Cement Model

download Cement Model

of 15

Transcript of Cement Model

  • 7/27/2019 Cement Model

    1/15

    Numerical modelling of flow and transport processesin a calciner for cement production

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

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

    Received 19 October 2005; received in revised form 1 September 2006; accepted 7 September 2006

    Available online 29 November 2006

    Abstract

    Controlling the calcination process in industrial cement kilns is of particular importance because it affects fuel consumption, pollutant emission

    and the final cement quality. Therefore, understanding the mechanisms of flow and transport phenomena in the calciner is important for efficient

    cement production. The main physico-chemical processes taking place in the calciner are coal combustion and the strongly endothermic

    calcination reaction of the raw materials. In this paper a numerical model and a parametric study are presented of the flow and transport processes

    taking place in an industrial calciner. The numerical model is based on the solution of the NavierStokes equations for the gas flow, and on

    Lagrangean dynamics for the discrete particles. All necessary mathematical models were developed and incorporated into a computational fluid

    dynamics model with the influence of turbulence simulated by a two-equation (k) model. Distributions of fluid velocities, temperatures and

    concentrations of the reactants and products as well as the trajectories of particles and their interaction with the gas phase are calculated. The

    results of the present parametric study allow estimations to be made and conclusions to be drawn that help in the optimization of a given calciner.

    2006 Elsevier B.V. All rights reserved.

    Keywords: CFD; Coal combustion; Calcination; Calciner modeling; Cement production

    1. Introduction

    The main processes of cement production include raw-mix

    preheating and calcination, clinker formation and cooling to

    achieve a crystalographic structure that meets the required ce-

    ment specifications. After cooling, the clinker is fed into grind-

    ing or finish mills and is mixed with plaster and ameliorating

    additives. The mills consume a very large amount of the total

    energy required for cement production.

    The raw-mix consists mainly of pulverized calcium carbon-ate and silicon dioxide. During its heating/drying at tempera-

    tures from 100 C to 500 C the moisture evaporates and at 850

    to 890 C the endothermous calcination reaction begins, where

    CaCO3 is converted into CaO and CO2. The activation energy

    for the calcination is provided by the combustion heat of the

    fuel.

    Dry heating of raw-mix in vertical suspension preheaters (see

    Fig. 1) is mostly used, where calcination also takes place. The

    innovation in the entire pyroprocess in modern cement plants is

    the use of an additional calcining vessel, in which the raw-mix

    undergoes calcination to a level of 90 to 95%. In this way, the

    calcined raw-mix enters the rotary kiln at a higher temperature,

    thus reducing the energy demand and the thermal load on the

    kiln. After being heated to the appropriate temperature, it enters

    the calciner together with the fuel and the hot tertiary air, Fig. 2.

    The combustion heat released by the fuel causes calcination ofthe raw-mix according to the chemical reaction:

    CaCO3 Y1160 K

    CaO CO2 178 kJ=mol 1

    The high fineness of the raw-mix and the good turbulent

    mixing cause uniform and fast coal combustion and calcination

    reactions. The products of the calciner are fed to the last cyclone

    that feeds the rotary kiln. The placement of calcination outside

    the cement kiln results in better quality of CaO and energy

    savings. For example, in the Olympus plant of AGET Hercules

    in Greece calcination takes roughly 60% of the total heat

    Powder Technology 171 (2007) 8195

    www.elsevier.com/locate/powtec

    Dedicated to the late Professor Shao-Lee Soo, for his pioneering work in

    multiphase dynamics. Corresponding author. Tel.: +30 2421074094; fax: +30 2421074085.

    E-mail address: [email protected] (N.S. Vlachos).

    0032-5910/$ - see front matter 2006 Elsevier B.V. All rights reserved.doi:10.1016/j.powtec.2006.09.011

    mailto:[email protected]://dx.doi.org/10.1016/j.powtec.2006.09.011http://dx.doi.org/10.1016/j.powtec.2006.09.011mailto:[email protected]
  • 7/27/2019 Cement Model

    2/15

    absorbed in the system, while 35% is spent for preheating and

    5% for clinkering [1]. This ratio of 60:40 is reversed in the case

    where the calcination is taking place inside the rotary kiln. In

    addition, the good mixing of fuel, air and raw-mix in the calciner

    results to faster calcination with good efficiency at relatively low

    temperatures.The advantages of using calcination devices are: a) The

    addition of a burner in the calciner increases the capacity of the

    rotary kiln in comparison to using simple preheaters, b) The

    reduction of thermal load and the increased rotational speed of

    the kiln (to achieve better mixing at increased capacity) extends

    the lifetime of the firebricks and, thus, the operational life of the

    kiln, c) The reduction of energy demand and the minimal calci-

    nation in the kiln reduce considerably the exhaust gases and the

    kiln heat losses to the environment because the exhaust gases

    absorb most of the radiation, d) The combustion at mediumlow

    temperatures (b1400 C) in the kiln reduces the production of

    NOx, although combustion control and kiln burner design is stillsignificant, e) The lower temperature required in the calciner

    allows the use of fuels with relatively low thermal capacity

    (usually bituminous coal), f) The reduction of the thermal load of

    the rotary kiln decreases the condensing of vapours (SO3, Na, K

    and Cl) in the combustion area. However, the volatile cycle is

    still a concern because now it will take place in the preheater/

    precalciner tower itself as opposed to the kiln), and g) The

    reduced calcification percentage in the rotary kiln, decreases its

    thermal load and improves its functional stability, as the kiln

    burners are now used only for clinkering.

    Calciners have become essential devices in cement produc-

    tion but have also disadvantages: a) The lower temperatures of

    the exhaust gases may cause condensation of volatile alkalis,

    while the higher rotational speeds can increase the quantity of

    alkaline dust in the kiln, b) Reduction of NOx emissions is not

    common in all cement production systems using calciners,

    mainly due to geometric and operational differences, as well as

    to different quality and quantity of raw-mix and fuels, and c)

    The utilisation of fuels with low energy value, although eco-

    nomically advantageous, requires particular attention in order toavoid undesirable emissions of polluting and erroding gases.

    From the above, it becomes apparent that control of cal-

    cination is important because it affects fuel consumption,

    pollutant emissions and the final cement quality. Therefore,

    understanding the mechanisms of flow and transport phenom-

    ena in the calciner may contribute to more efficient production

    and better quality of cement.

