. 5. The experimental pressure profile was calculated from the averaged pressure data of

over 200 successive cycles, and the apparent net heat release rate is calculated from the

12

217

218

219

220

221

222

223

224

225

226

227

2324

averaged pressure profile using the first law of thermodynamics. As it can be seen that the

adopted simulation models could reproduce the hybrid combustion process and shows good

agreement with the experimental data.

Fig. 5. The in-cylinder pressure and heat release rate of the baseline case from experiment

and simulation.

3. Results and discussions

3.1 Effect of piston shapes

Fig. 6 shows the effect of spark timings (ST) and piston shapes on the CA50, IMEP and

PRRmax. It should be noted that the IMEP of the simulation is calculated from 660 ºCA to 780

ºCA. As expected, the combustion phasing CA50 is significantly advanced with the earlier

spark timing. Accordingly, the IMEP increases at first and then shows a slight decrease.

Similarly, PRRmax gradually increases with the earlier spark timing.

Overall, the Flat Piston has the most advanced CA50 and highest IMEP. However, the

stratified mixture formed by the Flat Piston cannot effectively control the heat release process

and suppress the pressure rise rate, manifested by the highest PRRmax at each spark timing.

Although delaying the spark timing could achieve lower PRRmax for the Flat Piston, the IMEP

decreased more significantly. In addition, the delaying of the combustion phasing could lead

13

228

229

230

231

232

233

234

235

236

237

238

239

240

241

242

243

244

245

2526

to unstable combustion [17, 24]. Piston C shows the latest CA50 and lowest IMEP among all

pistons though its PRRmax is not the lowest. By contrast, Piston A and Piston B can achieve

relatively earlier CA50, higher IMEP and lower PRRmax at any spark timing. In addition, the

differences in CA50, IMEP and PRRmax between the three pistons gradually increase with the

retarded spark timing.

Fig. 6. Effect of spark timings and piston shapes on CA50, IMEP and PRRmax.

In order to analyse the effect of piston shapes in detail, the combustion processes with the

constant combustion phasing are compared. The spark timings are adjusted to achieve the

constant CA50 around 727.5 ºCA for all cases. Fig. 7 compares the pressure traces of the

combustion process with different piston shapes. The baseline case with the Flat Piston (FP)

and pure PFI is also included in Fig. 7. It can be seen that the early flame propagation

processes before 710 ºCA are similar and they gradually deviate from each other with the

proceeding of the combustion process. Significant differences appear during the later auto-

14

246

247

248

249

250

251

252

253

254

255

256

257

258

259

260

2728

ignition combustion process after 720 ºCA between the pure PFI and the mixed PFI and DI

cases. The introduction of direct fuel injection slows down the auto-ignition combustion

reactions in the second half of the heat release process and dramatically reduces PRRmax with

slightly lower IMEP as shown in Fig. 8. In particular, Piston A produces the lowest PRRmax of

1.79, which is nearly the half of that of the Flat Piston.

Fig. 7. In-cylinder pressure traces for different pistons with the fixed CA50 (727.5 ºCA).

Fig. 8. Comparison of the IMEP and PRRmax for different pistons with the fixed CA50 (727.5

ºCA).

15

261

262

263

264

265

266

267

268

269270

271

2930

In addition, it is noted from Fig. 7 that the spark timings for the non-flat pistons have been

advanced from the origin spark timing (ST) of 685 ºCA to maintain the combustion phasing

around 727.5 ºCA. The main reason could be attributed to the combined effect of fuel

enrichment and charge cooling effect caused by the fuel direct injection.

Fig. 9 compares the section views of equivalence ratio ( ) and temperature (T) distributions

for Flat Piston and Piston B at 680 ºCA just before the spark ignition. Compared to the Flat

Piston, Piston B could effectively direct the fuel to the spark plug and lead to obvious

enrichment of the mixture around spark plug. However, it can be found that most of the fuel

still concentrates near the piston top just before the spark ignition timing, leading to the over-

rich mixture with >1.1 in the central region. This over-rich mixture actually would

deteriorate the flame propagation speed [50]. In addition, the charge cooling effect of direct

injection would also decrease the temperature around the spark plug, as shown in Fig. 9,

which could also slow down the flame propagation speed and delay the combustion phasing.

Fig. 9. Section views of the equivalence ratio and temperature distributions at 680 ºCA for

Flat Piston and Piston B with the fixed CA50 (727.5 ºCA).

In order to reveal the spatially resolved in-cylinder conditions, the cylinder volume is divided

into seven cylindrical zones, as shown in Fig. 10. In addition, a Spark Zone is defined as a

spherical volume of 20 mm in diameter around the spark plug gap. Fig. 11 shows the average

16

272

273

274

275

276

277

278

279

280

281

282

283

284

285

286

287

288

289

290

291

3132

equivalence ratio and temperature of the mixture in the whole combustion chamber, Zone 1

and Spark Zone at 680 ºCA just before the spark discharge. Although the bowl pistons enrich

the mixture in Spark Zone with average equivalence ratio between 1.0 and 1.2, they lead to

the over-rich mixture with the average equivalence ratio over 1.5 in Zone1. Meanwhile, the

mixture temperature in Spark Zone and Zone 1 of bowl pistons drops by 7-10 K compared to

that of Flat Piston. Thus, to maintain the CA50, the advanced spark timing is required to

compensate for the slower flame speed due to the over-rich fuel mixture and charge cooling

effect of direct injection. However, as part of the fuel is consumed by the flame propagation

in the stratified charge region and remaining fuel by the fast auto-ignition combustion, the

maximum heat release rate is tamed in the SFI hybrid combustion process with bowl pistons.

Fig. 10. Schematic of the zones defined to reveal the spatially resolved in-cylinder conditions.

17

292

293

294

295

296

297

298

299

300

301

302

303

304

3334

Fig. 11. The average equivalence ratio and temperature of the mixture in the whole chamber,

Zone 1 and Spark Zone at 680 ºCA for different pistons with the fixed CA50 (727.5 ºCA).

