49
СИНТЕЗ И СВОЙСТВА НЕОРГАНИЧЕСКИХ СОЕДИНЕНИЙ ISSN 0036-0236, RUSSIAN JOURNAL OF INORGANIC CHEMISTRY, 2019, VOL. 64, NO. 13, PP. 1611–1624. © PLEIADES PUBLISHING, LTD., 2019. On selection of matrix (wasteform) material for higher activity nuclear waste immobilisation ©2019 г. C. M. Jantzen 1 , M. I. Ojovan 2,* 1 The University of South Carolina, USA, Aiken, South Carolina 2 Imperial College London, UK, London, SW7 2AZ, South Kensington Campus, Exhibition Road * e-mail: m.ojovan @ imperial.ac . uk Selection of wasteform materials for higher activity nuclear waste containment is considered. Utilisation of materials such as glasses, ceramics, glass composite materials and cements is discussed as practiced in different countries. Emphasis is on multiple parameter approach on selecting the wasteform where the durability is not solely the most important characteristic. Ключевые слова: nuclear waste, wasteform, glass, ceramics, cement DOI: 10.1134/S0036023619130047 INTRODUCTION There is a generic consensus that the radioactive waste end points (e.g. storage and final disposal) and conditioning methods (e.g. immobilization and packaging in containers) depend on the level of radioactivity and waste radionuclide lifetimes [1, 2]. Safety of storage and disposal is achieved mainly by 1

Imperial College London · Web viewСинтез и свойства неорганических соединений. ISSN 0036-0236, Russian Journal of Inorganic Chemistry, 2019,

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Синтез и свойства неорганических соединений

ISSN 0036-0236, Russian Journal of Inorganic Chemistry, 2019, Vol. 64, No. 13, pp. 1611–1624. © Pleiades Publishing, Ltd., 2019.

On selection of matrix (wasteform) material for higher activity nuclear waste immobilisation

©2019 г. C. M. Jantzen1, M. I. Ojovan2,*

1The University of South Carolina, USA, Aiken, South Carolina

2Imperial College London, UK, London, SW7 2AZ, South Kensington Campus, Exhibition Road

*e-mail: [email protected]

Selection of wasteform materials for higher activity nuclear waste containment is considered. Utilisation of materials such as glasses, ceramics, glass composite materials and cements is discussed as practiced in different countries. Emphasis is on multiple parameter approach on selecting the wasteform where the durability is not solely the most important characteristic.

Ключевые слова: nuclear waste, wasteform, glass, ceramics, cement

DOI: 10.1134/S0036023619130047

Introduction

There is a generic consensus that the radioactive waste end points (e.g. storage and final disposal) and conditioning methods (e.g. immobilization and packaging in containers) depend on the level of radioactivity and waste radionuclide lifetimes [1, 2]. Safety of storage and disposal is achieved mainly by concentration and containment. Containment uses many barriers around the radioactive waste to restrict the release of radionuclides into the environment e.g. the multi-barrier concept. The restricting barriers can be either natural or engineered, e.g. obtained via processing of radioactive waste and using relevant wasteforms. The generic approach is to use more reliable barriers for more hazardous waste, including engineered barriers which result from the radioactive waste treatment and conditioning process. Very low level waste (VLLW) typically does not require immobilization, although some forming packaging could be used. Low level waste (LLW) is typically immobilized using cements, although glasses provide better retention of contaminants. Intermediate level waste (ILW), with a higher hazard than LLW, can be immobilized using glasses, bitumen, cements and geopolymers. LLW and ILW can be packed in high-integrity metal and concrete containers. The most hazardous waste, such as high level waste (HLW), requires the most durable and reliable wasteforms, e.g. ceramics and glasses and containers such as stainless steel canisters for vitrified HLW and copper containers for spent nuclear fuel declared as waste (SNFW).

Immobilization of radioactive wastes in a stable, solid matrix reduces the potential for migration or dispersion of radionuclides. The product of incorporating a waste into a suitable matrix is a wasteform. Immobilization is the conversion of a waste into a wasteform by solidification, embedding or encapsulation. It facilitates handling, transportation, storage and disposal of radioactive wastes. Immobilization of waste is achieved by its chemical incorporation into the structure of a suitable matrix (typically glass or ceramic) so it is captured and unable to escape. Chemical immobilization is typically applied to HLW. Encapsulation of waste is achieved by physically surrounding it in materials (typically bitumen or cement) so it is isolated, and radionuclides are retained. Physical encapsulation is often applied to ILW but can also be used for HLW especially where chemical incorporation of radionuclides in the surrounding matrix is also possible. Choosing the wasteforms and packages in which to immobilize higher activity radioactive wastes (e.g. ILW and HLW) is a crucial decision depending largely on the waste streams and following proven technologies used elsewhere.

Purpose of a Wasteform

The purpose of a wasteform is to immobilize nuclear waste in a solid matrix that is chemically and physically more durable, and more heat and radiation resistant than the waste itself [3-5]. The wasteform is meant to stabilize the waste during packaging, an interim storage period, transportation, emplacement within a geological repository, and post closure disposal period. Notable that deep geologic disposal is considered to be the most suitable option for disposing higher activity radioactive wastes worldwide because of the predicted effectiveness of many natural geological systems in reducing the transport of radionuclides that can be augmented by robust engineered barriers, thereby enhancing disposal facility performance.

Within the repository, the wasteform is one part of a multiple engineered barrier system (EBS). During storage the wasteform is the primary barrier preventing the release of radionuclides into the environment while during post closure disposal, the wasteform will reduce the release of radionuclides from breached and compromised containers that could result due to corrosion, earthquake, human intrusion, igneous intrusion (volcano), or other disruptive phenomena.

The natural barrier system (NBS) primarily establishes the geochemical environment wherein the engineered barriers reside. In some repository concepts, chemical buffers are included in the EBS to reduce the mobility of radionuclides (e.g., using clay backfill), reduce the solubility of certain radionuclides (e.g., by adding phosphates), or reduce the wasteform degradation rates (e.g., by adding reactive materials such as ductile iron). The wasteform, as part of the EBS, is the source of radionuclides and controls the release of those nuclides when in contact with water. Figure 1 shows schematically the various barriers including the wasteform, waste package, and NBS.

Figure 1.

The performance of a repository is assessed using a performance assessment (PA) model in which the effect of each of the barriers on the release and migration of radionuclides is taken into account and coupled to calculate the radionuclide mass flux that could potentially reach the accessible environment should the engineered barriers fail. Typically, interactions such as those among thermal, hydraulic, mechanical, and chemical (THMC) processes are sufficiently complex that models fully describing their behaviour are not fully coupled and the processes are abstracted into an overall PA model. Each significant variable is assigned a probability distribution. Dose conversion factors are then applied to determine the impact on human health and the environment resulting from radionuclides reaching the environment.

The resistance of the wasteform to aqueous corrosion and release of radionuclides in the disposal environment – chemical durability – is a critical parameter.

