Sunteți pe pagina 1din 14

Direct Determination of the Thermodynamic Properties of Uranyl Minerals Important for the

Performance of the Geological Repository at Yucca Mountain

A Subcontract Proposal to Argonne National Laboratory


In Response to the 1-19-05 Solicitation from
the Office of Civilian Radioactive Waste Management (OCRWM)
and the Office of Science and Technology and International (OST&I)

Jeremy B. Fein 1, Peter C. Burns 1, and Alexandra Navrotsky 2

1
University of Notre Dame; Civil Engineering and Geological Sciences
156 Fitzpatrick Hall
Notre Dame, IN 46556

Fein contact info - email: fein@nd.edu; phone: 574-631-6101; Fax: 574-631-9236


Burns contact info – email: burns.50@nd.edu; phone: 574-631-7380; Fax: 574-631-9236

2
NEAT ORU & Thermochemistry Facility; Chemistry Annex; One Shields Avenue
University of California at Davis
Davis CA 95616-8779

Email: anavrotsky@ucdavis.edu; phone: 530-752-3292


Concept Summary
Uranyl mineral phases will be the primary alteration products of spent nuclear fuel in oxygenated waste
repository settings such as Yucca Mountain. It is probable that uranyl minerals forming due to the alteration
of spent fuel in Yucca Mountain will incorporate radionuclides such as Np and Pu, and this process may have
a profound impact upon repository performance. Second and third-generation Source Term models
developed by the S&T program will likely include radionuclide incorporation in uranyl minerals as a key
process. Currently, there is insufficient understanding of the thermochemical parameters of uranyl minerals
to support development of such a model.

The proposed research will determine the solubility and thermodynamic properties of a range of
environmentally-important uranyl mineral phases, and will quantify the effect of Np-substitution on these
parameters. We will conduct solubility and calorimetry experiments in order to provide rigorous constraints on
the Gibbs free energies of formation, enthalpies of formation, and standard entropies for these solids.
Solubilities of representative members of the uranyl oxide hydrates, uranyl silicates, uranyl carbonates, and
uranyl phosphates will be measured both for the end-member phases as well as for selected phases
containing a range of substituted Np within the crystal structure. The solubility experiments will be conducted
as functions of pH, ionic strength, and temperature, and will yield Gibbs free energies of formation of the
phases of interest. Solubility measurements conducted as a function of Np content of the uranyl minerals will
enable us to extract solid phase activity coefficient parameters to use in thermodynamic models that quantify
the effect of the Np on uranyl mineral stabilities and solubilities. The calorimetric data will enable
determination of the enthalpies of formation for these phases, so together the standard entropies can be
determined as well. The resulting internally-consistent thermodynamic dataset will enable determination of the
effect of Np-substitution on the solubilities and thermodynamic stabilities of environmentally-important uranyl
phases, at room temperature and at temperatures relevant to a repository setting.

Research Objective
The objective of this study is to determine the thermodynamic properties for a broad range of
environmentally-important uranyl mineral phases using solubility and calorimetric measurements, producing an
internally-consistent dataset that is useful in modeling spent nuclear fuel alteration and U mobility behavior in
near-surface environments.

Enabling Elements/Background
Safe, permanent disposal of the nation's nuclear waste in a deep geological repository involves unique
scientific and engineering challenges, reflecting the very long-lived radioactivity, chemical complexity and
unique material properties of the waste. The repository must retain for thousands of years a wide variety of
radionuclides with vastly different chemical characteristics. Most of the radioactivity that will be housed in the
proposed repository at Yucca Mountain will be associated with spent nuclear fuel, most of which is from
commercial reactors [about 70,000 tons (heavy metal) of UO2 (LWR) spent fuel by the year 2010]. Safe
disposal of spent fuel requires a detailed knowledge of its long-term behavior under repository conditions, as
well as the potential fate of radionuclides released from the spent fuel after waste containers are breached.

Studies of natural analogues, as well as laboratory experiments using spent UO2 fuel, unirradiated
UO2, and uranium metal, all indicate that spent fuels are unstable under the moist oxidizing conditions
expected in the proposed repository at Yucca Mountain (Finn et al., 1996; Wronkiewicz et al., 1996; Finch et
al., 1999; Pearcy et al., 1994, Finch & Ewing 1992). Both vapor hydration and hydrologically unsaturated drip
tests on spent fuel have demonstrated that alteration rates of the spent fuel are appreciable under repository-
relevant conditions, and that the predominant alteration products are uranyl (U6+) phases, reflecting the
predominance of U in spent fuel (~88 wt.% U) (Finn et al., 1996; Finch et al., 1999).

2
Dissolution of spent fuel may release radionuclides to groundwater and, ultimately, to the human-
accessible environment. Burns et al. (1997) predicted that radionuclides such as Np and Pu may be
incorporated into the uranyl phases that form where spent fuel is altered, thereby having a potentially
profound impact upon the mobility of these radionuclides and on repository performance. Burns et al. (2004)
demonstrated that uranyl phases can incorporate neptunium (as Np5+). In an ongoing S&T-funded program,
Burns is examining the incorporation of Np5+ (and to a lesser extent Pu), as a function of temperature, pH,
and the presence of counter ions, into the structures of uranyl phases that are likely to form as alteration
products of spent fuel in Yucca Mountain. Work to date clearly supports the hypothesis that uranyl minerals
will reduce the mobility of Np5+ in Yucca Mountain.

Development of second and third-generation Source Term models for Yucca Mountain will likely
include the impact of uranyl minerals on the release of Np, Pu and potentially other radionuclides such as Se
and I. Currently available thermodynamic parameters (solubility, Gibbs free energy of formation, heats of
formation) for uranyl minerals that are likely to incorporate radionuclides in Yucca Mountain fall far short of
what is required to develop a defensible Source Term model. In this proposal, we address this shortcoming
by proposing to develop an internally-consistent database of thermodynamic parameters for a range of uranyl
minerals that are likely to be important in Yucca Mountain. We will also take the unprecedented step of
determining the impact of incorporation of trace amounts of Np5+ on the solubility of uranyl minerals.
Giammar and Hering (2004) examined the effect of Na and Cs impurities on uranyl oxide hydrate solubility
products, and found that incorporation of the cations in the uranyl mineral structure dramatically decreases
the equilibrium concentration of uranium in solution. Despite the potential for trace element incorporation to
affect uranium and other radionuclide mobilities, few studies have focused on this effect experimentally, and
there currently are no quantitative models that have been developed to estimate the effect that trace element
incorporation will have.