The equivalence ratio and fuel mass fraction consumed by flame propagation (rSI) of a certain

cell just before the auto-ignition of this cell can be extracted to examine its dilution condition

before auto-ignition. Fig. 12 plots the locations of the auto-ignited cells (10% of total cell

numbers) with the lowest and highest equivalence ratio respectively from the simulations

with Flat Piston and Piston B. Compared to the Flat Piston, the auto-ignited cells with highest

equivalence ratio from the simulation with Piston B are distributed in a more compact region

close to the spark plug because of the effective guidance of the fuel by the piston bowl, while

the majority of auto-ignited cells with the lowest equivalence ratio are distributed at the rim

of the combustion chamber. This stratified mixture pattern would enable the fuel rich region

mainly consumed by the slow flame propagation while the lean mixture consumed by auto-

ignition, which facilitate the reduction of the PRRmax.

Fig. 12. The location of the auto-ignited cells with the lowest and highest equivalence ratio

for Flat Piston and Piston B with the fixed CA50 (727.5 ºCA).

18

305

306

307

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

3536

Fig. 13 shows the distribution characteristics of the fuel mass fraction consumed by flame

propagation (rSI) for the auto-ignited cells shown in Fig. 12. The comparison between Flat

Piston and Piston B confirms that more cells with highest equivalence ratio in Fig. 13 (a) are

mainly consumed by the flame propagation (rSI > 0.5) for Piston B, while more cells with

lowest equivalence ratio in Fig. 13 (b) would be mainly consumed by the auto-ignition (rSI <

0.2) for Piston B. The above results demonstrate that the optimal stratified mixture pattern

with rich mixture around spark plug while diluted mixture at the peripheral region formed by

the designed piston bowl enables the slowdown of the heat release rate and reduces the

maximum pressure rise rate as shown in Fig. 8.

Fig. 13. The distribution of rSI of the cells with (a) highest and (b) lowest equivalence ratio

for Flat Piston and Piston B with the fixed CA50 (727.5 ºCA).

3.2 Effect of SOI timings

Fig. 14 shows the effect of spark timing (ST) and injection timings (SOI) on CA50, IMEP

and maximum pressure rise rate (PRRmax). As the SOI timing is delayed, PRRmax is reduced

significantly. In addition, CA50 occurs earlier and IMEP values experience slight decrease.

This is mainly caused by the optimal stratified mixture pattern and increased stratification

19

324

325

326

327

328

329

330

331

332

333

334

335

336

337

338

339

340

341

3738

formed by the delayed SOI timing. Overall, the differences between CA50, IMEP and PRRmax

gradually reduce with more retarded spark timing. This can be attributed to the increased

mixing time of the direct injected fuel and in-cylinder air with the delayed spark timing. This

in turn reduces the difference of the in-cylinder stratified mixture condition and subsequent

combustion process with different injection timings.

Fig. 14. Effect of spark timings and SOI timings on CA50, IMEP and PRRmax.

Fig. 15 compares the pressure traces of the combustion process for different SOI timings with

CA50 at 727.5 ºCA. It can be seen that the late SOI timing of 660 ºCA leads to early

occurrence of the auto-ignition. However, the subsequent auto-ignition combustion process is

moderate compared to that with early SOI timings. This leads to significant reduction of

PPRmax with tiny deterioration of IMEP, as shown in Fig. 16.

In order to explain the underlying process for the reduction of PRRmax with the later injection

timing, the detailed analysis of the in-cylinder conditions and combustion process is

20

342

343

344

345

346

347

348

349

350

351

352

353

354

355

356

3940

performed. Fig. 17 shows the section views of the equivalence ratio at 676 ºCA for different

SOI timings with the fixed combustion phasing. It can be seen that a large amount of the rich

mixture distributes at the exhaust side with the early SOI timing of 600 ºCA. This also

enriches the spark plug region, as shown in the figure. The delayed SOI timing concentrates

the rich mixture at the central region with slight enrichment of the spark plug region while

leave the remaining peripheral region with diluted mixture.

Fig. 15. The in-cylinder pressure traces for different SOI timings with the fixed CA50 (727.5

ºCA).

21

357

358

359

360

361

362

363

364

365

366

367

4142

Fig. 16. Comparison of the IMEP and PRRmax for different SOI timings with the fixed CA50

(727.5 ºCA).

Fig. 17. Section views of the equivalence ratio at 676 ºCA for different SOI timings with the

fixed CA50 (727.5 ºCA).

Fig. 18 compares the average equivalence ratio and temperature of the mixture in the whole

chamber and Spark Zone at 676 ºCA for different SOI timings. With the SOI timing delaying

from 600 to 660 ºCA, the equivalence ratio of the mixture in Spark Zone gradually decreases

from 1.13 to 0.99, while the corresponding temperature increases from 684.2 K to 693.6 K.

The slight decrease of the equivalence ratio of the mixture in the whole chamber indicates

that the evaporation of the direct injected fuel is a little slower for the later SOI timing.

22

368

369

370

371

372

373

374

375

376

377

378

379

380

381

4344

Fig. 18. The average equivalence ratio and temperature of the mixture in the whole chamber

and Spark Zone at 676 ºCA for different SOI timings with the fixed CA50 (727.5 ºCA).

Fig. 19 shows the average turbulence kinetic energy (TKE) and the flow velocity magnitudes

(Vm) in the whole chamber and Spark Zone at 676 ºCA for different SOI timings. It can be

seen that the delayed SOI timing enhances the in-cylinder flow and elevates the in-cylinder

turbulence level. This benefits the mixing between the injected fuel and in-cylinder air, which

enhances the early flame propagation process. Therefore, the combustion phasing of the later

SOI timing would be advanced with the fixed spark timing, as shown in Fig. 14. This

explains why the spark timing should be delayed to achieve the same CA50 with the later SOI

timing in Fig. 15.