Immobilization and Matrix Requirements

Wasteform matrix selection and optimization cannot be dissociated from selection of technology and waste characteristics and is a compromise between various constraints including flexibility, ease of processing, waste loading, chemical characteristics of the waste and required lifetime performance which will be affected by irradiation and water alteration [6]. The waste hosting matrix is expected to possess certain features that are desirable for easy processing, handling, storage and eventual disposal:

•Fabrication should be accomplished under reasonable conditions, including thee lowest temperatures to enhance the life of processing equipment and reduce volatilization of fission products and ideally in an air atmosphere, using well established methods to minimize worker dose and capital cost of the facility.

•Adequate viscosity if the wasteform is molten during processing to facilitate homogenization and pouring.

•Adequate electrical conductivity in the case of molten liquids while in Joule Heated Ceramic and Cold Crucible Melters.

•Good mechanical strength and shock resistance so that handling during transportation, storage or disposal is safe.

•Compatibility with the storage container (canister).

•Compatibility with processing equipment (melting temperature, corrosion, refractory, etc.).

The general requirements for matrices which incorporate HLW in wasteforms include, but are not limited to [7]:

•High waste loading: the system must be able to accommodate a significant amount of waste (typically 20-35 %) to minimize the wasteform volume, saving on storage and transportation costs and space in a potential repository (waste minimization concept).

•High radiation stability: the wasteform should show a high tolerance to radiation effects from decay products to reduce phase changes (which may be detrimental to durability) and volume swelling (resulting in cracking).

•Chemical flexibility: if required, the system should be able to accommodate a mixture of radionuclide and other contaminant species, with minimum formation of secondary phases which may compromise wasteform performance or processability.

•Durability: the wasteform should be resistant to alteration under conditions relevant to storage and disposal to minimize release of radiotoxic species.

Wasteform durability is measured using laboratory and field tests [5, 8-10] which measure corrosion in aqueous environments, the better of which try to match the test to the likely conditions in an underground repository and/or determine parameters for PA durability rate law models. Laboratory testing of vitreous wasteform durability, both short- and long-term has been reviewed in [11-13]. Wasteforms also need to be acceptable for disposal in a repository and a range of procedures are used by different countries, so-called Waste Acceptance Criteria (WAC) or Waste Acceptance Product Specifications (WAPS).

The long-term behaviour of a waste disposal facility is a function of the entire disposal system, including the wasteform, engineered barriers, and surrounding environment. In order to assess the ability of a given disposal concept to meet regulatory requirements it is necessary to consider the influence of each of these system components on short-and long-term performance. This is accomplished through the PA process. For HLW many countries are proposing long storage life for the glass wasteforms in steel canisters during geological repository siting and preparation. During that time a great many of the radionuclides will decay leaving the long lived radionuclides as the primary sources that need be considered in a PA. Figure 2 is a schematic of a generic HLW repository which shows the relative role of the wasteform, the role of the multiple barriers (canisters, containers, overpacks, and casks) in the waste disposal system.

Figure 2.

Ultimately the role of the repository or disposal environment is to isolate the waste from the biosphere until all the barriers have failed, i.e. a canister is anticipated to remain intact for time periods such as 100000 years, at which time almost all of the radionuclides will have decayed. While the wasteform is the source term and should be as durable as reasonably possible, multiple barriers must corrode before the wasteform will be exposed to groundwater.

The final evaluation of any wasteform is done using safety assessment (SA) and/or PA cases. SA of nuclear waste disposal evaluates the wasteform and repository performance and its radiological impact on the environment and humans. SA results in data demonstrating either compliance with safety standards or unacceptability of the proposed concept. PA provides data on the performance of all repository subsystems and the repository itself. PA results in data which are used to improve the overall safety and to demonstrate compliance with safety standards and performance targets. PA comprises scenario development and potential impact calculations. As a rule, the impact is characterised by the most probable doses arising from radionuclides which may be released and interact with the biosphere. Scenarios considered are not limited to the most probable but also include abnormal events.

Mathematical models are key components of the SA and PA of waste disposal. A chain of mathematical models is used to describe the long-term degradation of the engineered barriers and the eventual release of radionuclides into ground water and their transport to the biosphere. All studies to date have shown that disposal of radioactive waste including geological disposal, can be safe; adequate isolation from the environment can be assured for hundreds of thousands of years and, thereafter, radionuclide releases are negligible in comparison to natural radiation exposures.

Types of Matrix

Once radioactive waste is generated its management involves a series of steps aimed at its eventual permanent disposal. These include: (1) pre-treatment which may involve collection, segregation and decontamination, treatment aimed at volume reduction and where possible radionuclide removal, (2) conditioning which can involve physical and chemical immobilisation of waste radionuclides into the structure of a stable solid (such as glass, glass composite material or ceramic), or physical encapsulation in a matrix (such as cement or bitumen), followed by packaging often in a steel container. The packaged waste can then be placed in temporary storage under control and where retrievability is ensured until a suitable permanent geological disposal facility is available [6, 10].

Choosing a suitable wasteform (matrix) to use for nuclear waste immobilisation is not easy and its durability is not the sole acceptance criterion. In any immobilisation process where radioactive materials are used, the process and operational conditions can become complicated, particularly if operated remotely and equipment maintenance is required. Therefore, priority is given to reliable, simple, rugged technologies and equipment, which may have advantages over complex or sensitive equipment and processes. A variety of matrix materials and techniques are available for immobilisation. The choice of the immobilisation technology depends on the physical and chemical nature of the waste and the acceptance criteria for the long-term storage and geological disposal facility (GDF) to which the waste will be consigned. A host of regulatory, process and product requirements has led to the investigation and adoption of a variety of matrices and technologies for waste immobilisation. The main immobilisation technologies that are available commercially and have been demonstrated to be viable for higher activity radioactive wastes including ILW and HLW are vitrification, cementation and ceramification [2, 5, 10, 14, 15]. Reports comparing the properties of crystalline ceramics to glass generated considerable controversy [16, 17] in the 1970s and 1980s as “low leachability” had become the main criteria for comparison of vitreous and crystalline wasteforms rather than the multiple criteria presented above. Although SYNROC (SYNthetic ROCk) ceramics had been reported to have superior product quality to glass wasteforms, the high temperature processing, especially for SYNROC-D for defence wastes, formed an intergranular glassy phase from the alkali-and silica in the wastes. This intergranular glass limited the ceramic product stability and durability as often the radionuclides (137Cs and 90Sr) migrated to the glassy phase [18-21]. While ceramics were credited with having higher chemical durability than glasses, if the radionuclides were incorporated in the intergranular glassy phases, they were determined to leach at similar rates as those from glassy wasteforms [22]. This illustrates the danger of considering only one of the general requirements for matrices which incorporate HLW as given above, instead of optimizing the general requirements against one another for the waste stream being considered.

