Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Modelling of combustion and calcination in a cement precalciner

D.K. Fidaros, C.A. Baxevanou, C.D. Dritselis, N.S. Vlachos∗


Department of Mechanical & Industrial Engineering
University of Thessaly
Athens Avenue, 38334 Volos, Greece

Abstract
A numerical study is presented of the reacting flow and transport processes taking place in a calciner for cement
production. The CFD model is based on the solution of the Navier-Stokes equations for the gas flow and on
Lagrangean dynamics for the discrete particles with turbulence incorporated via the k-ε model. Distributions of
velocities, temperatures and concentrations of the reactants and products as well as the trajectories of particles and
their interaction with the gases are presented, allowing assessment of a particular calciner design to be made.
Key Words: CFD, coal combustion, calcination, calciner modeling, cement production

1. Introduction
Cement production involves raw-mix preheating and
calcination, clinker formation and cooling to achieve the
required crystallographic structure. After cooling, the
clinker is mixed with plaster and other additives in
grinding mills that consume a very large amount of
energy. The raw-mix consists mainly of pulverized
calcium carbonate and silicon dioxide. During its
preheating the moisture evaporates and at 890oC the
endothermous calcination reaction begins, where CaCO3
is converted into CaO and CO2. The required activation
energy is provided by the fuel heat of combustion.
Dry heating of raw-mix in vertical suspension
preheaters is mostly used, where some calcination also Fig. 1 Schematic of the calciner
takes place. Modern cement plants use an additional
device, in which the raw-mix undergoes calcination to a (<1400 °C ) in the kiln reduces the production of Nox,
high level (90-95%), Fig.1. Thus, the calcined raw-mix e) The lower temperatures in the calciner allow the use
enters the rotary kiln at a higher temperature, reducing of fuels with relatively low heat value, and f) The
its energy demand. The placement of calcination outside reduced kiln thermal load decreases the condensing of
the cement kiln results in better quality of CaO and vapours (SO3, Na, K and Cl). Some disadvantages are:
energy savings. For example, in the Olympus plant of a) The lower temperatures of the exhaust gases may
AGET Hercules in Greece calcination takes roughly cause condensation of volatile alkalis, and b) The use of
60% of the total heat, while 35% is spent for preheating low heat fuels requires particular attention to avoid
and 5% for clinkering [1]. This 60:40 ratio is reversed undesirable emissions of polluting and eroding gases.
when calcination takes place inside the rotary kiln. It is important to control calcination as it affects
After heated to appropriate temperature, the raw- fuel consumption, emission of pollutants and the final
mix enters the calciner with the fuel and hot combustion quality of cement. Using CFD, calciners have been
air. The combustion heat causes calcination according to studied at different geometries and operating conditions.
the reaction: For example, Hu et al. [3] used a 3D Eulerian-
CaCO3 ⎯1160
⎯⎯ K
→ CaO + CO 2 + 178 KJ / mol Lagrangean model in a dual combustor and precalciner
to predict the burnout and decomposition ratio during
The high fineness of raw-mix and good turbulent simultaneous injection of two types of coal and raw
mixing produce faster coal combustion and raw-mix material. Iliuta et al. [4] investigated the influence of
calcination with good efficiency at lower temperatures. operating conditions on the level of calcination, burnout
Among the advantages of calcination devices (see and NOx emissions of an in-line low NOx calciner.
Fidaros et al [2]) are: a) The addition of a burner in the
calciner increases the kiln capacity, b) The reduction of 2. Specific Objectives
thermal load of the kiln extends its operational life, c) In the present work a numerical model and a
The reduction of energy demand and the minimal parametric study is presented for the flow and transport
calcination in the kiln reduce its exhaust gases and heat, processes taking place in an industrial calciner. The
losses, d) The combustion at medium–low temperatures model is based on the solution of the Navier–Stokes