The importance of quality thermochemical measurements for uranyl minerals for the second and
third-generation Source Term models for Yucca Mountain is the driver for the current proposal. However,
we also recognize the considerable importance of our proposed measurements for understanding the mobility
of uranium in the environment at sites other than Yucca Mountain. Factors that impact U(VI) mobility include
adsorption of aqueous uranyl species onto mineral and bacterial surfaces, the formation of colloids that
uranium is sorbed onto or incorporated into, biologic activity that may reduce uranium, and co-precipitation of
uranium into non-uranium minerals. However, the thermodynamic stability of uranyl minerals is a primary
control on the maximum concentration of uranium that can be present in the aqueous phase in any system of
geologic or environmental interest. Uranium contamination in the subsurface poses a serious risk to the
environment at numerous sites in the U.S. These include sites that were involved in the production of
uranium (uranium mine and mill sites, e.g., Tuba City, Arizona) and the processing and re-processing of
uranium (e.g., Fernald, Ohio; Hanford, Washington; Oak Ridge, Tennessee). DOE sites contain 6.4 billion
cubic meters of contaminated soil, groundwater, and other environmental media. The most common
radionuclides in these contaminated areas are the isotopes of uranium. For example, at the Hanford site
seven distinct uranium plumes have been identified in the 200 and 300 areas (DOE-RL, 1999). Some sites
contain concentrations of uranium in ground water as high as 20 mg/L (DOE-RL, 1999), which is much
higher than the acceptable level for uranium in ground water, 30 μg/L set by the Environmental Protection
Agency in 2000 (EPA, 2000).

Uranyl minerals are chemically and structurally complex, with about 200 species known (Burns,
1999; Finch and Murakami, 1999). The solubilities of these species are poorly characterized in general, but it
is apparent that their solubilities vary dramatically, with uranyl carbonates being among the most soluble, and
uranyl phosphates, vanadates and arsenates being the least soluble (Langmuir, 1978; Grenthe et al., 1992).
Although it is clear that uranyl minerals control the mobility and distribution of uranium in a wide range of
oxygenated environments, the solubilities and thermodynamic properties of these minerals in general are
poorly characterized. For many uranyl mineral phases of environmental interest, thermodynamic data
3
necessary to predict their stabilities and solubilities are either completely lacking or are contradictory and
inconsistent. This lack of internally-consistent thermodynamic data has prompted the formulation of estimation
procedures for critical parameters (e.g., Chen et al. 1999; Chen and Ewing, 2003). These approaches are
useful in the absence of reliable thermodynamic data, however their accuracy and applicability remain
untested. Wanner and Forest (1992) provided the most comprehensive review of the literature on
thermodynamic data for uranium, however they provide data for only a handful of uranyl oxides and uranyl
oxide hydrates. At that time, no experimental thermodynamic or solubility data existed for a large number of
important uranyl minerals, and no data existed at that time for any uranyl silicates. Nguyen et al. (1992)
performed solubility measurements on a number of uranyl silicate minerals, deriving solubility products and
Gibbs free energies of formation for the minerals. However, there is considerable uncertainty associated with
these measurements not only because the experiments were not reversed, so equilibrium was not
demonstrated, but also due to uncertainties associated with the identity and stoic hiometry of the mineral
phases that may have reached equilibrium in the experiments. Casas et al. (1994) derived thermodynamic
data from uranophane solubility measurements, however there is considerable data scatter ranging over one
to two orders of magnitude, and the uranophane was present in the experiments with significant quantities of
other mineral phases. Therefore, unequivocal thermodynamic interpretation of these data are impossible.
There are only a handful of well-constrained experimental studies that provide thermodynamic properties for
environmentally-important uranyl mineral phases, including studies of becquerelite (Casas et al., 1997; Rai et
al., 2002) and peroxide-containing uranyl minerals (Hughes Kubatko et al., 2003).

A number of studie s of uranyl mineral thermochemistry have been conducted in which experimental
difficulties such as a lack of experimental reversal, a high degree of data scattering, or uncertainties in
mineral composition have led to ambiguous thermodynamic interpretations. While there have been a few well-
constrained experimental studies to date, the range of uranyl minerals studied is extremely limited.
Thermodynamic properties have been determined for only a small subset of the potentially important uranyl
oxide hydrates, uranyl silicates, uranyl phosphates and uranyl carbonates. Virtually nothing is known
concerning the temperature dependence of these thermodynamic properties for any uranyl mineral phases,
and virtually nothing is known concerning the effects of solid solution or minor element incorporation on the
thermodynamic stabilities or aqueous solubilities. Despite this lack of data, the experimental studies that have
been conducted provide evidence that well-constrained, well-executed, and properly-interpreted solubility
measurements can provide high quality thermodynamic data for uranyl minerals. Because the Gibbs free
energies of formation are well-constrained for aqueous uranyl species, especially under low pH conditions,
measurements of uranyl mineral solubilities can yield rigorous constraints on the Gibbs free energies of
formation for the mineral phase in each experiment. This thermodynamic parameter, along with its
temperature dependence and/or calorimetric data that yield enthalpies of formation of the uranyl phases of
interest, can be used to estimate the relative and absolute stabilities, and aqueous solubilities, for
environmentally-important uranyl minerals under a range of conditions.