Wasteform Options and Requirements

Nearly every reprocessing nation uses glass as the primary HLW form; most of which use alkali-borosilicate glass; with the exception of Russia which produces an alkali-aluminophosphate glass [2, 5, 10]. Glasses were selected in the 1970s primarily for the following reasons (after References [10, 23]):

· A known and reliable large scale non-nuclear commercial technology;

· More cost-effective process than ceramics and other HLW options;

· High solubility of waste components in the glass;

· Tolerance to variation in waste composition;

· Low raw materials costs;

· Sufficiently high durability to assure long-term performance;

· Continuous, high-throughput, operation of glass melters;

· High resistance to damage from radiation and radioactive decay;

· Reasonably well understood corrosion/release mechanisms for a homogeneous wasteform;

· Existence of natural analogues.

While borosilicates are the predominant glass wasteform phosphate based glasses are used in Russia – a glass similar to alkali-borosilicate with a network of PO43- tetrahedra instead of SiO44- tetrahedra: usually aluminophosphate or iron phosphate. The advantages of these glasses include higher solubility for a number of waste components such as P, Mo, Tc, Fe, Cr, S, and halides. Two disadvantages are the relatively high corrosivity of the melt and higher tendency to crystallize during cooling. Several recent studies suggest that these short-comings have been overcome by new glass formulations [24-28]. However, other wasteforms have several advantages including higher durability in the case of some ceramics or ease of fabrication for wasteforms made at low-temperature. Vitrification is a particularly attractive immobilisation route because of the high chemical durability of the glassy product. HLW has been incorporated into alkali borosilicate or phosphate vitreous wasteforms for many years and vitrification is an established technology in France, Japan, Russia, USA and the UK. Large streams of Low and Intermediate Level Waste (LILW) are planned to be vitrified in the USA, South Korea and Russia [10]. The glass structure is able to accommodate most of the Periodic Table and radioactive species can go into the glass network or other sites [29, 30]. For example, alkalis and alkaline earths (137Cs, 90Sr) sit on the large Na sites in a sodium borosilicate glass. The key issue during production of vitreous wasteforms is to balance the waste loading (to minimise volume), viscosity (ease of pouring), volatilisation (of high vapour pressure species) with wasteform durability and simplicity and economics of the process. For a variety of reasons some species present in wastes including tetravalent actinides, 99Tc, 36Cl, 129I and 14C are inherently difficult to immobilise and dependent on melt pool REDuction/OXidation (REDOX) behaviour. In particular, halogens have low solubility in glasses and are volatile at typical glass processing temperatures of 1150-1200C while the melting temperatures of components such as Al2O3 in a HLW can create glasses that are well above these glass processing temperatures, leading to bubbles or crystals in the glass; the latter becoming de facto Glass Composite Materials (GCMs) [10, 14]. As illustrated in Table 1 a range of wasteform matrices can be considered for immobilizing higher activity waste e.g. HLW and potentially ILW. The choice of wasteform depends on the physical and chemical nature of the waste; the criteria for storage, transportation, and geological disposal; and a host of regulatory, policy, process, and cost considerations.

Table 1

Recent reviews of the developments in wasteform research have been provided in [31-41]. The most recent interest has been associated with the desire to create new nuclear materials as part of new or advanced nuclear fuel cycles that chemically process the Spent Nuclear Fuel Waste (SNFW) [42].

At the other end of the spectrum shown in Table 1 the use of predominantly crystalline ceramic wasteforms (ceramification) has also been proposed including single-phase ceramics such as zircon or zirconolite to accommodate a limited range of active species such as Pu and multiphase systems such as SYNROC to accommodate a broader range of active species [43]. To date these systems have not been extensively used to immobilise active waste.

Recently, however, there has been a trend to systems intermediate between the “completely” glassy or crystalline materials [7, 10, 14, 29, 44 ]. An illustration of nuclear wasteforms used and developed for industrial application is given in Figure 3.

Figure 3.

Typically nondurable crystals are Na2SO4, Na2MoO4, NaF and NaCl while durable crystals include BaAl2Ti6O7, CaZrTi2O7, CaTiO3, CaMoO4, TiO2, and spinels such as (Ni,Mn)(Fe,Cr)2O4.

GCMs sit between the fully amorphous glasses and fully crystalline ceramics and contain both glassy and crystal phases. GCMs include: (i) glass ceramics where a glassy wasteform is crystallised in a separate heat treatment (ii) materials in which a refractory waste such La2Zr2O7 is encapsulated in glass to immobilise minor actinides and (iii) wasteforms which form glass and crystals during processing such as those made at La Hague by cold crucible melting which partly crystallise on cooling [45] derived from difficult wastes such as the French HLW U-Mo-containing materials.

GCMs may be used to immobilise long-lived radionuclides (such as actinide species) by incorporating them into the more durable crystalline phases, whereas the short-lived radionuclides may be accommodated in the less durable vitreous phase. Historically, crystallisation of vitreous wasteforms had been regarded as undesirable as it has the potential to alter the composition (and hence durability) of the remaining continuous glass phase which would (eventually) come into contact with water.

However, there has been a recent trend towards higher crystallinity in ostensibly vitreous wasteforms so that they are more correctly termed GCMs. This is particularly apparent in the development of hosts for more difficult wastes or where acceptable durability can be demonstrated even where significant quantities of crystals (arising from higher waste loadings) are present such as the high sodium Hanford wastes. Acceptable durability will result if the active species are locked into the crystal phases that are encapsulated in a durable, low activity glass matrix. The GCM option is currently being considered in many countries including Australia, France, Russia, South Korea, UK and USA.

Cement matrices [46, 47] are included in this section because of the IAEA definition of HLW being anything slated for geological disposal. As a result much waste, which in the UK would be regarded as ILW as it is not heat generating, is regarded as HLW. Hydraulic cements are inorganic materials that have the ability to react with water under ambient conditions to form a hardened and water-resistant product. The most common cements are those based on calcium silicates, such as the Portland cements. Cementation of radioactive waste has been practised for many years basically for immobilisation of Low Level Waste (LLW) and Intermediate Level Waste (ILW) and commonly (in France and the UK and US in particular) composite systems are used with high proportions of blast furnace slag or pulverised fuel ash compared to Portland cement, typically a 9:1 ratio to reduce setting heat generation and improve wasteform durability. The waste is physically encapsulated by the (amorphous) calcia-silica-hydrate gel and other phases in the microstructure, hence its inclusion next to GCMs in Table 1. A summary of the different wasteforms currently used or under development in various countries is given in Table 2.

