Está en la página 1de 14

Energy 93 (2015) 188e201

Contents lists available at ScienceDirect

Energy
journal homepage: www.elsevier.com/locate/energy

A mathematical tool for predicting thermal performance of natural


draft biomass cookstoves and identification of a new operational
parameter
Milind P. Kshirsagar a, *, Vilas R. Kalamkar b
a
St Vincent Pallotti College of Engineering and Technology, Nagpur, India
b
Visveswaraya National Institute of Technology, Nagpur, India

a r t i c l e i n f o a b s t r a c t

Article history: The goal of the work was to develop a simple yet reliable and computationally inexpensive mathematical
Received 19 March 2015 tool for the performance prediction of ‘rocket’ type natural draft direct combustion stoves with
Received in revised form unshielded pot. The work included development of a novel heat and mass transfer model for natural
21 August 2015
draft biomass cookstoves and its integration with Excel® spreadsheet to develop a user-friendly math-
Accepted 5 September 2015
Available online xxx
ematical tool. The results from the mathematical tool were validated against the experimental data set
from literature. A new operational parameter named ‘Inlet area ratio’ was identified, and its effects on the
stove performance were investigated. It was concluded that this newly identified parameter has a
Keywords:
Biomass cookstove
considerable effect on the performance of a natural draft stove, and the findings are likely to change the
Heat transfer future cookstove modeling and design approach.
Efficiency © 2015 Elsevier Ltd. All rights reserved.
Inlet area ratio
Performance prediction

1. Introduction easy to use mathematical tool for the performance prediction of a


biomass cookstove is still missing.
As a result of about 160 cookstove programs running across the
world, a wide variety of improved cookstove designs with varying
2. Need of a performance prediction tool for a biomass
performance exists [1e4]. The majority of these stoves are of direct
cookstove
combustion type with shielded combustion and unshielded pot.
These stoves range from a simple traditional clay combustion
Every biomass stove maker/designer/artisan willing to construct
chamber to a factory made metallic or ceramic combustion cham-
a biomass cookstove, fulfilling specific local needs, requires a user-
ber stoves. The most famous and widely used among these is the
friendly mathematical tool for predicting the thermal performance
design by Larry Winiarski: the popular ‘rocket stove’. The ‘rocket
of the proposed cookstove, as a prerequisite to construction and
stove’ design, which is under research and use for last 30 years,
testing [1]. In the absence of any such tool, one may choose a stove
consists of a chimney shaped elbow as a combustion chamber.
design and its different parameters; without having a faintest idea
About a half million, ‘rocket stoves’ are still in use in households
of how these parameters are affecting the stove performance. The
across the world [5]. In several studies, ‘rocket stoves’ have proven
research work presented in this paper, aimed at providing this
themselves as the best in the category; in some cases, even better
much-needed ‘virtual lab’ for the thermal performance prediction
than the Gasifier stoves. Examples of “rocket” type stoves are
of a biomass cookstove, through its mathematical tool.
‘Envirofit’, ‘StoveTec’, and ‘WFP rocket’. The empirical studies con-
ducted on ‘rocket stoves’ have resulted in broad design guidelines
[6], but, with no clue on the possible performance of the design. An 3. Literature review of modeling of biomass cookstoves

Most of the phenomenons occurring inside the cookstoves were


* Corresponding author. St. Vincent Pallotti College of Engineering and
independently studied and modeled. These phenomenons include
Technology, Wardha Road, Nagpur, Maharashtra, India. Tel.: þ91 8600565759. drying, pyrolysis, combustion, turbulence; conductive, convective
E-mail address: milindpkshirsagar@rediffmail.com (M.P. Kshirsagar). and radiative heat transfer. The early efforts of stove modeling

http://dx.doi.org/10.1016/j.energy.2015.09.015
0360-5442/© 2015 Elsevier Ltd. All rights reserved.
M.P. Kshirsagar, V.R. Kalamkar / Energy 93 (2015) 188e201 189

Nomenclature ma kg of air supplied for combustion per kg of dry fuel


burnt
A cross sectional area of the chimney/combustion m_ act actual mass flow rate of flue gas (kg/s)
chamber (m2) m_ f fuel burn rate (kg of dry fuel/s)
Aaverage pot gap average pot gap flow area (m2) mflue kg of flue produced per kg of dry fuel burnt
Ai inner wall surface area (m2) ms stoichiometric air required per kg of dry fuel
Ain area unoccupied by fuel at the feed door (m2) m_ th theoretical mass flow rate of flue gas (kg/s)
Am entire bounding surface area of volume Vm (m2) NCV net calorific value of fuel (kJ/kg of dry fuel)
Ao outer wall surface area (m2) Nubottom average Nusselt no for impinging jet heat transfer for
Aoutlet area of flow at the outlet of stove (m2) entire pot bottom
Apg area of pot above pot gap (m2) % O2 percentage oxygen in flue gas
Aside area of pot side up to water height (m2) O oxygen content of the fuel as percentage of dry weight
C carbon content of the fuel as percentage of dry weight (%)
(%) P firepower of the stove (kW)
Cd coefficient of discharge Pr Prandtl number for air/flue gas
Cheat coefficient of distributed heat addition Qchar radiation radiative heat transfer from char bed to the pot
Cp specific heat of air/flue gas (kJ/kg K) bottom (kJ/kg of dry fuel)
Cviscous coefficient of viscous loss Q_char radiationradiative heat transfer from char bed to the pot
D diameter of the chimney (m) bottom (W)
Dh hydraulic diameter of the entrance area available (m)
Q_ conv 2 convective heat transfer to the pot from pot bottom
Dhg hydraulic diameter of pot gap (m)
zone (W)
Di inner diameter of the stove (m)
Do outer diameter of the stove (m) Q_ conv 3 convective heat transfer to the pot from pot gap zone
em error in mass flow rate predictions (W)
eO2 error in % O2 predictions Q_ conv 4 convective heat transfer to the pot from pot side zone
eT error in temperature predictions (W)
Fchardoor view factor between char bed and feed door Qdoor loss radiative heat lost through the feed door opening (kJ/
Fcharpot view factor between char bed and pot bottom kg of dry fuel)
Fcharwall view factor between char bed and inner wall Qheat loss heat lost from the stove through insulation (kJ/kg of
fe friction factor for length of elbow dry fuel)
GrH Grashoff number for outer surface QH2 moisture heat loss due to sensible heating of H2 related
H height of the chimney (m) moisture (kJ/kg of dry fuel)
H2 hydrogen content of the fuel as percentage of dry Q_flameinne wallheat transfer from flame to inner wall (W)
weight (%) Qflame radiation radiative heat transfer from flame to the pot
hbottom impinging jet heat transfer coefficient for pot bottom bottom (kJ/kg of dry fuel)
zone (W/m2 K) Q_ flame radiation radiative heat transfer from flame to the pot
hci convective heat transfer coefficient for inner wall (W/
bottom (W)
m2 K)
Qfuel moisture heat loss due to evaporation and sensible heating of
hco convective heat transfer coefficient for outer wall (W/
fuel moisture (kJ/kg of d fuel)
m2 K)
Q_heat loss heat lost from the combustion chamber through
hrch char radiative heat transfer coefficient for inner wall
insulation (W)
(W/m2 K)
hrfl flame radiative heat transfer coefficient for inner wall Q_ inner wallouter wall heat transfer from inner wall to outer wall
(W/m2 K) (W)
hro radiative heat transfer coefficient for outer wall (W/ Qmoisture loss heat lost due to evaporation of fuel moisture (kJ/kg
m2 K) of dry fuel)
hs enthalpy of superheated steam at atmospheric Q_outer wallsurroundungs heat transfer from outer wall to the
pressure (kJ/kg of steam) surroundings (W)
hside convective heat transfer coefficient for pot side (W/ Q_ rad 3 radiative heat transfer to the pot from pot gap zone
m2 K) (W)
hp height of pot (m) Q_ rad 4 radiative heat transfer to the pot from pot side zone
hpot gap impinging jet heat transfer coefficient for pot gap zone (W)
(W/m2 K)
Q_ side loss rate of heat loss from pot side to the surroundings (W)
hw height of water in the pot (m)
RaH Rayleigh number for outer surface
k thermal conductivity of air/flue gas (W/m K)
ReD Reynolds no for combustion zone
kbottom thermal conductivity of flue in pot bottom zone (W/
Ree Reynolds no inside elbow length
m K)
Rehw Reynolds no over pot side
ki thermal conductivity of stove insulation (W/m K)
S sulfur content of the fuel as percentage of dry weight
kpot gap thermal conductivity of flue in pot gap zone (W/m K)
(%)
kside thermal conductivity of flue in pot side zone (W/m K)
Tc temperature of flue at the corner of the pot (K)
L elbow length (m)
Tchar char temperature (K)
M moisture content of the fuel on weight basis (%)
Ta ambient temperature (K)
190 M.P. Kshirsagar, V.R. Kalamkar / Energy 93 (2015) 188e201

