Vous êtes sur la page 1sur 4

Waste Treatment & Minimization

Overview

The Pyro- and Hydrometallurgical


Processing of Uranium-Containing
Waste
R.G. Reddy

INTRODUCTION
Contamination caused by improper
disposal of radioactive waste1 is of great
concern to industrialized countries.
Many nuclear testing sites and training
grounds are polluted by high levels of
radionuclides. There are two major categories of radioactive waste: high level
and low level. High-level waste (both
solids and liquids) result from spent fuel
generated from commercial and defense
R&D reactors, isotope production reactors, fuel reprocessing, fuel fabrication,
and weapon production activities. Lowlevel waste (solids and liquids) are generated from reactor operation, fuel cycle,
reprocessing, defense R&D and testing,
institutions, industry, uranium mining
and milling, and extraction activities.
6.00

2.00
0.00
2.00
4.00
6.00
8.00
0.00

LV = BV

Inc
Bac reasi
teri ng
al A
dhe

Bacterial Adhesion not


Dependent on SV

LV < BV

Bacterial
Adhesion

20.00

on

ng esi
asi dh
cre l A
De cteria
Ba

BV = Constant
40.00
60.00
SV (ergs/cm2)

80.00

Figure 1. The change in Gibbs energy of


adhesion as a function of solid surface tension.

2001 January JOM

PYRO- AND
HYDROMETALLURGICAL
PROCESSING
Bio-Redemediation
Bio-remediation of radionuclides is a
technique for cleaning uranium-contaminated soils. In this context, Reddy and
coworkers4 proposed a theoretical approach to predict a selective bacterial
attachment on surfaces of uranium-contaminated soil. A thermodynamic model5
was developed to calculate the Gibbs
energy of adhesion for several bacteriamineral systems. The kinds of bacteria
that could be used in uranium-leaching
processes were evaluated by using the
Gibbs energies of adhesion and the proposed limiting conditions for adhesion.
Their results indicated, for systems containing higher LV liquids or SV solids
(where is surface tension, S is solid, L is
liquid, and V is vapor), bacterial adhesion is strongly favored for uraniumcontaminated soils having negatively
charged surfaces. Thiobacillus ferroxidans
has been recommended as a suitable
microorganism for the uranium-contaminated soil. The change in Gibbs energy of bacterial adhesion per-unit surface area is given as:
Gadh = BS BL SL

sio

Gadh (ergs/cm2)

LV < BV

No Bacterial Adhesion

4.00

Large quantities of these radioactive substances contaminate the soil and ground
water, especially at U.S. Department of
Energy facilities.2 For example, the concentration of uranium in contaminated
soil and tailings of mine waste3 was 0.012
and 0.0020.11 percent respectively. Due
to stringent environmental regulations,
methods for disposing of radioactive
waste are being thoroughly investigated
and several processes for containing the
radioactive waste have been attempted.

(1)

where BS, BL, and SL are bacteriumsolid, bacterium-liquid, and solid-liquid


interfacial tensions, respectively. The
bacterial adhesion is favored if the Gadh
is negative, but not if the Gadh is positive. The change of Gibbs energy of adhesion as a function of mineral surface
tension SV at a fixed BV (surface tension
of bacteria) is shown in Figure 1. This
figure also shows the effect of LV on

100

Bacteria

80
Uranium Extraction (%)

As large amounts of uranium-containing


(both high and low-level) waste are generated from activities such as fuel fabrication,
fuel reprocessing, and R&D, concern is growing over the safe disposal of these radioactive
materials. Over the past few decades, numerous disposal options have been investigated,
including pyrometallurgical high-temperature fusion/vitrification, hydrometallurgical processing, biological remediation, polymerization, clay back-filling, and zeolite adsorption. This paper reviews the techniques,
thermodynamics, and kinetic aspects of processing uranium-containing waste. In particular, the removal of uranium waste by
zeolite adsorption and mechanisms are discussed. In addition, the pyro- and hydrometallurgical conceptual flow sheet is proposed
for the disposal of uranium-containing waste.

60

40

20

0
0

Sterile

100

200
300
Time (hour)

400

500

Figure 2. Uranium extraction from contaminated soil using thiobacillus ferroxidans.

