You are on page 1of 8

Ind. Eng. Chem. Res.

2005, 44, 6591-6598 6591

Thermal Decomposition of Aluminum Chloride Hexahydrate


M. Hartman,* O. Trnka, and O. Šolcová
Institute of Chemical Process Fundamentals, Academy of Sciences of the Czech Republic,
165 02 Prague 6-Suchdol, Czech Republic

The decomposition rate of aluminum chloride hexahydrate (AlCl3‚6H2O) was measured as weight
loss at ambient pressure and elevated temperatures up to 270 °C. Such incomplete thermal
decomposition produces a porous and reactive basic aluminum chloride [Al2O3‚2HCl‚2H2O or
Al2(OH)4Cl2‚H2O] which dissolves in water to give poly(aluminum chloride) used as an efficient
flocculation agent. A slowly rising temperature method and very small sample masses, which
minimize heat and mass transfer intrusions, were employed to determine intrinsic reaction rates.
A fractional order kinetic equation of Arrhenius type was proposed for the decomposition and
tested also against the results amassed by experiment in a constant temperature mode. This
correlation allows the estimation of the reaction rate as a function of temperature and the extent
of decomposition. It can be readily employed in modeling and simulation of the decomposition
process. The contents of aluminum and chlorine in the decomposed solids were also explored in
the course of the decomposition process. Pore volume (porosity), pore-size distribution, and BET
surface area data were also collected on decomposed chloride particles.

Introduction 25978.8
log K ) 85.342 - (1)
T
Poly(aluminum chloride), PAC, has been known and
employed as an effective flocculating agent in water where
treatment processes. The term PAC denotes an aqueous
solution of basic aluminum chloride hydrate (approxi- K ) PHCl3‚PH2O4.5 (2)
mately Al2(OH)4Cl2‚H2O or Al2O3‚2HCl‚2H2O), the con-
centration of which corresponds to about 10 wt % Al2O3. The dissociation pressure of AlCl3‚6H2O, P, is taken
The basic aluminum chloride hydrate, BAC, is formed as
by the well-controlled, partial thermal decomposition of
aluminum chloride hexahydrate (AlCl3‚6H2O) at el- P ) PHCl + PH2O (3)
evated temperatures.1,2 Aluminum chloride hexahydrate
is produced, for instance, by the reaction of aluminum
hydroxide with an aqueous solution of hydrogen chloride With respect to stoichiometry it holds
and by subsequent crystallization.
PH2O ) 1.5 PHCl (3a)
It has been well-established that anhydrous alumi-
num salts cannot be simply formed by heating the
corresponding hydrates because of their amphoteric and we get
properties and tendency to hydrolyze. Aluminum chlo-
ride hexahydrate should be written as [Al(H2O)6]Cl3 3463.84
log P ) 8.07122 - (4)
rather than as AlCl3‚6H2O since the interaction force T
Al-O prevents forming the Al-Cl bonds.
At high temperatures,3-5 the following, well-defined, It should be noted that eqs 1 and 4 are based upon the
decomposition reaction takes place thermochemical data recently compiled and tabulated
by Barin.6 The decomposition temperature, Td, defined
as a temperature at which the dissociation pressure, P,
1
[Al(H2O)6]Cl3(s) ) Al2O3(s) + 3HCl(g) + 4.5H2O(g) is equal to the pressure of the surrounding atmosphere
2 (i.e., P ) 1.01325 bar), and predicted by eq 4 amounts
(A)
to 429.5 K (156.3 °C).
At temperatures below 1000 °C, the less compact
Its standard enthalpy, deduced from Barin’s6 ther- γ-Al2O3 (F ) 3.40 g/cm3) is formed. It exhibits the
modynamic data, is as large as defective, cubic, face-centered structure. This open form
constitues the basis of the so-called active aluminum
∆HA° (298 K) ) +996.03 kJ/mol Al2O3 oxide which is frequently used, for example, in catalysis,
ion exchange, and chromatography.
For the sake of PAC production,1,2 AlCl3‚6H2O as the
Thermodynamics constraints imposed on reaction A
precursor is incompletely decomposed at about 160-200
can be expressed by °C under ambient pressure. In such operation condi-
tions, the exact decomposition chemistry has not been
* To whom correspondence should be addressed. Tel.: +420 fully understood yet. The following reaction is usually
220390254. Fax: +420 220920661. E-mail: hartman@icpf.cas.cz. assumed in the literature4,5 to occur
10.1021/ie058005y CCC: $30.25 © 2005 American Chemical Society
Published on Web 07/15/2005
6592 Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005

