Você está na página 1de 9

Energy Conversion and Management 120 (2016) 197205

Contents lists available at ScienceDirect

Energy Conversion and Management


journal homepage: www.elsevier.com/locate/enconman

Numerical study on a multiple-channel micro combustor


for a micro-thermophotovoltaic system
Yang Su a, Qiang Cheng a,b,, Jinlin Song a, Mengting Si a
a
b

State Key Laboratory of Coal Combustion, Huazhong University of Science and Technology, Wuhan 430074, PR China
Shenzhen Institute of Huazhong University of Science & Technology, Shenzhen 518057, PR China

a r t i c l e

i n f o

Article history:
Received 28 January 2016
Received in revised form 4 April 2016
Accepted 24 April 2016

Keywords:
Micro combustor
Multiple-channel
Premixed H2/air combustion
Thermophotovoltaic

a b s t r a c t
As a key component, the design of a micro combustor is critical when determining the performance of a
micro-thermophotovoltaic (TPV) system. This study proposes a multiple-channel micro combustor for a
micro-TPV system. A three-dimensional computational fluid dynamics (CFD) model with a skeletal reaction mechanism embedded is established for premixed hydrogen/air combustion in the micro combustors. The numerical simulation results indicate that the temperature distribution along the outer wall
of the multiple-channel micro combustor is more uniform and higher than that of the traditional
single-channel micro combustor, which is highly beneficial for the micro-TPV system. Moreover, the
radiation energy and radiation efficiency of the multiple-channel combustor are higher than those of
the single-channel combustor at the same volume flow rate. When the flow rate is 100 cm3/s and
H2/air equivalence ratio is 0.8, the mean temperature and the wall temperature difference on the
cross-sectional direction for the multiple-channel combustor are 1294.9 and 86.6 K, respectively,
whereas those for the single-channel combustor are 1256.0 and 107.6 K, respectively. The radiation
energy and radiation efficiency are 53.57 W and 18.84% for the multiple-channel combustor, respectively,
and 47.40 W and 16.67% for the single-channel combustor, respectively. Moreover, the high equivalence
ratio assignment for the side channels can further improve the temperature uniformity of the multiplechannel combustor. Results show that the wall temperature difference on the cross-sectional direction
can decrease from 86.6 K to 68.8 K.
2016 Elsevier Ltd. All rights reserved.

1. Introduction
In the past decades, the micro electromechanical system technology has rapidly proliferated, which created a strong demand
for small-scale power supply devices that have high energy densities (i.e., small size, low weight, and long duration). Conventional
electrochemical batteries can no longer satisfy power supply
requirements because of certain disadvantages, such as low gravimetric and volumetric energy densities, short usage times, and
negative environmental impacts [1]. Several micro power generation devices that utilize the high specific energy from combustion
of hydrogen/hydrocarbon fuel, such as micro gas turbine [2], micro
thermoelectric device [3], micro swing engine [4], and microthermophotovoltaic (TPV) system [5], have been developed to
address the issue. The high surface-to-volume ratio of the micro
combustor is favorable for the micro-TPV system to achieve a high
Corresponding author at: State Key Laboratory of Coal Combustion, Huazhong
University of Science and Technology, Wuhan 430074, PR China.
E-mail address: chengqiang@mail.hust.edu.cn (Q. Cheng).
http://dx.doi.org/10.1016/j.enconman.2016.04.088
0196-8904/ 2016 Elsevier Ltd. All rights reserved.

power density and is different from other micro power generators


[6]. The micro-TPV system is easy to manufacture and assemble
because it does not include any moving components [7].
The micro-TPV system mainly consists of four primary parts,
namely, a heat source, an emitter, a dielectric filter, and a PV cell
[8]. The micro-TPV system utilizes the thermal energy released
by the chemical energy of fuel to heat up the wall of the micro
combustor through the combustion of hydrocarbon fuels. The
emitter, which is the outer surface of the micro combustor, emits
many radiation photons. Only the radiated photons with energy
higher than the band gap of the PV cell material can activate free
electrons and generate electricity under the action of a PN junction.
According to the StefanBoltzmann law, a high and uniform temperature distribution along the wall of the micro combustor favors
the performance of the micro-TPV system [9]. Therefore, the combustor, which significantly affects the performance of these
devices, is the most important part of the micro-TPV system. However, a series of challenges uniquely different from conventional
combustors resulted from the small confinement for combustion
[10]. One of the main problems is that heat loss increases because