Gadh. For liquids having LV < BV, a


critical point (X) is reached after which
bacterial adhesion is favored. Similarly,
for liquids having LV > BV, a point (Y) is
reached after which detachment of bacterium takes place. Regarding the change
in Gibbs energy of adhesion as a function of uranium-contaminated liquid
surface tension, the curve goes through
maxima. The adhesion is favorable for
LV < 40 erg/cm 2 and LV > 60 erg/cm2 for
BV in the range of 5568 erg/cm 2.
Using the same soil, Reddy et al.6 studied the effect of thiobacillus ferroxidans on
the uranium extractability. They analyzed the soil samples by x-ray diffraction and x-ray fluorescence. As shown in
Figure 2, remarkable enhancement in
the uranium extraction was observed in
the presence of bacteria: about 38% of
the uranium was extracted under sterile
conditions compared to 99% with bacteria addition after about 400 hours of the
remediation process. For the extraction
of uranium using thiobacillus ferroxidans
at a fixed temperature of 35C, AvaramiErofeer equation was found to be suitable in describing the rate of extraction.
[1n(1)]1/2 = R t

(2)

where, is the fraction of uranium extracted in time (t), and R is the rate
constant containing the specific rate-constant per-unit-mass per-unit-time. In a
study of the effect of the particle size of
uranium-contaminated soil on the rate
21

of uranium extraction, the rate of uranium extraction from soils increased as


soil particle size decreased. That effect
could be related to the uranium distribution in the soils. Fine soil particles contained more uranium,4 hence, the rate of
uranium extraction increased with increasing uranium concentration.
Microorganisms, including bacteria,
algae, fungi, and yeast accumulate heavy
metals and radionuclides with high efficiency. Khalid et al.7 attempted to recover microbial leached uranium present
in aqueous solutions by free and immobilized axenic cultures such as Aspergillus amsta and found that more than 90%
of the uranium present in the solution
was sorbed by few cultures. These studies found that the solutions pH was
critical to uranium uptake, suggesting
an optimum value of 5.0 for biosorption
of uranyl acetate. Microbial biosorption
of uranium from waste streams of nuclear
fuel processing plants was carried out
with two organisms:8 Saccharomyces
cerevisiae and Pseudomonas aeruginosa. It
was observed that the accumulation of
uranium on Saccharomyces cerevisiae was
extra-cellular and its rate and extent were
dependent on pH, temperature, and the
effect of anions and cations. Uranium
accumulation by Pseudomonas aeruginosa
occurred intra-cellularly, was extremely
rapid, and was not responsive to environmental parameters. The cells revealed
the presence of uranium during examination by electron microscopy and energy dispersive x-ray analysis.
The potential use of biomass to treat
toxic and radioactive waste with relevant mechanisms and recovery of metals was studied.9 Polyvinyl polyacraylate
has been used for the removal and recovery of uranium from uranium-refining wastewater,10 recovering uranium
completely from solution at pH 6 and
desorbed easily with 0.1 M HCl.
A few studies have been reported on
the leaching of waste by alkalis and acid
solutions to recover uranium.11,12 A silicate-based fused-cast ceramic waste was
tested for uranium recovery by direct
sulfuric-acid leaching.11 Although sulfuric-acid/hydrogen-peroxide leaching
is simpler to operate than alkali leaching, the recoveries are marginal. An eightstage, continuous-ion-exchange plant
100.0

U% Absorbed

80.0

60.0
40.0

20.0

Chabazite
Clinoptilolite
Erionite
Mordenite

0.0
1.0 2.0 3.0 4.0 5.0 6.0 7.0
pH

8.0 9.0 10.0

Figure 3. Uranium absorption on zeolites versus solution pH at 298 K.

22

Table I. Enthalpy Values (and Compositions) of Synthesized High-Silica Zeolites


at 298 K
Composition (wt.%)
Enthalpy
Zeolite
SiO2
Al
Na
K
Ca
Mg
(kJ/mol)
ZSM12
95.42
0.043
0.157

0.058
0.016
902.0 1.3
ZSM5
97.48

0.071

0.068
0.012
902.5 1.3
ZSM11
94.09

0.247

0.076
0.024
902.5 1.4
SSZ24
97.56
0.011
0.234

0.013
0.005
903.5 1.3
Hexagonal faujasite 86.75
3.542
2.152
0.229
0.239
0.032
900.2 1.3
Cubic faujasite
95.73
0.119
0.013

