Sie sind auf Seite 1von 14

0256_frame_C01 Page 1 Wednesday, August 9, 2000 6:54 AM

1 Introduction

1.1 GENERAL
In many areas of chemistry and chemical technology there exists a need to know the properties of
salts at high temperatures. These areas include chemical synthesis, metallurgy, corrosion by salts,
and catalysis.
Questions often relate to the thermal stability of the salt, usually phrased in terms of the
following questions: Up to what temperature is the salt stable? What happens to it above that
temperature — does it melt or decompose in either the solid or liquid state? What are its decom-
position products? If it is placed in a closed container, how much will the pressure build up?
This is true whether one needs to know the properties of a particular salt, or whether one wishes
to select salts fitting particular criteria, such as melting points or decomposition pressures.
The kind of information desired proceeds from the qualitative — congruent melting point or
decomposition products to the quantitative — melting points, decomposition pressures, and rates
of decomposition, i.e., thermodynamics and kinetics.
The topics covered in this book include the above kinds of information for oxyanions of the
elements which form discrete anions: carbon, sulfur, nitrogen, and the halogens including:

Carbon: carbonates
Sulfur: sulfites, sulfates
Nitrogen: nitrites, nitrates
Chlorine: chlorites, chlorates, perchlorates
Bromine: bromites, bromates, perbromates
Iodine: iodates, periodates

In order to avoid complicating factors (the subject is already complicated enough) coverage
has been restricted to anhydrous salts and monatomic cations, e.g., no ammonium and acid salts
are included. Even with these restrictions, nearly 300 salts fit these criteria.
There is also a large body of literature dealing with additives, such as catalysts, the effect of
radiation and mechanical treatment, which affects both the surface and internal structure of crystals,
which are not treated in detail in this book.
Between 1968 and 1972 the author published four monographs covering this subject under the
auspices of the National Standard Reference Data Program (NSRDS) (see below). This material is
no longer readily available and needs revision and updating.
This book generally follows the format of the monographs. Each class of salts, e.g., sulfates,
is treated in a separate chapter. Since a major question is always “Does this salt exist?” each section
includes a periodic table which provides this information. A blank space does not mean the salt
does not (or cannot) exist, but only that nothing is known of its existence, perhaps because no one
has tried to synthesize it.
The discussion of each salt includes both qualitative and quantitative information, what is
known of its behavior above ambient temperature, and quantitative data: phase transitions, densities,
and thermodynamic properties. When possible, high temperature properties, enthalpies and entro-
pies at 298.15 K and above, are listed. When these are available, they are used to calculate
equilibrium constants and partial pressures of the component gases. (For a discussion of this topic

© 2001 by CRC Press LLC


0256_frame_C01 Page 2 Wednesday, August 9, 2000 6:54 AM

see Chapter 1.2). Kinetic data for decomposition reactions, such as rates and activation energies,
are also included, but this information is inherently less reliable since these parameters depend not
only on the nature of the material, but also on its physical state, e.g., particle size and defect
structure. This topic is also discussed in a separate section.

1.2 EQUILIBRIUM CALCULATIONS


The main method for deciding what happens to substances when they are heated is thermodynamic
calculations. Before they can be carried out, however, the relevant decomposition reaction(s) must
be known. For some substances this is straightforward; for example, the decomposition of a carbonate
always results in the formation of a metal oxide, nearly always in the solid state, and CO2. For other
classes of salts the problem is more difficult. For sulfates the gaseous decomposition products may
be SO3, or mixtures of this gas, SO2, and O2, depending on the equilibrium SO3 = SO2 + 1/2O2 at
the temperature of interest. Some oxyhalogens decompose to halides, others to oxides, and some to
a mixture of both. Correspondingly, the gaseous products will also be different in the two cases.
Thus it is essential to decide first what the reactions are for which calculations are to be performed.
In most of these cases, where simultaneous equilibria obtain, calculations of partial pressures usually
require the solution of simultaneous equations. Examples will be given in the appropriate sections.
In order to calculate the equilibrium constant for a particular reaction, thermodynamic functions
of reactants and products as a function of temperature must be known. The calculation of these
functions is discussed in detail in the work of Wagman et al. cited below.
The G of the reaction is then calculated in one of two ways: (1) the appropriate addition and
substraction of Gf for reactants and products, or (2) the calculation of Gibbs energy functions
for reactants and products from enthalpy and entropy increments

0 0 0 0
G T  H 298 H T  H 298
gef  ------------------------
T
-  ------------------------
T
-  ( S0  S0 )  S0
T 298 298 (1.1)

Finally, the equilibrium constant is calculated from

0
GT = Tgef + H 298 and (1.2)

G = RT ln K (1.3)

These methods were used in this work when necessary. However, in many cases the equilibrium
constants could be calculated by a computer program developed by Outukompu Research, which
uses data stored in their program. These data are usually the most recent available. References used
in this program are listed below. When data above 298 K were not available, data at 298 K are
given. These are taken from Wagman, D. D.; Evans, W. H.; Parker, V. B.; Schumm, R. H.; Halow,
I.; Bailey, S. M.; Churney, K. L.; and Nuttall, R. L., The NBS tables of chemical thermodynamic
properties, J. Phys. Chem. Reference Data 11, Suppl. 2, 1982

HSC Chemistry for Windows


This program stores data for approximately 7800 compounds. Data come from a variety of
sources. Those chiefly used for calculations in this work include
Barin, I., Thermochemical Data of Pure Substances, VCH Velagsgesellschaft, Wein-
heim, 1989 and 1993.
Barin, I., Knacke, O., and Kubaschewski, O., Thermodynamic Properties of Inorganic
Substances, Springer-Verlag, Berlin, 1973, Supplement 1977.