2[Al(H2O)6]Cl3(s) ) Al2O3‚xHCl‚yH2O(s) +
(6 - x)HCl(g) + (9 - y)H2O(g) (B)

The stoichiometric coefficients x and y are usually


presumed to be near 2. This figure (x ) y ) 2) was
employed in all stoichiometric calculations throughout
this work.
In their recent studies, Park et al.1,2 focused on the
bench-scale decomposition of aluminum chloride hexahy-
drate and the production of poly(aluminum chloride).
Little is known about the intrinsic reaction kinetics,
chemistry, and the textural features of the reaction
product.
In this article we present experimental findings on
the thermal decomposition of aluminum chloride hexahy-
drate at elevated temperatures. The purpose of the work
was to elucidate the intrinsic kinetics and the chemistry
of the decomposition reaction, and the physicochemical
properties of the reaction product.

Experimental Section Figure 1. DTA and thermogravimetric results (TGA) for alumi-
num chloride hexahydrate (AlCl3‚6H2O) amassed at a heating rate
Material. The work was carried out with the alumi- of 3 °C/min between room temperature (22 °C) and 270 °C. Initial
num chloride hexahydrate, ACHH, (AlCl3‚6H2O) ob- mass of samples, 20.5 mg; flow rate of entrainer (air), 20 mL/min;
tained from Sigma-Aldrich, with purity above 99% and standard sample (DTA), Al2O3.
the density as large as 2.398 g/cm3. This material was
also subjected to X-ray powder diffraction analysis with sample. Furthermore, the sample pan contained small
the use of a model X’Pert Philips Co. (Philips Analytical holes which permitted gas to flow past the sample. The
X-ray B. V.). Results confirmed the presence of a single tip of the thermocouple was located about 1-2 mm
solid component in the sample. below the sample pan.
DTA and TGA curves of the employed, finely-ground Provided that the chemistry of a decomposition reac-
aluminum chloride hexahydrate were recorded (model tion is well-defined, thermogravimetric analysis (TGA)
TG-750, Stanton-Redcroft Co.). As can be seen in Figure offers solid data under well-controlled laboratory condi-
1, there is a single, large endothermic peak on the DTA tions. If small samples, fine-powdered solids, and low
curve at about 185-190 °C. It is believed that this heating rates are used, intrusive heat and mass transfer
massive peak reflects a maximum rate of the formation effects on the rate of reaction are minimized. Removal
of the basic aluminum chloride. It should also be noted of the gaseous product of reaction eliminates a possible
that the decomposition reaction is strongly endothermic. effect of equilibrium constraints.
There is no discernible reason for the minuscule, but Since the chemistry of the decomposition process of
noticeable, wave occurring at about 35 °C on the DTA aluminum chloride hexahydrate is not straightforward,
curve. thermogravimetric experiments need to be supple-
Crystals of the hexahydrate were first crushed, then mented with chemical analyses.
ground, and finally sieved to provide samples for the
kinetics experiments. To conform to common practice The extent of decomposition of the aluminum chloride
(Mu and Perlmutter,7,8 Hartman et al.,9-12 and Kim and hexahydrate (ACHH) was determined as weight loss.
Yim13), all experiments were run on finely powdered With respect to the uncertainties of the reaction chem-
samples with particle sizes between 50 and 80 µm. istry, the measured weight loss was transformed into
Procedure and Apparatus. Decomposition TGA the conversion of the chloride to the unequivocally
experiments were performed with the finely powdered defined oxide (Al2O3) by means of the relationship
samples at a constant temperature increase rate of 3
°C/min. Initial sample weights were close to 10 mg, and 2 MACHH wo - w(τ)
air flow of 20 mL/min was maintained. Larger samples X) ‚ ‚ (5)
z 9MH2O + 6MHCl wo
were only employed for chemical analyses and textural
measurements. Also, constant temperature experiments
were done under such conditions. In this work, temper- The symbols wo and w(τ) are the initial mass of the
ature was maintained at a desired level within a range sample and the mass of the sample at any moment of
of (0.1 °C. The same apparatus was employed for both time, respectively. The symbol z is the weight fraction
nonisothermal measurements and isothermal runs. The of AlCl3‚6H2O in the original (initial) reactant (z ) 0.99).
TGA module monitors the weight of a sample and its Complete conversion of AlCl3‚6H2O to Al2O3 (X ) 1)
rate of change continuously, when heated in any inert corresponds to a relative decrease in weight of 0.78095
(or reactive) gas, either as a function of increasing for z ) 0.99. Provided that this aluminum chloride
temperature or at a preselected temperature. The DTA hexahydrate is completely converted into the basic
unit measures the temperature difference between a chloride [Al2(OH)4Cl2‚H2O or Al2O3‚2HCl‚2H2O], the
sample and a reference. relative weight loss only amounts to 0.5576. The weight
The gas space over the sample was as large as about loss is a directly and well-determined quantity. That is
70 cm3. A special arrangement (a doughnut-shaped why it accompanies in any event the data on the
baffle) prevented the gas stream from bypassing the conversion in this work.
Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005 6593