Table 2

Glasses

The most common matrix for immobilizing higher activity wastes (HLW and ILW) is glass mainly due to its ease of processing, flexible structure able to accommodate a range of radionuclides, and high durability [5, 10]. Generically phase separated glasses (glass-in-glass phase separation) should be avoided. Control of REDuction/OXidation (REDOX) of the melt during the processing also helps retain radionuclides. Vitrified HLW is predominantly glass although it may contain some crystals and bubbles. Several types of glasses have been proposed for HLW conditioning, however two types of glasses namely borosilicate and phosphate glass are currently used for engineering scale HLW vitrification in various countries (Table 2). Borosilicate glass is widely accepted as a waste matrix due to the acceptable chemical durability of glass. The formation temperature is moderate, in the range 950-1100 oC. The modifiers including alkali oxide and boric oxide lower the formation and pouring temperatures of the glass but also reduce chemical durability. To maintain a balance between these two properties i.e., (i) lower formation and pouring temperatures and (ii) excellent chemical durability, multivalent modifiers such as MnO, TiO2, BaO, Fe2O3, ZnO, CaO and MgO can be added if not already present in the waste being solidified. However, too many divalent cation oxides such as BaO, CaO, and MgO can cause undesirable glass-in-glass phase separation in borosilicate glasses [48] which can negatively impact glass durability.

An important aspect in the use of borosilicate and phosphate glasses for HLW is the solubility of key species in the glasses. Table 3 shows the solubility of U, Pu, Am, Cm and Np in a range of glass compositions (see also [49]).

Table 3

It is important to note that the solubility is also a function of the glass REDOX as driven by the processing conditions. Phosphate glass matrices are suitable for wastes rich in molybdenum, chromium, sulphates, strontium, and uranium. These elements have lower solubility in borosilicate glasses. Typical phosphate compositions incorporate about 30% of waste oxides. These glasses have low melting temperature in the range of 900 oC and low viscosity at pouring temperatures. However, as the main raw material is phosphoric acid there are concerns about container and melter corrosion. Also, these glasses have poorer mechanical strength and are more prone to devitrification compared with borosilicate glasses. Some wastes are not amenable to incorporation into glass structures e.g., because they are volatile (e.g., alkalis) or too refractory to melt (e.g., Al2O3, ZrO2) or they do not dissolve in silicate melts (noble metals). These may require alternative matrices or if present in small quantities may stay suspended in the glass and not impact glass behaviour.

Ceramics

Almost fully crystalline matrices have been developed for immobilizing reprocessing waste but not yet commercially used such as SYNROC by ANSTO. In contrast there are many single phase ceramic matrices for stabilization of a variety of radionuclides. A more detailed description of a variety of ceramic wasteforms can be found in [43] including a discussion of radiation damage.

Multiphase systems have been developed to immobilize complex multi-radionuclide waste streams. These include SYNROC, which consists of the titanates hollandite (BaAl2Ti6O16), zirconolite (CaZrTi2O7), perovskite (CaTiO3) and TiO2. The hollandite mainly fixes the Cs and Rb and some process chemicals, whereas actinides and rare earth fission products (FPs) are bound in zirconolite and perovskite, and the latter also incorporates Sr. Important that SYNROC compositions can be altered so they contain phases which target radionuclides in a particular waste stream. These minerals can accommodate various actinides and FPs in their crystalline structures which results in good resistance to chemical alteration even under hydrothermal conditions. Single-phase ceramics have been examined targeting difficult species such as Pu. Pyrochlore-based ceramics such as those based on zirconolite have been developed for immobilization of Pu and zircon (ZrSiO4) was examined extensively [43, 50], although its tendency to go amorphous (metamictisation) under self-irradiation may limit its application. Ceramics tend to be stable in aqueous environments but the alteration in durability arising from loss of crystallinity by irradiation may impact this behavior [51]. Finally, it should be noted that SNFW is a crystalline ceramic (uranium oxide or mixed plutonium/uranium oxides in MOX) but one which was not designed with post irradiation stability in mind.

Glass Composite Materials

Between the almost fully glass and fully crystalline matrices lie Glass Composite Materials. Early work on GCMs [52] was performed at Pacific Northwest National Laboratory (PNNL) and the Pennsylvania State University (PSU) in the 1970s on “supercalcine ceramics”, regarded as the forerunner of SYNROC, with the aim of achieving a waste loading of 60% by additions to HLW liquids which on calcination and sintering at ~1100 oC formed durable crystalline phases [53]. For high density a hot press was needed along with silica which formed a liquid, i.e. liquid phase sintering, at the formation temperature and an intergranular glass on cooling. Often the intergranular glass contained the radionuclides rather than the desired ceramic phases. GCMs as alternative matrices are still being developed e.g. to enable increased waste loading or to incorporate difficult wastes. GCMs include: (i) Glass ceramics where a glassy wasteform is crystallized in a separate heat treatment, (ii) GCMs in which e.g. a refractory waste is encapsulated in glass such as hot pressed lead silicate glass matrix encapsulating up to 30 vol% of La2Zr2O7 pyrochlore crystals to immobilise minor actinides [54], (iii) GCM formed by pressureless sintering of spent clinoptilolite from aqueous waste processing [55], (iv) Some difficult wastes such as the French HLW UMo-containing materials were immobilized in a GCM termed UMo glass formed by CCIM and which partly crystallises on cooling [29], (v) alkali Mo-containing wastes are immobilised in Russia in a yellow phase GCM containing up to 15 vol.% of sulphates, chlorides and molybdates [56], and (vi) GCMs which immobilise ashes from incineration of solid radioactive wastes [57]. Note that alkali-rich wastes at the Hanford site are also immobilized in glassy wasteforms with high crystal contents which characterize them as GCMs [58].

GCMs may be used to immobilize long-lived radionuclides (such as actinide species) by incorporating them into the more durable crystalline phases, whereas the short-lived radionuclides may be accommodated in the less durable vitreous phase. Historically, crystallization of vitreous wasteforms has always been regarded as undesirable as it has the potential to alter the composition (and hence durability) of the remaining continuous glass phase which would (eventually) come into contact with water. That is why it is also important to determine that the grain boundaries created by these durable crystals do not create accelerated grain boundary dissolution which can impact the overall wasteform durability.

However, there has been a recent trend towards higher crystallinity in ostensibly vitreous wasteforms so that they are more correctly termed GCMs. This is particularly apparent in the development of hosts for more difficult wastes or where acceptable durability can be demonstrated even where significant quantities of crystals (arising from higher waste loadings) are present such as in some of the Hanford wasteforms. Acceptable durability will result if the active species are locked into the crystal phases that are encapsulated in a durable, low activity glass matrix or the active species are locked into the glass and the crystalline species contain no active species. The former occurs in the Mo GCMs while the latter occurs in the US HLW defense wastes. The GCM option is currently being considered in many countries including Australia, France, Russia, South Korea, UK and USA. The processing, compositions, phase assemblages and microstructures of GCMs may be tailored to achieve the necessary material properties.

GCMs are under consideration for a variety of partitioned difficult wastes such as 129I separated from SNFW [59] and 99Tc as pertechnetate as well as e.g. spent clinoptilolite filters. One option being examined for the zeolite wastes from the Fukushima clean-up is use of a mobile vitrification system using a form of in-can melting. In the Areva process for cleaning up contaminated sea water a sludge is created and the Crystalline Silico Titanate Ion Exchange inorganic resin CST IE-911 is also being used. No decision has yet been made on what to do with the spent resins but cementation and vitrification are being considered [60].

