Next Article in Journal
Theoretical Study on the Relationship of Building Thermal Insulation with Indoor Thermal Comfort Based on APMV Index and Energy Consumption of Rural Residential Buildings
Next Article in Special Issue
pH-Responsive Hollow Polymeric Microspheres from Irradiated Cyclic Ether Aqueous Solution
Previous Article in Journal
Efficacy and Toxicity of VarroMed® Used for Controlling Varroa destructor Infestation in Different Seasons and Geographical Areas
Previous Article in Special Issue
Radiation-Assisted Synthesis of Polymer-Based Nanomaterials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Alternative Conceptual Model for the Spent Nuclear Fuel–Water Interaction in Deep Geologic Disposal Conditions

1
Posiva Oy, Olkiluoto, 27160 Eurajoki, Finland
2
Radiation Laboratory, Department of Physics, University of Notre Dame, Notre Dame, IN 46556, USA
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2021, 11(18), 8566; https://doi.org/10.3390/app11188566
Submission received: 30 August 2021 / Revised: 8 September 2021 / Accepted: 10 September 2021 / Published: 15 September 2021
(This article belongs to the Special Issue Applied Radiation Chemistry: Theory, Methods and Applications)

Abstract

:

Featured Application

Long-term safety assessment of spent nuclear fuel deep geologic repositories.

Abstract

For the long-term safety assessment of direct disposal of spent nuclear fuel in deep geologic repositories, knowledge on the radionuclide release rate from the UO2 matrix is essential. This work provides a conceptual model to explain the results of leaching experiments involving used nuclear fuel or simulant materials in confirmed reducing conditions. Key elements of this model are: direct effect of radiation from radiolytic species (including defects and excited states) in the solid and in the first water layers in contact with its surface; and excess H2 may be produced due to processes occurring at the surface of the spent fuel and in confined water volumes, which may also play a role in keeping the spent fuel surface in a reduced state. The implication is that the fractional radionuclide release rate used in most long-term safety assessments (10−7 year−1) is over estimated because it assumes that there is net UO2 oxidation caused by radiolysis, in contrast with the alternative conceptual model presented here. Furthermore, conventional water radiolysis models and radiation chemical yields published in the literature are not directly applicable to a heterogeneous system such as the spent fuel–water interface. Suggestions are provided for future work to develop more reliable models for the long-term safety assessment of spent nuclear fuel disposal.

1. Introduction

The long-term safety of spent nuclear fuel disposal will have a large impact on the future use of nuclear reactors as a major energy source. Deep geologic disposal is currently the leading long-term strategy for the disposal of spent nuclear fuel, either as is (i.e., with no reprocessing, also referred to as “direct disposal”) or as high-level waste (after reprocessing). In the case of direct disposal of nuclear fuel in a deep geologic repository, the fuel assemblies are isolated from the surface environment by a series of engineered barriers embedded in the host rock at a depth of (typically) a few hundred meters [1]. A few disposal sites have been selected to date, for example, the Olkiluoto site in Finland and the Forsmark site in Sweden. In both countries, the KBS-3 method (Figure 1) is implemented to ensure safe disposal. The spent fuel assemblies (consisting of a number of rods containing fuel (UO2) pellets) are encapsulated in a canister that is designed to ensure long-term containment in the conditions expected at the disposal site. According to the KBS-3 disposal method, the canister consists of a cast iron insert (inner canister), providing mechanical stability and a copper shell (outer canister) providing corrosion protection. The canister is surrounded by buffer bentonite, a swelling clay providing favorable hydraulic, chemical, and mechanical conditions for the canister. The underground openings are backfilled with more swelling clay or a mixture of swelling clay and crushed rock in order to seal the man-made connections to the surface. The repository depth (in Finland, the nominal depth is 430 m below sea level) provides isolation from the surface environment and mitigates the likelihood of inadvertent human intrusion due to common activities, such as underground construction and installation of drinking water wells. The hydrogeochemical conditions at repository depth at the selected disposal sites are typically reducing, supporting the longevity of the disposal canisters.
However, groundwater will eventually penetrate the disposal canisters and come in contact with the spent fuel. Nuclear fuel used in most commercial nuclear power plants consists of uranium dioxide (UO2) enriched with U-235, which is the fissile material. Most radionuclides (e.g., actinides and fission products) produced during the irradiation of nuclear fuel in the reactor are trapped in the UO2 crystalline matrix. The UO2 crystalline matrix is a fluorite-type structure with small distortions and a slight excess stoichiometry (UO2+x) due to the effect of irradiation and the formation of fission and activation products. Leaching tests of spent nuclear fuel samples (or materials that are designed to simulate spent nuclear fuel) carried out over several decades show that one of the limiting factors in the release of radionuclides into the environment is the intrinsically low dissolution rate of UO2 [1,2]. Such leaching tests also showed that the release of uranium and other elements seems to be inhibited in the presence of a reducing agent in solution (typically H2 or metallic iron) conditions. The mechanism for such an inhibiting effect is still being debated.
Spent nuclear fuel is an intrinsic source of ionizing radiation (alpha, beta, and gamma) and neutrons, and the intensity of the radiation field depends on the irradiation history and the age of the spent fuel. Radiolysis plays an important role in the UO2 alteration mechanism because the energy deposited within the UO2 matrix in combination with reactive species produced by radiolysis in the water layer at or near the interface will both lead to the formation of reactive species governing the redox conditions at the UO2 interface. Understanding oxidation and reduction processes at the spent fuel–water interface will contribute to improving the conceptual model of radionuclide releases and achieve a more realistic assessment of the environmental effects of spent nuclear fuel disposal.

2. Spent Nuclear Fuel in Disposal Conditions

Alpha, beta, and gamma radiation are classified as ionizing radiation, which leads to the cleavage of chemical bonds in the traversed medium in a process called “radiolysis”. Neutrons are also emitted within the spent fuel matrix (by fissile radionuclides), but their flux is small and they will interact with water mostly by formation of proton recoils. Radiation (ionizing and not) deposits energy within the spent fuel and any surrounding media such as water. The question as to whether spent fuel self-irradiation affects the dissolution rates has been pondered for decades. It has been found that the activity of the fuel does have an effect on dissolution rates that depends on the fuel age (inversely proportional to its activity) and on the environmental conditions. These dependencies are more fully discussed below. After about 1000 years, the radioactivity of spent fuel will be dominated by alpha-emitting radionuclides. Given that spent fuel is not expected to be in contact with groundwater for at least several thousands of years [3], even in the presence of an initially defective canister [4], the relevant type of ionizing radiation to be considered for its effect on fuel dissolution is alpha radiation. The effects of alpha radiolysis are also relevant to fuel dissolution processes because of the relatively short range (up to 30 micrometers) from the fuel surface in which the energy is deposited, as discussed below.
When discussing the interaction between ionizing radiation and spent fuel, the heterogeneous nature of the system cannot be neglected (see Figure 2). The system is composed of the spent fuel matrix, the solid–water interface, and (bulk) water. The cladding surrounding the UO2 pellets is also in the immediate surroundings but its role in the spent fuel–water interaction will not be discussed in this paper.

3. Observations from Leaching Tests in Reducing Conditions

Table 1 shows a compilation of estimated fractional release rates based on experimentally measured leaching rates obtained in reducing conditions using either alpha-doped UO2 samples (i.e., UO2 samples containing different fractions of short-lived alpha emitters, usually U-233 or Pu-238) or spent fuel samples. The references refer to leaching tests of alpha-doped UO2 samples in reducing conditions, spent fuel, or irradiated MOX samples in reducing conditions, and alpha-doped UO2 samples in Olkiluoto natural or simulated groundwater in the presence of a metallic iron strip. The measured U concentrations in solutions are compared with the solubility limit of the amorphous (more soluble) UO2 form (2 × 10−9 ≤ U < 5 × 10−8 M). Table 1 compiles observations by the authors citing negative, fluctuating between positive and negative, or non-measurable release rates in reducing conditions (in the presence of H2 or metallic iron or Fe(II) in solution), even for highly active spent fuel samples and highly alpha-doped UO2 samples. These observations point to an extremely low dissolution rate, if any dissolution at all, in agreement with the conclusion by Shoesmith [2] “it can be said that based on the above studies there is a strong probability that the corrosion of spent fuel can be avoided either by long term containment in a sealed container or by the reducing influence of H2 in failed container. […] Commonly no rate can be measured when H2 is present.”
From the broader set of results reported in the literature (in the presence and absence of H2 as reducing species), we also know that radiation seems to play an important role in determining whether H2 is effective in suppressing UO2 dissolution. In the absence of radiation, no differences in the oxidation state of the UO2(s) surface under Ar or H2 were observed, whereas a reduction was reported in the presence H2 when external sources of gamma radiation [6] and alpha radiation [7] were used.
Furthermore, in the absence of radiation (e.g., UO2 leaching tests in H2 saturated solutions), oxygen contamination significantly influences the concentrations of U in solution. In the case of spent fuel and its intrinsic radiation, it appears that the system is not as sensitive to the presence of small amounts of oxygen from air contamination and the concentrations of U in solution, and radionuclide releases (e.g., Cs-137) from spent fuel are not as dramatically affected as may be expected if U(IV) had been oxidized to U(VI) [8,9,10].
Therefore, despite existing reports in the literature of oxidative dissolution in both gamma-irradiated UO2 and UO2 doped with alpha emitters, Table 1 shows that there is considerable evidence of the opposite if the conditions are truly reducing, meaning if H2, iron, or other reducing species are dissolved in the leaching solution.

