Return to the Main Document

4. DISCUSSION OF DATA RELATING TO THE POSTCLOSURE SAFETY OF THE SITE

Section 114(a)(1)(C) of the
Nuclear Waste Policy Act of 1982 (NWPA), as amended (42 U.S.C. 10134(a)(1)(C)), requires "a discussion of data, obtained in site characterization activities, relating to the safety of such site." This report presents a summary of the results of site investigations, design studies, and analyses of the performance of a potential repository at Yucca Mountain that began in 1978. Since 1986, the U.S. Department of Energy (DOE) has performed these studies as part of a formal program of site characterization that addresses regulations of the U.S. Nuclear Regulatory Commission (NRC). The program is also reviewed by the Nuclear Waste Technical Review Board, the State of Nevada, affected units of local government, and others. This section presents a summary of the data collected during site characterization as they relate to analyses of the postclosure safety of the site. The discussion is divided into six major parts:

This section emphasizes the scientific and engineering data and analyses related to the safety of the Yucca Mountain site. Most of the detailed scientific and engineering data is presented in Sections 4.2 and 4.3, which describe in detail the subsystem processes and the possible disruptive events that would control the performance of the potential repository. Although many of the concepts and descriptions presented are technically complex, the discussion is, to the extent possible, presented in nontechnical terms, so the information is accessible to non-technical readers.

4.1 THE POSTCLOSURE SAFETY ASSESSMENT METHOD

Assessing how a repository will perform over the next 10,000 years and beyond is a challenge for both the DOE and regulators. The limitations to the analyses and the uncertainties inherent in future system behavior cannot be completely eliminated by further testing or modeling. For this reason, the DOE has adopted an approach that relies on multiple lines of evidence to evaluate whether or not a repository at Yucca Mountain could adequately isolate and contain waste during the compliance period. This approach is documented in the Repository Safety Strategy: Plan to Prepare the Safety Case to Support Yucca Mountain Site Recommendation and Licensing Considerations (
CRWMS M&O 2001a, Volume 2). The postclosure safety case depends on combining sound science and engineering practice with informed judgment and planning. The postclosure safety case is described here because it provides a context for understanding how data and analyses presented throughout the rest of this section are related to safety.

The first element of the postclosure safety case is a thorough and quantitative evaluation of the possible future performance of the repository. This element is based on a comprehensive testing program that has evolved to address identified uncertainties, and an engineered barrier design developed specifically to work in combination with the natural barriers of the site. U.S. Environmental Protection Agency (EPA) and NRC regulations specify the method by which the DOE will analyze whether a repository can safely isolate spent nuclear fuel and high-level radioactive waste (i.e., a TSPA). The TSPA is described briefly in Section 4.4 but in more detail in Total System Performance Assessment for the Site Recommendation (CRWMS M&O 2000a) and several subsequent documents including FY01 Supplemental Science and Performance Analyses (BSC 2001a; BSC 2001b) and Total System Performance Assessment—Analyses for Disposal of Commercial and DOE Waste Inventories at Yucca Mountain—Input to Final Environmental Impact Statement and Site Suitability Evaluation (Williams 2001a). These reports include analyses of all the processes expected to operate at the repository that could affect its ability to isolate waste. They also explicitly consider both disruptive events and alternative process models that could result in unanticipated behavior (i.e., identifies what could go wrong). These evaluations directly address uncertainty in both the DOE's knowledge of the site and in future conditions and include numerical sensitivity analyses to test how the repository might perform if current or future conditions differ from those expected.

To capture the technical inputs used in developing the overall TSPA system level model, a set of analysis model reports have been prepared. These reports contain the detailed technical information regarding data, analyses, models, software, and supporting documentation that is used in the development of the process models. The analysis model reports provide the direct input into the TSPA analyses, as well as document the abstraction of the process level models for use in the overall TSPA system level model.

Using the analysis model reports as a basis, the descriptions of these process level models are documented in a suite of process model reports that cover the following areas:

The process model reports synthesize the information contained in the individual analysis model reports and provide an integrated perspective for understanding each of the process level models. This hierarchical process of documentation was used to ensure the traceability of supporting information from its source through the analysis model reports and process model reports to its eventual use in the TSPA. Supplemental analyses at both the process model level, and the total system level, are presented in FY01 Supplemental Science and Performance Analyses (BSC 2001a; BSC 2001b).

Because the DOE recognizes that uncertainty about the future performance of the repository cannot be completely eliminated, the postclosure safety case includes several additional measures designed to provide confidence and assurance that the repository will meet postclosure performance standards. These measures include:

This approach is similar to that recommended by many national and international professional organizations that have studied nuclear waste disposal. As a panel of the National Academy of Sciences observed, "Confidence in the disposal techniques must come from a combination of remoteness, engineering design, mathematical modeling, performance assessment, natural analogues, and the possibility of remedial action in the event of unforeseen events" (National Research Council 1990, pp. 5 to 6).

The various elements of the postclosure safety case contribute in different ways to building confidence in analyses of the long-term performance of the repository. Quantitative numerical models permit scientists to test their understanding of the site and assess the consequences of uncertainty or assumptions in their models. Observations of natural or man-made analogues can help scientists determine whether the results of repository models are consistent with the behavior of actual systems. They can also be used to qualitatively evaluate the reliability and uncertainties associated with modeling. Safety margin and defense in depth provide one way to compensate for uncertainty in analyses. Long-term monitoring can help scientists verify that the uncertainties in site and design performance have been appropriately characterized. In total, the elements documented in the Repository Safety Strategy: Plan to Prepare the Safety Case to Support Yucca Mountain Site Recommendation and Licensing Considerations (CRWMS M&O 2001a, Volume 2) support the postclosure safety case for the Yucca Mountain site. The sections that follow contain additional descriptions of each element of the safety case.

4.1.1 Total System Performance Assessment

Analysis of the future performance of the potential repository is fundamental to the DOE's understanding of the Yucca Mountain site. Therefore, the first element of the safety case is a thorough analysis of how a repository at Yucca Mountain would behave in the future. As noted previously, the methods used and the results of the TSPA for Yucca Mountain are described in
Sections 4.3 and 4.4.

Performance assessment is a method or tool defined and provided by the EPA and the NRC (in 40 CFR Part 197 and 10 CFR Part 63 [66 FR 55732], respectively) for the evaluation of a Yucca Mountain repository. The objective of the total system performance assessment for site recommendation (TSPA-SR) for Yucca Mountain is to provide a basis for evaluating whether the safety of the general public will be protected. However, the DOE has also used the performance assessment for broader purposes during site characterization of Yucca Mountain. For instance, it has been used as a tool to evaluate the effects of uncertainty on total system performance and to identify areas where further work is needed. This has been accomplished in an iterative manner. For this updated Yucca Mountain Science and Engineering Report, TSPA results are presented and discussed in one comprehensive report, summarizing several additional supplemental documents that describe analyses performed to address specific technical and/or regulatory issues. The key TSPA references include:

Figure 4-1 schematically presents the information flow within an iterative TSPA approach. Information gained during studies and testing flows upward through the development of conceptual and numerical models that provide the basis for the TSPA. Information gained through analysis flows downward in the form of specific information needs for the next iteration of the TSPA.

The foundations of the TSPA are site characterization and engineering design data. Project scientists and engineers use this information to formulate conceptual models of the FEPs that could affect the performance of the natural and engineered barriers at the site. An important step in the formulation of these conceptual models is the identification of uncertainties in the current state of knowledge. These conceptual models, the level of uncertainty associated with each, and the essential assumptions used in their formulation are documented in Sections 4.2 and 4.3, as well as supporting documents describing the DOE's understanding of the Yucca Mountain repository system.

The conceptual models are next cast into process-level models. These are typically numerical computer models that range from simple to quite complex representations designed to capture and simulate the fundamental physical phenomena that influence the process being modeled. The process models, together with important uncertainties and assumptions, are also described in Sections 4.2 and 4.3, as well as in supporting references.

Many process level models are needed to analyze the various subsystems that could affect the performance of the potential repository. Some of the individual models are so complex that it is not possible or desirable to include them in a single linked total system model, due to computing limits or because simplified models of certain processes may be equally defensible. A total system model that depended on a single representation of a process might overlook credible alternative models. Therefore, the total system model is generally based on simplified (abstracted) models that enable assessment of the effects of alternative representations of potentially important processes.

The process level models described in Sections 4.2 and 4.3 provide the foundation for the abstracted models contained in the TSPA described in Section 4.4. These abstracted models include the important details of the process level models, so they can be used to simulate or bound the results of the process level models. Scientists and engineers evaluate the output of the detailed process level models to identify key results, uncertainties, and assumptions that must be captured by the abstracted models. For example, it may be determined that, of the many processes and parameters contained within a process level model, only a few have a significant effect on overall behavior. Analysts use test results, comparisons with alternative process level models, and judgment to determine how best to incorporate the uncertainty associated with each specific process in the abstracted models. The complexity of an abstracted model is governed by how sensitive total system performance is to the specific process and by how well the model incorporates uncertainty. The abstracted TSPA models, associated uncertainties, and important assumptions are described in Section 4.4 and in Total System Performance Assessment for the Site Recommendation (CRWMS M&O 2000a). More recent analyses are described in FY01 Supplemental Science and Performance Analyses (BSC 2001a; BSC 2001b) and Total System Performance Assessment—Analyses for Disposal of Commercial and DOE Waste Inventories at Yucca Mountain—Input to Final Environmental Impact Statement and Site Suitability Evaluation (Williams 2001a).

The abstracted models are combined into a total system model, which is used both to assess the potential future performance of the repository and to evaluate how uncertainty in the understanding of FEPs might affect performance. Throughout site characterization, the DOE has used this information to identify and prioritize future work activities. In this manner, the testing program has been updated and the repository design modified to continually improve the DOE's confidence in assessments of future performance. In the past decade, the DOE completed comprehensive analyses of total system performance in 1991, 1993, 1995 and 1998 (Barnard et al. 1992; Eslinger et al. 1993; Wilson, M.L. et al. 1994; CRWMS M&O 1995; DOE 1998, Volume 3). Each of these represented a significant advance in the DOE's understanding of how the potential Yucca Mountain repository might perform, and each resulted in modifications to the site testing program or the repository design to address key uncertainties. This report and the Total System Performance Assessment for the Site Recommendation (CRWMS M&O 2000a) reflect the knowledge and insights gained through this process. Additional analyses and insight are presented in FY01 Supplemental Science and Performance Analyses (BSC 2001a; BSC 2001b) and Total System Performance Assessment—Analyses for Disposal of Commercial and DOE Waste Inventories at Yucca Mountain—Input to Final Environmental Impact Statement and Site Suitability Evaluation (Williams 2001a).

4.1.1.1 Total System Performance Assessment Methods and Objectives

TSPA is a systematic analysis that synthesizes information (data, analyses, and expert judgment) about the site and region with the design attributes of the engineered barriers of the repository system. As defined in 10 CFR 63.2 (66 FR 55732),

Performance assessment means an analysis that:

  1. Identifies the features, events and processes (except human intrusion) and sequences of events and processes (except human intrusion) that might affect the Yucca Mountain disposal system and their probabilities of occurring during 10,000 years after disposal;

  2. Examines the effects of those features, events, and processes and sequence of events and processes upon the performance of the Yucca Mountain disposal system; and

  3. Estimates the dose incurred by the reasonably maximally exposed individual, including the associated uncertainties, as a result of releases caused by all significant features, events, and processes, and sequences of events and processes weighted by their probability of occurrence.

Features are the physical components of the total repository system, including both the natural system (e.g., the geologic setting) and the engineered system (e.g., the waste package). Processes typically act more or less continuously on the features; for example, moisture flow through the geologic materials and corrosion of the waste package. Events also act on the features but at discrete times. Examples include seismic and volcanic events.

The TSPA approach and models are designed to address the processes that could lead to release and migration of radionuclides, and the radiological consequences to potential human receptors. The approach is intended to provide a transparent analysis of the geologic repository in terms of the performance of the natural and engineered barriers over long periods of time.

40 CFR Part 197 provides that the DOE and NRC should determine compliance with the radiation protection standard of 40 CFR 197.20 based on the mean of the distribution of the highest doses resulting from the performance assessment. In the background information accompanying their final rule (66 FR 32074, p. 32125), the EPA noted that they believe that a thorough assessment of repository performance should examine the full range of reasonably foreseeable conditions and processes. However, they also stated that quantitative estimates of repository performance should not be dominated by unrealistic or extreme situations or assumptions. Therefore, the EPA believed the use of the mean was reasonable but still conservative. They further noted that the use of the mean was consistent with the literal mathematical interpretation of the term "reasonable expectation" and with the approach used to certify Waste Isolation Pilot Plant.

During their consideration of the appropriate performance measure, the EPA evaluated other possible measures, such as the median value of the distribution, or more extreme measures, such as the 95th or 99th percentile. Their analysis showed that the use of either the mean or the median was reasonably conservative because both are influenced by the high exposure estimates, without reflecting only the high dose results.

Although the EPA selected the mean for the compliance determination, both the EPA and NRC provide that the DOE consider the uncertainties inherent in performance assessment results. One way that the DOE's method addresses this concern is by presenting and analyzing the full range of doses resulting from the performance assessment. In addition, the DOE has performed numerous sensitivity and uncertainty analyses to characterize the properties and processes that are particularly important to dose calculations.

The TSPA described briefly in Section 4.4 and in more detail in Total System Performance Assessment for the Site Recommendation (CRWMS M&O 2000a), Volume 2 of FY01 Supplemental Science and Performance Analyses (BSC 2001b), and Total System Performance Assessment—Analyses for Disposal of Commercial and DOE Waste Inventories at Yucca Mountain—Input to Final Environmental Impact Statement and Site Suitability Evaluation (Williams 2001a) examines the performance of the potential repository for a broad range of potential subsurface and surface conditions (e.g., hydrologic, geologic, climatic, and biosphere) and evaluates potential radiation doses to future generations. The radiation protection standards that would apply to the postclosure performance of a Yucca Mountain repository are found in regulations promulgated by the EPA (40 CFR 197.20, 197.25, and 197.30) and the NRC (10 CFR 63.113(b), (c), and (d) [66 FR 55732]). Specific technical requirements for a comprehensive TSPA are prescribed in the NRC regulation, 10 CFR Part 63.

The TSPA method requires evaluation of postclosure performance of the Yucca Mountain disposal system where there is no human intrusion into the repository. The final regulations also require evaluation of the performance of the system where there is human intrusion, in accordance with NRC regulations. The TSPA evaluation method in both cases is the same, except that the TSPA method for human intrusion provides prescribed assumptions about the human intrusion scenario (10 CFR 63.322 [66 FR 55732]).

Human intrusion refers to inadvertent intrusion into the repository as a result of exploratory drilling for groundwater. Limited intrusion means a single borehole that penetrates the repository and the underlying groundwater aquifer.

To present the assessment results transparently, the first case (without human intrusion) is further subdivided into:

The result of a TSPA analysis is a distribution (range) of possible outcomes of future performance. Because of the probabilistic nature of the method, TSPA results can capture and display much of the uncertainty associated with complex models and unknown future conditions. For this reason, however, the results should be regarded as indicative of future performance, not as predictive in a precise way.

The TSPA methodology (i.e., approach and models) described in this report is the culmination of research and development conducted over more than a decade. Moreover, reviews of previous TSPAs by internal and external professional organizations have been invaluable in enhancing the rigor of the approach and guiding the improvements of the models. Most notably, important advances in the TSPA methodology have been made in response to review comments from:

The TSPA methodology used for this report is very similar to that used in the compliance certification application for the Waste Isolation Pilot Plant (DOE 1996a, Section 6.1; Helton, Anderson et al. 1999), a bedded salt repository in southern New Mexico. The Waste Isolation Pilot Plant was certified by the EPA in 1998 and began receiving and disposing of transuranic nuclear waste in March 1999. The TSPA approach is also similar to approaches adopted by other countries currently conducting detailed siting studies for potential geologic repositories (NEA 1991; Thompson 1999). In addition, the computer techniques used in TSPAs to address uncertainties are rooted in the probabilistic risk assessment method applied in the safety assessments for commercial nuclear reactors (Rechard 1999).

4.1.1.2 Treatment of Uncertainty in the Performance Assessment

Inherent uncertainties will exist in any projections of the future performance of a deep geologic repository. These uncertainties must be addressed in a way that is both clear and understandable to ensure technical credibility and sound decision-making and must be reduced or eliminated if important. Most, but not all, of those uncertainties can be quantified and addressed in the TSPA; examples include:

  • Potential changes in climate, seismicity, and other processes, such as coupled thermal-hydrologic-chemical processes, over the compliance period for geologic disposal (i.e., 10,000 years)

  • Variability and lack of knowledge of the properties of geologic media over large spatial scales of the hydrogeologic setting (e.g., the flow path from the repository to a point of compliance)

  • Incomplete knowledge about the long-term material behavior of engineered components (e.g., corrosion of metals over many thousands of years).

Both the EPA and the NRC have recognized that uncertainty about the future performance of the repository will remain even after site characterization is complete. In the licensing context, both EPA regulations at
40 CFR 197.14 and NRC regulations at 10 CFR 63.304 (66 FR 55732) have incorporated "reasonable expectation" as the standard for the NRC to determine whether the DOE complies with EPA and NRC regulations. Characteristics of reasonable expectation include that it:

There are a number of ways to accommodate or address uncertainty in analyses of performance. The methods employed for the TSPA models include:

Whether uncertainties are incorporated quantitatively through probabilistic distributions, or by developing "conservative" or "bounding" estimates, it is important to define and document assumptions on how uncertainties are treated in performance analyses. Clear documentation of the rationale for the assumptions in the models will enable others to understand and evaluate their adequacy. For the total system model to be transparent and defensible, the selection of each probability distribution must also be defensible. In some instances, the insufficient information that exists on the subsystem model is so complex that a probability distribution cannot be defined defensibly. In these instances, conservative or bounding approaches are taken.

Individual sources of uncertainties may be either quantified or unquantified. Quantified uncertainty consists of those sources of uncertainties that can be and have been explicitly (i.e., mathematically) represented and evaluated through a probabilistic performance analysis. Unquantified uncertainties are those that are recognized but are not well suited for direct evaluation through a probabilistic analysis.

As described in Section 4.4.1.2, many uncertainties have been quantified and incorporated directly into the TSPA models. The analyses done to address quantified uncertainties in the TSPA-SR model include a variety of sensitivity studies that address how total system performance might be affected if individual or groups of processes or parameters differed significantly from the way they were represented in the TSPA-SR model. These are described further in Section 4.4.5 and in Total System Performance Assessment for the Site Recommendation (CRWMS M&O 2000a). Additional uncertainties have been identified, analyzed, and quantified in FY01 Supplemental Science and Performance Analyses (BSC 2001a; BSC 2001b). These analyses are described, as appropriate, throughout Section 4.

As noted by the National Research Council (1990, p. 13) and others, there are residual uncertainties with deep geologic disposal that cannot easily be quantified and incorporated into performance analyses. Nevertheless, their potential impact must be, to the extent practicable, addressed and, if important, mitigated to provide confidence in postclosure performance. Examples of residual uncertainties associated with geologic disposal that are difficult to quantify include:

The DOE has made a substantial effort to identify, characterize, and mitigate the potential impacts of residual uncertainties that could significantly affect long-term performance. Where practicable, additional tests have been conducted to collect information that would provide insight to analysts. For example, the DOE, along with Nye County and the National Park Service, is continuing research on the regional groundwater system. The program includes cooperative and joint analysis of soil, rock, and water samples provided by Nye County, which is drilling additional boreholes to characterize the saturated zone as part of its review of the project.

To address uncertainties associated with coupled thermal-hydrologic-geochemical processes, the DOE has performed numerous tests that were not envisioned when site characterization began. Where additional testing was not feasible (e.g., it is not possible to run tests over the same time period as the repository must perform) or of limited benefit (e.g., no amount of excavation or drilling could completely characterize the natural system), modelers used conservative assumptions to "bound" their analyses of uncertain processes. To do this, they have incorporated assumptions in their models that represent the observed range of properties and processes and that also include parameter values or processes likely to result in calculated dose assessments that are greater than actual dose. The DOE has also used empirical observations and qualitative lines of evidence from natural analogues to address uncertainties (see Section 4.1.2).

In some cases, more than one conceptual model may be consistent with available data and observations; if so, the analysis of uncertainty includes the identification of the basis for model selection. In the absence of definitive data sets or compelling technical arguments for any specific conceptual model, analyses or sensitivity studies may be performed to test whether the selection of a given model is likely to substantially affect analyses of system performance. When model uncertainty is unavoidable, analysts may develop simplified performance models for use in TSPA that reflect the range of outcomes predicted by more detailed and specific conceptual models. The selection of models, model parameters, and scenarios of potential future behavior are described in Total System Performance Assessment for the Site Recommendation (CRWMS M&O 2000a) and its supporting documents, as well as in FY01 Supplemental Science and Performance Analyses (BSC 2001a; BSC 2001b) and Total System Performance Assessment—Analyses for Disposal of Commercial and DOE Waste Inventories at Yucca Mountain—Input to Final Environmental Impact Statement and Site Suitability Evaluation (Williams 2001a).

In addition, consistent with the postclosure safety case described in Section 4.1.3, the DOE has implemented a safety margin/defense-in-depth approach to the design of repository facilities. This approach mitigates some of the residual uncertainties by providing additional confidence in the performance of the total system. The iterative approach to characterization and testing, design, and performance analysis remains a fundamental part of the DOE approach in evaluating whether Yucca Mountain is a safe site to host a repository.

One disadvantage of an approach that combines both conservative and representative parameter distributions is that it is difficult to assess the extent to which the total system results are conservative or realistic. More realistic representations may be drawn from literature data, analogue systems or processes, and the technical judgment of the broader scientific and engineering community. In addition, the DOE initiated several activities to improve the treatment of uncertainty in current models. These are described in FY01 Supplemental Science and Performance Analyses (BSC 2001a; BSC 2001b). These activities:

This approach provides recommendations for consistent treatment of uncertainties, which will lead to a more rigorous treatment of the different types of uncertainty. It will also lead to a more transparent description of uncertainties in the process models and in the overall system models.

4.1.1.3 Explicit Consideration of Disruptive Processes and Events that Could Affect Repository Performance

Most of this report, as well as the Total System Performance Assessment for the Site Recommendation (CRWMS M&O 2000a), the FY01 Supplemental Science and Performance Analyses (BSC 2001a; BSC 2001b), the Total System Performance Assessment—Analyses for Disposal of Commercial and DOE Waste Inventories at Yucca Mountain—Input to Final Environmental Impact Statement and Site Suitability Evaluation (Williams 2001a), and associated documentation, focuses on the processes that are expected to operate in and near the repository over time (e.g., water flow, the effects of heat on conditions near waste packages, and the slow degradation of the engineered barrier system). Because of the time frames involved, however, it is also important for analysts to understand the safety consequences of unexpected events or behavior. Consequently, the performance analysis also describes explicitly how disruptive events (i.e., possible but unlikely events that could negatively affect performance) and alternative models of processes could affect the performance of the total system. This thorough consideration of what could go wrong, how wrong the models could be, and what the effect of inaccuracy in the models would be is a key element of the postclosure safety case.

A comprehensive set of potentially disruptive events, ranging from meteor or comet impacts to unexpected flooding of the repository, has been identified and evaluated. Similarly, a wide variety of potentially harmful processes, such as unanticipated failure of the waste packages and damage to repository systems by seismic activity, have been identified and evaluated. Section 4.3 describes the identification and screening of these potentially adverse processes and events. Depending on the results of this analysis, the events and processes have been treated in one of three ways:

  1. Events or processes with very low probabilities (e.g., less than 1 chance in 10,000 of occurring in 10,000 years, such as meteor impacts) or very small consequences (e.g., potential radionuclide releases much lower than regulatory standards) have been screened out and not analyzed further.

  2. Events or processes that have probabilities greater than 1 chance in 10,000 of occurring in 10,000 years (e.g., intrusive igneous activity), have been included in the performance assessment probabilistically—that is, the consequence of each event is weighted by its probability of occurrence.

  3. Events or processes expected to occur during the period of regulatory compliance, such as climate change or ground shaking associated with earthquakes, are included directly in models of the performance of the repository.

The potential consequences of each event and process considered in the performance assessment models are analyzed. These analyses are summarized in Section 4.2, which considers alternative models of subsystem processes that could affect performance, and Section 4.3, which considers specific events or processes that are not part of nominal site behavior. Section 4.4.5 also describes a number of sensitivity studies that enable analysts to assess the importance of alternative models or unexpected events on total system performance.

Many different events and processes have been considered to assess whether they might have the potential to disrupt repository performance. Examples of potential events considered include:

Climate change and cladding damage caused by earthquake ground motion is directly incorporated into the nominal case performance analyses (CRWMS M&O 2000c) because climate change and seismic activity are expected to occur during the period of regulatory compliance.

Many other events and processes have been considered but screened out because their impacts are so inconsequential that inclusion in the TSPA-SR is not warranted. Examples include the potential for future rises in the water table that could inundate the repository and the potential for nuclear criticality in spent nuclear fuel after repository closure. These examples are discussed in Section 4.3.

It is not possible to foresee every event or process that could affect the potential repository. However, by recognizing and explicitly analyzing all identified events and processes that could affect repository performance, the DOE has provided a sound basis for evaluating the performance of the repository system.

4.1.2 Observations from Natural and Man-Made Analogues

An alternative means of analyzing the reliability of repository performance models is by comparing them with natural or anthropogenic (man-made) analogues. As defined in this report, analogues are systems in which processes similar to those that could occur in a nuclear waste repository have occurred over long time periods (decades to millennia) and large spatial scales (up to tens of kilometers) not suited to laboratory or field experiments. The concept of geologic disposal is based on an analogue observation that certain geologic environments have intrinsic properties that contribute to the isolation of waste and will continue to do so for a long time. Analogue studies can also help scientists understand how specific repository systems might behave by allowing them to compare possible future behavior with the known past behavior of an analogue system. A variety of specific analogue sites are described throughout
Section 4.2, including several in settings that provide insight into the flow of water in the unsaturated zone, as described in Section 4.2.1.2.13.

For example, observations of the hydrologic behavior of ancient man-made tunnels and natural caves can provide relevant information about water seepage into mined openings in an unsaturated zone over thousands of years. Similarly, observations of the past migration of radioactive contaminants in groundwater in similar environments can provide insight into the possible future transport of radionuclides away from a repository. Such information can be obtained from both natural analogues (e.g., natural deposits of uranium and other minerals) or from anthropogenic analogues (e.g., the Nevada Test Site, Hanford, or the Idaho National Engineering and Environmental Laboratory), where the movement of radionuclides in groundwater caused by past releases is currently being monitored. The archaeological and historical record can also provide qualitative information on the degradation of materials that may be relevant to the performance of the repository (e.g., the preservation of materials in Egyptian pyramids and tombs more than 5,000 years old).

The value of natural analogues is not restricted to locations that may mimic aspects of repository behavior. The study of natural analogues is an intrinsic part of scientific studies, particularly in the earth sciences. For example, in order to undertake studies of basaltic volcanism in the vicinity of the Yucca Mountain site, an investigator must be versed in basaltic volcanism and especially in basaltic volcanism in the southern Great Basin. All other known occurrences of basaltic volcanism become, to some degree, natural analogues for volcanism near Yucca Mountain. Each occurrence can tell the investigator something about the mechanisms and controls on volcanic episodes. The overall understanding of volcanic processes gained from regional studies provides a basis for analyzing trends and patterns in the Yucca Mountain area that are essential to evaluating the possibility that future volcanic events could occur and affect a repository. In fact, it would be impossible to understand basaltic volcanism as a site characterization issue without recourse to natural analogues. One of the fundamental tasks for the investigator is to recognize and appreciate the value of the information offered by this kind of analogue study.

The scientific community has endorsed the use of analogues as a means of assessing the potential future performance of systems, components, and processes related to nuclear waste disposal (Chapman and Smellie 1986). The National Academy of Sciences/National Research Council (National Research Council 1990) and the NRC (10 CFR 63.101(a)(2) [66 FR 55732]) have also encouraged analogue studies.

There are no close analogues to a total repository system at Yucca Mountain. Nevertheless, studies of a variety of analogues can and have been used to assess how well repository models represent processes known to be important to performance, as well as the magnitude and duration of the phenomena. Analogue information has also been used (1) to evaluate the validity of extrapolating from short-term field-scale experiments to longer time scales in which field-scale experiments are impractical and (2) to add confidence when spatially extrapolating processes studied at laboratory and intermediate-scale experiments to tests at larger spatial scales. Knowledge gained from natural analogues has helped refine performance assessment model assumptions and parameter ranges and has improved the robustness and consistency of process models.

Given the imprecise nature of the information gained from investigations of similar, but not identical, processes and sites, analogue studies alone cannot prove that process or total system performance models are valid in a strict sense. However, natural analogue observations can confirm that a model takes into account the relevant processes in appropriate ways. In this way, the analogues can build confidence in models of future behavior. This is consistent with the expectations of NRC regulations in 10 CFR 63.101(a)(2) (66 FR 55732), which state: "Demonstrating compliance will involve the use of complex predictive models that are supported by limited data from field and laboratory tests, site-specific monitoring, and natural analog studies that may be supplemented by prevalent expert judgment."

Throughout this report and its supporting documents, numerous analogues are analyzed to provide information on processes that may affect both engineered and natural system features of a geologic repository at Yucca Mountain. Specific examples of relevant analogues are presented in Table 4-1. Additional discussion of analogues is provided as appropriate throughout Sections 4.2 and 4.3, which discuss in greater detail the understanding of the Yucca Mountain site. Although the direct applicability of the analogues for each process model varies, the analogue observations generally suggest that the conceptual and numerical models that form the basis for analyses of repository performance are reasonable to conservative. For many of the analogues, a large body of literature exists.