Novel technologies

Apart from conventional cold isostatic pressing (CIP) and cold uniaxial pressing (CUP) of ceramic precursors (calcined waste and additives) followed by sintering, a number of novel thermal methods are at various stages of development for immobilising difficult wastes such as those in the legacy ponds and silos at Sellafield, UK and at the Hanford site in the USA (Figure 4).

Figure 4

Hot isostatic pressing (HIP) and hot uniaxial pressing (HUP) are mature technologies in the ceramics and metallurgical processing fields but have had limited application in immobilization of radioactive wastes. The joint operation of high temperature and pressure increases the speed of sintering and reduces the fraction of volatiles lost during processing compared to pressureless processing. Hot pressing is an ideal technology for production of ceramic wasteforms which would otherwise be difficult to obtain with full density and for immobilizing wastes containing highly volatile ions such as TcO4-, I-, or Cl-. The HIP process is typically operated up to ~1250 °C and ~500 atm. when producing multiphase ceramics, but can vary the temperature and pressure widely to accommodate waste-containing systems.

Novel non-thermal (cement like) wasteforms are discussed in References [61-65]. These can be used together or alone to treat some of the problematic legacy wastes [66]. The flexibility of cement and other non-thermal systems means that a toolbox of systems made at close to room temperature are being developed (Figure 5) applicable to the complex array of wastes which they must encapsulate [67].

Figure 5

In particular, some metals (such as Al, Mg and U) are unsuitable for immobilisation in the alkali composite cement systems currently employed as they react with them in an expansive manner forming gases (e.g. H2). Alternative non-thermal systems for encapsulating these reactive metals based on calcium sulpho-aluminate cements and geopolymers [68] have shown promise in early investigations. A number of non-thermally produced wasteforms with improved durability and compatibility with specific ILWs are under development.

Processing technologies

In any immobilization process where radioactive materials are used, the process and operational conditions are complicated by operation and maintenance in a remote environment. Therefore priority is given to reliable, simple, rugged technologies and equipment, over complex or sensitive equipment. As described above a variety of wasteform materials are available for immobilization. The technologies commercially available to process these forms summarized in Table 4 [2, 5, 6, 15].

Table 4

Conclusions

Section of wasteforms for nuclear waste immobilisation is a complex task. Choosing a suitable wasteform for nuclear waste immobilisation is difficult and durability is not the sole criterion. In any immobilisation process where radioactive materials are used, the process and operational conditions can become complicated, particularly if operated remotely and equipment maintenance is required. Therefore priority is given to reliable, simple, rugged technologies and equipment, which may have advantages over complex or sensitive equipment. A variety of matrix materials and techniques is available for immobilisation. The choice of the immobilisation technology depends on the physical and chemical nature of the waste and the acceptance criteria for the long-term storage and disposal facility to which the waste will be consigned. A host of regulatory, process and product requirements has led to the investigation and adoption of a variety of matrices and technologies for waste immobilization. For higher activity nuclear waste the main immobilisation technologies industrially used are cementation, vitrification and ceramication.

Acknowledgements

Authors acknowledge contribution to this overview of W.E. Lee, A.I. Orlova, S.V. Stefanovsky, F. Takats, and late E.R. Vance.

References

1. IAEA Nuclear Energy Series No. NW-G-1.1, Policies and Strategies for Radioactive Waste Management, IAEA, Vienna (2009).

2. Ojovan M.I. Handbook of advanced radioactive waste conditioning technologies. Woodhead, Cambridge, 512 p. (2011).

3. Lutze W., Ewing R.C. Radioactive Waste Forms for the Future, Elsevier Science Publishers, Amsterdam, The Netherlands. 778pp. (1988).

4. IAEA Technical Report Series 345. Concepts for the Conditioning of Spent Nuclear Fuel for Final Waste Disposal (1992).

5. National Research Council. Waste Forms Technology and Performance: Final Report, National Academies Press, Washington, DC (2011).

6. IAEA TECDOC-1817. Selection of Technical Solutions for the Management of Radioactive Waste. IAEA, Vienna (2017).

7. Lee W.E., Ojovan M.I., Stennett M.C., Hyatt N.C. // Adv. Applied Ceramics. 2006. 105 [1] 3-12.

8. IAEA TECDOC-1537. Strategy and Methodology for Radioactive Waste Characterization, IAEA, Vienna (2007)

9. Ebert W.L. Testing Protocols to Support Waste Form Development, Production, and Acceptance. U.S. DOE Report GNEP-WAST-WAST-AI-RT-2008-000302 (September 2008).

10. Ojovan M.I., Lee W.E. An introduction to nuclear waste immobilisation. Second edition, Amsterdam, Elsevier (2013).

11. Jantzen C.M., Brown K.G., Pickett J.B. // International Journal of Applied Glass Science, 1 [1], 38-62 (2010)

12. Jantzen C.M., Crawford C.L., Pareizs J.M., Pickett J.B. Accelerated Leach Testing of GLASS (ALTGLASS): I. Waste Glass Hydrogel Compositions and Accelerated Dissolution. US DOE Report SRNL-STI-2014-00274 (2014).

13. Jantzen C.M., Crawford C.L., Pareizs J.M., Pickett J.B. Accelerated Leach Testing of GLASS (ALTGLASS): II. Leachate-Hydrogel Interactions, Glass Structure, and Accelerated Dissolution. US DOE Report SRNL-STI-2014-00381 (2014).

14. Ojovan M.I., Lee W.E. // Metallurgical and Materials Transactions A, 42A, 837-851 (2011).

15. Lee W.E., Ojovan M.I., Jantzen C.M. Radioactive Waste Management and Contaminated Site Clean-up: Processes, Technologies and International Experience, Woodhead Publishing, Oxford, UK, 879 pp. (2013).

16. Kerr R.A. // Science 204, 289-291 (1979).

17. Kerr R.A. // Science 205, 287-289 (1979).

18. Clarke D.R. // J. Am. Ceram. Soc., 64 [6] C–89–C–90 (1981).

19. Zhang Z., Carter M. L. // J. Am. Ceram. Soc., 93[3] 894-899 (2010).

20. Buykx W.J., Hawkins K, Levins D.M., Mitamura H., Smart R. St. C., Stevens G.T., Watson K.G., Weedon D., White T.J. // J. Am. Ceram. Soc., 71 [8] 678–88 (1988).

21. Cooper J.A., Cousens D.R., Hanna J.A., Lewis R.A, Myhra S., Segall R.L., Smart R. St.C., Turner P.S., White T.J. // J. Am. Ceram. Soc., 69 [4] 347–52 (1986).

22. Jantzen C.M., Clarke D.R., Morgan P.E.D., Harker A.B. // J. Am. Ceram. Soc., 65 [6], 292-300 (1982).

23. Vienna J.D. Nuclear Waste Glasses. In: Properties of Glass Forming Melts, eds. L.D. Pye, I. Joseph and A. Montenaro, pp. 391-404. CRC Press, Boca Raton, FL (2005).