In the proposed research program, we will synthesize and measure the solubility of a wide range of
uranyl mineral phases, including representative phases from the uranyl oxide hydrates, uranyl silicates, uranyl
carbonates, and uranyl phosphates. In order to improve the accuracy and precision of the thermodynamic
interpretation of the data, solubility measurements will be conducted primarily under the low pH conditions at
which UO2+2 is the dominant aqueous uranyl species. Solubility experiments will be conducted as a function
of pH and temperature over the temperature range 25 – 100 oC, experiments will be reversed to rigorously
demonstrate the attainment of equilibrium, and experiments for some uranyl mineral phases will also be
conducted with varying concentrations of mineral impurities to ascertain the effects of solid solution on
thermodynamic stabilities and solubilities. Calorimetric measurements will be made using pure end-member
uranyl phases, as well as using uranyl phases with varying degrees of trace element incorporation. The
experimental results will yield an internally-consistent set of thermodynamic properties for a wide range of
environmentally-important uranyl mineral phases. The creation of this database represents a crucial step in
development of second and third-generation Source Term models.
4
Research Methods
1. Thermodynamic Approach. The proposed experimental solubility measurements will be valuable as a
direct indication of the concentration of uranium and other elements in solution in equilibrium with
environmentally-important uranyl mineral phases. In addition, the solubility experiments will enable
determination of the Gibbs free energies of formation for the mineral phases as well. For example, the
dissolution of soddyite, under the low pH conditions where UO2+2 is the dominant aqueous uranium species,
can be expressed as:

(1) 4 H+ + (UO2)2(SiO4)(H2O)2 (soddyite) ⇔ 2 UO2+2 + SiO2(aq) + 4 H2O

If we define standard states for minerals and for H2O to be the pure phases at the pressure and temperature
of interest, and the standard state for aqueous species to be a hypothetical one molal solution at the pressure
and temperature of interest that behaves as if it is infinitely dilute, then the mass action equation for reaction
(1), assuming that the soddyite in the experiment is pure, is:

a 2UO2 + 2 ∗ a SiO2 ( a q)
(2) K (1) = 4
a H+

where K(1) represents the equilibrium constant for reaction (1), and a represents the activity of the
subscripted aqueous species. If we conduct our solubility measurements significantly below pH 5, then UO2+2
is the dominant uranium aqueous species and we can safely assume that a measurement of total dissolved
uranium is a direct measure of the concentration of UO2+2. Similarly, under these low pH conditions, SiO 2(aq)
is the only aqueous silica species of importance, so a measurement of total dissolved silica yields the
concentration of SiO 2(aq). If we buffer ionic strength with a relatively inert background electrolyte such as
NaClO 4, we can convert the measured molality of UO2+2 to an activity using the extended Debye-Hückel
equation (e.g., Helgeson et al., 1981). Measurements of amorphous silica solubilities as a function of ionic
strength (Marshall and Warakomski, 1980) will be used to calculate activity coefficients for SiO 2(aq) under the
ionic strength conditions of the proposed experiments. The activity of H+ is measured directly through pH
determinations.

With a single solubility measurement (experimental determination of the activities of UO2+2, SiO2(aq) and
H+), we can determine the value of K(1), and additional solubility measurements, conducted as a function of
pH and ionic strength, will enable rigorous, statistically-significant constraints to be placed on this value. While
the value of K(1) is extremely useful on its own for the determination of uranium mobility under conditions
other than those directly studied in the laboratory, the value can also be used to provide constraints on the
Gibbs free energy of formation of the mineral soddyite. The equilibrium constant directly yields a value for the
change in standard state Gibbs free energy for the reaction, Δ Goreaction:

− ΔG o reaction
(3) log K (1) =
2.303 R T

where R and T represent the gas constant and absolute temperature, respectively. The parameter Δ Goreaction
is defined as:
(4) Δ Goreaction = 2 ΔG o f (UO2 +2 ) + ΔG o f ( SiO2 ( a q ) ) + 4 ΔG o f ( H2 O) − 4 ΔG o f ( H + ) − ΔG o f ( Soddyite)

where ΔG o f (i) represents the standard state Gibbs free energy of formation for the species or phase i in the
parentheses. Because the values of ΔG o f (i) are well-established for H2O and for all of the aqueous species

5
of importance under the experimental conditions, the experimental determination of Δ Goreaction enables us to
directly calculate Δ Gof (Soddyite).

Although in theory only one solubility measurement, conducted at a single ionic strength, is sufficient to
determine the value of K(1), and hence Δ Gof (Soddyite), significant improvement in the uncertainty associated
with these values can be obtained if the solubility measurements are conducted both as a function of pH and
as a function of ionic strength. The measurements conducted as a function of pH not only provide additional
constraints on the thermodynamic values, but they also provide constraints on the stoichiometry of the
dissolution reaction. That is, because reaction (1) depends on H+, a fixed relationship exists between the
extent of dissolution that occurs in terms of the uranyl concentration and solution pH, and solubility
measurements conducted as a function of pH serve to test whether such a relationship exists. Solubility
experiments conducted as a function of ionic strength serve to constrain the extrapolation of calculated
thermodynamic properties to zero ionic strength (the reference state). The extended Debye-Hückel equation
enables such extrapolations based on a single solubility measurement, but repeat measurements conducted as
a function of ionic strength place more rigorous constraints on the extrapolation, significantly reducing the
uncertainties associated with the calculated thermodynamic parameters.

If in the experiments we use soddyite that is not pure (UO2)2(SiO4)(H2O)2, but rather involves atomic
substitution at one or more sites (e.g., Burns et al. 2004), then Equation (2) changes to:

a 2UO2 + 2 ∗ a SiO2 ( aq)


(5) K (1) =
a 4 H + ∗ a Soddyite

where a Soddyite represents the activity of (UO2)2(SiO4)(H2O)2 in the solid phase. Therefore, conducting
experiments using uranyl mineral phases that vary in their trace element composition in a controlled and
measureable fashion will enable us to determine the effects of the a Soddyite term on the Gibbs free energies of
formation for the solid and to derive a relationship between the two. Because other radionuclides such as Np
(Burns et al., 2004) are likely to substitute into the structure of uranyl mineral phases in a repository setting, it
is crucial to determine how this substitution affects the stabilities and solubilities of the host uranyl mineral
phases.

Solubility measurements conducted as a function of temperature can be used to derive other


thermodynamic properties of the uranyl mineral phases. If Δ Horeaction and Δ Soreaction are the standard state
enthalpy and entropy changes associated with reaction (1), then the following relationship holds:

− ΔH o reaction ΔS o reaction
(6) ln K (1) = +
RT R

Therefore, determination of the value of K(1) as a function of temperature can be used to constrain values for
Δ Horeaction and Δ Soreaction. Similar to the approach described to constrain Δ Gof (Soddyite) values from calculated
values of Δ Goreaction, we can use the enthalpy and entropy changes of the reaction to constrain the enthalpy
of formation and the standard entropy for the uranyl phase of interest. Standard state enthalpies of formation
have been determined for some uranyl mineral phases (e.g., Hughes Kubatko et al. 2003, 2004a,b) and our
experimental results can serve as checks on these values; however, these enthalpies are known for only a
small number of uranyl phases, and standard entropies have not been determined for any uranyl mineral
phases of environmental interest.

Solution calorimetry in a molten oxide solvent is a well established technique to determine the enthalpy of
formation of a complex oxide from its binary components (Navrotsky, 1977; 1997). Recent calorimetric

6
studies (Hughes-Kubatko et al., 2003; 2004a, b), with calorimetry performed in the Thermochemistry Facility
at UC Davis, have shown that molten sodium molybdate at 700 oC is an effective solvent for uranyl phases.
The uranium remains in the hexavalent state and volatiles (H2O, CO2) are evolved as gases, resulting in a
well defined final state from which the enthalpy of formation can be calculated. Using soddyite,
(UO2)2(SiO4)(H2O)2, as an example, thermodynamic cycles can be written as :

(7) 2UO22+ (dissolved, 700 oC) + SiO 44- (dissolved, 700 oC) + 2H 2O (gas, 700 oC) =
(UO2)2(SiO4)(H2O)2 (solid, 25 0C) Δ H7

(8) UO3 (solid, 25 0C) + SiO 2 (solid, 25 0C) = 2UO22+ (dissolved, 700 oC)
+ SiO42- (dissolved, 700 oC) Δ H8

(9) 2H 2O (liquid, 25oC) = H2O (gas, 700 oC) Δ H9

Then, for the formation reaction from oxides

(10) UO3 (solid, 25 0C) + SiO 2 (solid, 25 0C) + 2H 2O (liquid, 25 oC) =


(UO2)2(SiO4)(H2O)2 (solid, 25 0C) Δ H10

Δ H10 =?Δ H7 + Δ H8 + Δ H9

Δ H7 is the negative of the measured enthalpy of drop solution of soddyite in the sodium molybdate solvent,
Δ H8 is the sum of the enthalpies of drop solution of uranium trioxide and quartz, and Δ H9 is the heat content
of water, while ?Δ H10 is the enthalpy of formation of soddyite from its constituent oxides in their standard
states at room temperature. Once the enthalpy of formation from oxides is known, the enthalpy of formation
from elements can be calculated using tabulated enthalpies of formation of the binary oxides.

While the direct determination of uranyl mineral solubilitie s are of use, the extraction of standard state
Gibbs free energies of formation for the uranyl mineral phases is perhaps of greater importance in modeling
stabilities and solubilities of uranyl minerals under conditions not directly studied in the laboratory. The
calculated ΔG o f (i) values for the uranyl minerals can be used to determine relative stabilities of uranyl
mineral phases under a wide range of fluid composition conditions of environmental and geologic interest. The
ΔG o f (i) values for the uranyl minerals can also be used to calculate equilibrium constant values for any
reaction involving the studied uranyl mineral phases, enabling estimation of the solubility of any mineral
assemblage or condition of interest. Because it is impossible to measure uranium concentrations in equilibrium
with all uranyl phases of environmental interest under all conditions of interest in the laboratory, it is crucial to
be able to estimate uranyl mineral solubilities in order to assess the mobility of uranium in systems not directly
studied in the laboratory. The thermodynamic properties that we determine in the proposed research program
will enable these extrapolation techniques.

2. Research Team and Preliminary Studies. The principal investigators for this project have extensive
experience synthesizing and conducting research with all of the uranyl mineral phases of interest, and with
conducting similar solubility and calorimetric experiments and using the data to extract thermodynamic
parameters. Fein and students have used mineral solubility experiments, conducted over a wide range of
temperatures and pressures and using a wide range of experimental apparatuses, to determine equilibrium
constants for a number of types of water-rock interactions (e.g., Fein, 1994; Fein et al., 1995). For example,
Fein and Hestrin (1994) measured the solubilities of gibbsite, amorphous silica, and quartz in oxalate-bearing
aqueous fluids to 80oC in order to determine the effect of aqueous oxalate complexation on the mobility of the
mineral-forming cations during sedimentary diagenesis. The solubility measurements constrained the
stoichiometries and equilibrium constant values for the important dissolution and complexation reactions, and a

7
similar approach to that outlined above was used to calculate Gibbs free energies and enthalpies of formation
for the species of interest.

Burns and his research group have extensive experience in synthesizing uranyl minerals in high yields
and purities, as well as the characterization of synthetic uranyl minerals (e.g., Burns et al., 2003; 2004). Burns
and his students have published more than 50 papers that reported synthesis techniques for uranyl oxide
hydrates, uranyl silicates, uranyl carbonates, uranyl phosphates, uranyl arsenates, uranyl sulfates, uranyl
chromates, uranyl molybdates, and uranyl nitrates. Many of these studies involved detailed characterization of
the synthesized materials, including complete crystal structure determinations.

As part of her Ph.D. thesis in Peter Burns’ group at Notre Dame, Karrie Ann Hughes-Kubatko visited
the Thermochemistry Facility at UC Davis several times and participated in calorimetric measurements of
enthalpies of formation of a number of uranyl phases. The calorimetry was straightforward and trouble -free.
The phases chosen were a selection of different compositions (including studtite, a peroxide-containing
mineral), dictated by material availability. The values shown in Table 1 were obtained. The proposed work
will target important materials more systematically.

Table 1. Enthalpies of Formation of Uranyl Phases from Binary Oxides

Phase Formula Δ Hof, 298 (from oxides )


kJ/mol
Studtite UO2(O2)(H2O)4 +22.9+3.9*,a
Rutherfordine UO2CO3 -99.1+3.1b
Andersonite Na2CaUO2(CO3)3(H2O)5 -710.6+5.6b
Grimselite K3NaUO2(CO3)3(H2O) -989.3+2.4b
Meta-schoepite UO3(H2O)2 +4.4+2.5c
Synthetic a-UO2(OH)2 -26.6+2.3c
Synthetic CaUO4 -143.1+2.0c
Becquerelite Ca(UO2)6O3(OH)6(H2O)8 -267.5+4.0c
Synthetic Ca(UO2)4O3(OH)4(H2O)2 +20.4+9.4c
Clarkeite Na(UO2)O(OH) -150.6+2.9c
Synthetic Na2(UO2)3O3(OH)2 -1359.5+20.3c
Synthetic Na2(UO2)6O4(OH)6(H2O)7 -320.8+10.0c
• formation from crystalline metal oxides, gaseous CO2, liquid H2O and gaseous O2 as appropriate to
stoichiometry
a. Kubatko-Hughes et al. 2003; b. Kubatko-Hughes et al. 2004a; c. Kubatko-Hughes et al. 2004b

Preliminary proof-of-concept solubility experiments have been conducted by the principal investigators
for this project. Mineral synthesis approaches have been perfected for a number of uranyl mineral phases of
interest, enabling the production of approximately 0.5 gm quantities of each mineral powder to be used in the
solubility experiments. An example of some preliminary solubility data is shown in Figure 1, which depicts
measured total U, Ca, and Si concentrations in a pH 3.0, 0.1 M NaClO 4 solution in contact with uranophane
(Ca[(UO2)(SiO 3OH)]2?5H2O). Although we have yet to test the reversibility of the dissolution reaction or
whether secondary minerals have formed during the experiment, the data shown in Figure 1 demonstrate that
steady-state element concentrations are reached in approximately 150 hours, and that dissolution is congruent
for U, Ca, and Si. That is, on a molar scale at steady state, the concentration of total U and Si in solution is
twice that of Ca, as is expected from the composition of the mineral phase. These element concentration
measurements, in conjunction with pH determinations for the system, enable precise calculation of the
equilibrium constant that governs the dissolution behavior.

8
2
U

Concentration (mM)
1.6

1.2
S

0.8

0.4 C
0

0 100 200 300


Time (hr)
Figure 1. Experimental solubility measurements for uranophane (Ca[(UO2)(SiO 3OH)]2?5H2O) in terms total
dissolved Ca, Si, and U concentrations in a pH 3.0, 0.1 M NaClO 4 solution at 298K. Note that the dissolution
attains a steady-state with respect to each of the mineral-forming cations at approximately 150 hours, and
that dissolution is close to stoichiometric under the steady-state conditions.

3. Research Design and Methods


3.1 Solubility Experimental Procedures - Successful extraction of thermodynamic data from
solubility measurements requires a range of measurements or controls on the experimental systems: 1)
rigorous determination of equilibrium reversal; 2) measurements of the equilibrium pH and the total
equilibrium concentrations of all mineral-forming cations in the system; 3) ionic strength control or
measurement; and 4) determination that secondary mineral phases do not form during experimentation.

Each solubility experiment in the proposed research program will be run in Teflon batch reaction
vessels. Significant concentrations of aqueous uranyl species, in particular, can adsorb onto glass and even
high density plastic reaction vessels, thereby perturbing the system chemistry and increasing the uncertainties
associated with the calculated thermodynamic parameters. In each batch solubility experiment, a synthesized
uranyl mineral powder will be placed in contact with a fixed ionic strength solution, and the pH of the
suspension will be adjusted to a desired value using minute volumes of concentrated perchloric acid or sodium
hydroxide solution. The mass of mineral powder used will be noted, but is not a crucial variable for
constraining the chemistry of the system as long as the mineral is present in excess of the amount that will
dissolve at equilibrium. Ionic strength will be buffered using the inert electrolyte NaClO 4, as the perchlorate
anion does not form aqueous complexes to any appreciable extent with any of the components of the uranyl
mineral phases. Perchlorate concentrations will be maintained low enough to prevent the formation of uranyl
perchlorate phases. Due to complexities in the aqueous speciation of the uranyl ion above pH 5, and to speed
the attainment of equilibrium in the experimental systems, all experiments will be conducted between pH 2
and 5.

Each reaction vessel will be sealed, and placed in a temperature-controlled water bath that is outfitted
with a shaking table assembly in order to simultaneously insure precise temperature control on the
experimental system and adequate mixing of the system during the experiment. The water bath system can
be either cooled or heated to maintain constant temperatures (± 0.5oC) in the range 5-90oC. Each reaction
vessel can hold up to 250 ml of solution, and small samples, each approximately 3 ml, will be taken from each
9
vessel as a function of time over the course of the experiment. At the time of each sample extraction, the
experimental system pH will be monitored and adjusted if necessary. The pH measurements for systems not
at 25oC will employ pH standards of known H+ activity at the temperature of interest. Prior to sampling, the
vessels will be shaken vigorously to insure even distribution of the mineral powder so that the water:mineral
ratio remains constant during the course of the experiment. Each sample will be centrifuged and filtered to
separate the aqueous phase from the mineral phase, centrifuge tubes and a filter apparatus that has been
tested to be inert to cation (especially uranyl) adsorption. The filtrate from each sample will be acidified to
prevent a decrease in aqueous cation concentration due to precipitation.

Each experimental sample will be analyzed for dissolved total metal concentrations, using inductively
coupled plasma optical emission spectroscopy (ICP-OES). Element standards will be made using the same
electrolyte and electrolyte concentration as is used in the experiment that generated the sample to avoid
matrix effects on the analysis. The machine will be calibrated repeatedly during analysis, standards of known
element concentration will be analyzed periodically during the analysis, and each sample and standard will be
analyzed in triplicate in order to quantify analytical uncertainty precisely. Some samples will have low
dissolved Si concentrations, and therefore Si in these samples will be analyzed using a standard molybdenum-
blue approach that increases the sensitivity of the analysis by approximately 2 orders of magnitude. Trace
element (non-major mineral-forming cation) concentrations for at least the final sample from each solubility
experiment will be conducted using an inductively-coupled plasma mass spectrometry (ICP-MS) approach,
using methods similar to those we use for the ICP-OES analyses to quantify analytical uncertainty.

Some of the proposed solubility experiments will be conducted using pure end-member uranyl mineral
phases, however we will also determine the effect of Np substitution in these uranyl phase structures on the
solubility of selected phases. Each uranyl phase of interest will be synthesized with variable concentrations of
Np, following similar procedures used by Burns et al. (2004). The extent of Np incorporation will be
determined by whole -phase acid dissolution of each phase of interest, and Np analysis by inductively-coupled
plasma mass spectrometry (ICP-MS). The Np concentration of aqueous samples from the solubility
experiments will also be analyzed by ICP-MS. The concentration of other cations in equilibrium with each
uranyl phase will be determined by ICP-OES, as described above.

The experimental systems will be controlled so that the approach to equilibrium is either from under-
saturation or from over-saturation, thereby providing necessary and rigorous constraints on the equilibrium
conditions. Under-saturated conditions are attained simply by placing the uranyl mineral powder of interest in
the electrolyte of interest. The initial concentrations of the mineral-forming cations in the aqueous phase will
be negligible under these conditions, so this experimental approach will create a strong Gibbs free energy
driving force for mineral dissolution. Over-saturated solution conditions will be obtained in separate
experiments by first equilibrating the batch systems at pH 1.5 and then adjusting the pH up to the final desired
value between 2 and 5. Because the solubilities of each of the uranyl mineral phases of interest increase with
decreasing pH, the initial pH 1.5 equilibration step will cause aqueous activities of the mineral-forming cations
to reach levels in excess of the activities for the cations at equilibrium under higher pH conditions. Therefore,
the final equilibration will occur from over-saturated solutions and the uranyl mineral phase will precipitate
from solution.

Whether the experimental system approaches equilibrium from under- or over-saturation, it is crucial to
check the solid run-products of each experiment to insure that new minerals did not form during the course of
the experiment. Furthermore, it is vital to constrain the stoichiometry of the mineral phase used in each
experiment so that the dissolution reaction can be properly formulated. We will characterize the mineral
phases both before and after the experiments. All specimens will be examined using powder X-ray
diffraction, which permits identification of the major phases, and detection of impurity phases present at ∼5
wt.%. Specimens will also be examined with a FTIR spectrometer mounted on a polarized-light microscope
using an attenuated total reflection (ATR) objective. The ATR permits collection of IR spectra for sample
10
areas as small as 12 µm, and will be used to systematically search for impurity phases. In each case, multiple
IR spectra will be collected for 12 µm regions of the specimen, which will permit detection of impurities
distributed heterogeneously in the specimen that may not have been detected by XRD.

3.2 Enthalpy of Formation Measurements - High temperature oxide melt solution calorimetry uses a
Calvet-type high temperature custom-built calorimeter (see Fig. 2), details of which are given elsewhere
(Navrotsky 1977, 1997). Solution enthalpies for each sample will be measured by dropping ~ 5 mg pressed
pellets of material from room temperature, 298 K, into the molten oxide solvent, 3Na2O-4MoO 3, at
calorimetric temperature, 976 K. Earlier experiments for UO3 confirmed that U6+ is the stable oxidation state
of U dissolved at low concentrations in 3Na2O·4MoO 3 at 976 K (Helean et al. 2003). The calorimeter will be
calibrated using the heat content of a-Al2O3. It will be flushed continuously with O2 throughout the
experiments to ensure an oxidizing atmosphere.
High temperature oxide melt
drop-solution calorimetry
silica glass
Figure 2. Schematic diagram of Calvet calorimeter
drop tube
platinum
and sample assembly for high temperature oxide
bubbling tube melt solution calorimetry dissolution of uranyl
25°C silica glass
minerals, likewise, involves no oxidation-reduction.
liner

platinum
Samples dissolved readily and quickly. H2O and
crucible CO2 were evolved into the gas above the solvent
702°C
solvent and swept out of the calorimeter by the flowing O2
(3NaO ·4MoO)
2 3
gas, as demonstrated previously (Navrotsky et al.,
1994).

Twinned Thermopiles

Prior to the calorimetric experiments, complete dissolution in the solvent at calorimetric conditions will
be established in a furnace at 976 K. A la rge and rapidly generated endothermic enthalpy of drop solution for
each phase, return of the calorimetric signal to its baseline value, and a solvent color change from white to
yellow will indicate that uranyl minerals dissolve fully in the melt. The evolution of water occurs quickly but
not too vigorously. The high density of the sample assures that the pellets drop quickly and sink in the solvent,
so that the sample does not float on the surface of the melt. Calorimetry is straightforward and problem free
(Hughes Kubatko et al. 2003, 2004a,b).

In the proposed work, the methodology will be essentially the same. The well characterized samples
used for solubility experiments will also be used for calorimetry. However, solution calorimetry is not sensitive
to trace impurities, such as Np doping on the 10-1000 ppm level. Furthermore, the Thermochemistry Facility
is not presently licensed to do Np work. Thus, we will not initially work with Np-bearing samples. However,
after calorimetry for the primary phases studied is completed, we will work with minor dopants at the 1-10%
level of substitution.

3.3 Mineral Synthesis and Preparation - Natural uranyl minerals are difficult to study because they often
occur as fine-grained intergrowths of multiple species, and because some species dehydrate or otherwise
alter under atmospheric conditions. Synthetic preparation of pure and well-characterized uranyl minerals
must therefore be the foundation for experimental determination of their thermodynamic parameters. A
variety of approaches are used depending on the species of interest. In almost all cases, synthesis is
achieved using an aqueous medium containing dissolved U6+ created by dissolving uranyl acetate, uranyl
11
nitrate, or UO3. Many uranyl minerals can be grown under mild hydrothermal conditions in sealed Teflon-
lined containers, by combining the appropriate reactants, adjusting the pH of the solution, and heating the
reactants for a specific time (e.g., Burns et al., 2003; 2004). This approach often results in a well-crystallized
product, and optimisation of pH, reactant concentrations, temperature, and the time of the reaction usually
results in a pure phase in high yield.

Various synthesis techniques may be used to acquire some uranyl minerals at room temperature. The
simplest approach that works in a few cases is batch precipitation from a uranium-bearing solution. In this
approach, the pH of the solution is adjusted to promote precipitation of the target phase. Typically, poorly-
crystallized materia l results, and recrystallization by mild hydrothermal treatment is often advantageous.

It is sometimes possible to grow crystals of uranyl minerals, such as uranyl carbonates, simply by
evaporating solutions containing appropriate nutrients, often in a controlled atmosphere (such as a CO2-rich
atmosphere) (e.g., Hughes Kubatko et al., 2003). In the case of minerals with very low solubility, this
approach gives fine grained, poorly crystalline precipitates. In these cases, diffusion of nutrients into a
chemically inert gel often results in excellent crystals. Typically, nutrients are introduced to a gel located at
the bottom of a U-tube, different nutrients diffuse from either side, and crystals nucleate and grow where the
nutrients mingle in the gel. This approach works especially well for uranyl phosphates and uranyl arsenates
(e.g., Locock et al., 2004; Locock and Burns, 2003).

Work Products and Schedule


The primary product of the proposed research will be an internally-consistent dataset of the
thermodynamic properties for a wide range of uranyl minerals of repository and general environmental
interest. Experimental work will be conducted according to the OCRWM Quality Assurance Requirements
and Description. Costs associated with compliance with quality assurance requirements are included, but not
itemized in the following proposed budget.

The solubility and heat of formation experiments will run concurrently throughout the course of the
project. For both types of experiments, we will begin by investigating relatively simple mineral compositions
first, adding complexities as we proceed and concentrating on Ca, Na, and K as the dominant mineral-forming
cations other than U and Si. We will measure solubilities and heats of formation for the uranyl oxide hydrates
first (schoepite [(UO2)8O2(OH)12](H2O)12, metaschoepite UO3(H2O)2, becquerelite
Ca[(UO2)3O2(OH)3]2(H2O)8, compreignacite K2[(UO2)3O2(OH)3]2(H2O)7), and using these results as a
foundation to interpret results from experiments examining simple uranyl silicates (such as soddyite
(UO2)2(SiO4)(H2O)2, boltwoodite K[(UO2)(SiO 3OH)](H2O)1.5, weeksite K2(UO2)2Si6O15(H2O)4, uranophane
Ca[(UO2)(SiO 3OH)]2(H2O)5), uranyl carbonates (rutherfordine UO2CO3, andersonite
Na2Ca[(UO2)(CO3)3](H2O)5, liebigite Ca2[(UO2)(CO3)3](H2O)11), and uranyl phosphates (chernakovite,
(H3O)2(UO2)(PO4)2.6H 2O, autunite Ca[(UO2)(PO4)]2(H2O)11), thereby yielding an internally-consistent
thermodynamic dataset. Beginning in Year 2, we will use our initial findings to select a mineral from each
group to use to determine the effects of Np substitution on the mineral solubilities.