Te temperature of flue at the entrance to the pot gap (K) a4 absorptivity of flue gas in pot side zone
Tg flame temperature/flue temperature inside chimney εc char emissivity
(K) εg emissivity of flue gas in combustion zone
To temperature of flue at exit (K) ε0 emissivity of outer wall surface
Tp pot surface temperature (K) ε3 emissivity of flue gas in pot gap zone
Ts temperature of surface in the zone (K) ε4 emissivity of flue gas in pot side zone
Twi inner wall surface temperature (K) l excess air factor
Two outer wall surface temperature (K) hc combustion efficiency (%)
V actual velocity of flue gases inside the chimney (m/s) ho overall thermal efficiency of the stove (%)
Vgap velocity of flue in pot gap (m/s) DP chimney draught (N/m2)
Vm volume of the participating medium (m3) m dynamic viscosity of air/flue gas (kg/m s)
w width of fuel sticks (m) n kinematic viscosity of air/flue gas (m2/s)
Wi inner pot gap width (m) ra density of atmospheric air (kg/m3)
Wo outer pot gap width (m) rg density of flue gas (kg/m3)
s Stefan Boltzmann constant (W/m2 K4)
Greek symbols q angle made by radius with the fuel upper surface
ag absorptivity of flue gas in combustion zone (radian)
a3 absorptivity of flue gas in pot gap zone

started in 1980s; where, Prasad, Visser, Bussmann and Verhaart 5. Why a spreadsheet?
developed preliminary heat transfer models for open as well as
shielded fires [7e9]. Baldwin [10] in 1987 provided rough guide- The aim of the work was to develop and provide a user-friendly
lines for design of biomass cookstoves, discussing different pro- performance prediction tool to the stove makers/artisans. Any such
cesses involved and investigating wall losses and heat transfer tool requires solving the heat transfer model with a competitive
correlations. Sharma in 1992 presented the basic design principles mathematical software/spreadsheet. Nowadays, programs like
for a cookstove; like combustion, fluid flow and heat transfer [11]. ANSYS® and MAT LAB®, which are computationally and economi-
After negligence by the technical community during 1990s, cally expensive, are increasingly used for the mathematical equa-
biomass cookstoves have captured the attention of many re- tion solving. In spite of all their advantages, these programs require
searchers in last 10 years, wherein the researchers conducted heat- a properly trained operator, and hence their use is limited to the
transfer studies and identified a number of important performance engineering community only. On the contrary, the other softwares
variables [12e19]. MacCarty developed a steady-state computa- suitable for the present work, with less demand of ‘expertise’, and
tional zonal model of a wood-burning natural draft, single pot, ‘system requirements’ are Engineering Equation Solver, WPS Office,
shielded fire, to predict the fluid flow and heat transfer behavior LibreOffice and Microsoft Excel®. Out of these, Excel® is the most
inside the stove [20]. Kausley and Pandit experimentally validated familiar spreadsheet, is practically available in most of the personal
theoretical model of solid-fuel combustion in a domestic ‘Harsha’ computers and is capable of handling complex mathematical cal-
stove [21]. Many researchers have used the modern computational culations involved here. Hence, Excel® was chosen for the purpose.
packages such as CFD (Computational Fluid Dynamics) for the
modeling, simulation and optimization of biomass stoves [22e31]. 6. Scope of the work
Agenbroad et al., who developed and experimentally validated
simplified model of a chimney stove for predicting flue flow rate, The goal was to develop a simple yet reliable and computa-
flame temperature, and excess air ratio; did the finest work so far in tionally inexpensive performance prediction tool for the direct
the cookstove modeling and its experimental validation [32,33]. combustion natural draft stoves with shielded combustion and
unshielded pot. A new heat and mass transfer model was devel-
oped and integrated with Excel® spreadsheet to develop a user-
4. Need of a new heat and mass transfer model friendly mathematical tool for the performance prediction.
Though, inherently simple in nature, the mathematical tool allows
The development of a trustworthy mathematical tool for the for variation in 20 input variables (of geometry, material and
performance prediction of biomass cookstove required a quanti- operational type) and to observe their effect on 31 stove output
tatively reliable heat and mass transfer model for the purpose. The parameters.
models discussed in Section 4 though very essential, were not The multiple processes involved in all the zones were given
ready to use for the purpose. Many of these studies were of pre- due consideration, and coupled through the standard guiding
liminary type and do not involve a wholesome set of mathematical principles of combustion, heat transfer, and fluid flow. Heat
equations required for in-depth analysis of a complex energy transfer and the losses in all the zones were thoroughly modeled
system like modern biomass cookstoves [7e18]. Many of them using the principle of energy conservation and detailed thermal
were not validated experimentally. Those, which were validated, resistance calculation for all three modes of heat transfer. The
included validation of only one or two variables with a high range convective heat transfer to/from the stove walls, the pot bottom
of deviation of predicted values from the experimental one and the sides were expressed using heat transfer correlations and
[20e28]. Even the model by Agenbroad, was not quantitatively related dimensionless numbers. The radiation heat transfer from
reliable when it comes to the flame temperature and mass flow the flame was modeled by considering flame as participating
rate predictions [32,33]. Hence, a new heat and mass transfer media between two solid surfaces. For predicting flue flow rate,
model for the performance prediction of a biomass cookstove was the conservation of mass principle was applied primarily to the
developed. combustion zone, and then, was combined with the chimney
M.P. Kshirsagar, V.R. Kalamkar / Energy 93 (2015) 188e201 191

Pot side zone (flames scraping against the pot sides). The
important geometrical parameters for the given stove geometry
are listed in Table 1.

7.2. Model assumptions

A cookstove broadly involves three phenomenon i. e. combus-


tion, heat transfer and fluid flow. These phenomenons include sub-
phenomenons like drying, pyrolysis, char combustion, volatile
combustion, turbulence; pressure drops across different sections,
conductive, convective and radiative heat transfers and impinging
jet heat transfer. To include all these fine details was not required
for the performance prediction of biomass cookstoves. Hence, some
simplifying assumptions were made, without neglecting any of the
important aspects of stove operation:

 A cookstove operates under steady-state condition.


 Flow within the stove is one-dimensional and axisymmetric.
 Fuel bed covers the entire surface area of the grate.
 The emissivity and the surface temperature of burning char are
Fig. 1. The model geometry. 0.85 and 1100 K respectively [26,37,38].
 Emissivity of the stove inner surface and pot outer surface is 1
(treating them as black bodies).
flow expression. The pressure drop calculations across various
 The conductive resistance of the cooking pot is negligible (High
geometrical variations were also given a rigorous treatment. For
conductivity Al pots).
the sake of simplicity, the combustion aspects of the process
 The average temperature of the cooking pot is 373 K (Boiling
were neglected. No model for drying, pyrolysis, char combustion,
temperature of water).
heat release or pollutant formation was included. An operational
 The flue gas follows ideal gas behaviors. The thermodynamic
variable with the name ‘combustion efficiency’ was introduced to
properties of the flue gas are same as that of the air, and were
take care of the combustion quality.
evaluated at the atmospheric pressure. The following correla-
tions were established for different air properties in the tem-
perature range of 300e1500 K at 1 atm [39]:
7. The model

7.1. Model geometry


m ¼ 0:0447105 *T 0:7775 (1)
A modern-day biomass stove essentially consists of a com-
bustion chamber for enclosing the fire, insulation for minimizing k ¼ 0:00031847T 0:7775 (2)
heat losses/surface temperature, air supply system for directing
the flow of air, the heat transfer zone for facilitating convective Cp ¼ 0:9362 þ 0:0002T (3)
and radiative heat transfers to the cooking pot, and the cooking
pot holding the food. The geometry as shown in Fig. 1 is a  
representative of the ‘rocket stove’ design and consists of a cy- n ¼ 0:0000644T 2 þ 0:0631*T  9:54 106 (4)
lindrical combustion chamber (an insulated chimney for
shielding fire) and a single unshielded pot. It consists of four r ¼ 353=T (5)
zones: 1) Combustion zone (air supply port/s, fuel bed, fuel
feeding door, and fire shield or chimney) 2) Pot bottom zone
Pr ¼ 0:685ðconstantÞ (6)
(impinging flue jet striking against the pot bottom) 3) Pot gap
zone (gap between the pot bottom and the stove body), and 4) Values for enthalpy of air were correlated from Cengel [40]:

Table 1
Geometrical input parameters.

S. N. Parameter Value Unit Remark Ref.

1. Chimney diameter 100 mm e [32,33]


2. Height of the chimney 210 mm e [34]
3. Stove outer diameter 125 mm Basic Fiberfrax insulation of 1/2 inch thickness [35]
4. Inlet area ratio 0.7e0.9 e Calculated [36]
5. Area factor for elbow length 1.67 e Calculated [36]
6. Feed area/Chimney area factor 1 e e [32,33]
7. Inner pot gap Width 18 mm e [33]
8. Outer pot gap width 18 mm e [33]
9. Pot diameter 225 mm e [33]
10. Pot height 162 mm For standard 5 L cooking pot [3]
11. Height of water in pot 126 mm Calculated for 5 L of water [33]
12. Insulation thermal conductivity e W/mK k ¼ 3.125*107*T20.000118125*T þ 0.0464578 [35]
Correlated from Fiberfrax product specification sheet
192 M.P. Kshirsagar, V.R. Kalamkar / Energy 93 (2015) 188e201

h ¼ 9:314105 *T 2 þ 0:94576T þ 7:0445 (7) _ act


m
Cd ¼ ¼ Cviscous Cheat (13)
m_ th

7.3. Mathematical equations for different zones where, the discharge coefficient Cd accounts for losses due to
viscous effects in the stove and distributed heat addition from the
7.3.1. Combustion zone combustion process. Value of Cd is always less than one, and usually
The main processes considered in the combustion chamber ranges between 0.35 and 0.42 for different type of chimneys
were combustion of fuel (volatiles and finally that of char), radiative [32,33,42,43]. Agenbroad et al. [34] derived a value of coefficient of
heat transfer to the pot from char and the flame, and heat loss from heat addition as Cheat ¼ 0.707, assuming that the density of flue gas
the stove walls and the door opening to the surroundings. In varies linearly with the height (due to gradual heat addition with
naturally aspirated wood-burning stoves, chimney effect due to rising flame), and half of the draught is lost due to it. However, from
buoyancy force drives the airflow against the different flow re- literature review it was observed, that the flame temperature and
sistances. This airflow, in turn, determines the wood burning rate as hence the flue density was achieved just above the fuel bed [27,29].
well as the overall heat transfer and combustion performance. Allowing a tolerance of 20% of the total height for the temperature
Using the methodology adopted by previous stove and solar and density achievement the resulting draught loss is 10%, which
chimney researchers, the chimney effect (for the entire heat added approximated to a value of Cheat ¼ 0.95.
at the base of the chimney and constant flue gas temperature and Applying pressure balance to the chimney:
density) [19,33,41e46] is:
DPTotal ¼ DPKinetic energy þ DPLoss (14)
  1 2
DPchimney ¼ H* ra  rg g ¼ rg Vth ½34 (8)
2
X

Density of flue gases inside the chimney: 1 2 1 1 rg A 2


*rg *Vth ¼ *rg *V 2 þ *rg *V 2 Kl * * (15)
2 2 2 rl Al
353 l
rg ¼ (9)
Tg
Rearranging the Eqs. (8) and (9) the theoretical velocity of flue V 1
Cviscous ¼ ¼ sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

2 (16)
gases inside a chimney: Vth P
1 þ Kl *rr * AA
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
  l
l l

Tg
Vth ¼ 2g  H  1 (10)
Ta Expressions for different loss coefficients, density ratios, and
area ratios are given in Table 2.
From continuity equation theoretical mass flow rate of flue Applying mass balance:
gases:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi _f
_ act ¼ mflue m
ffi
m (17)

353 Tg
_ th
m ¼ A  2g  h  1 (11)  
Tg Ta M
mflue ¼ ma þ 1 þ (18)
Accounting for pressure drop due to different reasons across the 100  M
stove, the actual mass flow rate:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi P
353 Tg _f ¼
m (19)
_ act
m ¼ Cd * A* 2g  h  1 (12) NCV
Tg Ta
Rearranging Eqs. (12) and (17), we get:

Table 2
Loss coefficients, density ratios and area ratios for pressure drop calculations.

S. N. Type of loss Kl r/rl A/Al Ref.

1 Sudden contraction at inlet ¼0.5 Ta/Tg 1/Ar [20]


Tg þTa
2 Friction loss in elbow length ¼ feD*Lh e ; fe ¼ R64
ee
4*Ar *A
; Dh ¼ d*½pqþSinq 2*Tg 1/Ar

q ¼ ð2:289503842*A2r þ 1:23599729*Ar þ 1:522230692Þ [48]


 
Correlated from the results obtained from expression: Ar *A ¼ D2 p q Sin2q
4 2

3 Expansion to chimney ¼[1/(Ar0.5)1]2 1 1 [19]


4 Fuel bed resistance (Viscous and Inertial) ¼ 3:5*H *ð1εÞ*rrb þ 300*H*m *ð1εÞ
2
1 1 [21,29]
d p ε3 d2 *r *V
g p g ε3
ε¼bed voidage ¼ 0.69
H ¼ D/4
5 Friction loss in chimney ¼f*h/D, 1 1 [16]
f ¼ (0.79*lnReD1.64)2,
for ReD > 900
f ¼ 64/ReD, for ReD < 900
6 Loss due to bend at pot bottom 1 1 1 [20]
7 Loss due to friction in the pot gap zone 4f ðdp dÞ Tc þTe 2
[48]
¼ Dh ; f ¼ 24
Re ; Dh ¼ W1 þ W2 2*Tg
d
ðdþdp Þ*ðW1 þW2 Þ
8 Loss at the outlet 1 1 1 [20]
M.P. Kshirsagar, V.R. Kalamkar / Energy 93 (2015) 188e201 193

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 
353 1 Tg B ¼ 0:9589 þ 4:8 *106 *T
mflue ¼ Cd * * *A* 2*g*h* 1 (20)
Tg m_ f Ta Absorptivity of flue gas was taken from Bejan and Kraus [48]:
Now applying heat balance to the combustion chamber:  0:5
T
ag ¼ εg * (29)
Qsupplied by the fuel combustion ¼ Qgain by flue þ Qchar radiation Ts
þ Qflame radiation þ Qheat loss  
Tg  Ta 1
þ Qdoor loss þ QH2 moisture Qheat loss ¼ * (30)
1 n ðDo =Di Þ
þ l2pHk þ ðh 1 _f
1000*m
þ Qfuel moisture (21) ðhci þhrfl þhrch ÞAi i co þhro ÞAo

  The individual heat transfer coefficients in Eq. (30) were taken


LHV*hc ¼ mflue  hg  ha þ Qchar radiation þ Qflame radiation from Cengel [49].
For fully developed turbulent flow in smooth tubes:
þ Qheat loss þ Qdoor loss þ QH2 moisture þ Qfuel moisture
(22) k
hci ¼ *0:023 * Re0:8
D *Pr
0:3
(31)
Rearranging, D
From average Nusselt number expression for natural convection

LHVhc  Qchar radiation  Qflame radiation  Qheat loss  Qdoor loss  QH2 moisture  Qfuel moisture
mflue ¼   (23)
hg  ha

Solving Eqs. (20) and (23) simultaneously, the flame tempera- over vertical cylinder:
ture was obtained.
The individual heat transfers in Eq. (23) were defined mathe- k k
hco ¼ *0:59 * Ra0:25
H ¼ *0:59 * ðGrH PrÞ0:25 (32)
matically as given below. Rearranging the expressions for radiative H H
heat transfer from J. P. Holman [39] and using radiation network  
method: s* εg Tg4  ag Twi
4

  hrfl ¼   (33)
4 Tg  Twi
sA  Tchar  Tp4 1
Qchar radiation ¼ * (24)  
ð1εc Þ
þ 2 _f
1000*m 2
εc ð1þFcharpot Þ hro ¼ s*εo  Two þ Ta2 * ðTwo þ Ta Þ (34)

 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  
D2 þ 2ðH þ Wi Þ2  2ðH þ Wi Þ D2 þ ðH þ Wi Þ2 2
s* Tchar 2  ðT
þ Twi char þ Twi Þ
Fcharpot ¼ hrch ¼ ð1ε Þ Fcharwall þFcharpot
(35)
D2 c
þF
εc ðFcharwall þ2Fcharpot Þ
charwall *
(25)
From expression 8e19 of J. P. Holman [39] the rate of heat loss Fcharwall ¼ 1  Fcharpot  Fchardoor (36)
from fuel bed to the feed door:
Applying heat balance to the heat transfer from flame to
  1 surroundings,
4
Qdoor loss ¼ Fchardoor sA  Tchar  Ta4 * (26)
_f
1000*m
Q_ heat loss ¼ Q_ flameinner wall ¼ Q_ inner wallouter wall
The value of the view factor between char bed and feed door
was obtained from the ‘Catalog of radiation heat transfer configu- ¼ Q_ outer wallsurroundungs (37)
ration factors’ by John R. Howell [47] by approximating the situa-
 
tion as ‘Rectangle to perpendicular circular segment’ as Tg  Twi
Fchardoor ¼ 0.2. Q_ flameinner wall ¼ 1
(38)
ðhci þhrfl þhrch ÞAi
  1
Qflame radiation ¼ s*A* εg Tg4  ag Tp4 * (27)
_f
1000*m ðTwo  Ta Þ
Q_ outer wallsurroundungs ¼ 1
(39)
Following general expression from Shah and Date [19] the ðhco þhro ÞAo
emissivity of flue gas across different zones was estimated.
Solving above equations, values for Two and Twi were obtained.
 

AþB*ln 0:2*3:6*VAm
m
QH2 moisture ¼ ðhs  2547Þ0:09H (40)
εg ¼ e (28)

M
Qfuel moisture ¼ ðhs  4:187ðTa  273Þ  (41)
A ¼ 0:848 þ 9:02*104 *T 100  M
194 M.P. Kshirsagar, V.R. Kalamkar / Energy 93 (2015) 188e201

The enthalpy of superheated steam was correlated by the (   )


expression [40]: _ Tc þ Te 4 4
Q rad 3 ¼ sApg  ε3  aTp (51)
2
hs ¼ 0:00032273*T2 þ 1:64833661388*T þ 2021:531847
(42)  

Tc þ Te
Q_ conv 3 ¼ hpot gap Apg   Tp (52)
The other output parameters considered from the combustion 2
zone were:
 

Excess air factor: Tc þ Te


_ act *ðhe  hc Þ*1000 ¼ hpot gap *Apg *
m  Tp
2
ma (  )
l¼ (43) 
ms Tc þ Te 4
þ s*Apg * ε3  a3 *Tp4
2
Stoichiometric air required per kg of dry fuel burnt:
  (53)
8 100
ms ¼ C* þ H*8  O þ S * (44)
3 23  0:33
kpg Wi
h pot gap ¼ *0:424 * Re0:57
D * (54)
Generally, C, H, O, and S are obtained from ultimate analysis of D D
fuel; in absence of which, the following average values can be used
for wood or crop residue: C ¼ 50%, H2 ¼ 6.5%, O ¼ 42% and S ¼ 0.2%
[10,11,50e58]. The presented work used composition of Douglas fir
wood, as given by Anbroad [33,34]. 7.3.4. Pot side zone
Percentage oxygen in flue gas: The main processes considered in this zone were convective
heat transfer (laminar fully deloped flow) radiative heat transfer
l1
% O2 ¼ *21 (45) from flame to the pot side and heat loss from pot side to the
l
surroundings.

Q_ supplied by flue gas ¼ Q_ transfer to the pot þ Q_ side heat loss (55)
7.3.2. Pot bottom zone
The main process considered in this zone was the heat transfer (   )
_ Tc þ T0 4 4
to pot via impinging jet heat transfer (assuming no heat loss from Q rad 4 ¼ sAside  ε4  a4 Tp (56)
this zone). 2

 

Q_ lost by flue gas ¼ Q_ transfer to the pot Tc þ T0


Q_ conv 4 ¼ hside Aside   Tp (57)
2
 

Tg þ Te
Q_ conv 2 ¼ h bottom A   Tp (46)  

2 Tc þ T0
_ act ðhc  ho Þ1000 ¼ hside Aside 
m  Tp
2
 
(   )
  Tg þ Te
_ act  hg  he 1000 ¼ h bottom A 
m  Tp (47) Tc þ T0 4
2 þ sAside  ε4  a4 Tp4
2
The expression for impinging jet heat transfer (average) was
þ Q_ side loss
taken from Zuckerman et al [59].
 0:33 (58)
Wi
Nu bottom ¼ 0:424 * Re0:57
D * (48)   
D Tc þ T0
Q_ side loss ¼ 10Aside   Ta (59)
 0:33 2
k Wi
h bottom ¼ bottom *0:424 * Re0:57
D * (49)
D D Aside ¼ pDp hw (60)
Assuming combined convective and radiative heat transfer
coefficient ¼ 10 w/m2 K.
7.3.3. Pot gap zone From the expression of average Nusselt number, for laminar
The main processes considered in this zone were impinging jet flow over a flat plate [49].
heat transfer and radiative heat transfer from flame to pot bottom
(assuming no heat loss). kside
h side ¼ *0:664 * Re0:5
hw *Pr
1=3
(61)
hw
Q_ supplied by flue gas ¼ Q_ transfer to the pot (50)
Finally the overall efficiency of a stove:

Q_ char radiation þ Q_ flame radiation þ Q_ conv 2 þ Q_ rad 3 þ Q_ conv 3 þ Q_ rad 4 þ Q_ conv 4


ho ¼ 100 (62)
P
M.P. Kshirsagar, V.R. Kalamkar / Energy 93 (2015) 188e201 195

Fig. 2. A sample excel spreadsheet.


196 M.P. Kshirsagar, V.R. Kalamkar / Energy 93 (2015) 188e201

Q_ char radiation ¼ Qchar radiation m


_ f 1000 (63) parametric variation was required. The other details like, geomet-
rical and operational parameters were also needed. Only one
parametric study fulfilling all the above criteria was available in the
Q_ flame radiation ¼ Qflame radiation m
_ f *1000 (64)
literature. The experimental data set that included 10 data point
measurements of flame temperature, mass flow rate, and % O2, for
variations in firepower as shown in Table 3 were taken from lab-
8. Integration and execution of the model with Excel®
oratory investigations carried out by J Agenbroad on a standard
spreadsheet
rocket elbow [33,34]. The readings are temporal averages of 15-min
runs for almost constant firepower. The publications [33,34] pro-
The above-discussed equations were inserted and linked prop-
vided all required information regarding operating and physical
erly into a spreadsheet. An iterative procedure based upon the
parameters. Some of the missing inputs for the present model were
SOLVER add-in in the Excel® program was used to establish energy
obtained from the related literature review and personal commu-
and mass balances [60]. The spreadsheet-integrated model simul-
nication with the J Agenbroad [36]. Table 4 shows the operational
taneous determines six temperatures i. e. flame temperature,
type input parameters as used for the model validation.
entrance temperature, corner temperature, exit temperature, outer
The experimental data published by Agenbroad [33,34] was
wall temperature, and Inner wall temperature. The other output
consistent, with few notable irregularities. To take a closer look at it,
parameters were then determined from these six temperatures, in
the data was rearranged in the order of increasing firepower in-
combination with input parameters and flue gas properties.
dependent of the test number. It was expected that for the same
As shown in Fig. 2, the program requires 20 different inputs in
stove geometry, fuel, environment and test procedure, the flame
the form of 9 operational and 11 physical parameters for the given
temperature will increase, and the mass flow rate and % O2 will
stove geometry and operating conditions. The iterative procedure
decrease with the increasing firepower. However, from Table 3, one
predicts 31 output parameters deciding the performance of the
can observe that the mass flow rate in general is fluctuating and
stove for the given inputs. In addition, intermediate parameters
hence shows no direct relationship with the firepower or temper-
like, view factors and emissivity for different zones, and pressure
ature. In fact, the firepower reading of 2.96 kW reverses all the
loss coefficients at different sections of the stove are calculated. A
trends. From the firepower of 2.81 kW to 2.96 kW, there is a sudden
chart establishing energy balance in the combustion chamber is
drop in temperature and rise in % O2, with the mass flow rate
also given in the tool. Finally, the tool generates the energy balance
jumping to the maximum value. Keeping in view that each data
sheet for the entire stove along with the overall efficiency.
point under discussion is the time average of 15-min long test,
performed in one of the reputed laboratories related with cook-
9. Model predicted results and validation stove testing; with sophisticated instruments, this could not be an
experimental error. Furthermore, the 2.96 kW reading is not an
9.1. Data set, its investigation and identification of a new ‘absurd’ one; it is just that ‘somehow’ the same rocket elbow
operational parameter allowed the highest mass flow rate for intermediate firepower,
which decreased flame temperature and increased % O2. This could
In order to validate the proposed heat and mass transfer model, only be the difference of ‘the way the stove was operated’ during
a cohesive and consistent experimental data with sufficient each reading. This ‘unidentified operational parameter’, which was

Table 3
Model predicted results and comparison with the experimental values.

S. N. Test no. P (kW) hc (%) Ar Tg (K) _ act (kg/s)


m %O2 Cd Cl h0 (%)
Exp Mpr %Error Exp Mpr %Error Exp Mpr %Error

1 Test 1 1.43 0.971 0.78 589 562 4.7% 0.00301 0.00288 4.4% 17.36 17.13 1.3% 0.308 0.18 28.7%
2 Test 2 1.85 0.976 0.76 697 680 2.4% 0.0026 0.00265 2.0% 15.52 15.47 0.3% 0.285 0.21 26.0%
3 Test 1 2.15 0.987 0.78 727 729 0.3% 0.0028 0.00283 1.2% 15.25 14.99 1.7% 0.307 0.21 24.6%
4 Test 2 2.43 0.99 0.77 799 802 0.4% 0.00263 0.00267 1.8% 14.04 13.72 2.3% 0.295 0.23 24.0%
5 Test 2 2.81 0.995 0.755 895 900 0.6% 0.00237 0.00244 3.1% 12.16 11.63 4.4% 0.277 0.27 23.6%
6 Test 1 2.96 0.995 0.85 806 819 1.6% 0.00336 0.00345 2.9% 14.3 14.17 0.9% 0.383 0.2 21.9%
7 Test 1 3.57 0.994 0.8 929 977 5.2% 0.00267 0.00281 5.3% 10.55 10.63 0.8% 0.326 0.27 22.1%
8 Test 1 3.69 0.928 0.686 1017 1106 8.7% 0.00148 0.00162 9.8% 1.16 1.15 1.3% 0.195 0.37 22.0%
9 Test 2 3.73 0.995 0.78 1018 1028 0.9% 0.00246 0.00258 4.6% 9.35 9.03 3.5% 0.303 0.3 22.4%
10 Test 2 4.08 0.908 0.703 1082 1126 4.1% 0.00153 0.00177 15.4% 0.88 0.84 4.80% 0.214 0.36 20.20%

Exp e experimental results from Agenbroad [33,34] Mpr e model predicted results.

Table 4
Operational input parameters.

S. N. Parameter Value Unit Remark Ref.

1 Firepower 1.43e4.08 kW Reading specific in the given range [33,34]


2 Time interval of the test 15 minute e [33]
3 Combustion efficiency 90.8e99.5 % Reading specific [33,34]
4 Lower or net calorific value of fuel (dry basis) 18,280 kJ/kg Douglas fir, composition C4.4H6.3O2.5 [34]
5 Moisture content of fuel (wet basis) 7.0 % e [36]
6 Surface temperature of pot 373 K e [33]
7 Ambient temperature 300 K e [34]
8 Char emissivity 0.85 e [26]
9 Char temperature 1100 K e [34,38]
10 Width of the fuel sticks 20 mm e [36, 61]
M.P. Kshirsagar, V.R. Kalamkar / Energy 93 (2015) 188e201 197

not given expected consideration during stove modeling, was fraction of CO while calculating % O2, and the temporal averaging of
thought to be the missing link. It was because of the ‘unobserved’ experimental values with considerable deviations. Considering the
change in the neglected operational parameter that, the mass flow above discussion, the values predicted by the present heat transfer
rate suddenly increased; increasing the % O2 and decreasing the model within a ±10% deviation over the experimental values were
temperature. Similarly, the existence of different ‘achievable considered valid.
maximum firepower’ i.e. 3.69 kW in test 1 and 4.08 kW in test 2,
and the unexplainable ‘outliers’ [33,34] of very low mass flow rates 9.3. Model predicted results and validation
and % O2 pointed towards the existence of any such operational
parameter. As discussed in Section 8, the mathematical model was executed
The 2 nearly same firepowers from both tests i. e. 3.69 kW and using the spreadsheet, against input parameters from the 10-point
3.73 kW were showing different behavior. Both the data points at experimental data set as given in Tables 3 and 4. All the energy and
the same firepower and flame temperatures display remarkably mass balances resulting accurate within almost zero percent;
different combustion efficiency, mass flow rates, and % O2. The validated the model mathematically. The Solver tool allows for
3.73 kW firepower with 99.5% combustion efficiency, and 32,767 iterations; however less than 100 iterations were sufficient
0.00246 kg/s mass flow rate was following the general trend. every time to reach the solution with precision of 0.001, tolerance
However, the 3.69 kW firepower has only 92.8% combustion effi- of 0% and convergence of 0.001. The model was executed using
ciency and the least mass flow rate. In fact, the 3.69 kW firepower ‘Quadratic’ estimates because of the non-linear nature of the
was the highest achievable for the test 1, and hence represents equations. The relative errors in flame temperature, mass flow rate,
overfeeding of the stove with fuel sticks, in the attempt (“An and % O2 were calculated as by Persson et al. [62].
attempt will be made to reach the highest possible firepower for this
stove configuration”) [34]. Overfeeding of the stove probably choked Tex  Tpr
eT ¼ *100 (65)
the entrance section, leading to the lowest mass flow rate and Tex
hence oxygen starved fire with poor combustion efficiency. It
showed that apart from the same geometrical and physical pa- mex  mpr
em ¼ *100 (66)
rameters, there exists some ‘operational parameter related to mex
feeding conditions’, the difference in which was responsible for this
drastic difference in stove behavior. This ‘operational parameter’ is %O2ex  %O2pr
obviously, ‘the area available for air entrance at the stove inlet’. The e%O2 ¼ *100 (67)
%O2ex
parameter Ar shown in Table 2 was defined as:
Table 3 shows the comparison of values predicted by the model,
A Area unoccupied by fuel at the feed door and the experimental data set from Agenbroad. All the model
Ar ¼ in ¼ predicted values (except one) are within ±10% of experimental
A Cross sectional area of chimney
values. With 97% values falling within the range of ±10%, the model
Agenbroad or the other stove researchers have not considered was considered valid for assessing the performance of a natural
Ar in their mathematical models, nor was the variation in it draft biomass stove with unshielded pot. The values of Ar given in
observed during testing. The Ar though, will roughly depend on the Table 3 are the best-fitted values; for the optimum simultaneous
number of sticks fed, the cross-section of the fuel sticks and that of match between experimental and model predicted values for all
the chimney. three parameters. The possible range of Ar during these experi-
ments was estimated from the data obtained from Agenbroad [36].
9.2. Uncertainty in cookstove performance measurement and For about 3e5 sticks (2e6 for the extreme firepower) of cross
allowable deviations section 2 cm  2 cm fed in the 10 cm diameter rocket elbow, the
resulting Ar could have been in the range of 0.9e0.7. As the model-
There exists many studies in biomass cookstove literature where fitted values of Ar were in this range, these were considered valid.
the predicted values for different parameters like temperature,
fuel-burning rate, mass flow rate and heat transfer deviate from the 10. Discussion
experimental values by a considerable amount (as high as 50%)
[20e24,26e28,62e63]. However, these deviations were treated The following section compares the model predicted results
acceptable, mainly because of the inherent sources of discrep- with the experimental values, and those predicted by the model of J
ancies, such as the transient response of equipments, averaging of Agenbroad [33,34].
values and the errors associated with the dominant cookstove
testing protocol. The errors associated with these testing protocol 10.1. Flame temperature
(WBT) are: dependence of latent heat of vaporization on local
pressure, presence of ash with the char, reference state for the Variation of the flame temperature with the firepower is shown
specification of heating value, and actual moisture content of the in Fig. 3. As expected the flame temperature is increasing with the
fuel at the time of test [61,64]. According to Taylor [64], “the actual firepower. Values predicted by the present model fit better with the
deviation of the test results from true values is much more likely to be experimental values; than the values predicted by the Agenbroad
on the order of 10%.” MacCarty while validating a zonal model for model; the values predicted by which were “consistently over pre-
household biomass cookstoves said, “As such, the experimental dicted with measured values 20e30% lower than those predicted”
values should be considered with an inherent potential error of ‘at [34]. The temperatures predicted by the present model deviated
least’ ±5%.” Christian L'orange [65] while developing numerical from the experimental values within the range of ±0.3e8.7%, which
tools for predicting biomass cookstove impact made a ‘conservative was well within the range of ±10%. However, most of the values,
assumption’ of measurement error of ±10% [65]. especially, for medium to higher firepower were over predicted.
There are some possible sources of error specifically related to This could be due to non-uniformity of flame temperature field in
the chosen experimental data, like use of single step reaction in the radial directions. The thermocouple during experimentation by
combustion of volatiles as well as of char, negligence of the mole Agenbroad was located centrally, and hence measured the central
198 M.P. Kshirsagar, V.R. Kalamkar / Energy 93 (2015) 188e201

Fig. 5. Variation of % O2 with the firepower.

Fig. 3. Variation of flame temperature with the firepower. 10.4. Other important parameters

10.4.1. Heat loss coefficient


or maximum temperature of the flame, unlike the average tem-
Unlike an assumed heat loss coefficient in Agenbroad model, the
perature of the flame predicted by the present model [34].
present model predicted the value of a heat loss coefficient unique
to the given situation; based on the heat losses to the surroundings
10.2. Mass flow rate of flue as per heat transfer principles. The general value of the heat
transfer coefficient was in the range 0.18e0.37, which matches well
Variation of the mass flow rate with the firepower is shown in with the corresponding range 0.25e0.35 assumed by Agenbroad
Fig. 4. The mass flow rate is not a function of firepower alone, but and was close to his guess of “Up to one third of the firepower over a
also depends upon the other parameters such as the coefficient of typical stove operation is thought to be lost to the stove walls” [34].
discharge and the ‘Inlet area ratio’ in particular.
The model predicted values were in good agreement with the 10.4.2. Coefficient of discharge
experimental values; however, a value predicted for the highest Agenbroad used a constant value for the coefficient of discharge;
firepower was showing deviation of around 16%. This could be due the present model treated it as a variable, which in itself depends
to uncertain nature of the extreme firepowers. These firepowers are upon the other variables, i.e. ‘Inlet area ratio’ and the other
harder to create and hold at a steady-state condition, in natural geometrical variations. The model using the principles of fluid
draft stoves. mechanics as discussed in Section 7.3.1 predicted a unique value of
Cd for the given situation. The model predicted the coefficient of
discharge in the range of 0.195e0.383, the higher end of which,
10.3. % O2 matched with the value of 0.35 introduced by Agenbroad for the
model calibration [34]. However, it has also shown that the model
Variation of the % O2 in the flue gas is shown in Fig. 5. As ex- predictions with constant value of Cd can deviate considerably from
pected, the % O2 in general was found to decrease with the the actual values.
increasing firepower. The values predicted by the present model
deviated from the experimental values within the range of 10.4.3. Thermal efficiency
±0.8e4.8%, which was well within ±10%. The present model with The model predicted the thermal efficiency of the stove for
the help of a new parameter ‘Inlet area ratio’ also successfully given input parameters. It also predicted different heat transfer
predicted the behavior of the stove for ‘two outliers’ of about 1% and coefficients and the amount of heat transfer in each zone. The
1.2%. model predicted values for the give input data, in the range of
20.2e28.7%, which was within the plausible range of thermal ef-
ficiency for a natural draft, insulated; rocket type biomass cook-
stoves with unshielded pot [2,16,19,20,27]. However, in the absence
of corresponding efficiency values for the chosen data set, and
unavailability of any such consistent and cohesive data set in
literature, the thermal efficiency values predicted by the model
cannot be verified.

10.5. Unveiling an important parameter in cookstove modeling and


its effects

Several experimental studies evaluating the effect of different


parameters on the performance of biomass stoves were conducted
[12,13,19,20,26,33,61,66e68]. Parameters like firepower, turndown
ratio, flame temperature, flue flow rate, % O2, fuel moisture, stove
Fig. 4. Variation of mass flow rate with the firepower. geometry, fuel geometry, pot geometry, fuel feeding rate, method of
M.P. Kshirsagar, V.R. Kalamkar / Energy 93 (2015) 188e201 199

ignition, stove type, type of cooking task, cooking pot temperature, First, to ensure good combustion efficiency, a very low inlet area
ambient conditions, and fabrication materials were studied. How- ratio must be avoided. The Ar being an operational parameter will
ever, no study has taken in to account the effect of an operational vary for the given stove depending on the type of cuisine, fuel used
parameter, the ‘Inlet area ratio’ on the stove performance. In-fact (and its condition), and of course the operator. Hence, the value of
the parameter itself was defined for the first time. The Ar was un- Ar cannot be fixed beforehand; however, its effect on the cookstove
consciously used by previous researchers as an intermediate performance can be predicted by using the present model. The
parameter for calculating pressure loss [19,20,34], but not given existence of Ar also explains the failure of repeatability of laboratory
due consideration and hence, its effects on cookstove performance performance in the field. An actual operator in the field will operate
were never investigated. the given stove according to his/her need and not the way it was
run for ‘obtaining’ maximum efficiency and minimum emissions in
10.5.1. Effect on the mass flow rate the lab. The operator along with the fuel type, and its condition at
Fig. 6 shows the plot of experimental mass flow rate against the the time of cooking, will decide the real value of Ar. Operated in a
model fitted values of Ar from Table 3. It shows that the mass flow wrong way the ‘Inlet area ratio’ may turn a ‘good stove’ in to a ‘bad
rate increases in general with the increasing Ar; and there exists a stove’. As we are not in the position to fix this ‘best or optimal’ value
good correlation between the two. A linear curve was fitted to the of Ar, we cannot ensure the way stove will perform. It means that
experimental results from Agenbroad with R2 value of 89.4%. The just designing energy and emission efficient stove and dispatching
correlation between Ar and the mass flow rate is: it to the consumers is not sufficient. We should also study the other
important aspect of it; that is, how it is operated? In other words, a
_ act ¼ 0:012*Ar  0:006
m (68) cookstove must be designed after studying the technical, but more
importantly; the operational requirements for the particular
cuisine, like range of firepowers for local cuisine, the fuels used and
their general moisture content (depends upon local climate).
10.5.2. Effect on the combustion efficiency
The results further stress the use of standard or processed fuel
The combustion efficiency is a function of three Ts, Temperature,
for the given stove, having standard type, composition, size, and
Turbulence (mixing with oxygen), and Time. In natural draft stoves
shape. The LPG stove, which is the ‘universally accepted’ stove, has
with air velocity of around 1 m/s the burning volatiles has enough
this inherent advantage of burning the standard fuel; for which it is
of time to ensure good combustion. However, lower temperatures
exclusively designed (beside burning gaseous fuel), and hence
and the unavailability of oxygen in the required amount coupled
perform better under all operating conditions. A stove should be
with low flow velocities (low turbulence) may disrupt the com-
designed for a standard fuel, fed in standard quantity, resulting in a
bustion process. The higher Ar allows a rapid flow of air through the
standard range of Ar. More importantly, we have to ensure that
stove inlet and fuel bed leading to good mixing condition and
users do not deviate from these standards, whether consciously or
abundant supply of oxygen. It can be observed from Table 3 that for
unconsciously. The discussion also leads to the need of using forced
very low values of Ar the combustion efficiencies were badly
draft and/or gasification principles in future cookstoves. These
affected, as expected. For the highest firepower with lowest Ar, the
stoves uses recommended fuel in a standard batch size, and ensure
combustion was at its worst for both the tests. Here, the low values
good mixing and hence better performance [69e72].
of Ar leads to chocked flow, poor mixing and inadequate supply of
oxygen, resulting in the lowest combustion efficiencies in spite of
11. Model limitations and future scope
the high flame temperatures.
The model is applicable only for chimney type of combustion
10.6. Consequences of findings for future cookstove modeling and chamber stoves working under natural draft conditions and for
design wood burning. It is not applicable for charcoal stove or fan assisted
forced draft stoves or gasifier type stoves. Furthermore, the model
The operational parameter Ar, in essence represents the is applicable only for the stoves with the diameter of the com-
entrance area available for air and subsequently the resistance to air bustion chamber smaller than that of the pot. The model is not also
flow. It directly affects the mass of air entering the stove, which in applicable for stoves with a skirt. Although the model can predict
turns decide the mass flow rate of flue, flame temperature, excess heat transfer and thermal efficiency, it does not predict any
air ratio, combustion quality, and the highest possible firepower for combustion-related parameters such as CO, CO2 and the combus-
given configuration. Several new things can be learnt from the very tion efficiency. The research work has proven the existence of an
existence of this parameter, and its effect on stove performance. important operational parameter in the biomass cookstove
modeling, testing and performance prediction. However, the vali-
dation of the effect of variation in ‘Inlet area ratio’ on the different
performance parameters of a biomass cookstove needs a separate
experimental investigation.