23

382

383

384

385

386

387

388

389

390

391

392

4546

Fig. 19. The average TKE and Vm in the whole chamber and Spark Zone at 676 ºCA for

different SOI timings with the fixed CA50 (727.5 ºCA).

In this study, the crank angle corresponding to the mode transition (CAT) from flame

propagation to auto-ignition combustion is defined as the crank angle when the auto-ignited

cells exceed 2% of the total cells in the cylinder. Fig. 20 shows CAT and the ratio of the

accumulated heat released (RCAT) at CAT for different SOI timings. It can be seen that as the

SOI timing is retarded, CAT is gradually advanced and the corresponding RCAT is gradually

reduced from 26.4 % to 23.3%. This indicates that the auto-ignition of the end-gas is

enhanced by the later SOI timing. The reasons can be attributed to the distribution of the

optimal stratified mixture.

24

393

394

395

396

397

398

399

400

401

402

403

404

4748

Fig. 20. The mode transition crank angle (CAT) and the ratio of the accumulated heat

released (RCAT) at CAT for different SOI timings.

As shown in Fig. 21, the later SOI timing of 660 ºCA enables the early auto-ignition

occurring at two small regions instead of a larger region with SOI timing of 600 ºCA,

indicating the late SOI timing render more regions suitable to achieve auto-ignition at the

early stage of auto-ignition combustion.

Fig. 21. The location of 2% of the firstly auto-ignited cells with SOI timing of 600 ºCA and

660 ºCA.

Fig. 22 plots the distribution of the equivalence ratio and temperature of 2% of the firstly

auto-ignited cells at 676 ºCA and the crank angle just before CAT with SOI timing of 600

25

405

406

407

408

409

410

411

412

413

414

415

416

417

418

4950

ºCA and 660 ºCA. For the early SOI timing of 600 ºCA, the first auto-ignited cells have a

wider range of equivalence ratios at 676 ºCA just before the spark discharge. The higher

equivalence ratio leads to relatively lower temperature because of the stronger latent heat of

vaporization. The higher equivalence ratio and lower temperature finally lead to the longer

auto-ignition delay time of these cells. With the proceeding of the flame propagation and

mixing process, the equivalence ratio of these first auto-ignited cells gradually decreases,

while the corresponding temperature significantly increases because of the heat from the

flame propagation. For the SOI timing of 660 ºCA, the initial equivalence ratio of the first

auto-ignited cells is around 0.72, indicating that the direct injected fuel has not reached these

regions. Thus the corresponding temperature is a little higher than that from the SOI timing of

600 ºCA. The subsequent mixing process with the central injected fuel gradually increases

the equivalence ratio, which further decreases the auto-ignition delay time. Finally, it can be

found that the temperature before the CAT for SOI timing of 660 ºCA is significantly lower

than that from SOI timing of 600 ºCA, indicating the less requirement of the heat from flame

propagation. Essentially, the fuel stratification pattern formed by the later SOI timing of 660

ºCA enables earlier auto-ignition of the peripheral lean mixture with relatively higher

temperature compared to that from the SOI timing of 600 ºCA. The above detailed analyses

of the first auto-ignited cells have explained the reason for the early mode transition crank

angle CAT and the reduced RCAT for the SOI timing of 660 ºCA.

26

419

420

421

422

423

424

425

426

427

428

429

430

431

432

433

434

435

436

437

5152

Fig. 22. The distribution of the equivalence ratio and temperature of 2% of the firstly auto-

ignited cells at 676 ºCA and the crank angle just before CAT with SOI timing of 600 ºCA and

660 ºCA.

Fig. 23 shows the flame propagation dominated combustion duration D1 (CA10-CAT), the

auto-ignition dominated combustion duration D2 (CAT-CA90) and the overall combustion

duration D0 (CA10-CA90) for different SOI timings. As speculated from the previous

analysis, the early flame propagation dominated combustion duration (D1) significantly

reduces with the delayed SOI timing. However, it is found that the auto-ignition dominated

combustion duration (D2) increases more dramatically, leading to a moderate increase of the

overall combustion duration (D0). This can be attributed to the increased inhomogeneity in

the equivalence ratio and temperature distributions with the later SOI timings, as shown in

Fig. 24.

27

438

439

440

441

442

443

444

445

446

447

448

449

450

451

5354

Fig. 23. The flame propagation dominated combustion duration D1, the auto-ignition

dominated combustion duration D2 and the overall combustion duration D0 for different SOI

timings.

Fig. 24. The inhomogeneity of equivalence ratio and temperature at 676 ºCA for different

SOI timings.

3.3 Effect of direct injection ratio (rDI)

Fig. 25 shows the effect of spark timings and direct injection ratios on CA50, IMEP and

PRRmax. As the direct injection ratio is increased from 16% to 35%, CA50 advances when the

28

452

453

454

455

456

457

458

459

460

461

462

5556

earlier spark timings are before 680 ºCA. But when the spark timing is retarded after 680

ºCA, the combustion phasing is delayed with the increased direct injection ratio. In addition,

for the largest direct injection ratio, i.e. rDI=50%, CA50 is significantly delayed with both

early and late spark timing. With the early spark timings, the enrichment of the central

mixture and the enhancement of the mixing effect from the increased direct injection ratios

would benefit the early flame propagation and in turn advance the combustion phasing.

However, the over-rich mixture in the central region and over-lean mixture in the peripheral

region would inhibit the whole combustion with the largest direct injection ratio of 50% and

in turn delay the combustion phasing. With the late spark timing, there is more time for the

sufficient mixing between the injected fuel and in-cylinder air. The degree of fuel

stratification gradually increases with the direct injection ratio increasing, leading to the

slower combustion process and later combustion phasing. The IMEP gradually decreases

with the increasing direct injection ratio, which is caused by the slower combustion process

and increased incomplete combustion of the over-lean peripheral mixture with the direct

injection ratio increasing. Meanwhile, PRRmax gradually decreases as more fuel is injected

directly into the cylinder and the presence of mixture stratification.