4.1.3 Use of Defense in Depth and Safety Margin to Increase Confidence in System Performance

The extensive testing program at Yucca Mountain and the thorough assessments of the future performance of the potential repository do not "prove," in the usual sense of the word, that the potential repository will be safe. To provide additional assurance of long-term safety, the third major element of the postclosure safety case relies on a complementary, but less analytical, approach that is based on engineering principles that have a proven track record for safety. These principles are known as "safety margin" (or design margin) and "defense in depth."

Safety margin refers to the standard engineering practice of including safety factors on the performance of engineered components to account for uncertainty and variability in material, fabrication, and use. These safety factors are typically expressed as a ratio of the calculated level of performance to an allowable or laboratory measured level of performance. They are developed to ensure that the component or system has ample reserve performance capability. A simple example of a safety factor would be the limiting of the stress in a component to a fraction of what would cause failure. The safety margin is then the reserve strength over and above what would actually be applied. A Yucca Mountain-specific example of safety margin is the design and selection of the waste package material. The assumed corrosion resistance of Alloy 22, used in the outer shell of the waste packages, was decreased (or the corrosion rate was increased) over the laboratory values to account for potential environmental conditions. The corrosion rate of Alloy 22 was increased by 2.5 times to address potential heat-accelerated corrosion and an additional 2.0 times to address potential microbial corrosion.

The defense-in-depth approach (or reliance on multiple system attributes) complements design margin in that it provides a method of ensuring overall performance if one or more components of the repository system fail to perform as expected. Defense in depth is provided by having safety components that do not share common failure modes. In other words, the processes or conditions (such as the geochemical environment) that might cause a degradation of performance of one component of the design will not similarly affect other components. In a repository, defense in depth is provided by the attributes of both the natural barriers and the engineered barrier system.

The safety margin/defense-in-depth approach is not specifically required by the regulations for a repository. However, the NRC regulation (10 CFR Part 63 [
66 FR 55732]) does contain statements that are based on a similar philosophy, adapted to the long-term requirements for postclosure safety. 10 CFR 63.113(a) provides that a repository include multiple barriers, consisting of both natural barriers and an engineered barrier system.

At Yucca Mountain, the potential geologic repository system would contain several different barriers. As defined in 10 CFR 63.2 (66 FR 55732), a barrier is any material, structure, or feature that prevents or substantially reduces the rate of movement of water or radioactive material from the Yucca Mountain repository to the accessible environment. Table 4-2 presents a summary of the natural and engineered barriers present at Yucca Mountain, along with a brief description of their intrinsic and intended functions for the design analyzed in this section. An analysis of each barrier's contribution to performance is presented in Section 4.5.

Implementation of the safety margin/defense-in-depth approach has resulted in several improvements in the repository design since the design described in 1998 in the Viability Assessment (VA) (DOE 1998). Examples include:

More detailed descriptions of the barriers and their performance functions are summarized in Section 4.5. Additional information is available in Volume 2 of Repository Safety Strategy: Plan to Prepare the Safety Case to Support Yucca Mountain Site Recommendation and Licensing Considerations (CRWMS M&O 2001a) and Total System Performance Assessment for the Site Recommendation (CRWMS M&O 2000a) as well as FY01 Supplemental Science and Performance Analyses (BSC 2001a; BSC 2001b) and Total System Performance Assessment—Analyses for Disposal of Commercial and DOE Waste Inventories at Yucca Mountain—Input to Final Environmental Impact Statement and Site Suitability Evaluation (Williams 2001a).

4.1.4 Mitigation of Uncertainties by Selection of a Thermal Operating Mode

One way of mitigating the uncertainties in modeling long-term repository performance is to operate the repository so the temperature of the host rock stays below the boiling point of water. Uncertainties in thermally driven processes are of special interest because of their complexity and because the current modeling approach may mask the importance of thermal effects on performance. Two key uncertainties about thermal effects on potential repository performance are (1) the way coupled processes in the mountain will respond to the heat generated by emplaced waste and (2) the long-term performance of waste package materials in the potential repository environment.

In the models and design described in this report, uncertainties related to the higher-temperature operating mode have been recognized and addressed. Since the VA, the design has evolved to include a thermal management strategy that limits the region of rock with temperatures above the boiling point of water, along with other features (such as drip shields) that mitigates uncertainty. Current models attempt to capture the remaining uncertainties well enough to understand their impacts; however, the DOE has considered additional options for mitigation. In particular, the performance characteristics of lower-temperature operating mode concepts such as those described in
Section 2.1 have been investigated in FY01 Supplemental Science and Performance Analyses (BSC 2001a; BSC 2001b) and Total System Performance Assessment—Analyses for Disposal of Commercial and DOE Waste Inventories at Yucca Mountain—Input to Final Environmental Impact Statement and Site Suitability Evaluation (Williams 2001a).

Because uncertainties due to thermally induced coupled processes cannot be eliminated through additional testing, the DOE's approach is to consider options for mitigating thermal uncertainties by lowering temperatures in the emplacement drifts and on the waste package surface. Keeping the host rock temperature below boiling may reduce uncertainties associated with coupled processes (Anderson et al. 1998, p. 18; Cohon 1999). In addition, uncertainties in localized corrosion rates may be mitigated by avoiding the conservatively defined window of increased susceptibility by keeping the temperature of the waste package below 85°C (185°F) or the relative humidity in the emplacement drifts below 50 percent. Section 2.1 provides additional discussions of lower-temperature operating modes.

4.1.5 Performance Confirmation, Postclosure Monitoring, and Site Stewardship

The EPA, NRC, and DOE have recognized that some uncertainty about repository performance cannot be eliminated. Furthermore, the DOE understands that ensuring public safety requires continued site stewardship, including a program for evaluating new information discovered during the construction and operation phases. Therefore, the final component of the postclosure safety case is a program of performance confirmation, monitoring, and site stewardship that accomplishes several goals related to the DOE's obligation to protect public health and safety and the environment. The program addresses 10 CFR Part 63 (
66 FR 55732) Subpart F provisions for performance confirmation to ensure consistency with license specifications after waste emplacement and before permanent closure. The program also includes activities necessary for the DOE to provide postclosure oversight, as specified by Section 801(c) of the Energy Policy Act of 1992 (Public Law No. 102-486), and post-permanent closure monitoring, consistent with 10 CFR 63.51(a)(2). Specifically, the DOE will continue to observe and test the performance of the repository during and after waste emplacement and will maintain the integrity and security of the repository through a variety of institutional controls. The DOE will also continue to participate in research on geologic disposal, in cooperation with other international programs. These activities will ensure that any new information discovered at Yucca Mountain (or elsewhere) that is relevant to future repository decisions is considered appropriately.

The performance confirmation program is the most important monitoring activity. NRC regulations provide for performance confirmation to continue for at least 50 years after the initiation of waste emplacement. The DOE will continue its performance confirmation program until the repository is permanently closed. The amount of information collected during this period may be more relevant for long-term analyses of the repository than any experiment or test that could be conducted now or in the near future. Any decision to close the repository would be based on the increased understanding and confidence derived from decades of testing and observation.

The performance confirmation program will provide data on the actual performance of the key natural and engineered systems and components of the repository as conditions evolve. The program will also provide data to confirm (after repository construction and operation) that subsurface conditions encountered, and any changes in those conditions during repository construction and waste emplacement, are consistent with the expected performance of the repository. A primary goal of the program will be to confirm, through observation, monitoring, and analysis, that the repository is performing in a manner that will contain and isolate waste.

As described in Section 4.6, the performance confirmation program will monitor the processes important to future waste isolation in the repository. Examples include the flow of water past and near the repository, the geochemical environment in and near emplaced waste, coupled thermal-hydrologic-chemical-mechanical processes, and the performance characteristics of engineered materials in the repository environment (e.g., drip shield and waste package degradation). In addition to technical monitoring of the performance of the site, the DOE will maintain the security and integrity of the site throughout the performance confirmation period and beyond as required by the Energy Policy Act of 1992, Section 801(c) (Public Law No. 102-486). A program will be developed to prevent any human activity, including deliberate or inadvertent human intrusion, that could affect engineered or geologic barriers.

Section 122 of the NWPA (42 U.S.C. 10142) requires the DOE to maintain the ability to retrieve any and all emplaced wastes "for any reason pertaining to the public health and safety, or the environment, or for the purpose of permitting the recovery of the economically valuable contents of such spent fuel" prior to closure. For example, if it were learned from the monitoring program that an engineered barrier had been damaged, the waste packages could be removed and repairs could be made, as necessary.

NRC regulations (10 CFR Part 63 [66 FR 55732]) anticipate that the repository could be closed as early as 50 years after initial waste receipt. Closing the repository would involve the sealing of shafts, ramps, exploratory boreholes, and other underground openings. These actions would discourage any human intrusion into the repository and prevent water from entering through these openings. If a decision to close the repository were made, the DOE would still be responsible for a program of site stewardship under the Energy Policy Act of 1992 (Public Law No. 102-486).

At the surface, all radiological areas would be decontaminated, all structures removed, and all wastes and debris disposed at approved sites. All disturbed areas would be restored as close as practicable to their preconstruction condition.

NRC regulations (10 CFR Part 63 [66 FR 55732]) require the DOE to submit a plan for postclosure monitoring with any application to close the repository. The DOE has also committed to maintain security and continue monitoring at the Nevada Test Site for the foreseeable future. A network of permanent monuments and markers would be erected around the site to warn future generations of the presence and nature of the buried waste. Detailed records of the repository would be placed in the archives and land records of local, state, and federal government agencies and archives elsewhere in the world that future generations would be likely to consult. These records would identify the location and layout of the repository and the nature and hazard of the waste it contains.

4.2 DESCRIPTION OF SITE CHARACTERIZATION DATA AND ANALYSES RELATED TO POSTCLOSURE SAFETY

This section discusses the data obtained during site characterization activities, as well as analyses of the safety of a potential Yucca Mountain repository. The DOE planned and conducted its site characterization program to collect data about the site and about those physical and chemical processes that would affect the ability of the repository system to isolate waste.

Section 1.3 presented a brief summary of the geology of Yucca Mountain based on the results of site investigations. It provided a framework for the descriptions of the repository and waste package designs contained in Sections 2 and 3. In this section, the results of studies focused on the characteristics and potential future behavior of the repository system are presented in additional detail. The discussion is organized to provide a description of the major processes that control the waste isolation capability of the potential repository. As shown schematically in Figure 4-2, Section 4.2 describes in sequence the data and analyses relevant to the processes that affect the movement of water through Yucca Mountain and relevant to the potential for that water to contact and mobilize radionuclides. Disruptive events could potentially affect these processes and, therefore, also need to be considered. The data and analyses related to potential disruptive events are presented in Section 4.3, and the combined analysis of the potential performance of the repository is summarized in Section 4.4.

The processes pertinent to performance include those physical processes that control the movement of water, beginning with precipitation as rain and snow at the surface, followed by infiltration into the mountain, flow through the unsaturated zone to the potential repository level, flow from the repository level to the saturated zone, and from there to the accessible environment. At the repository level, water moving past the engineered barriers would be affected by the physical and chemical processes associated with the decay heat and could interact with waste packages and waste forms. These processes could lead to the mobilization of radionuclides. Eventually, the water could move out of the repository horizon and further downward through the unsaturated zone. Subsequently, it could move into the saturated zone, where it could be transported to the accessible environment where humans could be exposed.

The data collected during site characterization have been used to develop conceptual and numerical models of the hydrologic, geochemical, thermal, and mechanical processes that will determine how a repository at Yucca Mountain may behave over the next 10,000 years. These process models have, in turn, been used to develop a TSPA model that has been used to assess quantitatively the potential for radionuclide releases to the public and, consequently, the safety of the Yucca Mountain site.

Attributes Important to Long-Term Performance—The potential repository system can be described in terms of five key attributes that would be important to long-term performance: (1) limited water entering waste emplacement drifts; (2) long-lived waste package and drip shield; (3) limited release of radionuclides from the engineered barriers; (4) delay and dilution of radionuclide concentrations by the natural barriers; and (5) low peak mean annual dose considering potentially disruptive events. These attributes are summarized below. The first four reflect the interactions of natural barriers and the engineered barriers in prolonging the containment of radionuclides within the repository and limiting their release. The fifth attribute reflects the likelihood that disruptive events would not affect repository performance over 10,000 years.

  1. Limited Water Entering Emplacement Drifts—The climate at Yucca Mountain is dry and arid, with precipitation averaging about 190 mm (7.5 in.) per year. Little of this precipitation percolates into the mountain; nearly all of it (above 95 percent) either runs off or is lost to evaporation, limiting the amount of water available to seep into emplacement drifts. For the higher-temperature operating mode described in this report, a thermal management strategy was developed that would take advantage of the heat of the emplaced wastes to drive water away from the emplacement drifts. The heat generated by the waste would dry out the rock surrounding the drift and decrease the amount of water available to contact the waste packages until the wastes have cooled substantially. Drainage of water in the rock pillars between drifts would be encouraged by keeping much of the pillar rock between the drifts below the boiling temperature of water. As long as the drift walls remained at temperatures above the boiling point of water, there would no liquid water in the emplacement drifts and very little in the nearby rock. Even after the drift walls cooled below the boiling point of water, the residual heat would increase evaporation in and near emplacement drifts, thereby continuing to limit the amount of water present in the rocks adjacent to waste packages. In lower-temperature operating modes, the waste packages would be exposed to water earlier. Because the rock would eventually cool in any operating mode, there does not appear to be a significant difference in the amount of water to which the waste packages would eventually be exposed.

    Over the range of operating modes, the design also takes advantage of the mechanical and hydrologic processes that divert water around emplacement drift openings in the unsaturated zone. Because of capillary forces, water flowing in narrow fractures tends to remain in the fractures rather than flow into large openings, such as drifts. If any water reaches an emplacement drift, it could flow down the drift wall to the floor and drain without contacting the drip shield or the waste packages (however, in taking a more conservative approach in modeling, this potential process is not included in the models). Thus, the natural and engineered features of a repository in the unsaturated zone will combine to limit the potential for water to enter the emplacement drifts.

  2. Long-Lived Waste Package and Drip Shield—To further reduce the possibility of water contacting waste, the DOE has designed a robust, dual-wall waste package with an outer cylinder of corrosion-resistant material, Alloy 22. Alloy 22 was selected because it will remain stable in the geochemical environment that would be expected in a repository at Yucca Mountain. In the operating mode described in this report, the repository environment would be warm, with temperatures at the waste package surface initially rising above the boiling point of water. Waste package surface temperatures are expected to gradually decrease to below boiling after a period of hundreds to thousands of years, depending on the waste package's location within the repository layout, spatial variation in the infiltration of water at the ground surface, and variability in the heat output of individual waste packages.

    Chemically, the environment is expected to be near-neutral pH (mildly acidic to mildly alkaline) and mildly oxidizing. Because most corrosion would occur only in the presence of water and because highly corrosive chemical conditions are not expected, both the titanium drip shield and the Alloy 22 outer barrier of the waste package are expected to have long lifetimes in an unsaturated zone repository. In lower-temperature operating modes, the waste packages would be exposed to water earlier.

  3. Limited Release of Radionuclides from the Engineered Barriers—Because of the characteristics of the natural system, the drip shields, and the waste packages, scientists do not expect water to come into contact with the waste forms for over 10,000 years. Even if water were to penetrate a breached waste package before 10,000 years, several characteristics of the waste form and the repository would limit radionuclide releases. First, because of the warm temperatures, much of the water that penetrates the waste package will evaporate before it can dissolve or transport radionuclides. Neither spent nuclear fuel nor glass waste forms will dissolve rapidly in the expected repository environment. Further, a large majority of the radionuclide inventory is insoluble in the geochemical environment expected within the repository. Although the performance of the cladding (metallic outer sheath of a fuel element) as a barrier may vary because of possible degradation, it is expected to limit contact between water and waste. The component of the engineered barrier system below the waste package, called the invert, contains crushed tuff that would also limit the transport of radionuclides into the host rock.

  4. Delay and Dilution of Radionuclide Concentrations by the Natural Barriers—Eventually, the engineered barrier systems in the repository are expected to degrade, and small amounts of water may contact waste. Even then, features of the geologic environment and the repository system will help decrease radionuclide migration to the accessible environment and slow it by hundreds to thousands of years. Processes that could be important to the movement of radionuclides include sorption, matrix diffusion, dispersion, and dilution. Rock units in both the unsaturated zone and the saturated zone at Yucca Mountain contain minerals that can adsorb many types of radionuclides (i.e., radionuclides would attach to and collect on the mineral surfaces). As water flows through fractures, dissolved radionuclides can diffuse into and out of the pores of the rock matrix, increasing both the time it takes for radionuclides to move from the repository and the likelihood that radionuclides will be exposed to sorbing minerals. Dispersion and dilution will occur naturally as potentially contaminated groundwater flows and mixes with other groundwater and lowers the concentration of contaminants.

  5. Low Mean Annual Dose Considering Potentially Disruptive Events—Yucca Mountain provides an environment in which hydrogeologic conditions important to waste isolation (e.g., a thick unsaturated zone with low rates of water movement) have not changed very much for at least hundreds of thousands of years. Analysts have identified and evaluated a wide variety of potentially disruptive processes and events that could affect the performance of the design and operating mode described in this report. These range from extremely unlikely events, such as meteor impacts, to events that are likely to occur, such as regional climate change. Although the probability of volcanic activity in or near the potential repository is low, volcanic activity was a consideration in TSPA in the disruptive scenario case. Performance assessment results to date show that potentially disruptive events are not likely to compromise the performance of the repository, and the probability-weighted mean dose for an igneous disruption is low.

Table 4-3 shows the relationship between the key attributes of the site and the physical barriers that comprise the repository system, the processes important to waste isolation, and the descriptions presented in Sections 4.2.1 through 4.2.10 [4.2.1, 4.2.2, 4.2.3, 4.2.4, 4.2.5, 4.2.6, 4.2.7, 4.2.8, 4.2.9, 4.2.10]and in Section 4.3. The table is based on similar information developed and presented in Repository Safety Strategy: Plan to Prepare the Safety Case to Support Yucca Mountain Site Recommendation and Licensing Considerations (CRWMS M&O 2001a) and the Total System Performance Assessment for the Site Recommendation (CRWMS M&O 2000a). The table incorporates the results of lower-temperature operating mode evaluations (BSC 2001b). In addition to the information presented in these sources, Table 4-3 relates the key attributes and processes to the natural and engineered barriers that would contribute to waste isolation at Yucca Mountain. The last column of the table also lists process model reports that contain more detailed descriptions of the processes summarized here and included in the TSPA. The column showing the processes important to repository performance is not intended to be comprehensive: it is presented at a summary level. More detailed definitions of the proposed repository system and relevant specific processes are included in the individual sections that follow.

Organization of Section 4.2—As organized below, and shown in Figure 4-2, the total system consists of numerous interdependent subsystems that are described in ten subsections, as follows:

Within each of these sections, a similar internal organization is used to present the DOE's understanding of the key processes and to present the data and analyses that provide the basis for that understanding. Each discussion includes:

  1. A summary description of how each process would operate in a potential repository

  2. A description of the investigations, tests, and other data (including analogue information) that provide the technical basis for the DOE's understanding

  3. A description of the conceptual and numerical models that have been developed to allow the DOE to assess potential future performance, including specific information about the sources and treatment of uncertainties, alternative models, and model calibration and validation

  4. A description of how the conceptual and numerical models have been abstracted (or represented) in the TSPA-SR (CRWMS M&O 2000a).

4.2.1 Unsaturated Zone Flow

This section summarizes the current understanding of water movement (i.e., percolation flux) through the unsaturated zone and into a repository (i.e., seepage into drifts) at Yucca Mountain. Fluid flow through the unsaturated zone at Yucca Mountain is described at length in Unsaturated Zone Flow and Transport Model Process Model Report (
CRWMS M&O 2000c, Sections 3.2 to 3.4 and 3.6 to 3.9), which is supported by 24 detailed analysis model reports. Yucca Mountain Site Description (CRWMS M&O 2000b, Sections 8.2 to 8.10) also provides a comprehensive summary of investigations performed to characterize flow and seepage in the unsaturated zone. Figure 4-3 shows the relationships between the main unsaturated zone processes, with those relevant to unsaturated zone flow highlighted.

The primary purpose of Unsaturated Zone Flow and Transport Model Process Model Report (CRWMS M&O 2000c) is to develop models for the TSPA that evaluate the postclosure performance of the unsaturated zone. Unsaturated Zone Flow and Transport Model Process Model Report supplies to the TSPA (1) ambient and predicted (i.e., future) three-dimensional flow fields based on different climate states and infiltration scenarios and (2) seepage rates into potential waste emplacement drifts. The flow fields are used directly in TSPA calculations of transport, while the seepage rates are used to calculate distributions of the fraction of waste packages in contact with seepage water and the volumetric flow rate to a waste package segment, taking into account possible flow focusing effects from site scale to drift scale (see Section 4.2.1.4).

As noted in Section 4.1.4, the DOE is evaluating the possibility of mitigating uncertainties in modeling long-term repository performance by operating the design described in this report at lower temperatures. Some analyses described in this section have been updated and expanded to capture the features and processes relevant to operating the design at lower temperatures. The updated analyses are described in Volume 2, Section 4 of FY01 Supplemental Science and Performance Analyses (BSC 2001b).

4.2.1.1 Conceptual Basis

On the most fundamental level, the important factors affecting unsaturated groundwater flow at Yucca Mountain are climate and rock hydrologic properties. Derived from these two basic components are estimates of percolation flux and seepage into potential waste emplacement drifts, both of which are key unsaturated zone processes.

Located in southern Nevada, one of the most arid regions of the United States, Yucca Mountain is underlain by a thick unsaturated zone (
CRWMS M&O 2000c, Section 3.2.1). A dry climate and a deep water table are considered favorable characteristics for waste isolation. In a desert environment, the total amount of available water is small. The potential repository would be designed to complement the hydrologic environment by diverting the small flow of water that does occur away from the waste packages. Multiple natural and engineered barriers are expected to limit contact between water and waste forms, and retard radionuclide migration. The climate data and unsaturated zone characteristics are discussed in Sections 4.2.1.2.1 and 4.2.1.2.2, respectively.

The major components of the unsaturated zone and the emplacement drifts that affect water movement are illustrated in Figure 4-4. Water movement starts with rainfall in the arid environment, which is subject to runoff, evaporation, and plant uptake, such that much of the rainfall never reaches the potential repository host rock units. The infiltration of water that penetrates into the rock units of the unsaturated zone is redistributed by flow processes in the fractured and faulted, welded and nonwelded tuff layers. When percolating water encounters an underground drift, much of it will be diverted by capillary barrier mechanisms around the opening and never contact the engineered barriers within.

Major issues related to unsaturated zone flow processes include:

Using these major issues as subheadings, a summary of the current conceptual understanding of unsaturated zone flow beneath Yucca Mountain is first presented. Following this summary are subsections that describe the data supporting the conceptual understanding of flow in the unsaturated zone (Section 4.2.1.2), the development and integration of numerical process models based on the conceptual interpretation of the available data (Section 4.2.1.3), and how the results of unsaturated zone numerical modeling studies have been abstracted for the TSPA (Section 4.2.1.4).

4.2.1.1.1 Climate and Infiltration

Climate is defined by the variation of meteorological conditions (such as temperature, pressure, humidity, precipitation rate and prevailing winds) over time. At Yucca Mountain, climate is important because it provides the boundary conditions for the hydrologic system—specifically, the amount of water available at the surface. Estimates of the precipitation rate and temperature taken from climate models have been used as information to determine the net infiltration of water into Yucca Mountain and the percolation flux at the repository horizon. Percolation flux within the unsaturated zone, governed by climate and rock hydrogeologic properties, is a key process affecting seepage in waste emplacement drifts and transport of radioactive particles below the repository. Three representative climates are forecast to occur within the next 10,000 years (i.e., the period of regulatory compliance): the modern (present-day) climate, a monsoon climate, and a glacial-transition climate (
USGS 2000a, Section 6.6). Beyond 10,000 years, the TSPA-SR extends the glacial-transition climate for base-case simulations and includes a revised long-term climate model in a sensitivity study, as discussed in Section 4.4.2.4 (CRWMS M&O 2000a, Section 3.2.1.2) and Volume 1, Section 3.3.1 of FY01 Supplemental Science and Performance Analyses (BSC 2001a).

Net infiltration refers to the penetration of liquid water through the ground surface and to a depth where it can no longer be removed by evaporation or transpiration by plants. Net infiltration is the source of groundwater recharge and water percolation at the potential repository horizon; it provides the water for flow and transport mechanisms that could move radionuclides from the potential repository to the water table. The overall framework of the conceptual model for net infiltration is based on the hydrologic cycle. Important processes that affect net infiltration include precipitation (rain and snow), runoff and run-on (flow of surface water off one place and onto another), evaporation, transpiration, and moisture redistribution by flow in the shallow subsurface (USGS 2000b, Section 6.1.3). Infiltration is temporally and spatially variable because of the nature of the storm events that supply precipitation and because of variation in soil cover and topography (CRWMS M&O 2000bq, Section 6.1.1). Surficial soils and topography are considered part of the natural barrier system because they reduce the amount of water entering the unsaturated zone by surficial processes (e.g., precipitation lost to runoff, evaporation, and plant transpiration) (Table 4-2). Net infiltration rates are believed to be high on sideslopes and ridge tops where bedrock is exposed and fracture flow in the bedrock is able to move liquid water away from zones of active evaporation (Flint, L.E. and Flint 1995, p. 15).

Within the limits of extant monitoring data, significant net infiltration occurs only every few years under the present climate (CRWMS M&O 2000bq, Section 6.1.1). In these years, the amount of net infiltration still varies greatly, depending on storm amplitude, duration, and frequency. In very wet years, net infiltration pulses into Yucca Mountain may occur over a period of a few hours to a few days. A detailed discussion of net infiltration processes can be found in Simulation of Net Infiltration for Modern and Potential Future Climates (USGS 2000b) and in Volume 1, Section 3.3.1 of FY01 Supplemental Science and Performance Analyses (BSC 2001a).

4.2.1.1.2 Fracture Flow and Matrix Flow within Major Hydrogeologic Units

An early conceptual hydrologic model of unsaturated flow at Yucca Mountain, developed by
Montazer and Wilson (1984), identified five major hydrogeologic units within the unsaturated zone. From the land surface to the water table, these units are the Tiva Canyon welded (TCw), the Paintbrush nonwelded (PTn), the Topopah Spring welded (TSw), the Calico Hills nonwelded (CHn), and the Crater Flat undifferentiated (CFu) units. Table 4-4 correlates major hydrogeologic, lithostratigraphic, and detailed hydrogeologic units with the layering scheme used in unsaturated zone modeling activities. These units are described in greater detail in the Development of Numerical Grids for UZ Flow and Transport Modeling (CRWMS M&O 2000br, Section 6.4.1), in the Geologic Framework Model (GFM3.1) (CRWMS M&O 2000bs, Section 6.4.1), and by Flint, L.E. (1998).

The texture of Yucca Mountain tuffs ranges from nonwelded to densely welded (CRWMS M&O 2000c, Section 3.2.1). Typically, the porosity and permeability of the rock mass are inversely proportional to the degree of welding, and the degree of fracturing is directly proportional to the degree of welding. The degree of welding observed in the individual tuff units is primarily controlled by their cooling history. Generally speaking, the slower a rock cools, the more densely welded the material becomes. This densely welded material (matrix) is usually quite brittle in nature and develops well-connected fracture networks. These extensive, well-connected fracture networks, in turn, provide numerous pathways for the flow of liquids and gases. Conversely, the nonwelded rocks, which experienced rapid heat dissipation, display high matrix porosity and possess few fractures. Flow through these rocks is dominated by matrix flow processes (CRWMS M&O 2000c, Section 3.3.3).

The partitioning of total flux between the fracture flow component and the matrix flow component is one of the most important processes to determine in the unsaturated zone. Percolation distribution determines the amount of water that could potentially contact the waste packages and other components of the engineered barrier system. Determination of the flow components is also important for chemical transport processes. Water flow in fractures is typically much faster than flow in the matrix, leading to much faster movements for radionuclides and other chemicals in fractures compared to the matrix (CRWMS M&O 2000c, Section 3.3.3). The characteristic flow behavior in each of the major hydrogeologic units is described in the following paragraphs.

Flow Above the Potential Repository—The TCw is the most prevalent hydrogeologic unit exposed at the land surface (CRWMS M&O 2000c, Section 3.2.2.1). The unit is of variable thickness because of erosion and is composed of moderately to densely welded, highly fractured pyroclastic flow deposits of the Tiva Canyon Tuff. The high density of interconnected fractures and low matrix permeability of the unit (CRWMS M&O 2000bt, Sections 6.1 and 6.2) are considered to give rise to significant water flow in fractures and limited matrix imbibition (water flow from fractures to the matrix). Therefore, episodic infiltration pulses are expected to move rapidly through the fracture system into the underlying PTn unit with little attenuation by the matrix.

At the interface between hydrogeologic units TCw and PTn, tuffs grade downward over a few tens of centimeters from densely welded to nonwelded, accompanied by an increase in matrix porosity and a decrease in fracture frequency (Figure 4-5a) (CRWMS M&O 2000c, Section 3.2.2.1). The relatively high matrix permeabilities and porosities, and low fracture densities, of the PTn (CRWMS M&O 2000bt, Sections 6.1 and 6.2) should convert the predominant fracture flow in the TCw to dominant matrix flow within the PTn unit (CRWMS M&O 2000bq, Section 6.1). This, along with the relatively large storage capacity of the matrix resulting from its high porosity and low saturation, is expected to give the PTn significant capability to attenuate infiltration pulses and smooth areal differences in infiltration from the overlying welded unit and result in approximately steady-state water flow below the PTn. Through-going fracture networks within the PTn unit are rare and typically associated with faults (Rousseau, Kwicklis et al. 1999, pp. 53 to 54), so only a small amount of water is expected to pass through the PTn by way of fast flow paths. Recent analyses indicates that some lateral diversion on the PTn is probable (BSC 2001a, Section 3.3.3).