thirty points were taken to fit eq 8. When the effective


order of reaction, n, was found by a trial and error
procedure, the other kinetics parameters were computed
by a least-squares method. The activation energy and
frequency factor were determined from the slope and
the ordinate intercept of the best-fit straight line,
respectively. The values of kinetic parameters inferred
from the temperature-increasing experiments amount
to

ko ) 1.61 × 1014 1/s, E ) 1.359 × 105 J/mol,


and n ) 3.56
Since the above relationship 6 is nonlinear, it is, in
general, sensitive to the variations in values of the
respective kinetics parameters. The conversions com-
puted for different times of exposure and temperatures
from eq 6 with the use of the above-mentioned kinetics
parameters are in Figure 2 compared to the correspond-
ing experimental values determined at temperatures
between 30 and 270 °C. As can be seen, the experimen-
tal data points plotted in this figure fit the computed
Figure 2. Thermal decomposition of aluminum chloride hexahy- curve fairly well.
drate (AlCl3‚6H2O) in a temperature-increasing experiment. Initial
mass of the sample, 14.82 mg; heating rate, 3 °C/min; flow rate of Park et al.1,2 assumed that the decomposition reaction
entrainer (air 20 cm3/min; O, experimental data points. The solid is first-order in conversion of the chloride (i.e., n ) 1)
line shows the predictions of eq 6. and fitted their laboratory-scale data to the integrated
form of eq 6
Results and Discussion
Nonisothermal Decomposition Kinetics. Prelimi-
nary experiments indicated that first traces of the
ln [
-ln(1 - X)
τ ] E 1
) ln ko - ‚
R T
(9)
commencing decomposition of AlCl3‚6H2O became de-
tectable at about 90 °C. The data provided by the rising- n)1
temperature method, some of which are plotted in From the straight line drawn to fit the data amassed
Figure 2, were employed as the practical basis for at 160, 180, and 200 °C, the authors1,2 calculated the
development of a kinetics equation describing the rate preexponential factor, ko, and the activation energy, Eo,
of the overall decomposition reaction. As can also be to be as large as 297.6 1/s and 5.55 × 104 J/mol,
seen in Figure 2, the decomposition of AlCl3‚6H2O respectively. The authors warn that the decomposition
becomes quite rapid at about 150 °C. is too slow still at 140 °C and eq 9 may not be applicable
It cannot be anticipated that a single kinetic rate at such temperatures.
equation can describe the thermal decomposition ex- Isothermal Decomposition Kinetics. The rate of
plored in this work. In general, the decomposition of a thermal decomposition of the aluminum chloride hexahy-
solid can be represented by a rate expression drate was also explored in the constant temperature
mode. The constant temperature runs were carried out
dX E
dτ [
) ko exp -
RT( )]
(1 - X)n (6) in the range 90-150 °C. Preliminary tests proved that
the particles did not melt at such temperatures. The
with the initial condition results of constant temperature runs are shown in
Figure 3. As can be seen, there is an appreciable
X ) 0 at τ ) 0 (7) difference in the course of the curves measured at 90-
110 °C and those obtained at 130-150 °C. While the
This equation accounts for possible effects of nucle- first three lines are nearly linear, the other three exhibit
ation and diffusion on the rate of decomposition. It a sigmoid shape. In accordance with the temperature-
reduces to the Jerofeev equation for n ) 1 (Mu and increasing experiments, the decomposition becomes
Perlmutter7); it conforms to the three-dimensional and quite rapid at about 140 °C. It should be noticed that
two-dimensional shrinking core models for n ) 1/3 and in the range 130-150 °C, the experimental curves level
n ) 1/2, respectively. The equation follows the Avrami off after about 2 h exposure. Also, the weight loss,
nucleation law (constant density of nuclei and one- attained after such a long exposure, slightly augments
dimensional nucleus growth) with n ) 0 (Mu and with the increasing temperature.
Perlmutter;7,8 Hartman et al.14). In Figure 3 are also presented the predictions of eq 6
The experimental data from the decomposition run for 140 °C with the use of the kinetics parameters
were tested empirically by fitting to the linearized form deduced from the results obtained by the nonisothermal
of eq 6 procedure. As can be seen, agreement between the
constant temperature results and the predictions ap-
dX/dτ E 1 pears to be reasonable.
ln ) ln ko - ‚ (8)
(1 - X)n R T The different modes of experiment seem to be among
major factors affecting differences between experiment
Values of the reaction rate, ∆X/∆τ, were determined and the predictions in Figure 3. It is a common experi-
from the curve X vs τ at equal conversion intervals, ence that the results of the temperature-increasing runs
6594 Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005

Figure 3. Thermal decomposition of aluminum chloride hexahy- Figure 4. Thermal decomposition of larger samples of aluminum
drate (AlCl3‚6H2O) in temperature-constant experiments. Initial chloride hexahydrate in a muffle furnace. Initial mass of the
mass of the samples, 14.45-15.86 mg; flow rate of entrainer (air), samples, 3.002-3.006 g; flow rate of entrainer (air), 50 cm3/min;
20 cm3/min. The solid lines represent the experimental curves; O, elapsed time of exposure, τ ) 60 min; O, this work, decomposition
predictions for the temperature 140 °C of eq 6, the parameters of in a muffle furnace; ×, this work, deduced from the TGA data
which were deduced from the temperature-increasing run. shown in Figure 3; b, results of Park et al.1

are notoriously sensitive to the rates of heating. Heat immersed into a heated silicon oil bath. Some results
transfer can affect the results provided by constant deduced from the authors’ experimental curves are also
temperature runs, particularly if the rapid rate of incorporated into Figure 4.
decomposition is coupled with the high heat of reaction. As can be seen in this figure, Park’s data are not in
Considerable sensitivity of the rate of reaction to tem- conflict with the trend of our experimental findings.
perature also should not be overlooked. In light of such Amounts of Aluminum and Chlorine in the
facts the found differences seem to be understandable. Partially Decomposed Solids. A natural question
It is believed that the proposed kinetic equation can be arises as to continuous changes in the solids composition
applied to modeling and simulation of suitable reactors in the course of the decomposition process. As apparent,
for the thermal decomposition of aluminum chloride the composition of the evolved gas inherently unfolds
hexahydrate. The empirical relationship developed in from the progress of the decomposition reactions occur-
this work has usual limitations and it should be applied ring within the particles.
with caution outside the experimental conditions from The chloride particles, first exposed to different tem-
which it was deduced. peratures (150-350 °C) and to the sweep gas for 60 min
Decomposition of Larger Samples. To secure in the muffle furnace, were then analyzed for aluminum
quantities of partially decomposed aluminum chloride and chlorine. The measured contents of aluminum
hexahydrate needed for chemical and textural analyses, increased from 12.4 to 44.2 wt %; those of chlorine
larger samples of the chloride were exposed to elevated decreased from 43.2 to 8.9 wt % as the relative weight
temperatures. Crystals of the chloride were crushed and loss increased from 13.2 to 75.5%. These values of
sieved, and the fraction of particles within a sieve size weight loss correspond to the extents of decomposition
range of 0.50-0.63 mm (d h p ) 0.565 mm) was investi- of 16.7 and 95.7%, respectively. The determined mass
gated in this segment of the work. Samples (3 g), fractions of both species are shown in Figure 5 in
dispersed in shallow corundum crucibles, were inserted dependence on the fractional weight loss.
into a muffle furnace and exposed to the stream of air The chlorine curve in Figure 5 indicates that hydro-
as a sweep gas at the temperature of interest (150- gen chloride commences liberating at very early stages
350 °C) for 60 min. Then the decomposed particles were of the decomposition reaction. Both curves in Figure 5
stored in airtight containers, and shortly afterward they become quite steep when the decomposition is nearing
were subjected to chemical and textural analyses. The completion.
weight loss of each sample was also determined and is As illustrated in Figure 6, the relationship between
shown in Figure 4. the content of aluminum and that of chlorine in the
For the sake of comparison, some data read from the decomposing solids is nearly a linear one. The depen-
TGA lines shown in Figure 3 are plotted along with the dence of the molar ratio of Cl to Al in the solids on
new decomposition results. As can be seen, the new weight loss shown in Figure 7 exhibits a slightly concave
decomposition curve outlined by the data amassed with shape.
the larger samples is distinctly shifted (30-80 °C) Textural Features of the Decomposed Particles.
toward higher temperatures. The collected data also In the course of the decomposition, water vapor and
suggest that even at the highest temperature employed hydrogen chloride are evolved and consequently numer-
(350 °C), the complete conversion to Al2O3 had not yrt ous pores are formed. Upon the loss of gases, the
been attained. pseudomorphs similar to the parent chloride tend to
Park et al.1 explored the extent of decomposition of remain at lower temperatures.15 At higher temperatures
AlCl3‚6H2O at different temperatures (140-200 °C) recrystallization or sintering of the reaction product
using 5-g samples placed in a small flask which was takes place.
Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005 6595

Figure 5. Amounts of aluminum (Al) and chlorine (Cl) in the Figure 7. Molar ratio of chlorine (Cl) to aluminum (Al) present
partially decomposed samples as functions of weight loss of the in the partially decomposed solids as a function of the fractional
solids. Experimental conditions are the same as those mentioned weight loss. Experimental conditions are the same as those
in the caption for Figure 4. mentioned in the caption for Figure 4.

Figure 6. Relationship between the weight fractions of aluminum


(Al) and chlorine (Cl) in the partially decomposed solids. Experi-
mental conditions are the same as those mentioned in the caption
for Figure 4.
Figure 8. True particle density (FHe) and particle density (FHg)
as functions of fractional weight loss. Experimental conditions are
At first, an effort was taken to explore how the volume the same as those mentioned in the caption for Figure 4.
of the solid phase is changed by the decomposition
process. Since the operation temperatures are not high, True and particle density of the decomposed products
it is feasible to assume that no shrinkage occurs and were determined by helium and mercury pycnometry,
the decomposed particle retains its gross external respectively. Pore-size distributions were determined by
volume. As considerable amounts of gases (water vapor measuring the volume of mercury penetrating the pore
and hydrogen chloride) are evolved in the course of the volume at increasing pressure.
reaction, it is evident that the porosity/pore volume of Figure 8 shows the results of the pycnometric mea-
the decomposing particle is increased as the decomposi- surements in dependence on the extent of decomposi-
tion advances. It should be noted, however, that the tion. It is obvious that the measured true densities
reaction product is not unambiguously defined. There- increase with increasing extent of decomposition. Nev-
fore, an analytical relationship cannot be developed ertheless, all values (1.64-1.95 g/cm3) are much lower
between the porosity of the decomposing particle and than the density of γ-Al2O3, which amounts to 3.40
the progress of decomposition.16,17 g/cm3. This fact suggests that some pseudomorphous
6596 Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005

Figure 11. Differential pore-size distribution of the particles


decomposed at 250 °C. Time of exposure, 60 min.
Figure 9. Dependence of the porosity of decomposed particles on
the extent of decomposition. Experimental conditions are the same
as those mentioned in the caption for Figure 6.

Figure 12. Differential pore-size distribution of the particles


decomposed at 350 °C. Time of exposure, 60 min.

Figure 10. Differential pore-size distribution of the particles


decomposed at 150 °C. Time of exposure, 60 min. sizes in the respective samples. The shift of the most
probable pores to larger ones with the increasing
temperature of decomposition is most likely linked to
structures can form and persist at the employed tem- sintering or recrystallization of the formed pseudomor-
peratures. phs similar to the parent chloride. The bimodal distri-
The curved line shown in Figure 9 displays how the bution displayed in Figure 11, with the second maxi-
porosity (pore volume) augments with the increasing mum at about 1000 nm, was only found at 250 °C.
extent of decomposition. The highest porosity deter-
mined with the particles decomposed at 350 °C amounted Conclusions
to 57%. With increase in the porosity of BAC its rate of
dissolution in water will most likely increase. The BET Both water vapor and hydrogen chloride are released
surface area of the decomposed particles was on the when aluminum chloride hexahydrate is gradually
order of 101 m2/g. heated. First noticeable signs of the thermal decomposi-
The pore size distributions of the solids decomposed tion of the chloride occur at 90 °C.
at 150, 250, and 350 °C were also determined. The The results provided by the nonisothermal method
amassed results are illustrated in Figures 10, 11, and showed that the reaction kinetics was more than third-
12. The measured curves indicate that the radii of pores order in mass of the chloride up to high extents of
are distributed over the 1-104 nm range. The derivative decomposition. A reaction rate equation educed from
curves shown in these figures exhibit their peaks at the these temperature-increasing thermogravimetric data
radii of 20, 300, and 400 nm as the most probable pore describes the course of decomposition with fair accuracy.
Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005 6597