Fig. 26 compares the pressure traces of the combustion process for different direct injection

ratios with CA50 at 727.5ºCA. In order to keep the combustion phasing constant, more

advanced spark ignition is needed when the direct injection ratio is increased. As a result, the

flame propagation in the early part of combustion process is enhanced whilst auto-ignition

combustion contribution is reduced, although the combustion mode transition is advanced.

As the direct injection ratio is increased from 16% to 50%, the IMEP is reduced by 15% from

2.9 to 2.45 bar, whilst the PRRmax is significantly reduced by 52.7% from 2.62 to 1.24 bar/

ºCA, as shown in Fig. 27.

29

463

464

465

466

467

468

469

470

471

472

473

474

475

476

477

478

479

480

481

482

483

484

485

486

5758

Fig. 25. Effect of spark timings and direct injection ratios on CA50, IMEP and PRRmax.

Fig. 26. The in-cylinder pressure traces for different direct injection ratios with the fixed

CA50 (727.5 ºCA).

30

487

488

489

490

491

492

5960

Fig. 27. Comparison of the IMEP and PRRmax for different direct injection ratios with the

fixed CA50 (727.5 ºCA).

Fig. 28 shows the section views of the equivalence ratio distribution at 680 ºCA for different

direct injection ratios. The results show that the equivalence ratio of the outer region is

gradually reduced with the increase of the direct injection ratio. Meanwhile, the central rich

mixture becomes more concentrated around the spark plug with increasing direct injection

ratio. Fig. 29 provides the quantitative data of the in-cylinder equivalence ratio distribution

patterns with different direct injection ratios. Overall, the equivalence ratio of the mixture in

the whole chamber shows a slightly decreasing trend with the increasing of direct injection

ratio because of the weaker evaporation of the over-rich region caused by the larger direct

injection ratio. In the Spark Zone, as expected, the equivalence ratio shows a clear increasing

trend with the increase in the direct injection ratio. The interaction between the direct

injection at 660 ºCA and the piston bowl of Piston B produces a significant stratification from

the richest mixture in Zone 1 to the leanest mixture in Zone 7, as indicated in both Figs. 28

and 29. Although the stratification pattern is similar for all direct injection ratios, the degree

of the stratification is significantly increased with the direct injection ratio. Specifically, the

31

493

494

495

496

497

498

499

500

501

502

503

504

505

506

507

508

509

510

6162

larger direction injection ratio leads to richer central mixture while leaner mixture at outer

region, which is responsible for the decreased PRRmax.

Fig. 28. The section views of the equivalence ratio distribution at 680 ºCA for different direct

injection ratios.

Fig. 29. The zone-to-zone equivalence ratio distribution at 680 ºCA for different direct

injection ratios.

Fig. 30 shows the flame propagation dominated combustion duration D1, the auto-ignition

dominated combustion duration D2 and the ratio of the accumulated heat released RCAT at

CAT for different direct injection ratios. The early flame propagation dominated combustion

32

511

512

513

514

515

516

517

518

519

520

521

522

523

6364

duration D1 is gradually prolonged with the direct injection ratio, which is caused by the

over-rich mixture in the central region formed by the larger direct injection ratio. However,

the auto-ignition dominated combustion duration (D2) is nearly constant before the

substantial increase with rDI =50%. When the direct injection ratio is less than 35%, the

increased RCAT with the larger direct injection ratio benefits the maintenance of auto-ignition

dominated combustion duration because of the increased heat released from early flame

propagation. However, it is also observed that RCAT remains nearly constant after rDI=35%,

due to the over-rich mixture. This in turn triggers the earlier auto-ignition of the mixture with

appropriate equivalence ratio close to the central region instead of the lean mixture at the

peripheral region. In addition, the auto-ignition of inhomogeneous mixture leads to longer

auto-ignition dominated combustion duration D2. The above analysis explains the early

combustion mode transition and longer auto-ignition dominated combustion duration D2 with

larger direct injection ratio observed in Fig. 26.

Fig. 30. The flame propagation dominated combustion duration (D1), the auto-ignition

dominated combustion duration (D2) and the ratio of the accumulated heat released (RCAT) at

CAT for different direct injection ratios.

33

524

525

526

527

528

529

530

531

532

533

534

535

536

537

538

539

540

541

6566

Fig. 31 shows the average temperature profiles of the mixture in the whole chamber, Spark

Zone and Zone 7 for different direct injection ratios. The average temperature of the mixture

in the whole chamber is a little higher at the early combustion stage for the larger direct

injection ratio because of the advanced spark ignition timing, while it increases slowly at the

later stage. Specially, the temperature of the mixture in Spark Zone is significantly higher

with the larger direct injection ratios. The average temperature of the mixture in Zone 7 for

different direct injection ratios is similar at the early combustion stage because of the absence

of both flame propagation and the auto-ignition. At the later stage, the temperature rise in

Zone 7 is significantly slowed down with the larger direct injection ratio.

Fig. 31. The average temperature profiles of the mixture in the whole chamber, Spark Zone

and Zone 7 for different direct injection ratios.

As indicated by the temperature profile of the mixture in the whole chamber, the larger direct

injection ratio leads to lower temperature at the end of combustion, resulting from the larger

charge cooling effect and the incomplete combustion of the over-lean mixture at the outer

region. This deteriorates the IMEP, as shown in Fig. 27. With the direct injection ratio

increasing from 16% to 50%, the unburned fuel fraction at the end of simulation increases

34

542

543

544

545

546

547

548

549

550

551

552

553

554

555

556

557

558

559

6768

from 8.4 % to 11.8 %. Fig. 32 shows the distribution of the temperature and equivalence ratio

of all cells and the incompletely burned cells with xb<0.8 at 680 ºCA for rDI=16% and 50%.

