supporting technical report - unene

60
Supporting Technical Report REVIEW AND GAP ANALYSIS OF THE CORROSION OF COPPER CONTAINERS UNDER UNSATURATED CONDITIONS Report No: 06819-REP-01300-10124-R00 December 2006 For further information contact: Ontario Power Generation, Nuclear Waste Management Division 700 University Avenue, Toronto, Ontario, Canada, M5G 1X6 Fax: (416) 592-7336 Fraser King Integrity Corrosion Consulting Ltd.

Upload: others

Post on 07-Dec-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Supporting Technical Report - UNENE

Supporting Technical Report

REVIEW AND GAP ANALYSIS OF THE CORROSION OF COPPER CONTAINERS UNDER

UNSATURATED CONDITIONS

Report No: 06819-REP-01300-10124-R00

December 2006

For further information contact: Ontario Power Generation, Nuclear Waste Management Division 700 University Avenue, Toronto, Ontario, Canada, M5G 1X6

Fax: (416) 592-7336

Fraser King Integrity Corrosion Consulting Ltd.

Page 2: Supporting Technical Report - UNENE

- ii -

Disclaimer of Liability: Opinions and conclusions contained in this report are those of the author(s) and do not necessarily coincide with those of Ontario Power Generation.

Page 3: Supporting Technical Report - UNENE

- iii -

ABSTRACT

Report Number: 06819-REP-01300-10124-R00 Title: REVIEW AND GAP ANALYSIS OF THE CORROSION OF COPPER

CONTAINERS UNDER UNSATURATED CONDITIONS Author(s): Fraser King Company: Integrity Corrosion Consulting Ltd. Date: December 2006 Abstract An assessment of the likely evolution of the near-field environmental conditions and corrosion behaviour of copper used fuel containers (UFC) during the unsaturated phase is described. The unsaturated phase can be conveniently divided into two periods; the period of increasing container temperatures resulting in surface dry-out and the subsequent cooling period when the surface re-wets. The environment on the container surface is determined by the evaporation of surface moisture and the efflorescence (precipitation) of salts during the heating cycle, and the deliquescence (dissolution) of these salts during the cooling phase. The deliquescence behaviour, in particular, is important as it determines the nature of corrosive solutions that might form during cool-down as well as the time-of-wetness. Wetting of the surface will occur first as droplets formed by the deliquescence of isolated salt crystals but will eventually spread over the entire surface as the relative humidity increases. Similar processes will occur in the near-field sealing materials resulting in an evolution of the pore-water chemistry. There is evidence that chloride and sulphate salts will be re-distributed during the thermal cycle and concentrate close to the container surface. Desiccation of the buffer and backfill materials will also affect the diffusivity of gaseous and dissolved species in the near field. Corrosion of the container during the unsaturated phase will take the form of localized or under-deposit attack. Localized corrosion will result from the spatial separation of anodic and cathodic sites in water droplets that first form on the container surface during the cooling cycle. Cathodic processes will be limited to the periphery of the water droplets or to surrounding microdroplets, areas for which the rate of supply of O2 is fast. Anodic processes will occur in the centre of water droplets which will remain relatively anoxic. However, as these water droplets spread out with increasing relative humidity, the cathodic:anodic surface area ratio will decrease and attack will become more uniform. Nevertheless, the surface of the container is likely to have undergone some surface roughening by corrosion by the end of the unsaturated phase. Other forms of corrosion, including stress corrosion cracking and microbiologically influenced corrosion, are considered to be unlikely during the unsaturated phase. Stress corrosion cracking requires the presence of residual tensile stress and a sufficient concentration of SCC agent. Microbial activity close to the container during the unsaturated phase is unlikely due to the lack of water. The rate of dry air oxidation is slow at temperatures <100oC and the extent of such attack is predicted to be <0.1 µm. An assessment of the possible depth of attack on a copper UFC during the unsaturated period is given. The mean depth of penetration is predicted to be of the order of 100-200 µm, with a maximum penetration of 150-300 µm. It is possible that the majority of the initially trapped O2 will be consumed prior to the establishment of saturated conditions. Gaps in the current understanding of the evolution of environmental conditions and of the corrosion behaviour of copper UFC during the unsaturated phase are identified. The conclusions of this study apply generally to repositories in both granitic and sedimentary host rocks, with the major differences being the higher salinity and possible longer saturation times of the latter.

Page 4: Supporting Technical Report - UNENE

- iv -

Page 5: Supporting Technical Report - UNENE

- v -

CONTENTS

Page

ABSTRACT ...............................................................................................................................iii

1. INTRODUCTION...................................................................................................................1

2. EVOLUTION OF THE ENVIRONMENT DURING THE UNSATURATED PHASE.................2

2.1 OVERVIEW .........................................................................................................2 2.2 EVOLUTION OF ENVIRONMENTAL CONDITIONS ...........................................3 2.2.1 Environment at the Container Surface .................................................................3 2.2.1.1 Evaporation and Efflorescence ............................................................................3 2.2.1.2 Deliquescence and Stability of Surface Films ......................................................7 2.2.1.3 Physical Nature of Moisture on Container Surface...............................................8 2.2.2 Environment in the Near-field Sealing Materials.................................................11 2.2.2.1 Properties of Clay-based Sealing Materials .......................................................11 2.2.2.2 Effect of Unsaturated Conditions on the Transport of Solutes and Gaseous

Species..............................................................................................................14 2.2.2.3 Evolution of the Pore-water Chemistry...............................................................16 2.3 EVIDENCE FROM LARGE-SCALE TESTS ......................................................17 2.3.1 Moisture Content Profiles...................................................................................17 2.3.2 Pore-water Chemistry ........................................................................................18

3. CORROSION BEHAVIOUR OF COPPER UNDER UNSATURATED CONDITIONS..........19

3.1 ATMOSPHERIC CORROSION OF COPPER....................................................19 3.1.1 Overview of the Mechanism of Atmospheric Corrosion......................................19 3.1.2 Studies and Models for Architectural and Electronic Components .....................21 3.1.3 Other Cases of Atmospheric Corrosion of Copper .............................................25 3.2 EVIDENCE FROM NUCLEAR WASTE MANAGEMENT STUDIES...................26 3.2.1 Laboratory Tests................................................................................................26 3.2.2 Large-scale Tests ..............................................................................................29

4. IMPLICATIONS FOR COPPER CONTAINERS IN A DGR .................................................29

4.1 EXPECTED EVOLUTION OF ENVIRONMENTAL CONDITIONS .....................29 4.1.1 Qualitative Description of the Repository Evolution During the Unsaturated

Phase ................................................................................................................29 4.1.2 Semi-quantitative Description of the Repository Evolution During the Unsaturated

Phase ................................................................................................................33 4.2 EXPECTED EVOLUTION OF CORROSION BEHAVIOUR OF COPPER

CONTAINERS ...................................................................................................37 4.2.1 Forms of Corrosion ............................................................................................37 4.2.1.1 Uniform Corrosion..............................................................................................38 4.2.1.2 Localized (Under-deposit) Corrosion..................................................................38 4.2.1.3 Stress Corrosion Cracking .................................................................................38 4.2.1.4 Microbiologically Influenced Corrosion ...............................................................39 4.2.1.5 Oxidation ...........................................................................................................39 4.2.2 Evolution of Corrosion Behaviour During Unsaturated Phase ............................39 4.2.3 Estimate of Corrosion Damage During Saturated Phase ...................................42

Page 6: Supporting Technical Report - UNENE

- vi -

4.3 EFFECT ON THE NATURE OF THE HOST ROCK...........................................42

5. GAP ANALYSIS..................................................................................................................43

6. CONCLUSIONS..................................................................................................................45

REFERENCES .........................................................................................................................46

Page 7: Supporting Technical Report - UNENE

- vii -

LIST OF FIGURES

Page Figure 1: Schematic Illustration of the Processes of Evaporation and Deliquescence.................3 Figure 2: Illustration of the Chemical Divide Theory ....................................................................4 Figure 3: Predicted Changes in Solution Composition Upon Evaporation of a Ca-Na-Cl-based

Natural Water Using the EQ3/6 Code .................................................................................5 Figure 4: Ternary Diagram Illustrating the Calcium Carbonate and Calcium Sulphate Chemical

Divides for the Evaporation of Dilute Groundwaters ............................................................6 Figure 5: Temperature Dependence of the Deliquescence Relative Humidity for Various Mineral

Assemblages.......................................................................................................................7 Figure 6: Schematic Illustration of the Action of Capillary Forces Within a Porous Dust Layer

Preventing Contact of the Deliquescent Brine Solution with the Waste Package Surface in the Proposed Yucca Mountain Repository...........................................................................9

Figure 7: Photograph of Water Droplets on a Surface Illustrating the Discontinuous Nature of the Wetted Surface of a Used Fuel Container During the Unsaturated Phase ..................10

Figure 8: Predicted Dependence of the Water Activity and Percentage Relative Humidity on the Degree of Saturation for Gapfill, Dense and 50:50 Buffer and Dense and Light Backfill Based on Experimental Suction Potential Measurements .................................................12

Figure 9: Impurities from Avonlea Bentonite Adhering to Copper Surface Following Exposure to Compacted 50:50 Bentonite:Sand Buffer Material.............................................................13

Figure 10: Dependence of the Ratio of the Effective Solute Diffusion Coefficient (De) to that in Bulk Solution (D0) on the Volumetric Water Content of Two Unsaturated Soils .................14

Figure 11: Dependence of the Effective Diffusion Coefficient of Oxygen in Soil on the Degree of Saturation..........................................................................................................................15

Figure 12: Vertical Section of Water Contents Transverse to the Axis of the Disposal Room and Mid-height Horizontal Section Measured During Decommissioning of the Buffer-Container Experiment ........................................................................................................................18

Figure 13: Proposed Mechanism for the Distribution of Anodic and Cathodic Reactions for Micro- and Large Droplets of Water ..................................................................................20

Figure 14: Photomicrograph Illustrating Localized Corrosion Due to Ant’s Nest Corrosion .......22 Figure 15: Conceptual Model for the Atmospheric Corrosion of Copper Contaminated by NaCl

in the Presence of Oxygen and Carbon Dioxide................................................................24 Figure 16: Surface Profile of Corroded Copper Coupon Following Exposure to Groundwater-

Saturated Compacted Buffer Material at 50oC for 733 d....................................................27 Figure 17: Proposed Mechanism for the Under-deposit Corrosion of Copper in Compacted

Buffer Material...................................................................................................................28 Figure 18: Qualitative Description of the Evolution of the Environment at the UFC Surface

During the Unsaturated Phase ..........................................................................................30 Figure 19: Qualitative Description of the Evolution of the Environment in the Near Field During

the Unsaturated Phase......................................................................................................32 Figure 20: Assumed Time-dependent Saturation of Inner and Outer Buffer Materials in a Deep

Geologic Repository Used by King and Kolar (2006).........................................................34 Figure 21: Predicted Time Dependence of the Container Surface Temperature and Interfacial

Relative Humidity During the Unsaturated Phase..............................................................35 Figure 22: Predicted Relative Humidity-Temperature Relationship at the Container Surface....36 Figure 23: Comparison of Container Surface Relative Humidity-Temperature Conditions with

Deliquescence Behaviour of Various Salt Systems ...........................................................36

Page 8: Supporting Technical Report - UNENE

- viii -

Figure 24: Qualitative Description of the Evolution of the Corrosion Behaviour of a Copper Used Fuel Container During the Unsaturated Phase ..................................................................40

Page 9: Supporting Technical Report - UNENE

- 1 -

1. INTRODUCTION Ontario Power Generation (OPG) and the Nuclear Waste Management Organization (NWMO) have studied approaches for the management of Canada’s nuclear fuel waste and several approaches include the emplacement and isolation of the used fuel in corrosion-resistant containers in a deep geologic repository (DGR) in a stable host-rock formation (NWMO 2005, Russell and Simmons 2003). Copper is being considered as a possible used fuel container (UFC) material (Maak 1999). The corrosion behaviour of copper in DGR environments has been the subject of much work over the past 20 years, both in Canada and internationally (King et al. 2001, 2002a). Although much is known about the corrosion of copper UFC under saturated conditions, the corrosion behaviour under unsaturated conditions is less well understood. In particular, the significance of the relatively short unsaturated transient period during repository evolution is not known and this period is often ignored in models for the prediction of UFC lifetimes. As with any assessment of the corrosion behaviour of materials, it is necessary to understand the nature of the corrosive environment during the unsaturated phase and how it evolves over time. Even though the maximum UFC surface temperature is expected to be below the boiling point of water, the presence of hygroscopic clay-based sealing materials and the thermal gradient produced by the heat-generating used fuel will result in an initial dry-out period. Water will be driven away from the container and any surface moisture films will dry out, leading to the precipitation of salts. As the magnitude of the thermal gradient diminishes and as groundwater saturates the sealing materials, the surface of the container will become wet and corrosion is possible. Eventually, the repository will be completely saturated by incoming groundwater. Although relatively little work has been performed on the corrosion behaviour of copper UFC during the unsaturated phase, much is known about the atmospheric corrosion of copper alloys. Atmospheric corrosion can occur at relative humidity (RH) values less than 100% and is affected by the presence of surface salts. Therefore, there may be information in the literature on atmospheric corrosion of use in determining the nature and extent of corrosion of copper UFC during the unsaturated phase of the evolution of the DGR environment. The purpose of this report is to review the available information on the expected environment and consequent corrosion of copper UFC during the unsaturated phase. The review is separated into a discussion of the evolution of environmental conditions during the unsaturated phase (Section 2) and existing corrosion information on the atmospheric corrosion of copper and the small amount of work that has been done on copper UFC under unsaturated conditions (Section 3). Based on this review, the likely evolution of the environment and of the corrosion of a copper UFC in a Canadian DGR is discussed (Section 4). A gap analysis is also provided, identifying areas of uncertainty in the current understanding (Section 5).

Page 10: Supporting Technical Report - UNENE

- 2 -

2. EVOLUTION OF THE ENVIRONMENT DURING THE UNSATURATED PHASE

2.1 OVERVIEW The evolution of the environment during the unsaturated phase involves a complex set of coupled thermal (T), hydraulic (H), mechanical (M), and chemical (C) processes. Although it is possible to describe the nature of each of the important T, H, M, and C processes, a detailed THMC model of the evolution of the system is not yet available. Therefore, the description of the evolution of the system given here is based on a qualitative discussion of the processes involved supported by the results of simpler numerical modelling, where appropriate. For the current purposes, the unsaturated phase is defined as the period from the emplacement of the containers until full saturation of the buffer and backfill sealing materials. In general, the unsaturated period can be divided into two phases: an initial dry-out phase followed by a subsequent re-wetting period. The dry-out phase is characterized by the following processes:

1. increasing temperature 2. the transport of water away from the container surface 3. evaporative concentration of salts in moisture on the container surface and in the pores

of the buffer and backfill sealing materials 4. the precipitation of salts (efflorescence) 5. the cessation of corrosion as the surface of the container dries

The subsequent re-wetting stage is characterized by:

1. decreasing UFC surface temperature 2. the transport of water towards the container surface 3. wetting of the UFC surface and sealing materials close to the container 4. deliquescence of salts in the buffer and backfill and on the UFC surface 5. possible acid degassing and destabilization of concentrated brine solutions on the

container, followed by subsequent drying 6. the re-establishment of aqueous corrosion processes

Corrosion (oxidation) of the container following dry-out of surface moisture and prior to re-wetting is not considered significant as the rates of oxidation of copper at the temperatures in question (<100oC) are minimal (King and Kolar 1997). Two key processes occurring during the unsaturated phase are evaporation and deliquescence (Figure 1). Evaporation leads to the concentration of solutes in the aqueous phase and the eventual precipitation of one or more solid salts, a process also referred to as efflorescence. Deliquescence is the reverse process, whereby salts progressively dissolve as water is absorbed from the atmosphere as the RH increases.

Page 11: Supporting Technical Report - UNENE

- 3 -

Figure 1: Schematic Illustration of the Processes of Evaporation and Deliquescence (BSC 2004a).

2.2 EVOLUTION OF ENVIRONMENTAL CONDITIONS

2.2.1 Environment at the Container Surface

2.2.1.1 Evaporation and Efflorescence Upon emplacement, the surface of the container is likely to be wetted by a thin layer of moisture. Water may be adsorbed on the surface as the container is transported to the emplacement room through the repository, in which the atmosphere is likely to be close to 100% RH. Alternatively, moisture from the partially saturated bentonite-based sealing materials may contact the container surface. Formation of an adsorbed water layer will be enhanced by the presence of surface contaminants that will lower the dew point of water. These contaminants may comprise organic carbon-based materials, from lubricants, exhaust, or other hydrocarbons used during repository construction, or inorganic salts, present either as aerosols in the atmosphere (possibly formed from the spalling of efflorescent salts on the repository walls) or as impurities in the clay-based sealing materials contacting the UFC surface. In an industrial mining environment, such as a deep geologic repository, it will be impossible to avoid some degree of surface contamination, regardless of the processes put in place to avoid such contamination.

Page 12: Supporting Technical Report - UNENE

- 4 -

Figure 2: Illustration of the Chemical Divide Theory (Drever 1997).

The thickness of any surface water layer present initially on the container surface will be limited, however. The thickness of adsorbed water layers increases with increasing RH, reaching a thickness of a few nm at 100% RH, equivalent to a few hundred water layers (Lee and Staehle 1997, Nagano et al. 1998). However, water films as thin as 1-10 water layers are sufficient to support electrochemical reactions and, hence, corrosion (Leygraf and Graedel 2000). Following emplacement of the container, any surface moisture present will evaporate as the temperature increases and moisture is driven away from the surface by the thermal gradient. Evolution of the chemical composition of the surface moisture film can be predicted based on the chemical divide theory illustrated in Figure 2 (Drever 1997). Simply stated, the chemical divide theory predicts the sequence of precipitation processes during the evolution of a natural water, in this case containing Na+, Ca2+, Mg2+, HCO3

-, SO42-, and Cl- for the example shown. As

the solution evaporates and loses moisture, the first solid to precipitate in natural waters is almost invariably calcite (CaCO3). If the initial solution contains an excess of Ca2+ ions over carbonate, then the resulting brine following calcite precipitation will be free of carbonate. Conversely, for carbonate-rich solution, Ca2+ will be removed from solution following the precipitation of calcite. Further evaporation results in the precipitation of either gypsum (CaSO4⋅2H2O) or of a magnesium silicate depending upon whether the initial solution was Ca2+- or carbonate-rich, respectively. For Ca2+-rich solutions, the resulting brine following gypsum precipitation with either be Cl--dominated (if the SO4

2- content is less than that of Ca2+) or will comprise a mixture of Na+, Mg2+, SO4

2-, and Cl- (if the SO42- content is greater than that of

Ca2+). For carbonate-rich solutions, the resulting brine following the precipitation of magnesium silicate will be either a Na+/Mg2+-dominated SO4

2-/Cl- solution (if the Mg2+ content is greater than that of CO3

2-) or a Na+ brine containing a mixture of CO32-, SO4

2-, and Cl- (if the Mg2+ content is less than that of CO3

2-). Further evaporative cycles can be considered until complete dry out of the solution. Natural analogs can be found in the deserts of California and Nevada for each of the four end points illustrated in Figure 2.

Na, Ca, Mg, HCO3, SO4, Cl

Na, Ca, Mg, SO4, Cl Na, Mg, CO3, SO4, Cl

Na, Ca, Mg, Cl Na, Mg, SO4, Cl Na, Mg, SO4, Cl Na, CO3, SO4, Cl

CaCO3 precipitates

Carbonate species < Ca contentCarbonate removed from solution

Carbonate species > Ca contentCa removed from solution

CO3 species > Mg content

EvaporativeConcentration

CO3 species < Mg content

Mg removed from solution

CO3 removed from solution

SO4 content > Ca content

SO4 content < Ca content

Mg silicate precipitatesCaSO4 precipitates

SO4 removed from solution

Ca removed from solution

Page 13: Supporting Technical Report - UNENE

- 5 -

Figure 3: Predicted Changes in Solution Composition Upon Evaporation of a Ca-Na-Cl-based Natural Water Using the EQ3/6 Code (after BSC 2004a).

Computational codes exist for predicting the consequences of evaporation of waters. The results of a “numerical evaporation” of a precursor Ca-Na-Cl-based solution using the commonly used code EQ3/6 is shown in Figure 3. In this case, because of the presence of highly soluble NO3

- ions in the initial solution, evaporation results in the development of a nitrate-dominated K-Na-NO3-Cl brine. (See BSC 2004a for details of the initial composition and the assumed conditions for the thermodynamic calculations). If the time-dependent %RH is known, it is possible to predict the evolution of the surface water chemistry with time. (The calculations in Figure 3 were performed for a constant temperature of 105oC). A large number of such calculation shave been performed for the U.S. Yucca Mountain Project for a range of natural water compositions (e.g., Alai et al. 2005; BSC 2004a,b). Based on the results of these simulations for various types of groundwater, it is possible to generalize the evaporative behaviour of dilute Na-Ca-HCO3-SO4-Cl groundwaters (Figure 4). Three types of brine are predicted:

1. Ca-Cl brine, with precipitated CaCO3 and CaSO4⋅2H2O, 2. Sulphate-rich brine, with precipitated CaCO3 and CaSO4⋅2H2O, and 3. Na-carbonate brine, with precipitated CaCO3 and Mg silicates

1.E-7

1.E-6

1.E-5

1.E-4

1.E-3

1.E-2

1.E-1

1.E+0

1.E+1

1.E+2

0%10%20%30%40%50%60%70%80%90%100%RH

Mol

ality

3

4

5

6

7

8

9

10

11

12pH

ISAlCCaClFKMgNNaSSiH2O(kg)pH

Page 14: Supporting Technical Report - UNENE

- 6 -

Figure 4: Ternary Diagram Illustrating the Calcium Carbonate and Calcium Sulphate Chemical Divides for the Evaporation of Dilute Groundwaters (Alai et al. 2005).

The complete evaporation of these three brine types could lead to the formation of the following mineral assemblages, respectively:

1. NaCl (halite), MgCl2, CaCl2⋅xH2O, CaCO3, and CaSO4⋅2H2O 2. Na2SO4, MgSO4, CaCO3, and CaSO4⋅2H2O 3. NaCl, NaHCO3 (nahcolite), Na2CO3 (natron), CaCO3, and Mg silicates

Other minerals are also possible depending upon the composition of the initial moisture film. Potassium salts such as sylvite (KCl) may form, depending upon the K+ concentration. Although nitrate ions are not typically found in deep groundwaters, NO3

- may be present as a residue from blasting or as a consequence of the microbially mediated oxidation of nitrite ions (nitrification) (King et al. 2002b, 2004). Because of their high solubility, nitrate ions tend to dominate the brine composition at the final stages of dry out, eventually precipitating as niter (KNO3) or soda niter (NaNO3). It is important to identify what species will precipitate during the evaporative stage, as these minerals will determine the corrosion behaviour of the container during the subsequent re-wetting phase. In general, the different salts in a mineral assemblage formed from the evaporation of a dilute moisture film will re-dissolve at different relative humidities. Consequently, the precipitated mineral assemblage will determine both the composition of the surface liquid film and the time-of-wetness. A possible dry-out scenario for a container in a DGR is described in Section 4.1.

Page 15: Supporting Technical Report - UNENE

- 7 -

Figure 5: Temperature Dependence of the Deliquescence Relative Humidity for Various Mineral Assemblages (after Rard et al. 2005).

2.2.1.2 Deliquescence and Stability of Surface Films The process of re-dissolution of the precipitated mineral assemblage formed during the dry-out phase is known as deliquescence (Figure 1). At the deliquescence point, water is transferred from the vapour phase to form a saturated aqueous phase in equilibrium with one or more precipitated minerals. At the deliquescence point, the thermodynamic water activity in the aqueous phase is the same as the relative humidity (water activity) in the vapour. Deliquescence thus occurs at a specific relative humidity, known as the deliquescence relative humidity (DRH), which is a function of temperature and of the mineral phase. Figure 5 shows the temperature-dependent DRH for a number of mineral assemblages. To illustrate the use of such figures, let us consider the case of a UFC with a maximum surface temperature of 100oC as the %RH slowly increases as the repository saturates with

Page 16: Supporting Technical Report - UNENE

- 8 -

groundwater. The %RH (or time) at which the container surface first wets depends on the nature of the surface deposits. Let us assume that the surface is contaminated by a mixture of NaCl and NaNO3 as a result of the evaporative processes described in the previous section. Based on the data in Figure 5, the surface would first wet when the %RH increases to 55% at 100oC, at which point a small amount of solid NaNO3 would dissolve. (Deliquescence would occur at a slightly higher %RH if the temperature is <100oC, as described by the line labelled “NaCl + NaNO3” in Figure 5). As the %RH continues to increase, the NaNO3 would continue to dissolve and the solution would become more dilute. With increasing RH some NaCl would also dissolve forming a Na-NO3-Cl brine. At some stage all of the NaNO3 would have dissolved and the brine would be in contact only with precipitated NaCl, which would also eventually dissolve as the RH further increased and the aqueous phase became more dilute. The precise sequence of events depends on the RH-T-t behaviour at the container surface. A similar evolution of conditions can be described for other mineral assemblages, including those for salts not illustrated in Figure 5. In all cases, however, it is important to note that deliquescence will occur at a higher %RH (and, therefore, at a higher temperature during the re-saturation phase) for an assemblage of two or more salts than for the individual salts themselves. Certain deliquescent brines are unstable and will undergo acid degassing if formed on the container surface (EPRI 2004). Calcium chloride brines especially will decompose with the evolution of HCl at elevated temperatures, with the formation of CaO or CaCO3 depending upon the nature of the atmosphere. The rate of acid degassing depends on the partial pressure of acid gas (typically HCl or HNO3 for chloride or nitrate brines, respectively), which in turn depends on the solution pH. In open systems, complete acid degassing of CaCl2 brines has been demonstrated, leading to the drying out of the surface (Hailey and Gdowski 2003). However, significant acid degassing is unlikely in a Canadian repository owing to the relatively low container surface temperatures (<100oC). A more-detailed discussion of the possible deliquescence scenario for a copper UFC in a Canadian DGR is given in Section 4.1. 2.2.1.3 Physical Nature of Moisture on Container Surface The physical nature of the moisture on the container surface will change with time. A continuous aqueous phase covering the entire UFC surface is only likely under nearly fully saturated conditions. During much of the unsaturated phase, any water present will be in the form of droplets. There are two major reasons for the formation of water droplets:

1. presence of porous clay-based sealing materials 2. discontinuous distribution of precipitated minerals and other deliquescent impurities

Page 17: Supporting Technical Report - UNENE

- 9 -

Figure 6: Schematic Illustration of the Action of Capillary Forces Within a Porous Dust Layer Preventing Contact of the Deliquescent Brine Solution with the Waste Package Surface in the Proposed Yucca Mountain Repository (Bryan 2006).

Like all porous materials, the clay-based sealing materials contacting the container surface will develop a capillary pressure within the pores of the gap backfill material. Depending upon the pore-size distribution, a certain fraction of the water near the container surface will preferentially condense within the pores of the gapfill rather than on the container surface (Figure 6). Wetting occurs preferentially close to points of contact between soil (dust) particles and the container surface. Although the illustration in Figure 6 was developed for the case of dust piles on waste packages in the proposed Yucca Mountain repository, a similar situation would occur for Cu UFC surrounded by bentonite-based gapfill material.

Page 18: Supporting Technical Report - UNENE

- 10 -

Figure 7: Photograph of Water Droplets on a Surface Illustrating the Discontinuous Nature of the Wetted Surface of a Used Fuel Container During the Unsaturated Phase (http://www.pbase.com/robertwhite/image/44870727 ©).

Even in the absence of particulate material, the initial wetting of metallic surfaces occurs as droplets (Bian et al. 2005, Xu et al. 2002, Zhang et al. 2005). The discontinuous distribution of deliquescent salts and other contaminants on the container surface will cause some areas of the surface to be wetted, whilst other areas are dry. Figure 7 shows a photograph of a polished surface wetted by droplets and illustrates the possible pattern of wetting of the container surface during much of the unsaturated phase. Regardless of the mechanism of droplet formation, the presence of a discontinuous water film has a number of consequences for the corrosion behaviour of the container. First, the precipitation of different salts over the surface will result in an inhomogeneous chemical environment. For example, let us assume CaCO3, CaSO4⋅2H2O, NaCl, and NaHCO3 precipitate on the container surface due to the evaporation of an initial dilute aqueous phase. If the precipitated minerals are not in intimate contact, then they will deliquesce independently at their respective DRH’s. Thus, prior to the development of a continuous liquid film over the entire surface, there will be patches of the respective dissolved electrolytes on different areas of the surface and at different times. Second, it is likely that there will be a certain degree of localized corrosion during the unsaturated phase due to the physical separation of anodic and cathodic reactions. The

Page 19: Supporting Technical Report - UNENE

- 11 -

environment will be aerobic for much, if not all, of the unsaturated phase and O2 (or Cu(II) formed by the oxidation of Cu(I) by O2) will be the primary oxidant. As a consequence, the periphery of the liquid droplets illustrated in Figure 7 will be more aerobic than the centre as a result of more effective O2 transport. Therefore, the cathodic reaction will tend to occur around the outside of the droplet, whilst the anodic reaction will occur in the relatively anoxic centre of the drop. The consequences for these spatial variations in the chemical composition and electrochemical activity over the container surface will be discussed in more detail in Section 4.2.

2.2.2 Environment in the Near-field Sealing Materials The near-field sealing materials are defined here as the gapfill, inner buffer (100% bentonite), outer buffer (50:50 bentonite:sand), and the light and dense backfill materials (Russell and Simmons 2003). Although all of these materials will affect the near-field environment to some degree, attention is focussed here on the gapfill and inner and outer buffer materials in closest proximity to the UFC. A number of properties of these materials will affect the corrosion behaviour of the containers during the unsaturated phase. First, the presence of hygroscopic clay-based materials will affect the near-field moisture distribution which will determine in part the time at which the container surface is wetted. Naturally occurring clays also contain a number of mineral impurities which may contaminate the container surface and affect the efflorescence and deliquescence characteristics, discussed in the previous section. Second, the diffusive mass transport of solutes and gaseous species to and from the container surface are affected by the degree of saturation of the sealing materials. Third, the sealing materials affect the composition of the aqueous phase contacting the container as a result of the conditioning of incoming groundwater by ion-exchange and dissolution/precipitation processes.