12
LITERATURE CITED

Basnakova, G., Finlay, J.A. & Macaskie, L.E. (1998): Nickel accumulation by immobilized biofilm of
Citrobacter sp. Containing cell-bound polycrystalline hydrogen uranyl phosphate. Biotech. Lett. 20, 949-
952.
Buck, E.C., Brown, N.R. & Dietz, N.L. (1996): Contaminant uranium phases and leaching at the Fernald site
in Ohio. Env. Sci. Technol. 30, 81-88.
Burns, P. C., Deely, Kathryn M. & Hayden, L.A. (2003): The crystal chemistry of the zippeite group. Can.
Mineral. 41, 687-706.
Burns, P.C., Deely, K.M. & Skanthakumar, S. (2004): Neptunium incorporation into uranyl compounds that
form as alteration products of spent nuclear fuel: Implications for geologic repository performance.
Radiochimica Acta 92, 151-159.
Burns P.C. (1999): The crystal chemistry of uranium. Rev. Mineral. 38, 23-90.
Casas, I., Bruno, J., Cera, E., Finch, R.J., and Ewing, R.C. (1994) Kinetic and thermodynamic studies of
uranium minerals: Assessment of the long-term evolution of spent nuclear fuel. SKB Technical Report 94-
16. Stockholm, Sweden: Swedish Nuclear Fuel and Waste Management Co.
Chen, F.R., and Ewing, R.C. (2003) Structure-configurational entropy and its effect on the
thermodynamic stability of uranyl phases: With special application for geological disposal of
nuclear waste. Science in China Series D – Earth Sciences 46, 39-49.
Chen, F.R., Ewing, R.C. & Clark, S.B. (1999): The Gibbs free energies and enthalpies of formation of U6+
phases: An empirical method of prediction. Amer. Mineral. 84, 650-664.
Fein, J. B. (1994) Porosity enhancement during clastic diagenesis as a result of aqueous metal-carboxylate
complexation: Experimental studies. Chem.Geol. 115, 263-279.
Fein, J. B., and Hestrin, J. E. (1994) Experimental study of oxalate complexing at 80oC: Gibbsite, amorphous
silica, and quartz solubilities in oxalate-bearing fluids. Geochim. Cosmochim. Acta 58, 4817-4829.
Fein, J. B., Yane, L., Jyoti, A., and Handa, T. (1995) Experimental study of Al- and Ca-malonate complexation
at 25, 35, and 80oC. Geochimica et Cosmochimica Acta, 59, 1053-1062.
Finch R. and Murakami T. (1999) Systematics and Paragenesis of Uranium Minerals. In Uranium:
Mineralogy, Geochemistry and the Environment, Vol. 38 (ed. P. C. Burns and R. Finch), pp. 91-180.
Mineralogical Society of America.
Finch R.J. and Ewing, R.C. (1992): The corrosion of uraninite under oxidizing conditions. J. Nucl. Mater.
190, 133-156.
Finch, R.J., Buck, E.C., Finn, P.A. & Bates, J.K. (1999): Oxidative corrosion of spent UO2 fuel in vapor and
dripping groundwater at 90°C. In: Scientific Basis for Nuclear Waste Management XXII (D.J.
Wronkiewicz and J.H. Lee, editors) Materials Research Society Symposium Proceedings 556, 431-438.
Finn, P.A., Hoh, J.C., Wolf, S.F., Slater, S.A. & Bates, J.K. (1996): The release of uranium, plutonium,
cesium, strontium, technetium and iodine from spent fuel under unsaturated conditions. Radiochim. Acta
74, 65-71.
Francis, A.J., Dodge, C.J., Gillow, J.B., & Papenguth, H.W. (2000) Biotransformation of uranium compounds
in high ionic strength brine by a halophilic bacterium under denitrifying conditions. Environ. Sci. Technol.
34, 2311-2317.
Frondel C. (1958) Systematic Mineralogy of Uranium and Thorium. U. S. Geological Survey Bulletin.
Fuller C.C. and Barger, J.R. (2001): Characterization of uranium sorption by apatite in a permeable reactive
barrier. Abstracts of Papers of the American Chemical Society 222, 41-GEOC.
Fuller C.C., Barger, J.R., Davis, J.A., Piana, M.J. (2002): Mechanisms of uranium interactions with
hydroxyapatite: implications for groundwater remediation. Env. Sci. Tech. 36, 158-165.
Giammar D. E. and Hering J. G. (2002) Equilibrium and kinetic aspects of soddyite dissolution and secondary
phase precipitation in aqueous suspension. Geochim. Cosmochim. Acta 66(18), 3235-3245.
Grenthe I., Fuger J., Konings R. J. M., Lemire R. J., Muller A. B., Nguyen-Trung C., and Wanner H. (1992)
Chemical Thermodynamics, Volume 1: Chemical Thermodynamics of Uranium. North-Holland.
14
Helean, K.B, Navrotsky, A., Vance, E.R. Carter, M.L. Ebbinghaus, B. Krikorian, O. Lian, J. Wang, L.M.
and Catalano, J.G. (2002) Enthalpies of formation of Ce-pyrochlore, Ca0.93Ce1.00Ti2.035O7.00, U-pyrochlore,
Ca1.46U4+0.23U6+0.46Ti1.85O7.00 and Gd-pyrochlore, Gd2Ti2O7: three materials relevant to the proposed
waste form for excess weapons plutonium. Journal of Nuclear Materials, 303, 226-239.
Helgeson, H.C., Kirkham, D.H., and Flowers, G.C. (1981) Theoretical prediction of the thermodynamic
behavior of aqueous electrolytes at high pressures and temperatures: IV. Calculation of activity
coefficients, osmotic coefficients, and apparent molal and standard and relative partial molal
properties to 600oC and 5 Kb. Amer. J. Sci. 281, 1249-1516.
Hughes Kubatko, K.A., Helean, K.B., Navrotsky, A., Burns, P.C. (2003): Stability of peroxide-containing
uranyl minerals. Science 302, 1191-1193.
Hughes Kubatko, K-A., Helean, K.B., Navrotsky, A. & Burns, P.C. (2004a): Thermodynamics of uranyl
minerals: Enthalpies of formation of rutherfordine, UO2CO3, andersonite, Na2CaUO2(CO3)3(H2O)5, and
grimselite, K3NaUO2(CO3)3H2O. American Mineralogist (submitted).
Hughes Kubatko, K-A., Helean, K.B., Navrotsky, A. Burns, P.C. (2004b): Thermodynamics of
uranyl minerals: Enthalpies of formation of uranyl oxide hydrates. Amer. Mineral. (submitted).
Langmuir D. (1978): Uranium solution-mineral equilibria at low temperatures with applications to sedimentary
ore deposits. Geochim. Cosmochim. Acta. 42, 547-569.
Locock, A.L. & Burns, P.C. (2003): The crystal structure of synthetic autunite, Ca[(UO2)(PO4)]2(H2O)11.
American Mineralogist 88, 240-244.
Locock, A.J., Burns, P.C., Duke, M.J.M. & Flynn, T.M. (2004): Monovalent cations in structures of the
meta-autunite group. Canadian Mineralogist 42, 973-996.
Macaskie, L.E., Bonthrone, K.M., Yong, P. & Goddard, D.T. (2000): Enzymically mediated bioprecipitation
of uranium by a Citrobacter sp.: a concerted role for exocellular lipopolysaccharide and associated
phosphatase in biomineral formation. Microbiology-UK 146, 1855-1867.
Marshall W.L., Warakomski J.M. (1980) Amorphous silica solubilities – II. Effect of aqueous salt solutions at
25oC. Geochim. Cosmochim. Acta. 44, 915-924.
Murakami T., Ohnuki T., Isobe H., and Sato T. (1997) Mobility of uranium during weathering. American
Mineralogist 82, 888-899.
Navrotsky, A. (1977) Recent progress and new directions in high T calorimetry. Physics and Chemistry of
Minerals, 2, 89-104.
Navrotsky, A. (1997) Progress and new directions in high T calorimetry revisited. Phys. Chem. Minerals,
24, 222-241.
Navrotsky, A., Rapp, R.P., Smelik, E., Burnley, P., Circone, S., Chat, L., Bose, K., Westrich, H.R. (1994)
The behavior of H2O and CO2 in high temperature lead borate solution calorimetry of volatile -bearing
phases. American Mineralogist, 79, 1099.
Nguyen, S.N., Silva, R.J., Weed, H.C., and Andrews, J.E., Jr. (1992) Standard Gibbs free energies of
formation at the temperature 303.15K of four uranyl silicates: soddyite, uranophane, sodium boltwoodite,
and sodium weeksite. Journal of Chemical Thermodynamics 24, 259-276.
Pearcy, E.C., Prikryl, J.D., Murphy, W.M., & Leslie, B.W. (1994): Alteration of uraninite from the Nopal I
deposit, Pena Blanca District, Chihuahua, Mexico, compared to degradation of spent nuclear fuel in the
proposed U.S. high-level nuclear waste repository at Yucca Mountain Nevada. Applied Geochem. 9,
713-732.
Roh, Y., Lee, S.R., Choi, S.-K., Elless, M.P., and Lee, S.Y. (2000) Physicochemical and mineralogical
characterization of uranium-contaminated soils. Soil Sediment Contamination 9, 463-486.
Wanner, H. and Forest, I. (1992) Chemical Thermodynamics of Uranium. Amsterdam, North-Holland:
Elsevier Science Publishers B.V.
Wronkiewicz, D.J., Bates, J.K., Wolf, S.F. & Buck, E.C. (1996): Ten-year results from unsaturated drip tests
with UO2 at 90°C: implications for the corrosion of spent nuclear fuel. J. Nucl. Mater. 238, 78-95.

15

S-ar putea să vă placă și