4. Current Conceptual Models to Explain the Results in Reducing Conditions

Several authors have proposed explanations for the nil or negative (or alternating between positive and negative) dissolution rates observed in reducing conditions, particularly in the presence of dissolved H2 used as a reducing agent [23,24,25,26,27,28,29].
Arguably the most cited conceptual model for the UO2 alteration is the Matrix Alteration Model (MAM), shown in Figure 3, which was initially formulated in the framework of the EU Spent Fuel Stability (SFS) project [30,31]. The MAM conceptual model (and similar ones) is based on the idea of “oxidative dissolution” of spent fuel.
As shown in Figure 3, water radiolysis (due to alpha, beta, or gamma radiation) produces both reductants and oxidants upon interaction with water. The MAM model always invokes oxidation of U(IV) to U(VI) as the first step of the water-spent fuel interaction.
H2O2 is often claimed to be the main radiolytic species involved in UO2 oxidation on the basis of either its higher concentration (compared with more reactive radical species such as the hydrated electron and the •OH radical) or faster kinetics compared with other molecular radiolytic species (e.g., H2). Thus, these models imply that the conditions at the spent fuel surface are oxidizing due to the presence of H2O2.
Furthermore, this conceptual model explains the H2 effect mostly by invoking the activation of dissolved H2 on the surface of UO2 without acknowledging the role of the primary radiolytic products formed in the spent fuel itself and at the water–spent fuel interface. According to the MAM model, the H2 dissolved in bulk water is activated on the surface of the spent nuclear fuel thereby reducing the UO3 in solid form. Aqueous ligands (such as bicarbonate, HCO3) are able to form complexes with the oxidized U(VI) leading to the dissolution of the UO2 matrix. Secondary solid phases (e.g., schoepite and studtite) can also precipitate on the surface of spent nuclear fuel, if the solubility limit is reached.
Of course, there is always a component of dissolution that is driven by the solubility of the solid (this process is called non-oxidative dissolution in Figure 3). The non-oxidative dissolution is part of the MAM model [31]; it involves the formation of an amorphous UO2 layer leading to the release of U(IV)aq according to the solubility of UO2 (solid phase). The amorphous form of UO2 is more soluble than the crystalline form so the solubility limit of UO2(amorphous) is typically used to compare with the concentrations of U in solution measured in leaching tests (see Section 3).
The H2 activation mechanism has not been fully explained. What has been clearly demonstrated is that metallic inclusions of Mo, Tc, Pd, Ru, and Rh present in spent fuel (also called epsilon particles) are able to activate H2 and act as catalysts in the reduction of oxidized U(VI). H2 activation on noble metal epsilon particles has been demonstrated electrochemically [24,32] and chemically [26] using SIMFUEL specimens with different levels of simulated burn-up (i.e., different number densities of particles) or added Pd particles, or using radioactive epsilon particles extracted from spent fuel [25]. This work well established that one of the key reactions involved is the catalysis of the reversible dissociation of H2 to •H radicals on the noble metal epsilon particles and this process galvanically protects UO2 matrix from oxidation.
However, this explanation is not sufficient to justify the lack of a measurable dissolution rate observed in reducing conditions and in the absence of epsilon particles, for example, using alpha-doped UO2 samples (see Table 1). Alternative mechanisms have been proposed in the absence of epsilon particles. For example, Lemmens et al. [33] proposed that the UO2 surface acts as a catalyst in the reduction of oxidized U(VI) back to UO2 or in the recombination of radiolysis products back to water.
The proposed explanations above typically invoke an oxidizing first step due to oxidizing radiolytic species, of which H2O2 is claimed to be the most important, although recently the role of •OH radicals has also been mentioned [9,34,35,36]. The role of the intrinsic ionizing radiation from spent fuel and the effects on the solid and at the spent fuel–water interface are not included in the current conceptual model.
Therefore, although there is consensus on the fact that radiolytic species are involved in the processes occurring at the UO2 surfaces, it is still to be demonstrated whether these species cause oxidation of U(IV) to U(VI) (and possibly involving the intermediate species U(V)) or whether the net balance of oxidizing and reducing radiolytic species in reducing conditions maintains the spent fuel surface in a reduced state so that no oxidizing dissolution occurs.

5. Proposed Alternative Conceptual Model

We propose an alternative conceptual model for the observations from leaching tests based on the processes occurring at the early stages of the interaction between ionizing radiation and matter (either spent fuel or water).

5.1. Processes Occurring at the Time of Interaction between Ionizing Radiation and Matter

It is well known that the mechanism for water radiolysis proceeds in stages [37,38], as illustrated in Figure 4 by Buxton and Swiatla-Wojcik [39,40]. The first stage is the physical stage, during which time there is an energy transfer between the ionizing radiation and the medium. In the case of pure water, an excited water molecule (H2O*) and an electron-water cation (H2O+) pair are formed during the physical stage. The physical stage lasts approximately 10−15 s.
The second stage is called the physico-chemical stage (lasting approximately from 10−15 s to 10−12 s) during which the excited water fragments into •H and •OH radicals, in addition to excited O atoms (in triplet or singlet state) and H2. The electrons solvate, and the transfer of a proton from the water cation to a neighboring water molecule yields the •OH radical. In summary, this stage leads to formation of the primary radiolytic species (•H, •OH, H2, and hydrated electrons (eaq).
During the chemical stage (lasting from 10−12 s to approximately 10−6 s), the precursor species either escape the non-homogeneous volume where the energy deposition took place or interact with other primary species to form molecular products such as H2, O2, and H2O2, in addition to the reformation of H2O. Depending on the radiation type, dose and dose rate, and solvents in water, the net balance of these reactions may lead to the recombination of an H2O molecule with no net radiolysis effect.
Finally, the homogeneous stage (from 10−6 s onward) is the stage in which radiolysis products become uniformly distributed in water or aqueous solutions and interact with the environment. In the case of spent nuclear fuel, the sources of radiation are the radionuclides embedded in the solid (UO2) and many of the precursors and primary species may form in the UO2 matrix itself, at its surface or in the environment immediately adjacent to the fuel surface. The energy distribution between the solid and the liquid is also important to determine the nature and fate of such radiolytic species. Alpha radiation in water, for example, will deposit most its energy within 13 microns from the surface and will be completely stopped within 35 microns from the surface [19]. The stopping distance of beta radiation in water ranges from a few mm to tens of mm, depending on the energy. Gamma radiation will travel the farthest from the fuel surface and even penetrate through the cladding surrounding the UO2 pellets and the inner metallic structures of the canister.
The yield of the various radiolytic products depends on the radiation type and energy, and varies as function of time [41]. The primary yield of a particular radiation product, commonly referred to as the G value, is the amount of radiolysis product produced (or destroyed) per unit of absorbed energy. The G value also depends on the linear energy transfer (LET) of the ionizing particle, which is the amount of energy that the particle transfers to the material traversed per unit distance or stopping power (–dE dx−1). For gamma radiation (low LET), the energy is absorbed in a large volume of water, whereas the short range of alpha particles leads to higher LET. The radiolytic yields (G values) obtained from traditional radiolysis studies in homogeneous systems typically reflect the quantity of species that escape from the spurs and tracks, and are able to interact with solutes during the homogeneous stage.