2.2.2.1 Properties of Clay-based Sealing Materials The hygroscopic nature of montmorillonite-based clays, such as bentonite, determines the relative humidity in the near field during the unsaturated phase. The relative humidity is equivalent to the water activity (aW). In unsaturated systems, the suction potential (Sr) is related to the water potential (φ), which in turn is related to aW

WW

r alnVRTS =φ=− (1)

where VW is the partial molar volume of water and R and T are the gas constant and absolute temperature, respectively. Experimental suction potential data as a function of moisture content for highly compacted bentonite (100% bentonite compacted to a dry density of 1.67 Mg⋅m-3) and 50:50 bentonite:sand have been analyzed by Baumgartner (2006). The suction potential (at constant volume) is found to be a function of the degree of saturation (S) and the effective montmorillonite dry density (EMDD):

Page 20: Supporting Technical Report - UNENE

- 12 -

Figure 8: Predicted Dependence of the Water Activity and Percentage Relative Humidity on the Degree of Saturation for Gapfill, Dense and 50:50 Buffer and Dense and Light Backfill Based on Experimental Suction Potential Measurements (from Baumgartner 2006).

α−

=−− n/1m/11

r]1)S[(S (2)

where α, m, and n are fitting parameters that are functions of the EMDD. (Note: Equation (2) has been corrected from the version given as Equation (15b) by Baumgartner (2006)). The EMDD is the dry density of the soil normalized by the montmorillonite content, and is found to be a useful parameter for characterizing the effect of the swelling clay content on the suction potential and other properties of clay-based soils (Baumgartner 2006). Russell and Simmons (2003) give the EMDD for the gapfill, dense and 50:50 buffer materials and the dense and light backfill materials currently being considered for use in a Canadian DGR. Based on a montmorillonite content of 80% for Avonlea bentonite (Baumgartner 2000), Equations (1) and (2) can be combined to predict the dependence of aW on saturation for each of these sealing materials. Figure 8 shows the predicted variation of aW (or RH) against the degree of saturation for the various clay-based sealing materials based on the measurements of the suction potential as a

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1Degree of saturation

Wat

er a

ctiv

ity

0

20

40

60

80

100

% R

elat

ive

hum

idity

Gap fill50:50 bufferDense backfillLight backfillDense buffer

Page 21: Supporting Technical Report - UNENE

- 13 -

function of moisture content. Desiccation of the dense and 50:50 buffer materials under the action of the thermal gradient will significantly lower the %RH in the near field. In turn, if it is assumed that the %RH at the container surface is similar to that of the gapfill or dense buffer, the data in Figure 8 can be used to estimate when the UFC may be wetted during the unsaturated phase. A more-detailed discussion of the time-dependent %RH in the near field is given in Section 4.1. As well as affecting the near-field %RH, the clay-based sealing materials are also a source of contaminants in the form of mineral impurities. Avonlea bentonite is known to contain ~2 wt.% gypsum (CaSO4⋅2H2O), ~1 wt.% calcite (CaCO3), and ~0.5 wt.% organic carbon (Baumgartner 2000, Oscarson and Dixon 1989, Quigley 1984). The presence of a chloride-bearing mineral, possibly halite (NaCl), has also been inferred (Oscarson and Dixon 1989). Contact of bentonite-bearing sealing materials with the container surface may lead to contamination by impurities, a process that has been demonstrated experimentally by King et al. (1989). Figure 9 shows two adherent particles on a copper surface following exposure to buffer material initially compacted to ~80% saturation with synthetic saline groundwater and immersed in a bulk groundwater solution at a temperature of 150oC for a period of ~2 weeks. The “golf ball” is a magnetite spherule, another common impurity in Avonlea bentonite, and the “tee” is a gypsum crystal. As noted in the previous section, such surface contaminants will promote wetting of the surface through deliquescence.

Figure 9: Impurities from Avonlea Bentonite Adhering to Copper Surface Following Exposure to Compacted 50:50 Bentonite:Sand Buffer Material (King et al. 1989). The “golf ball” is a magnetite spherule and the “tee” a crystal of CaSO4⋅2H2O (gypsum).

Page 22: Supporting Technical Report - UNENE

- 14 -

2.2.2.2 Effect of Unsaturated Conditions on the Transport of Solutes and Gaseous Species The degree of saturation of the sealing materials will also affect the rates of diffusion of solutes and gaseous species to and from the container surface. Hu and Wang (2003) have reviewed the effect of moisture content on the aqueous-phase diffusivity in a range of unsaturated geological media. Typically, the ratio of the effective diffusion coefficient of a solute (De) in the unsaturated soil to that in bulk solution (D0) is found to be a power-law function of the volumetric water content (θ), given by

n

0

e

DD

αθ= (3)

where α and n are fitting parameters. Figure 10 shows the dependence of this ratio on the volumetric water content for two different soils, indicating the rapid decrease in solute diffusivity as the water content decreases.

Figure 10: Dependence of the Ratio of the Effective Solute Diffusion Coefficient (De) to that in Bulk Solution (D0) on the Volumetric Water Content of Two Unsaturated Soils (after Hu and Wang 2003).

1E-12

1E-09

1E-06

0.001

1

0 0.1 0.2 0.3 0.4 0.5 0.6

Volumetric water content

De/D

0

SandLoam

Page 23: Supporting Technical Report - UNENE

- 15 -

Figure 11: Dependence of the Effective Diffusion Coefficient of Oxygen in Soil on the Degree of Saturation. Soil porosity 0.40, tortuosity factor 0.1.

In contrast, the diffusivity of gases increases with decreasing moisture content. A number of different models have been developed to describe the diffusivity of gases, particularly O2, in unsaturated soils (e.g., Collin and Rasmuson 1988, Weerts et al. 2000). King and Kolar (2000) have used the following expression for modelling the effective diffusion coefficient of O2 (DEFF) as a function of the degree of saturation (S) in the Copper Corrosion Model 0fair

3fEFF SDD)S1(D ετ+ε−τ= (4)

where Dair is the diffusion coefficient of O2 in air and τf and ε are the tortuosity factor and porosity of the soil, respectively. Figure 11 shows the dependence of DEFF on S, indicating a rapid decrease in diffusivity as the soil approaches full saturation. It is apparent from Figures 10 and 11, therefore, that the rates of diffusive transport of species to and from the container surface will change significantly during the unsaturated phase. During the driest period, O2 and other gaseous species will diffuse rapidly through the unsaturated sealing materials, whereas pore-water solutes and dissolved corrosion products will be relatively immobile. As the buffer and backfill become progressively saturated, the situation will reverse. The consequences of these changes in diffusivity of gaseous and dissolved species for the corrosion behaviour of the container will be discussed in more detail in Section 4.2.

0.0000001

0.000001

0.00001

0.0001

0.001

0.01

0.1

1

0 0.2 0.4 0.6 0.8 1

Degree of saturation

Effe

ctiv

e di

ffusi

on c

oeffi

cien

t of O

2 (cm

2 s-1

)

Page 24: Supporting Technical Report - UNENE

- 16 -

2.2.2.3 Evolution of the Pore-water Chemistry A number of studies have been performed to assess the changes to the pore-water chemistry as the near-field sealing materials saturate (Arcos et al. 2000, Bruno et al. 1999, Cuevas et al. 1997, Curti and Wersin 2002, Lemire and Garisto 1989, Luukkonen 2004, McMurry et al. 2003, Nagra 2002, Xie et al. 2006). A number of different processes are involved, including:

1. Dissolution of soluble mineral impurities (NaCl, CaCl2⋅xH2O, CaCO3, CaSO4⋅2H2O) in the water added for compaction of the sealing materials.

2. Ion exchange with the sodium bentonite, resulting in the uptake of Ca2+ and the release of Na+ ions.

3. Precipitation of the least soluble minerals as water is driven away from the container surface by the thermal gradient.

4. Control of the pore-water pH in the range pH ~7-8 by calcite precipitation/dissolution. 5. Re-dissolution of precipitated minerals as the sealing materials re-wet, starting with the

most soluble (most deliquescent) phases. 6. Eventual saturation of the sealing materials.

A common approach to modelling the evolution of the pore-water chemistry is to treat the time-dependent evolution as a series of steady-states in which the original pore water is gradually exchanged with incoming groundwater (Curti and Wersin 2002, Lemire and Garisto 1989). These approaches essentially treat the pore water composition as spatially homogeneous. To account for spatial variation in the pore-water composition, reactive-transport models (Curti and Wersin 2002) or a simplified discretized transport model (Luukkonen 2004) have been applied. However, these models fail to predict the spatial redistribution of solutes in the pore water observed experimentally (Cuevas et al. 1997), and will not be discussed further here. The most relevant experimental evidence for the evolution of the pore-water chemistry under the action of drying and wetting comes from the work of Cuevas et al. (1997). In these tests, a block of partially saturated compacted bentonite was heated on one side and exposed to a source of deionized water on the other side. At the end of the 109-d test, pore water was extracted and analyzed. The two sections of clay closest to the heater were too dry to extract pore water, but the three sections further from the heat source and closer to the water source were sufficiently saturated to yield pore water when the bentonite was subjected to high pressures at the end of the test. Analysis of the extracted pore-water samples indicated a re-distribution of solutes in the pore water, with higher concentrations closer to the heat source. Compared with a reference, unheated pore-water extract, the salinity of the pore-water of heated samples was lower. However, there was a significant gradient in ionic strength, ranging from 0.08 mol⋅dm-3 in the bentonite closest to the water source to 0.2 mol⋅dm-3 in the mid-section of bentonite closer to the heat source. The ionic strength of the unheated reference bentonite was 0.3 mol⋅dm-3. Although pore water could not be extracted from the bentonite closest to the heater, analysis of post-test aqueous extracts of the bentonite suggested Cl- contents up to ten times higher in these samples than in the other heated bentonite samples. Interestingly, the SO4

2- contents of the samples (based on the aqueous extract data) were uniform across the five bentonite segments. It appears, therefore, that Cl- salts were re-distributed during the test with higher concentrations closer to the heater, whereas SO4

2- salts were not re-distributed.

Page 25: Supporting Technical Report - UNENE

- 17 -

Xie et al. (2006) carried out detailed numerical modelling of the experiments performed by Cuevas et al. (1997). Processes included in the reactive-transport modelling included: bentonite swelling, changes in porosity, two-phase flow, ion exchange, dissolution/precipitation, and heat transport. The authors claim good agreement between predicted and measured parameters. In particular, Xie et al. (2006) were able to predict the increase in Cl- content closer to the heated surface. The observation of the re-distribution of mineral impurities in the bentonite under the action of the thermal gradient is an important conclusion from the experimental and modelling studies. Upon re-saturation of the bentonite by the incoming groundwater, higher [Cl-] can be expected near the container surface than if the bentonite had not been thermally treated.

2.3 EVIDENCE FROM LARGE-SCALE TESTS

2.3.1 Moisture Content Profiles The fact that water will be re-distributed under the influence of the thermal gradient from the container has been amply demonstrated by a number of full-scale heater tests. In Canada, the Buffer-Container Experiment (BCE) was conducted at the Underground Research Laboratory between 1991 and 1994 (Graham et al. 1997). A heater simulating the then reference container design was placed in a borehole and surrounded by compacted 50:50 bentonite:sand buffer (initial water content 85-89% saturated). The buffer material was extensively instrumented with thermocouples, psychrometers, pressure transducers, and other instruments with which to monitor the time-dependent changes in moisture content and swelling pressure as the buffer was first heated and then allowed to cool. At the end of the test, the buffer material was carefully excavated and measurements made of the buffer properties, including the distribution of water content. Figure 12 shows the post-test moisture content profiles in vertical and horizontal sections measured at the end of the test. The initial moisture content was 17.8% (85-89% saturation for the density of buffer used in the test). Significant drying of the buffer closest to the heater has occurred, with buffer towards the periphery of the borehole wetter than when emplaced. Water mass-balance calculations suggested that little water had been lost from the borehole during the test and that, instead, the water had simply been re-distributed in the buffer. Similar results showing lower water contents in the dense bentonite closest to the heater have also been reported from the LOT experiments at Äspö (Karnland et al. 2000). Detailed predictions of the time- and spatial-dependence of the moisture content in the various sealing materials of a Canadian DGR are not currently available. Based on a conceptual understanding of the processes leading to the re-distribution of water and the behaviour of the different materials, however, a set of proposed saturation profiles has been proposed and used in various copper corrosion models and performance assessment studies (e.g., King and Kolar 2006). These profiles will be used in Section 4 to predict the possible time-dependence of the corrosion behaviour of copper containers during the unsaturated phase.

Page 26: Supporting Technical Report - UNENE

- 18 -

Figure 12: Vertical Section of Water Contents Transverse to the Axis of the Disposal Room and Mid-height Horizontal Section Measured During Decommissioning of the Buffer-Container Experiment (Graham et al. 1997). The saturated moisture content for the 50:50 buffer material used in the test was ~20.5%.

2.3.2 Pore-water Chemistry The observed re-distribution of soluble salts under the effect of heating reported by Cuevas et al. (1997) discussed above has also be observed in large scale tests. Dixon et al. (1998) measured the concentration of soluble salts in the buffer material from the BCE as a function of distance away from the heater. The concentrations of Na+, Cl-, and particularly SO4

2- were higher nearest the heater, whereas the concentration of carbonate was highest further from the heat source. The concentrations of K+, Ca2+, and Mg2+ showed no specific variation with distance. These results are consistent with the concept that the more soluble Cl- and SO4

2- salts concentrate near the heat source because of their greater tendency to deliquesce at lower %RH. What is still not clear is how these ions are transported given the strong dependency of the diffusivity of dissolved species on water content (Figure 10). Karnland et al. (2000) also report some redistribution of ions during the first year of the LOT experiment, with small increases in the SO4

2- content closest to the heater. McMurry et al. (2003) report a laboratory study by O. Karnland for SKB in which gypsum and calcite were found to redistribute during water percolation under a thermal gradient, although the cited report does not appear on the SKB website.

Page 27: Supporting Technical Report - UNENE

- 19 -

3. CORROSION BEHAVIOUR OF COPPER UNDER UNSATURATED CONDITIONS

3.1 ATMOSPHERIC CORROSION OF COPPER

3.1.1 Overview of the Mechanism of Atmospheric Corrosion Conceptually, the mechanism of the atmospheric corrosion of metals is reasonably well understood (Leygraf and Graedel 2000, Shreir 1976). Adsorption of water molecules from the atmosphere produces a surface water layer thick enough to support electrochemical reactions. Layers thicker than ~3 monolayers possess the same properties as a bulk aqueous phase (Leygraf and Graedel 2000). Initial water deposition is by physisorption and chemisorption, the latter promoted by the interaction between the polar water molecules and hydroxyl groups of air-formed oxides on the surface. As the layer thickens, and the surface corrodes, continued water deposition results from capillary condensation. Surface wetting is promoted by the presence of dust or salt contaminants and by pre-existing corrosion products on the surface. The onset of atmospheric corrosion is often associated with a critical RH of 60-70% that is relatively insensitive to the nature of the substrate (Shreir 1976). As discussed in more detail below, however, the water coverage increases monotonically with increasing %RH and the concept of a “critical” humidity is probably more related to the ability to measure the consequent corrosion than it is a reflection of an RH below which no corrosion occurs. As noted above, the surface coverage by water increases monotonically with increasing %RH. Lee and Staehle (1997) measured the change in mass due to the adsorption of water as a function of %RH and temperature for a number of different metals, including Cu. Monolayer coverage was reported at RH as low as 5% at 25oC and 45oC, although a higher humidity (23% RH) was required to achieve the same coverage at 90oC. Both adsorption and capillary condensation were found to be important processes for water layer formation, with the overall “adsorption” curve described by a Type IV BET isotherm. The adsorption isotherm exhibited more rapid growth of the water film at low (0-10% RH) and high (>80% RH) humidity. This latter increase in water layer thickness at high %RH may be responsible for the observed “critical” RH for atmospheric corrosion. The adsorption/desorption isotherm exhibited hysteresis, with thicker water layers at a given %RH during the desorption cycle (Lee and Staehle 1997). Thus, at a given %RH or aW, the surface of a UFC may be covered by a thicker film during the heating cycle of the thermal transient than during the latter cooling/re-wetting phase. This form of adsorption/desorption hysteresis is commonly observed, and has been reported by others for other materials (Nagano et al. 1998). As discussed earlier, the surface may only be uniformly wetted at high %RH. Prior to the development of a continuous water film, the surface will be covered by water droplets. These droplets will tend to form at discrete salt or dust particles attached to the surface. Xu et al. (2002) propose that the various stages in the development of a continuous water film involve an initial deposition of microdroplets which coalesce into larger droplets as the RH increases, finally forming a continuous water layer at some unspecified %RH. However, Bian et al. (2005) measured the distribution of droplet sizes formed at near-100% RH and found that small droplets could persist for as long as 100 h, suggesting only a slow transition to complete surface coverage.

