You are on page 1of 8

Ceramic Waste Forms

for Actinides
Gregory R. Lumpkin*

he concept of nuclear waste forms based on minerals that contain


actinides has led to the development of polyphase and special-purpose
crystalline ceramics. These ceramics are considered by many to be
attractive media for the long-term storage of actinides in geological repositories.
The available data show that monazite, pyrochlore, zircon, and zirconolite
are all highly durable in both natural and synthetic aqueous systems at low
temperatures. In comparison, perovskite is prone to dissolution and conversion
to anatase and other secondary alteration products. The titanate and silicate
phases of interest become metamict (amorphous) as a result of irradiation.
Several compounds, including monazite, cubic zirconia, and the defect fluorite structure types with Zr on the B site, exhibit the attractive property of
radiation resistance. These results, together with other materials properties,
are discussed briefly with respect to criteria for waste form performance.

A natural example of
radiation damage,
mechanical behavior,
and chemical alteration.
The SEM image shows
allanite from Amelia,
Virginia, with high (green) and low (white) levels of
Th. Radiation damage in the main part of the crystal
has compressed the crystalline, low-Th zone, resulting
in brittle failure and the development of numerous
tension cracks. Late aqueous fluids have migrated
through the cracks and caused a preferential
alteration (blue areas) of the amorphous allanite. The
zoned allanite behaves as a composite material.

testing. The best of these waste


forms are superior to borosilicate
glass in terms of aqueous dissolution. However, a major concern is
that the actinide host phases will
undergo a crystalline to amorphous transformation due to alpha
decay of the actinides. This may
lead to volume expansion, cracking,
and reduced chemical durability
(Weber et al. 1998).

KEYWORDS: ceramics, nuclear waste form, actinides,

Radiation damage effects in these


materials have been studied using
synthetic samples doped with
short-lived 238Pu or 244Cm, natural
samples (containing Th and U),
and various heavy ion irradiation techniques. Actinide doping experiments are the most relevant to real waste forms,
and natural analogue studies provide complementary information on the long-term behavior in geological environments. However, many investigators have chosen to simulate the damage produced by alpha-decay processes by
using ion irradiation techniques under controlled experimental conditions (Begg et al. 2001; Lian et al. 2003, 2006;
Lumpkin et al. 2004b).

radiation damage, aqueous dissolution

INTRODUCTION
Beginning in the 1970s with research on alternative waste
forms to borosilicate glass (McCarthy 1977) and followed
by the invention of Synroca synthetic rock made up of
stable titanate minerals (Ringwood et al. 1988) considerable effort has been dedicated to the development and scientific evaluation of ceramics designed for the safe disposal
of nuclear wastes. These diverse wastes range from the
reprocessed spent fuels from commercial power reactors to
high-grade plutonium derived from decommissioned
nuclear weapons. Over the past 25 years or so, titanate
ceramics have evolved from the original polyphase assemblages to specific compositions based largely on a single
phase. Advances in waste form development and testing
have been complemented by numerous mineralogical
investigations of the analogous crystalline phases in geological environments (Lumpkin 2001; Lumpkin et al. 2004a).
Information obtained from these studies has important
implications for validation of the long-term performance of
nuclear waste forms for disposal in geological repositories.
One of the major goals of research on crystalline ceramic
waste forms is to provide a host matrix capable of providing
a much higher level of chemical durability than borosilicate
glass when placed in a geological repository. Tailored
ceramics (Harker 1988), the Synroc titanate waste forms,
and their special-purpose derivatives are reasonably well
developed and have been the subject of extensive dissolution
* Australian Nuclear Science and Technology Organisation
PMB1, Menai, NSW 2234, Australia
E-mail: grl@ansto.gov.au

ELEMENTS, VOL. 2,

PP.

365372

POLYPHASE WASTE FORMS


It is instructive to examine the polyphase waste forms first,
as they have been studied more extensively and have themselves spawned a number of the single-phase waste forms
for actinides (Stefanovsky et al. 2004). FIGURE 1A shows the
typical microstructure of a Synroc-C ceramic with numerous triple points and very low porosity. Other forms of Synroc,
e.g. polyphase ceramics for disposal of partially reprocessed
spent fuels, were designed around the use of pyrochlore as
the main actinide host phase (TABLE 1). Lumpkin et al.
(1995) provided an extensive summary of the partitioning
behavior of lanthanides, actinides, and other elements in
polyphase Synroc samples. Zirconolite has a strong preference for the smaller lanthanide ions and tetravalent
actinides (FIG. 2). Large lanthanide ions (e.g. Nd, Ce) and
trivalent actinides, however, prefer the larger Ca site of perovskite. In general, such partitioning may have a direct
impact on both the chemical durability and the mechanical
properties of any polyphase waste form. This may be especially important for proposed glass-ceramic waste forms,
wherein actinides partition strongly into the crystals as

365

D ECEMBER 2006

Analytical electron microscope data showing how


actinides and lanthanides (indicated as a function of their
ionic radii) partition between zirconolite and perovskite in Synroc. The
dark line is the average for Na-free Purex waste (PW). The actinides follow a different trend line (green). Increased Na, as in Japanese waste
(JW), lowers the partition coefficient D, pushing more lanthanides and
actinides into perovskite (orange region). This can be reversed by substitution of K for Na (blue region).