198

Y. Su et al. / Energy Conversion and Management 120 (2016) 197205

Nomenclature
A
C in
C out
Hc
h
mH2
P
q
Tw
T1
Vin

outer surface area of the combustor, m2


hydrogen mass fraction at the inlet
hydrogen mass fraction at the outlet
higher calorific value of hydrogen, 1.396  105 kJ/kg
convective heat transfer coefficient, W/(m2 K)
hydrogen mass flow rate, kg/s
radiation energy, W
total heat loss rate, W/m2
outer wall temperature, K
ambient temperature, K
volume flow rate, cm3/s

of the significantly increased surface-to-volume ratio of the micro


combustor, which leads to flame quenching because of micro
combustor size reduction. Second, the residence time of the
reactants for combustion is significantly reduced and limited at
micro-scale, which results in incomplete combustion [11]. Then,
extinction occurs when the reactant residence time approaches
the characteristic combustion time because of incomplete combustion. Radical quenching may occur when active species generated
during combustion undergoes adsorption and depletion at the
walls of the high surface-to-volume micro combustor [12,13].
Various methods have been implemented to improve the performance of micro combustors in the past decades. Fan et al.
[1416] experimentally and numerically investigated combustion
behaviors and flame propagation modes in radial microchannel
under different conditions. Their works supplement the knowledge
on combustion characteristics and provide important information
on the practical applications of micro combustion. Yang et al.
[17] investigated the characteristics of a micro cylindrical combustor with different configurations and discovered that a backwardfacing step effectively enhanced the mixing of the fuel/air mixture
and control flame position. Lee and Kwon [18] assessed the
thermal performance of a micro cylindrical combustor with heat
recirculation. Heat recirculation can enhance the flame stability
limit and increase the wall temperature significantly. Catalytic
combustion is a viable approach to stabilize flame under small
scales because the catalyst can accelerate the reaction and suppress radical depletion at the wall [19]. Moreover, Chou et al.
[20,21] examined the porous media combustion characteristics in
micro cylindrical and planar combustors. Their work confirmed
that flame can be effectively anchored by the inserted porous
media. Furthermore, Fan and Wan [22,23] presented a micro planar combustor with bluff body or cavity, which can significantly
expand the blow-off limit and have promising results in terms of
combustion stability, because of the formed low-velocity and recirculation zone. A planar combustor is better than cylindrical tubes
in terms of emitting radiation normal to the PV cell given that
the radiating wall is placed facing the receiver, that is, the PV cells
[9]. Yang et al. [24] conducted a detailed comparison study
between cylindrical and planar micro combustors, and pointed
out that the planar combustor has a high radiation efficiency. However, the planar combustor exhibits a temperature difference not
only on the axial direction but also on the cross-sectional direction,
which is detrimental to the micro-TPV system.
A new type of multiple-channel micro planar combustor is proposed and numerically investigated to design a micro combustor
that possesses good combustion stability and good performance
on the outer wall temperature. The performances of the singlechannel and multiple-channel micro combustors are compared
by simulation with a detailed chemical reaction mechanism.

Greek letters
H2/air equivalence ratio
e
surface emissivity
r
StefanBoltzmann constant = 5.67  108 W/(m2 K4)
gre
radiation efficiency
gce
combustion efficiency

Subscript
i
incident

Different equivalence ratio assignments for every channel of this


type of micro combustor are applied, and the combustion characteristics of these assignments are explored.
2. Methodology
2.1. Geometric model
The specifications of the single-channel and multiple-channel
micro combustors for the micro-TPV system are shown in Fig. 1.
The micro combustors have a planar structure appearance, and
the overall dimensions are 25 mm long (X direction), 11.2 mm
wide (Y direction) and 2.4 mm tall (Z direction). The combustor
is made of 316 stainless steel, which can withstand high temperatures, and the physical properties of 316 stainless steel are shown
in Table 1. A backward-facing step is designed to produce recirculation to enhance the mixing process and control flame position of
the two combustors, and the distance from the inlet to the step is
set as 5 mm. The origin of the coordinates is set on the step section
to facilitate the description. However, five identical micro circular
channels are equidistantly arranged inside the multiple-channel
combustor.
2.2. Computational model
Fluids can be reasonably treated as continuums given that the
characteristic scale of the combustor chamber remains larger than
the molecular mean-free path of gases that flow through the micro
combustor. Therefore, the NS equations are suitable in the present
study [26]. Kuo and Ronney [27] emphasized that turbulence models are appropriate in estimating the combustion characteristic in
the micro combustor when the Reynolds number exceeds 500.
The standard ke model is utilized as a flow model during the
simulation process in this study [28].
Hydrogen and air are selected as fuel and oxidant, respectively.
The density and specific heat of the gas mixture are calculated
using the incompressible ideal gas law and mixing law. Gas thermal conductivity and viscosity are calculated from the mass fraction weighted average of all species. A skeletal mechanism
reported by Giovangigli and Smooke [29] is applied to model
H2/air mixture combustion (Table 2), which involves 9 species
and 19 reversible elementary reactions. The surface reaction effect
is neglected in the computation process.
For boundary conditions, a uniform velocity distribution of premixed H2/air is selected at the inlet with the temperature of 300 K.
A pressure outlet boundary condition is specified at the outlet
zone. Dufour effects, gas radiation, and viscous dissipation are
neglected in the computation process [30]. A non-slip and a zero

199

Y. Su et al. / Energy Conversion and Management 120 (2016) 197205

Fig. 1. Schematic of the micro combustor: (a) single-channel combustor, and (b) multiple-channel combustor (units: mm).