5.2. Processes Occurring in the Bulk of UO2 (Direct Effect of Ionizing Radiation)

Ionizing radiation is known to create electronic defects (such as “holes” formed by the release of an electron from a valence band), or excited states such as polarons or excitons (bound excited states of electrons and holes), and phonons (vibrating atoms in a crystalline structure). Additionally, ionizing radiation can produce Frenkel-type defects created by the displacement of an oxygen atom in the UO2 lattice to form an interstitial-vacancy pair. They are especially common with heavy ions (such as alpha particles) because of the large relative momentum transfer possible in nuclear collisions. The local density of Frenkel defects in a heavy ion track is much greater than that found in gamma radiolysis and can lead to considerable self-trapping of excitons, the potential precursor to H2 [42,43]. A local increase in the temperature occurs along the high-energy ion path. This temperature increase likely induces an enhanced defect migration leading to defect rearrangement [44].
Defect rearrangement causes point defects to coalesce into a lower energy defect structure (dislocation loops), as observed in self-irradiated UO2, or alpha-doped UO2 or PuO2 samples [44,45]. The local density of Frenkel defects in a heavy ion track is much greater than that found in gamma radiolysis and can lead to considerable self-trapping of excitons, which are bound excited states of electrons and holes, the potential precursor to H2.
Furthermore, Mohun et al. [46] studied the response of UO2, PuO2, and ThO2 to an external source of alpha particles (He ions). Mohun et al. [46] propose that there is a coupling between electronic defects and phonons in UO2 and PuO2, whereas alpha irradiation only causes electronic defects in ThO2. The authors suggest that the alpha-induced irradiation defects may be involved in chemical reactions at the UO2–water interface. In doing so, the UO2 boundary layer acts as a sink and governs the annealing of defects.
Tocino et al. [47] and Corkhill et al. [48] also highlighted the importance of atomistic defects during the dissolution of uranium mixed oxides (U-Pu)xOy and cerium oxides (CeO2), respectively. The authors showed that the high leaching rate of the studied metal oxides was related to the reduced chemical durability of the materials due to the presence of defects in the crystalline lattice, such as oxygen vacancies.
Carbol et al. [11] suggest that the neutralization of the radiolytic oxidants produced near the oxide surface by alpha radiolysis may be due to oxygen vacancies created on the surface of UO2(s), which likely contributes to the release of hydrogen atoms (•H) through the dissociative adsorption of a water molecule. The authors could not rule out that this or other surface processes contribute to the orders of magnitude lower homogeneous radiolytic yields by alpha particles emitted by UO2(s) surfaces, compared to the yields measured with an external source of alpha radiation.
These findings indicate that species formed due to self-irradiation of UO2 (oxygen vacancies, electronic defects, or excited species) may play a significant role in the chemical reactions occurring at or near the fuel matrix–water interface.

5.3. Processes Occurring at the Metal Oxides-Water Interface

The radiolysis of water adsorbed at the surface of metal oxides, including UO2, shows high (~10-fold) radiolytic yields of molecular H2 in adsorbed water compared to that formed by radiolysis of pure water, whereas the yields of O2 were at least an order of magnitude lower for water adsorbed on UO2 or ZrO2 surfaces [42,43,49,50]. The possible mechanism for excess H2 production in the radiolysis of adsorbed water proposed by LaVerne and Tandon [42] (and references therein) involve dissociative electron attachment of low-energy electrons, recombination of electron-hole pairs, and exciton reactions. The negligible production of O2 in these systems may be explained by the decomposition of water bound at the interface, resulting in oxygen species attached to or near the oxide surface.
In 1971, Boehm observed that water associated with oxide surfaces is usually dissociated to form OH groups in the form of hydroxylated metal oxide surfaces (≡M-OH) [51]. Additional water will be hydrogen-bound to this first chemisorbed layer in the form of physisorbed water layers. The fate of O2 in these systems is unclear. Pulse radiolysis studies suggest that only the reducing species leave the surface to the bulk water [52]. XPS studies of irradiated materials report non-stoichiometric oxygen within the bulk material [53]. The oxygen may be filling up vacancies, but it is most likely breaking up the crystal structure, a process well known in radiation-induced corrosion studies [54].
As mentioned above, self-irradiation of UO2 can cause the formation of oxygen vacancies (Ov). Oxygen vacancies (Ov) are also known to play an important role in the surface chemistry of metal oxides such as CeO2 and AmO2 (which are fluorite-type structures often used as UO2 analogues) and TiO2 [55,56]. Hydroxylated species (≡M-OH) and hydride species (≡M-H) bonded to the surface can act as oxidizing or reducing species [55,56]. It is also known that oxygen vacancies have the tendency to spontaneously ionize, and thus automatically liberate electron polarons in the lattice and reduce the surrounding metal atoms [57].
First-principles studies of the reaction of water on the crystallographic surfaces of UO2 and PuO2 (and ThO2) help in explaining the excess formation of electrons and H2. Bo et al. [58] propose that H2 is formed from water dissociation. Water dissociation may undergo two pathways in the presence of surface oxygen vacancy on the reduced UO2(111) surface. One path is characterized by direct combination of two hydrogen atoms of one water molecule, and the other is characterized by dissociation of the first hydrogen atom and its combination with a neighboring surface hydrogen atom.
Wang et al. (2019) [59] propose that excess electrons are formed from radiation-induced oxygen vacancies. The excess electrons lead to the exothermic splitting of H2O and formation of molecular H2 on UO2 surfaces. This effect depends on the nature of the oxide; for example, for PuO2 the formation of H2 from interfacial water is thermodynamically unfavorable.
Calculations have reported that energies associated with adsorbed and dissociated water are highly dependent on the specific UO2 surface [58,60]. Water adsorbed on a pristine stoichiometric UO2 surface is experimentally found to be a reversible process at room temperature [61]. The evolution of H2 is not observed in temperature programmed desorption studies of these surfaces, indicating that the energy binding the water to the UO2 surface is relatively low [62]. However, water is dissociatively adsorbed on UO2 surfaces with U in a lower oxidation state or from highly defected surfaces. In these latter surfaces, temperature-programmed desorption reveals that the production of H2 is thought to be due to an oxygen vacancy at the surface [62].
The different types of ionizing radiation in spent fuel, because of the different stopping powers or LET, also play an important role in determining the energy distribution and resulting species formed within the UO2 matrix, at the water–spent fuel interface and in bulk water. When a gamma ray, a beta particle, or an alpha particle is released by the radionuclides embedded in spent fuel, part of the energy that is absorbed by the solid forms either defects or excited species that can interact with the water at the interface and beyond. The fraction of the ionizing radiation that is not used in direct effects in the solid causes ionizing events in water within a range that depends on its energy on leaving the surface and LET. Typically, gamma and beta radiation are able to cause ionizing events at a far greater distance than alpha radiation due to their lower LET. When modelling the effect of the intrinsic radiation on water radiolysis, the fraction of the dose absorbed by the UO2 itself should be assessed by taking into account the electron density and the solid/water ratio in the system because it is possible that most of the ionizing radiation energy is absorbed by the solid itself.
One of the more common methods for determination of the energy dissipation through a metal–water interface is the measurement of the yield of the main reducing molecular radiolytic species, H2. The H2 yield has been determined for a wide variety of compounds [42,43,50,53,63,64], including UO2 [43]. In quite a few cases, the production of H2 far exceeds that expected from energy deposited in the water layer alone. Later studies of UO2 found that the radiolytic yield of H2 was about 40 molecules 100 eV−1 energy directly absorbed by a few adsorbed water layers, compared to the value of 0.45 molecules/100 eV for bulk water [42]. Clearly, there is a transport of energy or charge from the bulk of UO2 to the interfacial water. The exact precursor to this excess H2 formation is not known but is believed to be due to the transport of excitons [43]. In the next section, we provide a possible mechanism for the production of H2 at the spent fuel–water interface.

5.4. Processes Occurring in Confined Water Volumes

The surface of spent fuel is characterized by micro and macro cracks caused by the heavy thermal transients during its service life in the nuclear reactor, as shown in Figure 2. When the fuel comes into contact with water (in scenarios involving canister failure), water penetrates into cracks and gaps, and even reaches grain boundaries. The same processes involved in the interaction between the ionizing radiation and water occur but the solid surface to water volume ratio is different in the radiolysis of confined water.
When studying the radiolysis of confined water, similar high H2 production yields as in the case of radiolysis of metal oxides (see Section 5.3) were observed by Rotureau et al. [65] and LeCaer et al. [66]. The authors explained these effects as due to the lack of scavenging of H2 by the •OH radicals and efficient energy transfer to the pore.
Experiments were performed by Traboulsi et al. [67] on UO2 in alpha irradiated distilled water to study the effect of accumulation of radiolytic H2 in confined water volumes. In those experiments, radiolytic H2 was allowed to escape from the open system but to accumulate in the closed system. In the closed system, the dissolved U concentration was suppressed to about one-third of that observed in the open system due to the accumulation of radiolytic H2. This result shows that, in confined water volumes, radiolytically generated H2 may be sufficient to maintain the surface in reducing conditions and suppress U oxidation without the need for any H2 contribution.
Modelling work by Liu et al. [27,28] also suggests that accumulation of radiolytic H2 in a closed system will completely suppress UO2 dissolution even without an imposed H2 overpressure.