FIGURE 2

approaching 0.001 g/m2/d or lower after about 90 days at


temperatures of 70100C. Dissolution rates of the rare
earth elements are about an order of magnitude lower than
those of the soluble elements, and Ti and Zr rates are lower
still, often reaching detection limits within a few days.
Ringwood et al. (1988) suggested that the dissolution rates
of the actinides (Np, Pu, Am, Cm) are approximately two
orders of magnitude lower than those of the more soluble
elements and exhibit typical values of less than 10-4 g/m2/d
in short-term tests (less than 60 days). Solubility calculations indicated that the solubility of Pu controls the dissolution behavior in all experiments except for those conducted under the most acidic conditions. The solubility
limits of Np and Am were not exceeded under any of the
experimental conditions.

opposed to the glass matrix. Actinide doping experiments


are required in order to demonstrate the performance of
these materials.

Alteration depths for certain phases have also been reported


(Smith et al. 1992; Lumpkin et al. 2004a). After one year at
150C in pure water, the average alteration depths are about
350 nm for Al-rich oxides, 160 nm for perovskite and intermetallic phases, and 15 nm for hollandite. Similar experiments conducted at 70C gave values of about 80 nm for Alrich oxides, 65 nm for the intermetallic phases, 60 nm for
perovskite, and 5 nm for hollandite. These estimates are in
good agreement with surface SEM and cross-sectional TEM
observations. The microscopy work also shows that, at 70C,
the major alteration product consists of an amorphous or
poorly crystalline TiOH film derived mainly from the dissolution of perovskite. At 150C the alteration of perovskite
to anatase + brookite is the dominant surface feature. Other
secondary phases include monazite, AlOH phases, and
poorly crystalline FeOH material. Zirconolite appears to
be resistant to chemical alteration under these conditions.

Most of the basic dissolution tests on polyphase materials


have been conducted in the temperature range of
25200C, with periodic replacement of the aqueous fluid.
Results of these earlier studies revealed that the dissolution
rates of the most soluble elements, like Al, Ca, Sr, Mo, Cs,
and Ba, exhibit a non-linear decrease with time, typically

Measurements of accelerated radiation damage and dissolution tests on polyphase Synroc samples (both Na-free and
Na-rich types) doped with 244Cm have been reported by
Mitamura et al. (1992, 1994). Up to the maximum dose
achieved during this study (1.3 1015 /mg), the Na-free
samples showed a consistent decrease in density with

(A) Transmission electron microscope image showing the


microstructure of Synroc-C prepared by hot pressing
under reducing conditions. Note the lack of porosity, the lamellar structure in zirconolite (Z), and the twin domains in perovskite (P). Other
phases are hollandite (H) and rutile (R). (B) High-resolution transmission
electron microscope image showing the microstructure of a pyrochlore
(Py) waste form containing plutonium, prepared by high-temperature
sintering. Note the structural intergrowth of zirconolite and pyrochlore.
Inserts are electron diffraction patterns of the two phases.

FIGURE 1

ELEMENTS

366

D ECEMBER 2006

increasing dose. A Na-rich specimen, however, showed a


significant change in the dosedensity behavior at a dose of
8.5 1014 /mg. This change was possibly related to the
onset of cracking (which was not a major problem in the
Na-free samples). For the longest dissolution period, the
release rates of Ca and Sr in the Na-free samples were within
error of one another and increased from approximately
0.0003 g/m2/d to 0.002 g/m2/d. Perovskite is clearly the
major source of these elements in solution, even though
the damage rate of zirconolite appears to be much higher
according to the X-ray diffraction results.

SINGLE-PHASE WASTE FORMS


These crystalline ceramics are primarily designed for the
encapsulation of actinides and other elements (e.g. short-lived
Sr and Cs) that would arise from the possible separation of
high-level wastes (HLW) into various fractions; single-phase
waste forms could also host Pu from dismantled nuclear
weapons (TABLE 1). Technically, they are not always singlephase materials, as demonstrated by the important class of
pyrochlore-based ceramics designed for Pu immobilization.
As shown in FIGURE 1B, these ceramics consist of pyrochlore
with some zirconolite, often in layered intergrowths. Depending upon the bulk composition, the pyrochlore ceramics
may also contain some perovskite and brannerite. It is also
common practice to include an excess component, e.g.
TiO2, in the formulation in order to improve the processing
flexibility or to eliminate an unwanted (e.g. soluble) phase.

Pyrochlore
Cubic pyrochlore is a derivative of the fluorite structure and
corresponds to the general formula A2B2X6Y, in which A
and B are 8- and 6-coordinated cation sites and X and Y are
4-coordinated anions sites. Typically, A = Na, Ca, Y, lanthanides, and actinides, and B = Ti, Zr, Nb, Hf, Ta, Sn, and
W. Many more elements are found in natural samples, and
hundreds of different pyrochlore compositions have been
synthesized. Prototype waste form compositions are usually
based on the CaUTi2O7 end-member. Other actinides substitute directly for U in this compound. The neutron
absorber Hf substitutes directly for Ti on the B site, whereas
Gd occupies the A site and substitutes for Ca and U.

Dissolution studies on ion-irradiated titanate samples generally indicate that maximum Y and lanthanide release
rates may be anywhere between no change and 20 times
higher as a result of amorphization (Begg et al. 2001). Similar pyrochlore samples with Zr on the B site have been
studied, and these samples retained crystallinity at a dose
that renders the titanate amorphous. After a maximum
time of 28 days, the dissolution rates of the irradiated and
unirradiated zirconate samples were the same, dropping
from 0.11 g/m2/d to about 0.050.09 g/m2/d for Ca and
from 0.010.1 g/m2/d to 0.0020.006 g/m2/d for Gd. These
values are similar to those of the unirradiated titanate
pyrochlore. The variability of results in the studies noted
above probably relates to the experimental details, including temperature and redox state, nature of the apparatus
used, and quality of the samples (e.g. porosity, surface area,
tendency to crack, soluble impurity phases).
Lumpkin and Ewing (1988) have shown that natural
pyrochlores with Nb, Ta, Ti, and minor Sn and Zr on the B
site are subject to amorphization at a critical dose of about
1016 /mg. This dose is about 23 times higher than that for
CaPuTi2O7 (Clinard et al. 1984a) and is consistent with conditions encountered during storage at elevated temperatures in the Earths crust (Lumpkin 2001). This dose, however, increases with the geological age of the sample for
unmetamorphosed samples; this points to long-term
annealing via atomic diffusion. This study also provided a
systematic evaluation of strain, crystalline domain size, and
microstructural details as a function of dose. An example of
natural pyrochlore, associated with zirconolite, from Ti-rich
veins of the Adamello massif in northern Italy is shown in
FIGURE 3. These actinide-rich, radiation-damaged, and compositionally zoned crystals have coexisted for forty million

