Combustion and direct energy conversion inside a micro-combustor

Combustion and direct energy conversion inside a micro-combustor

Accepted Manuscript Title: Combustion and direct energy conversion inside a micro-combustor Author: Yafeng Lei, Wei Chen, Jiang Lei PII: DOI: Referenc...

876KB Sizes 1 Downloads 99 Views

Accepted Manuscript Title: Combustion and direct energy conversion inside a micro-combustor Author: Yafeng Lei, Wei Chen, Jiang Lei PII: DOI: Reference:

S1359-4311(16)30116-8 http://dx.doi.org/doi: 10.1016/j.applthermaleng.2016.01.162 ATE 7725

To appear in:

Applied Thermal Engineering

Received date: Accepted date:

18-9-2015 31-1-2016

Please cite this article as: Yafeng Lei, Wei Chen, Jiang Lei, Combustion and direct energy conversion inside a micro-combustor, Applied Thermal Engineering (2016), http://dx.doi.org/doi: 10.1016/j.applthermaleng.2016.01.162. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

1 2 3 4 5

Combustion and Direct Energy Conversion inside a Micro-Combustor Yafeng Lei1, Wei Chen2, Jiang Lei3 1

General Electrical Company, Houston 77041, Texas, USA

6 7 8 9 10

2

3

School of Aerospace, Xi’an Jiaotong University, Xi’an, Shanxi, 710049, China

11

2

Author Correspondence:

12

Professor: Wei Chen

13

China University of Mining and Technology

14

Xuzhou, Jiangsu 221008, China

15 16 17 18 19 20 21 22 23 24 25

[email protected]

26

Abstract

27

Electrical energy can be generated by employing a micro-thermophotovoltaic (TPV) cell

28

which absorbs thermal radiation from combustion taking place in a micro-combustor. The

29

stability of combustion in a micro-combustor is essential for operating a micro-power system

30

using hydrogen and hydrocarbon fuels as energy source. To understand the mechanism of

School of Electric power Engineering, China University of Mining and Technology, Xuzhou 221008, China

Highlights    

The flammability range of micro-combustor was broadened with heat recirculation The quenching diameter decreased with heat recirculation compared to without recirculation The surface areas to volume ratio was the most important parameter affecting the energy conversion efficiency The maximum conversion efficiency (3.15%) was achieved with 1 mm inner diameter

1

Page 1 of 28

31

sustaining combustion within the quenching distance of fuel, this study proposed an annular

32

micro combustion tube with recirculation of exhaust heat. To explore the feasibility of

33

combustion in the micro annular tube, the parameters influencing the combustion namely,

34

quenching diameter, and flammability were studied through numerical simulation. The results

35

indicated that combustion could be realized in micro- combustor using heat recirculation.

36

Following results were obtained from simulation. The quenching diameter reduced from 1.3

37

mm to 0.9 mm for heat recirculation at equivalence ratio of 1; the lean flammability was

38

2.5%-5% lower than that of without heat recirculation for quenching diameters between 2

39

mm to 5 mm. The overall energy conversion efficiency varied at different inner diameters. A

40

maximum efficiency of 3.15% was achieved at an inner diameter of 1 mm. The studies

41

indicated that heat recirculation is an effective strategy to maintain combustion and to

42

improve combustion limits in micro-scale system

43 44

Keywords: Micro-combustor; Equivalence ratio; Flammability, Quenching diameter;

45

Quenching cross sectional area; Energy conversion efficiency

46 47 48

Nomenclature C

Tube perimeter (m)

d

Tube diameter (mm)

49

D

Diffusion coefficient (m2/s)

E

Total energy (kJ)

50

hc

Lower heating value (kJ/kmole)

hH

51

L

Tube length (mm)

T

52

Tout

Gas temperature in counter flow (K)

53

w F 

Chemical reaction rate (kmole/m3∙s)

u



2

Heat transfer coefficient (kW/m2∙K) Gas temperature (K) Velocity (m/s) Dynamic viscosity (Pa∙s)

Page 2 of 28

54



Kinematic viscosity (m2/s)

55

cP

Specific heat (kJ/kg∙K)

56







Thermal conductivity (kW/m∙K)



Efficiency

Fluid density (kg/m3)

Equivalence ratio

57

LFL

Lean flammability limit

MEMS Micro-electromechanical system

58

MIT

Massachusetts Institute of Technology

TPV

59

TE

Thermoelectric

Thermophotovoltaic

60 61

1. Introduction

62

The micro-electromechanical systems (MEMS) experienced growing interest during the

63

past few years. The combustion of hydrogen or hydrocarbon fuels such as methane in MEMS

64

to produce electrical power has several advantages over the batteries because of high specific

65

energy of liquid fuels. The specific energy of liquid hydrocarbons is 35 to 300 times higher

66

than that of batteries built on latest technology [1]. For example, the specific energy of

67

methane is 50 MJ/kg whereas for an alkaline battery it is only 0.6 MJ/kg [1]. Thus, the direct

68

conversion from chemical to electrical energy even at 10% efficiency is attractive.