© 2001 by CRC Press LLC


0256_frame_C01 Page 3 Wednesday, August 9, 2000 6:54 AM

Chase, M. W., Davis, C. A., Downey, J. R., Frurip, D. J., McDonald, R. A., and Syverud,
A. N., JANAF thermochemical tables, J. Phys. Chem. Reference Data, 14, Supplement
1, 1985.
Knacke, O., Kubaschewski, O., and Hesselman, K., Thermochemical Properties of Inor-
ganic Substances, Springer Verlag, Berlin, 1991.

In addition, some tables come from a 1997 edition of HSC. References for these tables appear
at the end of each chapter.
Previous work by the present author is given in the following monographs:
High Temperature Properties and Decomposition of Inorganic Salts
Part 1. Sulfates, Stern, K. H. and Weise, E. L., NSRDS – NBS, 1966.
Part 2. Carbonates, Stern, K. H. and Weise, E. L., NSRDS – NBS, 1969.
Part 3. Nitrates and Nitrites, Stern, K. H., J. Phys. Chem. Ref. Data, 1, 3, 1972.
Part 4. Oxysalts of the Halogens, Stern, K. H., J. Phys. Chem. Ref. Data, 3, 2, 1974.

When data from the HSC program were used, Hf°, Gf°, and S° have been listed for reactants
and products as a function of temperature. Gaseous equilibria are listed in separate tables in each
section, e.g., SO2 + 1/2 O2 = SO3 for sulfates. In many cases the equilibrium constants for the
decomposition reactions can be calculated directly from data in the program. When this was not
possible, other methods, using the equations given above, were used.
When the equilibrium constants are known, the partial pressures of the individual gases can be
calculated under certain restrictive conditions, e.g., closed system, pressures of some gases are
fixed, etc. For example, if one starts with a closed system that is initially pumped out and the salt
allowed to decompose so that pressure builds until equilibrium is reached, it is possible to calculate
the partial pressures of the gases in the system. Thus, for carbonates, the pressure of CO2 equals
the equilibrium constant. For other classes of salts the calculations can be much more complicated.
Appropriate methods are discussed in the individual chapters.

1.3 PHASE TRANSITIONS


Solid–solid phase transitions have been much studied and are generally reliable if they were carried
out appropriately, e.g., if equilibrium was approached from above and below.
Melting points are usually obtained by rapid heating of the salt to minimize decomposition,
but this cannot always be avoided. The lower values should therefore be regarded as lower limits.
Above the melting point most salts decompose so rapidly that boiling points cannot be measured.
When more recent or reliable data were not available, phase transitions were taken from the
following NBS publication: Rossini, F. D., Wagman, D. D., Evans, W. H., Levine, S., and Jaffe, I.
Selected Values of Chemical Thermodynamic Properties, NBS Circular 500, U.S. Govt. Printing
Office, Washington, D.C. 1952.

1.4 DENSITY
The densities of solid salts are generally determined directly by pycnometry, or are calculated from
crystallographic (X-ray diffraction) data. In the latter method atomic packing must be assumed.
Since no allowance is made for crystal imperfections, calculated densities are generally higher than
experimental values by as much as 10%. Since most crystallographers do not indicate the temper-
ature for which the density is calculated, in this work the experimental values have been given
when they were available. However, in some cases the experimental values seem to depend on the
method of preparation, perhaps because of variations in the defect density. Thus, although the
density of a given sample can certainly be determined to four or five significant figures, not too

© 2001 by CRC Press LLC


0256_frame_C01 Page 4 Wednesday, August 9, 2000 6:54 AM

much significance should be attached to figures beyond the third. This is particularly true because
the units of the reported density values are frequently in doubt. In some cases they are not stated
at all; in others it is not clear whether the units are g/cm3 or g/ml; in still others they are specific
gravity values and may be referred to water at 4°C or at some other temperature. The difference
between cubic centimeters and milliliters is so small (the ratio is 1.000028) as to be negligible.
However, the difference in the density of water between 4 and 20° is in the third significant figure
and is no longer negligible for better measurements.
In this work the units reported by the original authors are given. When no units are given it
can generally be assumed that g/ml is meant.
There are some density data for solid salts above ambient temperature which are given in
the form of thermal expansion coefficients. These have been listed when they seemed reliable.
Above the melting point, density data are scarce. Most are available for alkali halides but those
available for salts are taken from the critically evaluated compilation: Janz, G.J., Thermodynamics
and transport properties for molten salts. correlation equations for critically evaluated density,
surface tension, electrical conductance, and viscosity data, J. Phys. Chem. Reference Data, 17,
Suppl. 2, 1988.