    Recently, calciners have been studied with different geom-

    etries and operational conditions in 2D and 3D CFD simulations.

    Huanpeng et al [2] studied the influence of various physical

    parameters on the dynamics of gassolid two-phase flow in a

    precalciner using kinetic theory of granular flow to represent thetransport properties of the solidphase in a 2D model. Hu et al. [3]

    used a 3D model for a dual combustor and precalciner using a

    Eulerian frame for the gas phase and a Lagrangean one for the

    solid phase in order to predict the burn-out and the decompo-

    sition ratio during thesimultaneous injection of twotypes of coal

    and raw material into the device. Iliuta et al. [4] investigated the

    influence of operating conditions on the level of calcination,

    burn-out and NOx emissions of an in-line low NOx calciner, and

    made a sensitivity analysis of their model with respect to aero-

    dynamic and combustion/calcination parameters.

    In the present work a numerical model is described for the flow

    and transport processes taking place in an industrial calciner. The

    model is based on the solution of the NavierStokes equations for

    Fig. 1. Schematic of cement production.

    Fig. 2. Calciner device.

    82 D.K. Fidaros et al. / Powder Technology 171 (2007) 8195

  • 7/27/2019 Cement Model

    3/15

    the gas flow and on Lagrangean dynamics for the discrete

    particles, using a commercial CFD code. All necessary flow, heat

    and mass transfer and chemical reaction models are presented

    with the influence of turbulence simulated by a two-equation

    (k) model. Limited available measurements from the Olympus

    cement plant of AGET Hercules are used to verify the model.

    2. Mathematical models

    2.1. Gaseous phase

    The general form of the time-averaged transport equation for

    momentum, heat and mass of the gases is:

    A

    AtqU

    A

    AxqUU

    1

    r

    A

    ArqrVU

    1

    r

    A

    Ahq

    W

    r

    AU

    Ah

    A

    AxCU

    AU

    Ax

    1

    r

    A

    ArCUr

    AU

    Ar

    1

    r

    A

    AhCU

    1

    r

    AU

    Ah

    SU

    2

    where U, V, Ware the time-averaged velocities in the axial, radial

    and circumferential direction, respectively, the transport

    coefficient, and any time-averaged transported fluid property.

    2.2. Particle dynamics

    The particle trajectories are calculated from their corre-

    sponding motion equation:

    dUp

    dt FDUUp gi

    qpq

    qp fi 3

    where, the subscriptp denotes particle.For spherical particles, FD in the drag force term is:

    FD 3lCDRe

    4qpDp2

    4

    where the drag coefficient is calculated from:

    CD a1 a2

    Re

    a3

    Re25

    and 1, 2 and 3 are constants proposed by Morsi and

    Alexander [5].

    The additional force term fi

    in Eq. (3) may be due to pressure

    gradients, thermophoretic, Brownian or Saffman lift forces.

    2.3. Particle size distribution

    The particle sizes follow a RosinRammler distribution:

    MD eD=Do

    n 6

    where n is calculated from:

    n lnlnMD

    lnD=D7

    Each size interval is represented by an average diameter for

    which the trajectory calculations are performed.

    2.4. Particle heat transfer

    Particle heat transfer is due to convection, radiation and de-

    volatilization, as follows:

    Tpt t

    h ApTl dmpdt

    hfg ApeprH4

    R

    h Ap ApeprT3p

    Tpth ApTl

    dmpdt

    hfg ApeprH4

    R

    h Ap ApeprT3p

    0@

    1A

    e

    Ap hep rTe

    p

    mp Cpt

    82.5. Devolatilization model

    The devolatilization model of Kobayashi [6] is used:

    R1 A1expE1

    RTp ; 9a

    R2 A2expE2

    RTp

    9b

    where, R1 and R2 are competitive volatilization rates at different

    temperature ranges. These yield an expression for devolatiliza-

    tion:

    mvt

    mpomash

    Zt0

    a1R1 a2R2exp

    Zt0

    R1 R2dt

    dt 10

    The Kobayashi model requires known kinetic parameters

    (A1, E1) and (A2, E2) and the contribution of the two reactionsvia the factors a1 and a2. More specifically A1 =2.0e +07 s

    1

    and A2 =1.0e + 0 7 s1 are the pre-exponential factors, and

    E1 =1.046e +05 J/mol and E2 =1.67e + 05 J/mol are the

    activation energies. It is recommended that the value of a1should be equal to the fraction of volatiles that is determined by

    the proximate analysis, because this rate represents the volatile

    evaporation at low temperatures. The value of a2 should be

    equal to 1, as it expresses the contribution of the evaporation

    rate of volatiles at very high temperatures.

    2.6. Surface/coal combustion models

    After devolatilization is completed, there starts the surface

    chemical reaction of the coal particle which may be modelled as

    follows:

    2.6.1. Diffusion model

    The reaction rate is determined by the diffusion of the gas

    oxidant into the particle surface:

    dmp

    dt 4pDp Dim

    moTpqg

    SbTp Tl11

    In this model the particle diameter is assumed constant and,

    as its mass decreases, the active density decreases resulting in

    83D.K. Fidaros et al. / Powder Technology 171 (2007) 8195

  • 7/27/2019 Cement Model

    4/15

    a more porous particle. Eq. (11) proposed by Baum and

    Street [7] ignores the contribution of kinetics to the surface

    reaction.

    2.6.2. Kinetic/diffusion model

    The reaction rate is determined by the diffusion of gas

    oxidant into the particle surface or by the reaction kinetics. Themodel proposed by Baum and Street[7] and Field [8] is used, in

    which the diffusion rate is:

    R1 C1Tp Tl=2

    0:75

    Dp12

    and the kinetics rate:

    R2 C2exp E

    RTp

    13

    The kinetics rate incorporates the effects of chemical reactionin the internal surface of a coal particle and the epidermic

    diffusion. The rates R1 and R2 are combined to give the

    combustion rate of the coal (char) particle.

    dmp

    dt pD2pP0

    R1R2

    R1 R214

    The particle size is kept constant, until a significant reduction

    in its mass leads to a new size estimation.

    2.7. Particle radiation

    The radiation from the coal particles into the gas isincorporated via the P-1 model [910]:

    jdCjG 4p arT4

    p Ep

    a apG 0 15

    where, Ep and p are calculated from:

    Ep limVY0

    XNn1

    epnApnrT4pn

    pV16a

    ap limVY0

    XN

    n1

    epnApn

    V16b

    The quantity in Eq. (15) is:

    C 1

    3a ap rp17

    and p is calculated from:

    rp limVY0

    XN

    n1

    1fpn1epnApn

    V18

    The calculation ofp is repeated in the entire trajectory forn

    particles. Then, the source term that is introduced into the

    energy equation is:

    jqr 4p arT4

    p Ep a apG 19

    2.8. Chemical reaction models

    The present modelling of mixture fraction [11,12] with the

    method of probability density function (mixture fraction/PDF)

    requires the solution of transport equations for one or two

    conservative scalar properties. The effect of turbulence is also

    considered. The method of mixture fraction with PDF has been

    developed specifically for turbulent chemically reacting flow

    simulations. The chemical reaction is determined by turbulent

    mixing, which controls the limits of the kinetic rates. The PDF

    method offers many advantages compared to the method of

    finite reaction rate. The method of mixture fraction allows theexplicit intermediate calculation of chemical compound form-

    ing and the interlacing of turbulence and chemistry. The method

    is economic, because it does not require the solution of a large

    number of transport equations for each chemical species. More-

    over, it allows precise determination of auxiliary variables such

    as density, and it does not use average values, in contrast to the

    method of finite reaction rate.

    For a binary system such as fuel and oxidant, the mixture

    fraction can be formulated in terms of elemental mass fractions:

    f ZkZkO

    ZkFZkO20

    The value of f is calculated from the solution of a time-

    averaged transport equation:

    A

    Atqf

    A

    Axiqui f

    A

    Axi

    ltrt

    Af

    Axi

    Sm 21

    The source term Sm is present only when particle mass

    transport to the gaseous phase takes place.

    Simultaneously with the solution of Eq. (21), a conservative

    equation for the variance of mixture fraction, fV2, describing the

    interaction between chemistry and turbulence, is solved:

    A

    AtqfV2

    A

    Axiquif

    V2 A

    Axi

    ltrt

    AfV2

    Axi

    !

    CgltAf

    Axi

    !2Cdq

    e

    kfV2 22

    where, t, Cg and Gd are constants equal to 0.7, 2.86 and 2.6,

    respectively.

    2.8.1. Coal reaction mechanisms

    Coal combustion The most important physico-chemical

    change in the coal particle during heating is thermal frag-

    mentation (pyrolysis) at high temperatures. During this stage an

    84 D.K. Fidaros et al. / Powder Technology 171 (2007) 8195

  • 7/27/2019 Cement Model

    5/15

    important loss of weight occurs, because of dissolution of

    volatile matter, the quantity and composition of which depend

    on the ingredients of coal, its grain size and temperature. During

    dissolution of volatiles, a number of parallel reactions occur,

    with chemical combinations of reacting components or even

    species such as, for example, CH4, CHOH, C2H6, H2, and S2.

    After devolatization leading to production of water vapour, CO,CO2 etc, a series of progressive reactions of char and de-

    volatization gases take place as follows [1,1323]:

    Heterogeneous reactions

    Cs O2gCO2g 23a

    2Cs O2g2COg 23b

    Cs 2H2gCH4g 23c

    Cs CO2g2COg 23d

    Cs H2OgCOg H2g 23e

    Homogeneous reactions

    2COg O2g2CO2g 24a

    COg H2Og

    CO2g H2g 24b

    COg 3H2gCH4g H2Og 24c

    CH4g 2O2gCO2g 2H2Og 24d

    HCg 1:5O2gCO2g H2Og 24e

    The decomposition and polymerization reactions of the

    superior and unsaturated hydrocarbons are also added:

    Superior HCgY

    fragmentation

    Inferior HCg Cs

    Unsaturated HCgYSaturated HCg

    Unsaturated HCg H2gYPolymerization

    Superior HCg

    Pyrolysis As temperature increases, the humidity and the

    gases enclosed in the coal particles are released. The larger

    percentage of the non-chemically combined water is evaporated

    at temperatures below 105 C while the chemically combined at

    temperatures exceeding 350 C. At pyrolysis temperatures,

    certain types of coal melt, forming an intermediate product

    called metaplast. With the increase of temperature the metaplast

    is split, shaping the basic volatile products and semicoke,

    causing the coal particles to swell. This is described by a factor

    that depends on the composition of volatiles and the heating

    rate. The increase of particle volume does not influence the

    activity of pyrolysis, while the semicoke formed initially, is

    decomposed as temperature increases.The rate of thermal decomposition increases with increasing

    temperature up to a maximum value. Many researchers (for

    example, [15,17,23]) have found that pyrolysis ends around 850

    to 1000 C, while its duration is limited to a few seconds

    depending on the particle size. After the volatiles have been

    released, the remaining solid (char) still retains a small per-

    centage of volatiles (1.5%) like H2 and N2, requiring a tem-

    perature near 2000 C to be removed completely.

    Experiments show that the determination of volatiles in coal

    is demanding and time-consuming. Many measurements of

    volatiles based on the ASTM standard, present large differences

    in the percentage of volatiles depending on the rate of tem-perature increase and on the experimental method [10,15,16,

    18,19]. The solid remains of the particles formed during thermal

    decomposition are mainly fixed carbon, with high porosity and

    large internal surface, and the inorganic part is ash. The tem-

    perature varies between 1200 and 1800 C causing ash melting.

    The composition and the nature of ash as well as its properties

    (melting point, viscosity, etc) depend to a large extent on the

    pyrolysis conditions.

    In cases where the gaseous phase consists mainly of air, the

    pyrolysis and the combustion of char proceed simultaneously.

    However, in general, char combustion follows pyrolysis, with

    only a very small time overlap. In ordinary coal particles, volatiles

    tend to be emitted in concentrated but randomly distributed jetsfrom their surface. The larger jets reject volatiles during thermal

    decomposition while smaller jets begin and end during this

    period. When the gaseous phase is hot enough and rich in oxygen,

    the jets of volatiles ignite to form jet flames. In relatively large

    particles, the emission and combustion of volatiles can keep the

    char surface free of oxygen. When the surface of hot char is

    accessed by oxygen, there begins a heterogeneous combustion

    reaction with longer duration, lasting 15 to 20 times than the

    thermal decomposition of volatiles, depending on its evolution

    and combustion conditions [2129].