Weber et al. (1986) were among the first investigators to


conduct dissolution tests on single-phase pyrochlore samples (Gd2Ti2O7) doped with 244Cm. The dissolution tests
were limited to annealed, fully crystalline samples and fully
amorphous samples, and were exercised at 90C in pure
water for 14 days. Weight losses of 0.02% and 0.05% for the
crystalline and amorphous pyrochlore samples, respectively,
were measured. The dissolution rate of Cm increased by a
factor of 17 as a consequence of amorphization.
Significant new results are now available (Strachan et al.
2005; Icenhower et al. 2006) for synthetic pyrochlore samples doped with Ce or PuU and for single-phase Gd2Ti2O7
for comparison. The PuU pyrochlores were fabricated with
238Pu or 239Pu in order to compare samples with high and
low alpha-decay doses for a range of flow rate to surface
area ratios, while at the same time minimizing artifacts (e.g.
effect of F from teflon, effect of atmospheric CO2 on pH,
and effect of radiolysis at low flow rates). Results for solutions
buffered at pH = 2 and temperatures of 8590C reveal that a
steady-state, forward reaction rate can be determined for
high values of the flow rate to surface area ratio. Under these
conditions, release rates of 0.0013 to 0.0002 g/m2/d were
obtained for the different pyrochlore compositions. However,
the most important result of this work is that the dissolution
rates of the crystalline and X-ray amorphous (238Pu-doped)
samples are the same, within experimental error.

ELEMENTS

367

TABLE 1

SUMMARY OF WASTE FORMS FOR HIGH-LEVEL WASTES


FROM REPROCESSING AND SPECIAL-PURPOSE FORMULATIONS
FOR PU AND ACTINIDES

Waste form

Main phases

Application/waste loading

Synroc-C

Zirconolite, perovskite,
hollandite, rutile

HLW from reprocessing,


up to 20 wt%

Synroc-D

Zirconolite, perovskite,
spinel, nepheline

US defense wastes, 6070 wt%

Synroc-F

Pyrochlore, perovskite, uraninite

Conversion of spent fuel,


approximately 50 wt%

Tailored
ceramics

Magnetoplumbite, zirconolite,
spinel, uraninite, nepheline

US defense wastes, to 60 wt%


or higher

Pyrochlore

Pyrochlore, zirconolite-4M,
brannerite, rutile

Separated actinides or Pu,


up to about 35 wt%

Zirconolite

Zirconolite, rutile

Separated actinides,
up to about 25 wt%

Monazite

Monazite

Actinidelanthanide wastes,
up to about 25 wt% actinides

Zircon

Zircon

Pu-rich wastes from dismantled


nuclear weapons

Glass-ceramics

Titanite, zirconolite, pyrochlore,


perovskite, nepheline, sodalite,
alumino-silicate glass

Canadian wastes (low actinide


content), complex legacy
wastes, intermediate
level wastes

Others

Britholite, kosnarite, murataite,


crichtonite

Proposed as host phases for


actinides and lanthanides

D ECEMBER 2006

impurities. All are resistant to amorphization under most


conditions. However, these materials also require higher
processing temperatures, and they exhibit reduced chemical flexibility. The solubility of zirconia in pure water at low
temperatures and near-neutral pH values is quite low,
prompting proposals to employ such materials directly as
inert matrix fuels and nuclear waste forms that can be used
in nuclear reactors and then placed in geological repositories. Dissolution tests generally indicate elemental release
rates between 10-4 g/m2/d and 10-7 g/m2/d, depending on
the experimental conditions.

Zirconolite

Backscattered SEM image (colorized) showing intergrowth of natural pyrochlore (Py, gray) and zirconolite
(Z, blue, green, yellow) from Ti-rich veins, Adamello, Italy. Pyrochlore
(containing about 30 wt% UO2) occurs as overgrowths on zirconolite, is
amorphous due to alpha-decay damage, and exhibits a darker rim due
to minor hydration. Zirconolite is chemically zoned, with up to about
25 wt% Th + U oxide, and its structure ranges from crystalline to amorphous. Note the absence of cracking from differential swelling.