69

The hydrocarbon fuel based MEMS were found promising for application in micro

70

combustion. The MIT gas turbine laboratory developed MEMS based on gas turbine power

71

generator with an approximate total volume of 300 mm3 to produce 10~20 W of electric

72

power [2]. At the Combustion laboratories of UC Berkeley, research was conducted to

73

develop a liquid hydrocarbon fueled internal combustion rotary engine on millimeter scale [3].

74

A centimeter magnitude thermoelectric (TE) power generator integrated with plat-flame

75

micro combustor system was developed by Jiang et al. [4]. The maximum power output

3

Page 3 of 28

76

reached was 2 W, and the maximum overall chemical-electrical energy conversion efficiency

77

was 1.25% [4]. In addition to electric power generation, a micro-combustor was employed to

78

recirculate heat to produce hydrogen from ammonia [5]. Also, a thin-film-coated combustor

79

and a packed-bed combustor were designed and fabricated to operate as a heat source for a

80

methanol micro-reformer [6].

81

Even though the micro-power systems hold promise, the primary issue in them is

82

obtaining sustainable combustion in a micro-combustor. As the combustor size decreases, the

83

surface area to volume ratio increases. Because of the high surface area to volume ratio for

84

micro-combustor, there is a high amount of heat loss (proportional to area) compared to heat

85

energy generated (proportional to volume). Thus, the quenching and flammability problems

86

are more critical in a micro-scale combustor. On the other hand, the high surface area to

87

volume ratio characteristic of the micro combustor is most suitable for TPV systems. The

88

TPV system consists of three main parts: a heat source (combustor), a selective emitter, and a

89

photovoltaic array. The photovoltaic array converts the heat radiation absorbed from

90

combustion to electricity. Thus s smaller system will have higher energy conversion

91

efficiency due to its relatively larger surface area to volume ratio as long as the combustion is

92

sustainable. To maintain an optimal balance between the sustainable combustion and

93

maximum heat output is the main issue for micro TPV system.

94

The above concern necessitates innovative schemes to improve the performance of

95

micro-combustor. Several energy management methods including external heating,

96

backward-facing step, catalyzed combustion, and heat recirculation were employed to

97

improve the combustion in the MEMS[7]. The gas flow rates, equivalence ratio, and wall

4

Page 4 of 28

98

material are the main parameters which affect the combustion characteristics and heat loss in

99

a micro-combustor. Li et al. found that maximum heat was released when equivalence ratio

100

was slightly greater than one (and stainless steel wall resulted in more heat loss than ceramic

101

wall) [8]. Leach et al. investigated the effect of structural heat exchange and heat loss on the

102

power density and flame stability in order to optimize the design of silicon micro-combustors

103

[9]. Moreover, the geometry of combustor also has an important role in the flame stability.

104

Zhong et al. conducted experiments on micro Swiss-roll combustors with premixed CH4/air

105

mixture [10]. The combustion stability in the central regions of the combustors was enhanced

106

and the extinction limits of methane/air mixtures were significantly extended [10]. Li et al.

107

used backward-facing step in the combustor to effectively stabilize the flame position [11].

108

Fan et al. found that a triangular and semicircular bluff bodies significantly increased blow-

109

off limit of hydrogen/air flame in a planar micro-combustor [12]. Yang et al. investigated the

110

effect of wall thickness on the performance of three micro-cylindrical SiC combustors and

111

found that wall thickness of 0.4 mm gave maximum power output from micro-

112

thermophotovoltaic power generator [13].

113

Li et al. used porous medium in a 1 mm planar micro-combustor to enhance the flame

114

stability which included flame flashback, blow-off, and heat transfer [7]. Recently, catalysts

115

were employed to improve the combustion and heat transfer in a micro-combustor. Wang et

116

al. found that catalyst can effectively inhibit extinction and actively promote air-hydrogen

117

lean mixture reaction in the combustor [14]. In order to improve the efficiency and

118

performance of mico-combustors, the exhaust gas was recirculated to heat the outer wall of

5

Page 5 of 28

119

the micro-combustor and the incoming cold reactants. As result, the mean wall temperature,

120

total radiation energy emitted and useful radiation energy were improved [15].

121

In addition to experimental studies, extensive numerical simulations were developed by

122

various investigators to evaluate the combustion and heat transfer performance of the micro-

123

combustor. The flame stability in different numerical models was studied at different

124

Reynolds and swirl numbers for MEMS [16, 17]. Fanaee et al. used a two-dimensional model

125

to investigate the effects of reaction zone thickness, maximum temperature and quenching

126

distance on combustion phenomenon in micro-combustors under catalytic and non-catalytic

127

conditions and obtained acceptable agreement between the analytical and experimental data

128

[18]. Pan et al. investigated the effects of porous media, hydrogen to oxygen equivalence ratio,

129

porosity and fuel mixture flow rates on the performance of the micro-combustor [19]. Li et al.