Conceptualization of the character of flow at the interface between the PTn and the overlying TCw for the TSPA-SR model was based on findings from estimates of flux rates in the PTn from geochemistry (Yang and Peterman 1999) and on hydrogeological properties described by Flint, L.E. (1998). The implication of that characterization was that little or no lateral flow diversion of downward percolating unsaturated zone water could be expected as a result of a contrast of hydrologic properties across this surface and its gentle dip to the east. If diversion of downward percolating water to permeable fault zones outside the potential repository footprint at that elevation occurs, it implies that water might be diverted around a repository, thus benefiting repository performance. Earlier work by Montazer and Wilson (1984) supported the conceptual model of lateral diversion on the PTn. The TSPA-SR model did not include lateral diversion on the PTn on the assumption that if any such diversion exists, it is small and that to leave it out of the model is conservative.

Geochemical evidence collected since the conceptualization of the original TSPA-SR modeling appears to support the existence of lateral diversion in the PTn based on chloride abundance (BSC 2001a, Section 3.2.3). Recent modeling approaches that combine pneumatic data from above, below, and within the PTn, saturation and water potential data, and geochemical data were used to calibrate unsaturated characteristic parameters and to differentiate alternative conceptual models (BSC 2001a, Section 3.2.3). Using a fine grid spacing, calibrated to match the chloride distributions and the estimated percolation flux data in the unsaturated zone below the PTn, supports the lateral diversion of water around a potential repository and the relatively uniform distribution of percolation flux in the deep unsaturated zone (BSC 2001a, Section 3.3.3). The PTn exhibits inhomogeneous lithologic character and distribution and some evidence, in the form of inferred occurrences of bomb-pulse chlorine-36 at depth, that fast flow paths for relatively small volumes of water may exist along faults and perhaps in zones of fracture flow focusing. In most of the region of the potential repository footprint, the PTn is expected to damp flow surges in the percolation flux rate and smooths areal differences in flow, which originate in the temporal and spatial patterns of infiltration at the surface (CRWMS M&O 2000c, Section 3.3.3.2). Smoothing of flow is supported by evenly distributed chloride mass balance data (BSC 2001a, Section 3.3.3).

Although the PTn is predominantly nonwelded, rock-hydrologic properties are highly heterogeneous because of differing depositional environments, lateral variations in welding, and the variable distribution of mineralogically altered (e.g., smectitic and zeolitic) intervals within the individual PTn subunits (CRWMS M&O 2000c, Section 3.2.2.2).

The transition from the lower PTn into the upper TSw is marked by a decrease in matrix porosity and an increase in fracture frequency (Figure 4-5b) (CRWMS M&O 2000c, Section 3.2.2.2) as the tuffs grade sharply downward from nonwelded to densely welded. These changes in porosity and fracture characteristics may create saturated conditions above this contact that could initiate fracture flow into the TSw.

Lithostratigraphic units within the TSw (including the middle nonlithophysal, lower lithophysal, and lower nonlithophysal potential repository host units) are moderately to densely welded and are primarily distinguished by the relative abundance of lithophysae (cavities formed by bubbles of volcanic gases trapped in the tuff matrix during cooling), crystal content, mineral composition, pumice and rock fragment abundance, and fracture characteristics (CRWMS M&O 2000c, Section 3.2.2.3). Differences in lithophysal abundance and fracture characteristics are shown in Figure 4-6.

Unsaturated flow in the TSw is primarily through the fractures because of the magnitude of matrix hydraulic conductivity of the TSw relative to the estimated average infiltration rate. If the hydraulic gradient is assumed to be one (i.e., flow is vertically downward and gravity driven), the maximum matrix percolation rate is the same as the matrix hydraulic conductivity. Because the estimated matrix hydraulic conductivity of some TSw subunits is much lower than the average estimated infiltration rate (CRWMS M&O 2000bq, Section 6.1.2), the remainder of the flow must be distributed in the fracture network.

Flow Below the Potential Repository—Flow behavior below the TSw is important for modeling radionuclide transport from the repository horizon to the water table because transport paths follow the water flow pattern. The main hydrogeologic units below the TSw are the CHn and CFu (CRWMS M&O 2000c, Sections 3.2.2.4 and 3.2.2.5). The CHn contains primarily nonwelded layers whose initial vitric composition has been variably transformed by high and low temperature alteration to clays and zeolites. A portion of the lower half of the CHn (corresponding to the interior of the Prow Pass Tuff) is characterized by moderately welded to densely welded layers that have undergone devitrification (high-temperature conversion of glass to crystalline material). Devitrified, welded rocks show greater fracture intensity than the nonwelded layers and typically do not contain alteration minerals (Flint, L.E. 1998, p. 9). In the southern half of the potential repository footprint and to the south, a portion of the upper CHn (corresponding principally to the Calico Hills Formation, Tac) is largely unaltered (i.e., vitric). This volume of vitric material is believed to represent the part of the CHn that remained above past elevated saturated zone water levels (CRWMS M&O 2000c, Section 3.2.4).

The CFu unit (consisting of portions of the Bullfrog and Tram tuffs that occur above the water table) is a subset of the Crater Flat Group, which contains the Prow Pass, Bullfrog, and Tram tuffs (CRWMS M&O 2000b, Section 4.5.4.5). Lithostratigraphic units within the CFu are nonwelded to densely welded, with the nonwelded tuffs being pervasively altered to zeolites. The Prow Pass, Bullfrog, and Tram tuffs are all similar in that they each contain devitrified, densely welded interiors that grade above and below into nonwelded, zeolitically altered tuffs.

The nonwelded vitric, nonwelded zeolitic, and welded devitrified tuffs have significantly different properties and flow characteristics. The zeolitic rocks have very low matrix permeability and slightly greater fracture permeability; therefore, a relatively small amount of water may flow through the zeolitic units (primarily through fractures), while most of the water is diverted around these low-permeability bodies (Figure 4-7b) (CRWMS M&O 2000c, Section 3.3.3.4). Conversely, vitric portions of the CHn have relatively high matrix porosity and permeability and are characterized by low fracture frequencies (similar to layers within the PTn); therefore, matrix flow dominates, and fracture flow is believed to be limited in the vitric units. Devitrified tuffs have slightly lower matrix porosities than the nonwelded tuffs and increased fracturing (because of increased welding), yet little or no alteration, giving them relatively higher permeabilities than the zeolitic tuffs (Flint, L.E. 1998, pp. 29 and 32).

The high storage capacity of the vitric (unaltered) CHn matrix will attenuate the rate of water movement through the unsaturated zone (Figure 4-7a). Where extensive mineralogic alteration has occurred—for example, at the TSw–CHn contact—the downward flux of water may exceed the rock's transmissive capacity, leading to ponding above the flow barrier and the formation of perched water (Figure 4-7c). The presence of a low-permeability barrier to vertical flow can lead to lateral flow diversion, especially if the flow barrier is dipping and saturated moisture conditions (i.e., perched water bodies) exist above the barrier. Therefore, not all flow paths below the potential repository horizon are expected to be vertical. Lateral diversion of water at perching horizons may lead to flow focusing if the vertical flow barrier is intersected by a high-permeability feature, such as a fault, that could channel flow to the water table (CRWMS M&O 2000c, Sections 3.3.3.4 and 3.3.5).

4.2.1.1.3 Effects of Major Faults

Different kinds of faults with varying amounts of displacement exist at Yucca Mountain (
CRWMS M&O 2000c, Section 3.3.5). Fault hydrologic properties are variable and generally controlled by rock type and stratigraphic displacement. Because major faults have the potential to significantly affect the flow processes at Yucca Mountain, they are important features of the unsaturated zone.

A fault can act as a fast flow conduit for liquid water (Figure 4-7d). In this case, transient water flow may occur within a fault as a result of temporally variable infiltration. Major faults cut through the PTn unit, possibly reducing the attenuating effect of the PTn on transient water flow. However, fast flow along major faults is expected to carry only a small amount of water and may not contribute significantly to the flow of water above the potential repository horizon in the unsaturated zone (see Section 4.2.1.3.1.1). Faults intercepting the perched water bodies, however, can correspond to significant vertical water flow if fault permeability is relatively high because of the locally saturated conditions existing in the surrounding rock (see Figure 4-7c) (CRWMS M&O 2000c, Sections 3.3.5 and 3.7.3.2).

If faults within the CHn are relatively permeable features, they may provide a direct flow pathway to the water table. This is particularly significant because radionuclides released from the potential repository could bypass zeolitic or vitric layers within the CHn unit, where they could be retarded by sorption. Conversely, faults might be considered a positive feature of the site if they divert water around waste emplacement drifts or prevent laterally flowing water from focusing at the area of waste emplacement.

Alternatively, a fault can act as a barrier for water flow (CRWMS M&O 2000c, Section 3.3.5). Where a fault zone is highly fractured, the corresponding coarse openings will create a capillary barrier for lateral flow. On the other hand, a fault can displace the surrounding geologic units such that a unit with low permeability abuts one with relatively high permeability within the fault zone. In this case, the fault will act as a permeability barrier to lateral flow within the units with relatively high permeability. Montazer and Wilson (1984, p. 20) hypothesized that permeability would vary along faults, with higher permeability in the brittle, welded units and lower permeability in the nonwelded units where gouge or sealing material may be produced. While a fault sealed with gouge or other fine-grained material has much higher capillary suction (i.e., driving imbibition), it also has low permeability, retarding the movement of water.

Large lateral flow to the faults and/or focusing of infiltration near the fault zone on the ground surface are required to generate significant water flow in faults. Below the repository, low-permeability (zeolitic) layers in the CHn may channel some flow to faults that act as pathways to the water table. However, it is also possible that alteration of faulted rocks in the CHn and CFu causes the faults to be of low permeability (Figure 4-7e), slowing water movement from the TSw to the water table.

4.2.1.1.4 Fracture–Matrix Interaction

Fracture–matrix interaction refers to flow and transport between fractures and the rock matrix (
CRWMS M&O 2000c, Section 3.3.4). Owing to their different hydrologic properties, distinct flow and transport behavior occurs in each hydrogeologic unit. The extent of fracture–matrix interaction is therefore a key factor in assessing flow and transport processes in the unsaturated zone (BSC 2001a, Section 4.2.1).

4.2.1.1.5 Seepage into Drifts

Potential seepage of water into waste emplacement drifts is important to the overall performance of the potential repository system. The corrosion of drip shields and waste packages, the mobilization of radioactive contaminants from breached waste packages, and the migration of radionuclides to a receptor location all depend on the distribution of water seepage into the emplacement drifts.

Seepage is defined as flow of liquid water into an underground opening, such as a waste emplacement drift or exploratory tunnel (
CRWMS M&O 2000c, Section 3.9.1). Seepage does not include water vapor movement into openings or condensation of water vapor within openings. Seepage flux is the rate of seepage flow per unit area. The seepage percentage is the ratio of seepage flux to percolation flux in the surrounding host rock unit. Seepage threshold is defined as a critical percolation flux below which seepage into the opening is unlikely to occur. The seepage fraction is the proportion of waste packages that are located where drift seepage occurs. The drift shadow zone is a zone of reduced water saturation beneath the emplacement drift as a result of diversion of seepage around the drift opening by capillary forces.

Estimating seepage into underground openings excavated from an unsaturated fractured formation requires an understanding of processes on a wide range of scales (CRWMS M&O 2000c, Section 3.9.1). These scales range from the mountain-scale distribution of percolation flux, to the intermediate-scale channeling or dispersion of flow in an unsaturated fracture network, to the small-scale capillary-barrier effect, to the microscale phenomena within fractures, and specifically at the drift wall. Moreover, the thermodynamic environment in the drift (temperature, relative humidity, ventilation regime, etc.) must be considered. Figure 4-8 illustrates and summarizes seepage-related processes. The factors affecting drift seepage highlighted in Figure 4-8 are outlined below.

Capillary-Barrier Effect, Flow Diversion, and Seepage Threshold—For unsaturated conditions, the seepage flux is expected to be less than the percolation flux because the drift opening acts as a capillary barrier (Philip et al. 1989). When percolating water encounters the opening, capillary forces retain the water in the rock, preventing it from seeping into the drift. Water accumulates in the rock around the opening, and if the rock permeability is sufficient, the water flows around the drift opening. If the incident percolation flux is very high or the rock permeability is insufficient, complete water saturation occurs in the rock above the opening, and seepage occurs. The effectiveness of the capillary barrier to seepage is determined by the capillarity of fractures surrounding the drift and by the permeability and connectivity of the fracture network in the horizontal direction (BSC 2001a, Section 4.2.2). Note that even if seepage occurs, the seepage flux is generally less than the percolation flux unless flow is focused by fractures or other features. The seepage threshold indicates whether or not water seeps into the opening for a given average percolation flux in the surrounding rock. Seepage threshold behavior is controlled by drift geometry, fracture geometry, capillary properties of fractures, and the hydrologic properties of the fracture network (BSC 2001a, Section 4.2.2).

Distribution of Percolation Flux, Flow Channeling, and Episodic Flow—The magnitude and spatial distribution of percolating water in the potential repository host rock is the most important factor affecting seepage. The distribution of flow channels in the fracture network and the hydrologic properties of individual flowing fractures determine how seepage occurs (CRWMS M&O 2000c, Section 3.9.1). Depending on the flow of water within an individual channel, the seepage threshold may or may not be exceeded locally. This is important because seepage is sensitive to the magnitude of the percolation flux, which is moderated by flow processes between the ground surface and the potential repository horizon. For repository thermal loading conditions, the percolation flux will include downward flow of condensed water, in addition to water that infiltrates at the ground surface.

Hierarchal Fracture Network—The characteristics of the fracture network affect seepage because they determine the spatial distribution of percolation flux and the effectiveness of the capillary barrier (CRWMS M&O 2000c, Section 3.9.1). Intermediate-scale characteristics (between mountain-scale and drift-scale) of the fracture network control the potential focusing of flux in the unsaturated zone. Heterogeneity of the fracture network affects local percolation flux and, therefore, seepage. The capillary-barrier effect depends on the connectivity of the fracture network near drift openings and the capillary properties of individual fractures. Small fractures and microfractures, if interconnected, can decrease seepage because they have sufficient capillarity to hold water, but (unlike the rock matrix) sufficient permeability for flow diversion around the openings.

Drift Opening Geometry and Rock Surface Characteristics—The shape and size of underground openings also affects the likelihood of seepage. Partial collapse of the opening because of rockfall can affect seepage. Analytical solutions demonstrating the impact of drift geometry on seepage were developed by Philip et al. (1989). In addition, the geometry of the rock roof in drift openings and the characteristics of the rock surface control processes that could lead to dripping of seepage onto the engineered barriers below.

Ventilation, Evaporation, and Condensation—Until permanent closure of the potential repository, the emplacement drifts will be ventilated. The resulting temperature and humidity conditions in the drift will determine evaporation and condensation effects. Evaporation at the drift wall will generally decrease droplet formation and dripping (Ho 1997a) and create a dryout zone around the drift. When relative humidity in the drift is kept well below 100 percent by ventilation, seepage of liquid water will decrease, while water vapor movement into the drift will increase. Seepage flux and the moisture from increased vapor influx will be effectively removed by ventilation. After the thermal period, the relative humidity in emplacement drifts may be high enough to support condensation within the drift and engineered barrier system whose thermal properties are such that their temperature may be below the dew point. This would mostly occur after the diminution of all or most of the waste heat and the cessation of drift ventilation.

Excavation-Disturbed Zone and Dryout Zone Effects—The capillary-barrier effect that produces seepage diversion around openings occurs within a limited region around the opening (CRWMS M&O 2000c, Section 3.9.1). The extent of this zone is approximately given by the height to which water rises on account of capillarity. It is probably smaller than the zone affected by the mechanical effects of excavation; therefore, these effects may modify seepage behavior. Thermally induced stress changes may also cause changes in fracture permeability. These stress changes are currently under investigation in field-scale thermal testing at Yucca Mountain (BSC 2001a, Section 4.3.1.5). In addition, drift ventilation and heating will produce a dryout zone with associated dissolution and precipitation of minerals (CRWMS M&O 2000c, Section 3.9.1). The consequent alteration of hydrologic properties and its extent are also under investigation (BSC 2001a, Section 4.3.5).

Design—The layout and design of the potential repository and the engineered barrier system affect the probability of seepage water contacting waste packages. Orientation of the emplacement drifts with respect to natural fracturing and in situ stress directions affects the hydrologic behavior of the fracture network around the openings and the potential for rockfall that changes the drift geometry (see Section 2.3.4.1). Thermal loading controls temperatures, the duration and extent of dryout associated with heating, and the potential for coupled mechanical and chemical processes that can impact hydrologic properties of the host rock.

4.2.1.2 Summary State of Knowledge

Site data characterizing the ambient unsaturated system at Yucca Mountain have been collected since the early 1980s (CRWMS M&O 2000c, Section 2.1). The data are of numerous types (e.g., lithology, rock hydrologic properties, mineralogy, temperature, geochemistry, climate/infiltration) collected from surface-based activities (e.g., geologic mapping, installation of vertical boreholes) as well as subsurface mapping, sampling, and in situ testing in excavated tunnels.

Data collection has focused mainly on near-surface units down to the potential repository horizon. Site characterization data gathered from surface-based studies (including vertical boreholes and mapped pavements) represent the upper hydrogeologic units (i.e., TCw, PTn, and TSw), which are more readily accessible than the deeper units (i.e., the CHn and CFu) in the unsaturated zone. Exploratory tunnels excavated within Yucca Mountain, allowing scientists to collect large amounts of many different types of data, transect those layers above the lowest lithostratigraphic unit within which a repository would be constructed (i.e., the lower nonlithophysal unit of the TSw).

The discussion that follows summarizes the results of studies that collected many different types of data, including:

4.2.1.2.1 Climate and Infiltration Data

The southwestern United States is characterized by dry climates with evaporation exceeding annual precipitation. The dry climates are divided into an arid (or desert) type and a semiarid (or steppe) type from annual temperature and precipitation considerations (
Moran et al. 1997, pp. 438 to 443; Trewartha and Horn 1980, pp. 221 to 229). For Yucca Mountain's mean annual air temperature of approximately 17°C (63°F) (USGS 2000b, Section 6.4.2), the climate is arid if the precipitation is less than 243 mm (9.6 in.) per year and is semiarid if the precipitation is between 243 mm (9.6 in.) and 486 mm (19.1 in.) per year, based on a modified Köppen climate classification scheme (Trewartha and Horn 1980, p. 228). Seasonal variations and atmospheric conditions can be used to further modify and refine the classifications. The climate boundary definition depends on the classification scheme used and is best viewed as convenient approximation only (Trewartha and Horn 1980, p. 223).

Present day climate data have been collected in and around Yucca Mountain since 1988. The climate is dry and arid. The average annual potential evapotranspiration rate is approximately six times greater than the average precipitation rate (USGS 2000b, Section 6.1.4). On average over the unsaturated zone flow and transport model area, Yucca Mountain receives about 190 mm (7.5 in.) of rain and snow per year; nearly all the precipitation (above 95 percent) either runs off or is lost to evaporation (USGS 2000b, Table 6-9; CRWMS M&O 2000c, Table 3.5-2). The precipitation increases with elevation, so the higher portions of the mountain receive more precipitation than adjacent valleys. Over a larger area that includes valleys and flat lands around the Yucca Mountain model area, the average precipitation value can be lower (Hevesi et al. 1994, p. 2520, Figure 1).

Yucca Mountain is located in the rain shadow caused by the Sierra Nevada and other mountain ranges, which limits the number of storms that can generate precipitation throughout the Great Basin (USGS 2000a, Section 6.1). During the winter, regional precipitation comes from the occasional intrusion of frontal storms associated with polar fronts; during the summer, precipitation comes from the regular intrusion of hot, dry, subtropical high-pressure weather systems. Local vegetation on all but the highest mountains is limited by the poor soils and the amount of available water. The vegetation that grows there typically uses most of the available precipitation that is not removed by evaporation or runoff. Therefore, infiltration of precipitation from the surface into the underlying rock is modest and commonly associated with wet years that may be linked to El Niño cycles.

The climate study also evaluated the long-term records of analogue sites, such as calcite dating data from Devils Hole (Winograd et al. 1992) and microfossil records from Owens Lake (Forester et al. 1999, pp. 14 to 18). Geological information indicates that the regional climate has changed over the past million years and that long-term average precipitation (which reflects wetter glacial and monsoonal cycles in the past) is greater than modern conditions. Future climate scenarios use available climate data from wetter analogue sites in western states (CRWMS M&O 2000b, Section 6.4; USGS 2000a, Section 6.6.2).

Studies of past climates indicate that the climate oscillates between glacial and interglacial periods. The current climate is typical of interglacial periods, although paleoclimate records suggest that the present interglacial period is hotter and drier than some earlier interglacial periods. In contrast with the current climate, periods of more extensive glaciation have dominated the long-term climate for most of the past 500,000 years. Glacial periods characterized by wetter and colder conditions than now exist have prevailed during approximately 80 percent of that time (USGS 2000a, Section 6.2).

No glaciation has occurred in the Yucca Mountain region during these glacial periods; instead, the region has experienced climates characterized by increased rainfall and cooler temperatures (USGS 2000a). During glacial periods, winter storm activity is more common because the polar front moves far south of its average present-day position. Subtropical high-pressure systems in the summer are less frequent to nonexistent. Local vegetation receives more water from the atmosphere and average air temperatures are colder, leading to lower plant uptake rates and higher soil infiltration rates than today. The primary form of annual precipitation shifts from summer rain to winter precipitation (often snow). Wetlands are common on the valley floors, and local streams are active during all or most of the year. Large closed basins, such as Owens Valley and Death Valley, contain lakes. Although glacial climates are generally characterized by cooler temperatures and higher precipitation, particular glacial periods vary. Some are relatively warm and wet, whereas others are cold and may be either wet or dry.

In Future Climate Analysis (USGS 2000a, Section 6.6), three climatic states are forecast: (1) the current arid climate, (2) an interglacial monsoon climate of warm but wetter conditions, and (3) a cooler, wetter, glacial-transition climate typical of glacial conditions over much of the past several hundred thousand years. Within each of these general categories, conditions may vary from year to year and over decades and centuries. Monsoon climates would vary from climates like present-day Yucca Mountain to somewhat wetter climates like Nogales, Arizona, or Hobbs, New Mexico. Glacial-transition climates would likely range from conditions like the present-day central Great Basin (in central Nevada) to cooler conditions like the climate near Spokane, Washington. None of the likely future climates for Yucca Mountain is characterized by much larger annual precipitation rates at least for times on the order of 10,000 years. However, the cooler temperatures and decreased plant uptake of water would allow more water to infiltrate into Yucca Mountain during a glacial-transition climate.

The infiltration study at Yucca Mountain was initiated in 1984. To date, the infiltration study has used nearly 100 shallow boreholes located on ridgetops, on sideslopes, on stream terraces, and in/across stream channels to measure the changes in water-content profiles in response to precipitation and snowmelt events (Flint, L.E. and Flint 1995; Flint, A.L. et al. 1996, pp. 60 to 63; USGS 2000b, Sections 6.3.4 and 7.1). Weekly or monthly measurements were collected from 1984 through 1995. Four washes were instrumented for run-on and runoff measurements. Water-balance calculations from precipitation, evapotranspiration, run-on, and runoff along washes are used to derive the infiltration flux values over the ridge top, side slopes, and stream channels. Areas with exposed bedrock and no soil cover promote greater infiltration compared to areas with soil covers that have substantial storage capacity for excess water.

Deep soils and vegetation inhibit infiltration by allowing evaporation and plant transpiration to remove water. Steep slopes encourage rapid runoff, also limiting infiltration (USGS 2000b, Section 6.1.2). For these reasons, most infiltration probably occurs in areas with low slopes and relatively shallow soil cover, as is common at the higher elevations on the northern part of Yucca Mountain. Analyses of groundwater chemistry (especially oxygen and hydrogen isotopic compositions) indicate that much of the infiltration at Yucca Mountain occurs during the winter. During the cool rainy season, evaporation rates are low because of low temperatures, and low-intensity but sustained precipitation can saturate shallow soil or cover the soil with snow. Summer storms are more intense than winter rains, but higher runoff and evaporation combine to limit summer infiltration.

The infiltration distributions for different climate states are used as upper boundary conditions for the unsaturated zone flow model and TSPA models. A numerical, water-balance, infiltration model was developed for the Yucca Mountain area, including the area of the three-dimensional, site-scale unsaturated zone flow model. The infiltration model uses physiographic and hydrologic information and a daily precipitation record to calculate daily values of infiltration using a water-balance approach. The infiltration model was calibrated first by comparison to the total water-content change in the soil profile in neutron boreholes during 1984 to 1995 and then by comparison of model-simulated streamflow to discharge measures at stream-gauging sites on Yucca Mountain during 1994 to 1995 (USGS 2000b, Section 6.8.2).

The numerical infiltration model was used to simulate lower-bound, mean, and upper-bound infiltration associated with three climate scenarios determined to be pertinent to performance of the potential repository: modern, monsoon, and glacial-transition (USGS 2000b, Section 6.9). The modern, or present-day, climate conditions are expected to prevail for about the next 400 to 600 years. Monsoon climate conditions, with wetter summers than the modern climate, are expected to prevail for the following 900 to 1,400 years. Glacial-transition climate conditions, with cooler air temperatures and higher annual precipitation than the modern climate, are expected to begin in about 2,000 years and continue throughout the remainder of the 10,000-year period specified for performance analyses.

Average precipitation and average infiltration rates over the unsaturated zone flow model area are presented in Section 4.2.1.3.3. The distributions of mean present-day, monsoon, and glacial-transition infiltration are also shown in Section 4.2.1.3.1.3 in Figure 4-25.

4.2.1.2.2 Geologic Data

Depth below ground surface to the water table ranges from approximately 500 to 800 m (1,600 to 2,600 ft) within the potential repository area. The higher-temperature repository layout has the waste emplacement drifts at a depth ranging from about 200 to 500 m (660 to 1,600 ft) below ground surface and between about 210 and 390 m (690 and 1,300 ft) above the water table. These distances were estimated using the Site Recommendation Subsurface Layout (
BSC 2001d, Table V-1), topographic data from the Geologic Framework Model (GFM 3.1) (CRWMS M&O 2000bs, Section 4.1), and water table elevations (USGS 2000c, Table I-1; CRWMS M&O 2000g, Table 3).

Some of the most important site characterization data come from surface-based vertical boreholes. The first deep boreholes at Yucca Mountain were used to define site lithostratigraphy, to locate the water table, to collect core for rock-property analyses, and to test in situ borehole monitoring techniques. The lithostratigraphic description of the tuff units has been refined with coring, mapping, and geophysical logging data from additional surface-based boreholes and confirmed by data collected in the underground drifts (i.e., the Exploratory Studies Facility and Cross-Drift tunnels). Figure 4-9 illustrates the locations of selected deep boreholes, the underground Exploratory Studies Facility, and the Enhanced Characterization of the Repository Block (ECRB) Cross-Drift. A summary of the geology of Yucca Mountain is presented in Section 1.3.

The division of tuff units into members, zones, and subzones is based on variations in degree of welding (compaction and fusion at high temperatures), abundance of lithophysae (cavities formed by bubbles of volcanic gases trapped in the tuff matrix during cooling), depositional features, crystal content, mineral composition, pumice and rock fragment abundance, and fracture characteristics. These features of site lithostratigraphy are described by Moyer et al. (1996, pp. 7 to 54) and Buesch et al. (1996, pp. 4 to 16) and in the Unsaturated Zone Flow and Transport Model Process Model Report (CRWMS M&O 2000c, Section 3.2).

Early geologic mapping of Yucca Mountain was done by Scott and Bonk (1984) and was later updated and refined by Day, Potter et al. (1998) to include additional small faults at the land surface (such as the Sundance fault in the potential repository block) (Spengler et al. 1994, pp. 9 to 11). The fracture patterns in the bedrock were also mapped in pavement studies (i.e., with thin soil covers removed) and in shallow-pit studies of fractures exposed on the pit walls. Mapping and sampling data along transects, especially along washes and valleys, together with logs and core samples from deep boreholes and regional geophysical surveys, were used to construct early stratigraphic and structural models. Current three-dimensional geologic models rely heavily on surface-based vertical boreholes, as well as the Day, Potter et al. (1998) geologic map (CRWMS M&O 2000bs, Section 4). These geologic models are used as a framework for development of numerical models for simulating unsaturated zone flow and transport processes (CRWMS M&O 2000br).

4.2.1.2.3 Pneumatic Data

Several deep boreholes at Yucca Mountain have been instrumented in the unsaturated zone and continuously monitored to record changes in pneumatic pressure and gas flow with depth in response to changes in barometric pressure of the atmosphere (
CRWMS M&O 2000b, Section 8.4.2). Changes in atmospheric pressure are transmitted very rapidly throughout the TCw because of its high fracture permeability. In contrast, the PTn significantly attenuates the atmospheric-pressure signal and imposes a time delay to signal arrival because of higher porosity and water content and much lower fracture density and bulk permeability than the TCw. In general, attenuation of the atmospheric-pressure signal across the TSw is negligible and pressure signals are transmitted nearly instantaneously throughout most of the entire vertical section of the TSw. Nearly all the pressure data from the TSw indicate that the fractures within the TSw apparently are very permeable and highly interconnected within both the lithophysal and nonlithophysal units. In situ pressure records indicate that essentially all of the remaining barometric signal is attenuated by the CHn, primarily due to low permeability and the presence of perched-water zones.

When the Exploratory Studies Facility was excavated, the effects on in situ pneumatic pressure were carefully monitored to determine how the overall gaseous-phase system in the unsaturated zone was affected by direct exposure of deeply buried rock units to atmospheric pressure by way of the tunnel. Pneumatic monitoring data indicate that some faults, such as the Drill Hole Wash fault, transmit pneumatic-pressure signals over distances of several hundred meters nearly instantaneously. The data also indicate that gas-phase flow from the atmosphere into the TSw was essentially short-circuited when the Exploratory Studies Facility penetrated the PTn, locally removing this pneumatic barrier (CRWMS M&O 2000b, Section 8.4.2).

4.2.1.2.4 Matrix Properties

Rock matrix hydrogeologic properties (including porosity, bulk density, particle density, water content, saturated hydraulic conductivity, moisture retention characteristics, and saturation) have been measured for several thousand core samples recovered from 8 deep boreholes and over 30 relatively shallow boreholes. The deep boreholes penetrate to at least the bottom of the TSw, while the shallow boreholes penetrate only to the top of the TSw. Collection, processing, and preliminary analyses of the core data are described by
Flint, L.E. (1998). Results show that nonwelded tuffs (i.e., PTn and CHn) usually have large matrix porosities (typically 30 to 50 percent), while the densely welded tuffs (i.e., TCw and TSw) tend to have greatly reduced pore space (generally less than 15 percent matrix porosity). Because of variations in capillary strength, liquid saturations are usually lower in the nonwelded tuffs than in the welded tuffs, unless the nonwelded tuffs contain significant amounts of clay or zeolite. Mean matrix permeability ranges from about 10-15 to 10-12 m2 (10-14 to 10-11 ft2) for the unaltered nonwelded tuffs and from about 10-18 to 10-15 m2 (10-17 to 10-14 ft2) for the welded and altered (i.e., zeolitized) nonwelded tuffs.

Numerical models of flow and transport use matrix property data to estimate effective hydrogeologic properties for each layer within the model. Upscaling of rock properties (particularly permeability) from core scale to mountain scale is required for the larger, three-dimensional numerical models. Upscaling is an adjustment to the estimate of the effective property when the property data are collected on one scale but the estimate is intended for use in simulations and predictions on a much larger scale. Additional data from in situ water potential measurements are used to estimate moisture-retention characteristics. Collection and preliminary analysis of these data are described by Rousseau, Loskot et al. (1997, Section 4.2) and Rousseau, Kwicklis et al. (1999, pp. 143 to 151).

4.2.1.2.5 Fracture Properties

Fracture data are very important to the characterization of unsaturated flow at Yucca Mountain. Fracture hydrologic properties are estimated using (1) permeability data from in situ air-injection tests conducted in four surface-based boreholes and boreholes in alcoves of the Exploratory Studies Facility; (2) porosity data from gas tracer tests in boreholes in Alcove 5 of the Exploratory Studies Facility; and (3) fracture mapping from the Exploratory Studies Facility, ECRB Cross-Drift, and surface-based boreholes.

Air-injection testing is described by
LeCain (1997, pp. 2 to 9; 1998, pp. 5 to 11) and Rousseau, Kwicklis et al. (1999, pp. 63 to 67) and in In Situ Field Testing of Processes (CRWMS M&O 2000bu, Section 6.1). The data show the TCw unit as having the highest mean fracture permeabilities (as high as 10-10 m2 or 10-9 ft2) and the CHn as having the lowest mean fracture permeabilities (on the order of 10-14 m2 or 10-13 ft2). Mean values of fracture permeability within PTn and TSw layers are approximately 10-13 and 10-11 m2 (10-12 and 10-10 ft2), respectively. As stated previously in this section, most data available is from testing at the upper hydrogeologic units. There are limited air-permeability data for units below the TSw. Air-permeability data for the CHn are available from a single sampled interval in borehole UE-25 UZ#16. This interval lies within a zeolitically altered portion of the upper CHn.

Detailed line surveys and full peripheral mapping of fracture networks have been conducted along the Exploratory Studies Facility and ECRB Cross-Drift by Barr et al. (1996, pp. 133 to 135) and Albin et al. (1997, Appendix 1) (also see data sources listed in CRWMS M&O 2000c, Attachment I, Table I-4). Additional features have been mapped, including the observations of an intensely fractured zone in the southern part of the Exploratory Studies Facility main drift (Buesch and Spengler 1998, p. 19) and several recently discovered faults that show no surface expression in the western part of the ECRB Cross-Drift (see data sources listed in CRWMS M&O 2000c, Attachment I, Table I-4). The fracture density distributions from geologic mapping and geophysical imaging are illustrated in Figure 4-10. Nonwelded tuffs typically have few fractures (less than 1 per meter), while the densely welded tuffs generally have abundant fractures (approximately 1 to 4 per meter) (CRWMS M&O 2000bt, Section 6.1.2.3). Each lithostratigraphic unit generally has its own fracture network characteristics (fracture spacing, intensity, and connectivity) that are controlled by variations in lithology and degree of welding (Rousseau, Kwicklis et al. 1999, p. 23). Fracture characteristic data are used to estimate the potential for, and distribution and amount of, fracture flow within the welded and nonwelded layers in the unsaturated zone. It is important to reiterate that the exploratory tunnels do not penetrate units below the TSw; therefore, fracture characteristics below the TSw are available only from deep boreholes. There are limited fracture data from units below the TSw. For the CHn, vitric and zeolitic fracture frequencies are available from two boreholes (USW SD-12 and USW NRG-7a). Fracture frequencies and properties have not been determined for the CFu unit, which underlies the CHn and is present in the unsaturated zone only along the western margin and in the southwestern part of the repository area (CRWMS M&O 2000c, Section 3.2.2.5, Figure 3.2-3).

Some fracture property data were considered but not used in unsaturated zone flow modeling (CRWMS M&O 2000c, Section 3.6.3.2). In particular, fracture frequency data from the surface of the mountain, measured on outcrops and at the Large Block Test area, are not considered representative of fracture frequencies in the deep subsurface because unloading, or the absence of overburden at the surface, is likely to produce enhanced fracturing. Permeability data from air-injection testing in four Exploratory Studies Facility niches were not used because the scale on which the measurements were made was 0.3 m (1 ft); thus, the data may not be representative of the fracture permeability at the scale of interest. The scale of the air-injection test data that are used for flow modeling is from 1 to 12 m (3 to 40 ft) (CRWMS M&O 2000bt, Section 6.1.1.1).

4.2.1.2.6 Fault Properties

Fault permeability measurements are described by
LeCain (1998, pp. 19 to 22). Direct measurements of fault-specific properties have been conducted using air-injection tests in Exploratory Studies Facility Alcoves 2 and 6 (the Bow Ridge Fault Alcove and the North Ghost Dance Fault Access Drift, respectively). Analyses of cross-hole tests run in the Bow Ridge Fault Alcove (LeCain 1998, p. 21) and the North Ghost Dance Fault Access Drift (LeCain et al. 2000, Table 8) give estimates of fracture permeability in the TCw and TSw fault layers, respectively. These data indicate that, within the welded units, the fractures in the faults are more permeable and porous than the fractures in the formation (or nonfaulted rock).

4.2.1.2.7 Evidence for Fracture Flow

Currently, the estimates of percolation flux at Yucca Mountain range from 1 to 10 mm/yr (0.04 to 0.4 in./yr). Given the low matrix permeabilities of the welded tuffs (i.e., TCw and TSw), a large fraction of the percolation flux in the welded units must travel through fracture networks. This conceptual model is supported by the presence of relatively high fractional abundances of chlorine-36 measured in TCw samples from boreholes (
CRWMS M&O 2000bv, Section 6.6.3). The source of the elevated (bomb-pulse) chlorine-36 has been attributed primarily to nuclear testing in the Pacific Ocean conducted in the 1950s, and its occurrence in the TCw indicates the presence of fast pathways for water flow within the unit (CRWMS M&O 2000bq, Section 6.1.2).

Episodic infiltration pulses are expected to move rapidly through the fracture system of the TCw unit with little attenuation by the rock matrix. This conceptual model of minimal flow attenuation by the densely welded matrix is partially supported by pneumatic data for the TCw unit (CRWMS M&O 2000bq, Section 6.1.2), which show little attenuation of the barometric signal in monitoring boreholes relative to the barometric signal observed at the land surface (Rousseau, Kwicklis et al. 1999, p. 89).

Other evidence for fracture flow comes from calcite-coating data, which are signatures of water flow history and indicate that most of the deposition is found within the fractures in the welded units (Paces, Neymark, Marshall et al. 1998, p. 37). As discussed in Conceptual and Numerical Models for UZ Flow and Transport (CRWMS M&O 2000bq, Section 6.1.2), carbon-14 ages of the perched water bodies below the TSw unit also suggest fracture-dominant flow within the TSw. These ages range from approximately 3,500 to 11,000 years (Yang, Rattray et al. 1996, p. 34), which is much younger than if the major path for water flow within the TSw was through the matrix (CRWMS M&O 2000bq, Section 6.1.2).

Field testing within Exploratory Studies Facility alcoves and niches also supports prevailing fracture flow within welded units. Figure 4-11, for example, summarizes the results of liquid release testing in Alcove 6, located within the fractured, densely welded, middle nonlithophysal unit of the TSw.

4.2.1.2.8 Evidence for Flow Attenuation in the Matrix

Liquid release tests conducted in Alcove 4 of the Exploratory Studies Facility, situated in nonwelded tuffs of the lower PTn, support the matrix flow-damping conceptual model of the PTn (
CRWMS M&O 2000bu, Section 6.7). The Alcove 4 test bed includes a small fault within which tracer-tagged water was released, as illustrated in Figure 4-12. A mass-balance approach involving recovery of outflow in a slot was adopted. The matrix of the PTn effectively damped flow pulses along the fault. The data in Figure 4-12 show a slow decline in water intake rates with time. One explanation for the slow decline is the possible swelling of clays in the PTn layers and subsequent reduction of fault permeability. Another observation made during liquid release testing was that, with sequential wetting, detection of downgradient increases in saturation occurred faster (i.e., the wetting front moved faster with each liquid release test).

Geochemical data from the Exploratory Studies Facility also support the conceptual model of predominantly matrix flow through the PTn by showing a lack of widespread, elevated (bomb-pulse) chlorine-36 signatures at the base of the PTn (Fabryka-Martin, Wolfsberg, Levy et al. 1998).

4.2.1.2.9 Fracture–Matrix Interaction

Field observations show limited fracture–matrix interaction within welded units at Yucca Mountain. The chloride concentration data indicate that perched water is recharged mainly from fracture water, with a small degree of interaction with matrix water (
CRWMS M&O 2000bq, Section 6.1.3). The small degree of interaction between fractures and the matrix at locations associated with geologic features is also suggested by the presence of bomb-pulse chlorine-36 (CRWMS M&O 2000bv, Section 6.6.3) at the potential repository level in the Exploratory Studies Facility. Studies by Ho (1997b, pp. 407 and 409) show that the match between simulations and observed matrix saturation data is improved by reducing the fracture–matrix interaction significantly.

The concept of limited fracture–matrix interaction in welded tuff at the Yucca Mountain site is also supported by many other independent laboratory tests, as well as theoretical and numerical studies (CRWMS M&O 2000bq, Section 6.1.3). In a number of laboratory experiments without considering matrix imbibition, Glass et al. (1996, pp. 6 and 7) and Nicholl et al. (1994) demonstrated that gravity-driven fingering flow is a common flow mechanism in individual fractures. This can reduce the wetted area in a single fracture to fractions as low as 0.01 to 0.001 of the total fracture area (Glass et al. 1996, pp. 6 and 7), although the matrix imbibition can increase wetted areas of fingering flow patterns in individual fractures (Abdel-Salam and Chrysikopoulos 1996, pp. 1537 to 1538). A theoretical study by Wang and Narasimhan (1993, pp. 329 to 335) indicated that the wetted area in a fracture under unsaturated flow conditions is generally smaller than the geometric interface area between fractures and the matrix, even in the absence of fingering flow. This results from the consideration that liquid water in an unsaturated fracture occurs as saturated segments that cover only a portion of the fracture–matrix interface area. Liu et al. (1998, p. 2645) suggested that in unsaturated, fractured rocks, fingering flow occurs at both a single fracture scale and a connected fracture-network scale, which is supported by the field observations from the Rainier Mesa site (see Section 4.2.1.2.13) and by a numerical study of Kwicklis and Healy (1993). They found that a large portion of the connected fracture network played no role in conducting the flow. Studies have also shown that fracture coatings can either reduce or increase the extent of fracture–matrix interaction. Thoma et al. (1992) performed experiments on coated and uncoated tuff fractures and observed that the low-permeability coatings inhibited matrix imbibition considerably. In contrast, fracture coatings may in some cases increase the fracture–matrix interaction when microfractures develop in the coatings (Sharp et al. 1996, p. 1331).

4.2.1.2.10 Mineralogic and Perched Water Data

The spatial distributions of vitric and zeolitic material within the CHn, along with the characterization of the basal vitrophyre (Tptpv3) of the TSw, are important for understanding the distribution of perched water and for determining potential flow paths for radionuclides.

Perched water has been encountered in a number of boreholes (e.g., USW UZ-14, USW NRG-7a, USW SD-7, USW SD-9, USW SD-12, and USW G-2) at the base of the TSw and above units of zeolitic tuff within the CHn (
CRWMS M&O 2000bw, Sections 6.2.1 and 6.2.2).

The spatial distribution of low-permeability zeolites has been modeled using mineralogic and petrologic data from several boreholes and is presented in the Integrated Site Model (ISM3.1) of Yucca Mountain (CRWMS M&O 2000i, Section 3.4). Figure 4-13 shows the modeled distribution of zeolites for layers within the lower TSw (Tptpv3 and Tptpv2) and the upper CHn (Tptpv1 through Tcpuv). Areas with less than or equal to 3 percent zeolite by weight are considered vitric, or unaltered. The figure shows that zeolites within the CHn are prevalent in the northern and eastern portion of the model domain. The areal extent of the vitric region diminishes with depth and is considered to be largely confined to the fault block bounded in the north and east by the Sundance and Ghost Dance faults, respectively, and in the west by the Solitario Canyon fault. The northern half of the potential repository area is underlain by predominantly zeolitic CHn, while the southern half is underlain by the predominantly vitric upper portion of the CHn. However, below the vitric CHn (yet occurring above the water table) are nonwelded portions of the Prow Pass, Bullfrog, and Tram tuffs that are pervasively altered to zeolites. Thus, there is no evidence to support a model that includes a direct vertical pathway from the potential repository horizon to the water table that does not intersect zeolitic units, except, perhaps, within fault zones.

4.2.1.2.11 Geochemical and Isotopic Field Measurements

Samples have been collected in boreholes and along the Exploratory Studies Facility for geochemical and isotopic measurements. Pore water is obtained by physical extraction from samples of tuff and used for chemical analyses and age dating. Measurements of total dissolved solids and chloride concentration in pore waters are mainly available for nonwelded tuff samples, which generally yield sufficient amounts of water by compression for chemical analyses (
Yang, Yu et al. 1998, pp. 6 to 7). These pore waters tend to be more concentrated in dissolved solids than water flowing in fractures, which indicate that the percolation flux in the rock matrix is limited.

Chemical composition, including the total dissolved carbon dioxide concentration, is measured in pore water samples to determine the origin of calcite and amorphous silica solids found in the unsaturated zone and to characterize the evolution of carbonate as waters percolated downward. Ion-exchange reactions along flow paths generally increase the abundance of cations like calcium and strontium with depth in the tuff units. Deviation from the general trend of geochemical evolution with depth provides information about the effects of faults and other structural features. Strontium participates in ion-exchange with zeolites, and concentrations can be significantly depleted in the zeolitically altered CHn unit, signifying effectiveness of hydrogeochemical processes (Sonnenthal and Bodvarsson 1999, pp. 143 to 147). The calcite and opal deposits found on fracture surfaces (Paces, Neymark, Marshall et al. 1998) and within lithophysal cavities (CRWMS M&O 2000bv, Section 6.10.1.1; CRWMS M&O 2000c, Section 3.9.7.1) are analyzed to evaluate flow in fractures and seepage into cavities over millions of years. Figure 4-14 shows examples of geochemical studies of tuff samples.

Environmental isotope data are used to infer the presence of fast flows from the ground surface through the unsaturated zone. Elevated chlorine-36 isotopic ratios well above background levels are related to global fallout from thermonuclear tests conducted in the Pacific Ocean within the last 50 years. The presence of such elevated concentrations (referred to as "bomb-pulse") of chlorine-36 in the unsaturated zone therefore suggests transport from the ground surface in 50 years or less. The bomb-pulse chlorine-36 signals, first observed in surface-based boreholes and later in the Exploratory Studies Facility at faults and other features, have received much attention and continue to be analyzed (Fabryka-Martin, Wightman et al. 1993; Fabryka-Martin, Wolfsberg, Dixon et al. 1996, Section 5.3; CRWMS M&O 2000bv, Section 6.6.3).

Other environmental isotopes have also been analyzed (CRWMS M&O 2000c, Section 3.8.1). Bomb-pulse tritium concentrations are present at depth in several locations and are associated with pathways for liquid and gas flow. The carbon-14 apparent age of the gas phase in the unsaturated zone increases with depth in borehole USW UZ-1, with less obvious trends in other boreholes. Deuterium and oxygen-18 data reflect climatic conditions at the time of groundwater recharge. Based on the analysis of these two isotopes, pore water in the CHn unit is inferred to have originated either during winter precipitation or during a time of colder climate. Perched water appears to be up to 11,000 years old, based on apparent carbon-14 and chlorine-36 ages. The perched water does not appear to have equilibrated with matrix water, based on major constituent concentrations and uranium isotope data (Paces, Ludwig et al. 1998; CRWMS M&O 2000bv, Section 6.6). Figure 4-15 illustrates examples of isotopic studies of tuff samples.

4.2.1.2.12 Seepage Data

Diversion of flow around underground openings in the potential repository host rock units at Yucca Mountain has been investigated through a series of tests conducted in "niche" openings constructed off the exploratory tunnels. More detailed descriptions of these tests and the associated modeling are provided in supporting documentation (
Wang, Trautz et al. 1999; CRWMS M&O 2000bu, Section 6.2; CRWMS M&O 2000c, Sections 2.2.2 and 3.9).

Four niches were excavated along the Exploratory Studies Facility main drift to test drift seepage processes. The seepage tests were motivated by the observed absence of continuous seepage in the Exploratory Studies Facility drifts following their excavation. For each niche test, pretest characterization of the rock was performed and the openings were excavated using a boom-cutter excavating machine (Figure 4-16). The use of water during excavation was controlled so that it did not spread into the rock above the openings. Seepage into the niche openings was monitored while pulses of water were infused at very low injection pressure into the rock above. Staining dyes were used to monitor the presence of seepage and the extent of seepage water movement within fractures and openings.

Pre- and Post-Excavation Permeability Distribution—Air-injection tests were performed in horizontal boreholes drilled at niches, and air-permeability values were determined before and after excavations (CRWMS M&O 2000bu, Section 6.1). The average excavation-induced increases in permeability are in the range of one to two orders of magnitude (i.e., increases by factors of ten to a hundred). The permeability enhancements could be interpreted as the opening of preexisting fractures induced by stress releases associated with niche excavation (Wang and Elsworth 1999; Bidaux and Tsang 1991). A geostatistical analysis of these postexcavation air permeabilities provided a measure of statistical variability and spatial correlation. The air-permeability field at Niche 2 (located at Station 36+50) was found to be largely random, with fairly weak spatial correlation.

Seepage Threshold Tests—A series of short-duration liquid release tests was performed at Niche 2 (Wang, Trautz et al. 1999; CRWMS M&O 2000bu, Section 6.2). Any water that migrated from the injection boreholes and dripped into the niche opening was captured and weighed. The seepage percentage, defined as the mass of water that dripped into the capture system divided by the total mass of water injected, was used to quantify seepage into the drift from a localized water source of known duration and flow rate. The seepage threshold was estimated for each location from a series of injection tests with decreasing flow rates, continuing until no seepage occurred.

Dye tracers were injected as pulses designed to represent repeated, episodic percolation events, as illustrated in Figure 4-17, for the seepage tests at Niche 2. Seepage flow paths indicated by dye tracers were used to observe wetting-front movement through the fractures. Seepage threshold data and wetting-front movement data were used to estimate the unsaturated hydrologic properties for the fractures (Wang, Trautz et al. 1999; CRWMS M&O 2000bu, Section 6.2).

Drift seepage tests were also conducted at Niche 3 (located at Station 31+07), near the crossover point between the Exploratory Studies Facility Main Drift and the ECRB Cross-Drift, and at Niche 4 (located at Station 47+88), which is in an intensely fractured zone. All the niche studies were conducted behind bulkheads to isolate them from ventilation, so that high-humidity conditions were maintained.

Flow Path Associated with Fault—In the excavation of Niche 1 (located at Station 35+66) in the vicinity of the Sundance fault (Figure 4-18a), a damp feature was observed, as illustrated in Figure 4-18b (Wang, Trautz et al. 1999, pp. 331 to 332). The feature was nearly vertical and approximately 0.3 m (1 ft) wide by over 3 m (10 ft) long. It dried out before the bulkhead could be installed to prevent contact by ventilation air. Full rewetting of this feature was not observed after more than 2 years of observation with the bulkhead closed. Figure 4-18c shows that the Sundance fault is one of several faults and features with bomb-pulse signals detected from chlorine-36 isotopic measurements along the Exploratory Studies Facility, as discussed previously.

Moisture Monitoring—Ventilation needed for underground operations and construction activities may explain the lack of observed seepage in the exploratory tunnels. Ventilation can remove a large amount of moisture, producing a dryout zone around the openings, thereby suppressing seepage. Ventilation would be used during repository operations to remove heat during thermal periods (CRWMS M&O 2000c, Section 2.2.2.2.1). To further investigate seepage processes without the influence of ventilation, two additional bulkhead sealing studies are ongoing. The first is at Alcove 7 in an over 100-m (330-ft) long drift segment that intersects the Ghost Dance fault. The second is at the furthest extent of the ECRB Cross-Drift, in an over 1,000-m (3,300-ft) long drift segment that intersects the Solitario Canyon fault. Both studies use double bulkheads for isolation from ventilation effects. To date, no continuous seeps have been observed in either of these two drift segments (CRWMS M&O 2000c, Section 2.2.2). Results will be documented following the completion of the tests. There have, however, been as yet undocumented reports of observations of the effects of liquid water, including organic growths, in the closed section. The closed sections were selected based on inferences from the unsaturated zone models of areas most likely to show seepage under ambient conditions. The lack of observable seepage could be due to incomplete rewetting of the rock following the period of ventilation. The observed effects of liquid water could be due to processes other than seepage. Condensation related to the effects of heat sources or leakage of moist air through fractured rock around the bulkheads could also lead to these effects. Studies to address this issue are continuing, and results will be documented following the completion of these studies.

4.2.1.2.13 Natural Analogues

Natural and man-made analogues with geological or archaeological records can be used to support long-term predictions of future performance of a repository at Yucca Mountain. Short-term field experiments cannot predict long-term performance, but many examples from the geologic record can be used qualitatively to evaluate aspects of repository performance for tens of thousands of years or more. The Yucca Mountain Site Characterization Project has used analogue studies for testing conceptual models and evaluating coupled processes as both quantitative and qualitative tools. An analogue relates a process or set of properties to a similar process or set of properties, either in another place or another time. Natural and man-made analogues offer direct comparisons for some of the actual long-term processes and functions of a repository. The confidence in the safe emplacement of nuclear wastes in the unsaturated zone may be greatly enhanced with natural analogue studies to supplement field testing and modeling (
CRWMS M&O 2000bp; CRWMS M&O 2000b, Section 13; Stuckless 2000). Figure 4-19 illustrates examples of analogue sites for unsaturated flow, transport, and seepage process evaluations.

Climate and Infiltration—A set of natural analogues related to climate and infiltration is listed in Table 4-5. The current climate and paleoclimate records of several sites in the southwestern and western United States were used to derive the climate and infiltration fluxes for future periods. Infiltration at Yucca Mountain under wetter conditions cannot be measured directly, but infiltration under wetter conditions at a geologically similar location can be measured, and the results can be used quantitatively. Monsoonal climates have been studied at Nogales, Arizona, and Hobbs, New Mexico. Beowawe, Nevada currently has a climate similar to that expected at Yucca Mountain during a glacial-transition period (USGS 2000a, Section 6.6.2).

Future climates can be modeled based on determinations of past climates because climates worldwide have been cyclical for the past 2 million years and are expected to remain cyclical in response to astronomical cycles. There is a calcite deposit at Devils Hole, Nevada, that has been precipitated continuously for the last 500,000 years (Winograd et al. 1992). Minute changes in the composition of this calcite provide an accurate record of the climate in southern Nevada for that time. In a similar fashion, sedimentary deposits at Owens Lake, California (Forester et al. 1999) provide a nearly 425,000-year record of past climates in the region (USGS 2000a, Section 6.5.1). These two records can be used to model likely future climates at Yucca Mountain.

Unsaturated Flow and Fracture–Matrix Interaction—The series of natural analogues for unsaturated flow and seepage processes is listed in Table 4-6. The conceptual and numerical modeling methodologies for the unsaturated zone at Yucca Mountain are applicable to other sites, including Rainier Mesa, Apache Leap, Box Canyon, and other arid and semiarid sites where there is fractured rock. The Peña Blanca site evaluation of uranium migration and the Box Canyon, Idaho, infiltration test modeling are two examples of analogue studies presented in Natural Analogs for the Unsaturated Zone (CRWMS M&O 2000bp). Field sampling of borehole cores, evaluation of migration from veins, and assessment of seepage into mine adits was conducted at the Nopal I mine at Peña Blanca, Mexico (Figure 4-19b). For the Box Canyon site, the modeling approach used in the unsaturated zone model for Yucca Mountain is used to interpret ponded-infiltration results through fractured rock (Figure 4-19c).

The calculated rate and amount of radionuclides that could be transported away from a repository is, in part, dependent upon the interaction between fracture flow and pore water in the matrix. The amount of naturally occurring fracture water at Yucca Mountain has been too small to allow direct observations of these important processes. However, Rainier Mesa (Figure 4-19a), located about 40 km (25 mi) northeast of Yucca Mountain, is also composed of alternating welded and nonwelded tuffs, and both pore and fracture water can be collected due to the higher moisture content. With the matrix close to saturation at Rainier Mesa, there should be limited fracture–matrix interaction. Thordarson (1965, pp. 6, 7, and 75 to 80) noted that typically only portions of fractures carried water, and that the chemical composition of water obtained from fractures was substantially different from that of water samples extracted from the nearby rock matrix at that site, which supports the concept of limited fracture–matrix interaction under conditions of near saturation. Further support for this concept can be found at a field site in the Negev Desert, Israel, where man-made tracers were observed to migrate with velocities of several meters per year across a 20- to 60-m (66- to 200-ft) thick unsaturated zone of fractured chalk (Nativ et al. 1995). Such high velocities could only occur for conditions of limited fracture–matrix interaction.

Seepage—Hydrologic models indicate that much of the water moving through the unsaturated zone will preferentially move around openings such as waste emplacement drifts. This modeling result is supported by the natural preservation of paintings, artifacts, and other remains in caves, man-made openings, and rock shelters (Stuckless 2000). Well known examples of this phenomenon include the Paleolithic caves of France and Spain, which contain paintings as old as 30,000 years in an environment 3 to 4 times wetter than that of Yucca Mountain. Figure 4-19e shows a painting made with charcoal from Chauvet cave in southern France that has been dated by the carbon-14 method to be approximately 32,000 years old (Stuckless 2000, p. 4). In addition to the remarkable preservation of these fragile paintings, the figure shows a common phenomenon for seepage in underground openings: much of the seepage flows down the walls. Current models do not take credit for seepage that would be diverted around waste canisters by this mechanism, and the models are, therefore, conservative in their estimates of the effects of seepage.

Spirit Cave, about 75 miles east of Reno, Nevada, is another example of a natural analogue to a repository. In this cave, several small burial pits were discovered. In one of these pits was found an almost completely intact human body, mummified by the dryness of the climate. Subsequent analyses, including carbon-14 testing, have shown the body to be that of a 45- to 55-year-old male who died around 9,400 years ago. His scalp was complete with a small tuft of hair. He wore a breechcloth of fiber and was wrapped in mats woven of fibers from tule leaves, all highly preserved (Barker et al. 2000).

Spirit Cave Man is not alone in being preserved by the dry air that circulates in the caves and rock shelters of the desert Southwest. Among other preserved biologic remains are numerous packrat middens. These middens consist of twigs, fecal droppings, and other debris cemented together by dried urine. Some of the packrat middens analyzed have been preserved for over 50,000 years (Stuckless 2000, pp. 2 to 4).

Elsewhere, in the western United States, caves and rock shelters have preserved fragile biologic remains for tens of thousands of years (Spaulding 1985; Davis 1990). These examples of natural preservation, plus archaeological artifact preservation (Winograd 1986; Stuckless 2000), support the results of mathematical models that predict waste isolation at Yucca Mountain over similar time periods. In addition to flow and transport data and geological records from analogue sites, archaeological records can be more comprehensible to the public than hydrologic-geochemical data and mathematical-numerical models. Paintings and tools preserved in caves, man-made openings and rock shelters (partial openings), structure integrity of underground dwellings (Figure 4-19f), and artifacts found in underground chambers can be used to demonstrate the possibility of preserving man-made materials surrounded by dry air over geological time scales.

Man-made structures include the subterranean tombs of Egypt that are 3 to 4 thousand years old and contain perfectly preserved wood, fabric, and murals. Although this example is from a drier climate than that of Southern Nevada, similar preservations of murals carved into volcanic rock as much as 2,200 years ago can be found in Buddhist temples carved into volcanic rock in India. In this area of India, the climate is monsoonal, with over four times as much precipitation in just four months as occurs annually in the Yucca Mountain region (Stuckless 2000, p. 19).

A particularly well-studied analogue is the paleolithic cave near Altamira, Spain, which is located in the unsaturated zone of a fractured limestone formation that contains clay-rich layers (Villar et al. 1985). Annual precipitation at this site is approximately six times greater than at Yucca Mountain. Nevertheless, seepage rates into the caves are observed to be a small fraction—about 1 percent—of the calculated percolation flux (Stuckless 2000, p. 6). Also, there is virtually no fluctuation in the observed seepage rate despite monthly changes in the amount of precipitation, even though the unsaturated zone is only about 7 m (23 ft) thick (CRWMS M&O 2000c, Section 3.9.7.3), which supports the hydrologic modeling prediction of the buffering effect of the unsaturated zone.