The parameter xb indicates the burned fuel mass fraction at 740 ºCA in a certain cell. It can be

seen that the distribution of the equivalence ratio and temperature of all cells at 680 ºCA is

mainly divided into two groups: the central stratified mixture along the diagonal and the

peripheral lean mixture along the bottom. The peripheral lean mixture with rDI=50% become

the main part of the incompletely burned region because of the leaner mixture and the less

heat from the central flame propagation. Fig. 33 shows the locations of the incompletely

burned cells with xb<0.8 at 740 ºCA with rDI=16% and 50%. As expected, the area of the

incompletely burned region is larger and extends into the central region for rDI=50%. The

above analysis indicates that the very high direct injection ratio actually deteriorates the

performance of the stoichiometric SFI hybrid combustion because of incomplete combustion

of the over-lean peripheral mixture, although it can be used to suppress the PPRmax.

Fig. 32. The distribution of the equivalence ratio and temperature of all cells and

incompletely burned cells with xb<0.8 at 680 ºCA with rDI=16% and 50%.

35

560

561

562

563

564

565

566

567

568

569

570

571

572

573

574

575

576

6970

Fig. 33. The locations of the incompletely burned cells with xb<0.8 at 740 ºCA with rDI=16%

and 50%.

4. Summary and Conclusions

The stratified flame ignition (SFI) hybrid combustion process achieved by the combination of

port fuel injection (PFI) and direct injection (DI) was proposed in this study to enhance the

control of SI-CAI hybrid combustion and moderate the high maximum pressure rise rate

(PRRmax) at the high load operation. The newly developed 3-D CFD hybrid combustion

models were applied to the in-cylinder flow, mixture formation and combustion analysis. The

effect of the stratified mixture formed by different piston shapes, start of direct injection

(SOI) timings and direct injection ratios (rDI) on the stoichiometric SFI hybrid combustion

operation was studied using the 3-D CFD simulations. A sweep of the spark timing (ST) was

performed for all pistons, SOI timings and direct injection ratios. The SFI hybrid combustion

process with the same combustion phasing was investigated in details. The main findings can

be summarized as follows:

1. The shallow pistons (Piston A and Piston B) can significantly enhance the formation

of a stable stratified flame in the stratified mixture around the spark plug. The

stratified flame formed by the direct injection and piston design allows the start of

combustion and combustion phasing to be controlled. In addition, by altering the

36

577

578

579

580

581

582

583

584

585

586

587

588

589

590

591

592

593

594

595

596

597

7172

relatively amount of heat released by relatively slower flame propagation in the

stratified charge mixture and faster auto-ignition of diluted mixtures, the peak

pressure rise rate (PRRmax) can be significantly reduced.

2. When the SOI timing of direct fuel injection takes place later in the compression

stroke, the combustion phasing (CA50) is brought forward and PRRmax can be reduced

significantly with a little deterioration of IMEP. With the later SOI timings, the

formation and propagation of a stable flame in the stratified mixture accelerated the

occurrence of auto-ignition reactions of the peripheral lean mixture, leading to the

early mode transition. Although the flame propagation dominated combustion

duration D1 is reduced, the auto-ignition dominated combustion duration D2

significant is increased because of the increased inhomogeneity of equivalence ratio

and temperature, leading to the reduction of PRRmax. Overall, the later SOI timing

increases the whole combustion duration slightly.

3. When the amount of the direct injection fuel is increased, PRRmax is reduced

noticeably while the IMEP is also deteriorated because of the slower combustion

process and incomplete combustion of the over-lean peripheral mixture. The early

flame propagation dominated combustion duration D1 is gradually prolonged with the

increasing of direct injection ratio, while the auto-ignition dominated combustion

duration D2 is nearly constant when less than half of the fuel is injected directly into

the cylinder.

4. Spark timing is an effective tool to control the combustion phasing, IMEP and PRRmax

of the stoichiometric SFI hybrid combustion regardless of SOI timings and direct

injection ratios. However, the sensitivity of SFI hybrid combustion to the stratified

mixture is different for different spark timings. The differences in CA50, IMEP and

PRRmax between different pistons become more noticeable at the retarded spark

37

598

599

600

601

602

603

604

605

606

607

608

609

610

611

612

613

614

615

616

617

618

619

620

621

622

7374

timings. With the same stratified mixture pattern formed by Piston B, the differences

in CA50, IMEP and PRRmax between different SOI timings become less with the

retarded spark timings. With the direct injection ratio increasing from 16% to 35%,

the CA50 gradually advances with the earlier spark timings before 680 ºCA but delays

with the later spark timings after 680 ºCA. For the largest direct injection ratio of

50%, the CA50 is significantly delayed with both early and late spark timings.

5. It can be inferred from this study that both the early flame propagation and auto-

ignition process in SFI hybrid combustion can be controlled by the fuel stratification.

The reduction of the PRRmax caused by the stratified mixture enables the advance of

spark timing to achieve maximum IMEP. However, the performance of the

stoichiometric SFI hybrid combustion is critical to the degree of the fuel stratification

because the over-rich mixture in the central region and over-lean mixture at the

peripheral region would deteriorate the flame propagation and auto-ignition

respectively.

Funding

The study is a part of the State Key Project of Fundamental Research Plan (Grant

2013CB228403) supported by the Ministry of Science and Technology of China.