5.5. Redox Conditions at the Spent Fuel Surface

As observed in some of the experimental studies discussed in Section 3, we know that radiation plays an important role in explaining the role of H2 in suppressing UO2 dissolution. Although there is consensus on the fact that radiolytic species are involved in the processes occurring at the UO2 surfaces, it is still to be demonstrated which species cause oxidation of U(IV) to U(VI) (possibly involving the intermediate species U(V)) and whether there are species capable of reversing the effect of oxidation so that the net balance of oxidizing and reducing species maintains the spent fuel surface in a reduced state, i.e., no oxidizing dissolution occurs in reducing conditions.
We present below two possible mechanisms producing an excess of reducing species (either •H, eaq or H2) that may be responsible for maintaining the spent fuel surface in a reduced state, hence precluding the oxidative dissolution of UO2.The following mechanism is based on the work of Essehli et al. [68] for water adsorbed on TiO2. Radiation first induces the formation of electron (e) –hole (h+) pairs in UO2, as shown in Equation (1):
UO2 (intrinsic gamma, beta and alpha irradiation) → h+ + e
where h+ is an electronic defect vacancy (see Section 5.2). The charge carriers can combine to give an exciton (hole-electron coupled state on the surface)—Equation (2):
h+ + e → exciton
It has been suggested that excitons in hydroxylated metals react with the surface water groups ≡U-OH, as shown in Equation (3). The possible formation of hydroxylated groups (≡U-OH) also on the surface of spent nuclear fuel via the formation of oxygen vacancies is discussed in Section 5.3. Lastly, •H atoms combine to form H2—Equation (4):
exciton + ≡U-OH (+H2O) → ≡U-O• + •H
•H + •H → H2
An additional surface reaction producing H2 can be a hydrogen atom abstraction from a hydroxylated metal surface:
•H + ≡U-OH → ≡U-O• + H2
There is not normally enough adsorbed water to form the hydrated electron, but with nearby bulk water, Equation (3) can take place in the aqueous phase along with the following combination reactions:
eaq + eaq + 2H2O → H2 + 2OH¯
eaq + •H + H2O → H2 + •OH
Another pathway for •H production may be the electron attachment to ≡U–OH groups (Equation (8)) followed by dissociation (Equation (9)):
e + ≡U-OH → ≡U-OH-
≡U-OH →U-O + •H
H2 is then formed via •H radical recombination, Equation (4). These mechanisms imply that the conditions at the spent fuel surface are reducing due to the radiation-induced production of excess H2 or other reducing species at the surface. If this is the case, oxidative dissolution does not take place in reducing conditions even without any external input of H2.
The idea that the conditions at the spent fuel surface may be reducing rather than oxidizing has already been proposed by a few authors. King et al. [6], after studying the gamma radiolysis of UO2 in the presence of H2, concluded that hydrogen not only suppresses oxidation due to radiolytic oxidants but also reduces the extent of surface oxidation observed in either Ar or H2 atmospheres in the absence of radiation. The authors propose that this dual effect of H2 is most likely the result of reactions between reducing species and both oxidants in solution and reactive surface sites, the latter possibly located at grain boundaries. The authors could not identify the nature of the reducing species responsible for these effects on the basis of their data, although they invoke the hydrated electron (eaq) and the •H radical.
Carbol et al. [8] suggests that a combination of alpha-doped UO2, synthetic and anoxic groundwater, and H2 is able to: (a) reduce radiolytic oxidants to water, (b) readily reduce traces of external oxidants entering the system, and (c) stabilize the sample surface as stoichiometric UO2.00. The authors propose that H2 dissolved in water is adsorbed on the UO2 surface and acts as electron donor able to reduce oxidizing species formed by radiolysis, and U(V) is potentially formed as a result. Shoesmith [23] proposes the formation of a U(IV)-U(V) site that can then absorb H2. Thereafter, one •H radical is adsorbed and an electron is donated to the surface. Hence, the surface is slightly oxidized but, due to the relatively large amounts of dissolved H2, a back reaction occurs before the next oxidant brings the U(V) to U(VI). The authors propose that such a process would explain the absence of oxidants and oxidized uranium during the long-term leaching tests mentioned in Section 3, and leads to the overall recombination of oxidizing species in solution back to water. The formation of U(V) and subsequent reduction was recently observed at the surface of a thin UO2 film exposed to plasma in vacuum at 400 °C [34,35]. The authors also mention the formation of a surface-bound •OH radical as an intermediate species involved in the oxidation of the UO2 surface [34,35,69].
Hansson et al. [70] also suggested that the effect of dissolved H2 on preventing surface oxidation by alpha radiolysis is apparently a result of the interaction of ionizing radiation with water adsorbed on the surface of actinide oxides, without proposing an exact mechanism but invoking the need of dedicated interfacial radiolysis studies. Hansson et al. [70] recognized that, in addition to standard radiolytic processes, energy, charge, or matter can be transferred through the interface, and catalytic or steric effects can alter the decomposition or reactivity of adsorbed molecules.
Aqueous phase radiation chemistry can also occur should there be sufficient water near the fuel surface. In that case, homogeneous radiation chemistry will occur with the additional build-up of H2 due to the confined conditions. Under these circumstances, the Allen cycle [71] will lead to the destruction of oxidizing species produced in the aqueous phase:
•OH + H2 → H2O + •H
•H + H2O2 → H2O + •OH
Reactions (10) and (11) will cycle until all the oxidizing species are removed as long as there is excess H2.
Excess H2 can also be formed by reactive species formed during the interaction of metal oxides with water, as discussed above. The role of reactive species on metal oxide surfaces in accelerating the dissociation of water and producing H2 in neutral and alkaline solutions has been also proposed in the framework of different studies, such as the development of catalysts for fuel cells [72,73] or metal corrosion [55,74,75,76]. Betova et al. [74] proposed a model to explain the observed H2 generation in the presence of copper metal and deoxygenated water involving surface sites available for adsorption of hydroxide intermediates acting as catalysts in the splitting of adsorbed water molecules.
According to Li et al. [55,56], in addition to surface hydroxyl species (M≡OH), various reactive species are formed on metal oxide surfaces in contact with water and in the presence of a reducing agent, such as H2, such as hydride species (M≡H) and hydrated protons (M≡H3O+). These M≡OH and M≡H species are formed on the surface and in the bulk, accompanied by oxidation and reduction reactions of the metal [55,56]. According to the authors, oxygen vacancies have a strong influence on both the stability and reactivity of these species. Hydroxyl groups, together with hydride species and hydrated protons, can explain the formation of H2 from the interaction of a metal oxide and water via the formation of H-bonding networks on the surface of the metal oxide, based on previous work [75,76]. The photocatalytic properties of a Pt-TiO2 surface in reducing water to H2 can be explained by such H-bonding networks, which are formed by hydroxyl groups on solid surfaces [55,56]. TiO2 itself exhibits structural hydroxyl groups and limited surface oxygen vacancies where H2O dissociates to form hydroxyl groups. Moreover, the loading of Pt co-catalysts on TiO2 introduces a Pt–TiO2 interface capable of efficiently dissociating H2O to form hydroxyl groups, which greatly facilitates the formation of H-bonding networks. Epsilon particles that can be found in spent fuel contain d-metals (Pd, Mo, Tc, Ru, and Rh), which may play a similar catalytic role as that of Pt on a TiO2 surface in dissociating water with the net production of H2.
The alternative conceptual model presented in this paper is shown in Figure 5. The novelty consists in including the interaction of ionizing radiation with the UO2 matrix involving the formation of both oxidizing and reducing species (e.g., hydroxylated species, hydrides, and hydrated protons), which may explain the depletion of the oxidizing radiolysis products and a possible net production of H2 at the surface. These species may be responsible for maintaining the spent fuel surface in a reduced state even without any external input of H2.
In summary, the main source of radiation over long periods will be alpha radiation and its short range will lead to radiolysis at or near the surface. This region is most critical for dissolution. The arguments apply also for confined water in the presence of an intrinsic source of radiation, such as the water that is trapped in cracks at the spent fuel surface and even water that has penetrated in the grain boundaries.