0.005
0.001
897.1 1.2

operating on a counter-current fluidizedbed concept to recover uranium from


low-grade solutions (about 120 mg/L
U3O8) was investigated for the recovery
of uranium.13
HIGH-TEMPERATURE
FUSION, VITRIFICATION,
AND OTHER STUDIES
Buelt14 briefly reviewed the vitrification and bituminization processes currently available for the confinement and
stabilization of radioactive waste. Studies on uranium-waste confinement by
Davidovits15 indicated the advantages of
geopolymer cements. The technique enables fast dewatering of the structure,
and the material remains stable even up
to 1,273 K. As a result, the radioactive
elements are trapped within the zeolitic
geopolymeric framework, enhancing the
innocuity of the confinement.
The role of clay on the disposal of
nuclear waste has also been studied. 16-18
In one such study,16 the current repository methods for high-level waste were
reviewed with an emphasis on specific
functions of clay in the geomedia and as
barrier material. The study proposed that
physicochemical properties of clay be
determined pertinent to site characterization and application in waste-disposal
practices. The effectiveness of
phyllosilicates as a component of the
buffer material was evaluated by comparing bentonite and kaolinite, which
are the two important members of the
phyllosilicates.17 Phyllosilicates, often
referred to as clay materials, are composed of well-defined silicate layers.
Bentonite is a 2:1 layer silicate which has
a swelling potential, high specific-surface area, and high cation-exchange capacity (CEC). In contrast, kaolinite is a
1:1 layer silicate with a low swelling
potential, low specific-surface area, and
low CEC. Clay minerals such as illite,
with chemical and physical properties
between those of bentonite and kaolinite, fall within the bentonite/kaolinite
behavior.1921 Bentonite, due to its high
swelling potential, low hydraulic conductivity, low effective porosity (0.01
compared to 0.5 for kaolinite), high sorption capacity (surface area 80 hm2 /kg,
compared to 2 hm2/kg for kaolinite),
would be the most effective clay mineral
for use as a buffer component. Bentonite
was found to remain physically and
chemically stable under the expected

conditions of waste disposal.


Various vitrification processes, desired
composition of the glasses to vitrify the
various waste, and the desirable properties of the resulting glasses have been
reviewed.22,23 Of particular interest are
the rates of reaction of glasses with water under conditions that would likely
occur in the future disposal of glasses.
Moegling24 reviewed the status and storage of high-level waste in the form of
borosilicate glass and discussed the process by which it formed a convenient
storage matrix for high-level waste.
In studies of the borosilicate glass (noncrystalline) and synthetic minerals (crystalline) as a media for the immobilization of high-level radioactive waste,25
borosilicate glass comprising SiO2 (64%
by wt.), B2O3 (23%) and Na 2O (13%) was
found to be most suitable, meeting the
required properties for containment. The
choice of borosilicate glass as a highlevel nuclear waste-containment material was evaluated26 based on certain
observations, including natural and artificial glass hydration and reaction, experimental results of glass leaching and
reaction, and the calculated release of
radionuclides in the waste form.
Also studied was the immobilization
of radioactive elements in crystal lattice
of synthetic rock.27 On cooling, the melt
forms a crystalline material, SYNROC,
in which the radioactive material becomes incorporated.28 The mineral
perovskite (CaTiO3) is an example of a
typical constituent of SYNROC.
The structure of glasses depends on
the composition of the melt. Glasses have
a certain amount of free volume in the
form of interstitial sites or holes. These
holes of different sizes are bonded by
oxygen atoms and are suited for fixing
cations from the wide range of radioactive waste. The glass formation characteristics of spent pot liner, a hazardous
solid waste, was studied,29 including the
effects of sand, limestone, dicalcium silicate dust, and MgO bag-house dust additions on the viscosity, glass formation,
and leachability characteristics.
ZEOLITES
Thermodynamics
In studies30,31 on the thermodynamic
properties of zeolites, the enthalpies of
various synthesized high-silica zeolites
(e.g., ZSM-5, ZSM-11, ZSM-12, and SSZ24) as well as cubic and hexagonal
JOM January 2001