Nomenclature

SFI stratified flame ignition

PRRmax maximum pressure rise rate

PFI port fuel injection

DI direct injection

SOI start of direct injection

rDI direct injection ratios

CFD computational fluid dynamics

38

623

624

625

626

627

628

629

630

631

632

633

634

635

636

637

638

639

640

641

642

643

644

645

646

647

7576

IMEP indicated mean effective pressure

ST spark timing

CA50 crank angle of 50% total heat release

CAI controlled auto-ignition

SI spark ignition

SACI spark assisted compression ignition

CCV cycle-to-cycle variation

PVO positive valve overlapping

HCCI homogenous charge compression ignition

ASSCI assisted spark stratified compression ignition

PLIF planar laser-induced fluorescence

RANS reynolds-averaged Navier Stokes

PISO pressure-implicit with splitting of operators

4VVAS 4-variable valve actuation system

eEGR external exhaust gas recirculation

iEGR internal exhaust gas recirculation

NVO negative valve overlapping

fuel/air equivalence ratio

rSI fuel mass fraction consumed by flame propagation in a certain cell

TKE turbulence kinetic energy

Vm flow velocity magnitudes

CAT crank angle corresponding to the mode transition

RCAT ratio of the accumulated heat released at CAT

D1 flame propagation dominated combustion duration (CA10-CAT)

D2 auto-ignition dominated combustion duration (CAT-CA90)

39

648

649

650

651

652

653

654

655

656

657

658

659

660

661

662

663

664

665

666

667

668

669

670

671

672

7778

D0 overall combustion duration (CA10-CA90)

xb burned fuel mass fraction at 740 ºCA in a certain cell

Reference

[1] Zhao H, HCCI and CAI engines for the automotive industry. Woodhead Publishing: 2007.

[2] Alkidas A C, Combustion advancements in gasoline engines. Energy Conversion and

Management 2007; 48: 2751–2761.

[3] Zhang Y, Xie H, Zhao H. Investigation of SI-HCCI Hybrid Combustion and Control

Strategies for Combustion Mode Switching in a Four-Stroke Gasoline Engine. Combust. Sci.

and Tech 2009; 181: 782–799.

[4] Xie H, Yang L, Qin J, Gao R, Zhu H, He B, Zhao H. The Effect of Spark Ignition on the

CAI Combustion Operation. SAE Technical Paper 2005-01-3738.

[5] Wang Z, Wang J, Shuai S, Ma Q. Effects of Spark Ignition and Stratified Charge on

Gasoline HCCI Combustion With Direct Injection. SAE Technical Paper 2005-01-0137.

[6] Wang Z, He X, Wang J, Shuai S, Xu F, Yang D. Combustion visualization and

experimental study on spark induced compression ignition (SICI) in gasoline HCCI engines.

Energy Conversion and Management 2010; 51: 908-917.

[7] Milovanovic N, Blundell D, Gedge S, Turner J. SI-HCCI-SI Mode Transition at Different

Engine Operating Conditions. SAE Technical Paper 2005-01-0156.

[8] Santoso H, Matthews J, Cheng W K. Managing SI/HCCI Dual-Mode Engine Operation.

SAE Technical Paper 2005-01-0162.

[9] Chen T, Xie H, Le Li, Yu W, Li Z, Zhao H. Continuous Load Adjustment Strategy of a

Gasoline HCCI-SI Engine Fully Controlled by Exhaust Gas. SAE Technical Paper 2011-01-

1408.

[10] Manofsky L, Vavra J, Assanis D, Babajimopoulos A. Bridging the Gap between HCCI

and SI: Spark- Assisted Compression Ignition. SAE Technical Paper 2011-01-1179.

40

673

674

675

676

677

678

679

680

681

682

683

684

685

686

687

688

689

690

691

692

693

694

695

696

697

7980

[11] Olesky L M, Martz J B, Lavoie G A, Vavra J, Assanis D N, Babajimopoulos A. The

effects of spark timing, unburned gas temperature, and negative valve overlap on the rates of

stoichiometric spark assisted compression ignition combustion. Applied Energy 2013;

105(0): 407-417.

[12] Urushihara T, Yamaguchi K, Yoshizawa K, Itoh T. A Study of a Gasoline-fueled

Compression Ignition Engine ~ Expansion of HCCI Operation Range Using SI Combustion

as a Trigger of Compression Ignition. SAE Technical Paper 2005-01-0180.

[13] Wang X, Xie H, Li L, Xie L, Chen T, Zhao H. Effect of the Thermal Stratification on SI-

CAI Hybrid Combustion in a Gasoline Engine. Applied Thermal Engineering 2013; 61(2):

451–460.

[14] Wang X, Xie H, Zhao H. Computational Study of the Influence of In-cylinder Flow on

SI-CAI Hybrid Combustion in a Gasoline Engine. International Journal of Engine Research

2014; Accepted.

[15] Xie H, Le Li, Chen T, Yu W, Wang X, Zhao H. Study on spark assisted compression

ignition (SACI) combustion with positive valve overlap at medium–high load. Applied

Energy 2013; 101: 622–633.

[16] Szybist J P, Nafziger E, Weall A. Load Expansion of Stoichiometric HCCI Using Spark

Assist and Hydraulic Valve Actuation. SAE Technical Paper 2010-01-2172.

[17] Yun H, Wermuth N, Najt P. Extending the High Load Operating Limit of a Naturally-

Aspirated Gasoline HCCI Combustion Engine. SAE Technical Paper 2010-01-0847.

[18] Aroonsrisopon T, Werner P, Waldman J O, Sohm V, Foster D E, Morikawa T, Iida M.

Expanding the HCCI Operation With the Charge Stratification. SAE Technical Paper 2004-

01-1756.

41

698

699

700

701

702

703

704

705

706

707

708

709

710

711

712

713

714

715

716

717

718

719

720

8182

[19] Choi S, Lim J, Ki M, Min K, Choi H. Analysis of Cyclic Variation and the Effect of Fuel

Stratification on Combustion Stabilityin a Port Fuel Injection (PFI) CAI Engine. SAE

Technical Paper 2009-01-0670.

[20] Dahl D, Andersson M, Berntsson A, Denbratt I, Koopmans L. Reducing Pressure

Fluctuations at High Loads by Means of Charge Stratification in HCCI Combustion with

Negative Valve Overlap. SAE Technical Paper 2009-01-1785.

[21] Dec J E, Yang Y, Dronniou N. Boosted HCCI - Controlling Pressure-Rise Rates for

Performance Improvements using Partial Fuel Stratification with Conventional Gasoline.

SAE Technical Paper 2011-01-0897.