Corresponding author: vlachos@mie.uth.gr
Proceedings of the European Combustion Meeting 2007
equations for the gas flow and on Lagrangean dynamics where, Dim is the diffusion coefficient of the oxidant, mo
for the discrete particles, using a commercial CFD code the local fraction of gas oxidant, ρg the gas density, and
(Fluent). All the flow, heat and mass transfer and the Sb the stoichiometric factor of the reaction. Equation (6)
chemical reaction models are briefly described. proposed by Baum & Street [7] ignores the contribution
Turbulence is simulated by the k–ε model. The of kinetics in the surface reaction. The diffusion model
performance of a specific cement calciner is studied considers a constant particle diameter, thus, resulting in
using two different quality coal fuels. Limited available a more porous particle (char) as its mass decreases.
measurements from the Olympus cement plant of AGET The kinetic/diffusion model considers that the
Hercules were used to verify the model. surface reaction advances with a rate that is determined
by the diffusion (R1) of gas oxidant into the particle
3. Mathematical models surface or by the reaction kinetics (R2) [8]:
Using a Eulerian frame and cylindrical coordinates, [(
R 1 =C 1
) ]
T p + T∞ / 2 0.75 ⎛
and R 2 = C 2 exp ⎜ −
E ⎞⎟
(7)
the time-averaged transport equations of momentum, ⎜ ⎟
heat and mass of the gases may be expressed as follows: Dp ⎝ R Tp ⎠
∂ ∂ 1∂ 1 ∂ W ∂Φ The rates R1 and R2 are then combined to give the
( ρΦ ) + ( ρUΦ ) + ( ρrVΦ ) + (ρ )=
∂t ∂x r ∂r r ∂θ r ∂θ (2) combustion rate of coal (char):
∂ ∂Φ 1∂ ∂Φ 1 ∂ ∂Φ
(Γ Φ )+ (Γ Φ r )+ (Γ Φ )+ S Φ dm p R R
∂x ∂x r ∂r ∂r r ∂θ ∂θ = − π D p 2 P0 1 2 (8)
where, U, V, W are the time-averaged velocities in the dt R1 + R2
axial, radial and circumferential direction, respectively where, P0 is the partial pressure of the oxidant.
and Φ represents a time-averaged quantity such as Particle heat radiation is calculated using the P-1
velocity, enthalpy, turbulent kinetic energy, turbulent model [9]. For a grey, absorbent, luminous, scattering
dissipation rate, etc. medium containing absorbing, luminous and scattering
The trajectories of particles are calculated from particles, the incident radiation is given from [6,10-18]:
their corresponding motion equations: ⎡ σΤ 4 ⎤
dUp (ρ p ρ ) + F ∇ ⋅ (Γ ∇G ) + 4π ⎢α [ ]
+ Εp⎥ − a + ap G = 0 (9)
= FD (U − Up ) + g i i (3) ⎣ π ⎦
dt ρp
where, ΕP is the equivalent particle brightness, and αP
where, U, UP the gas and particle velocity, respectively. the equivalent absorption factor, α is the absorption
The coefficient, FD, in the drag force is calculated from factor of gas, σP is the equivalent scattering factor of the
the relationship proposed by Morsi and Alexander [5]. particle, and Γ is the transport coefficient.
Devolatilization kinetics of two competitive rates, Then, the source term for the energy equation of the
are incorporated via the Kobayashi model [6]: gaseous phase is:
mv (t )
t ⎛t ⎞ ⎡ σΤ 4 ⎤
m po - mash
= (a1 R1 + a 2 R2 )exp⎜ (R1 + R2 ) dt ⎟ dt (4)
∫ ⎜ ∫ ⎟
− ∇q r = −4π ⎢α [ ]
+ Εp ⎥ + a + ap G (10)
0 ⎝0 ⎠ ⎣ π ⎦
where, R1 and R2 are the competitive rates controlling The modelling of the mixture fraction is based on
the volatilization at two different temperature ranges, the probability density function (mixture fraction/PDF
mv(t) is the sum of volatiles evaporated up to time t, mpo approach) requiring solution of transport equations for
the initial particle mass, and mash the mass of ash in the two conservative scalar properties. For a binary system
particle. The devolatilization coefficients are set to (fuel-oxidant), the mixture fraction can be expressed in
a1=0.3 (slow reaction) and a2=1.0 (fast reaction). terms of element mass fractions:
The particle heat balance during devolatilization Z − Z kO
f = k (11)
includes contributions by convection, radiation and heat Z kF − Z kO
consumed for devolatilization according to the equation: where Zk is the instant mass fraction, ZkO the mass
dm p fraction of oxidant and ZkF the mass fraction of fuel.
h Ap T∞ + h fg + Ap ε p σ Θ4R
T p ( t + Δt ) = dt + The value of f at each computational point is
h Ap + Ap ε p σ T p3 calculated from the following transport equation:
(5)