Seepage was also measured in the tunnels at Rainier Mesa. Most of the tunnels excavated for nuclear testing are in a sequence of zeolitized and fractured tunnel bed units within or below perched water zones. When intersected by tunnels or boreholes, a fraction of fractures and faults yielded significant amounts of water, with the total discharge from one tunnel complex equal to 8 percent of the measured precipitation (320 mm or 12.6 in. per year) (Russell et al. 1987; Wang, Cook et al. 1993). In a tunnel excavated above the perched water bodies in the Paintbrush nonwelded unit, no measurable seepage was observed. While this analogue is not ideal because of the near-saturation of the tuffs and because most of the tunnel penetrations are through the lower nonwelded zeolitic tuff with perched water bodies, it provides a conservative example because of the high precipitation relative to that at Yucca Mountain. The potential waste emplacement drifts would be located in welded tuff above the perched water bodies.

4.2.1.3 Process Model Development and Integration

4.2.1.3.1 Unsaturated Zone Flow Model

The unsaturated zone flow model simulates present and future hydrologic conditions between the ground surface and the water table (CRWMS M&O 2000c, Section 3.7). The flow model integrates site characterization data into a single, calibrated three-dimensional model. It captures infiltration at the ground surface and estimates the distribution of percolating water in the unsaturated zone. It quantifies the movement of water through fractures and through the porous tuff matrix at the site. The model includes the occurrence of perched water, which has implications for lateral diversion of flow along interfaces between rock layers. The flow model incorporates the damping of infiltration pulses from the surface as water percolates downward through the rock layers and also includes the effects of faults on water movement.

The unsaturated zone flow model is a tool used to (1) quantify the movement of moisture as liquid and vapor through the unsaturated zone for present-day and future climate scenarios; (2) develop hydrologic properties and other model inputs for predicting seepage into drifts and radionuclide transport; and (3) provide the scientific basis for representing unsaturated zone flow processes in TSPA. Output from the unsaturated zone flow model that is used by other models consists primarily of infiltration and percolation flux distributions and hydrologic properties for hydrogeologic units and fault zones. The flow model supports these outputs with extensive analysis and interpretation of the flow regime, perched-water occurrence, and chemical and isotopic data from the site.

Input to the flow model includes Yucca Mountain geologic and rock properties data in addition to site characterization data from surface hydrology investigations, borehole testing, field tests performed underground in exploratory tunnels, geochemical sampling, and isotopic sampling of the unsaturated zone. These inputs are among the input data represented in Figure 4-20. Section 3.7.2 of Unsaturated Zone Flow and Transport Model Process Model Report (CRWMS M&O 2000c) describes the development of the component models shown in the figure based on site characterization data. Modeling changes associated with such issues as ventilation and transportation in the drift shadow, thermally coupled processes, with a change of the proposed repository footprint for analysis of a lower-temperature operating mode case are discussed in Volume 1, Section 11.3.4 of FY01 Supplemental Science and Performance Analyses (BSC 2001a). Output from the flow model directly feeds the assessments of drift seepage, thermal-hydrology, and radionuclide transport, which ultimately support the TSPA.

4.2.1.3.1.1 Modeling Assumptions

Numerous methods exist for mathematically representing the unsaturated flow processes occurring at Yucca Mountain. These representations range from the relatively simple to the extremely complex. The existing site characterization data for Yucca Mountain suggest that flow between the land surface and the water table is complex. Even though large amounts of site characterization data exist for certain areas of the mountain, for other areas information is based on natural analogues and other sources. Therefore, any mathematical representation of the unsaturated zone at Yucca Mountain must avoid any unjustified complexity, as well as oversimplifications of the flow behavior beneath the mountain.

Mathematical representations of complex, actual phenomena will always require some simplification. Assumptions are made, with justification provided to lend confidence to the model. Some of the important assumptions made in the development of the unsaturated zone flow model of Yucca Mountain are summarized below.

Selection of Continuum Models—Estimates of the number of potentially water-conducting fractures at Yucca Mountain are on the order of one billion (
Doughty 1999, p. 77). It is not feasible to model discrete fractures at this scale, so a continuum modeling approach is used. This is justified by considering two important aspects of flow and transport behavior that are expected to occur in the unsaturated zone. The first, as further discussed in this section, is that where fast flow occurs, it will do so along a few flow pathways that carry relatively small amounts of water. There are numerous slower flow paths dispersed throughout the unsaturated zone, so the continuum model is a reasonable choice for simulating flow and transport. The second aspect is the coexistence of matrix-dominated flow in nonwelded units and fracture-dominated flow in welded units. This is readily accommodated by continuum models but not by other approaches, such as fracture-network modeling. Because continuum models are relatively simple and straightforward to implement, they are preferred for most practical applications (National Research Council 1996b, p. 331).

Dual-Permeability Models—Flow processes in fractured rock have been studied intensively. In order to make continuum models represent observed flow behavior more accurately, dual-permeability models have been developed that represent the rock matrix and the fracture network by different continua. The two continua are coupled together to represent flow interaction between the fractures and the rock matrix (National Research Council 1996b, p. 380). Darcy's law is used to represent flow in each continuum. The van Genuchten model (van Genuchten 1980), which is widely used to relate capillary pressure and saturation in porous media, is employed for both the fracture and matrix continua. Fracture flow in unsaturated media is believed to occur only in a subset of all fractures (called active fractures), depending on the flux of water moving through the fracture network. For such media, the active fracture model describes the relationship between fracture saturation and the flow interaction between the fracture and matrix continua (Liu et al. 1998; CRWMS M&O 2000bq, Section 6.4.5).

Using the dual-permeability approach, one-, two-, and three-dimensional numerical grids of the unsaturated zone were developed. Figure 4-21 illustrates the flow of information associated with the gridding process, summarizing key input and output. Hydrogeologic data, mainly from the Integrated Site Model (ISM3.1) of Yucca Mountain (CRWMS M&O 2000i), are used to develop numerical grids of the unsaturated zone mountain-scale domain. These grids include layering of hydrogeologic rock units, faults, and different types of rock where the geology is laterally heterogeneous. Fracture properties (CRWMS M&O 2000br, Table 5) are used to formulate the dual-permeability grids. Figure 4-22 illustrates the structural and stratigraphic framework established in the numerical grids of the unsaturated zone domain.

Fracture–Matrix Interaction—An active fracture model was developed by Liu et al. (1998) to account for the portion of fracture surfaces interacting with the matrix. Limited fracture–matrix interaction occurs for a variety of reasons, such as fingering flow (gravity-dominated, nonequilibrium, flow channeling within fractures) and fracture surface coatings. The active fracture model is documented in Conceptual and Numerical Models for UZ Flow and Transport (CRWMS M&O 2000bq, Section 6.4.5). In the model, only a portion of connected fractures are considered to actively conduct liquid water at a fracture network scale.

Transient Versus Steady-State Flow—Temporal variation in the infiltration rate drives the time-dependent or transient nature of flow in the unsaturated zone. The temporal variation of infiltration may be short-term due to weather fluctuations that drive episodic flow or occur over much longer time periods corresponding to climate change. As discussed previously, the PTn is believed to greatly attenuate episodic infiltration pulses such that water flow below the PTn is approximately steady. However, water flow in a relatively small area near the Solitario Canyon fault may be transient because the PTn is not present in that area. Some transience is also expected for liquid flow through isolated fast flow paths that cut through the PTn because of the lack of a significant attenuation mechanism. However, these isolated flow paths are believed to carry only a small amount of water because (1) inferred bomb-pulse chlorine-36 isotopic ratios have been found in only a few locations in the Exploratory Studies Facility; (2) no significant correlation between high matrix saturation and elevated chlorine-36 isotopic ratios has been reported; (3) these discrete fast paths are not associated with large catchment areas involving large volumes of infiltrating water; (4) bomb-pulse signatures of chlorine-36 were not found in the perched water bodies (CRWMS M&O 2000bv, Section 6.6.3); and (5) post-bomb tritium was detected only in one sample from the perched water (in Borehole NRG-7a) but not in any of the other samples (CRWMS M&O 2000bv, Section 6.6.2).

The PTn is expected to damp episodic flow in the unsaturated zone resulting from the temporally occasional nature of storms in the site area and the local nature and wide scattering of storm tracks. Individual locations may receive significant rainfall only once in several years, even in the case of focused runoff along arroyos. Geomorphic character, such as conditions of soil cover and slope, further contribute to the variability of infiltration from storms. Larger-scale variation in infiltration is the result of altitude, which controls seasonal rainfall quantity by influencing storm frequency. In general, the northernmost part of the site area is highest, the central part is intermediate in elevation, and the southern part is lowest. The site-scale pattern of infiltration magnitude as well as that of percolation flux at the proposed repository level reflects the site altitude (Figure 4-23). The percolation flux patterns in the unsaturated zone below the PTn (Figure 4-24) do not reflect differences in infiltration magnitude resulting from local storm tracks or surface conditions in regions of the site that are about at equal altitude because of the damping influence of the PTn.

Percolation flux in the unsaturated zone is expected to be episodic on the scale of seconds, minutes, hours, or days. Percolation is periodic flow especially in unsaturated, interconnected fracture systems because water must accumulate to gravitationally overcome capillary and frictional resistance to flow at tight asperities. The rate of episodic pulses of water movement in such a system is regulated by the overall flux rate, by evaporation, and by imbibition of water flowing in the fractures by the matrix, which together limit the storage necessary to push flow episodes. The volume and period of episodic flow events, therefore, is limited by ambient conditions. No large, long-period episodic flow surge has been observed at Yucca Mountain (BSC 2001a, Section 4.3.5.6.2). Geologic evidence at Yucca Mountain (Whelan, Roedder, and Paces 2001), based on the distribution of secondary minerals deposited in the unsaturated zone, clearly supports a flow regime in which most flow is film flow and void spaces along fracture networks are never filled. Such a flow environment is not conducive to larger-scale, long-period episodic flow. Episodic flow severe enough to impact seepage fractions into drifts and transport rates from the repository to the saturated zone has been proposed as a tentative alternative conceptual model based only on numerical simulation of thermally influenced flux (BSC 2001a, Section 4.2.2). The existence of episodic flow of sufficient magnitude to cause such impact in the nonthermal ambient environment in the unsaturated zone remains very improbable (BSC 2001a, Section 4.3.5.6.2).

Inferred occurrences of bomb-pulse chlorine-36 may exist in the deep unsaturated zone only in association with small volumes of water that flowed rapidly down permeable fault zones. The representation of ambient percolation flux in the unfaulted regions of the deep unsaturated zone in the unsaturated zone flow model and the seepage model as steady flow is appropriate.

The principal variable contributing to the regulation of episodic fracture flow volume and period resulting from capillary resistance is the overall flux rate. Condensation resulting from evaporation due to waste heat may locally cause accelerated percolation flux above the potential emplacement drifts. This local acceleration of percolation flux has the potential to impact seepage flux by generating enhanced episodic flow pulses near the potential emplacement drifts (BSC 2001a, Section 4.3.5). In the cases of higher-temperature operating modes, the boiling front may provide some protection of the emplacement drifts from episodic flow surges. In both the higher- and lower-temperature operating modes, the enhanced episodic flow may impact the seepage fraction in the emplacement drifts during the heating and cooldown periods (BSC 2001a, Section 4.3.5). No impact is expected after cooldown is complete. Sensitivity studies indicate that variability in seepage flow rates has only minimal impact on performance.

Based on these considerations, flow and transport in the unsaturated zone from episodic infiltration is believed to have low consequence; therefore, transient flow behavior is not considered in the unsaturated zone process model.

4.2.1.3.1.2 Unsaturated Zone Flow Submodels

A number of applications of the flow model have been developed for detailed study of the effects of specific hydrogeologic features on flow in the unsaturated zone. These include (1) flow through the PTn unit (
CRWMS M&O 2000c, Section 3.7.3.1); (2) the occurrence of perched water and the effect of alteration within the CHn unit on flow (CRWMS M&O 2000c, Section 3.7.3.3); and (3) the role of major faults as potential conduits or as barriers to flow (CRWMS M&O 2000c, Section 3.7.3.2). Each of these topics is summarized below.

Flow through the PTn and at the Potential Repository Horizon—Flow behavior within the predominantly nonwelded PTn unit was evaluated using the mountain-scale flow model with the three climate states (i.e., present day, monsoonal, and glacial-transition). Table 4-7 compares the proportions of vertical fracture flow and matrix flow at the middle of the PTn unit and at the potential repository horizon. The model results support the hypothesis that matrix flow is dominant in the PTn, taking nearly 90 percent of the total flow, with little variation among the three climate scenarios. This is consistent with observed pneumatic responses to barometric pressure fluctuations, which showed attenuated response within the PTn unit (Ahlers et al. 1999). As indicated in Table 4-7, the conditions change to fracture-dominated flow at the interface between the PTn and the underlying TSw unit (potential repository host rock).

More recent studies with new geochemical field data (BSC 2001a, Section 3.3.3) showed that the PTn acted as a buffer, damping out variations in the transient net infiltration, so that flow beneath the PTn was essentially steady-state. Lateral flow diverted net infiltration above the potential repository area eastward to the Ghost Dance and Drill Hole Wash faults. Flow thus diverted bypassed the potential repository block.

Effects of Major Faults—The effect of faults is important for site characterization because faults could provide direct flow pathways from the potential repository to the water table, which could bypass sorptive layers within the CHn unit that have the capacity to retard migration of radionuclides. Consequently, radionuclides could enter the saturated zone where such faults intersect the water table. Alternatively, faults could benefit potential repository performance if they cause water to be diverted away from emplacement drifts.

Based on the representation of faults in the unsaturated zone flow model, the model simulations indicate that the fraction of flow occurring through the modeled faults, as a percentage of the total flow (through fractures, matrix, and faults), increases with depth. Table 4-8 lists these predicted percentages at four different depth horizons for the three climate scenarios and for the northern and southern parts of the model domain. A recent analysis with a refined model indicates that flow through faults at the PTn–TSw interface may be slightly higher than shown here (BSC 2001a, Section 3.3.3.4.2). The table shows that flow percentages through faults at the water table would be very different in the southern part (where the CHn unit is thinner and highly porous) than in the northern part (where the CHn unit is thicker and altered to zeolites). This indicates that more lateral flow diversion on or within the CHn unit would occur in the northern part of the site area.

The simulations predict that percolation flow in the unsaturated zone will converge into the faults as water flows downward through the geologic units, and that lateral diversion of water into faults occurs mainly in the CHn unit below the potential repository horizon. Some lateral diversion into faults is also predicted to occur in the PTn unit above the potential repository horizon, as indicated in Table 4-8. In addition, although the percentage of fault flow at the water table below the potential repository is predicted to increase as the average infiltration increases, the percentage of fault flow above and at the level of the potential repository horizon is predicted to decrease as infiltration increases.

Studies of the CHn and Perched Water Occurrence—The CHn unit lies between the potential repository horizon and the saturated zone; thus, for groundwater flow and radionuclide transport from the potential repository, the CHn unit has an important role in site performance. Prolonged rock–water interaction in the geologic past has produced low-permeability clays and zeolites within the CHn, particularly in the northern part of the site area. This alteration has important implications for occurrence of perched water, for groundwater flow paths, and for radionuclide transport.

Model simulations have been performed using the three climate states (i.e., modern, monsoonal, and glacial transition) (CRWMS M&O 2000c, Section 3.7). The results generally match the observed perched water table elevations from Yucca Mountain boreholes, provided that percolation flux greater than 1 mm/yr (0.04 in./yr) exists at the site. All modeling results indicate significant lateral flow diversion (40 to 50 percent of the total flow) just above or within the CHn where low-permeability zeolites occur.

The effect of perched water zones on flow through the CHn is best explained by comparing percolation fluxes simulated at the repository level and the water table. Figure 4-24 presents map views of the simulated vertical percolation flux distributions at the potential repository horizon and at the water table. Comparison of these views shows large differences in the percolation flux distributions at the potential repository horizon and at the water table, a result that is consistent with the expectation of significant lateral flow diversion occurring just above or within the CHn unit. Simulated results for the northern part of the site area show that percolation flux at the water table tends to be focused along faults. Simulated results for the southern part of the site area show that percolation tends to flow vertically through the CHn unit as matrix-dominated flow in relatively high-permeability vitric zones. Below these vitric zones, however, are zeolitic layers (above the water table) that laterally divert some of the flow eastward, where it is intercepted by faults (Figure 4-24).

4.2.1.3.1.3 Percolation Flux at the Potential Repository Horizon

Percolation flux through the unsaturated zone is important to performance of the potential repository because it directly controls drift seepage and radionuclide transport and also influences the evolution of temperature and humidity in the emplacement drifts (
CRWMS M&O 2000c, Section 3.7.4.1). Because the low percolation flux through the unsaturated zone at Yucca Mountain cannot be readily measured, results from the unsaturated zone flow model are used to estimate the flux.

At the mountain scale, percolation flux at the potential repository horizon is mainly a reflection of the infiltration distribution at the ground surface. Areas with greater infiltration (i.e., the northern part of the site area and the crest of Yucca Mountain) have greater percolation flux, while areas with less infiltration have correspondingly less percolation flux. Percolation throughout the unsaturated zone reflects the redistribution of infiltration below the ground surface because of lateral flow, fracture–matrix flow partitioning, flow into faults, and other flow processes. Over an infinite area, the average infiltration flux would be equal to the average percolation flux. Over a limited area, such as within the repository footprint, percolation flux may be either greater or less than infiltration because of flow redistribution over a larger area.

Figure 4-25 shows the simulated distribution of vertical percolation flux at the repository level for the three climate states (i.e., present-day, monsoonal, and glacial-transition). Areas of higher percolation (shown in blue) are located principally in the northern part of the site area and along the Solitario Canyon fault. The distribution of percolation flux at the potential repository horizon closely matches the distribution of the infiltration at the ground surface (Figure 4-26) because of the limited lateral flow diversion between the surface and the potential repository level, as discussed previously.

Table 4-9 lists summary statistics for the averaged percolation fluxes within the potential repository footprint. The table indicates that the average percolation within the potential repository horizon is similar to the average infiltration over the entire model domain.

4.2.1.3.1.4 Fracture and Matrix Flow Components

Fracture flow has important implications for seepage flow into emplacement drifts and radionuclide transport and will directly influence the performance of a repository (
CRWMS M&O 2000c, Section 3.7.4.3). The partitioning of flow between fractures and matrix is inferred from model results. Figure 4-27 shows the simulated steady-state distribution of the total percolation flux through both the matrix and fractures at the potential repository horizon for the present-day climate state. In the potential repository host rock, fracture flow controls percolation wherever the total flux exceeds the hydraulic conductivity of the matrix.

Table 4-10 lists the proportion of fracture flux at the potential repository horizon and at the water table as a percentage of the total flux. Calculated results are shown for nine scenarios (i.e., present-day, monsoonal, and glacial-transition climate states, each with upper bound, mean, and lower bound infiltration). Fracture flow dominates both at the potential repository horizon and at the water table. As expected, the percentage of fracture flow is somewhat higher for the future climate scenarios compared to the present-day climate scenario because the total flux is greater. The results from fracture–matrix interaction tests at Alcove 6 in the Exploratory Studies Facility (shown in Figure 4-11) illustrate the dominance of fracture flow for high fluxes.

4.2.1.3.1.5 Supporting Geochemical Analysis

Geochemical processes are useful for estimating and bounding infiltration rates and percolation flux in the potential repository host rock (
CRWMS M&O 2000bv, Section 6.9). Upper-bound limits on infiltration rates and percolation flux at the potential repository horizon have been estimated from geochemical data, using several analyses and models (CRWMS M&O 2000c, Section 3.8). Collectively, these conceptual models and analyses are referred to in supporting documentation as the ambient geochemistry model. The primary data used for calibration and validation in this overview are pore water chloride concentrations, relative abundance of chlorine-36 in pore water (or extracted salts), and calcite abundance.

Calcite Deposition Analysis—Calcite and carbon-14 ages in the unsaturated zone have been used as a tool for estimating percolation fluxes (CRWMS M&O 2000bv, Sections 6.6.4.3, 6.7.2.2, 6.10.1.1, and 6.10.3.9). Modeling studies incorporating reactive transport in unsaturated zone flow simulations were used to investigate the relationship of calcite deposition to infiltration rate, water and gas compositions, and reactive surface area (CRWMS M&O 2000bw, Section 6.5). Model results for borehole USW WT-24 indicated that the infiltration rate ranges from about 2 mm/yr to 20 mm/yr (0.08 to 0.8 in./yr), which bounds the observed range of calcite abundance. An infiltration rate of approximately 6 mm/yr (0.2 in./yr) can account for the average abundance of calcite in the TSw unit (Figure 4-14b).

Chloride Mass Balance—Small concentrations of chloride occur in rainfall at Yucca Mountain, averaging approximately 1 mg/L (CRWMS M&O 2000bv, Section 6.3.2). The chloride becomes more concentrated as evaporation occurs. The water that does not run off or evaporate infiltrates into the unsaturated zone, carrying the dissolved chloride. The resulting concentration of chloride in unsaturated zone pore waters indicates the extent of evaporative concentration relative to rainwater, and the infiltration flux is inferred using the average annual precipitation and runoff.

Chloride concentrations in unsaturated zone pore waters can be used to evaluate the unsaturated zone flow model. The comparison of concentrations for the present-day climate state and mean infiltration (CRWMS M&O 2000c, Section 3.8) are shown in Figure 4-14. The modeled concentrations were higher than measured concentrations toward the northeast end (left side of Figure 4-14a) and lower at the southwest end of the Exploratory Studies Facility (right side). The northeast end of the drift corresponds to an area of very low infiltration rates, whereas the southwest end is beneath the crest of Yucca Mountain, where infiltration is greater. Measured chloride concentrations exhibit a smaller range of variation than is predicted using the present-day, steady-state infiltration rates. The general agreement among these results indicates that the average of present-day infiltration over the model domain for the mean infiltration distribution, 4.6 mm/yr (0.18 in./yr), is accurate.

As an alternative interpretation of the observed chloride data, infiltration rates were adjusted in the UZ flow and transport models to match with the measured pore water chloride concentration data. The match can be achieved by adjusting infiltration rates in those areas where the match can be improved, while maintaining the average infiltration the same as the current model (USGS 2000b, Section 6.11; CRWMS M&O 2000bw, Section 6.4.3.1). The modified percolation flux map is shown in Figure 4-23. The domain was divided into 9 regions, and for those regions where pore water chloride data were unavailable (Regions I, II, and VIII), the map was filled in using average infiltration values from the unsaturated zone flow model.

The modified model has a more uniform spatial distribution of infiltration rates. However, it is possible to have an alternative interpretation if the PTn, between the shallow infiltration zone in TCw and the potential repository unit in TSw, has strong damping capacity and the large lateral diversion of flow occurs in the PTn. The spatial variation of infiltration can remain the same as the current model while flow redistribution occurs in the PTn. Significant damping and lateral diversion of flow by the PTn is strongly supported by the recent analyses (Section 4.2.1.1.2). It is therefore demonstrated that alternative models can be formulated to maintain both the heterogeneous distribution of infiltration near the surface and more uniform distribution of chloride content along the underground drifts in the potential repository unit.

Applying the chloride mass balance approach to estimating percolation (using measured pore water chloride concentration data), a modified percolation flux map was developed (Figure 4-23). The domain was divided into 9 regions, and for those regions where pore water chloride data were unavailable (Regions I, II, and VIII), the map was filled in using average infiltration values from the unsaturated zone flow model. The percolation flux values estimated in this manner were then evaluated as estimates of the infiltration rates (USGS 2000b, Section 6.11). The infiltration rates estimated by chloride mass balance are similar to the mean infiltration rates obtained by averaging the rates over the same area (CRWMS M&O 2000bw, Section 6.4.3.1). This provides confirmation that the infiltration data used in the unsaturated zone flow model represent the present-day climate state.

As an alternative interpretation of these data, chloride concentrations modeled from the mean infiltration rates were compared to the measured chloride concentrations (Figure 4-14). The modeled concentrations were higher toward the northeast end (left side of Figure 4-14a) and lower at the southwest end of the exploratory drift (right side). The northeast end of the drift corresponds to an area of very low infiltration rates, whereas the southwest end is beneath the crest of Yucca Mountain, where infiltration is greater. Measured chloride concentrations exhibit a smaller range of variation than is predicted using the present-day, steady-state infiltration rates. However, the general agreement among these results indicates that the average of present-day infiltration over the model domain for the mean infiltration distribution, 4.6 mm/yr (0.18 in./yr), is accurate.

Chlorine-36 Isotopic Analysis—Measured background chlorine-36 isotopic ratios in extracted pore waters, while highly variable, are uniformly lower over much of the south ramp compared to the north ramp of the Exploratory Studies Facility. Modeled results indicate that chlorine-36 isotopic ratios in the south ramp should be much lower than in the north ramp, even though the PTn unit is thinner, because the infiltration flux is less. The greater background chlorine-36 isotopic ratios in the north ramp could also be the result of mixing of bomb-pulse water with older matrix pore water. Bomb-pulse chlorine-36 isotopic ratios, indicating fast flow paths, are found in several locations in the vicinity of some fault zones in the exploratory tunnels. Elevated chlorine-36 signatures are confined to the immediate vicinity of faults and other structural features, fast flow zones are localized, and large areas of the potential repository appear to be unaffected by fast-path flow (CRWMS M&O 2000bv, Section 6.6). With regard to perched water, the lack of bomb-pulse chlorine-36 is difficult to interpret because of potential mixing with older water. However, low tritium signatures in perched waters below the potential repository horizon are consistent with the interpretation of limited fast flow at this depth.

Because of the important implications of the occurrence of bomb-pulse chlorine-36 to the site-scale unsaturated zone flow and transport model, the project has undertaken a study to confirm the occurrence of bomb-pulse chlorine-36 at two locations in the Exploratory Studies Facility (the Drill Hole Wash fault zone and the Sundance fault zone). Preliminary results from this ongoing study have not confirmed the presence of chlorine-36, and the analyses of the validation samples at two different laboratories are not consistent. The project has defined a path forward to understand the reasons for the differences: a common set of protocols will be developed, and analyses of the validation samples will continue. However, the unsaturated zone flow and transport model and the TSPA are based on the original data set used in the conceptualization of unsaturated zone behavior. Hence, the project approach conservatively bases the model on the presence of bomb-pulse chlorine-36.

4.2.1.3.1.6 Summary and Conclusions from Unsaturated Zone Mountain-Scale Modeling

Results from the unsaturated zone flow model and from analysis of supporting geochemical data are summarized in
Figures 4-28 and 4-29, respectively. In summary, available site data have been used to construct an unsaturated zone flow model that has provided infiltration and percolation flux distributions and hydrologic properties, as intended. The model accounts for the occurrence of perched water and enhances the understanding of the effects of the PTn unit and fault zones on flow in the unsaturated zone at the site. The flow model has been interpreted for comparison with chemical and isotopic data, confirming the flow model results in an average sense.

Upper-bound limits on infiltration rates and percolation fluxes at Yucca Mountain are estimated based on multiple approaches (CRWMS M&O 2000bw, Sections 6.2, 6.4, 6.5, and 6.6). These include analyses of chloride and chlorine-36 isotopic ratios, calcite deposition, and the occurrence of perched water. The analysis of chloride data indicates that the average percolation rate over the model domain at Yucca Mountain is about 4.6 mm/yr (0.18 in./yr). Analysis of calcite deposition gives infiltration estimates of 2 to 20 mm/yr (0.08 to 0.8 in./yr) in the vicinity of borehole USW WT-24. To match perched water occurrences, three-dimensional model calibrations require that the present-day average infiltration rate be greater than 1 mm/yr (0.04 in./yr), with an upper limit of about 15 mm/yr (0.6 in./yr).

4.2.1.3.2 Drift Seepage Model

A qualitative description of drift seepage processes was provided in
Section 4.2.1.1.5. The characterization and modeling approach presented here for the seepage calibration model focuses on obtaining effective hydrologic properties based on relevant seepage test data (CRWMS M&O 2000c, Section 3.9.1). Figure 4-30 schematically shows the relationships between the different seepage models, as well as data input and the exchange of information. Seepage test data, such as those from the niche studies, are used for developing the seepage calibration model. The modeling approach and parameters are then used in the seepage model for performance assessment. This model supports the abstraction of drift seepage for use directly in TSPA calculations.

The seepage calibration model is a porous-medium model of the fracture continuum that has spatially variable permeability based on air-permeability data (CRWMS M&O 2000c, Section 3.9.4.4) and is calibrated to the test data from Niche 2 in the Exploratory Studies Facility (CRWMS M&O 2000c, Section 3.9.4.5). Model calibration was used to determine effective hydrologic properties that represent the potential effects of individual fractures and microfractures on drift seepage. Simulations with multiple realizations of the heterogeneous property field were performed to account for the random nature of the fracture network (CRWMS M&O 2000bx, Section 6.3). The steps involved in development of the seepage calibration model are shown schematically in Figure 4-31.

Model Assumptions—All seepage models presented in this section are single-continuum models, explicitly representing only the fracture network (not the rock matrix). The justification for this approach is based on the conceptual framework that the fracture network is extensive and well connected, and that flow conditions are steady-state or very slowly varying throughout much of the potential repository host rock because of the moderating influence of the overlying PTn unit. Under steady flow conditions, the flow interaction between the fractures and the matrix in the vicinity of the emplacement drifts will be negligible because of the low permeability of the rock matrix relative to that of the fractures (CRWMS M&O 2000c, Sections 3.3.13 and 3.6.3.2).

The continuum approach is valid for simulating drift seepage, as well as percolation flux, based on the observation that the fracture network in the middle nonlithophysal unit of the Topopah Spring Tuff is well connected (CRWMS M&O 2000bx, Section 6.7). The appropriateness of using the continuum approach to simulate flow through fractured rock was also studied by Jackson et al. (2000) using synthetic and actual field data. They concluded that heterogeneous continuum representations of fractured media are self-consistent (i.e., appropriately estimated effective-continuum parameters can represent the underlying fracture network characteristics).