Page 28: Supporting Technical Report - UNENE

- 20 -

Figure 13: Proposed Mechanism for the Distribution of Anodic and Cathodic Reactions for Micro- and Large Droplets of Water (after Zhang et al. 2005).

As shown in Figure 7, the larger droplets are commonly surrounded by a number of smaller droplets and Zhang et al. (2005) have proposed an important role of these microdroplets in the atmospheric corrosion of metals. According to the proposed mechanism. These microdroplets act as the site of the cathodic reaction, supporting anodic dissolution processes under the main droplet (Figure 13). A critical RH was observed at which water would condense on an individual NaCl crystal to form the centre droplet. Provided the %RH was greater than the equilibrium RH for this droplet (which would depend on the nature and concentration of the solute) microdroplet formation is possible. Since the micro- and main droplets are connected by a thin continuous water layer (which enables ion transport), electrochemical reactions occurring in the microdroplets are coupled to those in the main droplet. Presumably because of the higher surface area:volume ratio (which promotes the supply and prevents the depletion of O2), the cathodic reduction of O2 occurs predominantly in the microdroplets and the periphery of the main droplet. Anodic dissolution of the substrate is focussed at the centre of the main droplet. The heterogeneity of the local environment, therefore, promotes the separation of anodic and cathodic reactions resulting in the possibility of localized corrosion. In addition to the potential for the separation of anodic and cathodic reactions, droplet and thin-film formation also affects other aspects of the corrosion process. Electrolyte concentrations tend to be higher because of the small volumes of water (Dante and Kelly 1993) and saturated solutions could be formed, especially during the initial stages of deliquescence of salt contaminants. In addition, the small volumes of water result in relatively rapid super-saturation by dissolved corrosion products, which in turn promotes the formation of protective surface films. For this reason, atmospheric corrosion rates can be smaller than those in bulk aqueous environment, despite the relatively rapid supply of oxidant in the form of atmospheric O2. Therefore, corrosion reactions are more likely to be anodically limited than cathodically limited in thin water films or in water droplets.

Page 29: Supporting Technical Report - UNENE

- 21 -

3.1.2 Studies and Models for Architectural and Electronic Components Because of the importance of copper in architectural and electronic components, there has been considerable effort to understand the mechanism of the atmospheric corrosion of copper (Leygraf and Graedel 2000). Various studies have been performed in different areas of the world to understand the corrosion behaviour in rural, urban, industrial, and marine environments. The parameters most commonly reported are the nature of the corrosion products, or patina, and the overall corrosion rate. In general, the atmospheric corrosion behaviour of copper is largely determined by the RH and the nature of contaminants present on the surface or in the atmosphere. The most commonly studied contaminants are: sulphur dioxide (SO2), nitrogen dioxide (NO2), ozone (O3), ammonium sulphate ((NH4)2SO4, a common pollutant in rural locations), hydrogen sulphide (H2S), and sodium chloride (NaCl). Other species considered include carbon dioxide (CO2) and acetic acid (CH3CO2H). The atmospheric concentrations of industrial pollutants expected in the repository would be typical of an urban environment (King et al. 2001, 2002a). Of the other contaminants, the effects of NaCl and CO2 are particularly interesting for the corrosion of Cu UFC in unsaturated bentonite. Sulphur dioxide has been reported to act as both an oxidant and as a reducing agent. Chawla and Payer (1990) observed the presence of a mixture of copper oxide and sulphide, implying that the S had been reduced (from +IV to –II) with the concomitant oxidation of Cu. Other workers, however, report either the conversion of SO2 to sulphite (SO3

2-) and the precipitation of copper sulphite (Persson and Leygraf 1990) or its oxidation to sulphate (SO4

2-, +IV to +VI), the latter catalyzed by the presence of either Cu(II) (Chen et al. 2005a) or O3 (Strandberg and Johansson 1997, Zakipour et al. 1995). The presence of H2S has little effect at low RH because the initially formed Cu2O film remains protective (Sharma 1980). However, at higher RH the Cu2O film is less protective resulting in additional corrosion. Nitrogen-containing contaminants, as might be present in a DGR due to microbial activity or as blasting residues) also lead to accelerated atmospheric corrosion and modified surface films. Nitrogen dioxide is oxidized to nitrate and precipitates as a Cu(NO3)2 species on top of an inner Cu2O layer (Persson and Leygraf 1990). Ammonium sulphate, present as an aerosol in rural environments, is likely to be highly deliquescent due to its high solubility. Corrosion products formed include Cu2O and one of several forms of basic cupric sulphate (e.g., antlerite and posnjakite), with the possible conversion of one form to another as the film ages and re-crystallizes (Lobnig et al. 2003, Odnevall and Leygraf 1995, Tidblad and Graedel 1997). In the presence of Cl- contamination, a wide range of Cl-bearing salts has been observed, similar to the range observed in bulk aqueous Cl- solutions (King et al. 2001, 2002a). Species observed include CuO, CuCl, and Cu(OH)3Cl (Chen et al. 2005b, Strandberg and Johansson 1998). The similarity of the corrosion products observed under atmospheric conditions and in bulk solution confirms that the individual corrosion processes in the two systems are essentially identical.

Page 30: Supporting Technical Report - UNENE

- 22 -

Finally, the presence of acetic acid vapour (another possible microbially produced contaminant in a DGR) results in a corrosion product layer comprising Cu2O, Cu(OH)2, and Cu(CH3CO2)2 (Cano et al. 2001). Another form of atmospheric corrosion involving acetic and other organic acids is known as “ant’s nest” or formicary corrosion (Corbett and Elliott 2000; Elliot and Corbett 1999; F. King and L. Yang, unpublished data 2006; Notoya 1991). Ant's nest corrosion is a form of localized corrosion of copper generally associated with cooling and air conditioning systems. It is characterized by a network of interconnected pores which intersect the pipe surface as a small pinhole, barely observable with the naked eye (Figure 14). The interconnected pore network resembles the structure of an ant's nest, hence the name. In addition, corrosion is often associated with the smell of formic acid, hence the alternative term "formicary corrosion." Ant's nest corrosion has been observed in a number of countries, and has been studied in some detail in Japan and the U.S. The phenomenon has been associated with both attack on the internal and external surfaces of copper tubing, in the latter case due to the condensation of a thin moisture film from the atmosphere. Typical applications involve copper tubing in air conditioning and ventilation units or in copper piping used to circulate cooling fluids. Penetration of the tubing is often reported to be quite rapid (of the order of weeks-months), with rates of 0.6-2.4 mm/yr apparent for failed components (Corbett and Elliot 2000). Pre-requisites for corrosion are reported to be oxygen, moisture (in the case of corrosion from the external surface), and a specific corrodent (generally formic or acetic acids).

Figure 14: Photomicrograph Illustrating Localized Corrosion Due to Ant’s Nest Corrosion (Elliot and Corbett 1999).

Page 31: Supporting Technical Report - UNENE

- 23 -

Not only are the constituents of the corrosion products similar under atmospheric and immersed conditions, so too is the film structure. As in bulk solution, corrosion films formed in atmospheric conditions exhibit a duplex or multi-layered structure, comprising an inner layer of Cu2O and outer layers of Cu(II) salts (Leygraf and Graedel 2000, Toyoda et al. 2004). The Cu2O layer forms initially (first few seconds to hours), followed by the subsequent precipitation of the basic Cu(II) salts over periods of weeks to years under atmospheric conditions (Leygraf and Graedel 2000). Corrosion rates are typically of the order of a few µm⋅a-1 (Leygraf and Graedel 2000), with minor variations between the different types of atmosphere: Rural ~0.5 µm⋅a-1 Urban ~1-2 µm⋅a-1 Industrial <2.5 µm⋅a-1 Marine ~1 µm⋅a-1 Corrosion kinetics are typically parabolic in nature and are anodically limited (due to the excess atmospheric O2). Based on a large database of atmospheric corrosion data, Graedel and co-workers have developed a conceptual and mathematical model to account for the atmospheric corrosion behaviour of different metals. The GILDES model separates the various components of the overall corrosion process as follows (Leygraf and Graedel 2000): G gas I interface L liquid D deposition E electrodic S solid The reactions occurring within each component or at each interface can be quantified in kinetic or thermodynamic terms and predictions made about the rate of corrosion and the nature of the deposited corrosion products. Conceptually, this description of the system in terms of its component parts is similar to the mechanistic descriptions that form the basis for the Copper Corrosion Model (King and Kolar 2000). Of the various systems studied, the atmospheric corrosion of Cu in the presence of NaCl and CO2 is perhaps of most relevance to the corrosion of a Cu UFC during the unsaturated phase in the DGR. In a series of papers, Chen et al. (2005a,c,d) proposed a mechanism to describe the effects of NaCl surface contamination on the corrosion of Cu in humid atmospheres containing O2 and CO2. The proposed mechanism is similar to that of Zhang et al. (2005) and is illustrated in Figure 15.

Page 32: Supporting Technical Report - UNENE

- 24 -

Figure 15: Conceptual Model for the Atmospheric Corrosion of Copper Contaminated by NaCl in the Presence of Oxygen and Carbon Dioxide (Chen et al. 2005a,c,d).

As with the mechanism of Zhang et al. (2005), Chen et al. (2005c,d) propose a physical separation of anodic and cathodic reactions. Chen et al. (2005c,d) observed lateral spreading of the central droplet, as opposed to the formation of microdroplets. The anodic reaction occurs in the central droplet, with the supporting cathodic reduction of O2 occurring around the periphery of the droplet and within the secondary film created by the lateral spreading. At low CO2 concentrations (< 5 vppm), the pH in the liquid film remains high and protective CuO/Cu(OH)2 deposits are observed in addition to the underlying Cu2O layer and precipitated basic Cu(II) chloride (Cu2(OH)3Cl). At higher CO2 concentrations (350 vppm), dissolution of CO2 in the droplet maintains a lower pH, preventing the formation of CuO/Cu(OH)2 and leading to the precipitation of Cu2(OH)2CO3 as well as Cu2(OH)3Cl. As a consequence, more extensive corrosion was observed at the higher CO2 concentration. Chen et al. (2005c) used a scanning Kelvin probe to measure the surface work functions under atmospheric corrosion conditions. These Volta potentials can be related to the corrosion potential (ECORR) commonly measured in bulk aqueous solutions. More-negative potentials were observed in the centre of the droplet, indicative of more-active conditions associated with the anodic dissolution reaction. The potential difference between the centre and outside of the droplet was of the order of 150 mV, corresponding to a difference in the rate of anodic dissolution of 1-2 orders of magnitude between these two locations. There was evidence for a gradual change in potential around the periphery and edge of the droplet, consistent with the location of the cathodic reaction in these areas.

Page 33: Supporting Technical Report - UNENE

- 25 -

An interesting conclusion from this work was the effect of contaminant concentration on the corrosion rate (Chen et al. 2005d). More extensive corrosion was observed the lower the surface coverage by NaCl particles. This observation can be rationalized by the fact that the cathodic:anodic surface area ratio is maximized if the anodic sites are widely separated. (Each anodic site corresponds to a single NaCl crystal on the surface). At higher NaCl coverage, the cathodic areas surrounding the central droplets interfere and coalesce, resulting in a lower cathodic:anodic surface area ratio. Consequently, corrosion is both less localized and less severe.

3.1.3 Other Cases of Atmospheric Corrosion of Copper Two other aspects of the atmospheric corrosion of copper deserve attention, namely: the oxidation of Cu in dry air and additional evidence regarding the propensity for the localization of corrosion in thin liquid layers. Most studies of the oxidation of Cu in dry air or O2 have been performed at elevated temperatures (>250oC), in order that measurable oxide thicknesses can be achieved in reasonable time periods (King and Kolar 1997). Various oxide growth laws have been reported (Rönnquist and Fischmeister 1960-61), including: direct logarithmic x = a1log(k1t + 1) (5a) inverse logarithmic 1/x = a2log(k2t + 1) (5b) cubic x3 = k3t (5c) parabolic x2 = k4t (5d) linear x = k5t (5e) where x is the film thickness, t is the time, and the a and k are constants in time but depend on temperature and oxygen concentration. The governing growth law typically changes as the film thickens. Gdowski and Bullen (1988) report that oxidation kinetics progress from logarithmic through cubic and parabolic rate laws with increasing thickness and/or temperature. The composition of the oxide depends on a number of factors. Thicker films tend to consist of an inner layer of Cu2O and an outer layer of CuO. In general, Cu2O is observed at T < 250oC, even though CuO is more thermodynamically stable (Rönnquist and Fischmeister 1960-61). The study of Cu oxidation at temperatures below 150oC is complicated by slow oxidation kinetics and the disproportionate influence of factors such as surface finish, cold work and impurity content of the air (or oxygen atmosphere). Roy and Sircar (1981) report logarithmic oxidation kinetics in dry air at temperatures between 75oC and 100oC. In contrast, Pinnel et al. (1979) observed parabolic kinetics for temperatures between 50oC and 150oC for periods up to 1000 h. King and Kolar (1997) estimated the maximum extent of oxide formation for a Cu UFC due to dry air oxidation based on the kinetic expressions of Roy and Sircar (1981) and Pinnel et al. (1979) and a characteristic temperature profile for a container in a DGR. The predicted film thicknesses appear to reach an effective limiting film thickness of ~100 nm for logarithmic growth and ~140 nm for parabolic kinetics. A limiting film thickness of 140 nm would correspond to a wall penetration of 84 nm for a Cu2O film and 80 nm for a CuO film. The amount of O2 consumed would correspond to 2.9 x 10-7 mol⋅cm-2 and 5.6 x 10-7 mol⋅cm-2 for Cu2O and CuO films, respectively, <0.1% of the total available of 6.2 x 10-4 mol⋅cm-2 O2 (for the

Page 34: Supporting Technical Report - UNENE

- 26 -

in-room emplacement configuration). Consequently, the extent of damage and of O2 consumption due to oxidation is minimal. As discussed above, the spatial heterogeneity of water droplets on the surface may induce localization of corrosion during the unsaturated phase. In addition, the possibility of highly concentrated electrolytes in thin surface films may affect the propensity for localized corrosion. King et al. (2001, 2002a) have reviewed the evidence for localized corrosion in concentrated aqueous electrolytes. In addition, King (2002) considered the effects of elevated pH on the pitting of Cu in Cl- environments. The likelihood for local film breakdown under saturated conditions can be estimated by comparing the value of ECORR to the film breakdown or pitting potential (Eb), the condition for film breakdown being ECORR ≥ Eb (6) Based on this comparison, it was concluded that pitting is unlikely to initiate via a classical pitting mechanism in repository environments. The predicted value of ECORR for a Cu UFC in the repository was found to be generally several hundred mV more negative than Eb, implying that film breakdown would not occur. Increasing pH, as might occur in a thin electrolyte layer, renders the surface less susceptible to film breakdown as the magnitude of the difference between ECORR and Eb increases with pH (King 2002). However, more-recent predictions of the ECORR during the early unsaturated period in the DGR evolution suggest more-positive ECORR values (-0.15 VSCE to –0.05 VSCE), raising the possibility of some early localized corrosion (King and Kolar 2006). Measurements of the value of ECORR in thin film solutions suggest that the conclusions above in saturated environments apply equally under unsaturated conditions. Zhong (2002) used a wire beam electrode to measure the ECORR of a Cu electrode in a thin (8.5-µm-thick) film of aerated electrolyte (5 wt.% NaCl, ~0.9 mol⋅dm-3). Using this technique it is possible to measure the spatial distribution of ECORR values over a number of sensors (121 in this case). At short times (of the order of a few hours), ECORR in the thin film was up to 50 mV more positive than the value measured in bulk solution. However, within ~7 h the two values were similar, with a mean ECORR of –0.26 VSCE to –0.27 VSCE. These values are more negative than those predicted for a Cu UFC in the early stages of the evolution of a DGR (King and Kolar 2006).

3.2 EVIDENCE FROM NUCLEAR WASTE MANAGEMENT STUDIES

3.2.1 Laboratory Tests No specific, detailed, study of the corrosion of Cu under simulated unsaturated DGR conditions has been performed. However, the early work of King and co-workers (King et al. 1992, 1997; Litke et al. 1992) included Cu corrosion tests in initially unsaturated 50:50 bentonite:sand (King et al. 1992, Litke et al. 1992) or in 100% bentonite with initial moisture contents of 60% or 80% of the saturated value. Although the buffer material inevitably saturated during the course of the exposure, the Cu coupons were exposed to unsaturated conditions for at least part of the test period.

Page 35: Supporting Technical Report - UNENE

- 27 -

Figure 16: Surface Profile of Corroded Copper Coupon Following Exposure to Groundwater-Saturated Compacted Buffer Material at 50oC for 733 d (Litke et al. 1992).

In these experiments, Cu coupons were exposed to compacted buffer material wetted with synthetic groundwater solutions and exposed to aerated conditions for periods of up to 18 months. At the end of the exposure period, the coupons were separated from the buffer material and examined to determine the morphology of corrosion and the total mass loss. The buffer material was sectioned and the copper and moisture content of each slice determined. King et al. (1997) performed three experiments in compacted 100% bentonite (effective clay dry density 1.2 Mg⋅m-3) to determine the effect of the initial moisture content on the corrosion behaviour. Tests were performed at 60%, 80%, and ~100% saturation at a temperature of 95oC. The corrosion rate was higher in unsaturated bentonite, although the highest rate was observed at 80% saturation (22 µm⋅a-1), compared with rates of 12 µm⋅a-1 and 8 µm⋅a-1 for 60% and 100% saturation, respectively. Higher rates in unsaturated bentonite were explained in terms of faster O2 transport, which typically is the rate-determining process in these systems (King et al. 1997). At 60% saturation there may have been insufficient moisture to sustain rapid corrosion and/or the limited amount of water resulted in rapid super-saturation and the precipitation of protective corrosion products. Although the nature of the corrosion products and the surface topography were not examined in detail in these latter tests, the appearance of the coupons was similar to that reported by King et al. (1992) and Litke et al. (1992). These earlier tests were performed in 50:50 bentonite:sand buffer initially saturated with synthetic groundwater at ~80% saturation. The surface profile of a coupon exposed for a period of 733 d at a temperature of 50oC is shown in Figure 16. Tests at 50oC saturated with synthetic groundwater more slowly than those at either 100oC or 150oC, so that the buffer material would have been unsaturated for much of the duration of the test.

Page 36: Supporting Technical Report - UNENE

- 28 -

Figure 17: Proposed Mechanism for the Under-deposit Corrosion of Copper in Compacted Buffer Material (King and Kolar 2000).

The nature of the surface profile shown in Figure 16 is inconsistent with a classical pitting mechanism involving the localized breakdown of a passive film and the permanent separation of anodic and cathodic sites. The entire surface has been corroded, with deeper penetration at certain locations. (Note the difference in scales on the x- and y-axes in Figure 16). Some spatial separation of anodic and cathodic reactions must have occurred to account for the “roughened” profile, but this separation was temporary. Growing pits must have died due to stifling of the cathodic reaction, loss of critical pit chemistry, or by coalescence. Pits would then initiate at a different location and the sequence repeat itself so that the entire surface became corroded to some degree. King and Kolar (2000) proposed a mechanism to account for the type of surface roughness observed (Figure 17). This mechanism involves the birth (initiation), growth, and death of localized corrosion (“pits”) and, repeated, a number of times would result in a roughened surface profile. This mechanism has similarities with those proposed for the localization of corrosion by droplet formation under unsaturated conditions (Figures 13 and 15).

Page 37: Supporting Technical Report - UNENE

- 29 -

3.2.2 Large-scale Tests The BCE and a number of other large-scale heated tests in compacted buffer material have employed materials other than Cu for the heater. The LOT experiment, however, does include Cu in contact with compacted bentonite in the presence of a thermal gradient (Karnland et al. 2000). Results of destructive analyses are available after 1-a exposure (Karnland et al. 2000), as well as from on-line electrochemical measurements (Rosborg et al. 2002, 2004). Preliminary examination of the coupons removed after 1-a exposure suggest uneven attack but no pitting (Karnland et al. 2000). These conclusions are consistent with the on-line electrochemical noise measurements which suggest some “tendency to localized attack, but an inability to maintain the attack” (Rosborg et al. 2004). In a separate set of experiments at the Äspö hard rock laboratory, Taxén (2004) exposed Cu coupons and Cu in contact with compacted bentonite to the humid laboratory atmosphere. One set of coupons in compacted bentonite were heated to 75oC. After three years exposure coupons exposed at ambient temperature retained their original bright colour, whereas those exposed at elevated temperature exhibited a green patina containing paratacamite (Cu2(OH)3Cl). Measured corrosion rates were <0.1 µm⋅a-1.

4. IMPLICATIONS FOR COPPER CONTAINERS IN A DGR

4.1 EXPECTED EVOLUTION OF ENVIRONMENTAL CONDITIONS

4.1.1 Qualitative Description of the Repository Evolution During the Unsaturated Phase Based on the information presented in earlier sections, there is a good qualitative understanding of the expected evolution of the environment at the UFC surface and in the near field during the unsaturated phase. Figure 18 presents a qualitative description of the evolution of the environment on the container surface. Prior to the emplacement of the UFC in the repository (defined here as time t = 0), the container surface may be wetted and contaminated by exposure to the humid repository atmosphere. Surface contamination is inevitable given the presence of airborne organics (lubricants, exhausts, etc.) and aerosols comprising efflorescent salts from the walls of the rooms and tunnels. The surface will become further contaminated by clay particles and salt impurities from the bentonite as the container is emplaced in the buffer material. During the heat-up phase in the evolution of the repository (0-30 a), any existing moisture film will evaporate and the surface will become dry. Evaporation will lead to the concentration of dissolved solutes and the progressive precipitation of salts, beginning with the least soluble phase. New mineral assemblages may be produced on the surface due to the precipitation of different salts from this mixed electrolyte solution. However, if the UFC surface is perfectly dry at the time of emplacement, no new salts will precipitate and the container surface will be contaminated only by those salts deposited from contact with the bentonite.

Page 38: Supporting Technical Report - UNENE

- 30 -

Figure 18: Qualitative Description of the Evolution of the Environment at the UFC Surface During the Unsaturated Phase.

Transportation of sealed UFC through

repository(< 0 years)

Emplacement of UFC

(0 years)

Increase in UFC surface temperature

(0-30 years)

Decrease in UFC surface temperature

(30-200 years)

Complete repository saturation

(>200 years(?))

Timeframe

Possible surface contamination by aerosols, airborne particulates, volatile organics, or gaseous residues from blasting,

leading to adsorption of moisture on surface of UFC in humid repository atmosphere

Contamination of UFC surface by clay minerals and salt impurities (Figure 9)

UFC surface wettedat time of emplacement

UFC surface dryat time of emplacement

Evaporative concentration of surface moisture (Figs 1,3).Progressive precipitation of

solids according to Chemical Divide theory (CO3

2- > SO42- >

Cl- > NO3-) resulting in new

mineral assemblages(Figs 2,3).

Possible acid degassing of concentrated brines.

Surface remains dry

Deliquescence of surface contaminants and mineral assemblages precipitated during period of T increase, with

deliquescence in the order NO3- > Cl- > SO4

2- > CO32-. Mixed salt

systems likely to deliquesce before pure salts (Fig. 5).Initial wetting of surface as isolated droplets centred on salt crystals, gradually spreading, and eventually coalescing and

covering entire surface as %RH increases (Figs. 6, 7).Possible acid degassing of deliquescent brines, resulting in loss

of solute and increase in pH.Dilution of brine solutions as %RH increases.

pH of surface electrolyte controlled in part by pCO2.Relatively oxidizing due to rapid supply of gaseous O2 and

presence of Cu(II) at interface.

Relatively saline environment due to precipitated Cl- and SO42-

salts in bentonite near UFC surface.Depletion of O2 at UFC surface due to rapid consumption but

slow transport through saturated buffer.Slow equilibration of interfacial environment with bulk

groundwater

Page 39: Supporting Technical Report - UNENE

- 31 -

The reverse process will occur during the cool down phase (defined here as 30-200 a, when complete re-saturation is assumed to have occurred). Deliquescence, starting with the most soluble salts in the mineral assemblage, will progressively occur as the %RH increases. In general, nitrate salts are more soluble than chloride salts, which are more soluble than sulphate salts, which are more soluble than carbonate salts. Mineral assemblages comprising different salts in intimate contact tend to be more deliquescent (soluble) than single salts (Figure 5). Initial wetting of the container surface will be as droplets formed by the deliquescence of individual salt crystals. As the RH continues to increase, however, these droplets will spread, additional droplets will form, and eventually droplets will coalesce to form a continuous liquid film on the container surface. Throughout this process additional salt phases will dissolve and the deliquescent brines will become more and more dilute. Depending upon the UFC temperature and the pH of the liquid film, the deliquescent brines may undergo acid degassing, in which HCl, HNO3, or H2SO4 are lost from the solution in the gaseous form, resulting in a loss of solute and an increase in the electrolyte pH. Throughout this period, environmental conditions will be relatively oxidizing due to the continued presence of O2 and its rapid diffusion to the container surface through the unsaturated buffer. Eventually, all of the clay-based sealing materials will become saturated (defined here as a period of 200 a). At this time, the surface environment may remain relatively saline because of the enhancement of the salinity of pore fluids near the container surface (see below). However, the near-surface environment will rapidly become anoxic (if it has not done so already), because of the continual consumption of O2 and, more importantly, the slow rate of transport of O2 to the container surface through saturated bentonite. Slowly over time, the near-surface environment will equilibrate with that of the incoming groundwater. A similar qualitative description for the near-field environment is presented in Figure 19. Many of the same processes that determined the container surface environment, such as progressive precipitation according to the Chemical Divide theory and deliquescence, will also control conditions in the near field (considered to comprise the clay-based sealing materials in the emplacement room). In addition, there is substantial evidence from laboratory and large-scale tests that soluble sulphate and chloride salts will accumulate in the pore water near the heat source (container), whereas carbonate salts will accumulate in the bentonite further away. Therefore, during the heat-up phase, there will be a separation of precipitated salts within the pores of the buffer which, upon re-saturation during the cool-down phase, will re-dissolve and produce regions with pore water of different chemical composition. The pore water near the container surface will then be relatively saline (high [Cl-] and [SO4

2-]) compared with the pore water further away from the UFC. Throughout the heat-up and cool-down phases, the transport of solutes and gases will be affected by saturation-dependent changes in the diffusivity. The diffusivity of gaseous species will increase during the heat-up phase as the buffer dries out, but will decrease as the buffer re-saturates during cool-down. The diffusivity of dissolved solutes will be oppositely affected.

Page 40: Supporting Technical Report - UNENE

- 32 -

Figure 19: Qualitative Description of the Evolution of the Environment in the Near Field During the Unsaturated Phase.

Transportation of sealed UFC through

repository(< 0 years)

Emplacement of UFC

(0 years)

Increase in UFC surface temperature

(0-30 years)

Decrease in UFC surface temperature

(30-200 years)

Complete repository saturation

(>200 years(?))

Timeframe

Conditioning of pore water by dissolution of soluble mineral impurities in clay and ion exchange with bentonite.

Movement of water away from UFC (Fig. 12), resulting in decreasing degree of saturation (S) and water activity (aW), but increasing suction potential (Sr) (Fig. 8). May result in either re-distribution of water or net loss/gain depending upon DGR

configuration and boundary conditions.Precipitation of dissolved solutes in sequence CO3

2- > SO42- >

Cl- > NO3-.

Re-distribution of solutes, with SO42- and Cl- accumulating

close to heat source and CO32- further away from UFC.

Decreasing diffusivity of dissolved solutes (Fig. 10), but increasing diffusivity of gaseous species (Fig. 11).

Cessation of microbial activity for aW < 0.96.Oxidizing environment due to ready supply of O2.

Movement of water towards UFC, resulting in increasing S and aW, but decreasing Sr (Fig. 8).

Re-dissolution of precipitated salts, initially by deliquescence (Fig. 5) and eventually by incoming groundwater. Salts dissolve

in sequence NO3- > Cl- > SO4

2- > CO32-, with mixed salt

assemblages the first to deliquesce. Gradual dilution of deliquescent brines as %RH (aW) increases.

Enhanced pore-water [SO42-] and [Cl-] close to container and

enhanced [CO32-] further away from UFC.

Progressive equilibration of pore water with incoming groundwater, involving ion exchange with bentonite.

Increasing diffusivity of dissolved solutes (Fig. 10), but decreasing diffusivity of gaseous species (Fig. 11).Re-establishment of microbial activity for aW > 0.96.

Oxidizing environment if O2 still present.

Continued slow equilibration of pore water with ground water.Increasing diffusivity of dissolved solutes (Fig. 10), but rapidly

decreasing diffusivity of gaseous species as bentonite saturates (Fig. 11).

Onset of anoxic conditions close to UFC due to slow rate of supply of O2, but environment could still be oxidizing due to

presence of Cu(II).Establishment of buffer swelling pressure.

Suppression of any microbial activity by mechanical forces for swelling pressures >2 MPa.

Page 41: Supporting Technical Report - UNENE

- 33 -

Microbial activity in the repository will also be affected by changes in the degree of saturation of the sealing materials during the unsaturated phase. A pore-water activity of 0.96 has been established as a threshold below which most, if not all, microbial activity ceases (King et al. 2002b, 2003, 2004). Therefore, microbial activity will be suppressed in any region of the near field in which the water activity falls below this value, equivalent to an RH of 96%. This condition is believed to exist in dense buffer at all moisture contents, but may also be achieved in 50:50 buffer and dense and light backfill at low levels of saturation (Figure 8). In addition, there is evidence that swelling pressures in excess of 2 MPa also suppress microbial activity (Hedin 2006), possibly due to the mechanical forces on cell walls due to the swelling clay, so that microbes will not be active in the dense buffer even following saturation. Eventually, once the repository is saturated, the composition of the pore water of the sealing materials will equilibrate with that of the surrounding groundwater. Equilibration is likely to be a slow process due to the limited rates of mass transport, and will involve dissolution of previously precipitated salts and ion exchange with the bentonite clay.

4.1.2 Semi-quantitative Description of the Repository Evolution During the Unsaturated Phase

Although detailed THMC calculations have not yet been performed for the unsaturated phase, it is possible to develop a semi-quantitative description of the evolution of environmental conditions at the UFC surface based on a set of assumed time-dependent re-saturation times. King and Kolar (2006) used the time-dependent saturation values for inner and outer buffer shown in Figure 20 for calculations of O2 consumption in a DGR, which were based on analyses for the Third Case Study performed by Gierszewski et al. (2004). For this analysis, the inner (100% bentonite) and outer (50:50 bentonite:sand) buffer materials were subdivided into three and two components, respectively, and the time-dependent saturation estimated for each sub-layer. In general, the buffer material is assumed to dry out during the heat-up phase (0-20 a), especially in the inner buffer closest to the container. Moisture is assumed to be re-distributed into the outer buffer to some degree, resulting in a smaller decrease in saturation further from the container. These assumptions are consistent with the observations from large-scale tests described in Section 2.3.1. The time-dependent saturation in Figure 20 can be used to predict the evolution of conditions at the container surface during the unsaturated phase. Let us assume that the water activity at the container surface is the same as that in the section of inner buffer closest to the container, namely section “Inner Buf 1” in Figure 20. The aW (%RH) in the inner (100% bentonite) buffer can then be estimated from Equations (1) and (2) and Baumgartner’s (2006) data illustrated in Figure 8. Figure 21 shows the time dependence of the %RH at the container surface (for both the heat-up and cool-down phases) and the container surface temperature. As is particularly evident from the linear-time plot (Figure 21(b)), the unsaturated phase is characterized by a rapid decrease in the %RH at the container surface during the heating part of the thermal cycle, and a much slower subsequent increase in %RH during the cool-down phase.

Page 42: Supporting Technical Report - UNENE

- 34 -

Figure 20: Assumed Time-dependent Saturation of Inner and Outer Buffer Materials in a Deep Geologic Repository Used by King and Kolar (2006).

The temperature and %RH data in Figure 21 are re-plotted in %RH-T space in Figure 22. These curves define the conditions for the efflorescence (precipitation) and deliquescence of salts during the heating and cooling periods, respectively. Thus, during the heating period, the %RH will decrease rapidly as the temperatures rises above 50oC. By reference with the evaporation data in Figure 3, the concentrations of solutes in any surface liquid film will increase significantly resulting in the precipitation of salts, the nature of which will be governed by the Chemical Divide Theory (Figures 2 and 4). It is likely that all but the most soluble salts will precipitate before the container reaches the maximum assumed temperature of 95oC. For salts to remain at least partially dissolved during heat-up, they must exhibit efflorescence relative humidities (ERH) to the left or below the zone defined by the heating curve. For example, for a solute to remain in solution at a temperature of 80oC during the heat-up phase, the ERH must be <8%. Very few salts exhibit such low ERH and the surface is likely to be completely dry at this point. For the cooling part of the thermal cycle, the curve in Figure 22 again defines the conditions of %RH and temperature at the surface. Salts on the surface will deliquesce (i.e., start to dissolve) if their DRH lies to the left or below the cooling curve in the figure. The %RH is predicted to increase relatively rapidly with decreasing temperature, suggesting that it is likely that a number of salts might deliquesce during this period.

0

0.2

0.4

0.6

0.8

1

1.2

0.1 1 10 100 1000

Time (yrs)

Deg

ree

of s

atur

atio

n

Inner Buf 1Inner Buf 2Inner Buf 3Outer Buf 1Outer Buf 2

Page 43: Supporting Technical Report - UNENE

- 35 -

(a) Logarithmic time plot.

(b) Linear time plot.

Figure 21: Predicted Time Dependence of the Container Surface Temperature and Interfacial Relative Humidity During the Unsaturated Phase.

0

20

40

60

80

100

120

0.1 1 10 100 1000

Time (yrs)

% R

elat

ive

Hum

idity

or

Tem

pera

ture

(o C)

Temperature%RH Heating%RH Cooling

0

20

40

60

80

100

120

0 50 100 150 200

Time (yrs)

% R

elat

ive

Hum

idity

or

Tem

pera

ture

(o C)

Temperature%RH Heating%RH Cooling

Page 44: Supporting Technical Report - UNENE

- 36 -

Figure 22: Predicted Relative Humidity-Temperature Relationship at the Container Surface.

Figure 23: Comparison of Container Surface Relative Humidity-Temperature Conditions with Deliquescence Behaviour of Various Salt Systems.

0

20

40

60

80

100

120

20 40 60 80 100

Temperature (oC)

% R

elat

ive

Hum

idity

HeatingCooling

0

20

40

60

80

100

120

20 40 60 80 100

Temperature (oC)

% R

elat

ive

Hum

idity

HeatingCoolingCaCl2.2H2OCaCl2.4H2OCaCl2.6H2ONaClNaNO3NaCl-NaNO3

Page 45: Supporting Technical Report - UNENE

- 37 -

Figure 23 shows the deliquescence RH-T behaviour of a number of salts, superimposed on the %RH-T curves for the container surface from Figure 22. Neither the composition of the liquid film that might be present on the UFC surface initially nor the nature of those salts that might be present on the container surface after the heat-up period are currently known. For illustrative purposes, however, let us assume that a thin liquid film containing Na+, Ca2+, Cl-, and NO3

- ions is present initially due to the adsorption of moisture during UFC transport in the humid repository and because of contamination by salt impurities in the bentonite clay and volatile nitrate residues from blasting operations. During the heat-up phase, this liquid film would begin to evaporate and solutes would sequentially precipitate from solution. Assuming the ERH and DRH are equivalent, the first salt that would precipitate of those considered in Figure 23 is NaCl, followed by NaNO3. Based on the times at which the ERH curves for these solids cross the %RH-T cooling curve, NaCl would precipitate within a matter of days and NaNO3 within 2-3 months of UFC emplacement. The solution that would then be present on the surface would be free of Na+ and NO3

- ions and, for the assumptions made here, be equivalent to a calcium chloride brine. This brine would continue to concentrate until such time that the solution is supersaturated and CaCl2⋅6H2O precipitates. Based on the data in Figure 23, final evaporation of the solution would occur approximately 10-12 months after container emplacement at which point the UFC surface temperature would be ~55oC. Thus, during the heat-up part of the thermal cycle, the container surface would be exposed to a continually concentrating Na-Ca-Cl-NO3 solution for a period of ~1 a. For the salts assumed to be present here, the UFC surface remains dry until some time after the surface temperature begins to cool. The deliquescence behaviour is approximately the reverse of the efflorescence behaviour. The first salt to deliquesce would be CaCl2⋅6H2O, which would occur approximately 30 a after container emplacement. Initially the container surface would be wetted by a saturated calcium chloride brine, but only at those locations at which CaCl2⋅6H2O crystals had been precipitated during the heat-up. As time progresses and the %RH increases, the calcium chloride brine will become increasingly dilute until, at a point defined by the intersection of the cooling curve and the NaNO3 and NaNO3-NaCl deliquescence curves in Figure 23, NaCl and/or Na NO3 will also begin to dissolve. This is predicted to occur at about 110 a after container emplacement at a surface temperature of ~89oC. Isolated NaCl crystals (i.e., those not in intimate contact with NaNO3 crystals) will deliquesce approximately 140 a after container emplacement at a surface temperature of ~88oC. From that point on, the surface liquid film will become increasingly dilute at the %RH continues to increase. Thus, during the cool-down phase, the surface of the container will be first wetted by a saturation Ca-Cl brine after ~30 a, which will become progressively diluted and mixed with NO3 and Na as additional salts deliquesce and dissolve.

4.2 EXPECTED EVOLUTION OF CORROSION BEHAVIOUR OF COPPER CONTAINERS

4.2.1 Forms of Corrosion Before describing the likely evolution of the corrosion behaviour of a copper UFC during the unsaturated phase, the impact of these conditions on each of the major corrosion processes of concern will be described.

Page 46: Supporting Technical Report - UNENE

- 38 -

4.2.1.1 Uniform Corrosion It is unlikely that, in the strictest sense, uniform corrosion will occur during the unsaturated phase. Uniform corrosion requires that the anodic and cathodic reactions occur equally over the surface of the container. For the reasons described above, it is likely that spatial separation of these reactions will occur, due initially to wetting of the surface by droplets. Precipitation of corrosion products will sustain this localized separation of reactions until such time that the system is completely saturated.

4.2.1.2 Localized (Under-deposit) Corrosion Localized corrosion is likely to be the major form of attack during the unsaturated period. Droplet formation and the precipitation of corrosion products will result in the spatial separation of anodic and cathodic sites. However, the degree of localization will diminish over time as the surface becomes progressively wetted and as a consequence of the coalescence of individual localized sites. Experimental evidence from laboratory and full-scale tests suggests that corrosion will take the form of surface roughening, rather than the propagation of deep, isolated pits. Surface roughening results from sequential birth (initiation), growth (propagation), and death (stifling) of individual localized sites. Propagating “pits” will cease to grow as a result of the loss of the critical “pit” chemistry or coalescence of neighbouring sites resulting in a decrease in the cathodic:anodic surface area ratio. The degree of surface roughening (as measured by the ratio of the depth of the deepest site to the average penetration depth) is likely to decrease with time. Initially, the ratio of maximum to mean penetration depth will be large as some areas of the surface will not be attacked at all, either because they are not wetted or because they correspond to cathodic areas associated with microdroplets or around the periphery of larger droplets. As corrosion continues, however, the cathodic:anodic surface area will decrease due to increasing wetting of the surface and the coalescence of adjacent sites. The surface profile illustrated in Figure 16 may represent an intermediate stage in this evolution and exhibits a maximum:mean penetration depth of ~1.4. Over time, it is expected that this ratio would decrease further and the degree of surface roughness would diminish.

4.2.1.3 Stress Corrosion Cracking Although SCC is feasible during the unsaturated phase, it is considered to be unlikely on both environmental and stress grounds. From an environmental perspective, there is a general absence of species known to cause SCC. The major sources of SCC agents (nitrite, ammonia, and acetate ions) during the unsaturated phase are residues from blasting operations and microbial activity occurring within the repository at a location remote from the container surface. It is assumed that the level of airborne ammonia and nitrite from blasting will either be too low to cause SCC or could be controlled by suitable engineering means to preclude the possibility of cracking. Microbial activity is expected to be minimal during the unsaturated phase because of the low water activity. If it does occur in wetter regions of the repository, the SCC agents must then diffuse to the UFC surface for cracking to occur. During the unsaturated phase, only the diffusion of

Page 47: Supporting Technical Report - UNENE

- 39 -

gaseous ammonia is likely to be sufficiently fast to result in an aggressive cracking environment. In the event that an SCC agent is present, however, the redox conditions are likely to be sufficiently oxidizing to support cracking. From a mechanical loading perspective, the only source of stress during the unsaturated phase is residual tensile stress from container manufacture and sealing. It is assumed that some form of stress relief will be applied following manufacture of the container shell and bottom end cap, so that the only source of residual stress will be the final closure weld between the shell and the lid. It is possible to induce compressive surface stresses to the final closure weld without heating the UFC using techniques such as laser-peening or low-plasticity burnishing (BSC 2004c). Such treatment would further minimize the probability of SCC during the unsaturated phase.

4.2.1.4 Microbiologically Influenced Corrosion Microbiologically influenced corrosion is considered to be unlikely at any time during the UFC service lifetime, but especially during the unsaturated phase. Among the various factors that can limit the extent of microbial activity, one of the severest stressors is the absence or lack of sufficient water (King et al. 2002b, 2003, 2004). The unsaturated phase, by definition, corresponds to the period of lowest water activity in the buffer and backfill sealing materials. Microbial activity is considered impossible in highly compacted bentonite at any degree of saturation. Based on the data in Figure 8, microbial activity would also be impossible in the 50:50 buffer material at saturation levels of <80-90%. Consequently, MIC is considered to be unlikely during the unsaturated period.

4.2.1.5 Oxidation Based on the analysis in Section 4.1.2, the surface of the UFC could be dry for the period between ~1 a and ~30 a following container emplacement. During this period, the container surface temperature will increase from ~55oC to the maximum temperature of 95oC (based on the nominal conditions assumed for the Third Case Study, Gierszewski et al. 2004). As discussed in Section 3.1.3, the rate of oxidation at these temperatures is low and the maximum extent of oxidation is predicted to be <0.1 µm. Therefore, oxidation is considered to be an insignificant contributor to UFC corrosion during the unsaturated phase.

4.2.2 Evolution of Corrosion Behaviour During Unsaturated Phase A qualitative description of the evolution of the corrosion behaviour of a copper UFC during the unsaturated phase is illustrated in Figure 24. This description follows the same sequence of events used to describe the evolution of the environment at the UFC surface (Figure 18) and in the near-field (Figure 19).

Page 48: Supporting Technical Report - UNENE

- 40 -

Figure 24: Qualitative Description of the Evolution of the Corrosion Behaviour of a Copper Used Fuel Container During the Unsaturated Phase.

Transportation of sealed UFC through

repository(< 0 years)

Emplacement of UFC

(0 years)

Increase in UFC surface temperature

(0-30 years)

Decrease in UFC surface temperature

(30-200 years)

Complete repository saturation

(>200 years(?))

Timeframe

Brief period of atmospheric corrosion possible if surface is wetted and contaminants (SO2, NO2, O3, (NH4)2SO4) are present.

Minimal oxidation if surface is dry (<0.1 m).

Possible onset of localized corrosion (LC) if UFC surface wetted during transportation (Figs. 13, 15-17).

Minimal oxidation if surface is dry (<0.1 m).

Limited corrosion as surface wet for only short period.Aqueous corrosion processes cease completely as surface dries.

Minimal oxidation if surface is dry (<0.1 m).No lower %RH threshold for aqueous corrosion as it depends on

nature of surface contaminants.

Re-establishment of wetted surface promotes LC in liquid droplets (Figs. 13, 15-17). ECORR may be generally elevated, and

especially around periphery of droplets, due to rapid supply of O2.Continuous cycle of birth, growth, and death of LC sites, resulting

in under-deposit corrosion mechanism.Tendency for corrosion to become less localized as droplets spread and coalesce, resulting in more uniform distribution of

anodic and cathodic sites. Corresponding decrease in value and spatial variability of ECORR.

Classical pitting mechanism (defined by ECORR > Eb) only possible during initial stages of re-saturation.

SCC only possible if NH3 or acetic acid produced in near-/far-field diffuses to UFC surface and if sufficient residual stress to support

cracking.Ant’s nest corrosion possible if appropriate organic acids are

present.No MIC under biofilm provided aW < 0.96.

Evolution of roughened surface to more uniform surface profile.Period of slow uniform corrosion supported by O2 and Cu(II)

reduction.Negative drift in ECORR as [Cl-] increases and as oxidant (O2 and

Cu(II)) is consumed.Gradual transformation to non-oxidizing conditions as Cu(II) is

reduced to Cu(I).SCC unlikely due to decrease in ECORR, consumption of oxidants,

and slow diffusion of SCC agents through saturated buffer.No MIC under biofilm provided aW < 0.96 and swelling pressure

>2 MPa.