faujasite were measured by high-temperature solution calorimetry. The results obtained in those studies are given
in Table I. These zeolites have enthalpies
of 897.1903.5 kJ/mol and are only 714
kJ/mol less stable than quartz32 (910.7
1.0 kJ/mol), and only 0.56.5 kJ/mol less
stable than silica glass. Amorphous silica
derived from gels is higher in activation
enthalpy than fused glass (010 kJ/mol).
Since zeolites and amorphous silica occupy an overlapping range of energies33 ,
their enthalpies are not a strong function
of molar volume or related parameters
such as framework density or pore size.
Adsorption Kinetics
Considerable interest has been shown
toward the adsorption of heavy metal
cations, such as uranium, by zeolites.3440
Zeolites, which are able to adsorb as
much as their own weight of aqueous
and organic liquids, are a family of complex aluminosilicates having a three-dimensional network structure with
unique structural and chemical characteristics.41 Zeolites are also crystalline
and have pores interconnected by large
and small channels. The structural openings in zeolites are occupied by alkali,
alkaline earth cations, and loosely held
water molecules. The cations are exchangeable and the exchange is very
similar to crystalline SYNROC.42
Zeolites have excellent ion-exchange,
selective-adsorption, and catalytic properties.43 Grant et al.44 developed a system
to treat a variety of low-level, liquid,
radioactive waste streams generated

100 m
Figure 4. Scanning-electron micrograph of
uranium-adsorbed chabazite at pH2 and
298 K.

100 m
Figure 5. Scanning-electron micrograph of
uranium adsorbed chabazite at pH6 and 298
K.

2001 January JOM

Table II. Reaction Rates for Uranium Adsorption on Natural Zeolites at pH 2 and 298 K
Rate of Reaction (min.1)
Chabazite
Clinoptilolite
Erionite
Mordenite
1.7
0.5
0.5
0.3
0.9
0.3
0.23
0.27
0.29
0.14
0.11
0.12
0.096
0.044
0.035
0.038

Time
(min.)
10
30
100
316

from a nuclear-fuel processing plant using a non-regenerative-zeolite ion-exchange process. The minimum fluidizing velocity was calculated using the
equation developed by Ergun.45 The required fluidization velocity increased
with the voidage of the bed until it was
equal to the terminal velocity of the particles as the voidage approached one.
Once the terminal velocity of a particle
was exceeded, the particle would be entrained and carried from the bed. Relationships were developed between laboratory and full-scale systems.
Reddy et al.34,46 studied the adsorption
of uranium from aqueous solutions on
various natural zeolites (i.e., chabazite,
clinoptilolite, erionite, and mordenite).
Their results showed that uranium species are strongly adsorbed on zeolites
between pH 69 and that the adsorption
strongly depended on the pH of the
solution and the type of zeolite. The
reaction rates of uranium adsorption on
different zeolites at pH 2 and temperature 298 K are given in Table II. The
values of reaction rates are higher for
chabazite in comparison with other zeolites. From these studies, it can be concluded that, of all the zeolites tested,
chabazite has the maximum adsorption
capacity for uranium. This observation
could be related to the Si/Al ratio. If all
the factors are kept equal, a lower Si/Al
ratio means a higher zeolite capacity. In
a study on the removal of heavy metals
and other cations from wastewater,
Zamzow et al.47 observed maximum recoveries of lead with chabazite, suggesting that higher Si/Al ratios have higher
zeolite capacities. Bulk density must be
considered when selecting a zeolite since
the capacities are measured per unit
weight. Among the four zeolites studied, chabazite has the least bulk density
(most porous).34
Reddy et al.34,46 also studied the temperature dependence of uranium adsorption on chabazite and mordenite. The
results showed an activation-enthalpy
value of 3.6 kJ/mol in the temperature
range 293313 K. These results indicate
that the reaction is more favorable at
slightly high temperatures (Table III).
The uranium adsorptions on various
zeolites as a function of solution pH
were investigated in nitric-acid solutions
and at 298 K. The experimental results
(Figure 3) show that the amount of zeolite adsorption increases as the pH of the
solution increases. Scanning-electron
micrographs of two chabazite samples