Adopting the continuum approach, unsaturated flow of liquid water is governed by Richards' equation, the governing equation of water content or capillary pressure based on mass conservation law (Richards 1931). Relative permeability and capillary pressure are described according to the van Genuchten-Mualem model (Luckner et al. 1989, pp. 2191 to 2192). Within the heterogeneous property distribution, capillarity of the effective medium is correlated to absolute permeability according to the Leverett scaling rule, with capillary strength inversely proportional to the square root of permeability (Leverett 1941, p. 159).

Since percolation flux cannot be directly measured in the field, the average percolation flux in the host rock is estimated by the unsaturated zone flow model and multiplied by flow-focusing factors derived using the active-fracture concept (CRWMS M&O 2000by, Section 6.3.3).

In-drift evaporation and ventilation effects are not included in the current seepage model, but neglecting them is conservative because it produces greater estimates of predicted seepage. The evaporation reduces drop formation and dripping (Ho 1997a) and enhances the vapor diffusion into the drift. Neglecting evaporation effects increases predicted seepage of liquid water (CRWMS M&O 2000c, Section 3.9.3.3).

Seepage Threshold Prediction—Steady-state seepage simulations were performed with the seepage calibration model (CRWMS M&O 2000bz, Section 6.6). In this application, percolation flux was applied at the top of the model, instead of from a borehole as in the niche tests. The ambient percolation flux was varied over a wide range, starting from a small value yielding zero seepage, increasing stepwise until seepage was predicted to occur, and increasing further to estimate seepage percentage. Using this procedure, a seepage threshold of approximately 200 mm/yr (7.8 in./yr) was obtained for the middle nonlithophysal unit of the Topopah Spring Tuff. The seepage-threshold prediction obtained for Niche 2 suggests that diversion of flow around the emplacement drift openings is an effective barrier to water that could otherwise contact waste packages (BSC 2001a, Sections 4.2.2 and 11.3.1.1.1).

4.2.1.3.3 Model Calibration and Validation

An important objective for the unsaturated zone flow model is to produce a model consistent with the available site characterization data. This is accomplished through an iterative process of model calibration to the data, adjusting the hydrologic properties that represent the rock units (
CRWMS M&O 2000c, Section 3.9.4.5). A combination of one-, two-, and three-dimensional numerical models is used to represent the lateral variation of hydrologic conditions in the unsaturated zone, for example, from the northern to the southern ends of the potential repository layout area. The data that constrain the calibration process include borehole-measured matrix saturation, water potential, temperature, the presence of perched water, pneumatic-pressure measurements, and geochemical data. The model calibrations require specification of the water infiltration at the ground surface and its variation throughout the site area (USGS 2000b, Section 6.11). To represent uncertainty in the infiltration estimates, separate infiltration distributions are used, representing upper-bound, mean, and lower-bound conditions. Separate model calibrations (i.e., hydrologic property sets) are developed for the three infiltration distributions. The distributions of infiltration for future monsoonal and glacial-transition climate conditions were developed using the infiltration model (CRWMS M&O 2000c, Section 3.5.2), which was calibrated using data for present-day infiltration rates. Table 4-11 summarizes average precipitation and infiltration rates. Figure 4-25 shows mean infiltration distributions over the model domain for each climate state.

Model calibrations were performed using one-dimensional numerical grids to estimate mountain-scale hydrologic properties for the hydrogeologic units. The one-dimensional calibration model consists of 11 vertical columns, shown schematically in Figure 4-32, representing the hydrostratigraphy at 11 boreholes for which suitable site characterization data are available. The use of a one-dimensional vertical model implicitly assumes that flow is adequately approximated as one-dimensional and vertical (i.e., that lateral flow is not important). In the TCw, PTn, and TSw rock units, this assumption is supported by the absence of perched water (CRWMS M&O 2000c, Sections 3.6.4.1). At the bottom of the TSw unit and below, perched water does exist in some areas, especially to the north. Perched water is investigated using the three-dimensional model (CRWMS M&O 2000c, Section 3.7.3.3).

The one-dimensional calibrated mountain-scale formation properties are then used as input to a two-dimensional model, which is used to calibrate the fault properties. The two-dimensional model consists of an east–west cross section, shown schematically in Figure 4-33. This cross section is located where there are borehole data available for the ambient hydrologic conditions in a fault at Yucca Mountain. Use of a two-dimensional model implies that flow constrained to the vertical and east–west directions adequately represents ambient conditions. The dip of the bedding and the strike of the fault are approximately parallel to the cross section; therefore, the assumption is reasonable. The same types of properties calibrated for the mountain-scale formation properties are calibrated for the faults. Figure 4-33 also shows the match between the calibrated simulation and the saturation, water potential, and pneumatic data for present-day ambient conditions (mean infiltration).

The resulting calibrated hydrologic properties (including fracture and matrix permeability, fracture and matrix van Genuchten parameters, and active-fracture parameters) are then used as input to mountain-scale and drift-scale hydrologic models. Calibration activities in three dimensions (involving hydrologic measurements, perched water, temperature, and ambient geochemical data) are carried out for additional refinement of mountain-scale properties. Figure 4-34 shows an example of the results from the three-dimensional calibration. This figure compares the observed and simulated matrix saturation and perched water elevation at borehole USW UZ-14, using the present-day ambient conditions (mean infiltration). Overall, the simulation results for this borehole (and others used in the calibration activity) are generally consistent with the observed saturation and perched water data.

Model Validation and Confidence Building—Validation of the calibrated hydrologic property sets constituting the calibrated properties model was performed using the three-dimensional model (CRWMS M&O 2000bw, Section 6.8). In situ water potential data measured from the ECRB Cross-Drift compare well with predicted water potentials, as shown in Figure 4-35. Although the predicted water potentials are generally lower, the difference is only a few tenths of a bar (1 bar = 105 Pa). Pneumatic pressure data also compare well to the predicted pneumatic pressures shown in Figure 4-35. These and other results presented in Unsaturated Zone Flow and Transport Model Process Model Report (CRWMS M&O 2000c, Section 3.6.5.3) show that the calibrated properties are valid for predicting ambient conditions, and that the assumptions of one-dimensional and two-dimensional flow are suitable for use in the calibration process.

Pore water chemical composition data have been used to validate the unsaturated zone flow model to bound the infiltration flux, flow pathways, and transport time through the unsaturated zone. Infiltration rate calibrations are performed using the pore water chloride concentration data. Agreement between the predicted chloride distributions and observed data are improved when the calibrated infiltration rates are used. Similar analyses have been performed using calcite deposition to further constrain hydrologic parameters, such as infiltration flux. These geochemical studies provide additional support for validation of flow and transport models (CRWMS M&O 2000bw, Sections 6.4 and 6.5; CRWMS M&O 2000c, Sections 3.7 and 3.8).

Seepage Model Calibration—Data from five tests (CRWMS M&O 2000bz, Table 6) were selected for calibration of the seepage model. The five tests were conducted in a 0.3-m (1-ft) long borehole interval at various injection rates to sample the dependence of seepage on flux. Approximately 1 liter of water was injected in each test. Seepage percentages from the test series demonstrate storage effects and seepage rates above and below a seepage threshold. Analysis of all five tests provided a match between model results and observed seepage. As shown in Figure 4-36, the heterogeneous three-dimensional seepage calibration model matches the observed seepage data better than the two-dimensional or homogeneous alternatives.

Seepage Model Validation—The seepage calibration model was used to make predictions of observed seepage percentages from liquid injection tests that were performed in a different borehole interval using different injection rates and varying other test conditions. The uncertainty of the model predictions was evaluated using linear error propagation analysis and Monte Carlo simulation. (This approach reflects the intended use of the resulting seepage models, in which seepage is treated statistically.) Observed seepage percentages lay within the uncertainty range of the model predictions. This favorable result provides confidence in the validity of the seepage model. More details about the seepage calibration model can be found in Seepage Calibration Model and Seepage Testing Data (CRWMS M&O 2000bz).

4.2.1.3.4 Alternative Conceptual Processes

In developing models of water movement through the unsaturated zone at Yucca Mountain, alternative conceptual processes were identified and considered (
CRWMS M&O 2000c, Section 1.2.4). This section summarizes some key alternative concepts for processes governing water movement through the unsaturated zone.

PTn Lateral Flow—In an early conceptual model of Yucca Mountain, Montazer and Wilson (1984, pp. 45 to 47) hypothesized that significant lateral flow occurs within and above the PTn unit because of the contrast in hydraulic properties at the contact between the TCw and PTn units. They also showed that vertical heterogeneities within the PTn may result in a much larger effective permeability of the unit in the direction of dip, compared with the effective permeability in the direction normal to the bedding plane. Montazer and Wilson (1984, p. 47) argued that the combination of this factor and capillary barrier effects might introduce considerable lateral flow within the unit. Recent modeling of the potential for diversion on or in the PTn supports the Montazer and Wilson conceptual model of lateral flow and diversion. Pneumatic measurement of permeability on, in, and below the PTn, geochemical data, and saturation and water potential data were used to calibrate unsaturated zone parameters and to differentiate alternative conceptual models. Modeling based on the capillary barrier effect using a fine grid spacing supports the concept of diversion of flow above the potential repository horizon (BSC 2001a, Section 3.3.3.4.2). Diversion of flow above the repository would be beneficial to the performance of the repository. It is not necessary for diversion large enough to result in diversion to faults to occur for the PTn to damp episodic flow in the unsaturated zone, but diversion might be an additional mechanism for uniformly distributing the areal variation of infiltration.

PTn Fracture Flow—An alternative conceptual model for water flow through the PTn is one that assumes pervasive fracture flow through this unit (CRWMS M&O 2000c, Section 3.3.3). However, the available data, which show high matrix porosity and storage capacity combined with relatively high matrix permeability and limited fracturing (see to Sections 4.2.1.2.3 through 4.2.1.2.5 [4.2.1.2.3, 4.2.1.2.4, 4.2.1.2.5])—in addition to geochemical data that indicate a lack of widespread bomb-pulse chlorine-36 signatures in the PTn (see Section 4.2.1.2.8)—support the preferred conceptual model of predominantly matrix flow through the PTn.

Episodic Flow Within the TSw—In the prevailing conceptual model, episodic flow into the TSw unit is considered to be damped by the PTn to the extent that flow can be regarded as steady-state when it enters the TSw. The exception is that at or near major faults, episodic flow may still persist through the PTn, though these isolated fast flow paths are considered to carry only a small amount of water. An alternative view to this conceptual process is one in which episodic flow is pervasive through the TSw unit (CRWMS M&O 2000c, Section 3.7.3.1). This alternative may be a more conservative conceptualization because a greater amount of episodic flow could lead to greater seepage into potential waste emplacement drifts. However, for the reasons presented in Section 4.2.1.3.1.1 (in the discussion of transient versus steady-state flow), flow and transport in the unsaturated zone has been found to have low consequence; thus, the approach that is regarded as more plausible has been taken.

Flow Through Faults—As discussed in Section 4.2.1.2.6, limited fault permeability measurements are only available for the welded units, TCw and TSw. From these data, it is inferred that faults within the PTn and CHn/CFu units have relatively higher permeabilities than the adjacent nonfaulted rock. An alternative view is one in which faults within the PTn and CHn/CFu units have low permeabilities and retard the movement of water because of the occurrence of low-permeability alteration minerals within the fault zones (CRWMS M&O 2000c, Section 3.3.5). This conceptual model would result in slow transport from the TSw to the water table. Available measured data are insufficient to confirm either conceptual model. However, the conceptualization of faults as higher-permeability structures in the PTn and CHn/CFu units is adopted because it is the conservative approach, providing fast flow pathways to the water table and allowing discharge from perched water bodies.

Discrete Fracture Network Model for Seepage—A discrete fracture network model is an alternative conceptual model to the heterogeneous fracture continuum model (CRWMS M&O 2000bz, Section 6.7). A high-resolution discrete fracture network model, in principle, should be capable of simulating channelized flow and discrete seepage events. However, the development of a defensible discrete fracture network model for the unsaturated zone at Yucca Mountain would require the collection of geometric and hydrologic property data on billions of fractures. While some of the required geometric information could be obtained from fracture mappings, the detailed description of the fracture network would be incomplete and highly uncertain. Moreover, measurement of unsaturated hydrologic parameters on the scale of individual fractures would be required, which would be largely impractical. The development of a discrete fracture network model at the mountain scale is therefore impractical (CRWMS M&O 2000bz, Section 6.7), so an equivalent fracture continuum model is used to represent the fracture system for the prediction of effective seepage quantities. The appropriateness of this model is demonstrated by Finsterle (2000).

4.2.1.3.5 Limitations and Uncertainties

The assessment of the current understanding of flow through the unsaturated zone needs to take into account the limitations and uncertainties in the unsaturated zone flow model. The model is limited mainly by the current characterization of the unsaturated zone flow system within Yucca Mountain, including geologic and conceptual models; by the applicability of the volume-averaging modeling approach; by the assumption of steady-state moisture flow; and by the available field and laboratory data (
CRWMS M&O 2000c, Section 3.7.4.5). Remaining uncertainties in the model parameters include (1) accuracy in the estimated present-day, past, and future net infiltration rates over the mountain; (2) quantitative descriptions of heterogeneity within the welded and nonwelded tuff units, the flow properties associated with these units, and the detailed spatial distribution of these units within the mountain, especially below the potential repository horizon; (3) sufficiency of field studies and data, especially for characterizing hydrologic properties of faults and fractures within the zeolitic units; (4) alternative conceptual models quantifying the fluid transmissive properties of faults; and (5) evidence for lateral diversion above or within the zeolitic portions of the Calico Hills nonwelded hydrogeologic unit beneath the potential repository horizon (CRWMS M&O 2000c, Section 3.7.4.5).

As noted in Section 4.1.1.2, the DOE has initiated several activities to improve the treatment of uncertainty in current models (BSC 2001a, Section 2.1). Some of the unsaturated zone models will be updated as a result of those activities. Those updates will be documented in future reports.

The identification and propagation of uncertainties is important for the appropriate treatment of the models used in TSPA-SR calculations. Uncertainties associated with the major unsaturated zone model components are described in detail in the Unsaturated Zone Flow and Transport Model Process Model Report (CRWMS M&O 2000c, Section 3.13).

The uncertainty and variability in the model parameters are due, in part, to the natural variability and heterogeneity in the geological, hydrologic, chemical, and mechanical systems that are difficult to characterize in situ, such as the precise fracture network in the unsaturated zone. Uncertainties in models may also be due to conditions that are difficult to predict, such as future climate states.

Uncertainties Associated with Climate—As described in Future Climate Analysis (USGS 2000a, Section 6.5), future climates cannot be predicted in any precise way. Uncertainty arises because of the complexity of global climate systems and because climate changes can be triggered by unforeseen circumstances, such as major tectonic events (e.g., volcanic eruptions) or human activity. However, studies of past climates (paleoclimate studies) demonstrate that changes can be correlated with cyclical variations in the earth's orbit and the tilt of the earth's axis of rotation, both of which affect the amount of solar radiation the atmosphere receives. The earth's orbit changes in a regular and predictable manner over a cycle of about 400,000 years. Paleoclimate studies, which include geochemical analyses of sediments deposited in lakes, minerals deposited in springs, fossils of microorganisms that lived in both lakes and springs, and plant and animal remains preserved in caves, suggest that the sequence of climate changes in the 400,000-year cycle is not random, and that future climate conditions will evolve systematically. The variability of the glacial-transition period has large impacts to the model predictions as the results of long duration and high precipitation values. In the TSPA-SR, a conservative approach is adopted, with the glacial-transition climate extending beyond 10,000 years and the shorter and drier interim periods not taken into account (CRWMS M&O 2000a, Section 3.2.1.2).

Although future climate conditions cannot be precisely predicted and climate varies considerably even within glacial and interglacial periods, studies provide a reasonable basis for forecasting the range of climates Yucca Mountain will probably experience in the future (USGS 2000a). This forecast, which incorporates the variability observed in studies of the past climate, has been used as input to models that assess the future performance of a repository at Yucca Mountain. The present warm, dry interglacial period will probably end in the next 400 to 600 years, and may be followed by a transition to a warm, wet monsoon climate for approximately 900 to 1,400 years. The climate would then shift to a glacial-transition period. The variability of the climate conditions is quantified by upper bounds for wetter climates and lower bounds for drier climates, as presented in Section 4.2.1.3.1.

Uncertainties Associated with Lower Tuff Units—The relatively high density of data in the potential repository area, particularly within and above the TSw, helps to reduce uncertainty in the understanding of flow behavior between the land surface and potential repository horizon. Greater uncertainty exists, however, below the TSw (i.e., within the CHn and CFu) because rock hydrologic properties are highly variable and few data are available to capture the spatial variability. Modeling uncertainties also increase rapidly with lateral distance away from the potential repository area as the density of data points is greatly reduced.

As a result, many of the hydrologic properties used in unsaturated zone modeling studies of Yucca Mountain for layers within the CHn and CFu have been estimated using analogue data from the PTn, the TSw, and portions of the CHn for which data are available. Despite similar welding characteristics, the PTn data (specifically, fracture permeability) used as an analogue to the CHn tend to be conservative because of many inherent differences in the depositional and postdepositional history of these tuffs. Fault properties were estimated for the CHn/CFu, derived from in situ fault testing in the TSw; however, there are different welding textures associated with each major unit. For example, faults may be more permeable within the TSw because the brittle nature of the densely welded tuffs can lead to the development of well-connected fracture networks. Within the CHn, however, the predominantly nonwelded tuffs are likely to exhibit more plastic deformation (producing fewer well-connected fractures) and are much more susceptible to alteration (producing low-permeability clays and zeolites that hinder vertical flow) when exposed to percolating water.

Uncertainties Associated with Calibrated Property Values—Calibrated rock hydrologic property values derived from core-sample measurements, fracture mapping, and in situ field data provide important input to the unsaturated zone model in that they define the hydrologic characteristics of each cell within the numerical grid. Uncertainties related to calibrated hydrologic properties include (1) variability in measured properties of rock core samples and uncertainty in cross-correlations between measured properties; (2) spatial variability in rock properties; (3) uncertainty in the initial estimates of rock properties and in upscaling of measured data to model grid blocks; and (4) nonuniqueness of results generated by the estimation procedure (CRWMS M&O 2000c, Table 3.13-1). With a total of 194 calibrated properties (all the calibrated hydrogeologic unit and fault properties), the quantitative estimation of uncertainty is complex. Cross correlations between some of the properties tend to compound the uncertainties assigned to individual properties. Furthermore, the statistical assumptions that underlie the uncertainty analysis implemented in the estimation procedure are not justified if the estimation uncertainties become very large because of cross correlations. However, the calibrated property values generally do not vary much from the initial values input to the estimation procedure, and the initial values are chosen to be plausible. Because the initial and calibrated properties are generally similar, the uncertainties of the initial property estimates can be used as surrogates for the uncertainties of the calibrated property values (CRWMS M&O 2000c, Section 3.6.5.2).

Where the calibrated properties change significantly with respect to the initial values, the numerical model produces results that differ from the initial interpretation. For example, the transitions between matrix-dominated and fracture-dominated flow, at interfaces between nonwelded and welded tuff, depend on processes that occur at scales smaller than the numerical grid spacing. Consequently, the calibrated property values at these interfaces reflect the response of the numerical model, which uses coarser spacing.

Uncertainties Associated with the Numerical Approach—An additional uncertainty in unsaturated zone modeling is the mathematical representation of complex flow phenomena. The volume-averaging approach used, and the model assumption of steady-state moisture flow (used in interpreting ambient conditions), simplify the representation of water flow through the site, adding uncertainty to the model results.

Uncertainties Associated with Geochemical Analyses—The amount of calcite precipitated over time is sensitive to the water and gas composition, the reactive surface area, and the thermodynamic and kinetic parameters used in the model (CRWMS M&O 2000bw, Section 6.5). Because calcite abundances are highly variable at the different locations sampled (CRWMS M&O 2000bv, Section 6.10.1.1; see also Figure 4-14 in Section 4.2.1.2.12), the calcite analysis approach is suitable for estimating a range of percolation fluxes, as discussed in Section 4.2.1.3.

The use of chloride to estimate percolation fluxes or infiltration rates is directly related to the initial estimate of the effective chloride concentration in precipitation, the spatial variation, and changes over time (Ginn and Murphy 1997, pp. 2065 to 2066). The long-term projection of spatial and temporal patterns of precipitation is uncertain because the patterns have been measured for less than 100 years, which is short compared to the time period over which the chloride concentrations have developed in the unsaturated zone (tens of thousands of years) (Sonnenthal and Bodvarsson 1999, pp. 107 to 111). Given these uncertainties, the chloride mass balance approach is used to estimate ranges in the infiltration rate for comparison to the flow model. The distribution of chloride in the unsaturated zone is also influenced by lateral flow diversion, as well as diffusion and dispersion processes, and thus may not accurately represent local infiltration conditions. However, the average chloride concentration for pore water in the unsaturated zone is a better indicator of the average infiltration flux.

Bomb-pulse chlorine-36 signatures can be clearly identified in isotopic data. However, the chlorine-36 background analysis is more uncertain because of the possibility for contamination of older chlorine by bomb-pulse chlorine-36. Accordingly, the chlorine-36 method is used to detect clear indications of modern fast pathways and to identify regions of the unsaturated zone with faster transport to the repository horizon (which may also be associated with fast pathways). Bomb-pulse tritium signatures can be attributed to liquid or vapor movement, both of which are prevalent in the unsaturated zone. The presence of bomb-pulse tritium can indicate fast liquid flow from the surface or redistribution of modern infiltration by evaporation and gas-phase movement of water vapor.

Uncertainties Associated with Seepage—Seepage threshold predictions are expected to vary with location. The seepage threshold value of 200 mm/yr (7.8 in./yr) applies to Niche 2 (CRWMS M&O 2000c, Section 3.9.4.7). Further abstraction analysis is used to extend the seepage calibration model to the repository area (CRWMS M&O 2000c, Section 3.9.6.4). Much of the potential repository area is located in the lower lithophysal zone of the Topopah Spring Tuff, which is more permeable than the middle nonlithophysal zone, with shorter fractures and pervasive lithophysal cavities. Some of the frames of Figure 4-18 for Niche 5 in the ECRB Cross-Drift show an example of a large lithophysal cavity and a borehole image of lithophysal cavities.

The greater permeability of the lower lithophysal unit of the Topopah Spring Tuff may enhance the capillary barrier effect (i.e., reduce seepage), either by the presence of more permeable fractures or by higher porosity and permeability in the rock matrix that can absorb more water. Geologic mapping along the ECRB Cross-Drift indicates that the lower lithophysal zone is very heterogeneous; investigations of the hydrologic properties of this rock unit are ongoing.

4.2.1.3.6 Summary of the Current Understanding of Unsaturated Zone Flow and Seepage into Drifts

Current understanding of unsaturated flow at Yucca Mountain has been gained through collection of site data and modeling of the relevant processes.
Table 4-12 summarizes the current understanding of flow parameters and processes. The table identifies features, events, and processes that are important to unsaturated zone flow processes and could affect the waste isolation performance of a repository. The statements listed under the "Current Understanding" column are a mixture of observations, hypotheses, and insights that constitute only abridged, summary information. Additional detail is provided in the Unsaturated Zone Flow and Transport Model Process Model Report (CRWMS M&O 2000c). Note that Table 4-12 summarizes current understanding only for flow processes occurring in the unsaturated zone hydrogeologic units and for seepage. A discussion of issues related to flow and transport processes occurring below the potential repository is presented in Section 4.2.8.

4.2.1.4 Total System Performance Assessment Abstraction

4.2.1.4.1 Unsaturated Zone Flow Abstractions for Total System Performance Assessment

A total of nine flow fields are used in the TSPA base case calculations. These consist of three infiltration cases (lower, mean, and upper) within each of the three climate states (present-day, monsoon, and glacial-transition).

Abstraction of Water Table Rise—The two future climate states (monsoon and glacial-transition) are expected to be wetter than the present-day climate, and, as a result, the water table is expected to rise. However, uncertainty exists regarding the amount of water table rise for each climate state. Therefore, as discussed in
Section 4.3.3.1 and in Abstraction of Flow Fields for RIP (ID: U0125) (CRWMS M&O 2000ca, Section 6.2), a conservative water table rise of 120 m (390 ft) is used for all flow fields using future climate states. Recent analyses described in Section 4.3.3.1.3 indicate that the maximum rise in the last 2 million years has been about 17 to 30 m (56 to 98 ft).

The impact of the water table rise on transport beneath the potential repository was evaluated in Analysis of Base-Case Particle Tracking Results of the Base-Case Flow Fields (ID: U0160) (CRWMS M&O 2000ca, Section 6.2.4). Results showed that the elevated water table reduces transport times beneath the repository (the median breakthrough for a sorbing tracker, neptunium, decreased by nearly a thousand years for the mean infiltration case).

Abstraction of Groundwater Breakthrough—The breakthrough time to the water table of a nonsorbing tracer (technetium) released uniformly in the repository region is simulated for TSPA to gain insight into the range of possible radionuclide transport times that can result based on the different possible infiltration cases for the present-day climate. Breakthrough for future climates is presented in Analysis of Base-Case Particle Tracking Results of the Base-Case Flow Fields (ID: U0160) (CRWMS M&O 2000cb, Section 6.2.6).

Breakthrough curves for technetium using the present-day climate and three infiltration cases show that median breakthrough times are approximately 400 years, 2,000 years, and 600,000 years for the upper, mean, and lower infiltration cases, respectively. As expected, the higher infiltration rates yield shorter breakthrough times relative to lower infiltration rates.

More recent modeling of unsaturated zone transport is presented in Volume 1, Section 11 of FY01 Supplemental Science and Performance Analyses (BSC 2001a). Influence of model refinements of five issues in the travel of radionuclides between the potential repository horizon and the saturated zone is treated (BSC 2001a, Table 11-1). These issues are the degree of advection-diffusion splitting in the drift shadow zone; the effects of the drift shadow concentration boundary on engineered barrier system release rates; the effect of matrix diffusion; the significance of three-dimensional transport modeling; and the effects of coupling of thermal-hydrologic, thermal-hydrologic-chemical and thermal-hydrologic-mechanical processes on transport. Results of model refinement of breakthrough time are presented in Volume 1, Section 11 of FY01 Supplemental Science and Performance Analyses (BSC 2001a, Figures 11.3.1-7, 11.3.1-8, and 11.3.2-8).

4.2.1.4.2 Seepage Model for Performance Assessment

Abstraction of seepage models for the TSPA, as documented in Abstraction of Drift Seepage (
CRWMS M&O 2000by), focuses on providing conservative seepage estimates for a wide range of hydrologic conditions.

Selection of Parameter Ranges and Case StudiesTable 4-13 shows four parameters identified for prediction of seepage into the potential repository drifts. Ranges of values are shown for each parameter, which were used as the basis for an extensive sensitivity analysis. The maximum and minimum values for each parameter were selected based on field data and modeling studies. For example, seepage is evaluated for percolation flux as small as 5 mm/yr (0.2 in./yr) and as great as 500 mm/yr (20 in./yr) (CRWMS M&O 2000bx, Section 6.3.6). The higher value accounts for a hypothetical future climate scenario with spatial and temporal focusing effects (CRWMS M&O 2000c, Sections 3.9.3.1, 3.9.3.2, and 3.9.6.4). The rationale for selecting the parameter ranges shown in Table 4-13 is further discussed in supporting documentation (CRWMS M&O 2000bx, Section 6.3). The results of this sensitivity analysis are used in seepage abstraction for the TSPA (CRWMS M&O 2000c, Section 3.9.6) to account for uncertainty in the seepage calibration model.

Results from Modeling of Seepage for Performance Assessment—The seepage percentage is defined as the seepage flux into a drift opening divided by the average percolation flux over the drift footprint (CRWMS M&O 2000c, Section 3.9.1). The seepage model was implemented to evaluate seepage percentage for multiple statistical realizations of the hydrologic property field representing fractured host rock and for many combinations of parameters in Table 4-13 (CRWMS M&O 2000c, Section 3.9.5.3). The results confirm the seepage behavior observed in testing: seepage increases with decreasing permeability, decreasing capillarity, and increasing percolation flux. For most of the realizations examined, the capillary barrier effect resulted in seepage flux that was substantially less than the percolation flux (i.e., seepage percentage much less than 100 percent). Zero seepage was obtained for a significant portion of the realizations calculated.

Abstraction of the Seepage Model for Performance Assessment—The seepage model for performance assessment was used to simulate seepage for a large number of realizations (CRWMS M&O 2000c, Section 3.9.5). Examination of the results revealed that seepage percentage is most sensitive to a combination of parameters (k/small greek alpha symbol): the product of fracture permeability, k, and the capillary strength parameter, 1/small greek alpha symbol. With this simplification, seepage can be treated as a function of just two variables (i.e., percolation flux and k/small greek alpha symbol).

The abstraction for TSPA focuses on two quantities: (1) the seepage fraction, which is the fraction of waste package locations (i.e., model realizations) for which seepage is predicted and (2) the seepage flow rate, which is the volumetric flow rate of seepage in a drift segment of specified length. Details of the abstraction procedure are provided in supporting documentation (CRWMS M&O 2000by, Sections 6.2.2 and 6.4). Table 4-14 summarizes the abstracted seepage distributions as they vary with percolation flux for ambient conditions (not the nearly dry conditions expected during repository heating). Seepage threshold values of approximately 200 mm/yr (7.8 in./yr), 15 mm/yr (0.6 in./yr), and 5 mm/yr (0.2 in./yr) are estimated for the minimum, expected (i.e., most likely), and maximum seepage conditions, respectively. Note that these values are different from the previously discussed seepage threshold of 200 mm/yr (7.8 in./yr) (CRWMS M&O 2000c, Section 3.9.4.7) for a single location in Niche 2 in the middle nonlithophysal zone.

Summary and Conclusions for the Drift Seepage Model—Seepage into waste emplacement drifts is important to the performance of a repository at Yucca Mountain. Numerical modeling, field testing, and observations at analogue sites suggest that seepage into repository emplacement drift openings would be substantially less than the local percolation flux. This performance results mainly from capillarity retaining the water in the rock and diverting the flow around the openings. The effectiveness of this capillary barrier principle depends on the percolation flux magnitude, the hydrologic properties of the rock, and the drift opening geometry.

A sequence of models was developed to predict the seepage percentage, seepage threshold, and seepage flow rate for waste emplacement drifts. The seepage model was calibrated against relevant data from liquid injection tests in the Exploratory Studies Facility. Seepage percentages and flow rates were then calculated for a wide range of parameter values representing uncertainty in the model and summarized in a probabilistic abstraction model for TSPA. The results indicate that only 13 percent of waste packages are likely to be subject to seepage (CRWMS M&O 2000a, Section 4.1.2). Alternative conceptual models leading to as much as 48 percent of waste packages subject to seepage have been considered (BSC 2001b, Section 4.2.2), although such a high percentage of impacted waste packages is supported only by an inference from an uncalibrated model. The qualitative and quantitative results from seepage testing and modeling, as reflected in the abstraction model, are summarized in Figure 4-37 (CRWMS M&O 2000c, Section 3.9.3).

4.2.2 Effects of Decay Heat on Water Movement

After permanent closure, the heat produced by radioactive decay of the nuclear waste will have an immediate effect on seepage into the repository drifts, water movement through the repository, and the patterns of natural water flow in the unsaturated rock layers. The nature and extent of these effects, however, will depend on thermal loading (or areal heat output), ventilation rates and durations, and attendant thermal operating conditions (i.e., above-boiling or below-boiling). The analytical and experimental studies conducted to date have examined these heat effects in detail but with emphasis on environmental conditions associated with the higher-temperature operating mode described in
Section 2.1.2.3 of this report. The data and analytical results presented in this section mainly describe the effects of higher-temperature conditions on water movement and, specifically, the process models and abstractions employed in the TSPA-SR model, as reported in Total System Performance Assessment for the Site Recommendation (CRWMS M&O 2000a).

As noted in Section 4.1.4, the DOE is evaluating operating the repository at lower temperatures, which may reduce the magnitude and duration of the effects of decay heat on water movement described in this section. Alternative thermal operating modes and supplemental uncertainty evaluation results related to thermal hydrology and thermally coupled models are documented or summarized in FY01 Supplemental Science and Performance Analyses (BSC 2001a, Sections 4.3.5, 4.3.6., 4.3.7, 5.3, 5.4, and 6.3; BSC 2001b, Sections 3.2, 4.2.2, 4.2.3, and 4.2.4).

This section explains the scientific understanding of how the decay heat from radioactive waste will affect natural water movement into and through the repository and water flow in the surrounding unsaturated rock layers. During the period in which decay heat strongly influences fluid flow, the potential sources of water seeping into emplacement drifts are heat-driven condensate flow (thermal seepage), as opposed to ambient percolation and drift seepage, as discussed in Section 4.2.1. Because of the thermal inertia of the heated rock and the continuing (though declining) heat source, the return to near-ambient temperatures may take many thousands of years (CRWMS M&O 2000al). The goals of the near-field thermal hydrology and thermally coupled process models are to assess the effects of the initial thermal pulse (and longer thermal period) on key environmental conditions, such as temperature and relative humidity in the emplacement drifts. These conditions, in turn, may affect the performance of the engineered barriers and the transport of radionuclides (CRWMS M&O 2000al; CRWMS M&O 2000as).

The abstraction of thermal-hydrologic data for use in TSPA represents the potential variability and uncertainty in thermal-hydrologic conditions. It provides a quantitative description of thermal-hydrologic variability (i.e., from variability in the host rock unit, edge proximity, waste package type, infiltration rate, and climate state) and also incorporates uncertainty associated with the infiltration (i.e, lower, mean, and upper). Multiscale model results that are used directly in the TSPA include waste package temperature, relative humidity at the waste package surface, and the percolation flux in the host rock 5 m (16 ft) above the emplacement drift. Temperature and relative humidity are used for the corrosion model, and percolation flux is used for the seepage model. Time-histories of waste package temperature, percolation flux, evaporation rates, and maximum and minimum waste package surface temperatures are also provided (CRWMS M&O 2000cc, Section 6.3).

4.2.2.1 Conceptual Basis

Decay heat generated by radioactive waste may affect the movement of water in the host rock units (CRWMS M&O 2000al, Section 1.1). The conceptual processes described in this section assume thermal loading high enough to result in conditions above the boiling point of water. Under these conditions, decay heat directly affects thermal-hydrologic processes (i.e., movement of water). Heat-driven thermal-chemical processes and thermal-mechanical processes may also affect the movement of water. The conceptual basis for each of these processes is discussed in this section.

4.2.2.1.1 Conceptual Basis of Thermal-Hydrologic Process

Evaporation will occur in the drifts and in the rock immediately surrounding the drift openings. A region of elevated temperature and rock dryout will form around each drift (
Pruess, Wang et al. 1990, p. 1241). Heating can change the flow properties of the rock, and the chemical composition of water and minerals in the affected region. These changes can also occur within the drifts in the engineered barrier system. Figures 4-38 and 4-39 illustrate the conceptual processes of heat-driven water movement.

Calculations indicate that the heat generation rate from radioactive decay decreases rapidly with age relative to the initial output. Heat generation continues at decreased output for thousands of years. Both the initial heat output and its rate of decrease with time depend upon the type of nuclear waste. Calculations indicate that the repository would initially produce approximately 80 MW of thermal power (CRWMS M&O 2000cd, Attachment II). The thermal power output will decrease to approximately 25 percent in 100 years, 12 percent in 300 years, and 2 percent of its initial value in 10,000 years (CRWMS M&O 2000ce, Table I-1).

Continuous forced ventilation of emplacement drifts during operations will remove 70 percent of the total decay heat generated during a period of 50 years after the first waste is emplaced. A conservative model, accounting only for heat removed as sensible heat in the ventilation air (and ignoring heat removed by evaporation of moisture), indicates that a ventilation air flow rate of up to 15 m3/s in each emplacement drift will provide this level of heat removal (CRWMS M&O 2000cd, Section 6.5). The remaining heat output during the preclosure period will be transferred to the host rock by radiation, conduction, and convection, increasing the host rock temperature. Some moisture, and the associated latent heat of evaporation, will be removed by ventilation during this period. At closure, ventilation will cease after drip shields have been placed over the waste packages. This would cause an abrupt increase of heat flux into the host rock.

The major effects of decay heat on water movement would occur after closure (CRWMS M&O 2000cf, Section 6.11.4). Initially, heat will be transported radially away from the drifts by heat conduction through the rock and movement of air through fractures. A portion of the heat will be transported by water that vaporizes near the heat sources and condenses in cooler rock farther away. If the heat flux is high enough, rock near the drifts will be heated to the boiling point of water (nominally 96°C [205°F] at the elevation of the repository) and then to higher temperatures after most of the water in this region has evaporated. The region within which substantially all the water has evaporated is called the dryout zone. Surrounding the dryout zone, a heat pipe zone would form, within which the temperature is essentially constant at the boiling point. The heat pipe zone in turn would be surrounded by a condensation zone of increased water saturation and temperature below the boiling point. After closure, these regions will first expand, then contract as the heat output diminishes over time. The timing of these events will depend on local thermal loading, percolation flux, and location in the potential repository layout (i.e., near the center or the edge). After sufficient time has passed, the temperature will return to preemplacement levels. For purposes of this section, "thermal pulse" is used to describe the development of above-boiling conditions and the heat pipe zone, a process which may last on the order of a few hundreds of years (CRWMS M&O 2000cf, Figure 6-53). "Thermal period" is generally used to describe the time required for temperatures to return to ambient and may last on the order of tens of thousands of years (CRWMS M&O 2000al). Figures 4-38 and 4-39 provide, respectively, conceptual drift-scale and repository-scale illustrations of thermally driven features and processes during the thermal pulse (CRWMS M&O 2000al, Section 3.2.1).

In the investigation of thermal-hydrologic processes, it has been assumed that heating, cooling, and the resulting movement of water will occur in a system with fixed thermal and hydrologic properties (such as porosity, permeability, and thermal conductivity). Properties of the rock may vary with temperature and water saturation but are assumed to return to preemplacement values after the temperature returns to ambient levels (CRWMS M&O 2000al, Sections 3.2.2 and 3.3.5). This concept is used to develop process models like those described in Section 4.2.2.3. Thermal-hydrologic processes in the near field will determine the environmental conditions in the drift, including temperature, relative humidity, and seepage at the drift wall.

4.2.2.1.2 Conceptual Basis of Thermal-Hydrologic-Chemical Process

Thermal-hydrologic-chemical processes involve liquid and vapor flow, heat transport, and thermal effects resulting from boiling and condensation; transport of aqueous and gaseous chemical species; mineralogical characteristics and changes; and aqueous and gaseous chemical reactions.
Figure 4-40 shows schematically the relationships between thermal-hydrologic and geochemical processes in the zones of boiling, condensation, and water drainage in the rock surrounding a repository, particularly in the rock above emplacement drifts.

Heat transfer to the drift wall and surrounding rock will evaporate water in fractures first, then the rock matrix. Vapor will migrate out of the matrix blocks and into fractures, where it will move away from the emplacement drifts because of pressure effects and buoyant convection (CRWMS M&O 2000al, Sections 3.1.5 and 3.2.1). In cooler regions further away from the emplacement drifts, vapor will condense on fracture walls. The condensate will then drain through the fracture network; some of this water will drain back toward the heat source. The resulting localized counter-flow of water vapor and liquid (thermal reflux) is called a heat pipe. A heat pipe zone will develop between the dryout zone and the condensation zone (Pruess, Wang et al. 1990), as discussed previously in Section 4.2.2.1.1.

Chemical evolution of waters, gases, and minerals is coupled to thermal-hydrologic processes. The distribution of condensate in the fracture system will determine where mineral dissolution and precipitation can occur in the fractures and where there can be direct interaction (via diffusion) between matrix pore waters and fracture waters. Investigation of the reactive-transport processes in the potential repository host rock accounts for different rates of transport in the very permeable fractures, compared to the less permeable rock matrix (Steefel and Lichtner 1998, pp. 186 to 187).

One important aspect of the system is release of carbon dioxide from the liquid phase as temperature increases. The release of carbon dioxide and its transport out of the boiling zone will cause pH to increase in the boiling zone and decrease in the condensation zone. Because gases are more mobile than liquids, the region of gas-phase carbon dioxide transport could be much larger than the region affected by thermally driven water movement (CRWMS M&O 2000al, Section 3.3.1.2).

Conservative species (i.e., those that are unreactive and nonvolatile), such as chloride, will become increasingly concentrated in waters undergoing evaporation or boiling but will be more dilute in the condensate zone. The concentrations of chloride and other constituents in condensate waters will be determined mainly by interaction of fracture waters with matrix pore waters via diffusion. Concentrations of aqueous species, such as calcium, will also be affected by mineral dissolution or precipitation and by reactions involving zeolites, clays, and plagioclase feldspar. Calcite may precipitate in fractures over a broad zone of elevated temperature. Silica precipitation will be confined to a narrower zone where evaporative concentration from boiling causes the silica concentration to exceed solubility limits. Alteration of feldspars to clays and zeolites will be most rapid in the boiling zone. As waters drain away from the emplacement drifts, mineral dissolution and precipitation may occur in fractures or in the adjacent rock matrix (CRWMS M&O 2000al, Section 3.3.1.2).

The composition of the percolating waters above the potential repository (before mixing with condensate) may be similar to matrix pore water, or it may reflect more dilute water that has traveled through fractures (CRWMS M&O 2000al, Section 3.3.1.3). The chemical composition selected for input to the thermal-hydrologic-chemical model is described in Sections 4.2.2.2.1.3, 4.2.2.3.3, and 4.2.3.3.1 of this report.

Changes in the percolation flux can affect the extent of mineral deposition and dissolution because of changes in the fluxes of dissolved species. For example, with more calcium transported toward the emplacement drifts, more calcite would tend to be precipitated. Also, a greater percolation flux will tend to increase the dissolution of minerals that are undersaturated in the fluid (CRWMS M&O 2000al, Section 3.3.1.3).

Mineral dissolution and precipitation in fractures and in the rock matrix can modify the porosity, permeability, and unsaturated hydrologic properties of the potential repository host rock in the vicinity of the emplacement drifts. The extent of mineral-water reactions will be controlled by the surface area of each mineral phase that is exposed to the liquid water. Other factors that may control property changes are the distribution of liquid saturation in fractures, the proportion of fractures with actively flowing water, and the rate of evaporation or boiling, which can control crystal growth and nucleation (CRWMS M&O 2000al, Section 3.3.1.4). Rock–water interactions will affect the chemistry of the water that may seep into the emplacement drifts. The effect of rock–water interactions on hydrologic properties and seepage is evaluated in Section 4.2.2.3.3.

4.2.2.1.3 Conceptual Basis of Thermal-Mechanical Process

The stress field in the rock mass surrounding emplacement drifts would be altered by excavation of drifts and by the heating/cooling cycle associated with emplacement of radioactive waste. The direction and magnitude of principal stresses will change significantly because of thermal loading and then will return to near-ambient values during cooldown—but not completely, since the rock mass will be changed permanently from deformations occurring from stress redistribution; however, the magnitude of changes in hydrologic properties will be limited (
CRWMS M&O 2000al, Section 3.5). Compressive stress will build up rapidly in the host rock, especially after the end of the ventilation period. The stress field generally will gradually decay as the temperature in the rock decreases. Potential seismic effects on the repository system are discussed in Drift Degradation Analysis (CRWMS M&O 2000e, Sections 6.3.4 and 7.1); potential effects of fault displacement on the repository system are discussed in Disruptive Events Process Model Report (CRWMS M&O 2000f).

The potential repository host rock is a fractured, densely welded, ash-flow tuff. These fractures are expected to deform as stress conditions evolve. Two types of fracture deformations will contribute to thermal-hydrologic-mechanical coupling: normal displacement perpendicular to a fracture plane and shear displacement parallel to a fracture plane.

Rock-mass permeability is an important thermal-hydrologic property for assessment of repository performance. Because the rock has low matrix permeability, the rock mass permeability is mainly associated with fractures, and large changes in rock mass permeability may result from fracture deformation.

The potential effect of fracture deformation on fracture permeability is discussed in Section 4.2.2.3.4. The potential for drift degradation to affect the drip shield (i.e., rockfall on the drip shield) is discussed in Section 4.2.3. The basis for screening these potential processes from the TSPA-SR model is referenced in the process and model and/or the TSPA abstraction sections.

4.2.2.2 Summary State of Knowledge

This section presents a summary of the state of knowledge of properties (rock and fluid) and processes tested in the laboratory and in the field. Much of the information presented in this section is extracted from Hardin and Chesnut (1997).

4.2.2.2.1 State of Knowledge of Laboratory Measured Properties

This section describes the available laboratory data for assessing matrix contributions to thermally coupled processes. Laboratory data for assessing fracture contributions are relatively limited and are described to a lesser extent. Because sample sizes are typically much smaller than the in situ fracture spacing, laboratory measurements generally provide properties of the rock matrix only and do not directly show the effects of thermally coupled processes on rock-mass behavior. Laboratory data therefore provide only part of the input data required to investigate coupled processes. The properties of fracture networks are also needed and have been inferred from observations and measurements in the field.

4.2.2.2.1.1 Thermal-Mechanical Properties

Variation of matrix thermal and thermal-mechanical properties with temperature is relatively well understood, and data are available. The dependence of matrix and fracture rheology on temperature, including deformation modulus and creep properties, is less well known, but some data at elevated temperature are available. These data do not indicate that rheology is important for prediction of long-term repository performance.

Rock Creep—Laboratory testing of the rock matrix showed that significant creep occurs only when samples are stressed to at least 50 percent, and in some cases more than 90 percent, of their ultimate strength (
Martin et al. 1995). Such stress conditions may be uncommon in the host rock, occurring only at fracture asperity contacts. These results also indicate a tendency for fractures in the host rock to close in response to heating and open in response to cooling.

Physical Properties (Porosity and Grain Density)—For the TSw2 welded tuff, these properties change little from ambient temperature up to at least 180°C (356°F). Change in the welded tuff matrix near the emplacement drifts is expected to be minor and will probably be caused mainly by mineral phase transitions (e.g., asymbol for right pointing arrowb cristobalite) and dehydration of hydrous phases, such as clinoptilolite and smectites. Of these, dehydration of hydrous fracture-lining minerals has a greater potential to affect host rock performance (Hardin and Chesnut 1997, Section 2.2).

Thermal Conductivity—Laboratory measurements of thermal conductivity have been performed on samples from the Exploratory Studies Facility, in conjunction with field-scale thermal testing (Brodsky et al. 1997). There is a slight increase of thermal conductivity with temperature for the Topopah Spring welded tuff. Water saturation apparently increases thermal conductivity of the TSw2 welded tuff by approximately 50 percent. A small pressure effect in other rock types, whereby conductivity increases with confining stress, has been observed and is probably caused by closing of microcracks. Pressure effects on thermal conductivity have not been examined for Yucca Mountain tuffs but are likely to be small or highly localized in the host rock.

Heat Capacitance—When measured on dried samples, the heat capacitance of the TSw2 welded tuff increases about 20 percent from ambient temperature to 200°C (392°F) (SNL 1996). Behavior at temperatures greater than 150°C (302°F) is affected by mineral phase transitions, notably that of cristobalite, which occurs at temperatures greater than 200°C (392°F). In polycrystalline rocks, the cristobalite transition apparently occurs over a temperature range of 20 to 50 C° (36 to 90 F°).

Thermal Expansion—The coefficient of thermal expansion for TSw2 welded tuff increases with temperature because of mineral-phase transitions and dilatancy caused by heterogeneous thermal expansion of different minerals. Linear unconfined expansion measurements have been reported for ambient pressure, temperatures to 300°C (572°F), and several saturation states (Brodsky et al. 1997, Table B-5) determined from samples that are somewhat heterogeneous and exhibit some variability between samples. Measured thermal expansion for samples of TSw2 welded tuff varies by a factor of about five. Thermal expansion is relatively insensitive to saturation. Hysteresis becomes apparent at temperatures greater than 200°C (392°F), probably because expansion produces irreversible changes in rock fabric.

Mechanical Properties—For the Topopah Spring welded tuff, long-term (3.5- to 6-month) changes in the mechanical properties of three samples were investigated at temperatures of 80°, 120°, and 180°C (176°, 248°, and 356°F) (Hardin and Chesnut 1997, Section 2.5.1). The results indicate that temperature effects on mechanical properties are smaller than the differences between the samples. More recently, a 0.5-m (1.6-ft) scale block of Topopah Spring welded tuff was subjected to uniaxial loading at temperatures as great as 85°C (185°F) (Hardin and Chesnut 1997, Section 3.8.1). The apparent Young's modulus for the tuff matrix at several locations in the block decreased significantly as temperature increased.

Compressive Strength Versus Saturation—It has been reported that a significant decrease in compressive strength could be associated with increased saturation (Nimick and Schwartz 1987, Section 3.4.2.2.1). This observation was based on early studies that may have been affected by different methods used to control sample saturation (Boyd et al. 1994, Section 4.3).

4.2.2.2.1.2 Hydrologic Properties

Thermal-hydrologic processes in fractured rock have been investigated theoretically and experimentally since the early 1980s (
Pruess, Tsang et al. 1984; Pruess, Wang et al. 1990; Buscheck and Nitao 1993; Pruess 1997; Tsang and Birkholzer 1999; Kneafsey and Pruess 1998). The laboratory work and early field studies are reported in Synthesis Report on Thermally-Driven Coupled Processes (Hardin and Chestnut 1997) and Near-Field Environment Process Model Report (CRWMS M&O 2000al, Sections 2.2 and 3). See also Section 4.2.1.2.5 of this document for hydrologic properties of fractures.

For matrix hydrologic properties, there are fundamental temperature-dependent responses that may be important to understanding the thermal-hydrologic process. These include the temperature effect on hysteresis of wetting and drying characteristic curves, Knudsen diffusion, and enhancement of vapor diffusion (CRWMS M&O 2000al, Section 3.6.3.1).

Matrix Permeability—Variations in matrix permeability of the Topopah Spring welded tuff that are associated with temperature changes have been found to be much less than natural variations between samples (CRWMS M&O 2000al, Section 3.6.3.1).

Unsaturated Hydraulic Conductivity—A limited number of measurements have been made of unsaturated matrix conductivity in Yucca Mountain tuffs. Changes in the properties of water at elevated temperature (viscosity and surface tension) suggest that unsaturated conductivity may increase by as much as an order of magnitude from 20° to 100°C (68° to 212°F). The viscosity effect is taken into account in current thermal-hydrologic simulations, but the surface-tension effect is not. In addition, changes in the water-rock-air contact angle at elevated temperature can also influence unsaturated conductivity (Hardin and Chesnut 1997, Section 2.10).

Enhanced Vapor Diffusion—No enhancement in vapor diffusion was observed in a limited investigation of the Topopah Spring welded tuff matrix (Wildenschild and Roberts 1999).

Knudsen Diffusion—Knudsen diffusion and its variation with temperature are possible mechanisms for transport of moisture in the host rock (Hardin and Chesnut 1997, Section 2.7.4). This is likely to be of little significance and has not been investigated experimentally.

4.2.2.2.1.3 Chemical and Transport Properties

Chemical reactions are strongly temperature-dependent, and laboratory measurements of reaction rates and surface areas under controlled conditions are sparse. However, experimental kinetic data are not generally needed for reactions that can be modeled satisfactorily using qualified thermodynamic equilibrium and reaction-path models, such as EQ3/6 and its associated chemical databases. Determination of which chemical processes in the potential repository can be modeled in this manner rely on results from laboratory and field-scale testing.

Thermodynamic equilibrium data for many aqueous and mineral species have been measured or estimated, and reviewed for accuracy and consistency in preparation for use with qualified analyses. For certain other types of reactions (e.g., surface complexation), equilibrium conditions at elevated temperatures are relatively unknown.

Seepage Water Compositions—Two water compositions have been considered for use in TSPA-SR modeling for various purposes. One is referred to as chloride-sulfate-type water that is based on the chemical analyses of matrix pore waters from near the Drift Scale Test (
CRWMS M&O 2000cg, Sections 6.5 and 6.7.4; BSC 2001o, Section 6.1.2). Another is referred to as bicarbonate-type water, based on the composition of J-13 well water (CRWMS M&O 2000cg, Section 6.5; see also Section 4.2.4). The chloride-sulfate-type water is more concentrated in total dissolved minerals and is selected for calculations that evaluate potential changes in fracture properties from precipitation of minerals and salts. The source of water and gas chemistry for use in the thermal-hydrologic-chemical model is based on the chemical composition of matrix pore water collected from Alcove 5 (BSC 2001o, Sections 4.1.3 and 6.1.2).

Behavior of Radionuclides in J-13 Water—Experimental data on the speciation and solubility of important radionuclides at elevated temperature are limited. However, the investigation of spiked J-13 water at temperatures as great as 100°C (212°F) indicates that plutonium solubility decreases, but uranium, neptunium, and americium remain soluble or become increasingly soluble at elevated temperatures (Nitsche 1991). Carbonate complexes appear to be important to the solubility of uranium, neptunium, and americium at elevated temperatures. Knowledge about complexation and solubility of nickel, zirconium, technetium, uranium, neptunium, plutonium, and americium in J-13 water at elevated temperatures was published in a recent review (Wruck and Palmer 1997).

Hydrothermal Tuff Alteration—Batch studies of hydrothermal alteration of wafers of Topopah Spring welded, devitrified tuff have been performed at temperatures from 90° to 250°C (194° to 482°F) and for durations to 120 days (Knauss 1987; Knauss and Beiriger 1984; Knauss, Beiriger et al. 1987; Knauss, Delany et al. 1985; Knauss and Peifer 1986; Oversby 1984a; Oversby 1984b; Oversby 1985). They show that changes in the composition of water in contact with the tuff are moderate at temperatures as great as 150°C (302°F), with slight alteration of the tuff over a few months. At higher temperatures, similar alteration products are produced, but reaction rates increase significantly. Accelerated experiments on crushed tuff at temperatures greater than 150°C (302°F) have produced more extensive alteration, including metastable phases.

Energetics of Zeolite Dehydration—Zeolites could have a significant effect on the heat and water balance where they are abundant because zeolite dehydration is more energetic than evaporation of water on a molar basis (Bish 1995; Wilder 1996, Section 3.4.3; Hardin and Chesnut 1997, Section 2.6.2.1). Zeolite hydration is apparently reversible at dehydration temperatures as great as 215°C (419°F) for clinoptilolite, so complementary effects will occur during repository cooldown. Altered units above and below the repository horizon contain a large fraction of zeolites; the data produced by these studies indicate that dehydration will cause some amount of shrinkage, increasing porosity and probably also increasing permeability. This could affect the water-perching behavior at altered zones associated with the upper and lower Topopah Spring vitrophyres.

Effect of Hydrothermal Alteration on Flow Paths—Plug-flow reactor studies involving flow-through reaction of J-13 water with crushed tuff at 240°C (464°F) resulted in significant dissolution of alkali feldspar and cristobalite (DeLoach et al. 1997, p. 5). This experiment produced significantly different results from those of batch reactor studies at similar temperatures (i.e., predominantly dissolution instead of alteration). The two approaches span the range of conditions likely to exist in the host rock: stagnant vs. flowing water in the tuff matrix or along fractures. Dissolution and alteration behavior of the major minerals constituting the host rock are temperature-dependent and much slower at temperatures near the boiling point of water.

Limitations of Available Kinetic Data—The available kinetic data for dissolution of mineral phases that may be important to repository performance are limited reflecting the general sparseness of laboratory data on kinetic interactions involving rock. Different investigators have used various investigation and measurement strategies, and test results are sensitive to methodology (e.g., batch methods versus flow-through methods).

Kinetics of Silica Dissolution and Precipitation—Reaction rates for dissolution of quartz and silica polymorphs, and precipitation of amorphous silica, are key parameters in estimating the extent and magnitude of thermal-hydrologic-chemical coupled effects in the host rock. Dissolution will be expressed in heat pipe zones where refluxing water is at approximately 100°C (212°F). Mineral species like silica will then be deposited where the reflux water evaporates or boils. A boiling front will expand outward from each emplacement drift, but eventually reverse because of less heat generation. Depending on how fracture properties and connectivity are affected by precipitated minerals, seepage into the drift openings may become more or less likely during cooldown.

Experimental Data for Silica Kinetics—Dissolution rates are key parameters for estimating the extent and magnitude of thermal-hydrologic-chemical coupled effects in the host rock. In a classic study, the dissolution rate for silica polymorphs increased by 2 orders of magnitude for each 100 C° (180 F°) temperature increase, with a factor of 300 increase in the dissolution rate between 25° and 70°C (77° and 158°F) (Rimstidt and Barnes 1980, Figure 9). In addition, upon cooling a saturated silica solution, decreasing solubility caused supersaturation, while the rate constant for precipitation decreased, producing a maximum precipitation rate at a temperature 25 to 50 C° (45 to 90 F°) less than the saturation temperature. A more recent investigation of quartz dissolution kinetics at 70°C (158°F) (Knauss and Wolery 1988) produced dissolution rate data that were similar, at neutral to mildly alkaline pH, to rates predicted for quartz by the classic model.

Interaction of Radionuclides with Alteration Products of Introduced Materials—Surface complexation reactions will be important for retardation of actinides, and possibly pertechnetate, in the host rock. Introduced materials, including structural steel, could be a source for potential high-affinity sorbents for radionuclides. If the sorbents are colloidal, the sorbed radionuclides may be transported. Limited test data for radionuclide sorption on these materials (e.g., goethite, clays, or silica polymorphs) are available, mainly for ambient temperature and simplified chemical systems (see, for example, Section 6.6 of Engineered Barrier System: Physical and Chemical Environment Model [CRWMS M&O 2000cg]).

Matrix Diffusion Effects—Diffusion of radionuclides into minerals and into the tuff matrix is an important temperature-dependent retardation mechanism. The tuff matrix has been shown to contain ubiquitous nanopores that support slow diffusion, plus a few connected paths through which diffusion is much faster but limited in overall effect. Effective diffusion coefficients have been estimated for uranium migration into polished wafers of Topopah Spring Tuff at ambient temperature (Wilder 1996, Section 7.4.1). Relative diffusivities of actinide and technetium species have been compared at 90°C (194°F) using "tuff cup" experiments (Hardin and Chesnut 1997, Section 2.10). Effective diffusion parameters for migration of strontium and cesium ions in clinoptilolite have been estimated (Hardin and Chesnut 1997, Section 2.6.5.1). These data generally indicate that the rate of diffusion in the tuff matrix and sorbent minerals is enhanced at elevated temperature.

4.2.2.2.1.4 Other Properties

Self-Potential—Naturally occurring electrical potentials were observed in the Single Heater Test and in the Drift Scale Test and were large enough to be considered as a factor in waste package corrosion analyses, but the source of these potentials and the amount of current generated have not been investigated (
Hardin and Chesnut 1997, Section 2.10).

Microbial Activity—Investigations have established that the natural microbes present in the unsaturated zone, plus those introduced by excavation, include species that can survive exposure to desiccation and elevated temperature. Some species produce metabolic products that could be important in determining rates of corrosion and radionuclide transport in the near-field environment. There are few data that can be used to describe microbial activity at elevated temperatures (Hardin and Chesnut 1997, Section 2.10).

4.2.2.2.2 Laboratory-Scale Process Investigations

Laboratory experiments have included comparison of vapor-phase and liquid-phase rewetting, fracture healing, fracture–matrix coupling with flow into heated tuff, fracture flow visualization, heat pipe formation, and rock–water interaction studies. These physical simulations of thermally coupled processes have advanced conceptual understanding and provided data for testing mathematical models.