24. Day D.E., Brow R.K., Ray C.S., Kim C.W., Reis S.T., Vienna J.D., Peeler D.K., Johnson F.C., Hansen E.K., Gevigny N.R., Solberg N.R., Pegg I.L., Gan H. // Proc. Waste Management 2012, paper #12240 Phoenix, AZ (2012).

25. Day D.E., Wu Z., Ray C.S., Hrma P. // Journal of Non-Crystalline Solids, 241:1-12 (1998).

26. Kim C.W., Ray C.S., Zhu D., Day D.E., Gombert D., Aloy A., Mogus-Milankovic A., Karabulut M. // Journal of Nuclear Materials, 322(2-3):152-164 (2003).

27. Kim C.W., Zhu D., Day D.E., Kim D.S., Vienna J.D., Peeler D.K., Day T.E., Neidt T. Iron Phosphate Glass for Immobilization of Hanford LAW. Environmental Issues and Waste Management Technologies in the Ceramic & Nuclear Industries IX, 309-318 pp. American Ceramic Society, Westerville, OH (2004).

28. Sengupta P. // Journal of Hazardous Materials, 235-236, 17-28 (2012).

29. Caurant D., P. Loiseau O. Majerus V. Aubin-Chevalsdonnet, I. Bardez, A. Quintas. Glasses, Glass-Ceramics and Ceramics for Immobilization of Highly Radioactive Nuclear Wastes, Nova Science Publishers, New York, 359 pp. (2009).

30. Vernaz E. Nuclear Waste Conditioning. CEA Monograph, CEA Saclay and Groupe Moniteur (Éditions du Moniteur), Paris, 150pp. (2009).

31. Jantzen C.M. Development of Glass Matrices for HLW Radioactive Wastes. In: Handbook of Advanced Radioactive Waste Conditioning Technologies, M. Ojovan, Woodhead Publishing, Oxford, 230-292 (2011).

32. Ewing R.C. // Proc. Natl. Acad. Sci. USA 96, 3432-3439 (1999).

33. Ewing R.C. //. Canadian Mineralogist, 39, 697-715, (2001).

34. Ewing R.C., Weber W.J., Lian J. // Journal of Applied Physics, 95, 5949-5971 (2004).

35. Lumpkin G.R. // Journal of Nuclear Materials, 289, 136-166 (2001).

36. Lumpkin G.R. // Elements, 2, 365-372, (2006).

37. Donald I.W., Metcalfe B.L., Taylor R.N.J. // J. Mater. Sci. 32, 5851-5887 (1997).

38. Donald I.W. Waste Immobilization in Glass and Ceramic Based Hosts, John Wiley & Sons, London, UK, 507 pp. (2010).

39. Yudintsev S.V., Stefanovsky S.V., Ewing R.C. Actinide Host Phases as Radioactive Waste Forms. Chapter 13 in S.V. Krivovichev, P.C. Burns and I. Tananaev (Editors) Structural Chemistry of Inorganic Actinide Compounds, Elsevier, Amsterdam, pp. 457-490 (2007).

40. Ewing R.C., Weber W.J. Actinide Waste Forms and Radiation Effects. In: L.R. Morss, N. Edlestein and J. Fuger. The Chemistry of the Actinide and Transactinide Elements, vol. 6, Springer, Amsterdam (2011).

41. Stefanovsky S.V., Yudintsev S.V., Giere R., Lumpkin G.R. Nuclear Waste Forms. Geological Society of London Special Publications, v. 236, 37-63 (2004).

42. Peters M.T., Ewing R.C. // Journal of Nuclear Materials, 362, 395-401 (2007).

43. Burakov B.E., Ojovan M.I., Lee W.E. Crystalline Materials for Actinide Immobilization, Imperial College Press, London (2011).

44. Ojovan M.I., Lee W.E. New Developments in Glassy Nuclear Waste forms, Nova Science Publishers, New York, USA (2007).

45. Henry N., Deniard P., Jobic S., Brec R., Fillet C., Bart F., Grandjean A., Pinet O. // J. Non-cryst. Solids 333, 199-205 (2004).

46. IAEA TECDOC-1791. The Behaviour of Cementitious Materials in Long Term Storage and Disposal of Radioactive Waste, IAEA, Vienna, (2013).

47. Abdel Rahman R.O., Rahimov R.Z., Rahimova N.R., Ojovan M.I.. Cementitious Materials for Nuclear Waste Immobilization. Wiley, Chichester (2015).

48. Volf M.B. Chemical Approach to Glass, Glass Science and Technology, V. 7, Elsevier Science Publishing Co., Inc, New York, 594 pp (1984).

49. Gin S., Jollivet P., Tribet M., Peuget S., Schuller S. // Radiochim. Acta; 105 (11): 927–959 (2017).

50. James M., Carter M.L., Ahang Z., Zhang Y., Wallwork K.S., Avdeev M., Vance E.R. // Jour. Am. Ceram. Soc. 93[10] 3464-3473 (2010).

51. Weber W.J., Ewing R.C., Catlow C.R.A., Diaz de la Rubia T., Hobbs L.W., Kinoshita C., Matzke Hj., Motta A.T., Nastasi M., Salje E.K.H., Vance E.R., Zinkle S.J. // J. Mat. Res. 13 [6], 1434-1484 (1998).

52. Hayward P.J. Glass Ceramics. In: Radioactive Waste Forms for the Future, (Eds. W. Lutze and R.C. Ewing), pp. 429-493. Elsevier Science Publishers, Amsterdam (1988).

53. McCarthy G.J. // Nuclear Technol. 32, 92-105 (1977).

54. Boccaccini A.R., Bernardo E., Blain L., Boccaccini D.N. // J. Nucl. Mater. 327, 148-158 (2004).

55. Juoi J.M., Ojovan M.I., Lee W.E. // J. Nucl. Mater. 372, 358-366 (2008).

56. Sobolev I.A., Ojovan M.I., Scherbatova T.D., Batyukhnova O.G. Glasses for Radioactive Waste, Moscow, Energoatomizdat (1999).

57. Sobolev I.A., Dmitriev S.A., Lifanov F.A., Kobelev A.P., Stefanovsky S.V., Ojovan M.I. // Glass Technology, 46, 28-35 (2005).

58. Kim D.S., Smith D.E., Vienna J.D., Sevigny G.J., Hrma P.R., Buchmiller W.C., Schweiger M.J., Tixier J.S., Matyas J., Yeager J.D., Crum J.V., Belew K.B. Development and Testing of ICV Glasses for Hanford LAW. US DOE Report PNNL-14351, Pacific Northwest National Laboratory, Richland, Washington, (2003).

59. Garino T.J., Nenoff T.M., Krumhansl J.L., Rademacher D.X. // J. Am. Ceram. Soc. 94 [8] 2412-2419 (2011).

60. Farid O., Shih K., Lee W.E., Yamana H. Fukushima: The Current Situation and Future Plans. In: W.E. Lee, M.I. Ojovan, C. M. Jantzen,Radioactive Waste Management and Contaminated Site Clean-up: Processes, Technologies and International Experience. pp 744-771. Woodhead UK, (2013).