absorbed with uranium at pH2 and pH6


(temperature 298 K) are shown in Figure
4 and Figure 5.
When investigating the sorption of
uranium on activated charcoal, similar
results were reported,48 whereas on pozzolan (a natural alumino-silicate with a
SiO2 /Al2O3 ratio of 4.56), 98% uranium
adsorption was observed.49
When experiments were conducted to
determine the required time and pH of
the aqueous solution for the adsorption
of metal ions by zeolites, the percent of
adsorption was found to increase with
increases in the pH of the aqueous solution, reaching a maximum value at pH >
5. Chabazite has the highest adsorptivity
on uranium (6+) at pH < 4, whereas
clinoptiololite, erionite, and modernite
have almost the same ability for uranium adsorption at pH > 6. The proposed adsorption mechanism involves
the prior hydrolysis of uranium (6+).
Based on solution thermodynamics, four
different hydroxyl complexes are considered in aqueous solutions depending
on the pH:46
UO22+, UO 2 (OH)+ , (UO2)2(OH)3 2+,
and (UO2)3(O H)5+
A general equation for the formation of
hydroxyl complexes can be given as:
xUO22+ + yH2O (UO2 )x (OH)y (2x-y) +
yH+
(3)
A significant change of the zeta potential from negative to positive for some
zeolites could be possible between pH 4
and pH 6. The zeta potential becomes
negative for chabazite at pH > 5, whereas
for clinoptilolite, erionite, and mordenite
the change occurs between pH 4 and pH
6. In all zeolites, pore diameters are determined by the free aperture resulting
from four-, six-, eight-, ten-, or 12-member rings of oxygen atoms, which have
maximum values of 2.6 , 3.4 , and 4.2
, 6.3 , and 7.4 , respectively. Due to
Table III. Uranium Adsorption on
Mordenite as a Function of Time at
Different Temperatures from Solutions
with an Initial Concentration of 50 ppm
and pH 6
Temperature
Time
Adsorption
(K)
(min.)
(%)
293
10
82
100
82
298
10
98
100
97
308
10
98
100
99

23

Uranium Wastes

Pyrometallurgy/
Electrometallurgy

Precipitation

Adsorption on
Zeolites

Solution to
Recycle

Solids to
Discard

Salt Cake

U-ppt. Solids

U-with Zeolites

Hydrometallurgy

Solids to
Discard

Concentrated
Cake

High Temp. Fusion/


Vitrification

Recovery of U

Storage/Disposal

Market
Figure 6. A pyro- and hydrometallurgical conceptual flow sheet for the disposal/recovery of
uranium from waste containing radionuclides.

the puckering of apertures, the effective


free aperture of the pore structure may
be reduced and it may be elliptical. The
ionic diameter of the non-hydrated UO2 2+
is 3.84. Eight and ten rings, having
diameters 4.0 5.5 and 4.4 7.2 ,
form the intercrystalline channels in
clinoptilolite. The porous structure in
mordenite is due to interconnected large
and small channels. The large channels,
which are parallel to c axis, have 12 linked input openings with a diameter of 6.7
7.0 , and the small eight-linked
channels have openings perpendicular
to b with a diameter of 2.9 5.7 . The
opening diameter in erionite is 3.6
5.2 with eight linked rings. Chabazite
is the most porous zeolite with a threedimensional system of channels having
eight linked windows of diameter 3.0
4.4 . An ion-exchange reaction is
possible between UO22+ in solution and
cation (such as Na+ ) in the inter-crystalline sites of zeolites.
Diffusion in zeolites is complex when
compared to conventional catalysts and
requires experimental work to understand the transport of molecules. Diffusion in zeolites with pore diameters of 4
9 has been termed as the configurational regime, and it has a large range of
diffusivities50,51. In this regime, small differences in configurational structures can
have a large effect on diffusivity. The
other two regimes, Knudsen and regular, are for pore sizes of 10 to 100 and
above 100 , respectively.
Zeolite particles contain both microand macro-scopic channels; hence any
or both kinds of channels can control the
diffusion rate. The diffusivities are
strongly dependent on molecular size
and shape; therefore, the rate of adsorption will exhibit a strong dependence on
molecular size.
24