    The heating rate of coal particles depends on their size and

    contact with the thermal source. For example, the heating rate ofcoal powder by a surrounding flame is 1000 C/s, but when the

    flame is from powder coal particles, the rate may increase to

    10000 C/s. The pyrolysis results in a number of products with

    large differences in molecular weight, from gaseous hydrogen

    up to heavy organic species (tar). The data provided by exper-

    iments concerning rapid pyrolysis is not sufficient to determine

    the composition and distribution of intermediate products for

    various coals [3032].

    Thus, the mathematical models developed for devolatiliza-

    tion are based on the initial coal particle composition. Many

    researchers, assume that the coal is considerably homogeneous,

    so it is possible to be assumed as a heated mass and altered

    gradually from volatilescharash to charash and finally to

    85D.K. Fidaros et al. / Powder Technology 171 (2007) 8195

  • 7/27/2019 Cement Model

    6/15

    ash. From tables of ultimate analyses of coal and pet coke [1,2], it

    appears that the main components of volatiles are CO, CH, H2O,

    and H2. Given that the atmosphere of the calciner is oxidant and

    assuming that all these components react with oxygen, the main

    reactions considered as taking place are:

    CO 1=2O2CO2283:2kJ=mol 25a

    H2 1=2O2H2O242kJ=mol 25b

    CH4 2O2CO2 2H2O802:86kJ=mol 25c

    Char combustion The mechanism of char combustion has

    been investigated more than pyrolysis, without definitive

    answers to questions concerning the quantitative origin of

    some constituents after the end of transformation. Qualitatively,

    however, it has been modelled satisfactorily by various mathe-

    matical models. These were developed in order to describe the

    solid coal combustion and have found important application inreactions of porous solids with gases.

    Two simple mathematical models describe the reaction of

    coal grain with oxygen: the simple film and the double film

    model. In the first model the oxygen is diffused via a constant

    boundary layer in the surface of the char particle, where it reacts

    to form CO and CO2. The CO is then diffused in the well-mixed

    environment. In the second model, char reacts with CO 2 and not

    with oxygen, in order to produce CO that is burned in a thin

    flame inside the boundary layer. The CO reacts with oxygen

    inside the boundary layer, and thus the oxygen never approaches

    the char surface. Small particles (b100 m) are considered to

    burn according to the first model and larger (up to N2 mm)

    according to the second. However, the two models constituteonly the two extreme cases of char combustion and cannot,

    therefore, establish a general theory [2936].

    The real mechanism of combustion is more complicated,

    because of many factors involved such as particle size, local

    temperature, local oxygen concentration and reaction controlling

    mechanism. Generally, the oxygen and CO are readily available

    on the coal surface and can, therefore, react simultaneously with

    coal and also between each other. The situation becomes more

    complex when the char porosity is taken into consideration

    (intrinsic model).

    A more complex model proposed by Essenhigh [37] describes

    better the above processes. In this model the distribution oftemperature and concentrations are extended to the center of the

    particle. The more usual diffusion controlled combustion of CO

    can be extremely fast, consuming all the local oxygen before it

    reaches the char surface and reacting only with CO2. In the

    chemically controlled combustion of CO2 and O2, these have

    equal probability to react with the char surface. Moreover, ex-

    perimental data by Field [8] and Borghi [38] showed that the

    reaction of charCO2 is very slow in comparison with the reaction

    of charO2. Therefore, the latter can be considered as the main

    reaction on the char surface when the essential quantity of oxygen

    is available. However, the presenceof CO2 cannot be ignored and,

    thus, there always exists the probability of parallel reactions

    [26,27,2931]. Based on a comparative analysis of existing data

    for coal combustion and on the constitution and granulometry of

    particles (average char diameter100 m), the selected model

    for these particles was that of the kinetic/limited diffusion rate.

    This is similar to that of shrinking-reactant particle core adopted in

    the general theory of surface heterogeneous chemical reaction.

    The diffusion coefficientDim of oxidant in the porous char used in

    the present model was 5.0e05 m 2/s.

    2.8.2. Calcination mechanisms

    The calcination of limestone particles includes several stages,

    with each one imposing different chemical kinetics rates: a) Heat

    transfer from the gases to the particle surface and from it to the

    reaction interface, b) thermal decomposition of CaCO3 in the

    reaction interface, c) mass flux of CO2 from the reaction in-

    terface to the gases.

    For small limestone particles moving in high temperatures

    gases, the internal and external heat and mass transfer rates are

    high. Specifically, for particles with diameter between 1 and

    90 m and gas temperatures between 748 and 1273 K, Borgwardt[39] has reported that the calcination is chemically controlled and

    its rate is proportional to the surface area of the particle as

    determined by the BET method (nitrogen absorption at 77 K).

    Because, the limestone microstructure is not completely crystalic

    and has a diverse form of porosity, the surface determined by the

    BET method, is the sum of the porous surfaces accessed by

    nitrogen. Under these conditions, the calcination happens on the

    total available surface, giving pseudo-volumetric characteristics

    to the reaction.

    From the analysis of calcination data of high fineness

    limestone in isothermal reactors, it is concluded that, for a better

    description of the reaction evolution, the model of shrinking core

    should be selected, with the size diameter raised to thepower 0.6.The value of the exponent (b1) is explained by the fact that the

    calcination proceeds radially to the particle core, without

    inhomogeneities in the reaction interface. When the raw-mix

    particles are small, the reaction interface of the calcination is not

    easy to determine. The internal thermal gradients and the partial

    pressure of CO2 are also difficult to estimate. For these reasons,

    most calcination models of high fineness particles consider that

    the surface temperature is equal to the gas temperature,

    neglecting the internal thermal gradients [40].

    The decomposition reaction of CaCO3 is strongly endother-

    mic. Its thermodynamic state is defined by the reaction enthalpy

    H and the equilibrium pressure PCO2,eq,:

    PCO2;eq exp H

    RTS

    R

    26

    These values depend on temperature and are influenced by the

    nature of limestone, its degree of cleanliness and mainly by its

    structural mesh. The lower the degree of cleanliness of raw

    material, the lower is the reaction enthalpy. Also, the function of

    temperatureequilibrium pressure PCO2,eq = f(T) develops to

    lower temperatures because of the chemical kinetics of the

    recently formed CaO and the impurities in the reaction

    environment. The values of reaction enthalpies provided by the

    open literature [4046] for the present endothermic reactions

    86 D.K. Fidaros et al. / Powder Technology 171 (2007) 8195

  • 7/27/2019 Cement Model

    7/15

    vary. Particularly, in the high interest range for calcination (800 to

    1000 C), the reaction enthalpy is not linearly dependent on

    temperature. Thus, for practical calculations 900 =1660 kJ/kgCaCO3 =396 kcal/kg CaCO3 can be assumed a constant value for

    this specific temperature range [1,41].

    The decomposition of limestone takes place in a reaction

    zone, where the core of unreacted CaCO3 and the newly formed

    CaO meet. This front moves from the perimeter to the center

    with a certain speed, while heat is transferred simultaneously to

    the core and CO2 is emitted to the outside. This reaction

    proceeds in the following stages: a) Heat is transferred from the

    surroundings to the particle surface, b) heat is conducted

    through the reacted layer to the reaction zone, c) chemical

    reaction occurs in the reaction zone, CO2 emission, nuclei

    creation and reforming of CaO, and d) CO2 is diffused through

    the CaO layer to the particle surface and the surroundings.The final reaction speed is a function of the rates of the

    above stages. Because these rates are of the same order of

    magnitude, a balance is achieved in the decomposition front,

    under the prevailing temperature and partial pressure of CO 2,

    so that the rates of the above stages become equal. If large

    limestone particles exist, diffusion of mass and conduction of

    heat will dominate, especially when the surrounding temper-

    ature is high and the partial pressure low. In the case of low

    temperatures and high partial pressures of CO2, the material

    transformation occurs by the diffusion of CaO. For a fine

    granulometry of ground limestone or raw-mix in ordinary

    calcination conditions, the chemical kinetics play a decisiverole [1].

    Thus, the proposed model, calculates the rates of particle

    calcination and heat transfer by considering: a) the heat transfer

    by convection from the gases to the particle and by conduction

    to the particle interior, b) surface decomposition of CaCO3, and

    c) mass transport of CO2 from the reaction interface via the

    porous particle to the gaseous environment.

    The calcination is a heterogeneous reaction and occurs at the

    lime surface when the local pressure exceeds the criterion of

    Baker [47]:

    Pe 1; 826 10

    7

    exp

    19; 680

    T

    27

    The reaction rate at the interface is expressed as follows,

    Borgwardt [39]:

    Rate ks ACaCO3 28a

    where,

    ks Aexp Ea

    RT

    28b

    The activation energy Ea of the decomposition reaction is in

    the range 165205 kJ/mol.

    The calcination of small limestone particles dispersed in the

    gaseous phase, can proceed at temperatures up to 1600 C. The

    effect of CO2 partial pressure is incorporated in the decompo-

    sition rate by modifying it as proposed by Darroundi and Searcy

    [48]:

    kVs ks for Pb102Pe 29

    k

    V

    s ksPe

    P=Pe for 102

    Peb

    Pb

    Pe 30

    The effect of temperature on calcination chemical kinetics is

    shown in Fig. 3. During calcination, the thermal conductivity of

    Fig. 3. Variation of calcination chemical kinetics with temperature.

    Fig. 4. Calciner side view.

    Table 1

    Mass flow rates at the inlets of the calciner

    Kind of

    mass flow

    Case 1 coal Case 2 pet coke

    Quantity [kg/s] Percentage Quantity [kg/s] Percentage

    mCaCO3 52.47 54.3% 52.47 54.6%

    mCoal 3.78 3.9% 3.17 3.3%

    mTertiary Air 39.36 40.7% 39.36 41.0%

    mAir Coal 0.97 1.0% 0.97 1.0%

    Total 96.59 100.0% 95.98 100.0%

    87D.K. Fidaros et al. / Powder Technology 171 (2007) 8195

  • 7/27/2019 Cement Model

    8/15

    lime depends on the state of the solid material and differs con-

    siderably for non-calcinated, partially calcinated or fully

    calcinated particles. This is mainly due to the different structure,

    but also to the change of the specific surface area. In the reaction

    region, the thermal conductivity of the particle is a linear

    function of specific surface area and temperature. Thus, the

    thermal conductivity of CaCO3 was 1.646W/(m.K) and of CaO

    0.860 W/(m.K). The mass fraction of CO2 is determined from a

    diffusion equation assuming a spherical particle.

    3. Computational details

    3.1. Calciner geometry

    The modeled calciner, Fig. 4, consists of a cylindrical and a

    conical section having three kinds of inlets at the bottom part

    and an outlet at the top, from where the products such as

    calcined raw-mix, CO2, and other gases exit. Raw-mix is fed

    into the calciner via two 0.6 m diameter pipes inclined at 60

    to the horizontal. The tertiary air enters axially from the bottom

    via a concentric 2.6 m diameter duct and the coal is fed at the

    lower conical part via two 0.2 m pipes at 30 to the horizontal.

    The physico-chemical processes take place in the main volume

    of the calciner, consisting of a 6.6 m diameter cylinder with

    20 m height. The upper conical part has 1.1 m height and leads

    to a cylindrical part with 4.3 m diameter and 5 m height. The

    total calciner volume is 850 m3. The coal entries are at 2.4 m

    height from the start of the cone and at 2.68 m from the

    calciner axis.The computational domain consists of a hybrid mesh of

    67.104 cells. Because of symmetry, the calculations were carried

    out for one half of the calciner using the FLUENT code. Two

    fuels (coal and pet coke) were considered and the total rate of

    mass (raw-mix, coal and air) fed into the three kinds of inlets was

    aproximately 100 kg/s. As shown in Table 1, the larger per-

    centage of mass rate is that of CaCO3, followed by tertiary air,

    coal and finally the coal feeding air.

    The RossinRammler distribution of the raw-mix size had

    an average value of d=16.6 m and a spread parameter of

    n =0.822, while the coal had d=34.5 m and n =1.248. The

    analysis of the raw meal and coal particles is given in Table 2.The tertiary air entered with a velocity 24 m/s, coal with 11.5 m/s

    and the raw-mix with 1.5 m/s. The coal was fed pneumatically

    while the raw-mix entered by gravity.

    All the geometric data and the initial and boundary conditions

    were supplied by Olympus plant of AGET Hercules in Volos,

    Greece.

    4. Results and discussion

    4.1. Case 1 (Good quality coal)

    Fig. 5 shows the calculated velocity distribution of the gaseous

    phase for Case 1 (good quality coal) in two vertical diametral

    Table 2

    Ultimate analysis for the solid feeds

    Components Case 1 coal Case 2 pet coke

    As used (%) Dry basis (%) As used (%) Dry basis (%)

    Humidity 1.45 0.0 1.45 0.0

    Volatiles 27.70 28.11 13.21 13.40

    Ash 12.05 12.23 0.14 0.14Fix carbon 55.80 56.62 81.70 82.90

    Total 97.00 96.96 96.50 96.44

    Fig. 5. Velocity distribution in vertical symmetry plane (left) and at 90 (right) for Case 1.

    88 D.K. Fidaros et al. / Powder Technology 171 (2007) 8195

  • 7/27/2019 Cement Model

    9/15

    planes normal to each other. The gases undergo an abrupt

    deceleration at the beginning of the lower conical part due mainly

    to the entry of the coal and secondly of the raw meal. 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, a fact due to the relative absence of particles.In Fig. 6 higher temperatures are observed in the opposite

    side of the raw-mix entries. This is due to the trapping of small

    coal particles while the concentration of CaCO3 particles is low.

    The main body of the calciner is maintained at temperatures

    little above the threshold for calcination, so that calcification

    takes place almost in the whole device. There are no spots of

    high temperature but rather regions of low temperature because

    of intense calcination, resulting to high heat absorption. The

    high temperatures in regions where higher velocities prevail are

    mainly due to the high concentration of burning coal and to the

    absence of CaCO3 particles.Fig. 7 shows concentration distributions of CO2, O2 and H2O

    in various horizontal cross-sections. Higher concentrations are

    observed a little after the raw-mix inlet. It should be noted that

    high CO2 concentrations result to high heat absorption, thus

    Fig. 6. Temperature field in the vertical symmetry plane (left) and at 90 (right) for Case 1.

    Fig. 7. Concentration distributions of CO2 (left), H2O (middle) and O2 (right) for Case 1.

    89D.K. Fidaros et al. / Powder Technology 171 (2007) 8195

  • 7/27/2019 Cement Model

    10/15

    limiting calcination. The CO2 concentration decreases progres-

    sively towards the exit, mainly because the available CaCO3 is

    also decreasing. The amount of CO2 produced by coal com-

    bustion is much smaller than that due to the calcination (ratio

    1:6). Smaller O2 concentrations are observed in regions of high

    temperature, mainly due to pyrolysis and the start of combustion

    of the coal remains (char). In the part of the cylinder whereintense calcination takes place, the O2 concentration remains at

    low levels because of the high raw-mix and CO2 concentrations,

    while in regions where coal particles have been trapped, lower

    CO2 concentrations are observed (mainly from coal combus-

    tion). Higher H2O concentrations are recorded after the fuel inlet

    in the regions of pyrolysis and coal combustion. These

    concentrations decrease along the height of the device and

    away from the combustion region.

    The trajectories of coal particles are presented in Fig. 8 and

    those of CaCO3 in Fig. 9. The average passage length covered by

    all particles is about 54 m, while the longest exceeds 75 m. The

    right parts ofFigs.8 and 9, show the parts of the trajectories wherethe 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 trajectory shows the

    cooling of the ashby thegases. Theaverage passage lengthof coal

    particles is 32 m, while the longest exceeds 45 m. Their average

    residence time is 5 s, mainly because the air moves faster

    sweeping the smaller coal particles. The average particle

    residence time is 10 s while the longest 15 s. Calcination is

    noticeable in the semi-cylinder defined by the raw-mix feeding

    pipes and the coal inlets, which agrees with the temperature field

    and the CO2 concentration. The predicted calcination for Case 1reaches 96.5%, and is realised in all the active calciner height.

    4.2. Case 2 (Pet coke)

    Fig. 10 shows the gas velocity field for Case 2 (pet coke

    fuel). In the start of the conical part, the gases undergo a strong

    deceleration, again for the same reasons as in Case 1. In the

    main cylindrical section of the calciner, a region with higher

    velocities opposite to the two raw-mix inlets and in the upper

    conical part is observed. The gas velocity reaches 8.5 m/s,

    encouraged by the relatively small particle load. In contrast, the

    velocities decrease to 56 m/s in the remaining part. This is

    mainly due to the higher particle concentration, which

    influences the exiting speed of the gases.

    In Fig. 11 higher temperatures relative to Case 1 are observed.

    This is attributed to the better quality and utilization of the fuel.

    Higher temperatures areobserved in the opposite side of the raw-

    mix inlets, mainly due to the trapping of small coal particles

    while the concentration of CaCO3 particles is exceptionallysmall. The main body of the calciner is kept at temperatures well

    above the effective calcination temperature, so that large CaCO3quantities are calcinated in short times. The high temperatures

    seen in regions where higher speeds prevail are attributed to the

    absence of CaCO3 particles that would consume the heat re-

    leased. Also at the exit, higher temperatures are observed, be-

    cause calcification there is exceptionally limited.

    Fig. 12 shows that higher concentrations of CO2, O2 and

    H2O are present at a small distance after the raw-mix inlet.

    Again it should be noted that high CO2 concentrations result to

    high heat absorption, as far as it does not limit the calcination.

    CO2 concentration decreases gradually along the 16.5m calcinerheight until the exit, where the available quantity of CaCO3 is

    also decreased. Smaller O2 concentrations are observed in the

    regions where the temperatures are high (pyrolysis and com-

    bustion of solid coal remains). Also, in the semi-cylindrical part,

    where intense calcination takes place, the concentration of O2remains at low levels because of high raw-mix and CO2concentrations, while in regions where coal particles have been

    trapped, still lower concentrations are observed. Higher con-

    centrations of H2O are seen little after the coal entry and in the

    pyrolysis and coal combustion regions. These concentrations

    are decreasing along the calciner height.

    From the predicted trajectories of pet coke particles (not

    shown), most calculated residence times did not exceed 10 sFig. 8. Trajectories of coal particles for Case 1 (combustion is marked in red).

    Fig. 9. Trajectories of CaCO3 particles for Case 1 (calcination is marked in red).

    90 D.K. Fidaros et al. / Powder Technology 171 (2007) 8195

  • 7/27/2019 Cement Model

    11/15

    with the longest reaching 16.2 s. The average trajectory length

    for all particles (raw-mix and pet coke) was roughly 52 m, while

    the longest exceeded 73 m. The average residence time of pet

    coke particles was smaller (4.5 s) than in Case 1. The average

    trajectory length of these particles was 30 m while the longest

    42 m. It should be noted that the combustion of most pet coke

    particles is completed inside the device. The higher tempera-tures observed are due to the intense and fast char combustion,

    and are accompanied by low concentrations of O2 and high

    water vapour levels. This results from the volatiles and

    hydrogen compounds in the char. The fast combustion of pet

    coke forces the ash to abandon the calciner faster than other

    particles.

    From the predicted trajectories of CaCO3 particles (not

    shown), their average residence time reached 10 s and the longest

    15.6 s, indicating that the speed of these particles is higher thanthat of the pet coke. The longer residence time corresponds to the

    particles that collide in the upper conical part (before the exit) and

    Fig. 10. Velocity distribution in vertical symmetry plane (left) and at 90 (right) for Case 2.

    Fig. 11. Temperature field in the vertical symmetry plane (left) and at 30 (right) for Case 2.

    91D.K. Fidaros et al. / Powder Technology 171 (2007) 8195

  • 7/27/2019 Cement Model

    12/15

    are trapped by the rising particles from the central part of the

    calciner. Calcination is realised fast and in a region close to the

    calciner axis. The calcination rate for the particular fuel (Case 2)

    reaches 98.7%, without taking advantage of the total active height

    by the majority of CaCO3 particles. The reason that a small raw-

    mix quantity is not being calcinated is the large diameter of

    CaCO3 particles and the rapid acceleration observed near the

    device exit.

    Fig. 13 shows details of the fuel particle trajectories. It is

    evident that the pet coke particles (Case 2) react faster than thoseof coal (Case 1), as soon as they enter the calciner.

    Finally, the evolution of CaCO3 calcination for Cases 1 and 2 is

    depicted in Fig. 14. The differences relate to the energy prevailing

    levels observed, the aerodynamics (mass density, particle load)

    and the CO2 concentrations that suppress calcination. The

    evolution of calcination is represented by the CaCO3 decompo-

    sition along the calciner height, starting from where calcification

    Fig. 12. Concentration distributions of CO2 (left), H2O (middle) and O2 (right) for Case 2.

    Fig.13. Details of fuelparticle trajectories for Case 1 (upper) and Case 2 (lower).(The pet coke particles (Case 2) react faster than the coal particles (Case 1)). Fig. 14. Evolution of calcination with height for Cases 1 and 2.

    92 D.K. Fidaros et al. / Powder Technology 171 (2007) 8195

  • 7/27/2019 Cement Model

    13/15

  • 7/27/2019 Cement Model

    14/15

    t time

    U gas velocity

    Up particle velocity

    U,V,W time-averaged axial, radial and circumferential

    velocities

    zk mass fraction for the chemical element k

    zKF mass fraction for fuel streamzKO mass fraction for the oxidizer stream

    S change of grammolecular entropy

    H reaction enthalpy

    Greek letters

    absorption gas factor

    p equivalent absorption factor

    transport coefficient

    p particle brightness

    R particle temperature produced by the intensity of

    thermal radiation

    gas molecular viscosityt turbulent viscosity

    p particle mass density

    g gas density

    Boltzmann's constant

    p particle scattering equivalent factor

    time-averaged transported fluid property

    Acknowledgements

    This work was partially supported by the General Secretariat

    for Research and Technology of Greece and the cement

    company AGET Hercules (EPET-II/96SYN121). The authors

    wish to thank Dr. I. Marinos and Mr. T. Pissias of AGET for thedata and for useful discussions.

    References

    [1] E. Kolyfetis, C.G. Vagenas, Mathematical modeling of separate line

    precalciner, ZKG International 2 (1988) 559563.

    [2] L. Huanpeng, L. Wentie, Z. Jianxiang, J. Ding, Z. Xiujian, L. Huilin,

    Numerical stu dy of gassolid flow in a precalciner using kinetic theory of

    granular flow, Chemical Engineering Journal 102 (2004) 151160.

    [3] Z. Hu, J. Lu, L. Huang, S. Wang, Numerical simulation study on gassolid

    two-phase flow in pre-calciner, Communications in Nonlinear Science &

    Numerical Simulation 11 (3) (2006) 440451.

    [4] I. Iliuta, K. Dam-Johansen, A. Jensen, L.S. Jensen, Modelling of in-line-low-NOx calciners a parametric study, Chemical Engineering Science

    57 (2002) 789903.

    [5] S.A. Morsi, A.J. Alexander, An investigation of particle trajectories in two-

    phase flow systems, Journal of Fluid Mechanics 55 (1972) 193208.

    [6] H. Kobayashi, J.B. Howard, A.F. Sarofim, Coal devolatilization at high

    temperatures, Proc. 16th Intl Symposium on Combustion, The Combustion

    Institute, 1976, pp. 411425.

    [7] M.M. Baum, P.J. Street, Predicting the combustion behavior of coal particles,

    Combustion Science and Technology 3 (1971) 231243.

    [8] M.A. Field, Rate of combustion of size-graded fractions of char from a

    low rank coal between 1200K2000K, Combustion and Flame 13 (1969)

    237252.

    [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 (3) (1996)

    289294.

    [10] M.Q. Brewster, T. Kunitomo, The optical constants of coal, char and

    limestone, ASME Journal of Heat Transfer 106 (1984) 678683.

    [11] Y.R. Shivatahnu, G.M. Faeth, Generalized state relationships for scalar

    properties in non-premixed hydrocarbon/air flames, Combustion and

    Flame 82 (1990) 211230.

    [12] W.P. Jones, J.H. Whitelaw, Calculations methods for reacting turbulent

    flows: a review, Combustion and Flame 48 (1982) 126.

    [13] R.A. Srinivasan, S. Sriramulu, S. Kulasekaran, P.K. Agarwal, Mathemat-ical modeling of fluidized bed combustion 2: combustion of gases, Fuel

    77 (1998) 10331049.

    [14] R.L. Backreedy, R. Habib, J.M. Jones, M. Pourkashanian, A. Williams, An

    extended coal combustion model, Fuel 78 (1999) 17451754.

    [15] F. Winter, M.E. Prah, H. Hofbauer, Temperatures in a fuel particle burning

    in a fluidized bed: the effect of drying, devolatilitazion and char

    combustion, Combustion and Flame 108 (1997) 302314.

    [16] A.S Hull, P.K. Agarwal, Estimation of kinetic rate parameters for coal

    combustion from measurements of the ignition temperature, Fuel 77 (1998)

    10511058.

    [17] L.D. Smoot, International research center's activities in coal combustion,

    Progress in Energy and Combustion Science 24 (1998) 409501.

    [18] L. Huilin, Z. Guangbo, C. Rushan, C. Yongjin, D. Gidaspow, A coal

    combustion model for fluidized bed boilers, Fuel 79 (2000) 165172.

    [19] L. Huilin, Z. Gungbo, B. Rushan, Y. Lidan, Q. Yukun, Modeling of coalcombustion in a 25-MW FBC power plant, Energy 24 (1999) 199208.

    [20] H.H. Liakos, K.N. Theologos, A.G. Boudouvis, N.C. Markatos, Pulverized

    coal char combustion: the effect of particle size on burner performance,

    Applied Thermal Engineering 18 (1998) 981989.

    [21] W. Sujanti, D. Zhang, A laboratory study of spontaneous combustion of

    coal: the influence of inorganic matter and reactor size, Fuel 78 (1999)

    549556.

    [22] C.N. Eastwick, S.J. Pickering, A. Aroussi, Comparisons of two com-

    mercial computational fluid dynamics codes in modeling pulverized coal

    combustion for a 2,5 MW burner, Applied Mathematical Modeling 23

    (1999) 437446.

    [23] L. Armesto, J.L. Merino, Characterization of some coal combustion solid

    residues, Fuel 78 (1999) 613618.

    [24] M.S. Khan, S.M. Zubair, M.O. Budair, A.K. Sheikh, A. Quddus, Fouling

    resistance model for prediction of CaCO3 scaling in AISI 316 tubes,International Journal of Heat and Mass Transfer 32 (1996) 7379.

    [25] J.R. Fan, X.H. Liang, Q.S. Xu, X.Y. Zhang, K.F. Cen, Numerical simulation

    of the flow and combustion processes in a three-dimensional, W-shaped

    boiler furnace, Energy 22 (1997) 847857.

    [26] J.K. Fink, Pyrolysis and combustion of polymer wastes in combination with

    metallurgical processes and the cement industry, Analytical and Applied

    Pyrolysis 51 (1999) 239252.

    [27] P. Basu, Combustion of coal in circulating fluidized-bed boilers: a review,

    Chemical Engineering Science 54 (1999) 55475557.

    [28] F.C. Lockwood, T. Mahmud, M.A. Yehia, Simulation of pulverized coal test

    furnace performance, Fuel 77 (1998) 13291337.

    [29] 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) 17371742.

    [30] C.A. Gurgel, J. Saastamoinen, J.A. Carvalho, M. Aho, Overlapping of thedevolatilization and char combustion stages in the burning of the coal par-

    ticles, Combustion and Flame 116 (1999) 567579.

    [31] W.B. Fu, B.L. Zhang, S.M. Zheng, A relationship between the kinetic

    parameters of char combustion and the coal's properties, Combustion and

    Flame 109 (1997) 587598.

    [32] T. Gentzis, F. Goodarzi, C.N. Koykoyzas, A.E. Foscolos, Petrology,

    mineralogy and geochemistry of lignitesfrom Crete, Greece,Coal Geology

    30 (1996) 131150.

    [33] J. Zhang, S. Nieh, Comprehensive modeling of pulverized coalcombustion

    in a vortex combustor, Fuel 76 (1997) 123131.

    [34] L. Reh, Challenges of circulating fluid-bed reactors in energy and raw

    materials industries, Chemical Engineering Science 54 (1999) 53595368.

    [35] 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) 267282.

    94 D.K. Fidaros et al. / Powder Technology 171 (2007) 8195

  • 7/27/2019 Cement Model

    15/15

    [36] I.E. Sarris, O. Giannopoulos, D. Fidaros, N.S. Vlachos, Modelling coal

    combustion in a rotary cement kiln, 3rd National Congress on Computational

    Mechanics, Volos 2426 (June 1999) 665672.

    [37] R.H. Essenhigh, in: M.A. Elliot (Ed.), Chemistry of coal utilization, Second

    Supplementary Volume, Wiley, New York, 1981, pp. 11531178.

    [38] G. Borghi, A.F. Sarofim,J.M. Beer, 70thAIChE Annual Meeting,New York,

    vol. 34c, Nov. 1977, pp. 1417.

    [39] R.H. Borgwardt, Calcination kinetics and surface area of dispersed limestoneparticles, AIChE Journal 31 (1) (1985) 103111.

    [40] F. Acke, I. Panay, Activation energy of calcination by means of a temperature

    programmed reaction technique, Thermochimica Acta 300 (1997) 7376.

    [41] V.L. Boris, Mechanism of thermal decomposition of alkaline-earth

    carbonates, Thermochimica Acta 303 (1997) 161170.

    [42] A.K. Galwey, M.E. Brown, Arrhenius parameters and compensation

    behavior in solid-state decompositions, Thermochimica Acta 300 (1997)

    107115.

    [43] F. Acke,I. Panas,Activation energy of calcination by means of a temperature

    programmed reaction technique, Thermochimica Acta 306 (1997) 7376.

    [44] J. Khinast, G.F. Krammer, C. Brunner, G. Staudinger, Decomposition of

    limestone: the influence of CO2 particle size on the reaction rate, Chemical

    Engineering Science 51 (1996) 623634.

    [45] N. Hu, A.W. Scaroni, Calcination of pulverized limestone particles under

    furnace injection conditions, Fuel 75 (1996) 177186.

    [46] S.V. Krishnan, S.V. Sotirchos, Effective diffusivitty changes duringcalcination, carbonation, recalcination and sulfation of limestones,

    Chemical Engineering Science 49 (1991) 11951208.

    [47] E.H. Baker, Calcium oxide-carbon dioxide system in the pressure range

    1300 atmospheres, Journal of Chemical Society (London) (1962) 464470.

    [48] T. Darroudi, A.W. Searcy,Effectof CO2pressure on the rate of decomposition

    of calcite, Journal of Physical Chemistry 85 (26) (1981) 39713974.

    95D.K. Fidaros et al. / Powder Technology 171 (2007) 8195