FIGURE 3

years without cracking or significant loss of U and Th. This


provides a nice illustration of the long-term performance of
the pyrochlore waste form for Pu shown in FIGURE 1B.
The radiation damage properties of many synthetic III-IV
pyrochlore and defect fluorite materials containing Na, Ca,
Cd, and lanthanides on the A site and Ti, Zr, Sn, Hf, Nb, and
Ta on the B site have now been studied in detail using Kr
ion irradiation. As noted previously (see Ewing et al. 2004),
the incorporation of Zr at the B site of the pyrochlore structure promotes retention of the crystalline structure to very
high ion fluences. By radiation resistance or tolerance,
it is understood that the heavy ion or alpha-recoil collision
cascade still forms but anneals back to the original crystalline configuration on a very short timescale (e.g. picoseconds). The effect of composition on the radiation tolerance
of pyrochlore is shown in FIGURE 4A. Here, we have plotted
the critical temperature above which the material remains
crystalline (Tc) versus the metaloxygen radius ratio (rM/rx,
in which rM is the average of the A and B cation radii). This
is a convenient way of plotting, as it reflects Paulings Rules
and also effectively separates the data into the main groups
consisting of the titanate and stannate pyrochlores. Within
these two groups, Tc increases as rM/rX increases due to the
effect of increasing A-site cation size. These results are
related in part to classical radius ratio arguments and to the
x(48f) anion positional parameter. In general, the radiation
tolerance of pyrochlore correlates with the gradual structural transformation toward the fluorite-like geometry,
energetic considerations, and increased ionic bonding (e.g.
Minervini et al. 2000; Lian et al. 2003, 2006; Ewing et al.
2004; Lumpkin et al. 2004b; Trachenko et al. 2005). These
effects can be modelled empirically, as shown in FIGURE 4B,
and used to predict the critical amorphization temperature
of compounds in the III-IV pyrochlore composition space
(Lumpkin et al. 2006).

Cubic Zirconia
These compounds are essentially fluoritedefect fluorite
structures based on the general formula MO2-x with M = Ca,
lanthanides, actinides, and other elements simulating
ELEMENTS

Zirconolite is a derivative of the fluorite structure type and


can be considered as a condensed version of pyrochlore.
The prototypical end-member composition is CaZrTi2O7 for
the 2M polytype. Different polytypes (e.g. 3T, 3O, 4M) arise
through solid solution toward lanthanide and actinide endmembers. Actinide and lanthanide elements may substitute
on both the Ca and Zr sites, with the appropriate chargebalancing mechanisms. Charge compensation may be
achieved in some cases via substitution of elements like Mg,
Al, Fe, Nb, Ta, and W for Ti. The observed natural compositions are complex, illustrating a high level of chemical flexibility, and this has been confirmed to a large extent
through experimental work. Zirconolite becomes amorphous when doped with 238Pu or 244Cm or when subjected
to heavy ion irradiation (Clinard et al. 1984b; Weber et al.
1986; Wang et al. 2000). The critical amorphization dose
and total volume swelling are nearly identical to those of
pyrochlore. Both natural and synthetic zirconolites exhibit
a high level of resistance to aqueous dissolution, and even
amorphous samples with accumulated doses up to ~2 displacements per atom retain U and Th and their daughter
products in the structure. Zirconolite also exhibits exceptional corrosion resistance below 250C in various acidic
and basic fluids over a wide range of pH. Between 250 and
500C at elevated pressure (50 MPa), zirconolite is subject to
partial corrosion in acidic and basic media, but significant
corrosion occurs only above 500C (Lumpkin et al. 2004a).
The dissolution kinetics of zirconolite have been determined as a function of pH using pure water in single-pass
flow-through tests at temperatures of 75C and lower
(Zhang et al. 2001). These authors studied a Ce-, Gd-, and
Hf-doped zirconolite containing about 16 wt% UO2, and
the measured release rates of Ti and U indicate that zirconolite dissolves congruently after about 20 days following an initial period during which U is released at a somewhat faster rate than Ti. The limiting rate constants are
equivalent to U release rates of 6.4 10-7 to 1.3 10-5
g/m2/d for zirconolite over the entire pH range of 212 and
temperature range of 2575C. The dissolution rate of zirconolite is characterized by a shallow v-shaped pattern with
a minimum near pH = 8, similar to the results obtained for
pyrochlore.
Ion irradiation studies have been performed on synthetic
samples using 1.0 MeV Kr or 1.5 MeV Xe ions (Wang et al.
2000). Results showed that the radiation response of endmember zirconolite-2M, CaZrTi2O7, varies with ion mass
and energy. In this example, the heavier Xe ions resulted in
a higher Tc value (654K for Kr and 710K for Xe). Additional
results for Kr ions showed that Tc ranges from 550 to 1020K,
with the highest values obtained from samples with the
highest Nd and Ce contents (substituting for Ca and Zr),
regardless of the polytype structure. Relative to the pure
end-member, two samples with Nb and Fe substituting for
Ti had the lowest Tc values. These results indicate that there
is a significant relationship between radiation tolerance and
cation mass in the zirconolite system.
368

D ECEMBER 2006

Perovskite

Perovskite is an ABX3 structure type based on a framework


of octahedral B-site cations and large A-site cations. The
major A-site cations include Na, Ca, Sr, lanthanides, and
minor actinides. The B site is occupied predominantly by Ti
and minor Al in nuclear waste form compositions. Ion irradiation and natural analogue studies indicate that the critical amorphization dose of perovskite is higher than that of
pyrochlore and zirconolite by a factor of about 25. Dissolution studies, however, demonstrate that perovskite is the
least durable phase present in polyphase waste forms
(Lumpkin et al. 2004a). Perovskite reacts quickly with aqueous fluids to form an amorphous TiOH film at temperatures below about 100C or crystalline TiO2 polymorphs at
higher temperatures. Using cross-sectional electron
microscopy and atomic force microscopy, Zhang et al.
(2005) illustrated the epitaxial nature of anatase growth on
single-phase perovskite specimens in aqueous dissolution
experiments conducted at 150C. Thermodynamic calculations show that perovskite is unstable with respect to titanite, titanite + quartz, rutile, and rutile + calcite (Nesbitt et al.
1981). The range of groundwater compositions used in this
study was representative of groundwaters emanating from
dunites, peridotites, serpentinites, rhyolites, granites, and
limestones. Therefore, from a natural systems and experimental viewpoint, perovskite is not the best host phase for
actinides (or radioactive Sr) in waste forms destined for a
geological repository.

Brannerite
Brannerite, ideally UTi2O6, consists of layers of Ti octahedra
connected by columns of U octahedra. It is a minor but
actinide-rich phase in some of the pyrochlore-based compositions designed for the disposal of Pu. Brannerite may
account for up to 20 percent of the U and 15 percent of the
Pu in these waste forms (Thomas and Zhang 2003). Natural
and synthetic brannerites can incorporate substantial amounts
of Ca, REE, Th, and other elements. In both cases, the incorporation of lower-valence cations on the A site may be charge
balanced by oxidation of some U4+ to U5+ and/or U6+ ions.
Synthetic samples are easily amorphized by ion irradiation.
Electron microscopy studies show that most natural brannerites with ages greater than about 10 Ma are fully amorphous due to alpha-decay damage and are commonly
altered by natural aqueous fluids. Altered natural brannerite typically loses U, and the concentration may decrease to
approximately 1 wt% UO2 in the most heavily altered areas.
The most important findings to emerge from experimental
studies (e.g. Zhang et al. 2003) are that certain species, e.g.
phthalate, will increase the solubility of titanium. In bicarbonate solutions, however, the uranium release rate is
strongly dependent on bicarbonate concentration. In acidic
solutions, the dissolution of brannerite involves a preferential release of uranium. TEM results have shown that brannerite exposed to a solution with a pH = 2 at 90C for four
weeks produces a relatively small amount of reaction product, which consists mainly of polycrystalline anatase.
When exposed to a pH 11 solution at 90C for four weeks,
a fibrous secondary phase is formed on the surface. This
phase is amorphous and contains a significant proportion
of U6+, as indicated by surface analytical studies. Thomas
and Zhang (2003) have shown that the aqueous dissolution
of brannerite in a open atmosphere can be modelled as a
function of pH using two reaction steps: oxidation of U4+ at
the surface followed by release of U6+ into solution, which
is catalyzed by protons under acidic conditions or carbonate species under alkaline conditions. This is an important
advance, as the release of uranium at 40C can be predicted
quantitatively from this model.

ELEMENTS

(A) Critical amorphization temperature of pyrochlore and


defect fluorite compounds versus the metal to oxygen
ionic radius ratio. Trends for samples with Ti and Sn on the B site are
shown in red and blue, respectively. Radiation tolerance is promoted by
smaller A cations in each series. For a given A cation, the critical temperature is lowered dramatically by substitution of Hf and especially Zr
on the B site. (B) Three-dimensional surface of the predicted critical
temperature for amorphization of III-IV pyrochlore and defect fluorite
compounds versus the individual A-site and B-site cation radii. This surface was calculated from an empirical model based on structural, energetic, and bonding terms.

FIGURE 4

Zircon
Zircon, ZrSiO4, is a common accessory mineral found in a
variety of geological environments. The major elemental
impurity in natural zircon is Hf, which substitutes for Zr.
Trace to minor amounts (generally 5000 ppm or less) of
other elements may be present, including Ca, lanthanides,
and actinides on the Zr site and P on the Si site. Higher concentrations have been reported but are exceptional. Amorphization, with a critical dose of about 4 1015 /mg in natural and actinide-doped samples, and total volume swelling
of up to 18 percent are well-known characteristics of zircon

369

D ECEMBER 2006

(e.g. Holland and Gottfried 1955; Weber et al. 1998). The


mineral is highly durable in most environments and commonly survives weathering to be recycled in the Earths crust.
To investigate the effect of radiation damage on dissolution
rates, Ewing et al. (1982) conducted experiments on natural
zircons at 87C in an aqueous solution containing 5 wt%
KHCO3. The results showed that the dissolution rate
increases by nearly two orders of magnitude, from 2 10-4
g/m2/d to 2 10-2 g/m2/d, as the alpha-decay dose increases

from 6 1013 /mg to greater than 1 1016 /mg, spanning


the crystallineamorphous transformation. Hydrothermal
experiments using chloride fluids containing H, Al, and Ca
at temperatures of 175 to 650C have produced reaction
rims on zircon grains, with rim thickness dependent on
alpha-decay dose. Experiments on heavily damaged natural
zircons also showed severe loss of Si, U, Th, and Pb and gain
of H2O species, Al, and Ca (Geisler et al. 2001). Studies using
natural samples have also dramatically illustrated the
effects of alpha-decay damage on physical properties (Chakoumakos et al. 1987, 1991; Zhang and Salje 2001) and
chemical behavior, with loss of up to 4050% of the U due to
aqueous alteration in certain zircon crystals (Lumpkin 2001).
Ion irradiation studies have been published for the orthosilicates (A = Zr, Hf, Th) using 0.8 MeV Kr or Xe ions. Only
minor differences in the dosetemperature response of
ZrSiO4 were observed when irradiated with the two different ions. Even the critical temperatures of these silicate
compounds indicate a similar behavior under Kr ion irradiation (Tc ~9001100K). Under certain conditions, silicate
zircons may decompose into the component oxides, tetragonal
ZrO2 or HfO2 and amorphous SiO2 (Meldrum et al. 1999).

Monazite
Like zircon and thorite, monazite also has ABO4 stoichiometry, but the crystal structure is monoclinic and consists of
chains of alternating BO4 tetrahedra and AO9 polyhedral
sites. These chains are cross linked by edge sharing with the
AO9 polyhedra, effectively closing off open tunnels and creating a structure that is approximately 10% denser than the
zircon structure type. Natural monazite contains up to
27 wt% UO2 + ThO2 and remains crystalline in spite of high
accumulated alpha-decay doses; however, monazite can be
amorphized by heavy ion irradiation, and the different
compounds have been studied in some detail as a function
of temperature. Irradiated, fully amorphous synthetic monazite has excellent aqueous durability and is roughly equivalent to the corresponding crystalline samples.
Monazite is highly insoluble in most hydrothermal and
low-temperature fluids; however, solubility may be
enhanced in aqueous fluids with low pH, low phosphate
content, or high F concentrations, which can lead to the
formation of REE-fluoride complexes. At temperatures
below 250C, the solubility of monazite in aqueous solutions decreases with increasing temperature (see Boatner
and Sales 1988; Lumpkin et al. 2004a). Experiments recently
reported by Oelkers and Poitrasson (2002) have provided
results on the steady-state dissolution rates of monazite at
temperatures of 50230 C, pH ranging from 1.5 to 10, and
variable flow rate and surface area. Using a natural sample
as the starting material, these authors showed that the
release rates of the REEs and U are essentially congruent for
all experimental conditions.
Systematic ion irradiation studies of the orthophosphates
(Meldrum et al. 1997) have been conducted for six monazite-structure compounds (A = La, Pr, Nd, Sm, Eu, Gd) and
six zircon-structure compounds (A = Sc, Y, Tb, Tm, Yb, Lu).
The critical amorphization temperatures of these materials
were found to increase systematically with cation mass

ELEMENTS

from about 330 to 490K for the monazites and from about
470 to 580K for the zircon-structure orthophosphates (compare this with the high values found for silicate zircons).

OTHER COMPOUNDS
A number of materials based on compounds such as britholite, crichtonite, garnet, kosnarite, murataite, and titanite
(sphene) have been proposed as waste forms for actinides,
lanthanides, and other elements (e.g. radioactive Sr and I,
and a range of impurities such as transition metals). With
the possible exception of britholite and titanite, none of
these compounds have been studied to the same extent as
those described above. All have natural analogues, although
the contents of Th and U can be quite low in many natural
samples. Stefanovsky et al. (2004) summarized the work on
these phases, and they will not be dealt with further in this
article (apart from the comparison shown in TABLE 2).

DISCUSSION AND CONCLUSIONS


Generally speaking, nuclear waste forms will be required to
conform to a number of performance criteria. Although
there are at present no universal criteria for acceptance,
numerous scientific articles have stated that waste forms
should exhibit low elemental release rates if and when they
are exposed to water in the geological repository, crystal
chemical flexibility, potential for high waste loadings, processing efficiency (both from an engineering and cost perspective), and suitable physical properties. The latter item
includes mechanical properties, thermal behavior, and
response to alpha-decay damage. Note that radiation damage
may dramatically affect the physical properties of materials
that become amorphous due to such damage. Hence, in
recent years we have seen the emergence of radiation tolerance
as an important research topic. The performance of potential
waste form phases for actinides is summarized in TABLE 2.
Starting with the requirement of aqueous durability, it is
clear that perovskite is the least durable phase present in
the polyphase waste forms. This is shown diagrammatically
in FIGURE 5, where dissolution rates are compared for different
phases as a function of solution pH. Alternatives to perovskite
as hosts for radioactive Sr include the magnetoplumbite
structure type and monazite. Among the titanates,
pyrochlore and zirconolite are the currently preferred
actinide host phases. They exhibit dissolution rates with
minimal variation as a function of pH, incorporate a large
range of impurities, are well established with regard to processing, and are capable of high waste loadings. Although
titanate pyrochlore and zirconolite are rendered amorphous
by alpha-decay processes, the total volume expansion is low
(nearly the same for both phases) and the effect of amorphization on aqueous durability appears to be within
acceptable limits.
Crystalline zircon is generally considered to be highly
durable in aqueous systems. However, there is a clear indication of a difference in the behavior of crystalline and
amorphous zircon in aqueous fluids, and this is supported
by microscopic and micro-analytical studies of natural samples. The use of zircon as a waste form may be compromised
by the unusually large volume swelling and anisotropic
unit cell expansion, leading to cracking and increased dissolution rates. Monazite has the particular advantage of
remaining crystalline to alpha-decay doses well above those
that render zircon, brannerite, pyrochlore, and zirconolite
amorphous. Monazite also exhibits low dissolution rates
and appears to have the unusual property of decreasing solubility up to about 250C. The crystal chemistry of monazite

370

D ECEMBER 2006

is reasonably flexible, but does not allow the large range of


substitutions possible in the titanate pyrochlore and zirconolite structure types.
How do the other radiation-resistant materials perform?
The existing experimental data indicate that cubic zirconia
and the zirconate pyrochlorefluorite phases have chemical
durability comparable to or slightly better than that of the
titanate pyrochlores and zirconolite. However, these phases
lack the crystal-chemical flexibility of the titanates and generally require significantly higher processing temperatures.
Furthermore, they have no direct natural analogues, the
closest minerals being baddeleyite and uraninite. Thus, we
cannot confirm the long-term behavior of these zirconiumbased materials to the same extent as successfully studied
natural pyrochlores, zirconolite, monazite, and zircon.

ACKNOWLEDGMENTS

Aqueous dissolution data for U release from pyrochlore,


zirconolite, and brannerite. Results are also shown for Ca
release from perovskite (green line). This figure shows the excellent performance of pyrochlore and zirconolite in aqueous fluids. Although Ca
release from perovskite is shown, it is clear that this compound should
be avoided as a major actinide host phase.

FIGURE 5

TABLE 2

I am grateful to numerous colleagues in Australia, France,


Germany, Switzerland, the USA, and the UK for their guidance, teamwork, and support over the years. Thanks to R.C.
Ewing and an anonymous reviewer for their assistance in
improving this manuscript. 

SUMMARY OF THE PERFORMANCE CHARACTERISTICS OF INDIVIDUAL PHASES MEASURED AGAINST POTENTIAL


SELECTION CRITERIA AND OTHER MEANS OF EVALUATION (E.G. NATURAL SAMPLES)
Aqueous
durability

Chemical
flexibility

Waste
loading

Radiation
tolerance

Volume
swelling

Natural
analogues

Perovskite
(Ca,Sr)TiO3

Low

Medium

Low

Medium

High

Yes

Pyrochlore
Gd2(Ti,Hf)2O7

High

High

High

Low-high

Medium

Yes

Zirconolite
CaZrTi2O7

High

High

Medium

Low-medium

Medium

Yes

Zircon
ZrSiO4

High

Medium

Low (?)

Low

High

Yes

Monazite
LnPO4

High

Medium

High

High

Low

Yes

Zirconates
Gd2(Zr,Hf)2O7

High

Medium

Medium

High

Low

No

Zirconia
(Zr,Ln,Act)O2-x

High

Medium

Medium

High

Low

No

Medium

Medium

High

Low

Yes

High

Medium

Low (?)

Yes

High

High

Medium

Medium

Rare

High

Medium

Low

Yes*

Titanite
CaTiSiO5

Medium

Medium

Low

Low

Medium

Yes

Minerals of the apatite


group (e.g. britholite)

Medium

Medium

Low

Low

Medium

Yes

Kosnarite
NaZr2(PO4)3

Medium

Medium

Medium

Low

Yes

Brannerite
UTi2O6
Crichtonite
Ca(Ti,Fe,Cr,Mg)21O38
Murataite
Zr(Ca,Mn)2(Fe,Al)4Ti3O16
Garnet
Ca3Zr2(Al,Si,Fe)3O12

* Natural garnets typically do not contain substantial amounts of Th or U.

ELEMENTS

371

D ECEMBER 2006

REFERENCES
Begg BD, Hess NJ, Weber WJ, Devanathan
R, Icenhower JP, Thevuthasan S, McGrail
BP (2001) Heavy-ion irradiation effects
on structures and acid dissolution of
pyrochlores. Journal of Nuclear Materials
288: 208-216
Boatner LA, Sales BC (1988) Monazite. In:
Lutze W, Ewing RC (eds) Radioactive
Waste Forms for the Future. NorthHolland, Amsterdam, pp. 495-564
Chakoumakos BC, Murakami T, Lumpkin
GR, Ewing RC (1987) Alphadecayinduced fracturing in zircon: The
transition from the crystalline to the
metamict state. Science 236: 1556-1559
Chakoumakos BC, Oliver WC, Lumpkin
GR, Ewing RC (1991) Hardness and
elastic modulus of zircon as a function of
heavy-particle irradiation dose: I. In situ
alpha-decay event damage. Radiation
Effects and Defects in Solids 118: 393-403
Clinard FW Jr, Peterson DE, Rohr DL,
Hobbs LW (1984a) Self-irradiation effects
in 238Pu-substituted zirconolite: I.
Temperature dependence of damage.
Journal of Nuclear Materials 126: 245-254
Clinard FW Jr, Rohr DL, Roof RB (1984b)
Structural damage in a self-irradiated
zirconolite-based ceramic. Nuclear
Instruments and Methods in Physics
Research B 1: 581-586

Lumpkin GR, Ewing RC (1988) Alphadecay damage in minerals of the


pyrochlore group. Physics and Chemistry
of Minerals 16: 2-20
Lumpkin GR, Smith KL, Blackford MG
(1995) Partitioning of uranium and rare
earth elements in Synroc: effect of
impurities, metal additive, and waste
loading. Journal of Nuclear Materials 224:
31-42
Lumpkin GR, Smith KL, Gier R, Williams
CT (2004a) Geochemical behavior of host
phases for actinides and fission products
in crystalline ceramic nuclear waste
forms. In Gier R, Stille P (eds) Energy,
Waste, and the Environment: a
Geochemical Perspective, Geological
Society of London Special Publications
236: pp 89-111
Lumpkin GR, Whittle KR, Rios S, Smith KL,
Zaluzec NJ (2004b) Temperature
dependence of ion irradiation damage in
the pyrochlores La2Zr2O7 and La2Hf2O7.
Journal of Physics Condensed Matter 16:
8557-8570
Lumpkin GR, Whittle KR, Rios S,
Trachenko K, Pruneda M, Harvey EJ,
Redfern SAT, Smith KL, Zaluzec NJ (2006)
Radiation damage in pyrochlore and
related compounds. In: Van Iseghem P
(ed), Scientific Basis for Nuclear Waste
Management XXIX, Materials Research
Society Symposium Proceedings 932: 549558

Ewing RC, Haaker RF, Lutze W (1982)


Leachability of zircon as a function of
alpha dose. In: Lutze W (ed) Scientific
Basis for Nuclear Waste Management V.
Elsevier, New York, pp. 389-397

McCarthy GJ (1977) High-level waste


ceramics: materials considerations and
product characterization. Nuclear
Technology 32: 92-104

Ewing RC, Weber WJ, Lian J (2004) Nuclear


waste disposalpyrochlore (A2B2O7):
Nuclear waste form for the immobilization of plutonium and minor actinides.
Journal of Applied Physics 95: 5949-5971

Meldrum A, Boatner LA, Ewing RC (1997)


Displacive radiation effects in the
monazite- and zircon-structure
orthophosphates. Physical Review B 56:
13805-13814

Geisler T, Ulonska M, Schleicher H,


Pidgeon RT, van Bronswijk W (2001)
Leaching and differential recrystallization
of metamict zircon under experimental
hydrothermal conditions. Contributions
to Mineralogy and Petrology 141: 53-65

Meldrum A, Zinkle SJ, Boatner LA, Ewing


RC (1999) Heavy-ion irradiation effects in
the ABO4 orthosilicates: decomposition,
amorphization, and recrystallization.
Physical Review B 59: 3981-3992

Harker AB (1988) Tailored ceramics. In:


Lutze W, Ewing RC (eds) Radioactive
Waste Forms for the Future. NorthHolland, Amsterdam, pp. 335-392
Holland HD, Gottfried D (1955) The effect
of nuclear radiation on the structure of
zircon. Acta Crystallographica 8: 291-300
Icenhower JP, Strachan DM, McGrail BP,
Scheele RD, Rodriquez EA, Steele JL,
Legore VL (2006) Dissolution kinetics of
pyrochlore ceramics for the disposition of
plutonium. American Mineralogist 91:
39-53
Lian J, Chen J, Wang LM, Ewing RC,
Farmer JM, Boatner LA, Helean KB (2003)
Radiation-induced amorphization of rareearth titanate pyrochlores. Physical
Review B 68: 134107
Lian J, Helean KB, Kennedy BJ, Wang LM,
Navrotsky A, Ewing RC (2006) Effect of
structure and thermodynamic stability on
the response of lanthanide stannate
pyrochlores to ion beam irradiation.
Journal of Physical Chemistry B 110:
2343-2350
Lumpkin GR (2001) Alpha-decay damage
and aqueous durability of actinide host
phases in natural systems. Journal of
Nuclear Materials 289: 136-166

ELEMENTS

Minervini L, Grimes RW, Sickafus KE


(2000) Disorder in pyrochlore oxides.
Journal of the American Ceramic Society
83: 1873-1878
Mitamura H, Matsumoto S, Hart KP,
Miyazaki T, Vance ER, Tamura Y, Togashi
Y, White TJ (1992) Aging effects on
curium-doped titanate ceramic containing sodium-bearing high-level nuclear
waste. Journal of the American Ceramic
Society 75: 392-400
Mitamura H, Matsumoto S, Stewart MWA,
Tsuboi T, Hashimoto M, Vance ER, Hart
KP, Togashi Y, Kanazawa H, Ball CJ,
White TJ (1994) -decay damage effects
in curium-doped titanate ceramic
containing sodium-free high-level
nuclear waste. Journal of the American
Ceramic Society 77: 2255-2264

Ringwood AE, Kesson SE, Reeve KD, Levins


DM, Ramm EJ (1988) Synroc. In: Lutze
W, Ewing RC (eds) Radioactive Waste
Forms for the Future. North-Holland,
Amsterdam, pp. 233-334
Smith KL, Lumpkin GR, Blackford MG, Day
RA, Hart KP (1992) The durability of
Synroc. Journal of Materials Research
190: 287-294
Stefanovsky SV, Yudintsev SV, Gier R,
Lumpkin GR (2004) Nuclear waste forms.
In Gier R, Stille P (eds) Energy, Waste,
and the Environment: A Geochemical
Perspective. Geological Society of London
Special Publication 236: pp 36-63
Strachan DM, Scheele RD, Buck EC,
Icenhower JP, Kozelisky AE, Sell RL,
Elovich RJ, Buchmiller WC (2005)
Radiation damage effects in candidate
titanates for Pu disposition: Pyrochlore.
Journal of Nuclear Materials 345: 109-135
Thomas BS, Zhang Y (2003) A kinetic
model of the oxidative dissolution of
brannerite, UTi2O6. Radiochimica Acta
91: 463-472
Trachenko K, Pruneda JM, Artacho E, Dove
MT (2005) How the nature of the
chemical bond governs resistance to
amorphization by radiation damage.
Physical Review B 71: 184104
Wang SX, Lumpkin GR, Wang LM, Ewing
RC (2000) Ion irradiation-induced
amorphization of six zirconolite
compositions. Nuclear Instruments and
Methods in Physics Research B 166-167:
293-298
Weber WJ, Wald JW, Matzke Hj (1986)
Effects of self-radiation damage in Cmdoped Gd2Ti2O7 and CaZrTi2O7. Journal
of Nuclear Materials 138: 196-209
Weber WJ and 11 coauthors (1998)
Radiation effects in crystalline ceramics
for the immmobilization of high-level
nuclear waste and plutonium. Journal of
Materials Research 13: 1434-1484
Zhang M, Salje EKH (2001) Infrared
spectroscopic analysis of zircon:
Radiation damage and the metamict
state. Journal of Physics Condensed
Matter 13: 3057-3071
Zhang Y, Hart KP, Bourcier WL, Day RA,
Colella M, Thomas B, Aly Z, and Jostsons
A (2001) Kinetics of uranium release from
Synroc phases. Journal of Nuclear
Materials 289: 254-262
Zhang Y, Thomas BS, Lumpkin GR,
Blackford M, Zhang Z, Colella M, Aly Z
(2003) Dissolution of synthetic
brannerite in acidic and alkaline fluids.
Journal of Nuclear Materials 321: 1-7
Zhang Z, Blackford MG, Smith KL, Vance
ER, Lumpkin GR (2005) Aqueous
dissolution of perovskite (CaTiO3): Effects
of surface damage and [Ca2+] in the
leachant. Journal of Materials Research
20: 2462-2473 

Nesbitt HW, Bancroft GM, Fyfe WS,


Karkhanis SN, Nishijima A (1981)
Thermodynamic stability and kinetics of
perovskite dissolution. Nature 289: 358362
Oelkers EH, Poitrasson F (2002) An
experimental study of the dissolution
stoichiometry and rates of a natural
monazite as a function of temperature
from 50 to 230C and pH from 1.5 to 10.
Chemical Geology 191: 73-87

372

D ECEMBER 2006

You might also like