Process Flow Sheet


Because there is no clear process flow
chart for the disposal/recovery of uranium-bearing waste based on the pyro/hydro metallurgical techniques, a conceptual flow sheet (Figure 6) is proposed.
Among these techniques, adsorption of
uranium by natural zeolites has good
potential, as it is simpler than available
techniques.
ACKNOWLEDGEMENTS
The author is pleased to acknowledge the
financial support from U.S. Department of
Defense, sponsored project No. DSWA 0197-10012.
References
1. D. Sanning and H. Stietzel, Proc. 4th Int. kfk / TNO Conf. on
Contaminated Soil 93, vol. I, ed. F. Arendt et al. (Dordrecht/
Boston/London: Kluwer Academic Publishers, 1993), pp.
1125.
2. R.G. Riley, J.M. Zachara, and F.J. Wobber, Chemical Contaminants on DOE Lands and Selection of Contaminant Mixtures
for Subsurface Science Research, U.S. Department of Energy
(April 1992), p. 29.
3. A.F. Oluwole et al., Proc. 8th Symp. Radiation Measurements
and Applications, vol. 353, N 1-3 (Amsterdam, Elsevier Science B.V., 1994), pp. 499502.
4. R.G. Reddy, S Wang, and A.E. Torma, Proc.
Biohydrometallurgical Technologies, vol. 1, ed. A.E. Torma, J.E.
Wey, and V.I. Lakshmanan (Warrendale, PA: TMS, 1993),
pp. 715729.
5. R.G. Reddy, Proc. Mineral Processing, ed. R.W. Smith et al.
(Warrendale, PA: TMS, 1991), pp. 453468.
6. S. Wang, R.G. Reddy, and A.E. Torma, Proc. 2nd Int. Symp.
Year 2000 and Beyond, ed. H.Y. Sohn (Warrendale, PA: TMS,
1994), pp. 583589.
7. A.M. Khalid et al., Proc. Bio-hydrometallurgical Technologies,
vol. I, ed. A.E. Torma et al. (Warrendale, PA: TMS, 1993), pp.
299307.
8. G.W. Strandberg, S.E. Shumate, II, and J.R. Parrott, Jr.,
Applied and Environmental Microbiology, 41 (1) (1981), pp. 237
245.
9. G.M. Gadd, Chemistry and Industry (London), 13 (1990), pp.
421426.
10. A. Nakajima and T. Sakaguchi, J. Chem. Technol. Biotechnol.,
36 (6) (1986), pp. 281289.
11. R.L. Miller, J.M. Welch, and J.E. Flinn, Miner. Metall.
Process., 2 (1) (1985), pp. 5152.
12. I.P. Smirnov et al., Atomnaya Energiya, 78 (4) (1995) , pp.
288289.
13. F. Delmas and C.J. Coelho (Paper presented at the 5th
Symp. on Ion Exchange held at Lake Balaton, Hungary, May
1986).
14. J.L. Buelt, Proc. 4th Int. kfk/ TNO Conf. on Contaminated Soil
93, vol. II, ed. F. Arendt et al. (Dordrecht/Boston/London:
Kluwer Academic Publishers, 1993), pp. 639645.15. J.
Davidovits, Concrete International, 16 (12) (1994), pp. 5358.