12. Conclusion

A heat-transfer model for natural draft stove with unshielded


pot was developed and validated against experimental data set
from literature. The model predicted values agreed well with the
experimental data and were in the plausible range. The present
model was found to be more effective than the previous models.
The work also uncovered an important operational parameter that
is ‘Inlet area ratio’. It is concluded that the Ar has a considerable
effect on the performance of a natural draft stove, and hence be
Fig. 6. Variation of mass flow rate with the inlet area ratio. given adequate attention during modeling, design, and testing of
200 M.P. Kshirsagar, V.R. Kalamkar / Energy 93 (2015) 188e201

biomass stoves. Furthermore, there is a need of standard stove [26] Varunkumar S, Rajan NKS, Mukunda HS. Experimental and computational
studies on a gasifier based stove. Energy Convers Manage 2011;53:135e41.
design, coupled with standard fuel and standard operating proce-
[27] Miller-Lionberg DD. A fine resolution CFD simulation approach for biomass
dure, to deliver a good cooking experience (stove performance) to cook stove development [MS thesis]. Colorado: Colorado State University,
the customer. Department of Mechanical Engineering; Spring 2011.
[28] Bryden KM, Ashlock DA, McCorkle DS, Urban GL. Optimization of heat transfer
utilizing graph based evolutionary algorithms. Int J Heat Fluid Flow 2003;24:
Acknowledgments 267e77.
[29] Burnham-Slipper H. Breeding a better stove: the use of computational fluid
dynamics and genetic algorithms to optimise a wood burning stove for
The authors gratefully acknowledge the friendly support of Josh
eritrea, doctoral dissertation. Nottingham: University of Nottingham; 2008.
Agenbroad, for sharing the information related to the experimental [30] Varunkumar S. Packed bed gasification-combustion in biomass based do-
data, for a good cause. mestic stoves and combustion systems [PhD thesis]. Bangalore: Department
of Aerospace Engineering, Indian Institute of Science; 2012.
[31] Scharler R, Benesch C, A. Neudeck, Obernberger I. CFD based design and
References optimisation of wood log fired stoves. In: Proceedings of 17th European
Biomass Conference and Exhibition, From Research to Industry and Markets,
[1] Kshirsagar MP, Kalamkar VR. A comprehensive review on biomass cookstoves Hamburg, Germany, 29 June e 03 July 2009; 2009.
and a systematic approach for modern cookstove design. Renew Sustain En- [32] Agenbroad J, DeFoort M, Kirkpatrick A, Kreutzer C. A simplified model for
ergy Rev 2014;30:580e603. understanding natural convection driven biomass-cooking stoves e part 1:
[2] Still D, MacCarty N, Ogle D, Bond T, Bryden M. Test results of cook stove setup and baseline validation. Energy Sustain Dev 2011;15:160e8.
performance. Cottage Grove, OR: Aprovecho Research Center; 2011. London: [33] Agenbroad J, DeFoort M, Kirkpatrick A, Kreutzer C. A simplified model for
Shell Foundation; Washington DC: U.S. Environmental Protection Agency. understanding natural convection driven biomass cooking stovesdpart 2:
[3] Jetter J, Zhao Y, Smith KR, Khan B, Yelverton T, DeCarlo P, et al. Pollutant with cook piece operation and the dimensionless form,. Energy Sustain Dev
emissions and energy efficiency under controlled conditions for household 2011;15:169e75.
biomass cookstoves and implications for metrics useful in setting interna- [34] Agenbroad J. A simplified model for understanding natural convection driven
tional test standards. Environ Sci Technol 2012;46:10827e34. biomass-cooking stoves [MS thesis]. Colorado, Summer: Colorado State Uni-
[4] MacCarty N, Still D, Ogle D. Fuel use and emissions performance of fifty versity, Department of Mechanical Engineering; 2010.
cooking stoves in the laboratory and related benchmarks of performance. [35] Refractory fiber products, technical datasheets. New York, USA: Unifrax I LLC;
Energy Sustain Dev 2010;14(3):161e71. 2009. Form A1-004 and U-112 EN.
[5] Mac Carty N, Ogle D, Still D, Bond T, Roden C. A laboratory comparison of the [36] Agenbroad J. Personal communication. Rocky Mountain Institute; January
global warming impact of five major types of biomass cooking stoves. Energy 2015.
Sustain Dev 2008;12(2):56e65. [37] Urbas J, Parker WJ. Surface temperature measurement on burning wood
[6] Bryden M, Still D, Scott P, Hoffa G, Ogle D, Bailis R, et al. Design principles for specimens in the cone calorimeter and the effect of grain orientation. Fire
wood burning cook stoves. Cottage Grove, OR: Aprovecho Research Center; Mater 1993;17:205e8.
2006. London: Shell Foundation, Washington DC: U.S. Environmental Pro- [38] Dasappa S, Paul PJ. Gasification of char particles in packed beds: analysis and
tection Agency. results. Int J Energy Res 2001;25:1053e72.
[7] Prasad KK, Bussmann P, Visser P, Delsing J, Claus J, Sulilatu W, et al. Some [39] Holman JP. Heat transfer. 10th ed. New York: The McGraw-Hill Companies
studies on open fires, shielded fires and heavy stoves. The Netherlands: The Inc; 2010.
Wood Burning Stove Group, Departments of Applied Physics and Mechanical [40] Cengel Y, Boles MA. Thermodynamics: an engineering approach. 6th ed.
Engineering, Eindhoven University of Technology and Division of Technology Boston: McGraw-Hill; 2006.
for Society Apeldoorn; 1981. [41] Nag PK. Power plant engineering. 3rd ed. New Delhi: Tata McGraw-Hill
[8] Prasad KK, Sielcken MO, Nieuwvelt C, Nievergeld P, Sulilatu W, Meyvis J. Publishing Company Ltd; 2008.
A study on the performance of two metal stoves. Apeldoorn, The Netherlands: [42] Ong KS, Chow CC. Performance of a solar chimney. Sol Energy 2003;74:1e17.
The Wood Burning stove group Departments of Applied Physics and Me- [43] Sakonidou EP, Karapantsios TD, Balouktsis AI, Chassapis D. Modeling of the
chanical Engineering, Eindhoven University of Technology and Division of optimum tilt of a solar chimney for maximum airflow. Sol Energy 2008;82:
Technology for Society; 1981. 80e94.
[9] Bussmann P, Prasad KK. Parameter analysis of a simple wood-burning cook- [44] Koonsrisuk A, Lorente S, Bejan A. Constructal solar chimney configuration. Int
stove. In: Proceeding of the 8th International Heat Transfer Conference, San J Heat Mass Transf 2010;53:327e33.
Francisco, CA, vol. 6; 1986. p. 3085e90. [45] Prapas Jason. Toward the understanding and optimization of chimneys for
[10] Samuel B. Biomass stoves: engineering design, development, and dissemi- buoyantly driven biomass stoves [PhD thesis]. Fort Collins, Colorado, fall:
nation. Arlington, VA: Volunteers in Technical Assistance; 1987. Colorado State University; 2013.
[11] Sharma SK. Improved solid biomass burning cookstoves: a development [46] Zou Z, Guan Z, Gurgenci H, Lu Y. Solar enhanced natural draft dry cooling
manual. Field Document No. 44. Bangkok, Thailand: Food and Agriculture tower for geothermal power applications. Sol Energy 2012;86:2686e94.
Organization of the United Nations; 1993. [47] Howell JR. A catalog of radiation heat transfer configuration factors. University
[12] Andreatta D. A report on some heat transfer experiments. In: Proceedings of of Texas at Austin. Retrieved from: http://www.thermalradiation.net/index.
ETHOS conference, Seattle, Washington; 2005. html [accessed 15.12.14].
[13] Andreatta D. Temperature and heat flux distributions around a pot. In: Pro- [48] Bejan A, Kraus A. Heat transfer handbook. Hoboken, New Jersey: John Wiley &
ceedings of ETHOS Conference, Seattle, Washington; January 27e29 2006. Sons, Inc; 2003.
[14] Kshirsagar MP. Experimental study for improving energy efficiency of char- [49] Cengel Y. Heat transfer: a practical approach. 2nd ed. Mcgraw-Hill; November
coal stove. J Sci Industrial Res 2009;68:412e6. 2002.
[15] Andreatta D, Wohlgemuth A. An investigation of skirts. In: Proceedings of [50] Werther J, Saenger M, Hartge E-U, Ogada T, Siag Z. Combustion of agricultural
ETHOS conference, Kirkland, Washington; January 29e31, 2010. residues. Prog Energy Combust Sci 2000;26:1e27.
[16] Zube DJ. Heat transfer efficiency of biomass cookstoves [MS thesis]. Fort [51] Obenberger I, Brunner J, Barnthaler G. Chemical properties of solid biofuels e
Collins, CO: Colorado State University, Department of Mechanical Engineer- significance and impact, Institute for resource efficient and sustainable sys-
ing; 2010. Summer. tems, working group “Thermal biomass utilization”. Graz, Austria: Graz Uni-
[17] Gupta R, Mittal ND. Fluid flow and heat transfer in a single-pan wood stove. versity of Technology.
Int J Eng Sci Technol 2010;2(9):4312e24. [52] Yin C, Rosendahl LA, Kær SK. Grate-firing of biomass for heat and power
[18] Castillo T. Experimental investigation of the velocities in three small biomass production. Prog Energy Combust Sci 2008;34:725e54.
cookstoves [MS thesis]. Ohio: Columbus, The Ohio State University; 2011. [53] Ragland KW, Aerts DJ. Properties of wood for combustion analysis. Bioresour
[19] Shah R, Date AW. Steady-state thermochemical model of a wood-burning Technol 1991;37:161e8.
cook- stove. Combust Sci Technol 2011;183(4):321e46. [54] Domalski ES, Jobe Jr TL, Milne TA. Thermodynamic data for biomass conver-
[20] MacCarty NA. A zonal model to aid in the design of household biomass sion and waste incineration, National Bureau of Standards. US: Solar Technical
cookstoves [MS thesis]. Ames, Iowa: Iowa State University; 2013. Information Program of the Solar Energy Research Institute; September 1986.
[21] Kausley SB, Pandit AB. Modeling of solid fuel stoves. Fuel 2010;89(3):782e91. [55] Duku MH, Essel SG, Hagan B. A comprehensive review of biomass resources
[22] Wohlgemuth A, Mazumder S, Andreatta D. Computational heat transfer and biofuels potential in Ghana,. Renew Sustain Energy Rev 2011;15:404e15.
analysis and design of third-world cookstoves. J Therm Sci Eng Appl 2009;1: [56] Telmo C, Lousada J. Heating values of wood pellets from different species.
1e10. Biomass Bioenergy 2011;35:2634e9.
[23] Weerasinghe WMSR, Kumara UDL. CFD approach for modeling of combustion [57] Kalschimit M, Thran D, Smith KR. Renewable energy from biomass, encyclo-
of a semi enclosed cooking stove. In: Proceedings of The International Con- pedia of physical sciences and technology. 3rd ed., vol. 14. Burlington, MA:
ference on Mechanical Engineering, Dhaka, Bangladesh; 2003. Academic Press/Elsevier; 2002.
[24] Misra JS. Considering value of information when using CFD in design [MS [58] Obernberger I, Biedermann F, Brunner T, Sippula O, Jokiniemi J, Finnan John,
thesis]. Ames, Iowa: Iowa State University; 2009. et al. Design and operation concepts for low-emission biomass grate furnaces
[25] Ravi MR, Kohli S, Ray A. Use of CFD simulation as a design tool for biomass based on advanced air staging, ERA-NET bioenergy, project“FutureBioTec”.
stoves. Energy Sustain Dev 2002;VI(2):20e7. Austria: BIOENERGY 2020þ GmbH Graz; October 2012.
M.P. Kshirsagar, V.R. Kalamkar / Energy 93 (2015) 188e201 201