130

found that in 1D cylindrical micro-combustor model hydrogen was superior to methane and

131

propane as a fuel owing to its higher flame temperature and lower flame thickness. [20]. Tang

132

et al. conducted studies on premixed hydrogen/air combustion in 3D model of a micro planar

133

combustor and found enhancement in heat transfer and increase in mean temperature of the

134

radiation wall by inserting plates in the micro-combustor chamber [21].

135

The effect of cross-sectional geometry on the ignition/extinction behavior of catalytic micro-

136

combustors using CFD models was studied by Benedetto [22]. He found that square cross-

137

section channel was more resistant to extinction compared to circular channel [22]. Moreover,

138

heat recirculation from the post-flame to the pre-flame in a micro combustor improved the

139

flame stabilization and enhanced burning rate on a 2D mode [23].

6

Page 6 of 28

140

The objective of this work is to develop a one-dimensional numerical model using

141

FORTRAN code to explore the feasibility of CH4/air combustion in a micro-combustor

142

(straight tube combustor with heat recirculation) and investigate on its application to TPV

143

power system. The effect of heat recirculation of a micro-scale counter flow combustor on the

144

lean flammability, quenching diameter, and overall energy conversion efficiency were

145

investigated in this study.

146 147 148

2. Numerical Modeling and Its Formulation 2.1 Physical Model of the Micro-combustor

149

Figure 1 illustrates the physical model of the micro combustor used in this study. It

150

includes a combustion tube made of carbon steel with inner diameter varying from 1 mm to 3

151

mm and a concentric outer tube of same material which served as a counter flow heat

152

exchanger. The incoming cold mixture in the inner tube was preheated by the hot combustion

153

gases in the outer tube resulting in reduction of the chemical reaction time and helping

154

complete combustion. For a micro-size combustor, the reduced diffusion time resulting from

155

small size has a crucial effect on making the concentration distributions uniform near the

156

flame in the combustor. In addition, the laminar flow prevented rapid mixing between fuel

157

and air due to the low mass and heat transfer coefficients in the combustor. Therefore, a

158

premixed combustion instead of diffusion combustion was assumed in this study. The flame

159

propagation in the current model is two-dimensional in nature. However, if interest is limited

160

to micro-scale burners with tube diameter and wall thickness much smaller than the tube

161

length, the problem could be simplified to one-dimensional. This study assumed one

162

dimensional laminar plug flow in the model.

7

Page 7 of 28

163

The baseline micro-combustor configuration consisted of two annular tubes with an inner

164

tube of 3 mm in diameter and 30 mm in length and an outer tube of 4.2 mm in diameter and

165

30 mm in length. Cold premixed fuel mixture which passes through inner tube will be ignited

166

by hot gases in outer tube. In the mean time the heat radiation from high temperature outer

167

tube wall will reach TPV system and be converted to electricity.

168

2.2 Heat Transfer:

169 170

The Nusselt number Nu for the inner tube was determined by using the following empirical correlation for constant wall temperature [24] Nu 

h in n e r d i k

171



4 .3 6 4  B i 1  0 .2 6 8 2 B i

(1)

172

In the above relationship, hinner is the convective heat transfer coefficient, di is the diameter of

173

inner tube, k is the thermal conductivity of the fluid, and Bi is the Biot number of the micro-

174

tube. The dimensionless Bi characterizes the heat transfer resistance "inside" a solid body and

175

it was defined as: Bi 

176

h Lc

(2)

k

177

Where LC is characteristic length, which is commonly defined as the volume of the

178

body. This empirical correlation is applicable to the forced convection laminar flow in

179

circular duct. The Biot number is 0.006 for steel tube of dimensions shown in Figure 1.

180

Figure 2 shows the heat transfer model for the micro-combustor proposed. The convective

181

heat transfer coefficient at the outer surface of inner tube is denoted as houter,i. Heat is

182

dissipated via radiation from the outer surface of outer tube with heat transfer coefficient

183

houter,o. The Bi number

in

equation1is close to zero as the micro-combustor is small

8

Page 8 of 28

184

enough. This corresponds to a heat transfer coefficient of 111 W/m2.k [24]. For outer tube

185

section with associated hydraulic diameter dh (the different between outer diameter do and

186

inner diameter di), the outer surface Nu is 5.38 and the inner surface Nu is 4.602 for limited

187

wall condition and Nu is 4.364 for the inner tube[25]. The estimation of heat transfer

188

coefficient is very coarse in the current model. In practice, the heat transfer coefficient can be

189

enhanced for rough and grooved wall. In this model, to be simple, it was assumed a constant

190

convective heat transfer coefficient for both inner and outer tubes (e.g., hH.outerr,o=

191

hH.inner,o=hH.inner,i=111 w/m2.K).