6. Outlook on Developments of the Conceptual Model for Spent Fuel–Water Interaction

The role of the direct effect of the different types of radiation in spent fuel and the interaction with the UO2 matrix and with the water at the spent fuel interfaces is just starting to be acknowledged. Although there is consensus on the fact that radiolytic species are involved in the processes occurring at UO2 surfaces, it is still to be demonstrated which species cause oxidation of U(IV) to U(VI) (possibly involving the intermediate species U(V)) and whether there are species capable of reversing the effect of oxidation so that no net UO2 oxidation occurs.
Unfortunately, there is no perfect analogue to study the releases of “aged” spent fuel. Spent fuel samples currently available have the right physical and chemical properties but still contain a significant amount of beta and gamma radiation, being only a few decades old. Alpha-doped UO2 samples simulate correctly the level of activity of “aged” spent fuel but do not have the right physical and chemical properties. Chemical analogues, such as CeO2, have the same crystalline (fluorite) structure as UO2 but no radiation. SIMFUEL, which is synthetic UO2 containing d-group noble-metal particles, similar to those formed in spent nuclear fuel, also does not emit radiation. All these materials together provide useful insights on the dissolution rate of UO2, but none individually.
Both Tocino et al. [32] and Mohun et al. [46] showed that metal oxides which have the same crystalline structure can respond differently to radiation (and/or interact differently with water). Therefore, care should also be taken when comparing UO2 crystalline analogues and drawing conclusions on the energy–water interaction mechanism.
Similarly, water radiolysis studies using external sources of ionizing radiation (typically alpha or gamma or accelerated electrons) in the absence of a solid–liquid interface cannot be used to simulate the effect of intrinsic (or self-) irradiation on spent fuel because typically they address the homogeneous stage of water radiolysis and not the interaction between the species formed due to the direct effect of radiation in the solid and those formed at the solid–water interface. For example, Pastina and LaVerne [77,78] investigated the decomposition of an initial H2O2 solution (5 × 10−5 mol L−1) in bulk water under external sources of pure gamma and alpha radiation. In the case of an external source of gamma irradiation, complete H2O2 consumption was observed, confirming the effect of radiolytic species and dissolved H2 in reducing oxidizing species such as H2O2. However, under an external source of alpha radiation, even if the solution was saturated with H2 (0.8 mM), the initial H2O2 could not be reduced. In this case, the explanation is that the escape yields of primary radiolytic products from the tracks formed by an external alpha source are too low to interact with solutes in bulk water, such as H2 and H2O2. In order to interact with the radical species in the alpha particle tracks, much higher concentrations of radical scavengers would be needed, which was not the case in this particular experiment by Pastina and LaVerne [78]. Alternatively, during bulk water radiolysis using a mixed source of gamma and alpha radiation at 25 °C (where the alpha particles were produced in situ by neutron capture in a solution containing B10), even modest levels of dissolved H2 (0.8 mM) can suppress the formation of molecular radiolytic oxidants [79].
The different responses for the alpha radiolysis are mainly due to the energy distribution of the alpha particles. In external radiolysis, much of the energy loss occurs at the higher energies and lower LET, so radicals can escape the track and interact with oxidizing species. By comparison, in situ radiolysis results in a substantial portion of the alpha particles being attenuated in energy before reaching the surface. The shift in average LET to lower values reduces the escape yields of radical species. Thus, radiolysis studies in bulk water can only provide limited information on the processes occurring at the spent fuel surface.
It can be argued that even the concept of radiolytic yield does not apply to interfacial water radiolysis. Radiolytic yields (G values) obtained from traditional radiation chemistry experiments cannot be applied in the model of spent fuel dissolution because they apply to the homogeneous phase of water radiolysis, whereas, in this case, the yields of radiolytic species formed during the physical and physical-chemistry stage are of relevance.
In summary, the spent fuel–water interface may play an active role in the radiolysis processes, and the primary yields of radiolytic products may be very different from the G values obtained from traditional radiolysis studies in homogeneous systems. Hence care should be used when applying such G values to model spent fuel dissolution as a function of the radiation dose.
To improve the understanding of the interaction between the spent fuel surface and water, the effect of the intrinsic ionizing radiation on the UO2 crystalline matrix, the confined water volumes present at the spent fuel surface, and the layers of water immediately adjacent to the spent fuel interface should be further investigated. The formation of both oxidizing and reducing species at the surface of UO2 and at the UO2–water interface leads to a large diversity of reaction pathways to be taken into account for better understanding of reactivity of spent fuel in reducing conditions.
The following are examples for potential future studies that could shed further light on the processes occurring at the spent nuclear fuel–water interface:
  • Comparison of the effects of alpha particles produced in situ (for example, by using neutron capture on soluble compounds containing B-10) to those of external alpha sources;
  • Further exploration of the field of radiolysis in confined water volumes and of water at the surface of metal oxides, in particular, fluorite-type solids;
  • Investigation of whether excess H2 can be produced at the spent nuclear fuel–water interface;
  • Modelling using Monte Carlo based techniques exploring physico-chemical processes at the spent fuel–water interface.

7. Conclusions

This work provides an alternative conceptual model to explain the leaching test results involving actual spent nuclear fuel or simulant materials in (confirmed) reducing conditions, which are relevant for deep geologic disposal.
The key findings are the following: the direct effect of radiation forms radiolytic species (including defects and excited states) in the solid and in the first water layers in contact with its surface; excess H2 may be produced due to processes occurring at the surface of spent fuel and in confined water volumes involving such species; such species may also play a role in keeping the spent fuel surface in a reduced state.
Conventional water radiolysis models cannot be used directly to estimate the radionuclide release rates because models and G values are designed for bulk (homogeneous) chemistry in the absence of a solid–liquid interface. The implication of the alternative conceptual model proposed in this work is that the long-term fractional dissolution rate (10−7 y−1) used in most long-term safety assessments is overly estimated because this rate assumes that there is a net UO2 oxidation rate caused by radiolysis, in contrast with the alternative conceptual model presented here. (Note that the source term for radionuclide release and transport models used in the long-term safety assessment typically involves a long-term fractional dissolution rate for radionuclides embedded in UO2 and an “instantly” released fraction of the radionuclide inventory to take into account the fraction of radionuclides that are typically released shortly after the spent fuel comes in contact with water. Instantly released radionuclides, such as Cs-137, Cl-36, C-14, and I-129, can be found in locations that are easily accessible to water, such as fractures, gaps, or grain boundaries.)
Because of uncertainties in experiments, in addition to the evolution of the surface area of the fuel in contact with water and long-term self-irradiation effects, adopting a pessimistic long-term fractional dissolution rate and maintaining it at a constant value throughout the assessment period (typically 1 M years) is a cautious approach, at present. This approach is currently used in the source term for long-term safety assessments by spent fuel management organizations, such as Posiva Oy [80], SKB [81], Nagra [82], and NWMO [83].
This paper also provides suggestions for future work that will yield more information on the interfacial water radiolysis occurring at the spent fuel surface. If it can be proven that the conditions at the spent fuel surface are indeed reducing due to the effect of radiolytic species formed in the spent fuel itself or in the close proximity of the surface, then it will be possible to develop more reliable models for the long-term safety assessment of nuclear fuel disposal.

Author Contributions

Conceptualization; writing—original draft preparation, writing—review and editing: B.P. and J.A.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

Posiva Oy’s contribution to this work (B.P.’s time and publication fees) is acknowledged. JAL acknowledges the Notre Dame Radiation Laboratory, which is supported by DOE BES Grant Number DE-FC02-04ER15533. This contribution is NDRL-5332 from the Notre Dame Radiation Laboratory.

Conflicts of Interest

The authors declare that they have no known competing financial or personal relationships that could have appeared to influence the work reported in this paper. B.P. is a staff member of a spent fuel management organization (Posiva Oy) which may be perceived as a non-financial potential conflict of interest.