[59] Zuckerman N, Lior N. Jet impingement heat transfer: physics, correlations, and [66] Raman P, Ram NK, Murali J. Improved test method for evaluation of bio-mass
numerical modeling. Adv Heat Transf 2006;39. http://dx.doi.org/10.1016/ cook-stoves. Energy 2014;71:479e95.
S0065-2717(06)39006-5. ISSN 0065e2717. [67] Bhattacharya SC, Albina DO, Khaing AM. Effects of selected parameters on
[60] Fylstra D, Lasdon L, Watson J, Waren A. Design and use of the microsoft excel performance and emission of biomass- red cookstove. Biomass Bioenergy
solver. Interfaces 1998;28(5):29e55. 2002;23:387e95.
[61] L'Orange C, DeFoort M, Willson B. Influence of testing parameters on biomass [68] Hudelson NA, Bryden KM, Still D. Global modeling and testing of rocket
stove performance and development of an improved testing protocol. Energy stove operating variations. Cottage Grove, Oregon: Department of Mechan-
Sustain Dev 2012;16:3e12. ical Engineering, Iowa State University, Ames and Aprovecho Research
[62] Persson T, Fiedler F, Nordlander S, Bales C, Paavilainen J. Validation of a dy- Center.
namic model for wood pellet boilers and stoves. Appl Energy 2009;86: [69] Smith KR, Dutta K. Cooking with gas, editorial. Energy Sustain Dev 2011;15:
645e56. 115e6.
[63] Dixit CBS, Paul PJ, Mukunda HS. Part II: experimental studies on a pulverised [70] Still D, MacCarty N. Developing stoves to achieve the “50%/90%” future: stoves
fuel stove. Biomass Bioenergy 2006;30:684e91. in use that address health and climate issues. In: Proceedings of ETHOS
[64] Taylor III RP. The uses of laboratory testing of biomass cookstoves and the conference, Kirkland, Washington; 2011.
shortcomings of the dominant U.S. Protocol [MS thesis]. Ames, Iowa: Iowa [71] MacCarty N, Cedar J. The side-feed fan stove. In: proceedings of ETHOS con-
State University; 2009. ference, January 29e31, Kirkland, Washington; 2010.
[65] L'orange C. The development of numerical tools for characterizing and [72] Witt MJ. An improved wood cookstove: harnessing fan driven forced draft for
quantifying biomass cookstove impact [Doctoral THESIS]. Fort Collins, Colo- cleaner combustion. Cottage Grove, OR: Hartford CT: Trinity College; Apro-
rado: Colorado state university; 2013. Summer. vecho Research Center; May 2005.

También podría gustarte