1.5 KINETICS OF THERMAL DECOMPOSITION

1.5.1 GENERAL
In contrast to the thermodynamics of decomposition, where a few parameters permit the calculation
of the equilibrium properties of the system, the determination of decomposition rates is largely
an experimental problem, i.e., there are no standard kinetic data from which these rates can be
calculated. This is particularly true for the decomposition reactions of solids which are “topochem-
ical," i.e., where the rate depends on structural factors. One reason for this situation is that it does
not yet seem possible to prepare duplicate samples of any solid inorganic salt that are identical
in all the properties that may determine the rate of decomposition, e.g., the dislocation density of
the crystals.
Nevertheless, the kinetics of decomposition are of sufficient interest, for both theoretical and
practical reasons, that it seems worthwhile to include the results of at least the more carefully done
studies. In this section we summarize features common to all decompositions and the factors
affecting decomposition rates. The discussion is essentially limited to solids, since most experi-
mental and theoretical work has been done in this area. It should be pointed out that many of the
theoretical foundations were laid several decades ago, and that newer developments deal mostly
with refinements. Kinetics of the individual compounds are treated in their respective sections.

1.5.2 MECHANISM OF ENDOTHERMIC SOLID-STATE DECOMPOSITION


Many authors have derived equations for the rate of decomposition from theoretical models. The
aim of these studies is to find functional relations between some measurable quantities of the
system, such as weight and time, which can be tested against experimental results. Although the
usual caveat against taking agreement between the predictions of the model and experimental results
as evidence for the correctness of the model applies, nevertheless such agreement, together with
other experimental evidence, e.g., microscopic examination, is frequently very helpful in at least
narrowing the possibilities.
In this section some of the models which have been proposed are briefly summarized. Greater
detail can be found in books by Garner,1 Young,2 and Tompkins3 and in the original references.
Most models try to account for the shape of the experimental (= fraction decomposed) vs.
time(t) curves. These are generally S-shaped, with a relatively slow induction period, a steep portion
during which the bulk of the material decomposes, and a slow final section. Experimentally

© 2001 by CRC Press LLC


0256_frame_C01 Page 5 Wednesday, August 9, 2000 6:54 AM

determined curves differ primarily in the relative length of the induction period. This induction
period is commonly identified with the formation of nuclei of the new phase B imbedded in A;
these nuclei grow, by spreading both over the surface and into the bulk of the decomposing particle.
After the surface is covered by B, this phase grows into the bulk of the decomposing particle.
Decomposition thus occurs only at the reactant/product phase boundary, as has been argued by
Langmuir from phase rule considerations.4 There seems to be general agreement that the evolution
of gas C during endothermic solid decomposition is not a rate-limiting step in the process because
the product B is sufficiently porous to permit C to escape readily. However, if the gas is generated
in the bulk of the crystal, e.g., by ionizing radiation, and must diffuse along grain boundaries, the
rate of decomposition may depend markedly on the diffusion constant of the gas. This case has
been discussed by Gafner.5
Although the measured decomposition rates are not rate-limited by diffusion of gas out of
the crystal, it is clear that in the case of readily reversible decompositions, such as those of
the carbonates, the measured rate will depend on the difference between the equilibrium
pressure pe and the actual pressure p at the reaction interface. Thus, although the functional
dependence of  on t may be independent of pe – p, the actual rates are not. It is largely for
this reason that it is not possible to tabulate “standard reference data” rate constants, since the
pressure p is not usually known. Instead, one finds decompositions carried out in streams of
various “sweep gases,” which may or may not remove product gases from the reaction interface,
or in a vacuum. In the latter case, the question as to whether the manometrically measured
pressure equals the pressure at the interface needs to be answered. Small sample size and a
porous oxide favor this condition.

1.5.2.1 Kinetic Equations

The most complete classification of solid decompositions has been worked out by Jacobs and
Tompkins.6 When the decomposition begins at some points on the A lattice where the energy is
most favorable, small fragments of B are imbedded in A. These may at first retain the lattice
parameters of A and will therefore possess some strain energy as a result of this deformation.
Jacobs and Tompkins show that the Gibbs energy accompanying the formation of a fragment of B
containing m molecules is

G = am2/3 – bm

where a is proportional to the strain energy and b is the negative of the bulk Gibbs energy per
molecule. G passes through a maximum at m = m* when a fragment has the critical size to be in
equilibrium with its surroundings. Smaller fragments are unstable and revert to A; larger fragments
are stable nuclei and grow. Whether decomposition proceeds from the growth of a few nuclei or
whether many small nuclei are formed depends on the relative magnitudes of the activation energies
   
for nucleus formation ( G 1 ) and the growth stage G 2 . If G 2  G 1 , growth of existing
 
nuclei predominates over formation of new ones. If G 2 = G 1 , many small nuclei are formed,
none of which grow to visible size. In that case the induction period is much reduced. The whole
surface then nucleates rapidly and the kinetics of the remainder of the reaction depends on the rate
at which the interface progresses into the crystal.
We next consider various mechanisms for each of the stages in the decomposition.
1.5.2.1.1 Nucleus Formation
Jacobs and Tompkins6 consider only structure-sensitive nucleation, which occurs at definite sites
in the lattice where the activation energy is least, such as lattice defects and dislocations. The rate
of nucleus formation thus depends on both the defect density and on the activation energy.
Jacobs and Tompkins distinguish two cases:

© 2001 by CRC Press LLC


0256_frame_C01 Page 6 Wednesday, August 9, 2000 6:54 AM

a. The decomposition of a single molecule leads to the formation of a nucleus. The prob-
ability of this unimolecular decomposition is


k 1   exp ( G 1 RT ) (1.4)


where  is the frequency of lattice vibrations and G 1 is the activational Gibbs energy
for nucleus formation. The rate of nucleus formation at time t then is

dN/dt = k1N0 exp (–k1t) (1.5)

or

N = N0 [1 – exp (–k1t)]

where N0 is the total number of potential nucleus-forming sites. The possible loss of
nucleus- forming sites through their ingestion by growing nuclei is neglected. In the

early stages of the reaction and especially for large G 1 and consequently small k1

N  k1 N 0t (1.6)

so that the number of nuclei increases linearly with time.


b. A power law results from two possible mechanisms: (1) a stable nucleus is formed in a
bimolecular combination of two active intermediates, and (2) several decompositions are
required to form a stable nucleus. The first of these possibilities may require the surface
migration of several B molecules until they coalesce into a stable nucleus. Some aspects
of this phenomenon have been discussed by MacDonald.7

1.5.2.2 Growth of Nuclei

As each nucleus is formed, it grows, generally hemispherically, i.e., it spreads across the surface
of the crystal and into its interior. Since the laws for the formation of nuclei are different from
those for their growth, the overall decomposition rate is likely to be complicated if the two rates
are of comparable magnitude. This is particularly true if the complications resulting from overlap-
ping nuclei are taken into account. As nuclei grow, they ingest or overlap sites which would
otherwise yield nuclei. Also, as nuclei grow, they impinge on one another and cease to grow along
the line of contact. Both of these effects result in rates different from those predicted from models
without these refinements.
The simplest decomposition kinetics, commonly observed for carbonates, result from a model
in which nucleation and subsequent surface growth are extremely rapid. The rate of decomposition
is then determined by the progression of the reaction interface into the crystal. Rate laws have been
derived for various geometries, but the ones for spherical particles are most useful and have been
given in several equivalent forms by various authors. In this derivation it is assumed that the reaction
interface proceeds at a constant velocity into the crystal; i.e., for a constant interfacial area, the
rate of decomposition is a constant. For a spherical particle, the reaction proceeds along a constantly
diminishing interface. For such particles of initial radius R0, the fraction decomposed at time t is

3 3 2 2 3
4 3
R 0  4 3
( R 0  k 2 t ) k2t 3 3k 2 t 3k 2 t k2t
-  1   1  ------
-----------------------------------------------------------------------  ---------
- ------------ ------
4 3
R 0
3
 R 0   R 0  R 20 R 30 (1.7)

© 2001 by CRC Press LLC


0256_frame_C01 Page 7 Wednesday, August 9, 2000 6:54 AM

and

3 3 3
d 3k 6k 2 t 3k 2 t
-------  --------2  ---------
- -----------
2 3
- (1.8)
dt R0 R1 R0

These equations were first derived by Hume and Colvin.8


The same, but superficially different, equation was derived by Fischbeck and Schnaidt9 in a
form now frequently used:

d/dt = k1A (1.9)

where A is the total area of the reacting interface. For a spherical particle A = k 2 (1 – )2/3, where
k 2 depends on R0, and hence

d
-------  k ( 1   ) 2 3 (1.10)
dt 2

In this form the equation is frequently referred to as the 2/3 power law. In terms of mass change
it can be written as19

dm 2 3
 ------
-
dt  k 2 m (1.11)

where k 2 again depends on R0. Equation 1.10 is usually given in integrated form

1 3
k 2 t  1  ( 1   ) (1.12)

The dependence of rate on particle size can be included by substitution k2 = k 2/R0.


The second limiting case arises when the rate of linear propagation of the reacting interface
is so great that each particle decomposes as soon as it has a nucleus. For this case, Hume and
Colvin give

Rate = k0Nt

i.e.,

dN/dt = k0 Nt (1.13)

where Nt is the number of particles at time t. Integration from N0 to Nt gives

ln(Nt/N0) = k0t (1.14)

a first-order reaction. This behavior is most likely to be approached by very small particles.
In general, the kinetics of decomposition is more complex than in the two limiting cases just
outlined. Thus, if nucleation is not extremely rapid, the number of nuclei will increase in time.
Each of these nuclei will then begin to grow at a different time. Some of them will spread over
sites at which nucleation would have occurred at some future time. These sites are referred to
as “phantom nuclei” by Jacobs and Tompkins.6 This problem has been studied in great detail by

© 2001 by CRC Press LLC


0256_frame_C01 Page 8 Wednesday, August 9, 2000 6:54 AM

Mampel.10 Both Jacobs and Tompkins and Mampel assume the rate of nuclei formation to be
first order,

dN/dt = k1N0 exp(–k1t) (1.15)

where N0 is the total number of possible sites. For the induction period, Mampel finds a t4 law for
, a relation derived by Jacobs and Tompkins for random nucleation and short times, and by
Fischbeck and Spingler11 for the case of no overlap, a constant rate of nuclei formation, and a large
total number of nuclei.
For large radii (R > k2t) Mampel finds

k2t
------ 1 3
R  1 ( 1   ) (1.16)

an equation (1.12) previously given for the contracting sphere case. For small radii the fraction
decomposed is