192 193 194 195

2.3 Combustion Methane consumption rate during combustion can be expressed by Eq.(3) [26]: w F   A T P n

m

exp(

E

a

)[C H 4 ] [O 2 ]

RT

b

(3)

196

Where a=0.3, b=1.3, n=0, and m=0; A is the pre-exponential factor which is dimensionless

197

and equals to 1.3x109 here. P is pressure, E is activation energy and E=202,408 kJ/kmole, R is

198

universal gas constant, and

199

kmol/m3 , respectively.

200

Fuel and oxygen mass conservation equations were established as follow:

201

[C H 4 ]

u

and

[O 2 ]

2

dyF

 D

d yF

dx

202

u

dyO dx

are concentration of CH4 and O2 with the unit of

dx

 w F

2

(4)

2

2

 D

d yO dx

2

2

 w O

2

9

(5)

Page 9 of 28

203

In the above relationships,  is the gas density, u is the velocity, D is diffusion coefficient,

204

and yF is fuel mass fraction and yO2 is the oxygen mass fraction in air-fuel mixture.

205

At the entrance to the combustor, where x=0, yF=yF.0, yO2=yO2.0, and T=T0.

206

The Eqs. (4) and (5) can be combined and written in terms of Schvab-Zeldovich variable as

207

shown by Eq.(6): u

d  2

 D

dx

208 209

d

dx

2

(6)

In the above relationship,   

o2  F

 yF 

210 o  2

yo



2

o2

2 k m o le O 2 k m o le C H

211

and

212

Based on the assumptions listed above, the fluid flow field was divided into a finite number

213

of small control volumes as illustrated in Figure 3. The first law of thermodynamics as shown

214

by Eq.(7) was applied to each control volume:

215

dE dt

4

 Q  W   m i h to t , i   m e h to t , e

Q

(7)

Where, E is the total energy of the fuel,

217

inlet and outlet mass flow rate, respectively, and

218

enthalpy of formation, and

219

fueled with diesel and natural gas was conducted by Gümüş et al[27].

h

is heat transfer rate,

W

216

h to t  h f  h

is work,

mi

and

is the total enthalpy,

mi

hf

are

is

is thermal enthalpy. Similar energy analysis for a CI engine

10

Page 10 of 28

220

The equations (8) and (9) express the combined combustion and heat transfer processes in the

221

inner and outer tubes respectively:  u c p Ai

222 223 224 225 226

 u c p Ao

dT dx

dT dx

2

  Ai

d T dx

2

 h H C i ( T g , o  T )  w F h c A i

(8)

2

  Ao

d T dx

2

 h H C i ( T g , i  T )  h H , o u te r , o C o ( T w ,1  T )  w F h c A o

(9)

The Eq. (10) can be obtained by combining the equations (4) and (8). u

227

d

d  2

 D

dx

dx

2



hH C i hc Ai

(T g , o  T )

(10)

228

Where T is the temperature, cp is the specific heat of gas species, Ai cross area of the inner

229

tube, Ao is the cross area of outer tube, and

  h

t F



ht hc

 yF

where

ht  c pT

230

The Eq. (9) was used to determine the temperatures of gas and outer surface of the wall which

231

served as a selective emitter. For a photovoltaic array made of low band gap material such as

232

GaSb (gallium antimony), the efficiency of heat radiation to electrical energy transfer was

233

determined by Planck’s radiation law.

234

2.4 TPV System Modeling

235

Figure 4 illustrates a typical TPV system using a combustion heat radiation source. The

236

micro-TPV

system

consists

of

two

main

237

thermophotovoltaic convertor. Fuels combust in the micro-combustor and electricity will be

238

generated through photovoltaic cells by absorbing heat radiation from combustor. To

239

determine the spectrum emissivity, the wall temperatures along the outer surface of outer tube

240

need to be calculated. The black body hemispherical spectral emissive power is given by

241

Planck’s radiation law as shown by Eq.(11) [28].

11

components:

micro-combustor

and

Page 11 of 28

E b 

242

C 1 e

5

C 2 /(  T )

1

W/m3

(11)

243

Where λ is the wavelength, C

244

The range of spectrum from which a photon can be converted into electricity depends on the

245

type of TPV cells. In this study, GaSb cells were used as photovoltaic converter for which the

246

band gap was 0.72 eV. This implies that only photons emitted from a heat radiation source

247

with a wavelength smaller than 1.7 μm can generate electricity from the GaSb cells.

248 249 250

3. Results and Discussion

251

velocity and heat transfer coefficient on combustion performance in a micro-combustor were

252

studied. To better understand the effect of heat recirculation, the quenching diameters and

253

lean limits for combustion were compared with and without heat recirculation. The base case

254

conditions and sensitivity analysis parameters for this study are given in appendix A.