[22] Yang D, Wang Z, Wang J, Shuai S. Experimental study of fuel stratification for HCCI

high load extension. Applied Energy 2011; 88(9): 2949 - 2954.

[23] Yang Y, Dec J E, Dronniou N, Berg M S. Tailoring HCCI heat-release rates with partial

fuel stratification: Comparison of two-stage and single-stage-ignition fuels. Proceedings of

the Combustion Institute 2011; 33(2): 3047–3055.

[24] Yoshizawa K, Teraji A, Miyakubo H, Yamaguchi K, Urushihara T. Study of High Load

Operation Limit Expansion for Gasoline Compression Ignition Engines. Journal of

Engineering for Gas Turbines and Power 2006; 128(2): 377-387.

[25] Li D, Wang Z, Liu H, Wang J. Combustion Mode Switch by Integrating Stoichiometric

ASSCI Mode in a Four-cylinder Gasoline SI/HCCI Engine. SAE Technical Paper 2014-01-

1288.

[26] Persson H, Sjöholm J, Kristensson E, Johansson B, Richter M, Aldén M. Study of Fuel

Stratification on Spark Assisted Compression Ignition (SACI) Combustion with Ethanol

Using High Speed Fuel PLIF. SAE Technical Paper 2008-01-2401.

[27] Berntsson A W, Denbratt I. Optical study of HCCI Combustion using NVO and an SI

Stratified Charge. SAE Technical Paper 2007-24-0012.

42

721

722

723

724

725

726

727

728

729

730

731

732

733

734

735

736

737

738

739

740

741

742

743

744

745

8384

[28] Wu X, Daniel R, Tian G, Xu H, Huang Z, Richardson D. Dual-injection: The flexible,

bi-fuel concept for spark-ignition engines fuelled with various gasoline and biofuel blends.

Applied Energy 2011; 88(7): 2305 - 2314.

[29] Joh M, Huh K Y, Noh S H, Choi K H. Numerical Prediction of Stratified Charge

Distribution in a Gasoline Direct-Injection Engine - Parametric Studies. SAE Technical Paper

1999-01-0178.

[30] Lippert A M, Tahry S H E, Huebler M S, Parrish S E, Inoue H, Noyori T, Nakama K,

Abe T. Development and Optimization of a Small-Displacement Spark-Ignition Direct-

Injection Engine – Stratified Operation. SAE Technical Paper 2004-01-0033.

[31] Ikoma T, Abe S, Sonoda Y, Suzuki H, Suzuki Y, Basaki M. Development of V-6 3.5-

liter Engine Adopting New Direct Injection System. SAE Technical Paper 2006-01-1259.

[32] Xu Z, Yi J, Curtis E, Wooldridge S. Applications of CFD Modeling in GDI Engine

Piston Optimization. SAE Technical Paper 2009-01-1936.

[33] Yi J, Wooldridge S, Coulson G, Hilditch J, Iyer C O, Moilanen P, Papaioannou G,

Reiche D, Shelby M, VanDerWege B, Weaver C, Xu Z, Davis G, Hinds B, Schamel A.

Development and Optimization of the Ford 3.5L V6 EcoBoost Combustion System. SAE

Technical Paper 2009-01-1494.

[34] Iyer C O, Han Z, Yi J. CFD Modeling of a Vortex Induced Stratification Combustion

(VISC) System. SAE Technical Paper 2004-01-0550.

[35] Oh H, Bae C. Effects of the injection timing on spray and combustion characteristics in a

spray-guided DISI engine under lean-stratified operation. Fuel 2013; 107(0): 225 - 235.

[36] Costa M, Sorge U, Allocca L. Increasing energy efficiency of a gasoline direct injection

engine through optimal synchronization of single or double injection strategies. Energy

Conversion and Management 2012; 60(0): 77 - 86.

43

746

747

748

749

750

751

752

753

754

755

756

757

758

759

760

761

762

763

764

765

766

767

768

769

8586

[37] Wang X, Zhao H, Xie H, He B. Numerical Study of the Effect of Piston Shapes and Fuel

Injection Strategies on In-cylinder Conditions in a PFI/GDI Gasoline Engine. SAE Int. J.

Engines 2014; 7(4).

[38] Jones W P, Prediction methods for turbulent flames. In Prediction Methods for

Turbulent Flow, Kollmann W, Ed Hemisphere: Washington, D.C., 1980; 1-45.

[39] Angelberger C, Poinsot T, Delhay B. Improving Near-Wall Combustion and Wall Heat

Transfer Modeling in SI Engine Computations. SAE Technical Paper 972881.

[40] Stach T, Schlerfer J, Vorbach M. New Generation Multi-hole Fuel Injector for Direct-

Injection SI Engines - Optimization of Spray Characteristics by Means of Adapted Injector

Layout and Multiple Injection. SAE Technical Paper 2007-01-1404.

[41] Gosman A D, Marooney C J. Development and validation of a computer code for Diesel

combustion. IDEA Project, Subprogram E1, October, 1991.

[42] Reitz R D, Diwakar R. Effect of Drop Breakup on Fuel Sprays. SAE Technical Paper

860469.

[43] Cd-adapco. Methodology, STAR-CD VERSION 4.14.

[44] Bai C, Gosman A D. Development of Methodology for Spray Impingement Simulation.

SAE Technical Paper 950283.

[45] Colin O, Benkenida A. The 3-Zones Extended Coherent Flame Model (ECFM3Z) for

Computing Premixed/Diffusion Combustion. Oil & Gas Science and Technology 2004;

59(6): 593-609.

[46] CRUZ A P D. Three-dimensional modeling of self-ignition in HCCI and conventional

diesel engines. Combustion Science and Technology 2004; 176(5-6): 867-887.

[47] Mehl M, Chen J Y, Pitz W J, Sarathy S M, Westbrook C K. An approach for formulating

surrogates for gasoline with application towards a reduced surrogate mechanism for CFD

engine modeling. Energy & Fuels 2011; 25(11): 5215-5223.