61. Gougar M.L.D., Scheetz B.E., Roy D.M. // Waste Mangement, 16 [4], 295-303 (1996).

62. Allen P.G., Siemering G.S., Shuh D.K., Bucher J.J., Edelstein N.M., Langton C.A., Clark S.B., Reich T., Denecke M.A. // Radiochimica Acta, 76[1-2], 77-86 (1997).

63. Bart F., Cau-Di-Coumes, C., Frizon F., Lorente S. Cement-Based Materials for Nuclear Waste Storage. Springer-Verlag, New York, 266pp. (2013).

64. Spence R.D. Chemistry and Microstructure of Solidified Waste Forms. CRC Press, 288pp. (1992).

65. Noyes R. Nuclear Waste Cleanup Technology and Opportunities, Noyes Publications, Park Ridge, NJ, 454pp (1995).

66. Lee W.E., Grimes R.W. // Energy Materials, 1, 22-24 (2006).

67. Milesone N.B. // Adv. Appl. Ceram. 105, 13-18 (2006).

68. Vance E.R., Perera D.S. Development of Geopolymers for Nuclear Waste Immobilisation”. In:M.I. Ojovan. Handbook of Advanced Radioactive Waste Conditioning Technologies. Woodhead Publishing, Cambridge, UK, 207-229 (2011).

Table 1. Classification of higher activity wasteforms

Glasses

Glass Composite Materials (GCM’s)

Cements

Crystalline Ceramics

Borosilicate

Magnox, Blend, UK.

Glass Ceramics: Zirconolite for MA or Pu.

Multiphase

Composite Cements

UK and France, WVDP US

Single Phase: zirconolite for MA or Pu, hollandite for Cs

Borosilicate

DWPF*, WVDP**, USA

Crystal Waste Encapsulated in Glass Matrix, glass bonded sodalite (PNNL/KAERI and ANL)

Novel cements,

e.g. calcium sulphoaluminates

Multiphase (e.g., SYNROC) compositions tailored for specific waste streams

Borosilicate

R7/T7 (France), IR111 (India)

Crystal-tolerant Glasses:

U-Mo CCM French wastes, Hanford US wastes

Flyash geopolymeric cements at SRS for low activity fraction of HLW wastes, US

Spent Nuclear Fuel Waste (SNFW)

Aluminophosphate, Russia

* Defense Waste Processing Facility; **West Valley Demonstration Project

Table 2. Summary of matrices (wasteforms) used in different countries

Country

SNFW

Glass

Cement

Ceramics

GCM

China

No

Yes

May be under development

Finland

Yes

No

No

No

No

France

No, closed fuel cycle

Yes for HLW

Yes for ILW

May be under development

Yes (CCIM) UMo wasteforms, zirconolite glass ceramics under development

Germany

May be under development

India

No, closed fuel cycle

Yes for HLW

Yes for ILW

Japan

No, closed fuel cycle

Yes for HLW

Yes for LLW and ILW

No

Under development for some of the wastes at Fukushima.

Korea, Republic of

Yes

Yes for LILW

Yes for LILW

No

No

Russia

No, closed fuel cycle

Yes for HLW and ILW

Yes at Mayak and Radon (ILW and LLW)

Under development

Under development in CCIM and also full scale plasma melting at Radon

Sweden

Yes

No

No

No

No

UK

Yes, AGR* and PWR from Sizewell B

Yes for HLW

Yes for ILW

Under development for Pu

Under development for ILW from Sellafield.

USA

Reference wasteform for commercial power reactor fuels

Wasteform for defense HLW and WVDP reprocessing waste.,

Potential wasteform for advanced fuel cycles.

Yes for LLW and LAW at SRS and WVDP.

Under development for some advanced fuel cycle waste streams

Under development for some advanced fuel cycle waste streams and Hanford WTP

*Advanced Gas-cooled Reactor

Table 3. Incorporation limits of key radioactive species in various glasses

Glass type

Processing temperature, oC

Specie

Oxide AnO2 solubility, mass %

Borosilicate

1150

U(IV)

9

1150

U(VI)

40

1200-1300

U

25 – 28

1100, air atmosphere

Pu(IV)

0.55

1200, air atmosphere

Pu(IV)

<2

1500, air atmosphere

Pu(IV)

<10

1400 +graphite

Pu(III)

>13, <25

1000

Am

2

1175

Am

5

1000, air atmosphere

Np(IV)

2

1250, air atmosphere

Np(IV)

3

1350

Np(III)/Np(IV)

>5

La-borosilicate

1450, air atmosphere

Pu

>7, <10

1500

Pu

>11.4

1450, Air atmosphere

Am

0.1

Sodium silicate

1250, CO/CO2

U(VI)

19

1250, reducing conditions

U(VI)

52.7

1250

Np(V)

33

Phosphate

1100

Pu(III)

>2, <10

Table 4. Wasteform processing technologies for higher activity nuclear wastes, after [5].

Processing Technology

Processing Mode

Treatment and Waste Stream Scale

Wasteforms Produced

Advantages

Disadvantages

Joule Heated Melter (JHM)

Continuous

Large

Boro Silicate Glass, Other Glasses (LaBS, FeP, AIP, chalcogenide and others)

Proven technology; typically operates with a "cold cap" to minimize volatility of species of concern

Electrode and refractory erosion may be a problem; solubility control of certain species

Advanced Joule Heater Melter (AJHM)

Continuous

Large

Boro Silicate Glass, Glass Ceramic Materials, Other Glasses (LaBS, FeP, AIP, chalcogenide and others)

Increases throughput and melt rate compared to JHM

Operates with minimal or no "cold cap" with associated increases in volatility of species of concern

Cold Crucible Induction Melter

(CCIM)

Continuous

Large

Boro Silicate Glass, Glass Ceramic Materials, Other Glasses (LaBS, FeP, AlP, chalcogenide and others), crystalline ceramics/simple oxides, Metal Matrix

Allows the processing of corrosive glasses; No refractories; No metallic or oxide electrodes;water cooled; self-cleaning; high purity,can be stirred if needed; increases capacity compared to JHM and AJHM;can operate at higher temperatures than JHM and AJHM; operates with a "cold cap" to minimize volatility

Higher temperature operation can increase volatilization of species of concern but "cold cap" coverage minimizes these impacts

Cold Press and Sinter

Batch

Small

Glass Ceramic Materials, crystalline ceramics/simple oxides, metal matrix,zeolites, hydroceramic

Higher waste loadings; minimum diposal volume

Usually small scale; may require pre-calcining or pre-treating waste to an oxide form to avoid shrinkage of form

Hot Isostatic Pressing (HIP)

Batch

Small

Borosilicate glass (lab demonstration only), Glass Ceramic Materials, crystalline ceramics/simple oxides, metal matrix,zeolites, hydroceramic