References

  1. Johnson, L.H.; Shoesmith, D.W. Spent Fuel. In Radioactive Waste Forms for the Future; Lutze, W., Ewing, R.C., Eds.; Elsevier: Amsterdam, The Netherlands, 1988; p. 635. [Google Scholar]
  2. Shoesmith, D.W. Fuel corrosion processes under waste disposal conditions. J. Nucl. Mater. 2000, 282, 1–31. [Google Scholar] [CrossRef]
  3. Posiva. Safety Case for the Disposal of Spent Nuclear Fuel at Olkiluoto—Performance Assessment 2012; Posiva Oy: Eurajoki, Finland, 2013; p. 520. [Google Scholar]
  4. Posiva Oy. Safety Case for the Disposal of Spent Nuclear Fuel at Olkiluoto—Formulation of Radionuclide Release Scenarios 2012; Posiva Oy: Eurajoki, Finland, 2013; p. 135. [Google Scholar]
  5. Fors, P.; Carbol, P.; Van Winckel, S.; Spahiu, K. Corrosion of high burn-up structured UO2 fuel in presence of dissolved H2. J. Nucl. Mater. 2009, 394, 1–8. [Google Scholar] [CrossRef]
  6. King, F.; Kolar, M.; Shoesmith, D.W. Modelling the oxidative dissolution of UO2. In Scientific Basis for Nuclear Waste Management Xxii; Wronkiewicz, D.J., Lee, J.H., Eds.; Materials Research Soc: Warrendale, PA, USA, 1999; Volume 556, pp. 463–470. [Google Scholar]
  7. Sunder, S.; Boyer, G.D.; Miller, N.H. XPS studies of UO2 oxidation by alpha-radiolysis of water at 100 degrees C. J. Nucl. Mater. 1990, 175, 163–169. [Google Scholar] [CrossRef]
  8. Carbol, P.; Fors, P.; Gouder, T.; Spahlu, K. Hydrogen suppresses UO2 corrosion. Geochim. Cosmochim. Acta 2009, 73, 4366–4375. [Google Scholar] [CrossRef]
  9. Ekeroth, E.; Granfors, M.; Schild, D.; Spahiu, K. The effect of temperature and fuel surface area on spent nuclear fuel dissolution kinetics under H2 atmosphere. J. Nucl. Mater. 2020, 531, 13. [Google Scholar] [CrossRef]
  10. Spahiu, K.; Devoy, J.; Cui, D.Q.; Lundstrom, M. The reduction of U(VI) by near field hydrogen in the presence of UO2(S). Radiochim. Acta 2004, 92, 597–601. [Google Scholar] [CrossRef]
  11. Carbol, P.; Cobos-Sabate, J.; Glatz, J.-P.; Ronchi, C.; Rondinella, V.; Wegen, D.H.; Wiss, T.; Loida, A.; Metz, V.; Kienzler, B.; et al. The effect of dissolved hydrogen on the dissolution of 233U doped UO2(s). In SKB Technical Report 05-09; Swedish Nuclear Fuel and Waste Management Co.: Stockholm, Sweden, 2005. [Google Scholar]
  12. Loida, A.; Grambow, B.; Geckeis, H. Anoxic corrosion of various high burnup spent fuel samples. J. Nucl. Mater. 1996, 238, 11–22. [Google Scholar] [CrossRef]
  13. Loida, A.; Grambow, B.; Geckeis, H. Spent fuel corrosion behavior is salt solution in the presence of hydrogen overpressure. In Proceedings of the ICEM’01, 8th International Conference on Radioactive Waste Management, Bruges, Belgium, 30 September –4 October 2001. [Google Scholar]
  14. Loida, A.; Metz, V.; Kienzler, B.; Geckeis, H. Radionuclide release from high burnup spent fuel during corrosion in salt brine in the presence of hydrogen overpressure. J. Nucl. Mater. 2005, 346, 24–31. [Google Scholar] [CrossRef]
  15. Spahiu, K.; Cui, D.Q.; Lundstrom, M. The fate of radiolytic oxidants during spent fuel leaching in the presence of dissolved near field hydrogen. Radiochim. Acta 2004, 92, 625–629. [Google Scholar] [CrossRef]
  16. Ollila, K.; Albinsson, Y.; Oversby, V.; Cowper, M. Dissolution Rates of Unirradiated UO2, UO2 Doped With 233U, and Spent Fuel under Normal Atmospheric Conditions and under Reducing Conditions Using An Isotope Dilution Method; SKB: Stockholm, Sweden, 2003. [Google Scholar]
  17. Spahiu, K.; Eklund, U.B.; Cui, D.Q.; Lundstrom, M. The influence of near field redox conditions on spent fuel leaching. In Scientific Basis for Nuclear Waste Management Xxv; McGrail, B.P., Cragnolino, G.A., Eds.; Materials Research Soc: Warrendale, PA, USA, 2002; Volume 713, pp. 633–638. [Google Scholar]
  18. Spahiu, K.; Werme, L.; Eklund, U.B. The influence of near field hydrogen on actinide solubilities and spent fuel leaching. Radiochim. Acta 2000, 88, 507–511. [Google Scholar] [CrossRef]
  19. Cera, E.; Bruno, J.; Duro, L.; Eriksen, T.E. Experimental Determination and Chemical Modelling of Radiolytic Processes at the Spent Fuel/Water Interface, Long Contact Time Experiments; Swedish Nuclear Fuel and Waste Management Compant Technical Report Stockholm, Sweden TR-06-07; Swedish Nuclear Fuel and Waste Management Co.: Stockholm, Sweden, 2006. [Google Scholar]
  20. Puranen, A.; Roth, O.; Evins, L.Z.; Spahiu, K. Aqueous leaching of high burnup UO2 fuel under hydrogen conditions. MRS Adv. 2018, 3, 1013–1018. [Google Scholar] [CrossRef]
  21. Odorowski, M. Etude de L’altération de la Matrice (U,Pu)2 du Combustible Irradié en Conditions de Stockage Géologique: Approche Expérimentale et Modélisation Géochimique. Ph.D. Thesis, Mines Paris Tech, Paris, France, 2015. [Google Scholar]
  22. Zetterström-Evins, L.; Juhola, P.; Vahanen, M. REDUPP Final Report; Posiva Oy: Eurajoki, Finland, 2014. [Google Scholar]
  23. Shoesmith, D.W. Used Fuel and Uranium Dioxide Dissolution Studies—A Review; Nuclear Waste Management Organization: Toronto, ON, Canada, 2007. [Google Scholar]
  24. Broczkowski, M.E.; Noel, J.J.; Shoesmith, D.W. The inhibiting effects of hydrogen on the corrosion of uranium dioxide under nuclear waste disposal conditions. J. Nucl. Mater. 2005, 346, 16–23. [Google Scholar] [CrossRef]
  25. Cui, D.; Low, J.; Sjostedt, C.J.; Spahiu, K. On Mo-Ru-Tc-Pd-Rh-Te alloy particles extracted from spent fuel and their leaching behavior under Ar and H2 atmospheres. Radiochim. Acta 2004, 92, 551–555. [Google Scholar] [CrossRef]
  26. Jonsson, M.; Nielsen, F.; Roth, O.; Ekeroth, E.; Nilsson, S.; Hossain, M.M. Radiation induced spent nuclear fuel dissolution under deep repository conditions. Environ. Sci. Technol. 2007, 41, 7087–7093. [Google Scholar] [CrossRef] [PubMed]
  27. Liu, N.Z.; Qin, Z.; Noel, J.J.; Shoesmith, D.W. Modelling the radiolytic corrosion of alpha-doped UO2 and spent nuclear fuel. J. Nucl. Mater. 2017, 494, 87–94. [Google Scholar] [CrossRef]
  28. Liu, N.Z.; Wu, L.D.; Qin, Z.; Shoesmith, D.W. Roles of Radiolytic and Externally Generated H2 in the Corrosion of Fractured Spent Nuclear Fuel. Environ. Sci. Technol. 2016, 50, 12348–12355. [Google Scholar] [CrossRef]
  29. Trummer, M.; Roth, O.; Jonsson, M. H2 inhibition of radiation induced dissolution of spent nuclear fuel. J. Nucl. Mater. 2009, 383, 226–230. [Google Scholar] [CrossRef]
  30. Riba, O.; Coene, E.; Silva, O.; Duro, L. Spent fuel alteration model integrating processes of different time-scales. MRS Adv. 2020, 5, 159–166. [Google Scholar] [CrossRef]
  31. Martínez Esparza, A.; Cunado, M.A.; Gago, J.A.; Quinones, J.; Iglesia, E.; Cobos, J.; Gonzalez de la Huebra, A.; Cera, E.; Merino, J.; Bruno, J.; et al. Development of a Matrix Alteration Model (MAM); ENRESA Report 01; Empresa Nacional de Residuos Radiactivos, SA (Enresa): Madrid, Spain, 2005. [Google Scholar]
  32. Broczkowski, M.E.; Keech, P.G.; Noel, J.J.; Shoesmith, D.W. Corrosion of Uranium Dioxide Containing Simulated Fission Products in Dilute Hydrogen Peroxide and Dissolved Hydrogen. J. Electrochem. Soc. 2010, 157, C275–C281. [Google Scholar] [CrossRef]
  33. Lemmens, K.; Cachoir, C.; Mennecart, T. Dissolution Behavior of Spent Fuel at Highly Alkaline Conditions; Studiecentrum voor Kernenergie (SCK.CEN): Mol, Belgium, 2019. [Google Scholar]
  34. El Jamal, G.; Gouder, T.; Eloirdi, R.; Jonsson, M. Time-dependent surface modification of uranium oxides exposed to water plasma. Dalton Trans. 2021, 50, 4796–4804. [Google Scholar] [CrossRef]
  35. El Jamal, G.; Gouder, T.; Eloirdi, R.; Jonsson, M. X-Ray and ultraviolet photoelectron spectroscopy studies of Uranium (IV), (V) and (VI) exposed to H2O-plasma under UHV conditions. Dalton Trans. 2021, 50, 729–738. [Google Scholar] [CrossRef]