44

770

771

772

773

774

775

776

777

778

779

780

781

782

783

784

785

786

787

788

789

790

791

792

793

794

8788

[48] Wang X, Xie H, Xie L, Zhang L, Le Li, Chen T, Zhao H. Numerical simulation and

validation of SI-CAI hybrid combustion in a CAI/HCCI gasoline engine. Combustion Theory

and Modelling 2013; 17(1): 142-166.

[49] Xie H, Hou S, Qin J, Zhang Y, Li N, Zhao H. Control Strategies for Steady and

Transient Operation of a 4-Stroke Gasoline Engine with CAI Combustion Using a 4-Variable

Valve Actuating System (4VVAS). SAE Technical Paper 2006-01-1083.

[50] Bradley D, Hicks R A, Lawes M, Sheppard C G W, Woolley R. The Measurement of

Laminar Burning Velocities and Markstein Numbers for Iso-octane–Air and Iso-octane–n-

Heptane–Air Mixtures at Elevated Temperatures and Pressures in an Explosion Bomb.

Combustion and Flame 1998; 115(1–2): 126 - 144.

45

795

796

797

798

799

800

801

802

803

804

8990

Table captions:

Table 1 Engine specifications.

Table 2 Operation conditions.

Table 3 Computational conditions.

Figure captions:

Fig. 1. Schematic diagram of (a) CAI combustion and (b) SI-CAI hybrid combustion.

Fig. 2. Schematic diagram of spray modelling.

Fig. 3. Newly designed bowl pistons.

Fig. 4. Engine mesh with Piston A.

Fig. 5. The in-cylinder pressure and heat release rate of the baseline case from experiment

and simulation.

Fig. 6. Effect of spark timings and piston shapes on CA50, IMEP and PRRmax.

Fig. 7. In-cylinder pressure traces for different pistons with the fixed CA50 (727.5 ºCA).

Fig. 8. Comparison of the IMEP and PRRmax for different pistons with the fixed CA50 (727.5

ºCA).

Fig. 9. Section views of the equivalence ratio and temperature distributions at 680 ºCA for

Flat Piston and Piston B with the fixed CA50 (727.5 ºCA).

Fig. 10. Schematic of the zones defined to reveal the spatially resolved in-cylinder conditions.

Fig. 11. The average equivalence ratio and temperature of the mixture in the whole chamber,

Zone 1 and Spark Zone at 680 ºCA for different pistons with the fixed CA50 (727.5 ºCA).

Fig. 12. The location of the auto-ignited cells with the lowest and highest equivalence ratio

for Flat Piston and Piston B with the fixed CA50 (727.5 ºCA).

Fig. 13. The distribution of rSI of the cells with (a) highest and (b) lowest equivalence ratio

for Flat Piston and Piston B with the fixed CA50 (727.5 ºCA).

46

805

806

807

808

809

810

811

812

813

814

815

816

817

818

819

820

821

822

823

824

825

826

827

828

829

9192

Fig. 14. Effect of spark timings and SOI timings on CA50, IMEP and PRRmax.

Fig. 15. The in-cylinder pressure traces for different SOI timings with the fixed CA50 (727.5

ºCA).

Fig. 16. Comparison of the IMEP and PRRmax for different SOI timings with the fixed CA50

(727.5 ºCA).

Fig. 17. Section views of the equivalence ratio at 676 ºCA for different SOI timings with the

fixed CA50 (727.5 ºCA).

Fig. 18. The average equivalence ratio and temperature of the mixture in the whole chamber

and Spark Zone at 676 ºCA for different SOI timings with the fixed CA50 (727.5 ºCA).

Fig. 19. The average TKE and Vm in the whole chamber and Spark Zone at 676 ºCA for

different SOI timings with the fixed CA50 (727.5 ºCA).

Fig. 20. The mode transition crank angle (CAT) and the ratio of the accumulated heat

released (RCAT) at CAT for different SOI timings.

Fig. 21. The location of 2% of the firstly auto-ignited cells with SOI timing of 600 ºCA and

660 ºCA.

Fig. 22. The distribution of the equivalence ratio and temperature of 2% of the firstly auto-

ignited cells at 676 ºCA and the crank angle just before CAT with SOI timing of 600 ºCA and

660 ºCA.

Fig. 23. The flame propagation dominated combustion duration D1, the auto-ignition

dominated combustion duration D2 and the overall combustion duration D0 for different SOI

timings.

Fig. 24. The inhomogeneity of equivalence ratio and temperature at 676 ºCA for different

SOI timings.

Fig. 25. Effect of spark timings and direct injection ratios on CA50, IMEP and PRRmax.

47

830

831

832

833

834

835

836

837

838

839

840

841

842

843

844

845

846

847

848

849

850

851

852

853

9394

Fig. 26. The in-cylinder pressure traces for different direct injection ratios with the fixed

CA50 (727.5 ºCA).

Fig. 27. Comparison of the IMEP and PRRmax for different direct injection ratios with the

fixed CA50 (727.5 ºCA).

Fig. 28. The section views of the equivalence ratio distribution at 680 ºCA for different direct

injection ratios.

Fig. 29. The zone-to-zone equivalence ratio distribution at 680 ºCA for different direct

injection ratios.

Fig. 30. The flame propagation dominated combustion duration (D1), the auto-ignition

dominated combustion duration (D2) and the ratio of the accumulated heat released (RCAT) at

CAT for different direct injection ratios.

Fig. 31. The average temperature profiles of the mixture in the whole chamber, Spark Zone

and Zone 7 for different direct injection ratios.

Fig. 32. The distribution of the equivalence ratio and temperature of all cells and

incompletely burned cells with xb<0.8 at 680 ºCA with rDI=16% and 50%.

Fig. 33. The locations of the incompletely burned cells with xb<0.8 at 740 ºCA with rDI=16%

and 50%.

48

854

855

856

857

858

859

860

861

862

863

864

865

866

867

868

869

870

9596