Zero off-gas emissions; higher waste loadings; minimum diposal volume; mature flexible technology; no major secondary wastes; mature industrial process

Processes small quantities; can overpressurize if large amounts of volatiles (e.g. nitrates/hydrates) are present; requires pre-calcining or pre-treating waste to an oxide form (shrinkage handled by dumbbell-shaped canisters)

Hot Uniaxial Pressing (HUP)

(largely supplanted by HIP)

Batch

Small

Glass Ceramic Materials, crystalline ceramics/simple oxides, metal matrix,zeolites, hydroceramic

Higher waste loadings; minimum disposal volume, mature flexible techology; mature industrial process

Usually small scale; may require pre-calcining or pre-treating waste to an oxide form for shrinkage control

In-Container Vitrification (ICV); also known as "Bulk Vitrification"

Batch

Depends on container size (could be medium or large)

Borosilicate glass; Glass Ceramic Materials, Other Glasses (LaBS, FeP, AlP, chalcogenide and others)

Relatively cheap and simple for low activity wastes or contaminated soils; not applicable to HLW

Inhomogeneous wasteforms; no temperature control so radionuclide vaporization is high; little convection in melt causes processing problems

Self-sustaining Vitrification (SSV)

Batch

Small

Glass Ceramic Materials

Low capital requirements; can be used to process small amounts of wastes at remote locations

May require some pre-processing, for example, grinding of the waste and pre-mixing

Fluidized Bed Steam Reforming (FBSR)

Continuous

Large

Crystalline ceramics/simple oxides, zeolites as formed; hydroceramics, geopolymers (as encapsulated)

Industrially proven technology; acidic or basic tank wastes processed without neutalization; destroys organics and nitrates;converts aqueous components in a waste to stable water insoluble mineral products; immobilizes S, Cl, and F are in a stable mineral form, with no secondary waste

Product is granular and requires a high integrity container (HIC) or encapsulation in a binder to make a glass ceramic material, a geopolymer, or a hydroceramic; radionuclide partitioning amonst the phases needs to be further studied

Cyclone Furnaces

Continuous

Large

Boro Silicate Glass, Glass Ceramic Materials, Other Glasses (LaBS, FeP, AlP, chalcogenide and others), crystalline ceramics/

simple oxides, metal matrix

Suitable for soils containing low volatility radionuclides;

Secondary Recovery Process to treat off gases may be necessary

Rotary Kilns

Continuous

Large

Boro Silicate Glass, Glass Ceramic Materials, Other Glasses (LaBs, FeP, AlP, chalcogenide and others), crystalline ceramics/

simple oxides, metal matrix

Established industrial practice

Secondary recovery process to treat off gases may be necessary; not proven for nuclear waste processing; used as pre-processing in French HLW vitrification

Electric Arc Furnaces

Batch

Medium/Large

Boro Silicate Glass, Glass Ceramic Materials, Other Glasses (LaBS, FeP, AIP, chalcogenide and others), Mineral Ceramic/Simple Oxides, Metal Matrix

Established Industrial Practice; Similar technology is used for ICV

High temperatures (1600°C); requires significant off-gas treatment; volatilization of constituents of concern

Plasma Furnaces

Batch

Small

Boro Silicate Glass, Glass Ceramic Materials, Other Glasses (LaBS, FeP, AIP, chalcogenide and others), Mineral Ceramic/Simple Oxides, Metal Matrix

Plasma generating electrode erosion; Efficient for the destuction of organics

No large scale practice; high tempeatures, increaced volatilization

Microwave Heating

Batch or Continuous

Small

Boro Silicate Glass, Glass Ceramic Materials, Other Glasses (LABS, FeP, AIP, chalcogenide and others), Mineral Ceramic/Simple Oxides, Metal Matrix

Suitable for mixed wastes; Can be used as a heat source in other equipment (e.g fluidized bed)

Limited to small scale; Microwaves need to be guided

Processing Technology

Processing Mode

Treatment and Waste Stream Scale

Wasteforms Produced

Advantages

Disadvantages

Joule Heated Melter (JHM)

Continuous

Large

Boro Silicate Glass, Other Glasses (LaBS, FeP, AIP, chalcogenide and others)

Proven technology; typically operates with a "cold cap" to minimize volatility of species of concern

Electrode and refractory erosion may be a problem; solubility control of certain species

Advanced Joule Heater Melter (AJHM)

Continuous

Large

Boro Silicate Glass, Glass Ceramic Materials, Other Glasses (LaBS, FeP, AIP, chalcogenide and others)

Increases throughput and melt rate compared to JHM

Operates with minimal or no "cold cap" with associated increases in volatility of species of concern

Cold Crucible Induction Melter

(CCIM)

Continuous

Large

Boro Silicate Glass, Glass Ceramic Materials, Other Glasses (LaBS, FeP, AlP, chalcogenide and others), crystalline ceramics/simple oxides, Metal Matrix

Allows the processing of corrosive glasses; No refractories; No metallic or oxide electrodes;water cooled; self-cleaning; high purity,can be stirred if needed; increases capacity compared to JHM and AJHM;can operate at higher temperatures than JHM and AJHM; operates with a "cold cap" to minimize volatility

Higher temperature operation can increase volatilization of species of concern but "cold cap" coverage minimizes these impacts

ПОДПИСИ К РИСУНКАМ

Fig. 1. Schematic representation a of multiple barrier system. THMC stands for thermal, hydraulic, mechanical, and chemical.

Fig. 2. A generic vitrified HLW waste disposal system, after [15]

Fig. 3. Phase composition of nuclear wasteforms, after [10]

Fig. 4. Range of thermal technologies under development for difficult waste streams. The products of the vitrification processes for such waste streams are often GCMs.

Fig. 5. Non-thermal technologies under development for wastes destined for geological disposal which are reactive in Ordinary Portland Cement (OPC) systems.

On selection of matrix (wasteform) material for higher activity nuclear waste immobilisation

C. M. Jantzen1, M. I. Ojovan2,*

1Savannah River National Laboratory, USA, Aiken, South Carolina 29803, Wood Valley Dr. 3922

2Imperial College London, UK, London, SW7 2AZ, South Kensington Campus, Exhibition Road

Figure 1. Schematic representation a of multiple barrier system. THMC stands for thermal, hydraulic, mechanical, and chemical.

Figure 2. A generic vitrified HLW waste disposal system, after [15]

Figure 3. Phase composition of nuclear wasteforms, after [10]

Figure 4. Range of thermal technologies under development for difficult waste streams. The products of the vitrification processes for such waste streams are often GCMs.

Figure 5. Non-thermal technologies under development for wastes destined for geological disposal which are reactive in Ordinary Portland Cement (OPC) systems.

1

Geologic SystemUnderlying StrataGeologic Formation

Conceptual High LevelWaste Repository -CanisterAssumed to Last 1000 years

OverlyingStrataBackfillGeologicFormationOverpackContainerWaste Form(glass)