255

3.1 Quenching Diameter

256

3.1.1 Effect of Equivalence Ratio on Quenching Diameter

257

In this study, the equivalence ratio was defined as:

258

1

 3 .7 4 2  1 0

16

W m

2

,C

2

2

 1 .4 3 8 8  1 0 m  K

In this investigation, the influence of equivalence ratio, inner diameter, fuel inlet

ER 

s to ic h io m e tr ic a ir m o le s p e r m o le e m p itic a l fu e l a c tu a l a ir m o le s p e r m o le e m p itic a l fu e l



 A     F  s to ic h i  A     F  a c tu a l



 F     A  a c tu a l  F     A  s to ic h i

(12)

259

Figure 5 shows the variation of quenching diameter with respect to equivalence ratio (ER) for

260

CH4/air mixture with and without heat recirculation for the base case conditions. The model

261

without heat circulation is just a straight tube that has the same diameter as the inner tube of

12

Page 12 of 28

262

the model with heat circulation. With heat recirculation, the quenching diameter decreased,

263

especially at low values of equivalence ratio, which means that heat recirculation can help

264

combustion to sustain in micro-combustor at smaller scale. When the combustor scales down,

265

the surface area to volume ratio increases and the circulating heat compensates for part of the

266

heat lost to the environment resulting in the combustion to sustain inside the inner tube. The

267

stoichiometric calculations of premixed CH4/air combustion indicated that the diameter can

268

be as small as 0.9 mm for a combustor with heat recirculation whereas the minimum

269

diameter was 1.3 mm for combustor without heat recirculation.

270 271

3.1.2 Effect of Mixture Inlet Velocity on Quenching Diameter

272

The fundamental time constraint can be quantified in terms of a homogeneous

273

Damkohler number, which is the ratio of the residence time and characteristic chemical

274

reaction time [29] as shown by Eq. (13):

275

D ah 

 r e s id e n c e

(13)

 c h e m ic a l

276

To maintain stable combustion inside the tube, more fuel residence time than chemical

277

reaction time is necessary to have the Damkohler number greater than 1. The residence time

278

τresidence is defined as:

279 280 281 282

 r e s id e n c e 

L

(14)

v

Where L is the length of the combustor and v is the gas velocity. For a micro-combustion tube, the residence time is very small due to relatively high flow velocity and short tube length. The chemical reaction time has an inverse relationship with

13

Page 13 of 28

283

combustion temperature. In the current model, the heat transfer from outer tube to inner tube

284

increased reactants temperature and hence decreased chemical reaction time. Higher fuel inlet

285

velocity would increase power output, but a very high inlet velocity requires a larger tube

286

diameter to sustain combustion inside the tube. Figure 6 illustrates the relationship between

287

air/fuel inlet velocity and quenching diameter with heat circulation.. With heat recirculation,

288

the mixture could be burnt in micro combustor at mean flow velocities four to six times

289

higher than the stoichiometric laminar burn velocity (.40~.50 m/s). From Figure 6, for

290

velocities less than 2m/s, increasing velocity would increase fuel consumption rate and lead

291

to a decreased quenching diameter. For velocities higher than 2 m/s, the flame advanced

292

towards the end of inner tube due to the decreased residence time. This caused reduction of

293

chemical energy release, and hence required a larger quenching diameter to sustain the

294

combustion. The high velocity extinction limit was certainly caused by insufficient residence

295

time rather than the chemical reaction completion time.

296

At a fixed tube diameter, either extremely high or low Reynolds number (e.g., inlet velocity)

297

would lead to quenching. This is due to the fact that the mixture at low Reynolds number has

298

less amount of heat energy to sustain the combustion, and mixture at high Reynolds number

299

leads to insufficient residence time compared to the time required to complete the combustion

300

reactions. The calculations show that the Reynolds number as low as 53 can sustain

301

combustion with heat recirculation at tube diameter of 2 mm and the energy input of 10 W.

302

For a fixed tube of diameter of 3 mm, the maximum heat input is 70 W and corresponding

303

Reynolds number is 930.

304

3.2 Lean Flammability Limit

14

Page 14 of 28

305

When the equivalence ratio decreased from base case value of 0.83 to a value called lean

306

flammability limit (LFL), the flame failed from being stabilized. It is reasonable to expect

307

that the limits of flammability would be widened if preheating of inlet mixture by heat

308

recirculation is available. Figure 7 shows the lean flammability of micro combustor with and

309

without heat recirculation at different inner tube diameters and with inlet air/fuel mixture

310

velocity of 1 m/s. Both LFL for cases with and without heat circulation decreased along with

311

increased diameter. The LFL with heat circulation was 2.5%-5% lower than that of without

312

heat circulation for tube diameters between 2 mm and 5 mm. It indicates that heat

313

recirculation can enhance combustion at lower equivalence ratio. Similar results were

314

reported by Veeraragavan et al. [23]. Heat recirculation was found as the governing

315

parameter in enhancing the flame temperature and speed [23].

316

3.3 Overall Energy Conversion Efficiency

317

3.3.1 The Effect of Heat Transfer Coefficient

318

For tmicro-combustion-TVP system, the conversion of thermal energy into electrical

319

energy is obtained through the thermophotovoltaic cells which cover the outer surface of

320

micro-combustor to absorb heat radiation. The overall energy conversion efficiency is defined

321

as following:

322 323

 con 

e le c tr ic ity g e n e r a te d

(15)

c h e m ic a l e n e r g y in p u t

324

The effect of heat transfer coefficient on conversion efficient has been studied. Figure 8

325

illustrates the variation of overall energy conversion efficiency ηcon with respect to heat

326

transfer coefficient for base case combustion with heat recirculation. Higher heat transfer

15

Page 15 of 28

327

coefficients would enhance heat recirculation which resulted in higher energy conversion

328

efficiency. In addition, a broader flame flammability and smaller quenching diameter were

329

obtained by increasing the heat transfer coefficient which can be simply realized through

330

grooved or rough surface of tube.

331

3.3.2 The Effect of Inner Tube Diameter

332

The effect of inner tube diameter on overall energy conversion efficiency has also been

333

studied. Figure 9 shows efficiency at various diameters for stoichiometric mixture at inlet

334

velocity of 1m/s with heat circulation. It is not surprised to see that for diameters less than 1.5

335

mm or greater than 3 mm, the efficiency dropped at larger diameter due to lower surface area

336

to volume ratio at larger diameter. However, for diameters located in between 1.5 mm and 3

337

mm, the efficiency increased with increased diameter. In a micro-combustor system, both

338

wall temperature and surface area to volume ratio would influence the heat radiation and

339

hence the overall energy conversion efficiency. For diameters less than 1.5 mm, wall

340

temperature increased with increased diameter as shown in Figure 9 but surface area to

341

volume ratio decreases and dominates heat radiation. This caused efficiency to decrease.

342

When diameter increases from 1.5 mm to 3 mm, the increased wall temperature dominated

343

the heat radiation which caused efficiency to increase. As diameter continued to increase, the

344

wall temperature increase slowed down. For this situation, the surface area to volume ratio

345

dominated heat radiation again and efficiency dropped. The influence of surface area to

346

volume ratio on conversion efficiency observed by numerical simulation is yet to be validated

347

by experiments. Generally, MEMS are suffering from high surface area to volume ratio

16

Page 16 of 28

348

which leads to high heat loss to wall of combustor and flame quenching. However, this

349

feature is advantageous for direct energy conversion MEMS.

350

4. Conclusions:

351

In this study, a theoretical model was developed to simulate the combustion of CH4/air

352

fuel mixture in a micro-combustor with and without heat recirculation. The energy

353

conversion efficiency of the micro-combustor-TPV system was evaluated. Following are the

354

conclusion from this study:

355

1.

356

by reducing the heat losses from the system. A flame could be established in the micro-

357

combustor even at a small Reynolds number (Re=53) and energy input of around 10 W with

358

heat recirculation and inner tube diameter of 2 mm. For a fixed inner tube diameter of 3 mm,

359

the maximum heat input can be up to 70 W without flame blow off.

360

2.

361

with heat recirculation, the flammability range was broadened compared to that of without

362

heat recirculation. Also, the lean limit of flammability in air was 2.55%-5% lower than that of

363

without heat recirculation. With heat recirculation, the combustion can sustain at mean flow

364

velocities four to six times the stoichiometric laminar combustion velocity.

365

3.

366

combustor without heat recirculation. The quenching diameter can be as low as 0.9 mm with

367

an equivalence ratio of 1. In addition, the higher heat transfer coefficient between gas and

368

wall improved the energy conversion efficiency.

Combustion in a micro scale combustor with inner tube diameter of 3 mm was achieved

The flammability range decreased with decreasing diameter of combustor. However,

The quenching diameter of combustor with heat recirculation was smaller than that of the

17

Page 17 of 28

369

4.

370

the most important parameters affecting the overall energy conversion efficiency. The

371

maximum conversion efficiency (3.15%) was achieved for case with 1 mm inner diameter.

372 373 374

Acknowledgement:

375

for the Central University- 2015Q NA11”.

376 377

5. Reference

378

[1]

The authors wish to acknowledge the financial support by “The Fundamental Research Funds

379 380

A. C. Fernandez-Pello, "Micropower generation using combustion: issues and approaches," Proceedings of the Combustion Institute, vol. 29, pp. 883-899, 2002.

[2]

381 382

The surface areas to volume ratio of the micro-combustor and the wall temperature were

Epstein A H, Senturia S D, Anathasuresh G, Ayon A, and B. K, "Power MEMS and microengines " Proc IEEE Transducers 97 conf. (Chicago) pp. 753-756, 1997.

[3]

K. Fu, A. Knobloch, F. Martinez, D. Walther, and A. C. Fernandez-Pello, "Design and

383

experimental results of small-scale rotary engines " Proc. 2001 ASME International

384

Mechanical Engineering Congress and Exposition IMECE2001/MEMS-23924, 2001.

385

[4]

L. Q. Jiang, D. Zhao, C. M. Guo, and X. H. Wang, "Experimental study of a plat-

386

flame micro combustor burning DME for thermoelectric power generation," Energy

387

Conversion and Management, vol. 52, pp. 596–602, 2010.

388

[5]

J H Kim and O. C. Kwon, "A micro reforming system integrated with a heat-

389

recirculating micro-combustor to produce hydrogen from ammonia," International

390

Journal of Hydrogen Energy, pp. 1974-1983, 2011.

18

Page 18 of 28

391

[6]

J. k. Jin and S. Kwon, "Fabrication and performance test of catalytic micro-

392

combustors as a heat source of methanol steam reformer," International Journal of

393

Hydrogen Energy, vol. 35, pp. 1 8 0 3 – 1 8 1 1, 2010.

394

[7]

J. Li, Y. Wang, J. Shi, and X. Liu, "Dynamic behaviors of premixed hydrogen–air

395

flames in a planar micro-combustor filled with porous medium," Fuel, vol. 145, pp.

396

70-78, 2015.

397

[8]

L. Junwei and Z. Beijing, "Experimental investigation on heat loss and combustion in

398

methane/oxygen micro-tube combustor," Applied Thermal Engineering, vol. 28, pp.

399

707-716, 2008.

400

[9]

T. T. Leach and C. P. Cadou, "The role of structural heat exchange and heat loss in

401

the design of efficient silicon micro-combustors," Proceedings of the Combustion

402

Institute, vol. 30, pp. 2437–2444, 2005.

403

[10]

Z. B. Jing and W. J. Hua, "Experimental study on premixed CH4/air mixture

404

combustion in micro Swiss-roll combustors," combustion and Flame, vol. 157, pp.

405

2222-2229, 2010.

406

[11]

J. Li, S K Chou, G Huang, W M Yang, and Z. W. Li, "Study on premixed combustion

407

in cylindrical micro combustors: Transient flame behavior and wall heat flux,"

408

Experimental Thermal and Fluid Science, vol. 33, pp. 764–773, 2009.

409

[12]

A. Fan, J. Wan, Y. Liu, B. Pi, H. Yao, and W. Liu, "Effect of bluff body shape on the

410

blow-off limit of hydrogen/air flame in a planar micro-combustor," Apply Thermal

411

Engineering vol. 62, pp. 13-19, 2014.

19

Page 19 of 28

412

[13]

Y. Wenming, C. Siawkiang, S. Chang, X. Hong, and L. Zhiwang, "Effect of wall

413

thickness of micro-combustor on the performance of micro-thermophotovoltaic power

414

generators," Sensors and Actuators A, vol. 119, pp. 441-445, 2005.

415

[14]

Y. Wang, Z. Zhou, W. Yang, J. Zhou, J. Liu, Z. Wang, et al., "Combustion of

416

hydrogen-air in micro combustors with catalytic Pt layer," Energy Conversion and

417

Management, vol. 51, pp. 1127–1133, 2010.

418

[15]

W. Yang, S. Chou, K. Chua, H. An, K. Karthikeyan, and X. Zhao, "An advanced

419

micro modular combustor-radiator with heat recuperation for micro-TPV system

420

application," Applied Energy, vol. 97, pp. 749–753, 2012.

421

[16]

B.-i. Choi, Yong-shikHan, Myung-baeKim, Cheol-hongHwang, and C. B. Oh,

422

"Experimentalandnumericalstudiesofmixingandflamestabilityina

423

cyclonecombustor," Chemical EngineeringScience, vol. 64, pp. 5276-5286, 2009.

424

[17]

micro-

H. Wang, K. Luo, S. Lu, and J. Fan, "Direct numerical simulation and analysis of a

425

hydrogen/air swirling premixed flame in a micro combustor," International Journal of

426

Hydrogen Energy, vol. 36, pp. 13838-13849, 2011.

427

[18]

S. A. Fanaee and J. A. Esfahani, "Two-dimensional analytical model of flame

428

characteristic in catalytic micro-combustors for a hydrogeneair mixture," I n t e rna t i

429

onal journal o f hydrogen energy, vol. 39, pp. 4 6 0 0-4 6 1 0, 2014.

430

[19]

J. F. Pan, D. Wu, Y. X. Liu, H. F. Zhang, A K Tang, and H. Xue, "Hydrogen/oxygen

431

premixed combustion characteristics in micro porous media combustor," Energy

432

Procedia, vol. 61, pp. 1279 – 1285, 2014.

20

Page 20 of 28

433

[20]

J. Li, S. K. Chou, Z. Li, and W. Yang, "Development of 1D model for the analysis of

434

heat transport in cylindrical micro combustors," Applied Thermal Engineering, vol. 29,

435

pp. 1854–1863, 2009.

436

[21]

A. Tang, J. Pan, W. Yang, Y. Xu, and Z. Hou, "Numerical study of premixed

437

hydrogen/air combustion in a micro planar combustor with parallel separating plates,"

438

International Journal of Hydrogen Energy, vol. 40, pp. 2 3 9 6-2 4 0 3, 2015.

439

[22]

A. D. Benedetto, V. D. Sarli, and G. Russo, "Effect of geometry on the thermal

440

behavior of catalytic micro-combustors," Catalysis Today, vol. 155, pp. 116-122,

441

2010.

442

[23]

A. Veeraragavan and C. Cadou, "Flame speed predictions in planar micro/mesoscale

443

combustors with conjugate heat transfer," Combustion and Flame, vol. 158, pp. 2178–

444

2187, 2011.

445

[24]

446 447

W. M. Rohsenow, Handbook of Heat Transfer: 3rd (Third) edition: New York, McGraw-Hill Companies, 1998.

[25]

448

I. F. P and D. D. P, Fundamentals of Heat and Mass Transfer 4th New York: John Wiley and Sons, 1996.

449

[26]

G. L. Borman, Combustion engineering: Boston : McGraw-Hill, 1998.

450

[27]

M. Gumus and M. Atmaca, "Energy and Exergy Analyses Applied to a CI Engine

451 452 453

Fueled with Diesel and Natural Gas," Energy Sources, vol. 35, pp. 1017-1027, 2013. [28]

T. L. Bergman, A. S. Lavine, F. P. Incropera, and D. P. DeWitt, Fundamentals of Heat and Mass Transfer: Wiley,7 edition, 2011.

21

Page 21 of 28

454

[29]

455 456 457

K. Annamalai and I. K. Puri, "Combustion science and Engineering " vol. Taylor& Francis Group, 2007.

[30]

S. K. Chou. (2004). Development of a Novel

Micro Thermophotovoltaic Power

Generator Available: http://www.eng.nus.edu.sg/EResnews/0310/rd/rd_4.html

458 459 460

22

Page 22 of 28

461

Appendix A

462 463 464 465

Base case conditions and sensitivity study This study started for a base case with parameters listed in Table 1. Table 1: Base case data Diameter of inner tube,

di

3

mm

Diameter of outer tube,

do

4.2

mm

30

mm

1

m/s

Tube length,

L

Inlet fuel mixture velocity,V0

u

Equivalence ratio

0.83

Emissivity of outer tube

0.93

Effective wavelength (GaSb)

< 1.7

m

466 467

In addition to the base case, sensitivity studies with different combustor configurations and

468

fuel properties as shown in Table 2 were also performed.

469 470

Table 2: Parametric studies Inlet velocity, u Equivalence ratio,  Inner diameter, d i Heat transfer coefficient, h H

m/s mm W/m2K

0.5~3 0.4~1.2 0.5~ 5 50~200

471 472 473

23

Page 23 of 28

inflow outflow

4.2 mm

3mm

outflow

30 mm

474 475 476

Figure 1. Baseline micro-combustor in the theoretical model (carbon steel)

477 478 479

Figure 2. Heat Transfer Model for Micro-combustor

480

24

Page 24 of 28

481

482 483

Figure 3.Energy conservation in a control volume inside the inner tube

484 485

486 487 488

Figure 4 . Micro-TPV system [30]

489

25

Page 25 of 28

490 Equivalence ratio vs quenching diameter with heat recirculation

without heat recirculation

3

Quenching diameter (mm)

2.5 2 1.5 1 0.5

0 0

491 492 493 494 495

0.2

0.4

0.6 0.8 Equivalence ratio

1

1.2

1.4

Figure 5. CH4/air quenching diameter at different equivalence ratios

1.1

Quenching diameter (mm)

1 0.9 0.8 0.7 0.6

0.5 0.4

0

0.5

1

1.5

2

2.5

3

Velocity (m/s)

496 497 498

Figure 6. Variation of quenching diameter Vs inlet velocity with heat circulation

26

Page 26 of 28

Flammability limit(% of fuel by volume)

499

9.5

lean limit with heat circulation

9

lean limit without heat circulation

8.5 8 7.5 7 0

500 501 502 503

1

2

3 Diameter (mm)

4

5

6

Figure 7. Lean limit of flammability in air (% of fuel by volume)

Efficiency %

3

2

1

0

504 505 506 507 508

0

50

100 150 200 Heat transfer coefficient(W/m2K)

250

Figure 8. Effect of heat transfer coefficient on overall energy conversion efficiency  with heat circulation

27

con

Page 27 of 28

509

1200

0.03

1000

Tw

0.025

Overall efficiency

800 0.02 600 0.015

overall

400 0.01 200

0.005 0

510 511 512 513 514

Wall temperature (K)

0.035

0

2

4

6

8

10

12

0

Diameter (mm)

Figure 9.Variation of energy conversion efficiency Vs inner tube diameter with heat circulation

28

Page 28 of 28