Page 49: Supporting Technical Report - UNENE

- 41 -

During the transportation of the sealed UFC through the repository, there is the possibility of a brief period of atmospheric corrosion, especially if the atmosphere contains airborne contaminants, such as SO2, NO2, O3, or (NH4)2SO4. These and other contaminants are ubiquitous in industrial environments, but the period of exposure of the container is likely to be so short that minimal corrosion is expected. However, contamination of the surface could lead to subsequent corrosion after emplacement as the container cools (see below). If the container surface is dry during transportation, the extent of dry air oxidation would be minimal as the temperatures will be low. If the UFC surface is wetted during transportation in the repository, localized corrosion could occur as the container is emplaced and sealed in the repository. Contamination of the surface by mineral impurities in the clay (e.g., CaSO4⋅2H2O, NaCl) could contribute to the development of localized environments in which the anodic and cathodic reactions are spatially separated. As the UFC surface temperature increases, aqueous corrosion processes will only be possible for a short period (of the order of 1 a) before the surface dries completely. There is no critical %RH below which corrosion does not occur; rather, the initial time-of-wetness depends on the composition of the surface moisture film and the efflorescence properties of the precipitated salts. However, because of the expected rapid decrease in the %RH at the container surface, the initial period of aqueous corrosion will be brief. Furthermore, although the surface may remain dry for a period of ~30 a, the extent of oxidation will again be minimal as the container surface temperature is relatively low. Following the peak in container temperature after ~30 a, the UFC surface will once again be wetted by the deliquescence of the salts precipitated on the surface during the heat-up phase. Corrosion will be localized as the surface is wetted by droplets formed by the deliquescence of individual salt crystals or mineral assemblages. The degree of localization will depend on the time dependence of the cathodic:anodic surface area ratio. If this ratio decreases rapidly, as would occur if the surface is heavily contaminated by salt crystals or if the buffer material rapidly saturates, then the degree of surface roughening will be less than if the physical separation of anodic and cathodic sites is maintained for long periods of time. Of the precipitated salts likely to be present, CaCl2⋅xH2O will deliquesce first (i.e., at the lowest %RH and highest temperature). The precipitation of basic cupric chloride salts can be expected. Sulphate salts may also be present and will dissolve at lower temperature/higher %RH, leading to the precipitation of the analogous basic cupric sulphate salt. As the %RH continues to increase, the surface will become more uniformly wetted and the tendency towards localized corrosion will diminish. Throughout this period, a spatial variation in the electrochemical potential will exist over the surface of the UFC. Cathodic sites under microdroplets or around the periphery of larger droplets will be more positive than those of anodic sites undergoing active dissolution at the centre of the water droplets. Stress corrosion cracking is only possible if an active cracking agent and a sufficiently high residual tensile stress are present. No biofilm will be present on the container surface as the water activity will be less than that required to sustain microbial activity. The diffusion of remotely produced metabolic by products is possible if other regions of the repository are sufficiently wet to sustain microbial activity.

Page 50: Supporting Technical Report - UNENE

- 42 -

Eventually, as the repository fully saturates, the localized, under-deposit corrosion will cease and the attack will become more uniform in nature. The degree of surface roughness will diminish over time as uniform corrosion is supported by the consumption of the remaining O2 and Cu(II). The electrochemical potential of the container surface will be more uniform and the propensity for SCC and localized corrosion will diminish as the value of ECORR decreases.

4.2.3 Estimate of Corrosion Damage During Unsaturated Phase There is currently insufficient information on which to base an accurate assessment of the extent of corrosion during the unsaturated phase. Based on the predicted time-dependence of the container surface temperature, it is clear that the extent of oxidation will be minimal (<0.1 µm). By analogy with the reported rates of atmospheric corrosion at ambient temperature, the rate of corrosion once the container surface is (re-)wetted during the cooling phase could be of the order of 1-2 µm⋅a-1. This mean rate must be multiplied by a surface roughness factor (with, perhaps, a value of 1.5) to account for the localized nature of the attack. If the UFC surface is wetted for a period of ~150 a before complete saturation (see Section 4.1.2), then the mean and maximum depths of penetration would be 150 µm and 225 µm, respectively (for a mean corrosion rate of 1 µm⋅a-1). The extent of corrosion during the unsaturated period, therefore, is small in comparison to the wall thickness and container failure will not occur during this period. It is interesting to compare this estimate of the mean wall penetration with the total inventory of oxidants in the repository. For the reference in-room disposal configuration (Gierszewski et al. 2004), the initial inventory of O2 (the only oxidant in the repository) per unit area of the container is ~6 x 10-4 mol⋅cm-2. This amount of oxidant is equivalent to a uniform wall penetration of 175 µm. Therefore, if the estimated depth of corrosion above is accurate, virtually all of the oxidant in the repository will be consumed during the initial unsaturated phase and corrosion will cease soon after the saturation of the buffer and backfill sealing materials.

4.3 EFFECT ON THE NATURE OF THE HOST ROCK The results of this review and gap analysis apply generally to both granitic and sedimentary host rock formations. The near-field environmental conditions are determined more by the nature of the buffer and backfill materials and less by the properties of the far-field. However, the nature of the host rock clearly affects the groundwater salinity and the saturation time. In particular, groundwaters in sedimentary deposits are more saline than those found at an equivalent depth in the Canadian Shield and the lower permeability of shale formations will result in longer saturation times for repositories located in these formations (King 2005). As a consequence, the unsaturated phase is likely to last longer in sedimentary formations. Nevertheless, the same processes of dry-out (evaporation/efflorescence) and deliquescence will occur regardless of the nature of the host rock and the general conclusions from this study remain valid for both granitic and sedimentary formations.

Page 51: Supporting Technical Report - UNENE

- 43 -

5. GAP ANALYSIS There are a number of areas in which our understanding of the corrosion behaviour of copper UFC during the unsaturated phase is incomplete. Some of these areas relate to the uncertainty in the evolution of the environmental conditions, particularly the time dependence of the degree of saturation of the sealing materials surrounding the container. To some degree the time-dependent saturation behaviour is specific to the actual conditions of the site chosen for the location of the repository. Other areas relate more generically to the corrosion behaviour of the container and could, therefore, be studied in more detail prior to the selection of the repository location. These areas include assessments of the evolution of both the environmental conditions and of the corrosion behaviour. Gaps in our understanding of the evolution of the container surface and near-field environments include:

1. Assessment of the degree of surface contamination of emplaced UFC The evolution of environmental conditions on the container surface are closely related to the nature of the surface contaminants. Knowledge of the nature (composition) and spatial density of salt deposits would permit a more-informed analysis of the time-dependent evolution of the surface environment during both the heating and cooling cycles of the thermal pulse. For example, knowledge of the nature of the surface contaminants on an emplaced container would indicate whether the surface will be wetted during the heating cycle and whether analysis of the evaporation and efflorescence of salts is necessary. Similarly, the deliquescence behaviour during the cooling cycle is determined by the nature and density of surface salts. The nature of the salts will determine when the surface will (re-)wet and the surface density of contaminants will determine how many water droplets will form and the degree of localization of the resultant corrosion. (The greater the surface contamination, the less localized the corrosion). Information on surface contamination could be obtained from either laboratory tests or from large-scale in situ tests such as those being carried out at the Äspö HRL.

2. Characterization of atmospheric impurities in a repository Surface contamination of the container prior to emplacement will be determined by the extent and nature of atmospheric pollutants in the repository. Although no repository is yet in operation, testing at an underground laboratory would provide useful information, in particular regarding the concentrations of pollutants such as SO2, NO2, (NH4)2SO4, salt aerosols, and airborne blasting residues.

3. Evolution of the pore-water chemistry Evolution of the pore-water chemistry is an important consideration for corrosion during the unsaturated and early saturated phases. The reported increase in pore-water salinity close to the container under the action of the thermal gradient is significant for the corrosion behaviour of the container. Additional experimental and/or modelling studies are warranted.

Page 52: Supporting Technical Report - UNENE

- 44 -

Gaps in our understanding of the evolution of the corrosion behaviour of the container under unsaturated conditions include:

1. Experimental studies with heated surfaces Although we have a conceptual understanding of the mechanism of corrosion under unsaturated conditions, this understanding is based on analogy with studies of atmospheric corrosion. Confirmatory experimental evidence is required using heated copper surfaces exposed to appropriate environmental conditions (%RH, salt contamination, etc.). Such information could be obtained from specifically targeted experimental studies or from examination of heaters from large-scale in situ tests.

2. Characterization of corrosion products on copper surfaces Part of our mechanistic understanding is based on the nature of the corrosion products formed on surfaces during atmospheric corrosion. Further examination of the nature and distribution of corrosion products should be performed.

3. Variation of the corrosion potential on copper surface under unsaturated conditions The corrosion potential ECORR is a primary parameter determining the corrosion behaviour of metallic surfaces. In the current context, the spatial variation and distribution of ECORR indicates the extent of localized corrosion resulting from the separation of anodic and cathodic processes in water droplets on the surface. Experimental measurements of ECORR as a function of %RH and the presence of salt contamination would provide useful information for estimating the duration and consequences of the localized attack.

4. Mathematical model for under-deposit corrosion The conceptual basis for an under-deposit corrosion model has been defined, but has yet to be implemented. The model would account for the time-dependent birth (initiation), growth (propagation), and death of localized corrosion sites and could be used to predict the evolution of the cathodic:anodic surface area ratio. The model would also predict the degree of surface roughness for comparison with experimental observations. In addition, the model could be used to predict the fraction of the initial inventory of trapped O2 that will be consumed prior to the establishment of saturated conditions. This latter prediction would serve as an important input parameter for the CCM suite of corrosion models.

5. Evolution of the degree of surface roughness Since corrosion during the unsaturated phase is likely to be localized, it is important to characterize the degree of surface roughness and how it evolves with time. Some data are already available, but a systematic study of the effects of %RH, temperature, and salt contamination is required. In particular, it has been suggested here that the degree of roughening will diminish with time (i.e., increasing %RH), but this has yet to be experimentally proven. Such evidence could be obtained from the examination of heated test sections from large-scale in situ testing or, preferably, from well-characterized laboratory tests.

Page 53: Supporting Technical Report - UNENE

- 45 -

6. CONCLUSIONS An assessment of the likely evolution of the near-field environmental conditions and corrosion behaviour of copper UFC during the unsaturated phase has been carried out. The unsaturated phase can be conveniently divided into two periods: the period of increasing container temperatures corresponding to surface dry-out and the subsequent cooling period when the surface re-wets. The environment on the container surface is determined by the evaporation of surface moisture and the efflorescence (precipitation) of salts during the heating cycle, and the deliquescence (dissolution) of these salts during the cooling phase. The deliquescence behaviour, in particular, is important as it determines the nature of corrosive solutions that might form during cool-down as well as the time-of-wetness. Wetting of the surface will occur first as droplets formed by the deliquescence of isolated salt crystals but will eventually spread over the entire surface as the %RH increases. Similar processes will occur in the near-field sealing materials resulting in an evolution of the pore-water chemistry. There is evidence that chloride and sulphate salts will be re-distributed during the thermal cycle and concentrate close to the container surface. Desiccation of the buffer and backfill materials will also affect the diffusivity of gaseous and dissolved species in the near field. Corrosion of the container during the unsaturated phase will take the form of localized or under-deposit attack. Localized corrosion will result from the spatial separation of anodic and cathodic sites in water droplets that first form on the container surface during the cooling cycle. Cathodic processes will be limited to the periphery of the water droplets or to surrounding microdroplets, areas for which the rate of supply of O2 is fast. Anodic processes will occur in the centre of water drops which will remain relatively anoxic. However, as these water droplets spread out with increasing %RH, the relative cathodic:anodic surface area ratio will decrease and attack will become more uniform. Nevertheless, the surface of the container is likely to have undergone some surface roughening by corrosion by the end of the unsaturated phase. Other forms of corrosion, including SCC and MIC, are considered to be unlikely during the unsaturated phase. Stress corrosion cracking requires the presence of residual tensile stress and a sufficient concentration of SCC agent. Microbial activity close to the container during the unsaturated phase is unlikely due to the lack of water. The rate of dry air oxidation is slow at temperatures <100oC and the extent of such attack is predicted to be <0.1 µm. An assessment has been made of the possible depth of attack on a copper UFC during the unsaturated period. The mean depth of penetration is predicted to be of the order of 100-200 µm, with a maximum penetration of 150-300 µm. It is possible that the majority of the initially trapped O2 will be consumed prior to the establishment of saturated conditions. Gaps in the current understanding of the evolution of environmental conditions and of the corrosion behaviour of copper UFC have been identified. The conclusions of this study apply generally to repositories in both granitic and sedimentary host rocks, with the major differences being the higher salinity and possible longer saturation times of the latter.

Page 54: Supporting Technical Report - UNENE

- 46 -

REFERENCES Alai, M., M. Sutton, and S. Carroll. 2005. Evaporative evolution of a Na-Cl-NO3-K-Ca-SO4-Mg-

Si brine at 95oC: experiments and modeling relevant to Yucca Mountain, Nevada. Geochem. Trans. 6, 31-45.

Arcos, D., J. Bruno, S. Benbow, and H. Takase. 2000. Behaviour of bentonite accessory

minerals during the thermal stage. Swedish Nuclear Fuel and Waste Management Company Technical Report, TR-00-06.

Baumgartner, P. 2000. Elemental composition of disposal vault sealing materials. Ontario

Power Generation Nuclear Waste Management Division Report No: 06819-REP-01300-10015.

Baumgartner, P. 2006. Generic thermal-mechanical-hydraulic (THM) data for sealing

materials. Volume 1: soil-water relationships. Ontario Power Generation Nuclear Waste Management Division Report No: 06819-REP-01300-10122.

Bian, L., Y. Weng, and X. Li. 2005. Observation of micro-droplets on metal surface in early

atmospheric corrosion. Electrochem. Comm. 7, 1033-1038. Bruno, J., D. Arcos, and L. Duro. 1999. Processes and features affecting the near field

hydrochemistry. Groundwater-bentonite interaction. Swedish Nuclear Fuel and Waste Management Company Technical Report, TR-99-29.

Bryan, C.R. 2006. Evolution of waste package environments in a repository at Yucca

Mountain. Presentation to the U.S. Nuclear Waste Technical Review Board, September 25-26, 2006, Las Vegas, NV. Available at http://www.nwtrb.gov/meetings.

BSC. 2004a. Engineered barrier system: physical and chemical environment model. Report

prepared for the U.S. DOE by Bechtel SAIC Company, ANL-EBS-MD-000033 Rev 02. Available from http://www.ocrwm.doe.gov/technical/index.shtml.

BSC. 2004b. In-drift precipitates/salts model. Report prepared for the U.S. DOE by Bechtel

SAIC Company, ANL-EBS-MD-000045 Rev 02. Available from http://www.ocrwm.doe.gov/technical/index.shtml.

BSC. 2004c. Stress corrosion cracking of the drip shield, the waste package outer barrier, and

the stainless steel structural material. Bechtel SAIC Company report to DOE, ANL-EBS-MD-000005 Rev 02, October 2004. Available from http://www.ocrwm.doe.gov/technical/index.shtml.

Cano, E., J.M. Bastidas, J.L. Polo, and N. Mora. 2001. Study of the effect of acetic acid

vapour on copper corrosion at 40 and 80% relative humidity. J. Electrochem. Soc. 148, B431-B437.

Chawla, S.K. and J.H. Payer. 1990. The early stage of atmospheric corrosion of copper bu

sulphur dioxide. J. Electrochem. Soc. 137, 60-64.

Page 55: Supporting Technical Report - UNENE

- 47 -

Chen, Z.Y., D. Persson, and C. Leygraf. 2005a. In situ studies of the effect of SO2 on the initial NaCl-induced atmospheric corrosion of copper. J. Electrochem. Soc. 152, B526-B533.

