Anda di halaman 1dari 17

Modeling and Experimental

Studies on Combustion
Characteristics of Porous
Coal Char: Volume
Reaction Model
ANUP KUMAR SADHUKHAN,1 PARTHAPRATIM GUPTA,1 RANAJIT KUMAR SAHA2
1
Department of Chemical Engineering, National Institute of Technology, Durgapur 713 209, West Bengal, India
2
Department of Chemical Engineering, Indian Institute of Technology, Kharagpur 721 302, West Bengal, India
Received 30 July 2008; revised 5 January 2009, 1 April 2009, 29 December 2009; accepted 4 January 2010
DOI 10.1002/kin.20483
Published online in Wiley InterScience (www.interscience.wiley.com).

ABSTRACT: A generalized single-particle model for the prediction of combustion dynamics of a


porous coal char in a fluidized bed is analyzed in the present work using a volume reaction model
(VRM). A fully transient nonisothermal model involving both heterogeneous and homogeneous
chemical reactions, multicomponent mass transfer, heat transfer with intraparticle resistances,
as well as char structure evolution is developed. The model takes into account convection and
diffusion inside the particle pores, as well as in the boundary layer. By addressing the Stefan
flow originated due to nonequimolar mass transfer and chemical reactions, this work enables a
more realistic analysis of the combustion process. The model, characterized by a set of partial
differential equations coupled with nonlinear boundary conditions, is solved numerically using
the implicit finite volume method (FVM) with a FORTRAN code developed in-house. The use
of a FVM for solving such an elaborate char combustion model, based on the VRM, was not
reported earlier. Experiments consisting of fluidized-bed combustion of a single char particle
were carried out to determine the internal surface area of a partially burned char particle and
to enable model validation. Predicted results are found to compare well with the reported
experimental results for porous coal char combustion. The effects of various parameters (i.e.,
bulk temperature and initial particle radius) are examined on the dynamics of combustion
of coal char. The phenomena of ignition and extinction are also investigated.  C 2010 Wiley

Periodicals, Inc. Int J Chem Kinet 42: 299–315, 2010

due to its importance in many industrial applications


INTRODUCTION
including electric power generation. Various models
were developed for porous char combustion in the
Combustion of coal char has been subjected to a large
past two decades including the earlier work by the
number of theoretical and experimental investigations
authors [1–7], and a comprehensive review on igni-
tion and combustion of porous char was published by
Correspondence to: Parthapratim Gupta; e-mail:
parthagupta2000@yahoo.com. Ballal et al. [8]. Formation of a CO flame due to the
c 2010 Wiley Periodicals, Inc. homogeneous reaction (Eq. (3)) during combustion is
300 SADHUKHAN, GUPTA, AND SAHA

referred to as ignition, whereas quenching of the CO of the internal pore surface area may not be accessible
flame is termed extinction. Linjewile and Agarwal [9] to the oxidizer either due to pore size limitations or due
and Biggs and Agarwal [10] investigated the CO/CO2 to covering of inert gaseous component like nitrogen.
product ratio for a single porous char particle, whereas A detailed study by Radovic et al. [19] revealed the
Stubington [11] measured the radial variation of tem- importance of pore surface area during the combustion
perature of char particle in a fluidized bed. The phe- of coal char at 0.1 MPa in air at 700 K.
nomenon of fragmentation during gasification of mi- In practice, combustion of coal char occurs both
croporous chars was reported by Feng and Bhatia [12], at the surface and inside, depending on the porosity/
based on experimental investigations, and a theoretical structure of the char particle. Nonporous coal char
analysis of peripheral fragmentation during char com- particles undergo combustion with varying sizes and
bustion and gasification was presented by Wang and constant density (shrinking sphere model) or with con-
Bhatia [13]. Waters et al. [14] determined the carbon stant size and a shrinking unreacted core at the center
mass loss with reactor height for a spherical carbon (shrinking core model). But for porous char particles,
particle, simultaneously measuring particle tempera- the combustion occurs with constant size and vary-
ture, size, and velocity in an entrained flow reactor. ing particle density (volume reaction model). Detailed
Most of these studies addressed different issues that single particle mathematical analyses using shrinking
are generally independent and unrelated. As a result, sphere and shrinking core models were already ad-
the application of conventional models is often for nar- dressed by the authors in earlier communications [3,4].
row ranges of operating conditions. In many cases, In the present study, a volume reaction model with
model predictions do not fit the temporal profiles of porous coal char is developed and the specific surface
experimental particle mass loss and temperature. area, accessible porosity, and conversion-time profiles
Coal, when inside a furnace, immediately pyrolyzes are estimated experimentally.
into devolatilized char and evolved volatiles. The A generalized fully transient, one-dimensional,
volatiles ignite via gas-phase combustion and help the nonisothermal, volume reaction model was used in the
process of ignition of devolatilized char. Combustion present work for combustion of a coal char with con-
of volatiles occurs relatively fast, whereas the combus- stant particle size. It consists of (1) multicomponent
tion of the resultant char (involving an interaction of mass transfer with chemical reaction in porous char
heterogeneous and homogeneous reactions with trans- particle, as well as in the gas boundary layer around
port resistances) is slow and often rate determining. For the particle (following the Maxwell–Stefan analysis
rapid pyrolysis of spherical coal particles, cenospheres of Krishna and Wesselingh [20]), (2) heat transfer be-
are formed and lead to generation of large pore area. tween the gas phase and the particle considering con-
For slow pyrolysis, less internal surface area is evolved ductive, convective and radiative modes of heat trans-
due to decreased yield of volatile mater. Burned Indian fer, (3) microstructure from the experimental findings
bituminous coal particles with 40% volatile matter, col- of surface area and accessible porosity, and (4) estima-
lected from pulverized fuel-fired furnaces, show evi- tion of physical properties as a function of temperature
dence of internal combustion, as reported by Smith and composition under simulated operating condition.
and Tyler [15]. Bar-Ziv and Kantorovich [16] experi- These coupled submodels were solved simultaneously
mentally studied the effect of porous structure (poros- based on the finite volume method (FVM). It may be
ity and shrinkage factor) on char reactivity. D’Amore noted that the use of the FVM for solving such an elab-
et al. [17] experimentally determined the pore size dis- orate char combustion model, consisting of the above-
tribution during carbon conversion of char particles. mentioned submodels, has not been reported earlier.
Hurt et al. [18] observed experimentally that internal Finally, the model is validated with the experimental
surface area of coal increased first with carbon conver- findings of coal char combustion in a fluidized bed. A
sion due to the opening up of new pores, but after a more elaborate mathematical analysis is available in
critical point the internal surface area decreased with Gupta and Saha [5–7], where multicomponent mass
carbon conversion due to pore-collapsing phenomena. transfer with chemical reactions of any order and heat
Micropores typically constitute more than 95% of the transfer for general gas-solid, noncatalytic reactions
internal surface area of char particles and provide a were investigated using the FVM.
large number of sites for heterogeneous reactions when
the reactant gas molecules can penetrate the pores. The
mean free path of oxygen molecules is in the order of THE MODEL
50–150 nm, and its molecular diameter is 0.4 nm only.
Hence oxygen molecule can penetrate into the microp- The different submodels including combustion chem-
ores and reacts with carbon internally. However, some istry, mass and energy balances, and pore surface area