⎜ T p (t ) −
h Ap T∞ +
dm p
dt
h fg + Ap ε p σ Θ4R
⎞ Α p (h +ε p σ T p3 )
⎟ −
⎟e mp Cp
t

∂t
( )
ρf +

∂xi
( )
ρu i f =
∂ ⎛ μt ∂ f ⎞
⎜ ⎟ + Sm
∂xi ⎜⎝ σ t ∂xi ⎟⎠
(12)
⎜ h Ap + Ap ε p σ T p3 ⎟
⎜ ⎟ where Sm represents the particle mass transfer to the
⎝ ⎠
gaseous phase.
After the completion of devolatilization, there starts
A conservation equation for the mixture fraction
the surface chemical reaction of the fuel part of the coal
particle. The diffusion model considers that the surface variance, f ′2 (describing the chemistry-turbulence
reaction advances at a rate that is determined by the interaction) is solved simultaneously with equation (12):
diffusion of the gas oxidant into the particle surface:
dm p mo T p ρ g
= −4 π Dp Dim (6)
dt S b T p + T∞( )

2

∂t
( )
ρ f ′2 +

∂x i
( )
ρu i f ′ 2 =
at 30° to the horizontal. Combustion and calcination
take place in the main volume (850 m3) of the device.
2
(13) Coal enters at 2.4 m height from the start of the cone
∂ ⎛ μt ∂ f ⎞ ⎛∂f ⎞ ε
⎜ ⎟ ⎜ ⎟ − C d ρ f ′2 and at 2.68 m from the calciner axis.
∂x i ⎜ σ ∂x ⎟ + C g μ t ⎜ ∂x ⎟ k
⎝ t i⎠ ⎝ i ⎠ The computational domain consists of a hybrid
where, σt, Cg and Cd are constants[7-8]. mesh of 67.104 cells. Due to symmetry, the calculations
were carried out for one half of the domain. Two
4. Chemical reaction mechanisms different quality coals were considered with a total rate
Coal combustion of mass (raw-mix, coal and air) fed into the three kinds
Thermal fragmentation (pyrolysis) is the most of inlets of approximately 100 kg/s. As shown in Table
important physico-chemical change in the coal particle 1, the largest mass rate is that of CaCO3, followed by
during heating at high temperatures. During this stage a tertiary air, coal and finally the coal feeding air.
significant loss of mass occurs, because of dissolution Table 1 Mass flow rates at the calciner inlets
of volatile matter, the quantity and composition of
Case 1 [ Coal I ] Case 2 [ Coal II ]
which depend on coal properties, its grain size and Component
[Kg/s] [%] [Kg/s] [%]
temperature. A number of parallel reactions occur, with
chemical combinations of reacting components or even mCaCO3 53.60 54.81 57.20 54.46
species such as, for example, CH4, CHOH, C2H6, H2, Coal 3.86 3.95 3.78 3.73
and S2. After devolatization leading to production of mTertiary Air 39.36 40.25 39.36 38.85
water vapour, CO, CO2 etc, a series of progressive
mAir Coal 0.972 0.99 0.972 0.96
reactions of char and devolatization gases take place.
Based on a comparative analysis of existing data Total 97.79 100.00 101.98 100.00
[10-13,15] for coal combustion and on the particle
The raw-mix had a Rossin–Rammler distribution
characteristics (average char diameter <<100 μm), the
with an average size of d=16.6 μm and a spread
kinetic/limited diffusion rate model was selected. This is
parameter of n=0.822, while the coal had d=34.5 μm
similar to that of a shrinking-reactant particle core
and n=1.248. The proximate analysis of the two types of
adopted in the general theory of surface heterogeneous
coal particles is given in Table 2. The tertiary air
chemical reaction. The diffusion coefficient Dim of
entered with a velocity 24 m/s, coal with 11.5 m/s and
oxidant in the porous char used in the present model
the raw-mix with 1.5 m/s. The coal was fed
was 5.0x10–5 m2/s [2,19].
pneumatically while the raw-mix entered by gravity. All
Calcination
geometric data and initial and boundary conditions