X = 1 – ce–kt (1.17)

where c is independent of t, but does depend on R in a rather complicated way. The rates are
dependent on particle size and both small and large radii give lower rates than intermediate radii.
A generalized equation has been derived by Erofeev12 in terms of probability theory. In his
treatment,

t
 
  1     p dt (1.18)
 
 0 

where p is the probability that a molecule will react in the time interval dt.
The general solution of (1.18) for thermal decomposition is

 = 1 – exp (– ktn) (1.19)

where the constant n depends on the shape of the nucleus and the number of electrons necessary
for the formation of a stable nucleus. For example, cylindrical nuclei (the centers of formation are
edges or surface cracks) give n = 3, flat nuclei give n = 4. Thus the evaluation of n, most easily
carried out by transforming (1.19) into

ln [–ln(1 – )] = ln k + n ln t (1.20)

and plotting the left side of (1.20) vs. ln t, gives some information about the nucleation mechanism.

1.5.2.3 Surface Area Changes Decomposition

Since the molar volume of the product oxides is less than that of the corresponding salts, decom-
position results in a porous structure through which the product gases readily escape. Although not
directly related to decomposition kinetics, the resulting increase in surface area during decompo-
sition is of some interest in the production of “active solids” and has been treated by several
authors. Accordingly, we briefly review it here.

© 2001 by CRC Press LLC


0256_frame_C01 Page 9 Wednesday, August 9, 2000 6:54 AM

The oxide is first formed at sites on the salt lattice and is therefore in a strained state. Gregg13
has described these as small crystallites or micelles. The strained pseudo-lattice will tend to
recrystallize into a stable oxide lattice with a resulting decrease in surface area. Three mechanisms
for this recrystallization are generally recognized. At temperatures T/Tm < 0.2, where Tm is the
melting point of the oxide, adhesion predominates. This mechanism results from contact between
parts of the micelles which are atomically contiguous. For 0.2 < T/Tm < 0.35, surface diffusion
predominates. Here mobile atoms or ions in the surface layer leave their position on the lattice to
migrate to sites of lower energies, i.e., to stable oxide lattice positions. Above T/Tm = 0.5, sintering,
i.e., lattice or bulk diffusion, is the most significant process leading to decreased surface area. Thus,
as decomposition proceeds, there are basically two processes leading to changes in surface area:
formation of the metastable oxide produces an increase in surface, and formation of stable oxide
decreases the area. Consequently, the “surface area” vs. “fraction decomposed” plot exhibits a
maximum before decomposition is complete. Nicholson14 has developed a mathematical model
which agrees well with experimental observation. If the purpose of the decomposition is the
preparation of an “active solid,” the reaction should not be carried to completion in order to achieve
the desired maximum surface area.

1.5.3 THE EFFECT OF PRESSURE ON RATE


Most of the work on this topic has been carried out on carbonates because only one gas is produced
and the analysis is thus inherently simpler. For other salts, several gases are present. Although the
principles are basically the same, the discussion has been given here in terms of carbonates.
When the pressure of CO2 in a carbonate–oxide system is equal to the equilibrium pressure
pe, no net reaction occurs. When p < pe, the thermodynamic driving force favors oxide
formation; conversely, when p > pe, carbonate formation is favored. In the actual system the
favored reaction may not occur, however, because kinetic factors prevent it. Particularly when
p is not far from pe, the reaction may not proceed because some rate-limiting process, such
as nucleus formation, is proceeding too slowly. The resulting spurious equilibria15 give rise to
hysteresis effects, i.e., decomposition stops for some p < pe, recombination stops for some p
> pe. It is for this reason that this work relies largely on thermodynamic methods for the
calculation of equilibrium pressures.
When the carbonate and oxide are present as pure phases, the value of pe is completely
unambiguous, i.e., at a given temperature pe = Kp, the equilibrium constant for the reaction MCO3
= MO + CO2. However, when the condensed phase consists of a solid or liquid solution, pe no
longer has a unique value since

K  ( a MO a MCO3 ) p e (1.21)

and pe now depends on the solution composition. Since the decomposition rate is generally some
function of pe – p, we consider only systems without solution formation. We have also omitted
discussion of the effect of “foreign” gases, i.e., gases other than CO2, on the rate. Such gases, e.g.,
O2, H2O, frequently function as catalysts or inhibitors, but these effects are difficult to relate to
well-defined properties of the system.16–18 In order to study the dependence of the reaction rate on
CO2 pressure, it is necessary that the pressure remains fixed throughout the course of the decom-
position while more CO2 is evolving. Since the significant pressure is that at the reaction interface,
it is possible that in a closed system where one depends on diffusion to remove the excess CO2,
the pressure at the interface will exceed the nominal value, particularly if the sample size is large
and the reaction is rapid. The effect of a pressure gradient has been investigated19 by the application
of Fick’s law. For both slow and fast reactions the rate is proportional to pe – p0 where p0 is the
pressure at the reaction site but the proportionality constants differ in the two cases. One way to
avoid the problem of pressure gradients is to use sweep gases made from a known mixture of CO2

© 2001 by CRC Press LLC


0256_frame_C01 Page 10 Wednesday, August 9, 2000 6:54 AM

and some inert gas. Independence of decomposition rate from flow rate is generally taken as
evidence that the reaction is not diffusion controlled, e.g.,20 if the surface area is constant, the
decomposition rate is proportional to pe – p0, i.e.,

v = k(p0 – p) (1.22)

Such a relation was also found to hold for both cylindrical pellets of calcite and powdered CaCO3.21
Cremer and Nitsch,22 in studying the decomposition of CaCO3, found that for samples which
followed a 2/3 power rate law the pressure dependence of the rate (in m1/3t–1) was given by

v = k(1/p – 1/p0) (1.23)

A similar equation has been derived20 from a two-step model in which the CaO is assumed to be
formed first in some “active” state CaO* which occupies a constant fraction of the surface. The
reactions then are

k1
*
CaCO 3 CaO CO 2
k2

1  P

k3
*
CaO CaO
k4

both of which are reversible. The rate

k 0 k 1 k 3  Pk 0 k 2 k 4
R   dw dt  ----------------------------------------------
k1 k3 k4 k2P
- (1.24)

where the constant k0 is introduced to convert the surface area to surface concentration in molecules
per square centimeters, to account for the roughness factor, and the fact that only a portion of the
total surface area may be potentially reactive. Equation 1.24 can be written as

1  CP
R  -------------------
VP D
- (1.25)

where P = 0, R = R0 = 1/D, P = Pe, R = R0, and C = 1/Pe


Hence

1  P Pe
R  ----------------------------
BP 1 R 0
- (1.26)

When 1/R0 >> BP,

R = 1/BP – 1/Pe

which is essentially the equation given by Cremer and Nitsch.22

© 2001 by CRC Press LLC


0256_frame_C01 Page 11 Wednesday, August 9, 2000 6:54 AM

1.5.4 ACTIVATION ENERGY


The activation energy E, as defined by the usual Arrhenius equation

k = Ae–E/RT

is usually at least as great as the thermodynamic enthalpy for the reaction, H, though it is frequently
greater. When E = H, the reaction is often described as occurring “without activation.” The
question as to whether E = H for the endothermic carbonate decomposition was first considered
by Fischbeck and Schnaidt.23 Shortly thereafter, Zawadzki and Bretsnajder15 showed that for a
reversible reaction of the type A(s) = B(s) + C(g), the experimental value of E for the forward
reaction, as determined from the temperature dependence of k, must necessarily increase with rising
pressure and will have its minimum value in the absence of the backward reaction, i.e., in a vacuum.
This dependence was verified for the decomposition of CaCO3 , CdCO3, and Ag2CO3. In a vacuum
the temperature dependence of k was the same as that of the equilibrium pressure pe. i.e., E = H.
The dependence of E on the pressure was reexamined by Pavlyuchenko and Prodan,24 who
distinguished two cases: (1) if E is determined from the temperature dependence of the rate, with
the pressure p held constant at the various temperatures, then E increases as p → pe. (2) If,
however, the ratio p/pe is held constant as the temperature is changed, then E will be independent
of the actual numerical value of the ratio and will have the same value as in a vacuum. This
conclusion was reached by a theoretical argument and verified by data on the decomposition of
CdCO3. Since condition (1) is the more usual experimentally, particularly in the older literature, it
is not surprising that widely varying values of E for the same reaction have been reported25 by
authors who failed to recognize the dependence of E on pressure.
The atomic mechanism responsible for monomolecular reactions, including thermal decompo-
sitions, was first discussed by Polanyi and Wigner.26 Their model assumes that decomposition occurs
when, due to energy fluctuations in the bonds of the molecule, the bond strength is exceeded; or
more precisely, the bond energy “resides in harmonic vibrations and that decomposition occurs
when their amplitude is exceeded.” The resulting expression for the first-order Polanyi–Wigner rate
constant is

k = v exp(–E/RT) (1.27)

where v is the atomic frequency of vibration and E is the activation energy. Thus, a molecule
having energy E would leave the surface.
Shannon27 analyzed in some detail the theory as it applies to the thermal decomposition of
solids. He found that of the 31 reactions for which he compared experimental rate constants with
those calculated from the Polanyi–Wigner equation, only a third showed order-of-magnitude agree-
ment. In Shannon’s view, the lack of agreement stems from neglecting rotational and other vibra-
tional degrees of freedom.
Shannon calculated the rate constant for thermal decomposition of a solid from absolute reaction
rate theory. The resulting equation is of the same form as Equation 1.27, but v is replaced by a
partition function ratio:

k =( kT/h)(Q*/Q) exp(–E/kT) (1.28)

where Q* is the complete partition function of the activated complex, excluding that for the reaction
coordinate, and Q is the complete partition complex for the reactant. Detailed calculations of the
partition functions were carried out for the decomposition of CaCO3 and MgCO3, using known
values of the rotational and vibrational partition functions and several models for the activated
complex. The best agreement with experiment was obtained for a model in which the carbonate

© 2001 by CRC Press LLC


0256_frame_C01 Page 12 Wednesday, August 9, 2000 6:54 AM

ion in the lattice is free to rotate and the CO2 molecule leaves directly from the surface of the solid,
rather than from a mobile surface layer.
It should be noted that over the past 50 years studies of solid decomposition kinetics have
progressed from the application of equations which were originally derived for gases to more and
more detailed studies on molecular events occurring at the reactant-product interface. This point
has been emphasized by Boldyrev28 who, with his co-workers, has amassed a large body of evidence
that shows the importance of defect structure on kinetics. This structure may be affected by radiation
or by mechanical forces, such as grinding, which affect both the surface and the internal structure
of crystals.