The results of the isothermal experiments exploring the t ) Celsius temperature, °C


extent of the decomposition reaction are in reasonable T ) thermodynamic temperature, K
accordance with the predictions of the proposed ki- Vp ) pore volume ) (1/FHg) - (1/FHe) ) [e/(1 - e)]‚(1/FHe) )
netic relationship based upon the rising-temperature e/FHg, cm3/g
data. wo ) initial (original) mass of sample (τ ) 0), g
The amount of chlorine in the decomposing particles w(τ) ) mass of sample of any moment of time τ, g
commences decreasing from very early stages of the X ) fractional extent of decomposition of aluminum
decomposition process. Thus, the chloride particles chloride hexahydrate (AlCl3‚6H2O) expressed as the
decomposed at 350 °C for 60 min still contained 8.9% fractional conversion of the chloride to aluminum oxide
Cl by weight. (Al2O3) given by eq 5
Original, dense crystals of the chloride become porous z ) mass fraction of species (AlCl3‚6H2O) in original
as their decomposition proceeds. The largest measured reactant
porosity amounted to 57% (0.63 cm3/g). The true density zi ) mass fraction of species “i” in solids
of decomposed solid slowly increases with the extent of
Greek Letters
decomposition from 1.65 to 1.95 g/cm3. The particle
(apparent) density, needed for fluidization calcula- F ) solid density, g/cm3, kg/m3
tions,18 decreased with the progress of reaction from 1.3 FHe ) skeletal, true (helium) solid density, g/cm3, kg/m3
to 0.82 g/cm3. FHg ) particle, bulk, apparent (mercury) solid density,
The pore-size distributions of the decomposed samples g/cm3, kg/m3
indicate that the radii of pores are distributed over the τ ) exposure time, s, min
1-104 nm range. The most probable pore size in the
Subscripts
solids increased from 20 to 400 nm in dependence on
the extent of decomposition. Cl ) chloride(s), (Cl-)
Al ) aluminum (Al)
i ) species “i”
Acknowledgment
This research was supported by Grant 203/02/0002 Literature Cited
from the Grant Agency ČR. The authors thank Mrs. H.
Součková and Mrs. H. Šnajdaufová for the textural (1) Park, K. Y.; Kim, J. K.; Seong, J.; Choi, Y. Y. Production of
measurements. Assistance of Mrs. E. Macháčková and Poly(aluminum chloride) and Sodium Silicate from Clay. Ind. Eng.
Mr. J. Horáček in the chemical analyses is also appreci- Chem. Res. 1997, 36, 2646.
(2) Park, K. Y.; Park, Y. W.; Youn, S. H.; Choi, S. Y. Bench-
ated.
Scale Decomposition of Aluminum Chloride Hexahydrate to
Produce Poly(aluminum chloride). Ind. Eng. Chem. Res. 2000, 39,
4173.
Nomenclature (3) Marchessaux, P.; Plass, L.; Reh, L. Thermal Decomposition
Abbreviations of Aluminum Hexahydrate Chloride for Alumina Production.
Proceedings of the 108th AIME Annual Meeting; Light Metals
ACHH ) aluminum chloride hexahydrate, AlCl3‚6H2O Section, New Orleans, LA, February 18-22, 1979;p 189.
BAC ) basic aluminum chloride, Al2(OH)4Cl2‚H2O or Al2O3‚ (4) Naumann, R.; Petzold, D.; Paulik, F.; Paulik, J. J. Investi-
2HCl‚2H2O gation of Thermal Decomposition of Aluminum Chloride Hexahy-
PAC ) Poly(aluminum chloride); basic aluminum chloride drate under Dynamic and Quasi-isothermal Condition. Thermal
dissolved in water Anal. 1979, 15, 47.
(5) Petzold, D.; Naumann, R. J. Thermoanalytical Studies on