For energy conservation,

Table 1
Physical properties of 316 stainless steel [25].

q (kg/m3)
8.0  10

k (W/(m K))

Cp (J/(kg K))

Melting point (K)

16.3

502.48

1670

No.

Reactions

Ar ((cm3/mol)n1/s)

br

Er (J/kmol)

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19

O2 + H = OH + O
H2 + O = OH + H
H2 + OH = H2O + H
OH + OH = H2O + O
H2 + O2 = OH + OH
H + OH + M = H2O + Ma
O2 + M = O + O + M
H2 + M = H + H + M b
H + O2 + M = HO2 + Mc
H + O2 + O2 = HO2 + O2
H + O2 + N2 = HO2 + N2
HO2 + H = H2 + O2
HO2 + H = OH + OH
HO2 + O = OH + O2
HO2 + OH = H2O + O2
HO2 + HO2 = H2O2 + O2
H2O2 + M = OH + OH + M
H2O2 + H = H2 + HO2
H2O2 + OH = H2O + HO2

5.10E+13
1.80E+7
1.20E+6
6.00E+6
1.70E+10
7.50E+17
1.90E+8
2.20E+9
2.10E+12
6.70E+13
6.70E+13
2.50E+10
2.50E+11
4.80E+10
5.00E+10
2.00E+9
1.30E+14
1.70E+9
1.0E+10

0.82
1.00
1.30
1.30
0.00
2.60
0.50
0.50
1.00
1.42
1.42
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00

6.91E+7
3.70E+7
1.52E+7
0.00
2.0E+8
0.00
4.001E+8
3.877E+8
0.00
0.00
0.00
2.90E+6
7.90E+6
4.20E+6
4.20E+6
0.00
1.905E+8
1.57E+7
7.50E+6

(a)
1300
1200
1100
1000
900

10

12

14

16

18

20

Distance from step (mm)


1600

Outer wall temperature, Tw (K)

For continuity conservation,

54, 000 cells


91 , 000 cells
184, 260 cells

1400

800

where the convective heat transfer coefficient, h, and surface emissivity, e, are set as 10 W/(m2 K) and 0.75, respectively [31]. Tw is the
outer wall temperature, T1 is the ambient temperature with
T1 = 300 K,
and
r is the StefanBoltzmann constant
(=5.67  108 W/(m2 K4)).
The commercial software FLUENT 6.3.2 is selected for solving
the mass, momentum, energy, and species conservation equations
and the conjugated heat conduction in the solid wall. The governing equations are expressed as follows [30]:

102 ,616 cells


211 ,176 cells
337 ,400 cells

(b)

1500
1400
1300
1200
1100
1000
900

2
800

For momentum conservation,

@quj ui
@Pi @ sij


@xj
@xi @xj

1500

species flux normal to the wall surface boundary condition are


applied to the inner walls. The total heat loss rate at all outer walls
is calculated by Eq. (1):

@quj
0
@xj

@P
Sh
@xj

where Ri is the net production rate of species i by chemical reaction


and Jij is the diffusion flux of species i.
The first-order upwind scheme is applied to discretize the
model, and the SIMPLE algorithm is employed to deal with

Enhancement factors: H2O = 20.0.


Enhancement factors: H2O = 6.0, H = 2.0, and H2 = 3.0.
Enhancement factors: H2O = 21.0, H2 = 3.3, O2 = 0.0, and N2 = 0.0.

q hT w  T 1 erT 4w  T 41

uj

@J ij
@quj Y i

Ri
@xj
@xi

Outer wall temperature, Tw (K)

where Sh is the fluid enthalpy source.


For species conservation,

Table 2
Chemical reaction mechanism of H2/air combustion.

N
@quj h
@
@T X

k

hi J
@xj
@xi
@xj i1 ij

10

12

14

16

18

20

Distance from step (mm)

Fig. 2. Outer wall temperature profiles of the combustor at different mesh


densities: (a) single-channel combustor, and (b) multiple-channel combustor.

200

Y. Su et al. / Energy Conversion and Management 120 (2016) 197205

pressurevelocity coupling. A three-dimensional model is used for


the numerical simulation given the structural characteristic of the
new type of combustor. Gambit 2.4.6 is used for grid generation.
Further refinement of local meshes is conducted, and a nonuniform structured grid system is employed. Grid independency
is checked for single-channel and multiple-channel combustors,
with the H2/air equivalence ratio U = 0.8 and Vin = 100 cm3/s. The
results are illustrated in Fig. 2(a) and (b). The single-channel combustor shows that the outer wall temperature obtained by the
mesh with medium size (i.e., 91,000 cells) is close to that with finer
mesh (i.e., 184,260 cells). Fig. 2(b) shows that a mesh with 102,616
cells is sufficiently fine to capture the outer wall temperature of
the multiple-channel combustor. Therefore, meshes with 91,000
cells and 102,616 cells are selected for the single-channel and
multiple-channel combustors, respectively, to obtain a solution
for improved accuracy and reduced computational load. The residuals of all governing equations are adopted as 103, except for
energy, which is set as 106.
2.3. Validation of the computational model

Fig. 4. Comparison of the temperature distributions on the combustor center


section: (a) single-channel combustor, and (b) multiple-channel combustor.