International Journal of Chemical Kinetics DOI 10.1002/kin


COMBUSTION CHARACTERISTICS OF POROUS COAL CHAR: VOLUME REACTION MODEL 301

evolution during coal char combustion are described The small amount of H2 O vapor (1% by volume)
below. present in the bulk gas is assumed only to catalyze the
oxidation reaction of CO without taking part in the
gasification reaction. A similar approach was adopted
Combustion Chemistry by Biggs and Agarwal [10] to study the CO/CO2 prod-
The char particle is assumed to be surrounded by a uct ratio around a porous char particle in an incipient
stagnant gas boundary layer through which oxygen dif- fluidized bed.
fuses from the bulk gas and reacts with the char particle.
The char consists of reactive carbon and nonreactive Pore Structure and Surface Area Evolution
ash. During char combustion only reactive carbon re- The coal char particle is assumed to contain both
acts with gaseous reactants by heterogeneous surface micropores and macropores. The macroporosity acts
reactions within the pores to produce CO2 and CO, as channels for transport of gaseous species, and the
which diffuse out toward the bulk phase. Owing to the micropores act as the main reaction sites. It is assumed
high temperature of char particles, CO reacts with O2 that both microporosity εμ and macroporosity ε in-
to form CO2 by homogeneous reaction. The reaction crease with carbon conversion. The microporosity at
occurs both within the particle pores and in the gas any location in the radial direction may be modeled
boundary layer around the char particle. In the case of with the smooth field approximation of Jackson [26].
fluidized bed combustion, the homogeneous reaction The macroporosity of the porous particle is often re-
may not be significant in the boundary layer, though it ferred to as “accessible porosity” and may be modeled
takes place inside the particle pores [10]. Hence, in the using the percolation approach of Shah and Ottino [27]:
present model no a priori assumption has been made
on this account to ensure the applicability of the model ⎧
in general. Combustion chemistry similar to Biggs and ⎪
⎨0 0 ≤ ε ≤ 0.32
Agarwal [10] is used in the present work. Two hetero- ε = 0.995(ε − 0.32)0.3938 0.32 ≤ ε ≤ 0.5 (4)


geneous reactions and one homogeneous reaction are ε 0.5 ≤ ε ≤ 0.68
considered:

η+1 2η However, in the present study we have used a simplified


k1
Reaction 1 : 2 C + O2 −→ CO relation
η+2 η+2
2 ε = ε0 + (εf − ε0 )ξ (5)
+ CO2 (1)
η+2
k2
ξ = (W − Wash )/(W 0 − Wash ) (global carbon
Reaction 2 : C + CO2 −→ 2CO (2) conversion for experimental studies)
k3
Reaction 3 : 2CO + O2 −→ 2CO2 (3) = 1 − Wc /Wc0 (local carbon conversion for
simulation studies). (6)
The proposed chemical mechanism, though quite
simple, is very effective for describing the combustion where ε0 and εf , the initial and final accessible porosity
behavior of coal char and has been used in a number for devolatilized char and fully converted particles, re-
of other studies [10,13,21]. The rate expression and spectively. These were determined experimentally us-
kinetic values are presented in Table I. ing a mercury porosimeter.

Table I Chemical Reaction Rate Expression and Kinetic Values


Reaction Rate Equation Value Source
(1) 0 exp(− E1 )(p ) moles
Rs1 = ks1 of O2 0 = 127,080 mol m−2 atm−1 s−1
ks1 Smith [22]
RT O2
consumed/s m2 of surface E1 = 179,400 J mol−1
η = 70 exp(−3070/T ) Tognotti et al. [23]
(2) 0 exp(− E2 )(p
Rs2 = ks2 0 = 5.301 × 107 mol K m−2 atm−1 s−1
RT CO2 /T ) moles of O2 ks2 Dutta et al. [24]
consumed/s m2 of surface E2 = 248,120 J mol−1
(3) Rv3 = kv3
0 exp(− E3 )c c0.5 c0.5 moles of
RT CO O2 H2 O
0 = 1.3 × 108 m3 mol−1 s−1
kv3 Howard et al. [25]
CO consumed/s m3 of gas E3 = 125,400 J mol

International Journal of Chemical Kinetics DOI 10.1002/kin


302 SADHUKHAN, GUPTA, AND SAHA

The parameter ξ is the carbon conversion in solid resents the generation of the property, mass or enthalpy,
char particle. For the experimental studies, the global per unit volume due to chemical reactions.
carbon conversion is calculated from the experimen-
tal values of the instantaneous mass of char, whereas Solid Particle Phase. Component mass balance
local carbon conversion, used in simulation studies, is equation:
estimated from calculated instantaneous values of the
mass concentration of carbon in char at a given location ∂ 1 ∂ 2

(ε · ρg,s mk,s ) + 2 r Ntg,s Mav mk,s


(Eq. (6)). ∂t r ∂r
The pore surface area per unit mass Sm of partially l=3
1 ∂ 2 ∂mk,s
burned char at different carbon conversion level is mea- = Dke,s ρg,s 2 r + Rvl γkl Mk
sured experimentally and converted to specific surface r ∂r ∂r l=1
area expressed per unit volume of particle by multiply- (11)
ing Sm by the particle bulk density ρs . The bulk density
of the char particle is estimated at different values of The convective term contains the molar flux Ntg,s ,
the carbon conversion ξ , using the following equation: which is the superficial value of the molar flux in the
 0  solid particle phase, and the diffusive term contains an
ρs = ρs0 Xash + Xc0 (1 − ζ ) kg/(m3 of char particle) effective diffusivity, Dke,s , which includes the porosity
(7) effect.
Total molar balance of gas mixture:
To express the reaction rates per m3 of solid char,
the heterogeneous reaction rates (reactions (1) and ∂(ε · ct,s ) 1 ∂ 2
3 5
+ 2 r Ntg,s = γk l Rv l (12)
(2) in Table I) are multiplied by the particle specific ∂t r ∂r l=1 k=1
surface area, whereas the homogeneous reaction rate
(reaction (3)) is multiplied by porosity. Mass balance of solid carbon reactant in solid char
particle:
Rv,i = Rs,i S, i = 1, 2 (8)

mol dWc η+1
Rv3 = 
Rv3 ε (9) =− 2 Rv1 + Rv2 Mc (13)
m3 of solid char · s dt η+2

The variation of specific surface area per unit volume Energy balance:
of solid char S with carbon conversion ξ in solid char
particle is modeled using the random pore model of ∂(cps ρs Ts ) 1 ∂ 2

+ 2 r Ntg,s Mav,s cpg,s Ts


Bhatia and Perlmutter [28] as ∂t r ∂r

1 ∂ 2 ∂Ts
S  = λe 2 r + Rvl (− Hl ) (14)
= (1 − ξ ) · 1 − ψ ln(1 − ξ ) (10) r ∂r ∂r
S0
In writing Eq. (14), it is assumed that the gas mixture
where S0 is the pore surface area at zero carbon conver- present inside the pore volume of the porous particle is
sion of char and ψ is the dimensionless pore parameter, always in thermal equilibrium with the solid particle.
indicating the nature of the pore structure. The param-
eter ψ may be obtained by fitting the experimental val- Gas Boundary Layer. Component mass balance
ues of S(ξ ) at different carbon conversion to Eq. (10). A equation:
novel method by Kassebaum and Chelliah [29] decou-
pled the internal surface area from the surface kinetic ∂ 1 ∂ 2

rates. (ρg mk ) + 2 r Ntg Mav mk


∂t r ∂r

1 ∂ 2 ∂mk 
= ρg Dke 2 r + γk3 Rv3 Mk (15)
Conservation Equations r ∂r ∂r
The conservation equations for both mass and energy Total molar balance of gas mixture:
are presented in this section along with the initial
and boundary conditions. In general, the conservation
1 ∂ 2

5
∂(ct )
equations include accumulation, convection, diffusion, + 2 r Ntg = γk 3 Rv 3 (16)
and the source terms, respectively. The source term rep- ∂t r ∂r k=1

International Journal of Chemical Kinetics DOI 10.1002/kin


COMBUSTION CHARACTERISTICS OF POROUS COAL CHAR: VOLUME REACTION MODEL 303

Energy balance: At r = rb , i.e., edge of boundary layer near the free


stream, the Dirichlet boundary conditions are
∂ 1 ∂ 2

(Cpg ρg Tg ) + 2 r Ntg Mav Tg Cpg mk = mk,0 (32)


∂t r ∂r

1 ∂ 2 ∂Tg  Tg = T0 (33)
= λg 2 r + Rv3 (− H3 ) (17)
r ∂r ∂r
Expression for Molar Fluxes. Calculation of gaseous
Initial Conditions. At t = 0 diffusion flux is a complex problem in the case of a
multicomponent mixture and is described by Krishna
and Wesselingh [20], using a general matrix method
mk,s (r, 0) = m0k,s (r) (18)
solution of the Stefan–Maxwell equations. The mo-
Ts (r, 0) = Ts0 (r) (19) lar fluxes of gaseous components in the gas boundary
layer and within the voids of porous particle consist of
mk (r, 0) = m0k (r) (20)
diffusive transport and the total flux generated due to
Tg (r, 0) = Tg0 (r) (21) nonequimolar counter diffusion. A simplified formula-
tion is used here for fluxes involving the total flux and
Ntg (r, 0) = Ntg,s (r, 0) = 0 (22)
effective diffusion coefficient of each species, which
Wc = Wc0 (23) arises out of the nonequimolar and nonisothermal con-
ditions.
Boundary Conditions
Nk = xk Ntg − Dke ct grad(xk ) (34)

Solid Particle Phase. At r = 0


5
Ntg = Nk (35)
k=1
Ntg,s = 0 (24)

∂mks   Other Equations
−Dke ρg,s  + Ntg,s Mav,s mk,s r=0 = 0 (25)
∂r r=0
 xk = mk Mav /Mk and xk,s = mk,s Mav /Mk
∂Ts  
−λe  + Ntg,s Mav,s cpg,s Ts r=0 = 0 (26) (36)
∂r r=0

5
5
5
5

At r = rs xk = mk = xk,s = mk,s = 1 (37)


k=1 k=1 k=1 k=1

mk,s = mk (27) ct = P /RT (38)

Ntg,s = Ntg (28) ρg = P Mav /RT = ct · Mav (39)



∂Ts  
− λe + Ntg,s Mav,s cpg,s Ts r=rs
∂r r=rs Estimation of Thermophysical Properties


− σ ∈r Ts4 − T04 r=rs The physicochemical and thermal properties involved
 in the various conservation equations are often con-
∂Tg   sidered as constant in the literature. However, in the
= −λg  + Ntg Mav cpg Tg r=rs (29) case of nonisothermal chemical reactions the variation
∂r r=rs
of these parameters with temperature and composi-
tion must be taken into account. It will therefore be
Gas Boundary Layer. At r = rs
considered that the specific heat of the gas mixture
 and the heat of reaction are a function of temperature
∂mk,s  
− Dke,s ρg,s  + Ntg,s Mav,s mk,s r=rs and composition. The binary diffusion coefficients Dkj
∂r r=rs are calculated as a function of temperature, using the
 Chapman–Enskog equation as described by Bird et al.
∂mk  
= −Dke ρg  + Ntg Mav mk r=rs (30) [30]. Considering the presence of a multicomponent
∂r r=rs
gas mixture, the dependency of the mole fraction on
Tg = Tg,s (31) Dkj can be estimated by the Stefan–Maxwell equation

International Journal of Chemical Kinetics DOI 10.1002/kin


304 SADHUKHAN, GUPTA, AND SAHA

and the effective diffusivity Dke may be calculated as Table II Thermophysical Properties Used for the
Present Study
Property Value Property Value

Nk − xk 5j =1 Nj cps (J kg−1 K−1 ) 1458 ε0 0.15
Dke = 5 Nk xj −Nj xk (40)
j =1
ρso (kg m−3 ) 1240 εf 0.65
Dkj
dp (mm) 4.36 εr 0.90
Sphericity 0.88 S0 (m2 g −1 ) 17.00
τ 1.5–2.2 σ (J m−2 5.67 × 10−8
A more elaborate discussion on Eq. (40) is available K−4 s−1 )
in [3,4]. The effective diffusivity of gaseous compo-
nents within the void volume of porous solid is calcu-
lated by considering the influence of porous texture of
the particle, and by incorporating the accessible poros- NUMERICAL METHOD
ity and tortuosity factor of particles at different stages
of carbon conversion. A similar equation was used by a The conservation equations (Eqs. (11)–(17)) are sim-
number of researchers including Rafsanjani et al. [31], ilar to those of the diffusion/convection problem dis-
Patission et al. [32], and Sotirchos and Amundson [33]. cussed by Patankar [37] and can be solved by an im-
plicit FVM. In implicit formulations, a larger time step
may be chosen at the expense of a higher computation
εDke time at each time step. FVM has successfully been
Dke,s = (41)
τ employed by Gupta and Saha [5–7] to solve general
fluid–solid, noncatalytic reactions, where the merits of
FVM in handling such problems have already been
The tortuosity factor τ was calculated using the ex- discussed.
perimental data of mercury porosimeter after Carniglia The governing equations (Eqs. (11)–(17)) contain
[34] for devolatilized char, partially burned char and unsteady, convective, diffusion, and source terms. Inte-
fully converted char. The value varied between 1.5 and grating the governing equations along with the bound-
2.9, and an average value of 2.2 was used in the present ary conditions, after multiplying by the volume of the
study. The viscosity of the gaseous components was subdomain, results in discretized equations that can
estimated by the Chapman–Enskog equation, whereas easily be presented by a set of linear equations. The
the Wilke method was used to determine the viscosity nonlinear source terms are linearized by the Taylor se-
of the gas mixtures, as cited by Perry and Green [35]. ries approximation. The resultant equations are solved
Although the thickness for the mass transfer boundary using tri-diagonal matrix algorithm. A FORTRAN
layer was assumed to be the same as the particle radius code was developed to solve the model equations. The
by Biggs and Agarwal [10], in the present study it is cal- convergence criterion was selected as 10−4 for all nor-
culated using the correlation developed by Palchonok malized principal variables.
et al. [36] for fluidized beds. The thermal conductiv-
ity of the gas mixture is a function of temperature
and composition and is estimated by the Wilke and
EXPERIMENTAL
Eucken equation presented in Bird et al. [30]. The ef-
fective thermal conductivity of the porous char particle
Coal Char Preparation from Coal
accounts for conduction through the solid and conduc-
tion through the gas trapped within the pores (possibly Subbituminous coal samples of Indian origin (6 mm
influencing the gas flow through the pores). There- in diameter) were heated in an atmosphere of flow-
fore, the effective thermal conductivity of the porous ing nitrogen at 1300 K for a period of more than
solid char depends not only on temperature and gas 1 h. The devolatilized char particles were screened
composition but also on the composition of the solid through Tyler mesh (#31/2 and #4) with particle di-
and the porosity of the particle. The detailed equations ameters in the range of 4.75–5.60 mm. Each particle
for thermophysical parameters are available in earlier was separately held and individually filed off to re-
communications [3,4]. The other thermophysical prop- move the sharp edges and made into spherical particles.
erties for coal char are presented in Table II. Sphericity The size of the particle was measured in three mutu-
is defined as the ratio of the surface area of a sphere ally perpendicular directions by a vernier scale. The
having the same volume as the given particle to the equivalent diameter of the particle was estimated to be
surface area of the particle. the cube root of the products of the three measured

International Journal of Chemical Kinetics DOI 10.1002/kin


COMBUSTION CHARACTERISTICS OF POROUS COAL CHAR: VOLUME REACTION MODEL 305

Table III Proximate and Ultimate Analysis of the Coal


Sample 3 3
Proximate (db) Ultimate (daf) 2
Analysis (wt%) Analysis (wt%)
Parent coal Carbon 86.46
Volatile 27.0 Nitrogen 1.82
Fixed carbon 58.1 Hydrogen 4.51
Ash 14.9 Sulfur 0.15
Oxygen (by diff.) 7.06 3 1 14
1
Resulting char Resulting char 15
Volatile 3.9 Carbon 89.24 4
13
Fixed carbon 75.1 Nitrogen 1.43 11
Ash 21.0 Hydrogen 0.50 5
Sulfur 0.03 6 6
12 9
Oxygen (by diff.) 8.80
8 7

10
dimensions, and the sphericity of the particle was esti-
mated as the ratio of equivalent diameter to the longest
dimension of the particle [38]. Particles with a nar- Figure 1 Fluidized bed combustor: 1. Refractory brick
row size range having equivalent diameter of 4.36 ± wall, 2. feeding pulley, 3. mirror arrangement, 4. thermo-
couple, 5. distributor, 6 liquid petroleum gas entry, 7. orifice
0.08 mm and sphericity of 0.88 ± 0.02 were chosen for
meter, 8. air control valve, 9. Screen supporting glass beds,
experiments. 10. air blower, 11. digital multimeter, 12. digital display, 13.
The resulting char samples were stored in a closed bed material, 14. basket, and 15. char sample inside basket.
container under an atmosphere of nitrogen. Table III
shows the proximate and ultimate analysis of the
coal/char samples. employed for this purpose. Prior to the gas adsorption
measurements, char samples were degassed at 200◦ C
for a period of 2 h. Nitrogen adsorption isotherms
Combustion of Char Samples were measured over a relative pressure (P /P0 ) range
The combustion of char sample was carried out in a of 1.67 × 10−6 to 0.995. The sample contained in an
batch fluidized bed reactor (Fig. 1). A basket contain-
ing the char sample was dropped into a hot fluidized Table IV Operating Condition in a Fluidized Bed
bed at a desired bed temperature. Two to three parti- Combustor
cles each weighing about 0.055 g were used in each
experiment run with a total sample mass of about 0.11– Material of construction Stainless steel
0.165 g. The fluidized bed combustor consisted of a Inside diameter (mm) 82
stainless steel column, 82-mm diameter, fitted with a Bed inert Crushed refractory with a
stainless steel gas distributor, having 1.6-mm-diameter mean diameter 0.43 mm
holes with 1.2% open area. A static bed, 10 cm in and density 2100 kg/m3
height, made of crushed refractory of 0.43-mm mean Distributor Stainless steel plate, hole
diameter, was used. Air (21% O2 , 78% N2 , and 1% diameter 1.6 mm and open
area 1.12%
H2 O) was the fluidizing medium; the operating condi-
Fluidizing medium Air
tions are presented in Table IV.
Operating temperature, 973, 1073
Tb (K)
Minimum fluidization 0.097 at 973 K and 0.092 at
Characterization of Partially Burned Char
velocity (ms−1 ) 1073 K
The evolution of pore structure of partially burned char Total column height (mm) 700
was characterized by measuring the specific pore sur- Static bed height (mm) 100
face area and the accessible porosity at various carbon Expanded bed height 200
conversion levels. A nitrogen adsorption technique, (mm)
Superficial gas velocity 0.25 at 973 K and 0.23 at
using an automatic sorption analyzer Autosorb-1
(ms−1 ) 1073 K
(model ASIC-9) by Quantachrome Instruments, was

International Journal of Chemical Kinetics DOI 10.1002/kin


306 SADHUKHAN, GUPTA, AND SAHA

evacuated sample tube was cooled to −196◦ C and was RESULTS AND DISCUSSION
then exposed to the analysis gas at a series of pre-
cisely controlled pressures. The pressure at which ad- Internal Structure of Partially Burned Char
sorption equilibrium occurred was measured, and the Samples by SEM Images
universal gas law was applied to determine the quan-
tity of gas adsorbed. The pressure was increased to The structure of partially burned char particles at dif-
the point of bulk condensation of the analysis gas. The ferent carbon conversion is presented in Figs. 2 and 3.
desorption process was then initiated in which pres- The structure of ash (white contrast) and the carbon
sure was reduced gradually. As with the adsorption matrix (gray contrast), which covers the entire char,
process, measurements were carried out of the chang- is clearly visible. Figure 2 (carbon conversion 0.12)
ing quantity of gas on the solid surface. These two shows that the ash and carbon matrix are distributed
sets of data represented the adsorption and desorption randomly. The reactant, O2 , diffuses from the bulk gas
isotherms, and their analysis yielded information about through the gas boundary layer around the particle to
the surface characteristics of the material. Finally, the the particle surface, from which it diffuses through the
specific surface area was determined by application internal pores of the char particle and reacts with car-
of Brunauer–Eemmett–Teller (BET) analysis software bon to form CO/CO2 by Reaction (1). Owing to the
available with the instrument. porous structure of the char the gaseous reactants, O2
A mercury porosimeter (pore-master-GT, model and CO2 , react with carbon throughout the entire vol-
PM 33–6) was used to determine the accessible poros- ume of the char (Figs. 2 and 3), leaving behind the inert
ity of partially burned char. The char sample, con- ash. Figure 3 supports the hypothesis that char samples
tained in a sample holder called pentrometer tube, was in the present investigation follow the volume reaction
weighed and placed into the pentrometer cell attached model where the reactions take place throughout the
with the equipment. The instrument had two built-in, volume of the char particle.
automated low-pressure ports for filling of penetrome- As the carbon reacts during the char combustion,
ter by mercury. The penetrometer tube was connected the micropores are converted to mesopores, which are
automatically with one of the low-pressure port where finally transformed into macropores, leading to pore
the samples were degassed under vacuum. After de- opening phenomena [39]. This causes an increase in
gassing, the penetrometer was automatically filled up effective diffusivity of gaseous components within the
with liquid mercury under pressure, and the analysis pore structure of char particle (Eq. (41)), decreasing
proceeded from 0.1 to 50 psia. The volume of the liq- consequently the mass transport resistance. However,
uid mercury penetrating into the pore volume of char at large carbon conversions, the active surface area
particle was automatically determined, and the pres- per unit volume of char falls steeply (Fig. 4), decreas-
sure vs. intrusion volume graph (adsorption curve) was ing the heterogeneous reaction rates (Eq. (8)) dras-
generated by the instrument. The pressure was then re- tically and the remaining carbon content in the char
leased very slowly and again brought back to 0.1 psia. reacts very slowly. The phenomenon is clearly evident
The liquid mercury was extruded from the pore, and in Fig. 3 (carbon conversion 0.85), where the entire
graph of pressure vs. volume of mercury occupying matrix is occupied by the ash and a few unreacted is-
the pores (desorption curve) was generated again dur- lands of carbon surrounded by a thick layer of ash.
ing the extrusion phase. The porosity of the sample For a char particle of diameter 4.36 mm, at the bulk
was calculated from this intrusion/extrusion data by gas temperature of 973 K, the carbon conversions of
the instrument. 0.12 (Fig. 2), 0.78 and 0.85 (Fig. 3) are achieved after
Partially burned char samples were stored after liq- reaction times of 53, 314, and 425 s, respectively. The
uid N2 quenching, and SEM images were obtained corresponding reaction times for the same conversions
using a scanning electron microscope (JEOL, model at the bulk gas temperature of 1073 K are slightly lower
JSM-5800) for internal structure. The images were an- (Fig. 5).
alyzed for the presence of ash and carbon matrix. Two The experimentally measured internal surface area
SEM images with the char particles with the same car- at different carbon conversion levels for chars are
bon conversion were analyzed for internal structure, presented in Fig. 4 at two experimental fluidized bed
and no significant difference was observed. Images bulk gas temperatures of 973 and 1073 K. Bulk gas
were analyzed for partially burned coal chars at low temperature is the gas-phase temperature in the flu-
and high carbon conversion levels (0.12 and 0.85, re- idized bed, measured by a thermocouple inserted into
spectively), and the results are discussed in the next the bed. The value of the pore parameter  was esti-
section. mated by regression analysis to be 7 at 973 K and 5 at

International Journal of Chemical Kinetics DOI 10.1002/kin


COMBUSTION CHARACTERISTICS OF POROUS COAL CHAR: VOLUME REACTION MODEL 307

Figure 2 Internal structure of coal char particle at carbon conversion 0.12: ash (white); carbon matrix (gray).

Figure 3 Internal structure of coal char particle at carbon conversion 0.85: ash (white); carbon matrix (gray).

1073 K and was incorporated in the char combustion Comparison of the Char Combustion Model
model. The details of the estimation of pore parameter Prediction with Experimental Results
from experimental data were discussed by Sadhukhan
Model simulations were compared with the experi-
et al. [40].
mental results of this study in the fluidized bed and

1.0
30
1073 K
(ψ = 7, 973 K) 0.8
Surface area (m 2 cm )

Carbon conversion

20 0.6
973 K
(ψ = 5, 1073 K)
0.4
10
Experim ental 0.2 Experimental
Random pore m odel Model prediction

0 0.0
0.0 0.2 0.4 0.6 0.8 1.0 0 100 200 300 400 500 600 700 800
Carbon conversion Tim e (s)

Figure 4 Change in specific surface area with carbon con- Figure 5 Comparison of model prediction with experimen-
version: experimentally measured (rs = 2.18 mm, Tso = tally observed carbon conversion for coal char (rs = 2.18
300 K). [Color figure can be viewed in the online issue, mm, Tso = 300 K). [Color figure can be viewed in the online
which is available at www.interscience.wiley.com.] issue, which is available at www.interscience.wiley.com.]

International Journal of Chemical Kinetics DOI 10.1002/kin


308 SADHUKHAN, GUPTA, AND SAHA

2400 0.8 2400 0.8


Expt. tem perature
Expt. temperature
2000

Particle temperature (K)


Particle temperature (K)

Expt. conversion 2000


Expt. conversion

Carbon conversion
0.6 0.6

Carbon conversion
1600 1600

1200 0.4 1200 0.4

800
800
0.2 0.2
400
400
0 0
15 20 30 0 0
0 5 10 25
0 5 10 15 20 25 30
Reactor height (cm )
Reactor height (cm)
Figure 6 Comparison of model prediction with the exper-
Figure 7 Comparison of model prediction with the experi-
imental data of Waters et al. [14] : Run 5.  and : experi-
mental data of Waters et al. [14]: Run 6.  and : experimen-
mental; —: predicted.
tal; —: predicted. [Color figure can be viewed in the online
issue, which is available at www.interscience.wiley.com.]

with those published by Waters et al. [14]. It was


observed visually during the experiments with coal Simulation Study of Coal Char Samples
char (rs = 2.18 mm) that the char particles glowed at
bulk gas temperatures above 950 K, indicating com- A fully transient model was used to predict the combus-
bustion in ignited regime and the combustion rate was tion characteristics of porous coal char under different
significantly high. Hence the carbon conversion pre- process conditions. The variation of the particle center
dicted by the model was compared with the exper- and surface temperature, CO2 fraction at the edge of the
imentally measured values at bulk gas temperatures boundary layer, and carbon combustion rate with con-
of 973 and 1073 K (Fig. 5). Fairly good agreement version was investigated. The effects of various process
is obtained with a root mean square value of relative parameters that predominantly influence the combus-
error of 0.027. The model prediction, based on the tion of char, such as bulk gas temperature, particle
present mathematical analysis, is also compared with size, char structure property like tortuosity factor, etc.
the experimental results of carbon conversion and par- on the combustion of char were evaluated from simu-
ticle temperature for porous spherical carbon particles lation studies, carried out for highly porous char IGT
(spherocarb) with a high degree of sphericity and uni- HT 155 (Institute of Gas Technology (IGT), Chicago,
form composition, as published by Waters et al. [14]. IL). The physical properties of char sample were taken
Combustion of 140-μm spherocarb particles was car- from Dutta and Wen [24].
ried out in an entrained flow reactor at laminar condi-
tions. The experimental results for runs 5 and 6, as re- Effect of Tortuosity Factor. The tortuosity factor typ-
ported by Waters et al. [14], are summarized in Table V. ically depends on the structure of the porous medium,
Figures 6 and 7 show good agreement between the ex- i.e. interconnected spherical pores or tubular structures,
perimental and model predicted results both for car- and influences the diffusion of gaseous reactants into
bon conversion and particle temperature at different the pores of the solid reactant. In the present study, the
reactor collection heights. It may be noted that the tortuosity factor τ was calculated from experimental
model input included the experimental conditions and results of mercury porosimeter for devolatilized char,
physical properties of spherocarb particles from Waters partially burned chars at different carbon conversion
et al. [14]. and fully converted char. τ was found to vary between

Table V Experimental Results for Runs 5 and 6 (Waters et al. [14])


Run 5: O2 = 24%, H2 O = 16%, and CO2 = 2%; Run 6: O2 = 24%, H2 O = 17%, and CO2 = 2%

Particle Bulk Carbon Particle Bulk Carbon


Velocity Temperature Conversion Velocity Temperature Conversion
Height (cm) (cm s−1 ) (K) (%) (cm s−1 ) (K) (%)
12.70 215 1508 8 225 1584 16
19.05 240 1465 24 250 1532 31
25.40 270 1416 40 270 1478 45

International Journal of Chemical Kinetics DOI 10.1002/kin


COMBUSTION CHARACTERISTICS OF POROUS COAL CHAR: VOLUME REACTION MODEL 309

1.5 and 2.9, and an average value of 2.2 was used in 1.0
Bulk gas tem perature (K) 1123
the present study. Patission et al. [32] used a value
0.8
1.4 for the analysis of gas-solid reactions in a porous

CO2/(CO+CO2)
1003
particle, whereas Kassebaum and Chelliah [29] used a Ignition
0.6
value of 2.0 for the analysis of oxidation behavior of an Extinction
isolated porous carbon particle. A sensitivity analysis 0.4

of the tortuosity factor τ on the 50% conversion time 953


0.2
was carried out for rs = 1.5 mm and Tb = 1073 K. For
±20% perturbation from the base value of τ = 2.2, 0.1
923

the deviations in t50 were +7.60% and −7.51% of 0.0 0.2 0.4 0.6 0.8 1.0
Carbon conversion
the unperturbed value, respectively. This confirms the
important role played by the tortuosity factor in the Figure 9 Effect of the bulk gas temperature on CO2 fraction
combustion of char. at the edge of the boundary layer: combustion of IGT char
no. HT 155 (rs = 1.5 mm, Tso = 600 K).

Effect of Bulk Gas Temperature. The effect of bulk


gas temperature on combustion characteristics was
studied with two particle sizes (rs = 1.5 mm and heat loss, and an increase in the particle temperature
37 μm) for two different applications, fluidized bed initiates the homogeneous reaction (Eq. (3)) and forms
combustion and pulverized fuel combustion, respec- a CO flame. The highly exothermic CO oxidation re-
tively. action heats up the particle with the particle center
At low bulk gas temperatures, the center tempera- temperature rising much above the bulk gas tempera-
tures does not exceed the gas bulk temperature much ture (Fig. 8) and increases the CO2 fraction at the edge
during the entire combustion period (Fig. 8) for the of the boundary layer (Fig. 9) as most of the CO pro-
larger particles (rs = 1.5 mm), applicable in fluidized duced by heterogeneous reactions (Eqs. (1) and (2))
bed combustion. CO produced by the heterogeneous is converted to CO2 due to ignition. However, when
reactions does not get ignited by the homogeneous re- the carbon conversion increases beyond 98%, the spe-
action (Eq. (3)) in the gas boundary layer around the cific surface area of the coal char drops drastically
particle, leading to a small mole fraction of CO2 at the (Fig. 4) resulting in a sharp fall in both heterogeneous
edge of the boundary layer (Fig. 9). When there is no and homogeneous reaction rates. It causes an extinc-
CO flame in the gas phase, the combustion reaction is tion of the coal char (Fig. 10), and the absence of a
sustained by the heterogeneous reactions only, and in CO flame leads to a decrease in CO2 fraction at the
the absence of the highly exothermic homogeneous re- edge of the boundary layer (Fig. 9) and the particle
action, the gas-phase and the particle temperature are temperature (Fig. 10). Similar ignition in the combus-
small. Although the overall heterogeneous reactions tion regime and extinction near the end were observed
are exothermic, heating up of the particle is compen- experimentally by Ubhayakar and Williams [41].
sated by the heat loss to the ambience, and the particle Most of the earlier analysis of coal char combustion,
temperature remains reasonably constant. including Sotirchos and Amundson [33], assumed neg-
Above a critical bulk temperature, 995 K, the heat ligible temperature gradient within the particle during
generation due to heterogeneous reactions exceeds the combustion. However, in the present study, the energy

1800 1800

Bulk gas tem perature (K) Bulk gas tem perature (K)
1600 1123 1600
Center temperature (K)

1123
Temperature (K)

1400 1400
1003
1200 1200 1003
953
1000 1000

923 953
800 800 Center
Surface
600 600
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Carbon conversion Carbon conversion

Figure 8 Effect of the bulk gas temperature on the particle Figure 10 Effect of the bulk gas temperature on the center
center temperature: combustion of IGT char no. HT 155 and surface temperature: combustion of IGT char no. HT 155
(rs = 1.5 mm, Tso = 600 K). (rs = 1.5 mm, Tso = 600 K).

International Journal of Chemical Kinetics DOI 10.1002/kin


310 SADHUKHAN, GUPTA, AND SAHA

equation was solved within the particle and Fig. 10 in- 500

dicates that a considerable temperature gradient does


400
exist between the center temperature and the surface

99% Burnout time (s)


temperature of the larger particles (rs = 1.5 mm). At a 300
low bulk temperature, the difference between the cen-
200
ter and surface temperatures is insignificant, as ignition
of the CO flame does not take place at all. At a higher 100
bulk gas temperature, combustion proceeds with CO
ignition, causing a high rate of heat generation. The 0
900 950 1000 1050 1100 1150 1200
surface of the particle, being exposed to the ambient, Bulk gas tem perature (K)
loses heat rapidly by radiation, while the center of
Figure 12 Effect of the bulk gas temperature on the burnout
the particle faces additional conductive heat transfer time: combustion of IGT char no. HT 155 (rs = 1.5 mm,
resistance inside the particle and attains a higher tem- Tso = 600 K).
perature (by as much as 350 K) than at the surface.
Extinction takes place beyond a carbon conversion of
is not much affected by further increase in the bulk gas
0.98 when the CO flame disappears, and because of
temperature.
the absence of highly exothermic homogeneous reac-
The rate of carbon combustion varies significantly
tion (Eq. (3)), both the surface and center temperatures
with conversion for a given bulk gas temperature and
drop to the bulk gas temperature and overall rate of
particle size (Fig. 13). Below a critical bulk gas tem-
combustion diminishes significantly.
perature (995 K) for a particle with a radius of 1.5 mm,
For small particles (rs = 37 μm), applicable in pul-
the combustion rate is very slow and uniform. How-
verized fuel combustion, Fig. 11 indicates that even
ever, above the critical temperature, the combustion
at higher bulk gas temperatures, the particle temper-
proceeds in the ignition regime, where large evolu-
ature hardly exceeds the bulk gas temperature due to
tion of heat shoots up the particle temperature, and the
the absence of CO ignition. This clearly indicates that
carbon combustion rate increases rapidly with conver-
the combustion of very small particles is sustained by
sion. A maxima is reached at carbon conversions of
heterogeneous reactions (1) and (2) only without the
0.20 and 0.42 for bulk gas temperatures of 1123 and
formation of a CO flame within/around the particle due
1023 K, respectively, beyond which the combustion
to very thin gas boundary layer.
rate decreases due to depletion of carbon content of
The burnout time for a particle of radius 1.5 mm at
the particle. It is also evident that beyond a particular
different gas bulk temperatures is presented in Fig. 12.
conversion level, the carbon combustion rate becomes
It shows that a critical gas bulk temperature of 995
independent of the bulk gas temperature.
K exists, below which the CO flame is not formed,
and the burnout time decreases significantly with an
increase in the bulk gas temperature. Above this critical Effect of Particle Radius. For fine particles, the parti-
temperature, ignition of the particle occurs with the cle temperature is nearly the same as the bulk gas tem-
formation of a CO flame and the particle burnout time perature since ignition does not take place. Combustion
of larger particles in an ignited regime is associated
with high particle temperature, and the difference be-
2600 tween the center and the gas bulk temperatures could be
Bulk gas tem perature (K)
2200
Center temperature (K)

0.25
1673
1800 Bulk gas tem perature (K)
Carbon comb. rate (mg s−1)

1123
0.20
1400 1473
0.15
1023
1000
953 0.10

600
0.0 0.2 0.4 0.6 0.8 1.0 0.05 953
Carbon conversion
0.00
Figure 11 Effect of the bulk gas temperature on the particle 0.0 0.2 0.4 0.6 0.8 1.0
center temperature: combustion of IGT char no. HT 155 (rs = Carbon conversion

37 μm, Tso = 600 K). [Color figure can be viewed in the Figure 13 Effect of the bulk temperature on the carbon
online issue, which is available at www.interscience.wiley combustion rate: combustion of IGT char no. HT 155 (rs =
.com.] 1.5 mm, Tso = 600 K).

International Journal of Chemical Kinetics DOI 10.1002/kin


COMBUSTION CHARACTERISTICS OF POROUS COAL CHAR: VOLUME REACTION MODEL 311

140 0.0012
Carbon conversion
120 1023

Carbon comb. rate (mg s )


Bulk gas tem perature (K) 0.1
99% Burnout time (s)

100 0.0009
0.2
80

60 0.0006 0.5

40 0.7
0.9
1123 0.0003
20
0.99
0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 0.0000
Initial particle radius (m m ) 0.0 0.3 0.6 0.9 1.2 1.5
Radial distance (mm)
Figure 14 Effect of the initial particle size on the burnout
time: combustion of IGT char no. HT 155 (rs = 1.5 mm, Figure 16 Variation of the carbon combustion rate with
Tso = 600 K). radial distance: combustion of IGT char no. HT 155 (rs = 1.5
mm, Tb = 973 K). [Color figure can be viewed in the online
issue, which is available at www.interscience.wiley.com.]
as high as 350 K for rs = 1.5 mm. Figure 14 indicates
that the particle burnout time decreases with an de-
crease in the particle size until a critical particle ra- consumed near the surface, the reaction zone gradu-
dius is reached. Below this critical particle radius, the ally recedes toward the center of the particle with time.
burnout time increases with reduction in particle size. However, at low bulk gas temperature without the for-
This transition is attributed to the shift of combus- mation of a CO flame, the reaction rate is much smaller
tion regime from ignition to extinction. The same phe- and the reaction zone extends throughout the volume
nomenon was observed by Sundaresan and Amundson of the particle at all carbon conversion levels (Fig. 16).
[21] for combustion of carbon particles. This is also corroborated by Fig. 17, which shows ra-
dial profiles of the local particle macroporosity at 50%
carbon conversion for different bulk gas temperatures.
Reaction Zone. The reaction zone thickness in a gas-
It may be noted that the macroporosity varies from 0.3
solid noncatalytic reaction depends on the operating
(devolatilized char) to 0.75 (fully converted char). At
temperature, particle size, reaction kinetics, and prop-
high bulk gas temperature (say 1123 K), the macrop-
erties such as effective diffusivity, porosity, and so on.
orosity near the particle surface is as high as 0.75 and
An analysis was made to investigate the effect of bulk
the change in porosity is restricted to a thin layer near
gas temperature and particle size on the heterogeneous
the particle surface. Most of the inner core of the parti-
reaction zone thickness.
cle shows a porosity of around 0.3, indicating no reac-
Figure 15 shows the variation of carbon combus-
tion there, even though the overall carbon conversion
tion rate along the particle radial distance at different
reached 50%. However, at a gas bulk gas temperature
carbon conversion for large particles at a high bulk gas
of 993 K, the macroporosity changes uniformly from
temperature of 1073 K, when combustion takes place
0.45 at particle center to 0.58 at the surface, indicat-
in the ignition regime. Oxygen diffuses into the par-
ing carbon consumption throughout the volume of the
ticle and reacts with carbon, and a reaction zone gets
particle. Similar trends were observed in Figs. 15 and
formed extending upto about 30% of the particle radius
16. The observations are compatible with the fact that
into the particle from the surface. As carbon gets fully

0.030 At carbon conversion 50%


0.8
Carbon comb. rate (mg s )

0.025 0.1
Carbon conversion
Local porosity

0.020 Bulk tem perature (K)


0.2 0.6

0.015 993
0.5
0.010 0.4 1003
1013 1023
1123
0.005 0.7
0.99 0.9
0.000 0.2
0.0 0.3 0.6 0.9 1.2 1.5 0.0 0.3 0.6 0.9 1.2 1.5
Radial distance (mm) Radial distance (mm)

Figure 15 Variation of the carbon combustion rate with Figure 17 Variation of local porosity with radial distance:
radial distance: combustion of IGT char no. HT 155 (rs = 1.5 effect of the bulk temperature (Ts◦ = 600 K). [Color fig-
mm, Tb = 1073 K). [Color figure can be viewed in the online ure can be viewed in the online issue, which is available at
issue, which is available at www.interscience.wiley.com.] www.interscience.wiley.com.]

International Journal of Chemical Kinetics DOI 10.1002/kin


312 SADHUKHAN, GUPTA, AND SAHA

0.08 0.0025 0.30


At carbon conversion 50%
At carbon conversion 50%
0.25
Carbon comb. rate (mg s )

0.0020

Carbon comb. rate (mg s )


6

Oxygen mole fraction


0.06
Particle size (mm)
0.0015 0.20
0.04
0.15 973
0.0010
993 1023
0.02 0.10
0.0005
3
1 0.05 1003 1013 Bulk tem perature (K)
0.00 0.0000 1123
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.00
Normalized radial distance 0.0 0.5 1.0 1.5 2.0 2.5
Figure 18 Effect of the particle size on the carbon combus- Radial distance (mm)
tion rate at radial locations: combustion of IGT char no. HT Figure 20 Variation of oxygen mole fraction with radial
155 (Tb = 1073 K). [Color figure can be viewed in the online distance: effect of the bulk temperature (rs = 1.5 mm, Ts◦ =
issue, which is available at www.interscience.wiley.com.] 600 K). [Color figure can be viewed in the online issue, which
is available at www.interscience.wiley.com.]
at higher bulk gas temperature, the reaction is pore dif-
fusion controlled leading to a thin reaction zone that products within the particle pores are CO, which dif-
gradually recedes from the particle surface to the cen- fuses radially outward and gets burned to CO2 . A max-
ter. At the low bulk gas temperature, the reaction is ima is observed with respect to CO2 . Similar trends for
kinetically controlled and reaction occurs throughout species and temperature were reported by Kassebaum
the volume of the particle. and Chelliah [29] for oxidation of porous coal char par-
The effect of particle radius on the reaction zone is ticle under quasi-steady state condition. For very small
shown in Fig. 18, where the rate of carbon combus- particles or low bulk gas temperatures, all profiles are
tion is plotted against normalized radial distance in the relatively flat and the CO2 is hardly produced in the
particle. The reaction zone occupies the entire particle absence of CO ignition.
volume for small particles (1-mm diameter) even at Oxygen mole fraction profiles (Fig. 20) also sup-
high bulk gas temperature of 1073 K (Fig. 18) due to ports the conclusion that for a given particle size an
low rate of combustion in the absence of ignition. At increase in bulk gas temperature decreases the reaction
the same temperature, larger particles (6- and 3-mm zone thickness. At a bulk gas temperature of 1123 K,
diameter) undergo combustion in ignited regime and the oxygen mole fraction drops to zero within a short
the reaction is limited to a narrow zone. radial distance from the surface indicating a narrow
The above observations are further corroborated reaction zone. At a bulk gas temperature of 973 K,
with radial profiles of the mole fraction of different the oxygen mole fraction is significant even at particle
species (Fig. 19) for a 1.5-mm-radius char particle at a center, indicating that the reaction zone occupied the
bulk temperature of 1073 K when the carbon conver- entire particle volume.
sion is 50%. Oxygen diffuses from the bulk to particle The variation of oxygen mole fraction with normal-
surface and then from surface to the interior and gets ized radial distance is presented in Fig. 21 for different
consumed in the reaction zone. Most of the combustion particle sizes at a carbon conversion of 50% in the char.

0.4 1600 0.30


At carbon conversion 50% At carbon conversion 50%
1500 0.25
Particle diameter (mm)
Species mole fraction

Oxygen mole fraction

0.3
Temperature (K)

Tem perature 1400


0.20
CO
0.2 1300 1
0.15
1200
0.1 0.10
CO2
1100 3
O2
0.05
4
0.0 1000 6
0.0 0.5 1.0 1.5 2.0 2.5 0.00
Radial distance (mm) 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8

Figure 19 Variation of species and temperature with ra- Normalized radial distance

dial distance combustion of IGT char no. HT 155 (rs = Figure 21 Variation of oxygen mole fraction with radial
1.5 mm, Ts◦ = 600 K, Tb = 1073 K). [Color figure can distance: effect of particle size (Tb = 1073 K, Ts◦ = 600 K).
be viewed in the online issue, which is available at [Color figure can be viewed in the online issue, which is
www.interscience.wiley.com.] available at www.interscience.wiley.com.]

International Journal of Chemical Kinetics DOI 10.1002/kin


COMBUSTION CHARACTERISTICS OF POROUS COAL CHAR: VOLUME REACTION MODEL 313

0.30 the burnout time predicted by equimolar analysis is


0.25 Nonequimolar lower. For the smaller particles, the interior temper-
Equimolar ature is comparable similar to that of the bulk, and
0.20
reaction rates are low; thus the effect of Stefan flow is
0.15 not pronounced (Fig. 23) and the two analyses produce
0.10 identical results.
0.05

0.00 CONCLUSION
0 5 10 15 20 25 30 35 40
Tim e (s)

Figure 22 Comparison of equimolar and nonequimolar A generalized volume reaction model was developed to
analysis on the carbon combustion rate: combustion of IGT analyze the transient combustion behavior of a single
char no. HT 155 (rs = 1.5 mm, Ts◦ = 600 K, Tb = 1123 K). spherical porous coal char particle. Both heterogeneous
and homogeneous chemical reactions were considered.
It is evident that at a bulk gas temperature of 1073 K, the No a priori assumptions were made with respect to the
particles above 3-mm diameter undergo combustion location of the combustion of CO. Carbon conversion
with ignition, resulting in a narrow reaction zone thick- was found to be influenced by the evolution of the in-
ness. It is also observed that in the ignition regime the ternal surface area and char structure properties such
normalized reaction zone thickness decreases with an as the tortuosity factor. The model predictions com-
increase in particle size (Fig. 21). This effect become pared favorably with the experimental results from this
insignificant beyond a critical particle size for a given study and Waters et al. [14]. The S-shaped profile for
bulk temperature, when the reaction zone thickness conversion that is often observed experimentally was
is not further affected by an increase in the particle predicted by the model.
size. However, at the bulk gas temperature of 1073 K, The simulations confirmed the presence of bound-
1-mm-diameter particles undergo combustion without ary layer ignition of CO in consistent with the con-
ignition with a slow reaction rate and the reaction zone clusions drawn by Biggs and Agarwal [10]. It was
occupies the entire particle radius. observed that the formation of a CO flame in the bound-
ary layer affected the combustion dynamics, increas-
ing the particle temperature, the fraction of CO2 at the
Effect of Equimolar and Nonequimolar Analysis on
edge (near the bulk) of the gas boundary layer, and the
the Carbon Combustion Rate. The effect of inclusion
combustion rate and reducing the burnout time signif-
of Stefan flow in the analysis is shown in Figs. 22 and
icantly. Fine particles burn without the formation of
23. For the larger particles, combustion takes place
the CO flame, whereas large particles at the high bulk
with ignition of the CO flame, causing higher interior
gas temperature burn with a CO flame and a high parti-
particle temperature. Consequently, the rates of chem-
cle temperature in ignited regime. The critical bulk gas
ical reactions of both heterogeneous surface reactions
temperature and particle radius above which ignition
and homogeneous CO ignition increase. At this higher
takes place were identified. Even combustion in ignited
reaction rate, the nonequimolar flow is prominent.
regime was found to undergo extinction near the end
Figure 16 shows that for the larger particles equimo-
as the specific surface area of the coal char decreased
lar analysis always predicts a higher carbon combus-
drastically. This was in line with experimental obser-
tion rate than nonequimolar analysis. Consequently,
vations of Ubhayakar and Williams [41]. The burnout
time was found to attain a minima with respect to the
Nonequimolar particle radius due to shift in combustion regime from
Equimolar extinction to ignition, which was in conformity with
the observations of Sundaresan and Amundson [21]
for combustion of carbon particles.
Simulation studies exhibited that at higher bulk gas
temperature a thin reaction zone was formed due to
pore diffusion control, which gradually receded from
0.0E+00 the particle surface to the center. At the low bulk gas
0.0 1.0 2.0 3.0 4.0
Time (s) temperature, reaction occurs throughout the volume of
Figure 23 Comparison of equimolar and nonequimolar the particle due to kinetic control. Larger particles at
analysis on the carbon combustion rate: combustion of IGT the high bulk gas temperature are usually associated
char no. HT 155 (rs = 37 μm, Ts◦ = 600 K, Tb = 1123 K). with thinner reaction zone due to the ignition regime.

International Journal of Chemical Kinetics DOI 10.1002/kin


314 SADHUKHAN, GUPTA, AND SAHA

The effect of Stefan flow on combustion was found to Hl heat of reaction, J mol−1
be significant for the larger particles only. ε char macroporosity
εμ char microporosity
NOMENCLATURE εr solid char emissivity
εf solid char porosity at full carbon
ct total concentration of gaseous mixture, conversion
mol m−3 ε0 solid char porosity at zero carbon
ck concentration of gaseous component k, conversion
mol m−3 η mole ratio of product CO/CO2
cp mean heat capacity, J kg−1 K−1 λ thermal conductivity, J m−1 s−1 K−1
dp particle diameter, m μ viscosity of gas mixture, kg m−1 s−1
Dke effective diffusivity of component k ξ local carbon conversion in solid char
through the solid pore, m2 s−1 ρ density, kg m−3
El activation energy, J mol−1 σ Stephen Boltzmann constant,
km mass transfer coefficient, m s−1 J m−2 s−1 K−4
0 0
ks1 , ks2 surface reaction rate constant, τ tortuosity factor of char
reference Table II ψ pore parameter
0
kv3 volume reaction rate constant,
reference Table II
M molecular weight, kg mol−1 Subscripts
mk mass fraction of component k b edge of boundary layer
Nk molar flux of component k, c carbon
mol m−2 s−1 e effective physical property
Nt total molar flux of gas mixture, g gas
mol m−2 s−1 k index for gaseous component
pCO2 partial pressure of CO2 , atm l reaction number
pO2 partial pressure of oxygen, atm o bulk condition
r radial distance from axis of s solid phase
symmetry, m
R universal gas constant, J mol−1 K−1
Rs1 reaction rate for oxygen consumption, Superscript
mol m−2 s−1 0 initial condition
Rs2 reaction rate for CO2 consumption,
mol m−2 s−1

Rv3 , Rv3 reaction rate for CO consumption,
BIBLIOGRAPHY
mol m−3 s−1
S specific pore surface area, m2 kg−1
1. Gupta, P.; Saha, R. K. Int J Chem Kinet 2004, 36(1),
Sp , Sc terms for Taylor series approximation 1–11.
t time, s 2. Gupta, P.; Saha, R. K. Can J Chem Eng 2004, 82(5),
T temperature, K 1096–1103.
V volume of solid char particle, m3 3. Gupta, P.; Sadhukhan, A. K.; Saha, R. K. Int J Chem
W instantaneous mass of solid char, kg Kinet 2007, 39(6), 307–319.
Wash mass of ash in char sample, kg 4. Sadhukhan, A. K.; Gupta, P.; Saha, R. K. Int J Chem
Wc instantaneous mass concentration of Kinet 2008, 40(9), 569–582.
carbon in char, kg m−3 5. Gupta, P.; Saha, R. K. In Proceedings of Chemical
xk mole fraction of component k Engineering Congress (CHEMCON), Chandigarh,
Xash mass fraction of ash in solid char India, 1999.
6. Gupta, P.; Saha, R. K. J Chem Eng Japan 2003, 36(11),
Xc mass fraction of carbon in solid char
1298–1307.
7. Gupta, P.; Saha, R. K. J Chem Eng Japan 2003, 36(11),
1308–1317.
Greek characters
8. Ballal, G.; Li, C.; Glowinski, R.; Amundson, N. R.
α, β empirical constants Comput Methods Appl Mech Eng 1989, 75(1–3), 467–
γkl reaction stoichiometry for component k, 479.
in reaction no. (l) 9. Linjewile, T. M.; Agarwal, P. K. Fuel 1995, 74, 5–15.

International Journal of Chemical Kinetics DOI 10.1002/kin


COMBUSTION CHARACTERISTICS OF POROUS COAL CHAR: VOLUME REACTION MODEL 315

10. Biggs, M. J.; Agarwal, P. K. Chem Eng Sci 1997, 52, 26. Jackson, R. Transport in Porous Catalyst; Elsevier:
941–952. Amsterdam, 1977.
11. Stubington, J. F. Chem Eng Res Des 1985, 63, 241–249. 27. Shah, N.; Ottino, J. M. Chem Eng Sci 1987, 42, 63–72.
12. Feng, B.; Bhatia, S. K. Energy Fuel 2000, 14, 297–307. 28. Bhatia, S. K.; Perlmutter, D. D. AIChE J 1980, 26, 379–
13. Wang, F. Y.; Bhatia, S. K. Chem Eng Sci 2001, 56, 386.
3683–3697. 29. Kassebaum, J. L.; Chelliah, H. K. Combust Theory
14. Waters, J. B.; Mitchel, R. E.; Squires, R. G.; Laurendeau, Model 2009, 13(1), 143–166.
N. M. Combust Flame 1988, 74(1), 91–106. 30. Bird, R. B.; Stewart, W. E.; Lightfoot, E. N. Transport
15. Smith, I. W.; Tyler, R. J. Fuel 1972, 51(4), 312–321. Phenomena; Wiley: New York, 1960; p. 639.
16. Bar-Ziv, E.; Kantorovich, I. Appl Therm Eng 1998, 18, 31. Rafsanjani, H. H.; Jamshidi, E.; Rostam-Abadi,
991–1003. M. Carbon 2002, 40, 1167.
17. D’Amore, M.; Tognotti, L.; Sarofim, A. F. Combust 32. Patisson, F.; Francois, M. G.; Ablitzer, D. Chem Eng Sci
Flame 1993, 95(4), 374–382. 1998, 53, 697.
18. Hurt, R.; Sun, J. K.; Lunden, M. Combust Flame 1998, 33. Sotirchos, S. V.; Amundson, N. R. AIChE J 1984, 30
113, 181–197. (4), 537–549.
19. Radovic, L. R.; Walker, J. P. L.; Jenkins, R. G. Fuel 34. Carniglia, S. C. J Catal 1986, 102, 401–418.
1983, 62, 849–856. 35. Perry, R. H.; Green, D. W. Chemical Engineers’
20. Krishna, R.; Wesselingh, J. A. Chem Eng Sci 1997, 52, Handbook, 6th edn.; McGraw-Hill: New York, 1984;
861–911. pp. 3–279.
21. Sundaresan, S.; Amundson, N. R. Ind Eng Chem Fun- 36. Palchonok, G. I. Ph.D. Thesis, Chalmers University of
dam 1980, 19(4), 351. Technology, Göteborg, Sweden, 1998.
22. Smith, I. W. Combust Flame 1971, 17, 303–314. 37. Patankar, S. V. Numerical Heat Transfer and Fluid Flow;
23. Tognotti, L.; Longwell, J. P.; Sarofim, A. F. In 23rd Hemisphere: New York, 1980.
International Symposium on Combustion, Orleans, 38. Pabis, S.; Jayas, D. S.; Cenkowski, S. Grain Drying:
France, 1990, p. 1207. Theory and Practice; Wiley: New York, 1998; p. 149.
24. Dutta, S.; Wen, C. Y. Ind Eng Chem Process Des Dev 39. Mahamud, M. M. Appl Surf Sci 2007, 253, 6019–6031.
1977, 16(1), 31–38. 40. Sadhukhan, A. K.; Gupta, P.; Saha, R. K. Fuel Process
25. Howard, J. B.; Williams, G. C.; Fine, D. H. In 14th Inter- Technol 2009, 90(5), 692–700.
national Symposium on Combustion; The Combustion 41. Ubhayakar, S. K.; Williams, F. A. J Electrochem Soc
Institute: Pittsburgh, PA, 1973; pp. 975–986. 1976, 123(5), 747–755.

International Journal of Chemical Kinetics DOI 10.1002/kin

Anda mungkin juga menyukai