Symbols the Decomposition of Aluminum Chloride Hexahydrate. Thermal
Anal. 1981, 20, 71.
∆HA° (298 K) ) standard enthalpy of reaction A at 298.15 (6) Barin, I. In Thermochemical Data of Pure Substances, 3rd
K, kJ/mol ed.; Collaboration with Platzki, G.; VCH: Weinheim, Germany,
∆w ) (weight) mass loss [) wo - w(τ)], g 1995.
∆w/wo ) relative (weight) mass loss (7) Mu, J.; Perlmutter, D. D. Thermal Decomposition of Inor-
h p ) average particle size determined by sieving, mm
d ganic Sulfates and Their Hydrates. Ind. Eng. Chem. Process Des.
e ) fractional porosity of solid ) (FHe - FHg)/FHe ) (VpFHe)/ Dev. 1981, 20, 640.
(1 + VpFHe) ) VpFHg (8) Mu, J.; Perlmutter, D. D. Thermal Decompositions of
E ) effective (apparent) activation energy, fitted param- Carbonates, Carboxylates, Oxalates, Acetates, Formats, and Hy-
eter, J/mol droxides. Thermochim. Acta 1981, 49, 207.
ko ) preexponential factor, Arrhenius constant, fitted (9) Hartman, M.; Veselý, V.; Jakubec, K. Thermal Decomposi-
tion and Chemism of Hydronium Jarosite. Collect. Czech. Chem.
parameter, 1/s Commun. 1987, 52, 939.
K ) quantity defined by eq 2, bar7.5 (10) Hartman, M.; Trnka, O.; Svoboda, K.; Kocurek, J. Decom-
ln ) base-e, natural (Napierian) logarithm ) 2.30259 log position Kinetics of Alkaline-Earth Hydroxides and Surface Area
log ) base-ten (Briggsian) logarithm ) 0.434294 ln of Their Calcines. Chem. Eng. Sci. 1994, 49, 1209.
Mi ) molar mass of species (MH2O ) 18.015; MHCl ) 36.461; (11) Hartman, M.; Trnka, O.; Veselý, V. Thermal Dehydration
MAlCl3‚6H2O ) 241.431; MAl2O3‚2HCl‚2H2O ) 210.913; MAl2O3 ) of Magnesium Hydroxide and Sintering of Nascent Magnesium
101.961; g/mol Oxide. AIChE J. 1994, 40, 536.
n ) effective (apparent) order of reaction, fitted parameter (12) Hartman, M.; Martinovský, A. Thermal Stability of the
ni ) amount of species “i”, mol Magnesian and Calcareous Compounds for Desulfurization Pro-
cesses. Chem. Eng. Commun. 1992, 111, 149.
P ) dissociation pressure given by eq 3, bar
(13) Kim, J. H.; Yim, Y. J. Effect of the Particle Size on the
PHCl ) partial pressure of HCl, bar Thermal Decomposition of -Hexanitroisowurtzitane. J. Chem.
PH2O ) partial pressure of water vapor, bar Eng. Jpn. 1999, 32, 237.
r ) pore radius, nm (14) Hartman, M.; Trnka, O.; Veselý, V.; Svoboda, K. Thermal
rp ) radius of pore, nm Dehydration of the Sodium Carbonate Hydrates. Chem. Eng.
R ) ideal gas-law constant ) 8.31441, J/(mol K) Commun. 2001, 185, 1.
6598 Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005

(15) Hartman, M.; Svoboda, K. Physical Properties of Magnesite (18) Yates, J. G. Fundamentals of Fluidized-Bed Chemical
Calcines and Their Reactivity with Sulfur Dioxide. Ind. Eng. Processes; Butterworth: London, 1983.
Chem. Process Des. Dev. 1985, 24, 615.
(16) Hartman, M.; Pata, J.; Coughlin, R. W. Influence of
Porosity of Calcium Carbonates on Their Reactivity with Sulfur Received for review January 3, 2005
Dioxide. Ind. Eng. Chem. Process Des. Dev. 1978, 17, 417. Revised manuscript received May 2, 2005
(17) Hartman, M.; Svoboda, K.; Trnka, O.; Čermák, Ji. Reaction Accepted June 12, 2005
between Hydrogen Sulfide and Limestone Calcines. Ind. Eng.
Chem. Res. 2002, 41, 2392. IE058005Y

You might also like