The numerical results of the outer wall temperature of the


single cylindrical micro combustor have been compared with the
experimental data from [32] (Fig. 3) with the H2/air equivalence
ratio (U) equal to 0.8 to verify the accuracy of the computational
model. The average relative errors are approximately 2.8% and
2.2%, whereas the inlet velocities are 12 and 16 m/s. This result
indicates the applicability of the computational model adopted in
the present study.
3. Results and discussion
3.1. Comparison of the thermal performance of single-channel and
multiple-channel micro combustors
Fig. 4 shows the temperature distributions on the center section
(XY plane) of the two combustors when the total volume flow rate
(Vin) is fixed at 100 cm3/s, where the flow rates for every channel of
the multiple-channel combustor are the same in all tested cases,
and the H2/air equivalence ratio (U) is fixed at 0.8. Fig. 4 also shows
that the multiple-channel combustor has five separate flames and
the flame front presents an inverted V shape, whereas the singlechannel combustor has only one overall flame and the flame front
presents a semicircle shape. The mixture of hydrogen and air can

be ignited easily given the rapid flame speed [33]. The reactions
inside the two combustors can be completed in a small region,
which is only approximately 3 mm away from the step. Moreover,
the maximum luminance of the multiple-channel combustor is
slightly lower than that of the single-channel combustor. Therefore, the temperature is at a relatively low peak because of the
slightly incomplete combustion and high heat transfer in the
multiple-channel combustor.
Fig. 5 shows the outer wall temperature distributions of the two
combustors. The multiple-channel combustor is significantly
brighter than the single-channel combustor, which indicates a high
outer wall temperature. The mean temperatures are 1294.9 and
1256.0 K for the multiple-channel and single-channel combustors,
respectively, with an increment of 38.9 K. Where the mean
temperature means the area-weighted average temperature
of the selected face (0 6 X 6 20 mm, 5.6 mm 6 Y 6 5.6 mm,
Z = 1.2 mm). The high-temperature zone of the multiple-channel
combustor is significantly larger than that of the single-channel
combustor. The high-temperature zone of the multiplechannel combustor is slightly far from the step because the

1300
12 m/s, simulation
12 m/s, experiment [32]
16 m/s, simulation
16 m/s, experiment [32]

Outer wall temperature,Tw (K)

1200
1100
1000
900
800
700
600

10

12

14

16

18

Distance from step (mm)


Fig. 3. Comparison of the outer wall temperature distributions between the
simulation results and experimental data [32].

Fig. 5. Comparison of the temperature distributions on the combustor outer wall:


(a) single-channel combustor, and (b) multiple-channel combustor.

201

Y. Su et al. / Energy Conversion and Management 120 (2016) 197205

Fig. 6. Comparison of the temperature fields on the combustor outer wall at different flow rates: (a) Vin = 85 cm3/s, (b) Vin = 100 cm3/s, (c) Vin = 115 cm3/s, and (d)
Vin = 130 cm3/s.

configuration of multiple channels decreases the effective flow


area and increases the flow velocity of the gas mixture.
Numerical calculation is conducted at various volume flow rates
(Vin = 85, 100, 115, and 130 cm3/s) to explore this phenomenon
further. Fig. 6 shows the outer wall temperature distributions of
the two combustors for all tested cases. Similar trends can be
observed for different volume flow rates. That is, the multiplechannel combustor is significantly brighter than the singlechannel combustor, and the area of the high-temperature zone of
the multiple-channel combustor is extended compared with that
of the single-channel combustor for all investigated cases. Meanwhile, in contrast to the single-channel combustor, a more uniform
outer wall temperature has been obtained for the multiple-channel
micro combustor. The flame front shifts slightly away from the
step with the increase of flow rate because of reduced residence
time. However, this shift tendency is unclear for the multiplechannel combustor, which indicates an improved flame stability
performance.
Table 3 shows the mean temperature of the outer wall of the
single-channel and multiple-channel micro combustors at different flow rates with Vin = 85, 100, 115, and 130 cm3/s. Table 3 shows
that the mean wall temperature of the two combustors increases
with the increase of the rate. The mean temperature of the
multiple-channel combustor at all tested flow rates is higher than

that of the single-channel combustor, and the relative temperature


increase is 35.8, 38.9, 44.9, and 62 K. The mean temperature of the
multiple-channel combustor can reach 1275 K at the lowest flow
rate with 85 cm3/s, which is higher than the highest value of the
single-channel combustor among all investigated cases. Moreover,
the mean temperature of the single-channel combustor decreases
when the flow rate increases from 115 cm3/s to 130 cm3/s because
an increased amount of cold gas is blown into the combustion
chamber, which decreases the mean temperature.
Combustion efficiency has been widely recognized to exert a
crucial effect on micro scale combustion performance, which is calculated by Eq. (6):

gce 1 

C out
C in

where C in and C out are hydrogen mass fractions at the inlet and outlet of the micro combustor, respectively [34]. Table 4 shows the
combustion efficiency of the single-channel and multiple-channel
micro combustors at different flow rates. The combustion efficiency
of the multiple-channel combustor is slightly lower than that of the
single-channel combustor. This finding can be attributed to the relatively high flow velocity and subsequent incomplete combustion
in the multiple-channel combustor. Thus, the main reason for the
increasingly uniform outer wall temperature distribution formed
in the multiple-channel combustor is the increased heat transfer
intensity. The reduced flow area results in a higher flow velocity,

Table 3
Temperature performances under different flow rates.
Vin (cm3/s)

85
100
115
130

Table 4
Combustion efficiencies under different flow rates.

Mean temperature of the outer wall (K)


Single-channel

Multiple-channel

Relative increase

1239.5
1256.0
1260.9
1259.1

1275.3
1294.9
1305.8
1321.1

35.8
38.9
44.9
62

Combustion efficiency (%)

Single-channel
Multiple-channel

Vin (cm3/s)
85

100

115

130

99.54
98.82

99.31
98.63

99.07
98.33

98.25
97.91

202

Y. Su et al. / Energy Conversion and Management 120 (2016) 197205

1500

(a)
Outer wall temperature,Tw (K)

1400
1300
1200

Single-channel, Vin = 85 cm 3/s

1100

Multiple-channel,Vin = 85 cm 3/s

1000

Multiple-channel, Vin = 100 cm 3/s

Single-channel, Vin = 100 cm 3/s


Single-channel, Vin = 115 cm 3/s

900

Multiple-channel,Vin = 115 cm 3/s


Single-channel, Vin = 130 cm 3/s

800

Multiple-channel,Vin = 130 cm 3/s


700

10

12

14

16

18

20

Distance from step (mm)


1400

(b)
3.2. Comparison of the radiation performance of the single-channel
and multiple-channel micro combustors

1350

Outer wall temperature, Tw (K)

the high temperature gas has more chance to transfer heat to the
inner wall of the micro combustor, which increases the wall temperature. Moreover, the thermal conductivity of the inner ribs is
higher than that of the gas mixture. Thus, the inner ribs can help
deliver heat to the entire combustion chamber.
Fig. 7(a) and (b) shows the outer wall temperature distributions
along the axial (X) direction (Y = 0 mm) and the cross-sectional (Y)
direction (X = 11 mm) of the two combustors for all tested cases.
Fig. 7 indicates that the outer wall temperature of the multiplechannel combustor is higher and more uniform than that of the
single-channel combustor on the X and Y directions. The wall temperature difference on the Y direction is 86.6 K for the multiplechannel combustor given that the flow rate Vin is 100 cm3/s. The
difference is lower than that of the single-channel combustor with
a difference of 107.6 K. Figs. 6 and 7 also show that a higher and
more uniform wall temperature can be obtained with the increase
of the flow rate because more fuel is burned per unit time, which
results in a larger heat release rate.

1300
1250

For a micro-TPV system, the increasing of outer wall temperature is the most efficient way to enhance the power output and
energy conversion efficiency. The radiation energy from the outer
wall of the micro combustor can be calculated by Eq. (7). Moreover,
the radiation efficiency is defined as the ratio of the radiation
energy to the input chemical energy from the fuel, as shown in
Eq. (8) [35].

Multiple-channel,Vin = 85 cm3/s

1200

Single-channel,Vin = 85 cm3/s

1150

Single-channel,Vin = 100 cm 3/s

Multiple-channel,Vin = 100 cm3/s


Multiple-channel,Vin = 115 cm3/s

1100

Single-channel,Vin = 115 cm3/s


Multiple-channel,Vin =130 cm3/s

1050

P er

Single-channel,Vin = 130 cm /s

XN

T4 A
i1 w;i i

1000
-5

-4

-3

-2

-1

gre

Y (mm)
Fig. 7. Comparison of the outer wall temperature at different flow rates: (a) X
direction (Y = 0 mm), and (b) Y direction (X = 11 mm).

which further enhances the heat transfer between the hot gas and
the inner wall of the combustor. Because the forced convection heat
transfer is proportional to flow velocity. The setting up of multiple
channels will increase the effective convection heat transfer area,
the inner area of the multiple-channel combustor is approximately
1.23 times as much as that of the single-channel combustor. Thus,

24
Radiation energy of Single-channel
Radiation energy of Multiple-channel

60

Radiation efficiencyof Single-channel


Radiation efficiency of Multiple-channel

20

56

18
52
16
14

48

Radiation energy,(W)

Radiation efficiency, (%)

22

P
mH2 Hc

PN

er

4
i1 T w;i Ai

mH2 Hc

where e is the emissivity of the outer surface area of the combustor,


r is the StefanBoltzmann constant (=5.67  108 W/(m2 K4)), Ai is
the area of each element surface of the combustor, T w;i is the outer
wall temperature of each element surface, mH2 is the mass flow rate
of hydrogen, and Hc is higher calorific value of hydrogen. Fig. 8
shows the total radiation energy and radiation efficiency of the
two combustors at different flow rates. Radiation energy is proportional to the four power of the outer wall temperature. Thus, the
radiation energy of the two combustors shares an increasing trend
with the increase of the flow rate. The radiation energy of the
multiple-channel combustor is higher than that of the singlechannel combustor for all testes cases. Nevertheless, Fig. 8 shows
that the radiation efficiencies of the two combustors decreased with
the increase of the flow rate because of the decrease of the effective
residence time of the fuel as flow rate increases, which leads to
incomplete combustion. However, the radiation efficiency of the

12
44
10

85

100

115

130

Flow rate, Vin ( cm3/s)


Fig. 8. Comparison of the radiation energy and radiation efficiency at different flow
rates.

Fig. 9. Schematic of the heat loss network of the micro combustor (the arrows
indicate the possible direction of heat flux).

203

Y. Su et al. / Energy Conversion and Management 120 (2016) 197205

Fig. 10. Comparison of the temperature distributions on the outer wall with different equivalence ratio assignments: (a) Case 1, (b) Case 2, and (c) Case 3.

Table 5
Equivalence ratio assignments for different cases.
Total H2 mass flow rate (106 kg/s)

Case

H2/air equivalence ratio (U)


Channel #1

Channel #2

Channel #3

Channel #4

Channel #5

1
2
3

0.80
0.85
0.85

0.80
0.7675
0.80

0.80
0.7675
0.7045

0.80
0.7675
0.80

0.80
0.85
0.85

multiple-channel combustor remains higher than that of the singlechannel combustor. For example, when flow rate Vin is 100 cm3/s,
the radiation efficiency of the multiple-channel combustor
increases from 16.67% to 18.84%, with an increment of 2.2%. The relative radiation efficiency increase can reach 2.7% when flow rate Vin
is 130 cm3/s in the tested cases.
3.3. Effect of assigning different H2/air equivalence ratios for every
channel
The wall temperature difference on the cross-sectional (Y)
direction of the multiple-channel combustor is lower than that of
the single-channel combustor. However, this temperature difference is still detrimental to the micro-TPV system. This section
investigates the assignment of different equivalence ratios for
every channel to obtain a more uniform temperature.
Fig. 9 depicts the schematic of the heat loss network of the combustor in the YZ plane. The side effect, that is, the heat loss area in
the unit combustion volume of the side zone is larger than that of
the center zone. Thus, the calculated results show that the area/
volume ratio for Channel #1 is approximately 1.59 times that of
Channel #2, which leads to a larger heat loss of the side zone than
that of the center zone. Therefore, the wall temperature on the central zone is higher than that of the side zone, as shown in Fig. 7(b).
Given the significant effect of the equivalence ratio on combustion
temperature, an approach with different equivalence ratio assignments for every channel of the multiple-channel combustor is
applied to obtain a more uniform temperature, that is, a relatively
larger equivalence ratio for the side channels.
Fig. 10(a)(c) displays the outer wall temperature distributions
of the multiple-channel combustor with different equivalence ratio
assignments, as shown in Table 5, when the flow rate (Vin) is fixed

2.034
2.034
2.034

at 100 cm3/s and total H2 mass flow rates is 2.034  106 kg/s.
Fig. 10 shows that the shapes of the high-temperature zone for
all tested cases are almost the same. However, Fig. 10(a) shows
that the central region is relatively brighter, which indicates a high
wall temperature. This phenomenon is attributed to the increased
amount of thermal energy released in the central channel with
high equivalence ratio. Fig. 10(b) and (c) shows a more uniform
temperature distribution on the Y direction. The maximum temperatures in the center zone are slightly lower than that shown
in Fig. 10(a). Therefore, assigning different equivalence ratios for
every channel is an effective way to improve temperature uniformity further, particularly along the Y direction.
Fig. 11 shows the outer wall temperature distributions on the Y
direction, with X = 5, 11, and 17 mm of the multiple-channel combustor for different tested cases. Fig. 11 shows that the center wall
temperature of Case 1 is higher than that of the other cases. However, the side wall temperature of Case 1 is the smallest, which is
mainly caused by the H2/air equivalence ratio assignments. The
wall temperature profiles of Cases 2 and 3 for different X locations
are relatively smoother than that of Case 1, particularly in the central zone, given the different equivalence ratio assignments. Moreover, the wall temperature of the two sides of Cases 2 and 3 is
slightly higher than that of Case 1 because of the higher equivalence ratio assignments. Thus, the wall temperature difference on
the Y direction of Cases 2 and 3 is smaller than that of Case 1.
Table 6 shows that the temperature differences are 86.6, 72.8,
and 68.8 K for Cases 1, 2, and 3, respectively, where X is equal to
11 mm. The mean temperature of the outer wall, radiation energy,
and radiation efficiency of Cases 2 and 3 are slightly lower than
those of Case 1. Therefore, assigning different equivalence ratios
for every channel can further improve the outer wall temperature
uniformity and obtain relatively high radiation efficiency.

Outer wall temperature,Tw (K)

Outer wall temperature,Tw (K) Outer wall temperature,Tw (K)

204

Y. Su et al. / Energy Conversion and Management 120 (2016) 197205

(a)

Case 1
Case 2
Case 3

(b)

Case 1
Case 2
Case 3

(c)

Case 1
Case 2
Case 3

1350

1320

1290

1260

1350

1320

1290

1260

1290

system. When the flow rate is 100 cm3/s and H2/air equivalence ratio is 0.8, the mean temperature is 1294.9 K for the
multiple-channel combustor and 1256.0 K for the singlechannel combustor.
(2) The multiple-channel micro combustor can achieve a preferable temperature uniformity performance on the axial (X)
direction and on the cross-sectional (Y) direction compared
with the single-channel combustor. The wall temperature
difference on the Y direction is only 86.6 K for the
multiple-channel combustor, with a reduction of 21 K compared with the single-channel combustor with the same
flow rate.
(3) The radiation energy and radiation efficiency of the
multiple-channel micro combustor are higher than those of
the single-channel micro combustor. Numerical results
show that when flow rate Vin is 100 cm3/s, the radiation efficiency of the multiple-channel combustor reaches 18.84%
and that of the single-channel combustor reaches 16.67%.
(4) The approach that assigns a higher equivalence ratio for the
side channels and a lower equivalence ratio for the center
channels can further improve the outer wall temperature
uniformity. The wall temperature difference on the Y direction can decrease from 86.6 K to 68.8 K, and the relative high
radiation efficiency can be obtained.

Acknowledgments
This study was supported by the Foundation of State Key
Laboratory of Coal Combustion (FSKLCCB1601), the Key
fundamental Research Project from Shenzhen Research Council
(JCYJ20140819154343380 and JCYJ20150630155150193) and the
National Basic Research Program of China (No. 2015CB251505).

1260

1230

1200

-5

-4

-3

-2

-1

References

Y (mm)
Fig. 11. Comparison of the outer wall temperature with different equivalence ratio
assignments: (a) X = 5 mm, (b) X = 11 mm, and (c) X = 17 mm.

Table 6
Performances of the combustor with different equivalence ratio assignments.
Case

Temperature
difference (K),
X = 11 mm

Mean
temperature of
outer wall (K)

Radiation
energy
(W)

Radiation
efficiency
(%)

1
2
3

86.6
72.8
68.8

1294.9
1290.9
1293.3

53.57
52.91
53.30

18.84
18.60
18.74

4. Conclusions
A multiple-channel micro combustor for the micro-TPV system
was proposed. Numerical simulations with detailed chemistry
were conducted to compare the performances of the singlechannel and multiple-channel micro combustors. The assignment
of different equivalence ratio for every channel of the multiplechannel micro combustor is also investigated. Several conclusions
can be drawn, as follows:
(1) The temperature distribution along the outer wall of the
multiple-channel micro combustor is more uniform and
higher than that of the traditional single-channel micro
combustor, which is highly beneficial for the micro-TPV

[1] Chou SK, Yang WM, Chua KJ, Li J, Zhang KL. Development of micro power
generators a review. Appl Energy 2011;88:116.
[2] Cao HL, Xu JL. Thermal performance of a micro-combustor for micro-gas
turbine system. Energy Convers Manage 2007;48:156978.
[3] Jiang LQ, Zhao DQ, Guo CM, Wang XH. Experimental study of a plat-flame
micro combustor burning DME for thermoelectric power generation. Energy
Convers Manage 2011;52:596602.
[4] Dahm W, Mijit J, Mayor R, Qiao G, Benajmin A, Gu Y, et al. Micro internal
combustion swing engine (MICSE) for portable power generation systems;
2002.
[5] Yang WM, Chua KJ, Pan JF, Jiang DY, An H. Development of microthermophotovoltaic power generator with heat recuperation. Energy Convers
Manage 2014;78:817.
[6] Yang WM, Chou SK, Shu C, Li ZW, Xue H. A prototype microthermophotovoltaic
power generator. Appl Phys Lett 2004;84:3864.
[7] Su Y, Song J, Chai J, Cheng Q, Luo Z, Lou C, et al. Numerical investigation of a
novel micro combustor with double-cavity for micro-thermophotovoltaic
system. Energy Convers Manage 2015;106:17380.
[8] Wenming Y, Dongyue J, Kenny CKY, Dan Z, Jianfeng P. Combustion process and
entropy generation in a novel microcombustor with a block insert. Chem Eng J
2015;274:2317.
[9] Li J, Chou SK, Li ZW, Yang WM. A potential heat source for the microthermophotovoltaic (TPV) system. Chem Eng Sci 2009;64:32829.
[10] Sahota GPS, Khandelwal B, Kumar S. Experimental investigations on a new
active swirl based microcombustor for an integrated micro-reformer system.
Energy Convers Manage 2011;52:320613.
[11] Norton DG, Vlachos DG. Combustion characteristics and flame stability at the
microscale: a CFD study of premixed methane/air mixtures. Chem Eng Sci
2003;58:487182.
[12] Bagheri G, Hosseini SE, Wahid MA. Effects of bluff body shape on the flame
stability in premixed micro-combustion of hydrogenair mixture. Appl Therm
Eng 2014;67:26672.
[13] Norton DG, Vlachos DG. A CFD study of propane/air microflame stability.
Combust Flame 2004;138:97107.
[14] Fan A, Minaev S, Sereshchenko E, Fursenko R, Kumar S, Liu W, et al.
Experimental and numerical investigations of flame pattern formations in a
radial microchannel. Proc Combust Inst 2009;32:305966.

Y. Su et al. / Energy Conversion and Management 120 (2016) 197205


[15] Fan A, Minaev S, Sereshchenko E, Tsuboi Y, Oshibe H, Nakamura H, et al.
Propagation dynamics of splitting flames in a heated microchannel. Combust
Explosion Shock Waves 2009;45:24550.
[16] Fan A, Minaev S, Kumar S, Liu W, Maruta K. Experimental study on flame
pattern formation and combustion completeness in a radial microchannel. J
Micromech Microeng 2007;17:2398406.
[17] Yang W, Chou S, Shu C, Li Z, Xue H. Combustion in micro-cylindrical
combustors with and without a backward facing step. Appl Therm Eng
2002;22:177787.
[18] Lee KH, Kwon OC. Studies on a heat-recirculating microemitter for a micro
thermophotovoltaic system. Combust Flame 2008;153:16172.
[19] Yan Y, Tang W, Zhang L, Pan W, Yang Z, Chen Y, et al. Numerical simulation of
the effect of hydrogen addition fraction on catalytic micro-combustion
characteristics of methaneair. Int J Hydrogen Energy 2014;39:186473.
[20] Chou SK, Yang WM, Li J, Li ZW. Porous media combustion for micro
thermophotovoltaic system applications. Appl Energy 2010;87:28627.
[21] Li J, Chou SK, Li ZW, Yang WM. Experimental investigation of porous media
combustion in a planar micro-combustor. Fuel 2010;89:70815.
[22] Fan A, Wan J, Liu Y, Pi B, Yao H, Maruta K, et al. The effect of the blockage ratio
on the blow-off limit of a hydrogen/air flame in a planar micro-combustor
with a bluff body. Int J Hydrogen Energy 2013;38:1143845.
[23] Wan J, Yang W, Fan A, Liu Y, Yao H, Liu W, et al. A numerical investigation on
combustion characteristics of H2/air mixture in a micro-combustor with wall
cavities. Int J Hydrogen Energy 2014;39:813846.
[24] Yang WM, Chou SK, Pan JF, Li J, Zhao X. Comparison of cylindrical and modular
micro combustor radiators for micro-TPV system application. J Micromech
Microeng 2010;20:085003.

205

[25] Peckner D, Bernstein IM. Handbook of stainless steels. New York: McGraw-Hill
Book Co.; 1977 (Chapters paged separately).
[26] Ali Beskok GEK. Report: a model for flows in channels, pipes, and ducts at
micro and nano scales. Microscale Thermophys Eng 1999;3:4377.
[27] Kuo CH, Ronney PD. Numerical modeling of non-adiabatic heat-recirculating
combustors. Proc Combust Inst 2007;31:327784.
[28] Hosseini SE, Wahid MA. Investigation of bluff-body micro-flameless
combustion. Energy Convers Manage 2014;88:1208.
[29] Giovangigli V, Smooke MD. Extinction of strained premixed laminar flames
with complex chemistry. Combust Sci Technol 1987;53:2349.
[30] Jiang D, Yang W, Chua KJ, Ouyang J, Teng J. Effects of H2/CO blend ratio on
radiated power of micro combustor/emitter. Appl Therm Eng 2015;86:17886.
[31] Yang W, Jiang D, Chua K, Chua Y. Research on combustion process in a novel microcombustor with block inserts. In: Paper presented at the 15th international heat
transfer conference, IHTC15-8615. Kyoto; August 1015, 2014.
[32] Li J, Chou SK, Yang WM, Li ZW. Experimental and numerical study of the wall
temperature of cylindrical micro combustors. J Micromech Microeng
2009;19:015019.
[33] Tang A, Pan J, Yang W, Xu Y, Hou Z. Numerical study of premixed hydrogen/air
combustion in a micro planar combustor with parallel separating plates. Int J
Hydrogen Energy 2015;40:2396403.
[34] Wan J, Fan A, Yao H, Liu W. Effect of thermal conductivity of solid wall on
combustion efficiency of a micro-combustor with cavities. Energy Convers
Manage 2015;96:60512.
[35] Bagheri G, Hosseini SE. Impacts of inner/outer reactor heat recirculation on the
characteristic of micro-scale combustion system. Energy Convers Manage
2015;105:4553.

Você também pode gostar