Chen, Z.Y., S. Zakipour, D. Persson, and C. Leygraf. 2005b. Combined effects of gaseous

pollutants and sodium chloride particles on the atmospheric corrosion of copper. Corrosion 61, 1022-1034.

Chen, Z.Y., D. Persson, A. Nazarov, S. Zakipour, D. Thierry, and C. Leygraf. 2005c. In situ

studies of the effect of CO2 on the initial NaCl-induced atmospheric corrosion of copper. J. Electrochem. Soc. 152, B342-B351.

Chen, Z.Y., D. Persson, F. Samie, S. Zakipour, and C. Leygraf. 2005d. Effect of carbon

dioxide on sodium chloride-induced atmospheric corrosion of copper. J. Electrochem. Soc. 152, B502-B511.

Collin, M. and A. Rasmuson. 1988. A comparison of gas diffusivity models for unsaturated

porous media. Soil Sci. Soc. Am. J. 52, 1559-1565. Corbett, R.A. and P. Elliott. 2000. Ant-nest corrosion – digging the tunnels. Proc.

CORROSION 2000, NACE International (Houston, TX), paper no. 00646, Cuevas, J., M.V. Villar, A.M. Fernandez, P. Gomez, and P.L. Martin. 1997. Pore waters

extracted from compacted bentonite subjected to simultaneous heating and hydration. Appl. Geochem. 12, 473-481.

Curti, E. and P. Wersin. 2002. Assessment of porewater chemistry in the bentonite backfill for

the Swiss SF/HLW repository. Nagra Technical Report, 02-09. Dante, J.F. and R.G. Kelly. 1993. The evolution of the adsorbed solution layer during

atmospheric corrosion and its effects on the corrosion rate of copper. J. Electrochem. Soc. 140, 1890-1897.

Dixon, D.A., N.A. Chandler, A.W-L. Wan, S. Stroes-Gascoyne, J. Graham, and D.W. Oscarson.

1998. Pre- and post-test properties of the buffer, backfill, sand and rock components of the buffer/container experiment. Atomic Energy of Canada Limited Report, AECL-11786, COG-97-287-I.

Drever, J.J. 1997. The Geochemistry of Natural Waters: Surface and Groundwater

Environments, 3rd edition (Prentice-Hall, Englewood Cliffs, NJ). Elliott, P. and R.A. Corbett. 1999. Ant nest corrosion – exploring the labyrinth. Proc.

CORROSION/99, NACE International (Houston, TX), paper no. 99342. EPRI. 2004. Comments regarding in-drift chemistry related to corrosion of containment

barriers at the candidate spent fuel and HLW repository at Yucca Mountain, Nevada. Electric Power Research Institute Technical Update, 1010941.

Page 56: Supporting Technical Report - UNENE

- 48 -

Gdowski, G.E. and D.B. Bullen. 1988. Survey of degradation modes of candidate materials for high-level radioactive-waste disposal containers. Volume 2. Oxidation and corrosion. Lawrence Livermore National Laboratory Report, UCID-21362, Vol. 2.

Gierszewski, P., J. Avis, N. Calder, A. D'Andrea, F. Garisto, C. Kitson, T. Melnyk, K. Wei and L.

Wojciechowski. 2004. Third Case Study - Postclosure Safety Assessment. Ontario Power Generation, Nuclear Waste Management Division Report 06819-REP-01200-10109-R00. Toronto, Canada.

Graham, J., N.A. Chandler, D.A. Dixon, P.J. Roach, T. To, and A.W.L. Wan. 1997. The

buffer/container experiment: results, synthesis, issues. Atomic Energy of Canada Limited Report, AECL-11746, COG-97-46-I.

Hailey, P. and G. Gdowski. 2003. Thermogravimetric thin aqueous film corrosion studies of

Alloy 22; calcium chloride solutions at 150C and atmospheric pressure. In Scientific Basis for Nuclear Waste Management XXVI, R.J. Finch and D.B. Bullen (eds.), Mat. Res. Soc. Symp. Proc. 757 (Materials Research Society, Warrendale, PA), pps. 765-770.

Hedin, A. 2006. Safety function indicators in SKB’s safety assessment of a KBS-3 repository.

In Proc. 11th International High-level Radioactive Waste Management Conference, American Nuclear Society (La Grange Park, IL), pps. 1004-1010.

Hu, Q. and J.S.Y. Wang. 2003. Aqueous-phase diffusion in unsaturated geologic media: a

review. Crit. Rev. Environ. Sci. Technol. 33, 275-297. Karnland, O., T. Sanden, L.-E. Johannesson, T.E. Eriksen, M. Jansson, S. Wold, K. Pedersen,

M. Motamedi, and B. Rosborg. 2000. Long term test of buffer material. Final report on the pilot parcels. Swedish Nuclear Fuel and Waste Management Company Report, SKB TR 00-22.

King, F. 2002. Corrosion of copper in alkaline chloride environments. Swedish Nuclear Fuel

and Waste Management Company Report, TR-02-5. King, F. 2005. Evolution of environmental conditions in a deep geologic repository in the

sedimentary rock of the Michigan Basin, Ontario. Ontario Power Generation Nuclear Waste Management Division Report No: 06819-REP-01300-10102.

King, F. and M. Kolar. 1997. Corrosion of copper containers prior to saturation of a nuclear

fuel waste disposal vault. Atomic Energy of Canada Limited Report, AECL-11718, COG-96-566.

King, F., and M. Kolar. 2000. The copper container corrosion model used in AECL’s second

case study. Ontario Power Generation Nuclear Waste Management Division Report No: 06819-REP-01200-10041.

King, F. and M. Kolar. 2006. Simulation of the consumption of oxygen in long-term in situ

experiments and in the Third Case Study repository using the copper corrosion model CCM-UC.1.1. Ontario Power Generation Nuclear Waste Management Division Report No: 06819-REP-01300-10084.

Page 57: Supporting Technical Report - UNENE

- 49 -

King, F., C.D. Litke, and K.J. George. 1989. The corrosion of copper in synthetic groundwater

at 150°C. Part II. The characterization of surface films. Atomic Energy of Canada Limited Technical Record, TR-464.

King, F., C.D. Litke and S.R. Ryan. 1992. A mechanistic study of the uniform corrosion of

copper in compacted Na-montmorillonite/sand mixtures. Corros. Sci. 33, 1979-1995. King, F., S.R. Ryan and C.D. Litke. 1997. The corrosion of copper in compacted clay. Atomic

Energy of Canada Limited Report, AECL-11831, COG-97-319-I. King, F., L. Ahonen, C. Taxén, U. Vuorinen and L. Werme. 2001. Copper corrosion under

expected conditions in a deep geologic repository. Swedish Nuclear Fuel and Waste Management Company Report, SKB TR 01-23.

King, F., L. Ahonen, C. Taxén, U. Vuorinen and L. Werme. 2002a. Copper corrosion under

expected conditions in a deep geologic repository. Posiva Oy Report, POSIVA 2002-01. King, F., M. Kolar, and S. Stroes-Gascoyne. 2002b. Theory manual for the microbiological

copper corrosion model CCM-MIC.0. Ontario Power Generation Nuclear Waste Management Division Report No: 06819-REP-01200-10091.

King, F., M. Kolar, and S. Stroes-Gascoyne. 2003. Preliminary simulations of the long-term

activity of microbes in a deep geologic repository using CCM-MIC.0 and the implications for corrosion of copper containers. Ontario Power Generation Nuclear Waste Management Division Report No: 06819-REP-01200-10116.

King, F., M. Kolar, S. Stroes-Gascoyne, and P. Maak. 2004. Model for the microbiological

corrosion of copper containers in a deep geologic repository. In Scientific Basis for Nuclear Waste Management XXVII, V.M. Oversby and L.O. Werme (eds.), Mat. Res. Soc. Symp. Proc. 807 (Materials Research Society, Warrendale, PA), pps. 811-816.

Lee, S. and R.W. Staehle. 1997. Adsorption of water on copper, nickel, and iron. Corrosion

53, 33-42. Lemire, R.J. and F. Garisto. 1989. The solubility of U, Np, Pu, Th and Tc in a geological

disposal vault for used nuclear fuel. Atomic Energy of Canada Limited Report, AECL-10009.

Leygraf, C. and T.E. Graedel. 2000. Atmospheric Corrosion. Wiley-Interscience, New York. Litke, C.D., S.R. Ryan and F. King. 1992. A mechanistic study of the uniform corrosion of

copper in compacted clay-sand soil. Atomic Energy of Canada Limited Report, AECL-10397.

Lobnig, R., J.D. Sinclair, M. Unger, and M. Stratmann. 2003. Mechanism of atmospheric

corrosion of copper in the presence of ammonium sulphate particles. Effect of surface particle concentration. J. Electrochem. Soc. 150, A835-A849.

Page 58: Supporting Technical Report - UNENE

- 50 -

Luukkonen, A. 2004. Modelling approach for geochemical changes in the prototype repository engineered barrier system. Posiva Working Report, 2004-31.

Maak, P. 1999. The selection of a corrosion-barrier primary material for used-fuel disposal

containers. Ontario Power Generation, Nuclear Waste Management Division Report 06819-REP-01200-10020-R00.

McMurray, J., D.A. Dixon, J.D. Garroni, B.M. Ikeda, S. Stroes-Gascoyne, P. Baumgartner, and

T.W. Melnyk. 2003. Evolution of a Canadian deep geologic repository: base scenario. Ontario Power Generation, Nuclear Waste Management Division Report 06819-REP-01200-10902-R00.

Nagano, H., T. Doi, and M. Yamashita. 1998. Study on water adsorption-desorption on metal

surfaces and the early stage of atmospheric corrosion in steels. Materials Science Forum 289-292, 127-134.

Nagra. 2002. Project Opalinus Clay. Safety Report. Nagra Technical Report 02-05. Notoya, T. 1991. Localized corrosion in copper tubes and the effect of anti-tarnishing

pretreatment. Journal of Materials Science Letters 10(7). NWMO. 2005. Choosing a way forward. The future management of Canada’s used nuclear

fuel. Final study. Nuclear Waste Management Organization, Toronto, Ontario. Odnevall, I. and C. Leygraf. 1995. Atmospheric corrosion of copper in a rural atmosphere. J.

Electrochem. Soc. 142, 3682-3689. Oscarson, D.W. and D.A. Dixon. 1989. Elemental, mineralogical, and pore-solution

composition of selected Canadian clays. Atomic Energy of Canada Limited Report, AECL-9891.

Quigley, R.M. 1984. Quantitative mineralogy and preliminary pore-water chemistry of

candidate buffer and backfill materials for a nuclear fuel waste disposal vault. Atomic Energy of Canada Limited Report, AECL-7827.

Persson, D. and C. Leygraf. 1990. Vibrational spectroscopy and XPS for atmospheric

corrosion studies on copper. J. Electrochem. Soc. 137, 3163-3169. Pinnel, M.R., H.G. Tompkins and D.E. Heath. 1979. Oxidation of copper in controlled clean air

and standard laboratory air at 50oC and 150oC. Appl. Surf. Sci. 2, 558-577. Rard, J.A., K.J. Staggs, S.D. Day, and S.A. Carroll. 2005. Boiling temperature and reversed

deliquescence relative humidity measurements for mineral assemblages in the NaCl + NaNO3 + KNO3 + Ca(NO3)2 + H2O system. Lawrence Livermore National Laboratory Report, UCRL-JRNL-217704.

Rönnquist, A. and H. Fischmeister. 1960-61. The oxidation of copper. A review of published

data. J. Inst. Metals 89, 65-76.

Page 59: Supporting Technical Report - UNENE

- 51 -

Rosborg, B., O. Karnland, G. Quirk, and L. Werme. 2002. Measurements of copper corrosion in the LOT project at the Äspö hard rock laboratory. In Prediction of Long Term Corrosion Behaviour in Nuclear Waste Systems, D. Féron and D.D. Macdonald (eds.), European Federation of Corrosion Publications Number 36, pps. 412-423.

Rosborg, B., D. Eden, O. Karnland, J. Pan, and L. Werme. 2004. Real-time monitoring of

copper corrosion at the Äspö HRL. In Proc. 2nd Int. Workshop on Prediction of Long Term Corrosion Behaviour in Nuclear Waste Systems, Nice, September 2004, European Federation of Corrosion.

Roy, S.K. and S.C. Sircar. 1981. A critical appraisal of the logarithmic rate law in thin-film

formation during oxidation of copper and its alloys. Oxid. Metals 15, 9-20. Russell, S.B. and G.R. Simmons. 2003. Engineered barrier system for a deep geologic

repository in Canada. In Proc. 10th Int. High-Level Radioactive Waste Management Conf., Las Vegas, NV, March 30-April 2, 2003 (American Nuclear Society, La Grange Park, IL), pp. 563-570.

Sharma, S.P. 1980. Reaction of copper and copper oxide with H2S. J. Electrochem. Soc. 127,

21-26. Shreir, L.L. 1976. Corrosion. 2nd edition, Newnes-Butterworths, London. Strandberg, H. and L.-G. Johansson. 1997. Role of O3 in the atmospheric corrosion of copper

in the presence of SO2. J. Electrochem. Soc. 144, 2334-2342. Strandberg, H. and L.-G. Johansson. 1998. Some aspects of the atmospheric corrosion of

copper in the presence of sodium chloride. J. Electrochem. Soc. 145, 1093-1100. Taxén, C. 2004. Atmospheric corrosion of copper 450 metres underground. Results from

three years exposure in the Äspö HRL. In Scientific Basis for Nuclear Waste Management XXVII, V.M. Oversby and L.O. Werme (eds.), Mat. Res. Soc. Symp. Proc. 807 (Materials Research Society, Warrendale, PA), pps. 423-428.

Tidblad, J. and T.E. Graedel. 1997. GILDES model studies of aqueous chemistry. IV. Initial

(NH4)2SO4-induced atmospheric corrosion of copper. J. Elecrochem. Soc. 144, 2666-2675.

Toyoda, E., M. Watanabe, Y. Higashi, T. Tanaka, Y. Miyata, and T. Ichino. 2004. Depth

profiling analysis of tarnish films on copper formed during short-term outdoor exposure by glow discharge optical emission spectroscopy. Corrosion 60, 729-735.

Weerts, A.H., J.I. Freijer, and W. Bouten. 2000. Modeling the gas diffusion coefficient in

analogy to electrical conductivity using a capillary model. Soil Sci. Soc. Am. J. 64, 527-532.

Xie, M., S. Bauer, O. Kolditz, T. Nowak, and H. Shao. 2006. Numerical simulation of reactive

processes in an experiment with partially saturated bentonite. J. Contaminant Hydrology 83, 122-147.

Page 60: Supporting Technical Report - UNENE

- 52 -

Xu, N., L. Zhao, C. Ding, C. Zhang, R. Li, and Q. Zhong. 2002. Laboratory observation of dew formation at an early stage of atmospheric corrosion of metals. Corros. Sci. 44, 163-170.

Zakipour, S., J. Tidblad, and C. Leygraf. 1995. Atmospheric corrosion effects of SO2 and O3

on laboratory-exposed copper. J. Electrochem. Soc. 142, 757-760. Zhang, J., J. Wang, and Y. Wang. 2005. Micro-droplets formation during the deliquescence of

salt particles in atmosphere. Corrosion 61, 1167-1172. Zhong, Q. 2002. Study of corrosion behaviour of mild steel and copper in thin film salt solution

using the wire beam electrode. Corros. Sci. 44, 909-916.