Limestone calcination undergoes several stages at
corresponded to the Olympus plant of AGET Hercules.
different rates: a) Heat transfer from the gases to the
particle surface and thereof to the reaction interface, b) Table 2 Ultimate analyses for the solid feeds
thermal decomposition of CaCO3 in the reaction zone,
c) transfer of CO2 from the reaction zone to the gases. Case 1 [ Coal I ] Case 2 [ Coal II ]
For small particles moving in high temperature gases, Component As used Dry basis As used Dry basis
the internal and external heat and mass transfer rates are (%) (%) (%) (%)
high. For particles with diameters 1 to 90 μm and gas Humidity 1.45 - 1.45 -
temperatures 748 to 1273 K, the calcination is Volatiles 27.70 28.11 13.21 13.40
chemically controlled and its rate is proportional to the
Ash 12.05 12.23 0.14 0.14
particle surface area as determined by the BET method
(nitrogen absorption at 77K). Because, the limestone Fix Carbon 55.80 56.62 81.70 82.90
microstructure is not completely crystalic and has a Total 97.00 96.96 96.50 96.45
diverse form of porosity, the surface determined by the The low calorific value of Coal I (low quality) was
BET method is the sum of the porous surfaces accessed 26.3 MJ/kg and that of Coal II (high quality) 27 MJ/kg.
by nitrogen. Under such conditions, calcination happens
on the total available surface, giving pseudo-volumetric 6. Results and discussion
characteristics to the reaction [20-24].
Case I (Low quality coal)
Fig. 2 shows higher temperatures in the opposite
5. Geometric and computational details
sides of the raw-mix entries, which is mainly due to the
The modelled calciner consists of a cylinder with
trapping of small coal particles while the CaCO3 particle
6.6 m diameter and 20 m height and two (upper-lower)
concentration is low. The calciner is at temperatures
conical sections having three kinds of inlets at the
little above the threshold for calcination, so that
bottom part and an outlet at the top, from where the
calcification takes place almost in the whole device.
products such as calcined raw-mix, CO2, and other
There are regions of low temperature because of intense
gases exit. Raw-mix is fed via two pipes inclined at 60°
calcination, resulting to high heat absorption. The high
to the horizontal. The tertiary air enters axially from the
temperatures in high velocity regions are mainly due to
bottom via a concentric 2.6 m diameter duct and the
the high concentration of burning coal and to the
coal is fed at the lower conical part via two 0.2 m pipes
absence of CaCO3 particles.

3
Case II (Good quality coal)
Fig. 5 shows the velocity magnitude distribution of
the gaseous phase for Case II (good quality coal) in two
vertical diametral planes normal to each other. The
gases undergo a high deceleration at the lower conical
part due to the increasing cross sectional area and to the
entering coal and raw-mix materials. In the main
cylindrical part, the velocity remains at 7 to 8 m/s, with
regions of higher velocity in front of the two raw-mix
inlets and in the upper conical part. At the exit, a region
with higher velocities is observed.

Fig. 2 Temperature distribution for Case I


The trajectories and residence times of coal and
CaCO3 particles are presented in Fig. 3. The average
trajectory length of all particles (~8000) is about 48 m,
while the longest exceeds 70 m. The particle mean
residence time is about 10s while the longer is 16.3 s.

Fig. 5 Distribution of gas velocities (Case II)


The trajectories of coal particles are presented in
Fig. 6 and 7 (detail). Figs. 6 and 7 also show the parts of
the trajectories where the coal particles are activated
thermochemically. Thus, the blue colour in the start of
the trajectory corresponds to coal warming up, the green
and yellow to evaporation and combustion of volatiles,
respectively, and the red to the combustion of fixed
carbon. Finally, the blue colour in the end of the
Fig. 3 Particle trajectories and residence times (Case I) trajectory shows the cooling of the ash by the gases. The
Fig. 4 shows the parts of the trajectories where the average passage length of coal particles is 32 m, while
CaCO3 particles are activated thermochemically. Thus, the longest exceeds 45 m. Their average residence time
the blue colour in the start of the trajectory corresponds is 5 s, mainly because the air moves faster sweeping the
to CaCO3 warming up and the red to calcination. smaller coal particles.
Finally, the blue colour in the end of the trajectory
shows the cooling of the inactive remainder. Calcination
is noticeable in the main cylindrical part of the device.
The predicted calcination is 91.9% for Case I and occurs
in all the active height of the calciner.

Fig. 6 Coal particle trajectories (Case II)

From the details of the trajectories in Fig. 7 it is


evident that the coal particles of Case II react with small
hysteresis after their entrance in the calciner.
Fig. 4 CaCO3 particle trajectories (Case I)

4
7. Conclusions
A CFD model was described for the gaseous flow
and transport phenomena and for the trajectories of
particles in a calciner for cement production. Submodels
for radiation, chemical reactions and turbulence effects
were also included.
The small recirculation region observed near the
inlets of raw-mix and tertiary air increases the active
length of the device and the particle residence time. The
rapid calcination near the raw-mix inlet produces high
local CO2 concentrations, which limit calcination.
Higher temperatures are observed near the coal
inlet where combustion of volatiles occurs. The high
Fig. 7 Coal trajectories and devolatilization (Case II) temperature regions observed along the calciner are due
to coal particles with orbits that are not intermingled
The trajectories of CaCO3 particles are presented in with the raw-mix, suggesting that more attention should
Fig. 8. The average particle residence time is 10 s and be paid to its granulometry than to that of coal.
the longest 15 s. Calcination is noticeable in the semi- The high unevenness of the used Rossin–Rammler
cylinder defined by the feeding pipes of the raw-mix particle distribution is mainly responsible for the small
and the coal inlets, which agrees with the temperature percentage of the non-calcinated CaCO3. Despite the
field and the CO2 concentration (not shown here). The energy surplus (especially in Case II), calcination could
calcination for Case II reaches 96.5% and occurs also in not be completed because of the large amounts of CO2
all the active calciner height. released immediately after the raw-mix entries.
The upper conical part of the calciner causes an
acceleration of gases and particles, reducing calcination.
The observed high temperatures can cause thermal
stress problems leading to erosive by-products such as
SiO2 chemical activation, NOx formation. Also, small
recirculations are formed acting as a bumper wall for
the particles, and resulting to local high temperatures
and wall erosion.
The present model predicted the range of velocities
and gas temperatures and the asymmetries in the exit of
the calciner as observed in the Olympus cement plant of
AGET Hercules.

Acknowledgements
Fig. 8 CaCO3 particle trajectories (Case II) The support of the Greek General Secretariat for
R&T and AGET Hercules is gratefully acknowledged.
Finally, Fig 9 shows the evolution of calcination for the The authors thank Dr. I. Marinos and Mr. T. Pissias for
two fuel cases. It is evident from the figure that Case II the plant data and for useful discussions.
(good quality fuel) gives a higher level of calcination
(96.5%) compared to the calcination level of 91.9% for References
Case I (low quality fuel). [1] E. Kolyfetis, C.G. Vagenas, Mathematical modeling
of separate line precalciner, ZKG Intl. 2 (1988)
559-563.
[2] D.K. Fidaros, C.A. Baxevanou, C.D. Dritselis, N.S.
Vlachos, ‘Numerical modelling of flow and
transport processes in a calciner for cement
production’, Powder Technology 171 (2007) 81-95
[3] Z. Hu, J. Lu, L. Huang, S. Wang, Numerical
simulation study on gas-solid two-phase flow in
pre-calciner, Communications in Nonlinear Science
& Numerical Simulation 11 (2006) 440-451.
[4] I. Iliuta, K. Dam-Johansen, A. Jensen, L.S. Jensen,
Modelling of in-line-low-NOx calciners – a
Fig. 9 Evolution of calcination for the two coal types parametric study, Chemical Engineering Science 57
(2002) 789-803.
[5] S.A. Morsi, A.J. Alexander, An investigation of
particle trajectories in two-phase flow systems, J.