  36. Maier, A.C.; Kegler, P.; Klinkenberg, M.; Baena, A.; Finkeldei, S.; Brandt, F.; Jonsson, M. On the change in UO2 redox reactivity as a function of H2O2 exposure. Dalton Trans. 2020, 49, 1241–1248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Mozumder, A. Charges particle tracks and their structure. In Advances in Radiation Chemistry; Burton, M., Magee, J.L., Eds.; Wiley-Interscience: New York, NY, USA, 1969; Volume 1, pp. 1–102. [Google Scholar]
  38. Mozumder, A. Early production of radicals from charged particle tracks in water. Radiat. Res. 1985, 104, S33–S39. [Google Scholar] [CrossRef]
  39. Swiatla-Wojcik, D.; Buxton, G.V. Modeling of radiation spur processes in water at temperatures up to 300-degrees-C. J. Phys. Chem. 1995, 99, 11464–11471. [Google Scholar] [CrossRef]
  40. Buxton, G.V. The Radiation Chemistry of the Liquid State: (1) Water and homogeneous Aqueous Solutions. In Radiation Chemistry: Principles and Applications; Farhataziz, Rodgers, M.A.J., Eds.; VCH Publishers, Inc.: New York, NY, USA, 1987; pp. 321–349. [Google Scholar]
  41. Allen, A.O. Radiation Chemistry of Aqueous Solutions. J. Phys. Colloid. Chem. 1948, 52, 479–490. [Google Scholar] [CrossRef]
  42. LaVerne, J.A.; Tandon, L. H2 Production in the Radiolysis of Water on UO2 and Other Oxides. J. Phys. Chem. B 2003, 107, 13623–13628. [Google Scholar] [CrossRef]
  43. Petrik, N.G.; Alexandrov, A.B.; Vall, A.I. Interfacial energy transfer during gamma radiolysis of water on the surface of ZrO2 and some other oxides. J. Phys. Chem. B 2001, 105, 5935–5944. [Google Scholar] [CrossRef]
  44. Bricout, M.; Onofri, C.; Debelle, A.; Pipon, Y.; Belin, R.C.; Garrido, F.; Lepretre, F.; Gutierrez, G. Radiation damage in uranium dioxide: Coupled effect between electronic and nuclear energy losses. J. Nucl. Mater. 2020, 531, 10. [Google Scholar] [CrossRef]
  45. De Bona, E.; Benedetti, A.; Dieste, O.; Staicu, D.; Wiss, T.; Konings, R.J.M. Radiation effects in alpha-doped UO2. Nucl. Instrum. Methods Phys. Res. Sect. B-Beam Interact. Mater. Atoms 2020, 468, 54–59. [Google Scholar] [CrossRef]
  46. Mohun, R.; Desgranges, L.; Canizares, A.; Raimboux, N.; Duval, F.; Omnee, R.; Jegou, C.; Miro, S.; Simon, P. Investigating the role of irradiation defects during UO2 oxidative dissolution. J. Nucl. Mater. 2018, 509, 305–312. [Google Scholar] [CrossRef] [Green Version]
  47. Tocino, F.; Szenknect, S.; Mesbah, A.; Clavier, N.; Dacheux, N. Dissolution of uranium mixed oxides: The role of oxygen vacancies vs. the redox reactions. Prog. Nucl. Energy 2014, 72, 101–106. [Google Scholar] [CrossRef]
  48. Corkhill, C.L.; Bailey, D.J.; Tocino, F.Y.; Stennett, M.C.; Miller, J.A.; Provis, J.L.; Travis, K.P.; Hyatt, N.C. Role of Microstructure and Surface Defects on the Dissolution Kinetics of CeO2, a UO2 Fuel Analogue. ACS Appl. Mater. Interfaces 2016, 8, 10562–10571. [Google Scholar] [CrossRef] [Green Version]
  49. LaVerne, J.A. H2 Formation from the Radiolysis of Liquid Water with Zirconia. J. Phys. Chem. B 2005, 109, 5395–5397. [Google Scholar] [CrossRef]
  50. LaVerne, J.A.; Tandon, L. H2 Production in the Radiolysis of Water on CeO2 and ZrO2. J. Phys. Chem. B 2002, 106, 380–386. [Google Scholar] [CrossRef]
  51. Boehm, H.P. Acidic and Basic Properties of Hydroxylated Metal Oxide Surfaces. Disc. Faraday Soc. 1971, 53, 264–289. [Google Scholar] [CrossRef]
  52. Schatz, T.; Cook, A.R.; Meisel, D. Capture of Charge Carriers at the Silica Nanoparticle-Water Interface. J. Phys. Chem. B 1999, 103, 10209–10213. [Google Scholar] [CrossRef]
  53. Reiff, S.C.; La Verne, J.A. Gamma and He Ion Radiolysis of Copper Oxides. J. Phys. Chem. C 2015, 119, 8821–8828. [Google Scholar] [CrossRef]
  54. Was, G.S. Fundamentals of Radiation Materials Science; Springer: Berlin/Heidelberg, Germany, 2007. [Google Scholar]
  55. Li, Z.; Huang, W. Reactivity of hydrogen species on oxide surfaces. Sci. China Chem. Open Access 2021, 64, 1076–1087. [Google Scholar] [CrossRef]
  56. Li, Z.R.; Werner, K.; Chen, L.; Jia, A.P.; Qian, K.; Zhong, J.Q.; You, R.; Wu, L.H.; Zhang, L.Y.; Pan, H.B.; et al. Interaction of Hydrogen with Ceria: Hydroxylation, Reduction, and Hydride Formation on the Surface and in the Bulk. Chem.-Eur. J. 2021, 27, 5268–5276. [Google Scholar] [CrossRef]
  57. Noutak, M.; Freyss, M.; Jomard, G.; Geneste, G. Electron polarons and donor point defects in americium dioxide AmO2. Phys. Rev. B 2020, 101, 024108. [Google Scholar] [CrossRef]
  58. Bo, T.; Lan, J.H.; Zhao, Y.L.; Zhang, Y.J.; He, C.H.; Chai, Z.F.; Shi, W.Q. First-principles study of water adsorption and dissociation on the UO2 (111), (110) and (100) surfaces. J. Nucl. Mater. 2014, 454, 446–454. [Google Scholar] [CrossRef]
  59. Wang, G.X.; Batista, E.R.; Yang, P. Excess Electrons on Reduced AnO2 (111) Surfaces (An = Th, U, Pu) and Their Impacts on Catalytic Water Splitting. J. Phys. Chem. C 2019, 123, 30245–30251. [Google Scholar] [CrossRef]
  60. Weck, P.F.; Kim, E.; Jove-Colon, C.F.; Sassani, D.C. On the role of strong electron correlations in the surface properties and chemistry of uranium dioxide. Dalton Trans. 2013, 42, 4570–4578. [Google Scholar] [CrossRef] [PubMed]
  61. Senanayake, S.D.; Idriss, H. Water reactions over stoichiometric and reduced UO2 (111) single crystal surfaces. Surf. Sci. 2004, 563, 135–144. [Google Scholar] [CrossRef]
  62. Stultz, J.; Paffett, M.T.; Joyce, S.A. Thermal evolution of hydrogen following water adsorption on defective UO2 (100). J. Phys. Chem. B 2004, 108, 2362–2364. [Google Scholar] [CrossRef]
  63. Le Caer, S. Water Radiolysis: Influence of Oxide Surfaces on H2 Production under Ionizing Radiation. Water 2011, 3, 235–253. [Google Scholar] [CrossRef] [Green Version]
  64. Reiff, S.C.; La Verne, J.A. Radiation-Induced Chemical Changes to Iron Oxides. J. Phys. Chem. B 2015, 119, 7358–7365. [Google Scholar] [CrossRef] [PubMed]
  65. Rotureau, P.; Renault, J.P.; Lebeau, B.; Patarin, J.; Mialocq, J.C. Radiolysis of Confined Water: Molecular Hydrogen Formation. Chem. Phys. Chem. 2005, 6, 1316–1323. [Google Scholar] [CrossRef]
  66. Le Caer, S.; Rotureau, P.; Brunet, F.; Charpentier, T.; Blain, G.; Renault, J.P.; Mialocq, J.C. Radiolysis of Confined water: Hydrogen Production at Low Dose Rate. Chem. Phys. Chem. 2005, 6, 2585–2596. [Google Scholar] [CrossRef]
  67. Traboulsi, A.; Vandenborre, J.; Blain, G.; Humbert, B.; Barbet, J.; Fattahi, M. Radiolytic Corrosion of Uranium Dioxide: Role of Molecular Species. J. Phys. Chem. C 2014, 118, 1071–1080. [Google Scholar] [CrossRef] [Green Version]
  68. Essehli, R.; Crumiere, F.; Blain, G.; Vandenborre, J.; Pottier, F.; Grambow, B.; Fattahi, M.; Mostafavi, M. H2 production by gamma and He ions water radiolysis, effect of presence TiO2 nanoparticles. Int. J. Hydrog. Energy 2011, 36, 14342–14348. [Google Scholar] [CrossRef]
  69. Fidalgo, A.B.; Kumagai, Y.; Jonsson, M. The role of surface-bound hydroxyl radicals in the reaction between H2O2 and UO2. J. Coord. Chem. 2018, 71, 1799–1807. [Google Scholar] [CrossRef] [Green Version]
  70. Hansson, N.L.; Tam, P.L.; Ekberg, C.; Spahiu, K. XPS study of external alpha-radiolytic oxidation of UO2 in the presence of argon or hydrogen. J. Nucl. Mater. 2021, 543, 10. [Google Scholar] [CrossRef]
  71. Allen, A.O.; Hochanadel, C.J.; Ghormley, J.A.; Davis, T.W. Decomposition of Water and Aqueous Solutions under Mixed Fast Neutron and Gamma Radiation. J. Phys. Chem. 1952, 56, 575–586. [Google Scholar] [CrossRef]