1.5.5 THE EFFECT OF PREPARATION METHOD ON RATE


Various authors have recognized that the rate of thermal decomposition may depend on the method
of preparation and the thermal “history” of the sample, but systematic studies in this area are
inherently difficult. However, there has been progress in this field, called “tribochemical decom-
position.” Boldyrev28 has distinguished between physical and chemical effects. The main process
in the former is the introduction of defects by diminution, e.g., grinding. The latter arise from
chemical processes at the contact points of solid particles. Boldyrev28 has provided a framework
for the analysis of these effects and tested his hypotheses with an experimental study of some
nitrate and bromate decompositions. Additional problems arise if impurities present in the samples
at very low concentrations, such as H2O, catalyze the decomposition.29 In that case, the resulting
change in mechanism will affect not only the rate but also the activation energy.
If the carbonate is a single crystal, the reaction will propagate from the surface to the interior.
For polycrystalline materials the reaction may additionally propagate along grain boundaries.30
A systematic study of the effect of sample constitution on the decomposition rate was carried
out by Cremer and Nitsch.22 They decomposed four calcite samples at 850°C and CO2 pressures
ranging from 120 to 220 torr. The CaO product exhibited increasing grain size with increasing
pressure. The four CaO samples were then reacted with CO2 at 850°C and 429 torr to produce four
CaCO3 samples with similarly increasing grain size. When these four samples were then decom-
posed under identical conditions (850°, 200 torr CO2) the rate of decomposition increased in inverse
order of crystal size. The disappearance of the induction period with decreasing crystal size was
particularly noticeable. These results are explained in terms of more rapid nucleus formation at
lower pressures, leading to smaller CaO crystals and consequently a larger number of active sites
in the CaCO3 formed subsequently.
It is evident from these results that the course of the decomposition is affected by the micro-
structure of the sample and that reproducible preparation and characterization of nominally identical
materials are important prerequisites for understanding and systematizing decomposition kinetics.

1.5.6 THE EFFECT OF HEAT TRANSFER ON RATE


An endothermic solid-state decomposition proceeds as the product–reactant interface advances into
the interior of the sample. The interface can, however, advance only if the necessary heat of reaction
is applied to it. Since the source of heat is outside the sample, the rate of heat transfer to the
interface may become rate-determining if the inherent rate of reaction exceeds the rate of heat
transport. The rate of heat transport depends not only on the properties of the product through
which heat must be transported, but also on the general experimental arrangement. It is for this
reason that one expects and finds the literature to be conflicting.
For example, Narsimhan30 derived an equation for the thermal decomposition rate based on the
idea that the entire particle rapidly reaches the decomposition temperature, so there are no internal
temperature gradients. All the heat reaching the reaction interface is used up in the decomposition,
so the reaction rate depends on the rate of heat transport. For both spherical and cylindrical particles,

© 2001 by CRC Press LLC


0256_frame_C01 Page 13 Wednesday, August 9, 2000 6:54 AM

good agreement with some literature values is found. Under these conditions the chemical steps
are no longer rate-determining. It can thus be seen that in any solid-state reaction whose enthalpy
is appreciable, the establishment, maintenance, and measurement of the temperature at the reaction
interface is a complicated problem. Thus, careful studies of the decomposition of PbCO331 and
several other inorganic salts32 have shown that the induction period may disappear if the entire
sample is rapidly brought to the reaction temperature. There is, therefore, some question now as
to whether the commonly reported induction periods in the decomposition of so many substances
are artifacts of the experimental arrangements. Considerable care in the experiment’s design is
certainly required to insure a constant and known temperature at the reaction interface throughout
the course of the decomposition.

1.5.7 THE INFLUENCE OF CRYSTAL DEFECT STRUCTURE ON DECOMPOSITION RATE


The idea that the defect structure of a solid reactant affects the rate of decomposition seems to be
generally accepted; but of all the factors influencing the kinetics, this is the one most difficult to
characterize quantitatively. Boldyrev and his many co-workers28 have made considerable progress
in elucidating the factors which affect the decomposition rate of solids; however, at the level of
detail required for the understanding of a particular reaction, it seems difficult to make connections
to the standard kinetic equations derived by earlier workers.
Searcy and co-workers32–34 have addressed the problems that arise from the use of large and/or
powdered samples by measuring decomposition from specific faces of small single crystals, using
the Langmuir method in which weight loss in a vacuum is measured. The authors cite several
advantages of this approach: (1) the possibility that intergranular diffusion may be slow enough to
influence the measured rate is eliminated; (2) the area and morphology of the interface at which
the reaction proceeds can be precisely measured as a function of time; (3) the ratio of the measured
flux of gaseous reaction product which leaves the reactant surface to the maximum (equilibrium)
flux which can be predicted from equilibrium data and the kinetic theory of gases, provides a
potentially valuable parameter for describing and understanding reaction kinetics, just as it does
for congruent vaporization.
The maximum flux is given by the Hertz—Knudsen–Langmuir (HKL) equation

0
P eq exp (  G d RT )
J max  -------------------------------
-
1 2 
--------------------------------------
1 2 (1.29)
( 2
MRT ) ( 2
MRT )

The ratio of the experimental flow J to Jmax, v, the decomposition coefficient, may be useful in
assessing the decomposition mechanism. However, account must be taken of formation of the solid
product in a metastable state. The formation of this material may be the rate-limiting step. Beruto
and Searcy33,34 have carried out a study of CaCO3 decomposition using these principles. They have
also studied the nature of the CaO formed. Thus, the field of solid decomposition kinetics seems
to be moving toward a detailed examination of the products formed, with less emphasis on the
overall kinetic equations.

REFERENCES
1. Garner, W.E., Chemistry of the Solid State, Academic Press, New York, 1955.
2. Young, D. A., Decomposition of Solids, Pergamon Press, Oxford, 1966.
3. Tompkins, F. C., Decomposition Reactions, in Reactivity of Solids, Treatise on Solid State Chemistry,
Vol 4, Hannay, N. B., Ed., Plenum Press, New York, 1976.
4. Langmuir, I., J. Am. Chem. Soc., 38, 2263, 1916.
5. Gafner, G., Trans. Faraday Soc., 55, 981, 1959.

© 2001 by CRC Press LLC


0256_frame_C01 Page 14 Wednesday, August 9, 2000 6:54 AM

6. Jacobs, P. W. M. and Tompkins, I., Ch. 7 in reference 1.


7. MacDonald, J. Y., Trans. Faraday Soc., 47, 860, 1951.
8. Hume, J. and Colvin, J., Phil. Mag., 8, 589, 1929.
9. Fischbeck, K. and Schnaidt, K., Z. Elektrochem., 38, 649, 1932.
10. Mampel, K., Z. Phys. Chem., A187, 43, 235, 1940.
11. Fischbeck, K. and Spingler, H., Z. Anorg. Chem., 241, 209, 1939.
12. Erofeev, B. V., Dokl. Akad. Nauk SSSR, 52, 511, 1946.
13. Gregg, S. J., J. Chem. Soc., 3940, 1953.
14. Nicholson, D., Trans. Faraday Soc., 61, 990, 1965.
15. Zawadzki, J. and Bretsnajder, S., Z. Elektrochem., 41, 721, 1938.
16. Bischoff, F., Z. Anorg. Chem., 252, 10, 1942; ibid. 262, 18, 1950.
17. Huttig, G. F. and Hermann, E., Kolloid Z., 92, 9, 1940.
18. Huttig, G. F. and Strial, K., Z. Anorg. Chem., 289, 249, 1932.
19. Barret, P. and Bonnefair, L., Bull. Soc. Chim. France, 576, 1961.
20. Hyatt, E. P., Cutler, I. B., and Wadsworth, M. E., J. Am. Ceram. Soc., 41, 70, 1958.
21. Ingraham, T. R. and Marier, P., Can. J. Chem. Eng., 41, 170, 1963.
22. Cremer, E. and Nitsch, E., Z. Elektrochem., 66, 697, 1962.
23. Fischbeck, K. and Schnaidt, K., Z. Elektrochem., 38, 649, 1932.
24. Pavlyuchenko, M. M. and Prodan, E. A., 5th International Symp. Reactivity of Solids, Schwab, G.
M., Ed., Elsevier, Amsterdam, 407, 1965.
25. Stern, K.H. and Weise, E.L., High Temperature Properties of Inorganic Salts, Part 1, Sulfates, NSRDS-
NBS - 7, 1966.
26. Polanyi, M. and Wigner, E., Z. Phys. Chem., A 139, 439, 1928.
27. Shannon, R. D., Trans. Faraday Soc., 60, 1902, 1964.
28. Boldyrev, V. V., Thermochim. Acta, 100, 315, 1986; Z. Phys. Chem. (Leipzig), 256, 342, 1975.
29. Prodan, E. A. and Pavlyuchenko, M. M., Geterogennye Khim. Rektsii Inst. Obshch. i Neorg. Khim.
Akad. Nauk Belorusski SSSR, 64, 1965.
30. Narsimhan, G. V., Chem. Eng. Sci., 16, 7, 1961.
31. Zhabrova, G. M. and Gordeeva, V. A., Izvest. Akad. Nauk SSSR Otdel Khim. Nauk, 569, 1959.
32. Basu, T. K. and Searcy, A.W., J. Chem. Soc. Faraday Trans., I 72, 1889, 1976.
33. Beruto, D. and Searcy, A.W., J. Chem. Soc. Faraday, 70, 2145, 1974.
34. Searcy, A.W. and Beruto, D., J. Phys. Chem., 80, 425, 1976.

© 2001 by CRC Press LLC

Das könnte Ihnen auch gefallen