16. S.Y. Lee and R.W. Tank, Applied Clay Science, 1 (1985), pp.
145162.
17. D.W. Oscarson and S.C.H. Cheung, Atomic Energy Can
Ltd. Rep. (1983), p. 37.
18. F. Bucher and M. Muller-Vonmoos, Applied Clay Science,
4 (1989), pp. 157177.
19. R.E. Grim, Clay Minerology (New York: McGraw-Hill,
1975).
20. R.N. Yong and B.P. Warkentin, Soil Properties and Behavior
(New York: Elsevier Scientific Publishing Co., 1975).
21. J.K. Mitchell, Fundamentals of Soil Behavior (New York:
John Wiley and Sons, 1975).
22. R. Soundararajan, Journal of Hazardous Materials, 24 (1990),
pp. 199212.
23. J.A.C. Marples, Glass Technology, 29 (6) (1988), pp. 230
247.
24. S.D. Moegling, Proc. Specialty Conference on Environmental
Engineering (New York: ASCE, 1989), pp. 15459.
25. P.A. Tempest, Nuclear Technology, 52 (1981), pp. 415425.
26. M.E. Morgenstein and D.L. Shettel, Jr., Proc. of the 4th
Annual Intl. Conf., American Nuclear Society and American
Society of Civil Engineers (Reston, VA: ASCE,1993), pp. 1728
1734.
27. A.E. Ring Wood, Safety Disposal of High-Level Nuclear
Reactor Waste: A New Strategy (Canberra, Australia: Australian National University Press, 1978).
28. H.D. Megaw, Proc. Phys. Soc., 85 (133) (1946).
29. R.G. Reddy and K.N. Hebbar, Proc. Symp. on New
Remediation Technology in the Changing Environmental Arena,
ed. B.J. Scheiner et al. (Littleton, CO: SME, 1995), pp. 171175.
30. I. Petrovic et al., Chem. Mater., 5 (1993), pp. 18051813.
31. A. Navrotsky, MRS Bulletin, 22 (5) (1997), pp. 3541.
32. CODATA Task Group, J. Chem. Thermodyn., 8 (1976), p.
603.
33. J.Y. Ying, J.B. Benziger, and A. Navrotsky, J. Am. Ceram.
Soc., 76 (1993), p. 1465.
34. M.K. Mohan and R.G. Reddy, Global Symposium on Recycling, Waste Treatment and Clean Technology, vol. I., ed. I.
Gaballah, J. Hager, and R. Solozabal (Warrendale, PA: TMS
1999), pp. 857868.
35. F.C.V. Renaud, J. Chem. Soc. Faraday Trans., 86 (24) (1990),
pp. 40954099.
36. R.T. Pabalan et al., Mat. Res. Soc. Symp. Proc., 294 (1993),
pp. 777782.
37. M.T. Olguin and M. Solache, Separation Science and Technology, 29 (16) (1984), pp. 21612178.
38. H. Mimura and K. Akiba, J. of Nuclear Science and Technology, 30 (5) (1993), pp. 436443.
39. H. Mimura, K. Akiba, and K. Kawamura, J. of Nuclear
Science and Technology, 31 (5) (1994), pp. 46369.
40. S. Forberg, T. Westermark, and L. Falth, Proc. of 3rd Int.
Symp. on the Scientific Basis for Nuclear Waste Management
(New York: Plenum Press, 1980), pp. 227234.
41. D.E.W. Vaughan, Properties of Natural Zeolites: Occurrence,
Properties, Use, ed. L.B. Sand and F.A. Mumption (New York:
Pergamon Press, 1978), pp. 353355.
42. R.A. Munson, Properties of Natural Zeolities, U.S. Bur. of
Mines Rept. of Invest. 7744 (1973), pp. 13.
43. J.V. Smith, Zeolite Chemistry and Catalysis, Monogr. 171,
ed. J.A. Rabo (New York: Am. Chem. Soc., 1976).
44. D.C. Grant et al., Proc. of AIChE Symp. Series, Adsorption
and Ion Exchange: Fundamentals and Applications, 84 (264)
(New York: AIChE, 1988), pp. 1322.
45. S. Ergun, Chem. Eng. Prog., 48 (89) (1952).
46. R.G. Reddy and Z. Cai, Light Metals 1996, ed. W.R. Hale
(Warrendale, PA: TMS, 1996), pp. 11731180.
47. M.J. Zamzow et al., Separation Science and Technology, 25
(13-15) (1990), pp. 15551569.
48. M. Saleem and M. Afzal, Separation Science and Technology,
27 (2) (1992), pp. 239253.
49. U. Cetin and R.K. Mehta, Pre-print (Paper presented at
the SME Annual Meeting, Orlando, Florida, March 1998).
50. P.E. Eberly, Jr., Zeolite Chemistry and Catalysis, ed. J.A.
Rabo, ACS Monogr. 171 (Washington D.C.: ACS, 1976), p.
392.
51. D.M. Ruthven, Diffusion in Zeolites: A Review of Recent
Progress, in the Properties and Applications of Zeolites, ed. R.P.
Townsend (London: Chem. Soc., 1980), p. 44.
52. M.K. Mohan and R.G. Reddy, Recent Res. Devel. Metallurg.
& Materials Sci., 3 (1999), pp. 4969.

R.G. Reddy is the ACIPCO Chair Professor at the University of Alabama, Tuscaloosa.
For more information, contact R.G. Reddy, The
University of Alabama, A-129 Bevill, Tuscaloosa,
Alabama 35487-0202; phone: (205) 348-4246;
fax: (205) 348-2164; e-mail: rreddy@coe.
eng.ua.edu.

Visit JOM
on the
World Wide Web at
www.tms.org/pubs/
journals/JOM/jom.html
JOM January 2001

Vous aimerez peut-être aussi