  72. Martins, P.F.B.D.; Lopes, P.P.; Ticianelli, E.A.; Stamenkovic, V.R.; Markovic, N.M.; Strmcnik, D. Hydrogen evolution reaction on copper: Promoting water dissociation by tuning the surface oxophilicity. Electrochem. Commun. 2019, 100, 30–33. [Google Scholar] [CrossRef]
  73. Strmcnik, D.; Lopes, P.P.; Genorio, B.; Stamenkovic, V.R.; Markovic, N.M. Design principles for hydrogen evolution reaction catalyst materials. Nano Energy 2016, 29, 29–36. [Google Scholar] [CrossRef] [Green Version]
  74. Betova, I.; Bojinov, M.; Lilja, C. Influence of ionic strength on hydrogen generation during interaction of copper with deoxygenated neutral solution. Corrosion Sci. 2021, 188, 13. [Google Scholar] [CrossRef]
  75. Carrasco, J.; Hodgson, A.; Michaelides, A. A molecular perspective of water at metal interfaces. Nat. Mater. 2012, 11, 667–674. [Google Scholar] [CrossRef] [Green Version]
  76. Hodgson, A.; Haq, S. Water adsorption and the wetting of metal surfaces. Surf. Sci. Rep. 2009, 64, 381–451. [Google Scholar] [CrossRef]
  77. Pastina, B.; LaVerne, J.A. Hydrogen Peroxide Production in the Radiolysis of Water with Heavy Ions. J. Phys. Chem. A 1999, 103, 1592–1597. [Google Scholar] [CrossRef]
  78. Pastina, B.; LaVerne, J.A. Effect of Molecular Hydrogen on Hydrogen Peroxide in Water Radiolysis. J. Phys. Chem. A 2001, 105, 9316–9322. [Google Scholar] [CrossRef]
  79. Pastina, B.; Isabey, J.; Hickel, B. The influence of water chemistry on the radiolysis of the primary coolant water in pressurized water reactors. J. Nucl. Mater. 1999, 264, 309–318. [Google Scholar] [CrossRef]
  80. Posiva Oy. Safety Case for the Disposal of Spent Nuclear Fuel at Olkiluoto—Models and Data for the Repository System 2012 Parts 1 & 2; Posiva Oy: Eurajoki, Finland, 2013. [Google Scholar]
  81. SKB. Data Report for the Safety Assessment SR-Site; Swedish Nuclear Fuel and Waste Management Co. (SKB): Stockholm, Sweden, 2010; p. 457. [Google Scholar]
  82. Johnson, L. A Model for Radiolnuclide Release from Spent UO2 and MOX Fuel; NAGRA NAB 13-37; Nationale Genossenschaft für die Lagerung radioaktiver Abfälle: Wettingen, Switzerland, 2014; p. 20. [Google Scholar]
  83. Gobien, M.; Garisto, F.; Kremer, E.; Medri, C. Seventh Case Study: Reference Data and Codes; NWMO-TR-2018-10; Nuclear Waste Management Organization: Toronto, ON, Canada, 2018. [Google Scholar]
Figure 1. Schematic illustration of management and direct disposal of spent nuclear fuel according to the KBS-3 method (Copyright: Posiva Oy).
Figure 1. Schematic illustration of management and direct disposal of spent nuclear fuel according to the KBS-3 method (Copyright: Posiva Oy).
Applsci 11 08566 g001
Figure 2. The macro- and microstructure of a fuel pellet after its irradiation showing the main features, such as cracks, the surrounding cladding, and the pellet–cladding gap, in addition to the grain boundaries (shown in the inset, at a larger scale). The sizes of the pellet–cladding gap and pellet–pellet gap are exaggerated for clarity [5].
Figure 2. The macro- and microstructure of a fuel pellet after its irradiation showing the main features, such as cracks, the surrounding cladding, and the pellet–cladding gap, in addition to the grain boundaries (shown in the inset, at a larger scale). The sizes of the pellet–cladding gap and pellet–pellet gap are exaggerated for clarity [5].
Applsci 11 08566 g002
Figure 3. Mechanism for UO2 dissolution at macroscopic level according to the Matrix Alteration Model, initially formulated in the framework of the EU Spent Fuel Stability project [31] and further developed by Riba et al. [30].
Figure 3. Mechanism for UO2 dissolution at macroscopic level according to the Matrix Alteration Model, initially formulated in the framework of the EU Spent Fuel Stability project [31] and further developed by Riba et al. [30].
Applsci 11 08566 g003
Figure 4. Processes occurring upon the interaction of ionizing radiation and a water molecule leading to the production of radiolytic species in solution. Figure prepared based on schemes for the radiolysis of water reported in refs [39,40].
Figure 4. Processes occurring upon the interaction of ionizing radiation and a water molecule leading to the production of radiolytic species in solution. Figure prepared based on schemes for the radiolysis of water reported in refs [39,40].
Applsci 11 08566 g004
Figure 5. Alternative conceptual model for spent fuel–water interaction, modified from Riba et al. [30].
Figure 5. Alternative conceptual model for spent fuel–water interaction, modified from Riba et al. [30].
Applsci 11 08566 g005
Table 1. Summary of release rates in reducing conditions.
Table 1. Summary of release rates in reducing conditions.
Sample (Reducing Agent)Activity (Bq g−1 UO2)Observations by the AuthorsReference
PWR segment (H2)
50 MWd kg−1 U
5 × 108 aDecreasing rates but propose a fractional release rate of 4 × 10−7 y−1Carbol et al. [11] citing Loida et al. [12,13,14]
PWR powder (H2)
43 MWd kg−1U
4.3 × 108 a3 × 10−7 y−1 based on spent fuel powdersSpahiu et al. [15]
PWR fragment (Fe)
41 MWd kg−1 U
4.1 × 108 aDecreasing dissolution ratesOllila et al. [16]
PWR segments (H2)
43 MWd kg−1 U
4.3 × 108 aDecreasing rates over >1 yearSpahiu et al. [17,18]
PWR segment (H2)
43 MWd kg−1 U
4.3 × 108[U] concentrations 10−5 to 5 × 10−10 M practically constant over > 2 yearsEkeroth et al. [9]
PWR fragment (H2)
67 MWd kg−1 U
1.2 × 109 bDecreasing rates followed by steady-state after 502 daysFors [5]
PWR fragments (H2)
40 MWd kg−1 U
4 × 108 aInhibited dissolution rates over 3 yearsCera et al. [19]
PWR fragments (H2)
65 MWd kg−1 U
1.3 × 109 aDecreasing U concentrations reaching U solubility limits, inhibition of fuel dissolutionPuranen et al. [20]
MOX irradiated (H2)
48 MWd t−1 HM
3 × 109 aDecreasing U concentrations after 494 days, lower than U solubility limitCarbol et al. [8]
Irradiated MOX (Fe)
47 MWd t−1 HM
3 × 109
(beta, gamma 2.4 × 1010) b
No signs of oxidative dissolutionOdorowski 2015 [21]
UO2 un-doped (Fe)1 × 104 bNo measurable dissolution rate over 1 yearOdorowski 2015 [21]
UO2 Pu-doped (Fe)3.8 × 109 bNo measurable dissolution rate over 1 yearOdorowski 2015 [21]
UO2 Pu-doped (Fe)1.8 × 107 bNo measurable dissolution rate over 1 yearOdorowski 2015 [21]
Fresh MOX (7%Pu) (Fe)1.3 × 109 (beta, 9.1 × 109) bNo measurable dissolution rate over 1 yearOdorowski 2015 [21]
UO2 un-doped (Fe)1 × 104 b8.3 × 10−8 y−1 (min) 1.6 × 10−6 y−1 (max)Zetterström Evins et al. [22]
UO2 U-233-doped 5% (Fe)1.6 × 107 b7.5 × 10−8 y−1 (min) 1.1 × 10−6 y−1 (max)Zetterström Evins et al. [22]
UO2 U-233-doped 10% (Fe)3.1 × 107 b4.4 × 10−8 y−1 (min) 1.2 × 10−6 y−1 (max)Zetterström Evins et al. [22]
UO2 U-233-doped 10% (H2)3.3 × 107 bDecreasing total U concentrations over 328 days. No sign of surface oxidation observed.Carbol et al. [8]
a Approximation based on Figure 2-1 in Carbol et al. [11] and assuming 10 years after discharge. b Reported by the authors.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pastina, B.; LaVerne, J.A. An Alternative Conceptual Model for the Spent Nuclear Fuel–Water Interaction in Deep Geologic Disposal Conditions. Appl. Sci. 2021, 11, 8566. https://doi.org/10.3390/app11188566

AMA Style

Pastina B, LaVerne JA. An Alternative Conceptual Model for the Spent Nuclear Fuel–Water Interaction in Deep Geologic Disposal Conditions. Applied Sciences. 2021; 11(18):8566. https://doi.org/10.3390/app11188566

Chicago/Turabian Style

Pastina, Barbara, and Jay A. LaVerne. 2021. "An Alternative Conceptual Model for the Spent Nuclear Fuel–Water Interaction in Deep Geologic Disposal Conditions" Applied Sciences 11, no. 18: 8566. https://doi.org/10.3390/app11188566

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop