Sei sulla pagina 1di 24

Chemical Engineering Science, 1968, Vol. 23, pp. 297-320.

Pergamon Press.

Printed in Great Britain.

The mechanism of the thermal decomposition of calcium carbonate


A. W. D. HILLS John Percy Research Fellow in Process Metallurgy, Department of Metallurgy, Imperial College, London (First received 6 June 1967; in revisedform 14 September 1967) Abstract-The decomposition of single sintered spheres of calcium carbonate has been investigated using a thermo-balance, modified to allow the simultaneous measurement of the temperature. and weight of a decomposing sample. It has been shown that the decomposition reaction takes place on a definite boundary between the undecomposed carbonate, and the layer of porous lime formed outside it. This boundary moves towards the centre of the compact but remains spherical in shape. The rate of the reaction is controlled by the transfer of heat to this reaction boundary, and by the transfer of CO, away from it. This reaction mechanism is at variance with the mechanism proposed by previous workers who have suggested that the reaction is controlled by a chemical step at the reaction interface. Their conclusions have been based on the experimental observation of certain characteristics of the reaction that have been assumed to be symptomatic of chemically controlled reactions. This paper shows that transport controlled reactions can also display these characteristics, and so they do not provide any evidence as to the mechanism controlling the reaction. The modified thermo-balance has been used to measure the diffusion coefficient and thermal conductivity of the porous lime layer, and the mass and heat transfer coefficients to the surface of the compact. The values so obtained agree with values obtained from the literature. Three reaction models have been developed from which it is possible to predict the progress of the reaction for a single sphere in a constant environment. Predictions made by the two more complicated models are shown to agree very well with the experimental results. INTRODUCTION

industrial processes which involve reactions between gases and packed beds of solids result in the formation of a new solid phase. The existing theoretical treatments for packed bed processes do not apply to these reactions since their mathematical formulationsinvolve strongly non-linear equations. Since they are of considerable interest to the metallurgical industries the John Percy Research Group in Process Metallurgy initiated a programme to study numerical methods for predicting the rates at which these reactions will occur in packed and moving bed reactors. The decomposition of calcium carbonate was chosen as suitable for the initial work in this programme since it is a relatively simple exam.ple of this type of reaction. Moreover, it had been the subject of many investigations, and its mechanism appeared to be firmly established. However, the mechanism was found to be more complicated than had been thought previously, and its

MANY of the high temperature

determination turned out to be a major project in itself. It is this work that is reported here. After a brief review of previous work, this paper describes the experimental method used here and discusses the results obtained. These showed that the reaction was controlled by heat and mass transfer, and three increasingly complex mathematical models for the reaction are developed and their predictions compared with the results. The reaction mechanism proposed here is then compared with the mechanisms proposed by previous workers.
PREVIOUS WORK

A review [l] of the many investigations into the decomposition of calcium carbonate shows that the reaction occurs on a definite boundary between the lime and carbonate phases. Moreover, this boundary moves at an approximately constant rate towards the centre of a particle decomposing in an isothermal enclosure that contains a constant partial pressure of CO,. 297

A. W. D. HILLS

Almost all the samples used in the previous work have been spherical, or can be treated as if they were spherical. Simple conservation considerations show that the following identity applies:

If drldt is constant, this identity shows that the reaction rate per unit area of reaction interface is also constant. The approximate constancy of dr/dt has led many workers to propose that the reaction is controlled by a chemical step at the reaction interface. Ingraham and Marier[2] suggest that this step is the diffusion of CO, through a thin layer of active lime. Hyatt, Cutler and Wadsworth[3] have proposed a mechanism similar to the surface catalysis of gas reactions, involving the reaction of lime, carbonate, active lime, and absorbed carbon dioxide on the interface. Satterfield and Feakes[4] have also proposed a chemical control mechanism which involves the cracking of the original carbonate lattice. These three groups of workers consider that the effects of CO, partial pressure and of temperature on the reaction rate act as confirmatory evidence for the proposed mechanisms. Ingraham and Marier find that the rate decreases linearly as the CO, partial pressure around the sample is increased, whereas Hyatt, Cutler and Wadsworth find a slightly non-linear relationship. Both workers find the Arrhenius activation energy for the reaction to be about 40 kcals g-mole-lapproximately equal to the heat of the reaction. Satterfield and Feakes find a much larger activation energy - 360 kcals g-mole-. Only Satterfield and Feakes and Azbe[5] have measured the actual temperature of decomposing calcium carbonate samples. They both found that the temperature was considerably below that of the surroundings, and remained constant during the reaction. Their experiments were conducted in pure CO, at one atmosphere pressure and very high temperatures were therefore involved. They concluded that heat transfer was important in controlling the reaction rate at high temperatures and Haslam and Smith [6] and

Narsimhan [7] have advanced theoretical treatments for the reaction which assume it to be exclusively heat transfer controlled. None of the proposed mechanisms was adequate to account for the results of the authors preliminary experiments, and so it was necessary to conduct a full scale investigation into the mechanism of the reaction. Single spheres of calcium carbonate were decomposed whilst supported from a thermo-balance modified to allow the samples temperature to be measured at the same time as its weight. The experimental method is described below, and the results are discussed in the next section.
EXPERIMENTAL PROCEDURE

The apparatus is illustrated in Fig. 1. Carbonate samples were hung in a high temperature furnace tube (5 cm (i.d.)) from a Stanton 1 mg recording thermo-balance which was mounted on a water-cooled plate above the furnace. The supporting frames for the balance and furnace were completely independent, and the furnace could be raised and lowered hydraulically for the insertion of samples. The furnace temperature was controlled from a thermocouple placed in a twin bore alumina sheath supported from the gas outlet assembly so that the thermocouple bead was just above the sample. The output from this thermocouple was fed to the galvanometric-type temperature controller incorporated in the thermo-balance, which switched a bias resistance in and out of the circuit feeding the furnace windings. The output from a similar thermocouple was fed to a potentiometer, or a potentiometric recorder, so that the furnace temperature could be accurately determined. Air and air/CO* mixtures were swept through the furnace from a series of rotameters to keep the composition of the gas phase around the sample constant during each run. So that each run could be started with the sample at thermal equilibrium, pure COP was swept through the furnace while the sample and furnace were being heated to temperature. A double acting valve at the bottom of the furnace allowed this CO1 to be abruptly changed to the desired air/CO, mixture. The gases passed through a small bed of broken solids at the bottom of the furnace tube to even the flow, and through a porous silica disc just below the sample to heat them up to the temperature of the furnace. Throughout the experiments the total flow rate of gases was 1 1. mitt-. A convection shield was suspended just above the sample. This consisted of a cylinder of M.I. 28 brick, 4 cm in height, containing a central passage, and it effectively isolated the sample from natural convection currents originating at the water-cooled top of the furnace. This shield, and the porous silica disc, made a sample enclosure 8 cm in height, and the temperature within this enclosure could be maintained constant to ?O.PC, in the range 600-900C. So that its temperature could be recorded a platinumplatinum/ 13 per cent rhodium thermocouple was compacted

298

Thermal decomposition

of calcium carbonate
Balance pan support wire

&td plated Plug and 8ocket lead Terminal block on balance frwmr 0301 copper lvin

screw support Balance pan

Balance cam

floor

Alumina twin bore insulating tubing

Funacr

tubr Twin bore silica Control tube

thermocouple

Refractory shield

convectton

fvbe

thermocouple

Rhodium /#at/n Thermocouple leads

Reacting sphere

Fig 1. Diagram of apparatus for the simultaneous measurement temperature. into the centre of each sample. Two short tails were left protruding, and these were hooked over platinum-platinum/l3 per cent rhodium leads that were supported from the balance pan. These leads were contained in alumina sheaths 2 mm dia., and were attached to gold plated plugs fitting into insulated sockets attached to the pan support wires. These sockets held gold-plated screw terminals from which 0401 in. thick copper wires carried the thermal E.M.F. to terminals on the balance case. These thin wires had no effect on the sensitivity of the balance. The thermocouple tails were brazed onto the ends of the thermocouple leads, using white gold solder-an alloy of platinum and gold with a convenient melting point. A very small heat source was used in this

of the samples weight and

brazing so that there was no chance of the sample decomposing prematurely. The thermal E.M.F.s from the sample thermocouple, and from the furnace thermocouple, were biased by a fixed and measurable E.U.F. and fed to a sensitive O-l mV potentiometric recorder. The sample and furnace temperatures could then be calculated from the reading of this recorder, the fixed biasing E.M.F. and the temperature within the balance case. The samples were prepared from Analar grade calcium carbonate manufactured by Hopkin and Williams Ltd. The carbonate had been prepared by precipitation, and consisted of rhombs of calcite about 5-8~ across. The spherical samples were prepared by compacting this powder under

299

A. W. D. HILLS pressure in a three part tnould made from Vinamould.t A thermocouple was placed at the centre of the, mould cavity, which was then filled with the carbonate powder, and the mould was placed in a conventional 2 in. die. The sample was compacted at a pressure of 30 tons in-*, removed from the mould and then sintered in pure CO, at 820C for 2 hr. The spheres produced in this way had a density at least 90 per cent of the theoretical maximum. The thermocouple could not be placed accurately in the centre ,of the sphere but it was never far from the centre. Prior to each experiment, ten separate determinations of the sample diameter were made and its thermocouple tails were then brazed onto the thermocouple leads hanging from the balance. After the experiment, the solder was melted, the sample removed and its weight and diameter were determined. The final weight and the total weight loss were used as a stoichiometric check on the purity of the original sample. The thermo-balance and the potentiometric recorder produced continuous records of the samples weight and temperature during the experiment. The analysis of these records is described in the next section. ANALYSIS OF THE EXPERIMENTAL RESULTS

Many previous workers [2, 8-113 .have determined the position of the lime/carbonate interface within decomposing spherical carbonate samples from their weight. This can only be done if the interface is sharply defined, and if it remains spherical throughout the reaction. To show that these conditions were met, partially reacted samples were sectioned and examined. The position of the interface was made visible by painting the section with an alcohol solution of thymolphthalein, a colourless indicator which turns dark blue in the nresence of alkali. The dark blue lime layer contrasted strongly with the white undecomposed carbonate. The interface was seen to remain spherical and was sharply defined. Moreover, the sample weight calculated from the position of the interface agreed with the measured weight within the limits of experimental error. This shows that the following identity could be assumed in analysing the results: * ..*
. t .
, 9, .

r*

=L = (,*)I3 = {;;I>}13. r,

cc

(2)

tThis material is a jelly-like solid that can cast into any desirgd shape. Under pressure a liquid transmitting pressure equally in Powders can thus be compacted isotatically using a conventional die and ram.

be melted and it behaves like all directions. in Vinamould

If the reaction is chemically controlled, and if the interface moves at a constant speed, (m*)j3 should vary linearly with time. However, graphs showing the variation of (m*)13 with time were not straight lines but had a distinct S shape. Figure 2 shows a series of typical curves. These curves suggest that the reaction is not controlled by a chemical step at the interface, but that transport phenomena play an important role. Attempts were made to interpret the experimental results in terms of reaction mechanisms involving the mass transfer of CO, away from the interface, with and without a chemical step at the interface. Although the diffusion of CO, appeared to be a. major rate determining factor during the latter stages of th.e reaction, none of the reaction models considered adequately explained the experimental results. It was decided, therefore, to investigate the importance of heat transfer by compacting a thermocouple at the centre of a sample, and measuring its temperature during the reaction. The sample was allowed to come to thermal equilibrium in pure CO, at 860C. Figure 3 shows that, once the reaction was started by changing the COP for air, the temperature of the sample rapidly dropped to below 82OC, climbing slowly back to its equilibrium value during the rest of the reaction. It is interesting to note that the temperature of the furnace thermocouple which was a few millimeters from the surface of the sample, gave no indication of thii variation. If the reaction were controlled exclusively by a chemical step at the reaction interface, and if the activation energy of the reaction were as high as has been reported, the temperature variation shown in Fig. 3 would cause the reaction front velocity to vary significantly throughout the reaction. However, we have seen that the reaction front velocity is almost constant, and so the reaction must be controlled by some other mechanism. The simplest such mechanism involves heat and mass transfer alone. The diffusion process through the porous lime layer is an unsteady state process involving a moving interface. However, it can be analysed as a quasi-steady state process because the

300

Thermal decomposition

of calcium carbonate

0.7

0.b

rt
0.8 -

0.4 -

0.8 -

0.a -

0.1 -

Fig. 2. Experimentally determined variation of r* (= m*3) with time. Results taken from the following runs. Run (C) 26 28 877 876 0 0.106
I I

30 874 0.154
1

35 877 0.323

41 45 52 962 829 720.5 0 0 0

(atm)
870 , I I I I

[pcollc
I

REE,,NB

SPER.5

under a pressure of one atmosphere). Murray, Fischer and Sabean [ 121 and Azbe [ 1 I] have investigated the structure of the lime layer and shown it to be extremely fine. The- mechanism of CO, diffusion through it would therefore be Knudsen diffusion, and so the rate at which CO, is transferred from the reaction interface to the bulk gas phase around the sample can be represented by an equation of the form:

810 '

I
10

I
20

I
So

I
40

I
so

I
20

I
70

R8,{ (l/r*) - 1 +(D/cyr,)}


TThis equation is not strictly accurate since the temperature appearing in the denominator should be the varying solid temperature, and since the diffusion coefficient will alter during the reaction due to changes in this temperature. This latter effect will be small and the ratio D/ O , will change insignificantly-calculations show that the change in this ratio is less than 3 per cent under the most adverse conditions. Errors in the equation are therefore negligible.

t i m *( m i n r)

Fig. 3. Temperature of sample, and inert temperature probe, during the decomposition of a 1 cm Sample in a furnace held at 863C.

rate of accumulation of CO, in the pores of the lime is an infinitesimal part of the total flux (the decomposition of a carbonate sphere O-5 cm3 in volume liberates about 350 cm3 of CO, at 700C
C.E.S.-B

301

A. W. D. HILLS

For the reaction to be controlled by heat and mass transfer alone the chemical step at the reaction interface must be sulhciently rapid for equilibrium to be maintained there. If this is so, the partial pressure of CO, in the gas phase at the interface, [PC&, can be predicted from the reaction front temperature, and from the heat and entropy changes of the reaction. The average rate of rise shown by the temperature curve in Fig. 3 is 0*X min+. With such a slow rate, only negligible temperature gradients can exist within the undecomposed carbonate sphere enclosed by the reaction interface. The temperature indicated by the thermocouple at the centre of the sample is thus the reaction interface temperature, and can be used to calculate [PC&. t Since a and r* could be calculated from the continuous trace of the samples weight, all the variables appearing in Eq. (3) could be determined excepting the diffusion and mass transfer coefficients. The equation can be rearranged in the form:

[PCOSIR - rPcorlc =~[(-.gh


and so a graph of

I}+-$
(4)

Figure 4 shows a typical plot obtained. Remembering that the differentiation of experimental results is not very accurate, the linearity of these plots is sufficient to confirm that the chemical step at the reaction interface cannot play any significant role in determining the rate of the reaction. This being so, these graphs can be used to determine the diffusion coefficient-from the slope; and the mass transfer coefficient -from the intercept. The experimental results can also be used to determine the thermal conductivity of the lime layer, and the heat transfer coefficient to the outer surface of the sample. The heat of decomposition of calcium carbonate is about 40,000 cal g-mole-, and so 400 cal must be provided for each gramme of decomposing carbonate. Certainly no more than 10 cal for each gramme of the original carbonate can be involved in heating the sample during its slow temperature rise. Very little error is introduced, therefore, if we assume that heat transferred to the reaction front is all absorbed by the reaction, and if we analyse its transfer in steady-state terms. Thus we can write: i= -Hi (5) and 47&(8,-e&, 4=(llr*)-l+(k/h$)* (6)

l&& -. rhslG

These equations can be rearranged in a similar way to Eq. (3). As a result, we have: ~=&..(--$)Y}+&. (7)

should be a straight line. Graphs were prepared from the results of each run, several values for the rate of reaction, ir being obtained by differentiating the weight of the sample numerical1y.S
tPrevious me$surements of the equilibrium decomposition pressure of calcium carbonate were found to be insufficiently accurate for this and so it was necessary to make independent equilibrium experiments on samples identical to those used in the reaction rate experiments [ 131. &Stirlings central difference formula was used for this since it is of greater accuracy than any Corresponding formula involving simple difference[ 141.

Thus the thermal conductivity of the lime layer can be determined from the slope obtained by plotting (0, -&)/iz against {(l/m*)13-1}, and the heat transfer coefficient can be obtained from the intercept. Figure 5 shows a typical plot of Eq. (7). The heat transfer results are more scattered than those for mass transfer largely due to the relative unimportance of the thermal resistance of the lime layer. The four transport parameters were measured

302

Thermal decomposition
I
12 eg l o1 10 x [F t%l$%je 0 t . m if Y g m-l e 7 4 c

of calcium carbonate
I I I
(r n -

I
0 . 0 2)

I
x /_

p== 0 134 A t m. /
Xix

/_

e 6/

4X

*4

*A

& 2$

:. O

2!4

Fig. 4. Plot of ([PC,&-

[P,,],)/ri

a g a i n s t (l/m*)lrJresults of run 0.

I} calculated from the

2 . 8R U N 33

2 . 4-

$Ii

x 1 0-3 C.Min.g d

1 . 2-

0.8 -

0.4

. t

Fig. 5. Plot of (&-&)/A

against { (l/m * ) Is- I } calculated from the results of run 33.

303

A. W. D. HILLS

304

at various gas temperatures and compositions, and in samples of different diameters. The only significant variation was found to be that between the diffusion coefficient and the gas temperature. The actual nature of the relationship could not be ascertained from the results and so a linear correlation was used. The diffusion coefficient results are plotted in Fig. 6, and the mean values of the other transport parameters are set out in Table 1. Values for the mass transfer parameters could be determined from the results of 30 experiments, but only 24 could be used to determine the heat transfer parameters. The measured values of the diffusion coefficient were found to be correlated by the following equation: where A = O-083+ O-002cm*sec-l
B = ONI

f O.oooO3 cmzsec-lC-l.

The results were obtained with 28of freedom. The mass transfer coefficient makes very little contribution to the overall mass. transfer resistance, and the thermal conductivity makes very little contribution to the thermal resistance. As a consequence, these parameters were not determined very accurately and their standard deviations are large. The shape of the temperature/time curve for the sample is a further consequence of the small contributions made by these two paramters. The heat transfer rate is largely determined by factors outside the sample, whereas the mass transfer rate is determined by diffusion within the lime layer. Further evidence that the reaction is exclusively controlled by heat and mass transfer
Table

is provided by the agreement between values of the transport parameters determined from the results, and values determined from the results of independent experiments reported in the literature. The two sets of values are compared in Table 1, and it can be seen that the agreement is within the limits of experimental error in all cases. (Calculation of the literature values is described in the appendixi and was carried out for the mean temperature of the experiments830C.) The linearity of mass transfer plots such as that shown in Fig. 4, and the agreement between the measured values and literature values for the transport parameters, constitute strong evidence that the reaction is controlled by heat and mass transfer alone. Further evidence is provided by the accuracy with which theoretical models based on this mechanism predict the actual course of the reaction, arid their ability to explain the results previous workers. The derivation of these theoretical models is described in the next section, and their predictions are then compared with the experimental results.
THE THEORETICAL MODELS REACTION FOR THE

Derivation of models

The formal mathematical problem presented by the proposed heat and mass transfer mechanism is an extremely intricate moving boundary problem. Its rigorous solution would entail the use of finite difference methods to solve the differential mass and heat balance equations within the lime phase, and the differential heat balance equation within the carbonate phase. The packed bed problem could not be

1. Comparison between experimental and literature values for the transport


properties calculated at 830C

Transport parameter

Experimental 8.36 2 0.22 2.122 I.67 2.39 2 I .26 246~0.39

Literature 8.5 3.60 I.7 l-99

Source Appendix Appendix Tadokaro [ 151 Appendix

D X IV (cm%set-*)
a (cm set-I)

k X 10s(4 h X IOa(4

cm- set%?) cm-* sec-LC-l)

305

A. W. D. HILLS

tackled if such involved mathematical methods were required to predict the behaviour of single spheres, and so a simpler approach has been used. Three approximate models have been derived, in order to examine the effects of the different simplifying assumptions. In the first model, the heat and mass transfer processes are analysed in steady-state terms, and the thermal mass of the sphere is ignored. The CO, pressure at the reaction front is assumed to be a linear function of temperature, thus enabling explicit equations to be derived for the progress of the reaction. This model is termed the linear pressure model. The second model only differs from it in that the true exponential relation is used to link the CO, pressure at the reaction front to the temperature there. This model is termed the exponential pressure model. The third model makes approximate allowance for the unsteady nature of the heat transfer process by assuming that the

The equilibrium relationship provides another equation linking these two variables. When the temperature difference between the reaction front and the gas phase is small, we can use the following linear form:

r_PCO*lR = bcorlea+4&t--G)
where

(9)

Eliminating [pCOJR between Eqs. (8) and (9) gives the following equation for the temperature of the reaction front:
0,-e,=

bco!2le*-bcozIc
(l/r*)-l+(D/arJ RfI&+, I (l/r*) - I+ (k/hr,) I HD (I)

Using this equation to eliminate 0, - 0, from Eq. (6) and substituting into Eq. (5) gives the reaction rate as: (12)

bcalt?*- bcalc = l/4~ro{RB~/D[(l/r*)-1+(D/arro)]+aH/k[(l/r*)-1+(k/hr~)]}

thermal mass of the sample is concentrated at the reaction front. It is termed the semiunsteady model. Numerical methods are needed to solve the ordinary differential equations involved in both the last two models. The three models are described below. 1. Linear pressure model. Since the thermal mass of the sphere is ignored, Eq. (5) can be used to relate the heat transfer rate given by Eq. (6), to the rate of reaction given by Eq. (3). Thus the CO, pressure at the reaction front, must be related to the temperature there, by the following equation:
CPCOJR- CPco21c = -{ y&.y_:y(y;;;]

which can be rearranged

~mUk~~l~~- bcoJ~~ (aHik)l{(llr*)-l+r~j~~~~~~~~~~k,l)

(13)

A= [(R&/D)+

This equation can be simplified in form if we define two new transport parameters: A diffusive parameter, l? 1 1 -_=_+- aH (14) I- D R&k and a convective parameter, A: 1 1 aH -_=-+-_. A a R&h Equation (13) then becomes 4rrJ{
CPCOJ~~- [PCOJGI

(15)

=R&{(l/r*)-l+(r/rOA)}

(16)

(8)

If the reaction were controlled by mass transfer alone, the temperature of the reaction front would be that of the gas phase around the sample.

306

Thermal decomposition of calcium carbonate

Thus the CO* pressure at the reaction front would be the decomposition pressure at the gas phase temperature, [pc,JeW The rate of reaction would then be given by Eq. (3), except that [pcO& would replace [PCOnlR. Comparison between Eq. (3) and (16) shows that the transport parameters r and A are apparent diffusion and mass transfer coefficients respectively and that their use combines the effects due to heat transfer and mass transfer. Combining Eq. (16) with the conservation equation, Eq. (l), gives an ordinary differential equation for the fractional radius of the reaction front. Using the relationship between r* and m*, the solution of this equation gives: f(m*) = 3m*21a-2m*-2$(1 -m*)

mechanism suggested here with the mechanisms suggested by previous workers. 2. Exponentidpressure model. This only differs from the previous model in that the true exponential equilibrium relation, Eq. (18), is used instead of Eq. (9). Eliminating [pVo2],, between Eqs. (18) and (8), gives a non-linear equation for the temperature of the reaction front. The algebra is simplified if Eq. (18) is first rearranged:

= exp

H --_ Rec

Re, I bC*lw4.

(22)

The temperature of the reaction given by the equation:

front is then

= 1_ fW~co,le,- bcozldf R&pCaro2 = l-r*.

(17)
(l/r*) - I + (k//u-,,)

The values of T and A can be evaluated by differentiating the true exponential equilibrium relation.

In terms of the dimensionless variables defined in the list of symbols, this equation becomes:

hosIR = Yew
Thus

I$--$-}.

(18)
I -r*(l

a=

{~),=,,= PO[ {&$j--]]$ exp


(19) are given by:

-BIJI) I-r*(l-BiM) .

k*(I),*-&*)

= 0. (24)

The two transport parameters 1 1 -_=-+ I? D

bco*le~2
RRe$k

WV
(21)

1 1 -_=-+ [f%o*lt?c#2 a RRe$A A

Newtons method can be used to solve this equation to give t&* as a function of P, once the value of p* is set. There are no anomalies in the equation and so its numerical solution is straightforward. Combining the heat transfer equation, Eq. (6), with the conservation Eqs. (1) and (5), gives the following differential equation for the rate of advance of the reaction front {r*-r*2( 1 -BiH)}T= kficir:). (25)

This model has the great advantage that it leads to a simple equation for the progress of the reaction which forms a convenient basis for comparing the heat and mass transfer

In order to compare the numerical results obtained with this model and those obtained from the previous model the same dimensionless 307

A. W. D. HILLS

time is used. In terms of the variables defined in the list of symbols, Eq. (25) becomes:

&*-OR*

(l/r*)-l+&
- exp((1/8~*)-(1/8R*))-p*

{r*
(26)
separating the variables gives
t*=-6

k*{(l/r*)-l+BiM}

(30)

I
1

{r*-r**(l-BiH)}(l-p*)r*dr*
k*(&;*-eR*)

(27) The integrand appearing in this equation is a function of r* only, since f&* is given by Eq. (24). Thus the integral can be evaluated by numerical quadrature to give t* as a function of r*. A range of values is required and so the quadrature is best performed using Simpsons rule. The results can be expressed to show r* as a function oft*. Should the temperature of the reaction front be required, it can be calculated from Eq. (24) as a function of r* and also expressed as a function of t*. Figure 8 shows temperatures calculated in this way. 3. Semi-unsteady model. For the previous model to predict the progress of the reaction accurately, the reaction front temperature would have to change in a step-wise fashion at the beginning of the reaction. This it cannot do since the thermal mass of the sample slows down its rate of response to the temperature changes imposed by the reaction conditions. To examine the importance of this effect, the semi-unsteady model was developed. The thermal mass of the sphere was taken into account, but it was assumed to be concentrated at the reaction front. -Thus the heat balance equation gives:
4 =

The specific heats of lime and calcium carbonate do not vary greatly with temperature. Over the temperature range 600-900C the variation is about 3 per cent [ 161.Thus they may be assumed to be constant, and the thermal mass of the sphere is:
mTh = +w?{ ( Ir*? x PCaOCCaO

(31) The L.H.S. of Eq. (30) can be rearranged in terms of the previously defined dimensionless variables, to give:

2ccaot bco*lea- [PCOIIGI


R&k
x

(l__r*3)

!s!!2+r*SPChX~}~.
Pea PC02

Thus Eq. (30) becomes:

2Gaor*

R$l

-P*)

eG*- eR*

= (l/r*)-l+&
-

-Hti-

mThT

d@R

exp weG*)-wR*)h* k*{(l/r*)-l+&}

(32)

(28) Since the molar densities of the species are


equal, Eq. (32) can be rearranged to give: de,*

Using Eqs. (3) and (6) we have


hi -mTh% I (l/r*)e, -- I+& =4?rkr, I

At* _.

ec*-eR
= (l/r*)-l+&-

exp {i/e,* - i/e,*} -p* k*{(l/r*)-l++i&


*

In terms of the dimensionless variables defined in the list of symbols, this becomes: 308

2+~(1--p*){l+r*3[~-l]}

(33)

Thermal decomposition of calcium carbonate

Combining the mass transfer Eq. (3) with the conservation Eq. (1) gives:

of the reaction front. They can be solved by the method due to Runge [ 171.
Comparison between theory and experiment

In terms of the dimensionless variables, this equation can be rearranged to give: dr*
dt*=-

exp {l/O,* - l/e,*} -p* 6x (l-p*){r*-(1-&&**I.

(35)

Equations (33) and (35) constitute a pair of simultaneous non-linear differential equations for the fractional radius and the temperature
I

Errors resulting from the approximations made in the three models will become more pronounced as the reaction temperature is increased. Thus the most meaningful comparisons between theory and experiment are to be made at the highest reaction temperatures. Such comparisons are made in Figs. 7 and 8. The fractional mass is plotted against dimensionless time in Fig. 7, and Fig. 8 compares reaction front
I
I I I I

---I+ X

Semi- unsready model Exponenlid pressure model Linear pressure model Experinnntal results

Fig. 7. Variatior of fractional mass with dimensionless time-experimental run 4 1 and theoretical predictions of the three models.

results for

309

A. W. D. HILLS

Semi-unsteady model ----Exponential pnrrure model X Exprlmentol nrultr 820 98 - 909c

800

780

760

740 Fig. 8. Variation of reaction front temperature with dimensionless time-experimental results for run 41 and theoretical predictions of the exponential pressure and semi-unsteady models.

temperatures predicted by the two more complicated models with the measured temperature at the centre of the sphere. The experimental results are taken from run 4 I, in which the sweep gas was pure air at 902C. Values of the parameters used in the calculations were taken from the same run. The fractional mass curves predicted by the exponential-pressure and semi-unsteady models are extremely close together, and bracket the experimental results. It is obvious that a more intricate theoretical model is not warranted. The semi-unsteady model predicts a higher reaction front temperature than the exponential-pressure model, owing to the effect of the thermal mass. Thus it gives a slightly higher initial reaction rate ~ in Fig. 7. This model over-estimates the effect of the thermal mass, whereas the exponential
310

pressure model ignores it. Thus we would expect the reaction front temperature and therefore the reaction rate, to fall between the predictions of the two models. That this occurs can be seen in Fig. 7. The temperature in the centre of the sample will be above the reaction front temperature in the initial stages as is shown in Fig. 8. The theoretical and experimental temperatures come together towards the end of the reaction. It is during these stages that the property values are measured. That the theoretical fractional mass curves calculated from these property values follow the experimental results so closely during the entire reaction is additional confirmation of the proposed reaction mechanism. The curve predicted by the linear pressure model in Fig. 7 is somewhat removed from the

Thermal decomposition of calcium carbonate

experimental results, but the overall reaction time it predicts is only 10 per cent in error. We would expect the accuracy of the linear pressure model to improve as the temperature difference between the furnace and the reaction front is decreased. This will occur when the reaction temperature is decreased, or when the CO, partial pressure around the sample is increased. Under these conditions, the fractional mass function defined by Eq. (17) will vary in an increasingly linear manner with time. That this is so can be seen in Figs. (9) and (lo), which respectively show the effects of decreasing the reaction temperature, and increasing the COz partial pressure. Experimental results are plotted in these figures, together with the theoretical predictions of the semi-unsteady model. These curves show that the two more complicated theoretical models predict the decomposition rate of calcium carbonate to a high I.0

degree of accuracy. They also show that the simple linear-pressure model, which has the great advantage of -yielding an analytic expression, is sufficiently accurate to be used in the next section of the paper in comparing the reaction mechanism proposed here with the mechanisms proposed by previous workers.

COMPARISON

WITH

PREVIOUS

WORK

This work has shown that the decomposition of calcium carbonate takes place on a receding reaction interface, and that the reaction rate is controlled by the transfer of heat to this interface, and of CO, away from it. This conclusion rests on the following evidence: (1) Sections of partially reacted compacts. (2) The large temperature drop experienced by the compacts as soon as reaction commences.

+RonS2:8p-721F

pr p* p*

I 0. - 0 . I 0 .

*s
*6 -

O Run

46 : e,-S2oc 41:8,=002*C

XRun

*4 t I 5 (m*) *2 -

O-

_. 2-

-. 49

Fig. 9. Variation of fractional mass function with dimensionless time-experimental results for runs 41,45 and 52 and theoretical predictions of the semi-unsteady model.

311

A. W. D. HILLS

l*O
THEORETICAL ----.*-**--.--1--1--1-IX Run

-0

EXP ERIMEN TAL no.28 : 8$878% no. 28 : 6++876.C no. 35 : e,=878%

p*10
p*= 00151

Run

-6

-----

+ Run

pw- 0.48i

l 4

s I 02 ( m *) 0

-. 2

-. 4

_. 6

Fig. 10. Variation of fractional mass function with dimensionless time -experimental results for runs 26,28 and 35 and theoretical predictions of the semi-unsteady model.

(3 1

The linearity of plots relating the rate of reaction, divided by the CO, partial pressure difference between the reaction front and the bulk gas, to the diffusion resistance through the porous lime shell. (4) The agreement between relevant transport parameters determined from these plots, and values determined from independent experiments reported in the literature. (5) The extremely good agreement between the theoretical predictions for the progress of the reaction, and the experimental results.

The proposed reaction mechanism is considerably different from the mechanisms proposed by previous workers who have concluded that the reaction rate is determined by some step of a chemical nature occurring on the lime carbonate interface. It is necessary, therefore, to review the evidence on which these workers have based their conclusions. The chemical control models stem largely from the constancy of the interface velocity, but the effects of temperature, and CO, partial pressure, on this velocity have been taken to provide additional confirmatory evidence. As Eq. (1) shows, the interface velocity can only be constant if the reaction rate is pro312

Thermal decomposition of calcium carbonate

-L

313

A. W. D. HILLS

portional to the interface area. Previous workers have tacitly assumed that a transport controlled reaction cannot show this characteristic, and have therefore concluded that the reaction must be chemically controlled. However, it has been shown by Warner[ 181, and by the present author [ 191, that the rate of a diffusion and mass transfer controlled reaction can be proportional to the interface area. Equation (17) can be used to show that this is also true for a reaction controlled by heat and mass transfer. This equation shows that the fractional radius of the reaction front is given by a relation of the form:

r* =

(m*p=f[t*,A}. * 0

(36)
.04l 680 700
1 I I

Thus we can consider Tlr,A to be a parameter whose value determines the shape of plots of r* against t*. Figure 11 shows a series of such plots calculated from Eq. (17). When I/roA is small, the solid phase resistance is most significant, and the curves have a pronounced S shape. When it is large, convective resistance is most important, and the curves have a uniform curvature-concave downwards. However, for intermediate values, the curves are substantially straight for much of the reaction. When I/roA is 6 the curve is straight until I* = 0.4, at which time m* is 0X164. Under these conditions the reaction interface moves at a constant rate for the first 94 per cent of a reaction controlled by heat and mass transfer. It is therefore incorrect to conclude that such behaviour is symptomatic of chemically controlled reactions. Figure 12 shows the variation of I?, A, and F/r,-,A with temperature for the decomposition of calcium carbonate. These curves have been calculated from the mean property values set out in Table 1, except that the mass transfer coefficient has been calculated from the equation of Rowe, Claxton and Lewis[20]. We can see that T/roA varies between O-15 and O-4 as the gas temperature varies between 780 and 900C. Figure (11) shows that we would expect substantially linear r* plots with these values, and it is this fact that has mislead previous workers into concluding that the decomposition of

I
860 900

740

780 e,v

820

Fig. 12. Variation of transport parameters with temperature.

calcium carbonate is chemically controlled. The value of A varies with the heat transfer coefficient to the surface of the sample. As will be shown in the appendix, this coefficient is largely determined by the radiation conditions within the furnace. Thus the value of F/r,A, and consequently the shape of the r* curves, will vary from apparatus to apparatus. Slight nonlinearities in the r* curves have therefore been more apparent in some investigations than in others e.g. [2 11. Had more workers measured the actual temperature of decomposing carbonate samples, the chemical control explanation for the reaction rate would not have gained such universal acceptance. The temperature variation shown in Fig. 3 proves, in itself, that the constant interface velocity. cannot be explained in terms of a constant rate chemical step at the interface. However, most workers have assumed that the temperature of the sample is equal to that of the furnace, and have calculated an Arrhenius activation energy for the reaction by investigating how the interface velocity varies with this temperature when the CO, partial pressure in

314

Thermal decomposition of calcium carbonate

the furnace is zero. The activation energy is defined by the equation: (37) Values obtained have generally been about 40 kcals g-mole-l, and have been assumed to be too high for a transport controlled reaction. This fact has been taken as additional evidence that the reaction is chemically controlled. However, the activation energy determined in this way is virtually equal to the heat of the reaction,t a fact which helps to cot&m the reaction mechanism proposed here. Equation (17) shows that the total time for the reaction is given by

6N~c0,1,R ~cPcoJo2

bco,lcIt = 1+2 r
r

r ob

(38)

Dividing the radius of the sample by this total reaction time gives a mean value for the interface velocity. If we put Cpc& equal to zero, and substitute for [pco3]eqfrom Eq. (18), we get:
[uR1e=~Re&.&x

[;:*(rhlA)J
(39)

Xdexp(i-&).

As the gas temperature increases from 700 to 9OOC, the expression in curly brackets in this equation decreases from 1.40 X 10m4 to 0.95 x 10v4cm s-atm-I. Over the same temperature range, the decomposition partial pressure increases from 0.03 to l*Oatm. Since Fig. 9 shows that Eq. (39) underestimates the velocity of the reaction front at high temperatures, we can see that the variation of this velocity with temperature is all but entirely due to the change in the decomposition pressure. Thus the activation energy determined from Eq. (37) is virtually equal to the heat of the reaction- the result obtained by most previous workers. Satterfleld and Feakes [4], however, have obtained a much higher value, 360 kcals g-mole-. Their work deserves special mention, since they
tThe heat of the reaction = 39.1 kcal g-mole- [ 131.

concluded that the reaction was chemically controlled, although they did measure temperatures within their decomposing samples. Their experiments were conducted in a furnace containing one atmosphere pressure of C02. Under these conditions the equilibrium partial pressure of CO, changes so rapidly with temperature that very little temperature change occurs during the reaction. The Arrhenius plot that results is thus not very accurate, and this accounts for the unusually high activation energy obtained. However, Sattertield and Feakes adopted a chemical control explanation for the reaction largely because transport considerations were unable to account for the high reaction front CO, pressures calculated from the measured temperatures of their samples, and the thermodynamic data of Johnston[22]. Moreover, they found that samples prepared from freshly precipitated calcium carbonate decomposed at lower temperatures than samples prepared from commercial grade reagent. It has been shown, however, that decomposition pressures predicted from the thermodynamic data of Johnston are too high[23. 131, and that equilibrium pressures of calcium carbonate obtained from different sources differ [24]; freshly precipitated carbonate giving higher decomposition pressures [25]. Thus the arguments advanced by Satterfleld and Feakes to support their proposed heat transfer and chemical control mechanism are somewhat suspect. To test the validity of this mechanism, however, some experiments were performed with spheres of different sizes. Were the reaction to be controlled by heat transfer, and by a chemical step at the reaction interface, the reaction rate per unit area should vary with the reaction front temperature according to the Arrhenius equation:
if=noexp

I I
~e,

-E

(40)

A plot of Ln(n) against 1/OR would be a straight line, the results obtained with different sized spheres all falling on the same straight line. Figure 13 shows results obtained with 1 cm and 2 cm spheres. It can be seen that the results do

315

A. W. D. HILLS
4.2 I I I \ -4.3 *xT, X. -4.4A \ B / . x x\ X 1 I I I I I

x RUN N9 41. r,~OGbcm. 0 RUN N9 61. I-~- 1.14cm.

X \K ;cPL *-y_X_.ax~

-4.3 T *,+a *E cu LO $4.7 v -4.8

k h \o h

-4.9

b k 0 0~;-+0--0I 0.7 hR SC I a.75 * 104 I 8.0

P I 8.85 I 8.9 I 8.95

-so-

8.5

l 8.55

I 8.6

I 8.65

Fig. 13. Arrhenius plots for experimental results obtained with spheres 1 cm and 2 cm dia.

not fall on a single straight line; the results obtained with the different sized spheres fall on completely separate curves. This figure shows that the reaction mechanism proposed by Satterfield and Feakes cannot be correct,, The reaction front velocity at any one temperature, de&eases when the partial pressure of CO* around the sample is increased. Both Ingraham and Marier[2] and Hyatt, Cutler and Wadsworth [3] claim that the relationship between these variables confirms reaction mechanism that they propose. It is not possible, however, to distinguish between chemically controlled and transport controlled reactions in this way. Were the reaction to be controlled by mass transfer or by a first order chemical step at the reaction interface the reaction rate would vary linearly with partial pressure, as reported by Ingraham and Marier. On the other hand, the non-linear relationship reported by Hyatt, Cutler and Wadsworth, could be taken to indicate that the reaction was controlled by an adsorption mechanism of the type they propose, or by heat

and mass transfer. In fact, as Hashimoto[ lo] reports, the reaction rate decreases linearly with CO* partial pressure at low temperatures, but the relationship is non-linear at high temperatures. The reason for this can be appreciated from Eq. (17). This equation is accurate at low temperatures when it shows that the reaction rate is directly proportional to [Pc&JeP [pcol] w As Figs. (9) and (10) show the equation is inaccurate at high temperatures, but its accuracy is increased as [p,&jc increases. Thus the reaction rate varies non-linearly with CO, partial pressure at these high temperatures. This section shows that the evidence advanced by previous workers in support of chemical control mechanisms for the reaction can be easily explained in terms of the heat and mass transfer mechanism advanced in this paper.
CONCLUSION

This work has shown that the decomposition of calcium carbonate is controlled by heat and mass transfer and not, as reported by many

316

Thermal &composition ofcalcium carbonate

previous workers, by a chemical step at the reaction interface. These workers have based their conclusions on certain characteristics of the reaction that have been assumed to prove that it is chemically controlled. These assump tions were incorrect, and the proposed heat and mass transfer mechanism has been shown to explain how the reaction displays these characteristics. Mathematical equations based on the heat and mass transfer mechanism have been shown to predict the rate of decomposition to a high degree of accuracy. These equations should enable packed bed decomposition rates of calcium carbonate to be predicted
AckwwIs&ms~~-This work formed part of the authors Ph.D. investigation which was supervised by Professor A. V. Bradshaw. It gives the author great pleasure, thesefore, to thank Professor Bradshaw for the academic assistance and moral support that he provided during the course of the investigation. The author also wishes to thank Mr. A. J. Haynes of the John Percy Group and Mr. J. Whitcroft of the Stanton Instrument Company for their technical advice, and Professor Bradshaw and Professor F. D. Rich&&on for providing iinancial support. NOTATION

m o,m t,

moo

a A B

G
D h

h cow

hRad
H

rate of change of equilibrium pressure with temperature constant in correlation equation for diffusion coefficient gradient in correlation equation for diffusion coefficient molar specific heat of species (i) diffusion coefficient of CO, through porous lime heat transfer coefficient between the outer surface of the carbonate sample and the surroundings convective heat transfer coefficient to outer surface of sample radiation heat transfer coefficient to outer surfaceof sample molar heat of decomposition of calcium carbonate thermal conductivity of porous lime layer chemical rate constant 317

initial, interim and 6nal mass of a decomposing sphere thermal mass of decomposing sphere molecular weight rate of decomposition of carbonate sphere rate of reaction per unit area standard pressure-one atmosphere partial pressure of CO* in equilibrium with lime and calcium carbonate at the temperature of the furnace gases actual partial pressure of CO, in the gas phase partial pressure of CO, at the reaction front within a decomposing sample heat transfer rate to decomposing sample gas constant in mechanical units ( cm8 atm g-mole+ K-l) gas constant in heat units (cals g-mole-l K-l) radius of decomposing sample radius of reaction front specific surface area in porous lime layer entropy change of decomposition reaction time total reaction time average velocity of reaction front at temperature 0 mass transfer coetlicient to surface of carbonate sample voidage in a porous solid diiTusive parameter
=

[A+ $$$-l

emissivity of lime gas temperature temperature of reaction front temperature of sample temperature of furnace walls convective parameter
r.Pcoal~

RRQh

-l
I

c.es.-c

A. W. D. HILLS

Pt molar density of species V in original carbonate sphere apparent density of a porous solid PP Stefan-Boltzman constant u tortuosity factor in a porous solid 7

cc viscosity

mt-moo sphere = mo-m, P*

[PCOJC fractional pressure = [p coeQ

Dimensionless variables BiH Biot modulus for heat transfer X- k hr, BiM Biot modulus for mass transfer =- D
ar0

Pr Prandtl number r* fractional radius of reaction front = r/r0 Re Reynolds number SC Schmidt number Sh Sherwood number t* dimensionless time = 6N_~c,l,ho&It r* = l-/D eG*, eR* dimensionless temperatures =--Re, Re, HH NU Nusselt number

RfUwo2

k*

RW RDbco,lCp

m*

dimensionless mass of decomposing

REFERENCES HILLS A. W. D., Thesis, Londoq University 1966. INGREHAM T. R. and MARIER P., Can. J. Chem. Engng 1963 41170. HYATT E. P., CUTLER I. B. and WADSWORTH M.E.,J. Am. Cerum. Sot. 1958 4170-74. SAITERFIELD C. N. and FEAKES F-,/l. I. Ch. E. Jll959 5 115. AZBE V. J.. Rock Prod. Stone Edn 1944 47 July, 53. August, 70. HASLAM k. T. and SMITH V. C., fnd. En& Chem. 1928 20 170. NARSIMHAN G., Chcm. Engng Sci. 1961617. SPLICHAL J., SKRAMOVSKY St. and GOLL J., Cohn Czech. Chem. Commun. 1937 9 302. FREEMAN E. S. and CARROLL B. J., J. phys. Chem. 1958 62 394. HASHIMOTO H., J. them. Sot. Japan 1962 64 1162,1166. AZBE V. J., Rock Prod. Stone Edn 1944 47 September, 68. MURRAY J. A., FISCHER H. C. and SABEAN D. W., Proc. Am. Sot. Test. Mater. 1950 SO1263. HILLS A. W. D., Trans. Instn Min. Metall. 1967 76 C241. KOPALZ., NumericalAnalysis, p, 49. Chapman&Hall 1955. TADOKARO Y., Sci. Rep. Tohoku Uniu.Japan 1921~22 10,339 KUBASCHESKI and EVANS, Metallurgical Thermochemistry. PergamonPress 1956. KOPALZ.. NumericalAnalysis, p, 211. Chapman&Hall 1955. WARNER N. A., Aus. 1. M. M. Proc. 1964 210 3 1. HILLS A. W. D., Heat and Mass Transfer in Process Metallurgy._ Hills_ W.__ _(Ed.), _ . 39. I.M.M. 1967. A. D. _ . _ _ p. ROWE P. N., CLAXTON K. T. and LEWIS J. B., Trans. Instn them. Engrs 1%5 43 114. [21] JOSEPH T. L., BEATTY H. M. and BITSIANES G., TransAIMME 1943 154 148. [22] J0HNST0NJ.,J.Am.Chem.Soc.1910321910. 1231 BAKER E. H., J. them. Sot. 1962 87 464. 1241 BAKER E. H., private communication. 1251WHITING G. H. and TURNER W. F. S., J. Sot. Gfuss Technol. 1930 14 409. 1261PERRY J. H. (Editor), Chemical Engineers Hundbook, 4th edn, p. 14. McGraw-Hill 1963. [27] McADAMS H. W., Heat transmissions. McGraw-Hill 1954. 1281 HIbD K., Mitt. K.-Whelm-Inst. Eiserlforsch. Dhsseld. 1932 14 59. [29] SAlTERFIELD C. N. and SHERWOOD T. K., Role of Digusion in Cut&s& p. 17. Addison-Wesley 1963. [1] [2] [3] [4]

318

Thermal decomposition

of calcium carbonate

APPENDIX
CALCULATION OF TRANSPORT PARAMETERS FROM THE LITERATURE Certain transport parameter values calculated from experimental results reported in the literature are shown in Table 1. The calculation of these values is summaiised here. Convective heat and mass transfer coeflcients Table A.1 shows the values used in calculating the convective heat and mass transfer coefficients to the sample. Table A. 1. Calculations of convective heat and mass transfer coefficients Total gas flow rate Diameter of furnace tube Density of air at NTP Viscosity of air at 832C Diameter of sphere Reynolds number Thermal conductivity of air Diffusion coefficient of CO, in air, calculated from the Gilliland formula 1261 1 1. min- 5cm 1.29 g I- 4.48 X lo- g see-* cm-l 1.1 cm 2.4 1.7 X lo- cal cm-set-C- the sample. The sample enclosure can be treated as an isothermal enclosure whose surface area is very much greater than that of the sample. Under these conditions, it can be shown[27] that the radiation heat transfer coefficient is given by hRP,j ue(ed+ = e,r) (ew+ es). (42) Hild[28] has measured the emissivity of lime. At 900C he obtained a value of O-27,but this value drops with temperature. Thus we may take the emissivity of lime at 832C as 0.2. With this value we get: cals [hwliw = 146 X lOma cm-zsec-lC-l. Thus the total heat transfer coefficient is: hmP= 199 X 10m5 cals cm-*see-C-l. Diffusion coefficient in porous lime The diffusion ceefficient of CO% through porous lime is very difficult to estimate. However, Murray, Fischer and Sabean [ 121 found that lime whose porosity was 60 per cent had a specific surface area of 16 m g-l. Such a large value shows that we may estimate the diffusion coefficient of CO, through the lime from the equation for Knudsen diffusion _ 1291 DL 19,400 & (43)

1a19cm%ec-

Using these values in the formula of Rowe, Claxton and Lewis [20]: (41) and putting Pr = SC = 1, we get: a=2 = 3.60 cm set- and [hconvlans 5.3 x lo- cals cm-*set-C-l. = Total heat transfer coegicient Radiation plays a dominant role in the transfer of heat to

in which D is measured in cm*sec-*, y is the porosity, pr is the apparent density of the solid in gem-J, s is the specific surface area in cm* g-l, 0 is the absolute centigrade temperature, M is the molecular weight, and 7 is an empirical tortuosity factor. The density of the lime was about 1.3cm-3 whereas the theoretical density of time is 3.3cme3, and so the porosity is about 60 per cent. Values for 7 are very scattered, but Sattertield and Sherwood[29] show that the best value is about 2.0. Using this value with the specific surface measurement of Murray, Fischer and Sabean gives: D,,, = 0.075 cm*sec-*.

R&mm& La dkomposition de sphkres de carbonate de calcium simples frittkes a et6 Ctudib dans un equilibre thermique modifik pour permettre les calculs simultaks de la tempkrature et du poids dun dchantillon de dkcomposition. 11est dkmontrt que la rkaction de dkomposition a lieu sur une limite dCfine entre le carbonate non form6 et la couche de chaux poreuse formte B Iexttrieur. Cette limite se dkplace vers le centre du comprimt mais garde sa forme sphkrique. Le taux de r&action est contr6lC par le transfert de chaleur vers cette limite de rkaction et par le transfert de CO* loin delle. Ce mkanisme de r&&on est diffkrent du mkcanisme propok par des chercheurs pr&zCdents qui avaient suggkrk que la kaction soit contrblke par une action chimique g Iinterface de la r&action. Leurs conclusions ttaient bakes sur Iobservation expkimentale de certains caracttristiques de la rkaction qui ktaient supposdes &tre symptomatiques des &actions chimiquement contrSICes. Cet article montre que les &actions contr8ltes de transport peuvent aussi presenter ces caracttristiques, et elles ne four&sent done aucune Cvidence du contr6le de la &action par le mtcanisme. LCquilibre thermique modifit a et& utilisC pour mesurer le coefficient de diffusion et la conductivitt thermique de la couche de chaux poreuse et les coefficients de transfert de masse et de chaleur vers la surface du comprimt. Les valeurs ainsi obtenues saccordent avec celles de la litbrature. Trois mod&es de &action ont tti dtveloppks B partir desquels on peut p&dire le progrks de la rkaction dune seule sphere dans une ambiance stable. On montre que les prkdictions faites par les deux modkles les plus compliqds saccordent bien avec les rksultats expkrimentaux.

319

Zusammenfassung- Es wurde die_Zersetzung einzelner gesinterter Ktigerlchen von Kalzium karbonat unter Verwendungeiner Tbermowaage untersucht, die moditiziert war, urn die gleichzeitige Messung der Temperatur und des Gewichtes einer sich zersetzenden Probe zu ermoglichen. Es konnte gezeigt werden, dass die Zersetzungsreaktion an einer bestimmten Grenze zwischen dem unzersetzten Karbonat und der ausserhalb desselben gebildeten Schicht von poriisem Kalk stattflndet. Diese Grenze verschiebt sich gegen den Mittelpunkt des Presslings hin, behiilt jedoch kugelfdrmige Gestalt. Die Reaktionsgeschwindigkeit wird durch die W%meiibertmgung auf diese Reaktionsgrenze zu, und die Ubertragung von CO* von dieser Reaktionsgrenze weg, gesteuert. Dieser Reaktionsmechanismus widerspricht dem von friiheren Fonchem vorgeschlagenen Mechanismus, demzufolge die Reaktion durch einen chemischen Vorgang an der ReaktionstrennAbbe gesteuert wird. Diese Schliisse grtindeten sich auf die versuchsmiissige Beobachtung gewisser Reaktionsmerkmale, von denen angenommen wurde, dass sie ftt chemisch gesteurte Reaktionen symptomatisch sein. Es wird in diesem Artikel dargelegt, dass such durch Ubertragung gesteuerte Reaktionen diese Merkmale aufweisen konnen, so dass dieselben keineswegs einen Beweis fir den die Reaktion steuemden Mechanismus liefem. Die abgebnderte Thermowaage wurde dazu verwendet, den Diffisionskoeffizienten und die WimeleitfXhigkeit der porosen Kalkschichte, sowie die Massen- und W&rmetibertragungskoeffizienten zur OberlXiche des Presslings, zu messen. Die so erhaltenen Werte stimmen mit den aus der Literatur gefundenen Werten tiberein. Es wurden drei Reaktionsmodelle entwickelt, die es ermiiglichen, Voraussagen tiber den Reaktionsverlauf eines einzelnen Kiigelchens in einer gleichbleibenden Umgebung zu machen. Es koMte gezeigt werden, dass die auf Gnmd der zwei komplizierteren Modelle gemachten Voraussagen gut mit den Versuchsresultaten iibereinstimmen.

320

Potrebbero piacerti anche