5
Fluid Mechanics 55 (1972) 193-208. [22] B.V.L’vov, Mechanism of thermal decomposition
[6] H. Kobayashi, J.B. Howard, A.F. Sarofim, Coal of alkaline-earth carbonates, Thermochimica Acta
devolatilization at high temperatures, Proc. 16th Intl 303 (1997) 161-170.
Symposium on Combustion. The Combustion [23] A.K. Galwey, M.E. Brown, Arrhenius parameters
Institute 1976, 411-425. and compensation behavior in solid-state
[7] M.M. Baum, P.J. Street, Predicting the combustion decompositions, Thermochimica Acta 300 (1997)
behavior of coal particles, Combustion Science & 107-115.
Technology 3 (1971) 231-243. [24] J. Khinast, G.F. Krammer, C. Brunner, G.
[8] M.A. Field, Rate of combustion of size-graded Staudinger, Decomposition of limestone: the
fractions of char from a low rank coal between influence of CO2 particle size on the reaction rate,
1200K-2000K, Combustion & Flame 13 (1969) Chemical Engineering Science 51 (1996) 623-634.
237-252.
[9] S.S. Sazhin, E.M. Sazhina, O. Faltsi-Saravelou, P.
Wild, The P-1 model for thermal radiation transfer:
advantages and limitations, Fuel 75 (1996) 289-
294.
[10] J.K. Fink, Pyrolysis and combustion of polymer
wastes in combination with metallurgical processes
and the cement industry, J. of Analytical and
Applied Pyrolysis 51 (1999) 239-252.
[11] P. Basu, Combustion of coal in circulating
fluidized-bed boilers: a review, Chemical
Engineering Science 54 (1999) 5547-5557.
[12] F.C. Lockwood, T. Mahmud, M.A. Yehia,
Simulation of pulverized coal test furnace
performance, Fuel 77 (1998) 1329-1337.
[13] R.G. Bautista-Margulis, R.G. Siddall, L.Y.
Manzanares-Papayanopoulos, Combustion modeling
of coal volatiles in the freeboard of a calorimetric
bed combustor, Fuel 75 (1996) 1737-1742.
[14] C.A.G Veras, J. Saastamoinen, J.A. Carvalho, N.
Aho, Overlapping of the devolatilization and char
combustion stages in the burning of the coal
particles, Combustion & Flame 116 (1999) 567-
579.
[15] W.B. Fu, B.L. Zhang, S.M. Zheng, A relationship
between the kinetic parameters of char combustion
and the coal's properties, Combustion & Flame 109
(1997) 587-598.
[16] J. Zhang, S. Nieh, Comprehensive modeling of
pulverized coal combustion in a vortex combustor,
Fuel 76 (1997) 123-131.
[17] L. Reh, Challenges of circulating fluid-bed reactors
in energy and raw materials industries, Chemical
Engineering Science 54 (1999) 5359-5368.
[18] H. Wang, J.N. Harb, Modeling of ash deposition in
large-scale combustion facilities burning pulverized
coal, Progress in Energy and Combustion Science
23 (1997) 267-282.
[19] D.K. Fidaros, C.D. Dritselis, N.S. Vlachos
‘Aspects of numerical modeling of chemically
reacting flows in cement calciners’, Proc. 3rd
Meeting of the Greek Section of the Combustion
Institute, Patra, Nov. 2003.
[20] R.H. Borgwardt, Calcination kinetics and surface
area of dispersed limestone particles, AIChE J. 31
(1985) 103-111.
[21] F. Acke, I. Panas, Activation energy of calcination
by means of a temperature programmed reaction
technique, Thermochimica Acta 306 (1997) 73-76

You might also like