Rewetting Behavior—Testing of wafers of welded tuff has indicated (
Wilder 1996, Section 2.1.1) that water-retention hysteresis varies at elevated temperatures. Typical wetting/drying hysteresis at ambient temperature was nearly zero at 78°C (172°F) and reversed at 94°C (201°F). Rewetting behavior at elevated temperature is also summarized by Hardin and Chesnut (1997, Section 2.7.5). The effect is probably related to changes in surface tension and the rock–water–air contact angle at elevated temperatures. Hysteresis behavior is generally ignored, for computational expediency, in thermal-hydrologic models, and this appears to be defensible. The possible effects of negative hysteresis have not been considered.

Vapor Resaturation—Tuff matrix rewetting due to the presence of saturated water vapor (100 percent relative humidity) has a different result than rewetting by liquid at the same zero potential. Experimental data (Buscheck et al. 1992) show that rewetting of dry tuff in the presence of water vapor occurs much more slowly than does rewetting by imbibition of liquid water. This effect is incorporated in thermal-hydrologic models by adjusting the matric potential versus saturation relationship so that matrix saturation of 30 to 40 percent or greater corresponds to a relative humidity of nearly 100 percent. The vapor resaturation effect strongly influences the timing of rewetting in the repository (Wilder 1996, Section 10.1) and tends to increase the relative humidity calculated at waste packages during cooldown.

Fracture Healing—An understanding of fracture-permeability reduction has been developed, and observations reported in the literature can be explained by three mechanisms (CRWMS M&O 2000al, Section 3.6.3.3):

  1. Dissolution of fracture asperities by flowing water and consequent aperture reduction under the influence of confining stress

  2. Dissolution/precipitation reactions that clog porosity by redistributing silica or by creating alteration products with greater molar volume

  3. Migration of heated pore water from the rock matrix, toward fractures, where the pressure is lower and evaporation or boiling occurs, clogging fractures or the matrix porosity.

Experiments have shown that flowing water or steam promotes permeability reduction, and the effect is strongest at temperatures greater than 90°C (194°F) (Hardin and Chesnut 1997, Section 3.10) (Figure 4-41). All these mechanisms can lead to changes in fracture porosity and permeability in the host rock where there is sufficient water.

Fracture–Matrix Coupled-Flow Visualization—Fracture flow studies in the laboratory (Hardin and Chesnut 1997, Section 3.3.1) have physically demonstrated fracture–matrix flow coupling in welded tuff, using x-ray imaging to visualize the flow (Figure 4-42). By varying the water injection pressure and the resulting flow velocity, the nonequilibrium nature of flow coupling was demonstrated. When the experiment was repeated with a thermal gradient, a different flow regime resulted, with localized precipitation of the solute tracer (Figure 4-42).

Flow Channelization—Visualization experiments (Hardin and Chesnut 1997, Section 3.4) showed, among other findings, that fracture transport in response to constant boundary conditions can be unsteady and produce intermittent rivulets that "snap off" and reform episodically. The authors related the average repetition rate for episodic flow with fracture aperture and wetting properties and the inclination of the models to gravity. These ambient-temperature experiments demonstrated that simple, simulated fractures can produce unsteady fracture flow in response to constant boundary conditions. Similar flow can be expected in the fractures of a heat pipe zone. These data have important implications for fracture–matrix interaction (i.e., there is limited contact time available for fracture–matrix interaction).

Physical Models of Heat Pipes—Fracture thermal-hydrology visualization studies by Kneafsey and Pruess (1997) examined conditions (e.g., fracture saturation, temperature difference, and fracture dimensions) that support heat pipe development. Heat pipes were observed in parallel plate fractures containing obstacles, heat sources, and vents. Film flow as well as meniscal flow were observed to produce heat pipes (Figure 4-43). Unsteady rivulet-flow behavior, analogous to episodic fracture flow at ambient temperature, was observed (Hardin and Chesnut 1997, Section 3.5.1). Rapid evaporation events occurred when "islands" of fluid became superheated and suddenly boiled, constituting another mechanism for unsteady flow with the potential to rapidly disperse solute. A few of these observations were repeated with a half-cast model incorporating welded tuff as one fracture wall.

Water–Rock Interaction—Chemical analyses of effluent water from matrix flow and fracture flow experiments have indicated water-rock interaction. When J-13 water was flowed through an intact core sample, the concentration of most major anions and cations first increased, then approached influent concentrations (Hardin and Chesnut 1997, Section 3.6.2.2). Anions such as chloride and sulfate were leached in quantities that may be significant to the in-drift chemical environment. Other reported data for similar tests (Hardin and Chesnut 1997, Section 3.6.2.1) included chemical analyses of water that flowed through a healed, natural fracture at elevated temperatures.

Fracture Flow in a Heated and Stressed Block—Laboratory tests were conducted on 0.5-m (1.6-ft) scale blocks to monitor fracture flow and mechanical deformation properties under conditions that approximated the near-field environment expected in a repository at Yucca Mountain (Costantino et al. 1998). In this test, a rectangular block was bisected by an artificial fracture perpendicular to the fabric of the tuff. Water was supplied at a point source at the center of the fracture under various pressures. Both fluid flow and mechanical properties were found to be anisotropic and strongly correlated with the ash-flow fabric of the sample. Fluid flow measurements revealed that only minor imbibition of water occurred through the fracture surfaces, and that flow rates were independent of normal stress up to 14 MPa across the fracture, and at temperatures to 140°C (280°F). Flow through the fracture occurred largely through uncorrelated porosity that intersected the fracture plane.

4.2.2.2.3 Field-Scale Processes

This section briefly describes selected results from a number of field-scale tests and natural processes having some features in common with processes expected at Yucca Mountain. Emphasis is given to thermally driven coupled processes. The field-scale tests include those presented by
Hardin and Chesnut (1997) and those presented in the Near-Field Environment Process Model Report (CRWMS M&O 2000al, Section 3.6.1). These tests are grouped into three groups: non-Yucca Mountain tests, Yucca Mountain tests, and natural analogues.

4.2.2.2.3.1 Non-Yucca Mountain Tests

Climax Spent Fuel Test (
Hardin and Chesnut 1997, Section 4.1.1; Wilder and Yow 1987)—Acoustic emissions responded to the rate of thermal energy production and may be useful for monitoring the stability of a repository. No significant changes in mineralogy or microfracturing occurred, as a result of heat or irradiation, near the electrical heaters or spent nuclear fuel canisters. Nitric acid formed by radiolysis of atmospheric nitrogen accelerated corrosion of the carbon steel emplacement hole liners. Corrosion was also observed in alloys such as stainless steel, Inconel 600, and super-Invar.

Edgar Mine, Colorado School of Mines (Hardin et al. 1982)—Heating of a fractured gneiss caused significant reductions in the loading and unloading moduli and reductions in the permeability of a test fracture. The largest permeability change occurred during excavation. Compressive loading reduced the permeability, but the permeability did not return to the preexcavation condition. Permeability reduction at elevated temperatures was smaller in magnitude than the effect of excavation.

G-Tunnel Small-Diameter Heater Tests (Zimmerman and Finley 1987)—For the first month of heating in the horizontal heater test in welded tuff, small amounts of water collected in the heater borehole and wetted a sensor located immediately under the heater. Relative humidity approached saturation within hours after the start. Total pressures remained at ambient. Neutron-probe measurements of moisture content in the heated rock showed that significant changes occurred in the temperature range 70° to 120°C (158° to 248°F). Dewatering apparently began at temperatures less than boiling.

G-Tunnel Heated Block Test (Zimmerman et al. 1986)—A slight dependence of modulus on stress was indicated, but there was no significant temperature effect. Thermal expansion behavior of the heated block was well represented by measurements on intact rock samples. The largest changes in permeability of a test fracture were associated with excavation. Subsequent compressive loading and increased temperature lowered the apparent permeability of a test fracture. Saturation declined, in steps corresponding to successive cycles, from 60 to 80 percent and down to approximately 15 percent as a result of heating. Rehydration upon cooling was not significant on a time scale of weeks.

G-Tunnel Prototype Engineered Barrier System Field Test (Hardin and Chesnut 1997, Section 4.1.3; Ramirez et al. 1991; Lee and Ueng 1991)—The drying front penetrated most rapidly along fractures, and rewetting occurred most rapidly near fractures during the ramping down and cooling phases. Water vapor that condensed below the heater drained away from the boiling zone, and rock below the heater dried out more quickly than it did above the heater. During cooling, rewetting above the heater occurred slightly more quickly than it did below the heater. After heating and cooling back to ambient temperature, measured permeability in the heater borehole increased by 10 to 1,800 percent. The increase was greatest in intervals with the smaller values of preheating permeability. The boiling zone acted as an "umbrella," shielding rock below the heater from drainage of condensate generated above the heater.

Underground Tests at Stripa, Sweden—Fracture closure in response to heating was confirmed by observation of diminished water inflow to heater and instrument boreholes (Nelson et al. 1981).

4.2.2.2.3.2 Yucca Mountain Tests

Large Block Test—The Large Block Test was described by
Wilder et al. (1997). Figure 4-44 shows the Large Block Test during its construction, and a schematic of the test showing the instrument boreholes is given in Figure 4-45. One-dimensional heating geometry and moisture movement in the block were achieved as planned. Boiling of the pore water was indicated by temperatures measured near the heaters. Figure 4-46 shows the temperature measured at the TT1-14 and TT2-14 temperature sensors. Note that the boiling temperatures apparent from these two figures are slightly different, indicating heterogeneity in the pore pressure and/or concentration of chemicals in the pore water. Heat pipe activity was observed along the two vertical temperature holes in the block. Figure 4-47 illustrates one aspect of that heat pipe activity, whereby condensate drainage toward the heaters was most evident during the two thermal refluxing episodes of the Large Block Test. Figure 4-48 illustrates one of those refluxing episodes and the episodic water movement that followed its onset. Cooler liquid water apparently penetrated the heated interval, causing the temperature in a wider zone extending above and below the heaters to converge to near the boiling point. This was followed by episodic thermal refluxing, which caused the temperature to fluctuate. The redistribution of moisture was monitored by electrical resistivity tomography and neutron logging, as shown in Figure 4-49 and Figure 4-50, respectively. These refluxing events are believed to have been caused by rainstorms, which introduced water into the test.

Drying of the Large Block Test was also nearly one-dimensional. Neutron logging shows that localized dryout reached its maximum after about 334 days of heating. Subsequent heating extended the dryout zone but did not dry out the rock further. Mechanical displacement measurements on the block indicate that during the June 1997 thermal refluxing episode, a major near-horizontal fracture near the top of the block opened approximately 0.0094 to 0.011 cm (0.0037 to 0.0043 in.) at the northern and eastern sides, with a 0.0058 cm (0.0023 in.) shear displacement on the western side. Deformation data also indicate that during heating the block experienced horizontal expansion that increased linearly with height above the base. These deformations may have affected the hydrologic properties of the block. Simulation of the Large Block Test using the thermal-hydrology modeling code NUFT was used to capture the major characteristics of the measured temperatures of the Large Block Test (CRWMS M&O 2000al, Section 3.6.1.1) (see additional discussion below).

Single Heater Test (CRWMS M&O 2000al, Section 3.6.1.2)—The Single Heater Test was described in Single Heater Test Final Report (CRWMS M&O 1999m). A schematic of the Single Heater Test is shown in Figure 4-51. Measurement of mechanical displacements in the Single Heater Test showed expansion along the heater hole, with compressive movement at the beginning of the heating, followed by expansion perpendicular to the heater. These results were simulated by calculations using the continuum mechanical modeling code FLAC. Simulation using the thermal-hydrology modeling code TOUGH2 was used to represent the evolution of measured temperatures. Modeling of this test provided important insights into the hydrologic properties and responses of the fractured tuff. The test confirmed that water moves away from heat sources as vapor, condenses where it is cooler, and tends to drain downward. Condensate was collected in a borehole, where it intersected a fracture drainage pathway. The apparent rapidity of drainage through fractures indicates that rock–water interaction between condensate and fracture surfaces is limited to relatively short residence times. The chemistry of sampled water was similar to, but more dilute than, J-13 water. Solution equilibrium modeling indicates that the gas-phase carbon dioxide fugacity during the heated portion of the test was about two orders of magnitude greater than atmospheric.

Drift Scale Test—The Drift Scale Test, the largest thermal test conducted by the project to date, involves heating a drift that is approximately 48 m (157 ft) long. The rock mass being heated in this test will experience a thermal environment somewhat hotter than the repository design, with rock surface temperatures reaching a maximum of about 200°C (390°F). This test is designed to produce above-boiling temperatures in about 10,000 m3 of rock. The Drift Scale Test was described in Near Field Environment Process Model Report (CRWMS M&O 2000al, Section 3.6.1.3). A schematic of the Drift Scale Test is shown in Figure 4-52. After 4 years of heating, the Drift Scale Test is to begin a 4-year cooling phase beginning in early 2002 (CRWMS M&O 2000ch Section 1; CRWMS M&O 2000al, Section 3.6.1.3). The test results have provided and will continue to provide important insights into analyses of coupled processes, as described in this and following sections. Figure 4-53 shows the effect of boiling, as indicated by the temperature history at a single location. Boiling of pore water is represented by a relatively flat region at about the boiling point of pore water in the temperature-time curves. Variations in heating rates are due to different distances between the sensors and the heaters. The duration of boiling is apparently affected by the availability of pore water to evaporate and the extent to which condensate flows back toward the heaters at each location.

Drying in regions near the heated drift and the wing heaters has been measured by electrical resistivity tomography, neutron logging, and ground-penetrating radar. Results obtained using these methods are shown in Figures 4-54, 4-55, and 4-56, respectively.

Heat pipe activity caused by thermal refluxing of condensate water has been observed at some temperature measurement locations, as shown in Figure 4-57.

Thermal Tests Thermal-Hydrological Analyses/Model Report (CRWMS M&O 2000ch) reports that both the TOUGH2 and NUFT modeling codes have predicted the temperature evolution and the dryout (liquid saturation) for the rock around the heated drift, consistently and with reasonable success. These thermal-hydrologic simulations support high confidence in using the thermal conductivity of rock mass to predict average temperature distributions. In addition, Drift-Scale Coupled Processes (DST and THC Seepage) Models (BSC 2001o, Section 6.2.7) shows that the TOUGHREACT modeling code was able to produce reasonable predictions of the evolution of carbon dioxide and composition of collected water samples as affected by heating.

4.2.2.2.3.3 Natural Analogues

Natural analogues related to the effects of heating on water movement and rock mass response were discussed by
Hardin and Chesnut (1997) and in Near Field Environment Process Model Report (CRWMS M&O 2000al, Section 3.6.2). Additional information on natural analogues is provided in Natural Analogs for the Unsaturated Zone (CRWMS M&O 2000bp) and in Yucca Mountain Site Description (CRWMS M&O 2000b, Section 13). Figure 4-58 illustrates selected examples of analogue information for heat induced and coupled processes. The information from geothermal studies can be used to demonstrate that the effects of decay heat on water movement can be evaluated. The available information ranges from core-scale and outcrop evaluations of fossil thermal-hydrologic-chemical coupled processes, understanding of efficient heat transfer mechanism with vapor-liquid counter flow phenomena (heat pipe), induced changes in operating geothermal fields, to dynamic or even disruptive processes with mechanical changes. Some of these phenomena are described in this section.

Geothermal Fields—Heat pipes in the repository will be short-lived, transient features compared to geothermal systems. Thermal-hydrologic characteristics of the repository host rock are similar to vapor-dominated geothermal systems (Hardin and Chesnut 1997; Simmons and Bodvarsson 1997), and similar modeling methods are applicable. Geothermal systems are apparently self-sealing, but conditions for development of a mineral cap above the emplacement drifts in a repository are different and less likely to result in sealing.

Table 4-15 lists selected geothermal systems evaluated in many countries. A photograph of a tuff core with indication of fracture sealing and opening from the Yellowstone geothermal field and a photograph of topographic deformations associated with the high-temperature geothermal field at Krafla, Iceland, are examples illustrated in Figure 4-58. Also included in Figure 4-58 is a photograph of the concrete-lined tunnel, subjected to heating by nuclear reactor processes, at the Kransnoyarsk-26 site in Russia. This is one example of the different kinds of site studies of anthropogenic (human-induced) analogues (CRWMS M&O 2000b, Section 13.3.6.3). Additional information on natural analogues and model calibration and validation for thermal-hydrologic applications is provided below in Section 4.2.2.3.6.

The Geysers—The Geysers is a geothermal field located about 65 km (45 miles) north of Santa Rosa, California. The pressure and temperature of the vapor-dominated geothermal system at The Geysers plot along the liquid-vapor phase boundary for water, whereas the repository host rock will be free-draining, and the pressure of refluxing water will be constrained at about 1 atm (Hardin and Chesnut 1997, Section 4.3.1; Wilder 1996, Chapter 1). The maximum heat flux from the repository will be 50 to 70 times that calculated for The Geysers. Therefore, the calculated reflux magnitude at The Geysers, 4 to 5 mm/yr, does not imply that there will be no dryout zone above the repository. However, heat pipe activity is likely in the repository host rock and may be important for an extended period during cooldown.

Taupo Volcanic Zone and Geothermal Fields, New Zealand—The Taupo volcanic zone analogue was used to compare observed water composition and mineral occurrence in a geothermal well with chemical modeling calculations. In accordance with observations, major phases such as quartz and calcite were calculated to be at near-equilibrium. Other phases were predicted less accurately for several reasons: (1) uncertain kinetics, (2) completeness of thermodynamic data, (3) solid-solution effects (such as heterogeneity), and (4) the influence of boiling on precipitation (Glassley and Christenson 1992; Bruton et al. 1994). Differences were observed between precipitation rates measured in the laboratory and those measured in the field. Earlier studies by Rimstidt and Barnes (1980, p. 1691) argued that precipitation rates for silica polymorphs are the same, but this approach does not explain certain natural analogue observations (Carroll et al. 1995, pp. 7 to 13).

Figure 4-58 includes an example of geothermal fields in New Zealand. At the Wairakei geothermal field illustrated in Figure 4-58, rates of amorphous silica precipitation measured in the field were 400 times faster than those obtained in the laboratory measurements. Similar rates are found at the Broadlands field. Both field data and model results indicate that stable mineral assemblages are sensitive to small differences in fluid chemistry, temperature, and pressure. Silica precipitation under potential repository conditions at Yucca Mountain could exhibit rate behavior somewhere in a range between the laboratory and field experiments (CRWMS M&O 2000b, Section 13.4.1.3).

Grants Ridge, New Mexico—Despite the high-temperature basaltic intrusion, there was no evidence of pervasive hydrothermal circulation and alteration in the country rock that could have resulted from thermally driven liquid-phase convection (CRWMS M&O 2000al, Section 3.6.2.1).

Yucca Mountain—Conceptual models for mineral evolution at Yucca Mountain (Carey et al. 1997) suggest that the most likely mineralogical reactions caused by repository heating would include dissolution of volcanic glass and precipitation of clinoptilolite, clay, and opal-CT; dissolution and precipitation of silica polymorphs (cristobalite, opal-CT, tridymite, and quartz); alteration of feldspars to clays; and finally, reactions involving calcite and zeolites. Thermodynamic modeling results indicate that the stability of various zeolites is a function of silica activity, temperature, aqueous sodium concentration, and the mineralogy of silica polymorphs. Increasing temperature or sodium concentration causes the alteration of zeolites to other phases. Kinetic data suggest that water saturation conditions are necessary for significant progress in these reactions. Therefore, under ambient conditions the reactions are likely to proceed more slowly in the Yucca Mountain unsaturated zone (excluding perched water zones) than below the water table. However, if prolonged boiling occurs in water-saturated tuffs, significant progress in such reactions can occur (CRWMS M&O 2000al, Section 3.6.2.2).

Valles Caldera, New Mexico—The contact between the Banco Bonito obsidian flow and underlying Battleship Rock tuff on the southwest rim of the caldera was studied by Krumhansl and Stockman (1988) and Stockman et al. (1994). No evidence of hydrothermal alteration was noted, suggesting that the area was unsaturated at the time of contact. The effects of heating in this unsaturated environment appeared to have been slight and were limited to the tuff nearest the contact. The principal mineralogic change in tuff near the contact was the development of feldspar-silica linings on voids in the pumiceous tuff matrix; no significant development of zeolites was found.

Paiute Ridge, Nevada Test Site—This characterization has been further developed as a distribution of "key blocks" that are defined by the fractures. An example of a sill observed on the outcrop of Paiute Ridge is illustrated in Figure 4-58. Matyskiela's (1997) study of alteration surrounding one intrusion, the 50-m-wide (160-ft-wide) Papoose Lake sill, found alteration of glass shards to cristobalite and clinoptilolite within 60 m (200 ft) of the intrusion. He observed complete filling of pore spaces with silica at fracture–matrix interfaces, thus enhancing flow along fractures by inhibiting fracture–matrix interaction. Matyskiela (1997) estimated enhanced fracture flow to be as much as five times ambient conditions. A different kind of response to heating in the repository host rock (i.e., formation of a silica cap) was predicted by Nitao (reported in Hardin 1998, Section 5.6). Those simulations show that both the fractures and the adjacent matrix (at the fracture–matrix interface) could be plugged by dissolved solids transported into the boiling zone by fracture flow. This result depends on the total quantity of dissolved solids (e.g., silica) transported to the boiling zone and on the value of the fracture porosity (a smaller value of the fracture porosity makes plugging more likely). Another set of calculations reported in Near Field Environment Process Model Report (CRWMS M&O 2000al, Section 3.2.5) indicates that only minor changes in fracture permeability would result from mineral precipitation during the thermal pulse, based on a larger estimate for fracture porosity and consideration of more complete ranges of minerals and chemical species.

Scoping analyses by Lichtner et al. (1999) provide additional perspective on the potential for plugging of porosity in the fractures and matrix. The approach considers the dissolved silica in the pore water present initially (before heating) in the tuff matrix. During heating, the matrix pore water will evaporate in a region proximal to the drift openings. If the silica contained in the pore water migrates to the fractures as evaporation occurs (ignoring evaporation in the matrix), then the degree of fracture plugging depends on the initial fracture porosity, as well as the particular silica polymorph produced. Based on this analysis, the fracture porosity used in Nitao's models (reported in Hardin 1998, Section 5.6) would be completely filled with amorphous silica for tsw33 and tsw34 units and about 50 percent filled in the tsw35 unit (CRWMS M&O 2000al, Section 3.6.2.2).

4.2.2.3 Process Model Development and Integration

Two models, namely the mountain-scale thermal-hydrologic model and multiscale thermal-hydrologic model, differing primarily in the spatial scales of interest, are used to model the effects of decay heat on water movement (
CRWMS M&O 2000al, Section 3.2.2). Both models solve heat and mass balance equations written for a fracture continuum interacting with a matrix continuum. The mountain-scale model evaluates changes in temperature and water movement at greater distance from the potential repository and is used for comparison to analogous geothermal sites. The multiscale thermal-hydrologic model provides detailed output describing temperature and moisture movement in the emplacement drifts and is abstracted directly for use in performance assessment. Both models rely on hydrogeologic properties developed for the unsaturated zone flow model.

Another model is used to calculate the drift-scale thermal-hydrologic-chemical processes and to address the change of chemistry of water entering drifts and the flow properties of the surrounding rock that control the seepage flux (CRWMS M&O 2000al, Section 3.3). A final model is used to calculate the normal and shear displacement of distinct blocks of rock surrounding the heated emplacement drift and to address the consequent change in flow properties of the near-field rock that control the seepage flux (CRWMS M&O 2000al, Section 3.5).

4.2.2.3.1 Mountain-Scale Thermal-Hydrologic Model

The mountain-scale thermal-hydrologic model simulates the effects of repository heating on the unsaturated zone, including a representation of thermally driven processes occurring in regions far away from the potential repository (
CRWMS M&O 2000ci, Section 7; CRWMS M&O 2000c, Section 3.12). This model is used to understand the mountain-scale effects of repository heating on water movement and to conclude that the mountain-scale effects of heating on water movement would not degrade the function of the host rock to divert water from the engineered barriers and the waste form. Thus, the assessment of environmental conditions within the emplacement drifts can be focused on the drift-scale effects of heating on water movement.

The simulations provide predictions for thermally driven flow of liquid water and water vapor, and distributions for temperature and moisture in the unsaturated zone, for a period of 100,000 years. Approaches for simulation of the two-phase heat and mass transport processes in fractured systems are generally based on geothermal and petroleum reservoir simulation methods. Early multiphase modeling studies identified the importance of heat pipe effects in models that include heat sources. The heat pipe phenomenon is associated with boiling conditions and can occur if sufficient liquid is transported back toward the heat source (by gravity or capillarity) to sustain the boiling (Pruess, Wang et al. 1990, p. 1241). In the dryout zone, temperature variation is controlled by heat conduction and the thermal output of the heat source. In previous mountain-scale thermal-hydrology models, planar heat sources were used to represent closely spaced emplacement drifts, with the result that thick heat pipes and condensate zones were predicted (Buscheck and Nitao 1992). In the mountain-scale model, heat sources are more discrete, and drainage can occur through the pillars between emplacement drifts.

The mountain-scale thermal-hydrologic model uses the mathematical formulation employed in previous models (Pruess 1987, pp. 2 to 11; Pruess 1991, pp. 5 to 26). The approach involves volume averaging, the dual-permeability, continuum approach for modeling coupled fluid and heat flow, and fracture–matrix interaction (CRWMS M&O 2000br, Section 6.7; Liu et al. 1998). The model was developed from the unsaturated zone flow model and uses the mathematical formulations employed in the TOUGH2 code to describe flow under thermal-loading conditions (CRWMS M&O 2000c, Section 3.12.1). For the flow model, predictions of the three-dimensional ambient temperature distribution were calibrated against measured temperature profiles in boreholes (CRWMS M&O 2000ci, Section 6.1). The same thermal properties are used in the mountain-scale thermal-hydrology model, and the ambient temperature distribution is used as the initial condition for modeling thermal effects.

Several numerical models were developed, including a three-dimensional mountain-scale model and two cross-sectional models. Plan views of the three-dimensional grid and the two-dimensional cross-sectional grids are shown in Figures 4-59 and 4-60, respectively. The three-dimensional model represents each emplacement drift and the adjacent pillars as one element, which does not resolve thermal-hydrologic processes in the vicinity of the potential repository. The two-dimensional models are more detailed and therefore allow drainage of condensed water through the pillars between drifts. The two-dimensional cross-sectional models are finer than the three-dimensional model but coarser than the drift-scale models described in Section 4.2.2.3 (see Figure 4-72 in Section 4.2.2.3.6 for a comparison). Accordingly, only the elements outside the drift elements are evaluated in the mountain-scale studies. More detailed discussion of these numerical models is presented in supporting documentation (CRWMS M&O 2000br, Section 6; CRWMS M&O 2000ci, Sections 6.2 and 6.3).

The mountain-scale thermal-hydrologic model uses input parameters that are fully consistent with the unsaturated zone flow model (CRWMS M&O 2000c, Section 3.6; CRWMS M&O 2000cj, Section 6), including hydrologic properties, the mean infiltration distribution, and the evolution of future climate states (CRWMS M&O 2000c, Section 3.5; USGS 2000b, Section 6). The simulations use an average initial thermal loading of 72.7 kW/acre (18.0 W/m2, equivalent to average areal mass loading of 60 MTHM per acre), based on a potential repository area of 1,050 acres (4,250,000 m2) (CRWMS M&O 2000c, Section 3.12.2.3; CRWMS M&O 2000ci, Section 5.1). The thermal load is scaled down by the natural decay curve over a total simulation period of 100,000 years. To account for ventilation, only 30 percent of this heat is used during the first 50 years (CRWMS M&O 2000c, Section 3.12; CRWMS M&O 2000ci, Sections 6.10 and 6.11). Figure 4-61 depicts the relationship between the thermal-hydrologic model, the input data, and supporting models.

Presented below are some of the results obtained for the two-dimensional cross-sectional models, focusing on the case with ventilation. Discussion of the other two-dimensional and three-dimensional cases is presented in supporting documentation (CRWMS M&O 2000ci, Sections 6.10 and 6.11). The results presented below are based on the higher-temperature operating mode.

TemperatureFigure 4-62 shows the distribution of temperature along cross section NS#2 after 1,000 years of thermal load (a) without ventilation and (b) with ventilation for the first 50 years. The variations in net infiltration across the model influence the evolution of temperature in both cases. Higher temperatures are predicted to occur in areas with less infiltration. The plots show dryout only in the immediate vicinity of the potential repository. At a lateral distance of 100 m (330 ft) or more from the potential repository, no substantial increases in temperature are predicted, which suggests that buoyant, mountain-scale convection of the gas-phase, driven by production of water vapor and the heating of the gas-phase, will be important within or near the potential repository.

With or without ventilation, the predicted maximum temperatures at the centers of the pillars between emplacement drifts do not exceed boiling. Results also show that the long-term temperature response anywhere in the unsaturated zone is unaffected by 50 years of ventilation. Temperatures at the base of the PTn unit may increase to boiling conditions if the potential repository is not ventilated. Such temperatures could induce property changes in the PTn unit from mineral alteration, particularly in areas with less infiltration. With 50 years of ventilation as planned, temperatures in the PTn unit are predicted to increase to an average of 40° to 45°C (104° to 113°F). At the top of the CHn unit, below the potential repository, the models predict a maximum temperature of 70° to 75°C (158° to 167°F), which will occur after 2,000 and 7,000 years, depending on location. This temperature range suggests minimal potential for thermally induced mineralogical degradation of zeolites in the CHn unit. At the water table, the models predict a maximum temperature of 65° to 70°C (149° to 158°F), compared to the average ambient temperature of 30°C (86°F).

Saturation—With or without ventilation, all fractures in the drifts at the potential repository horizon are completely dry within the first few years of thermal loading. The dryout zone expands and reaches a maximum thickness between 600 to 1,000 years after thermal loading (CRWMS M&O 2000c, Sections 3.10.5.2 and 3.12.3.2). Thereafter, the dryout zone contracts and persists for several thousand years (i.e., 1,000 to 3,000 years), well after cessation of above-boiling conditions (BSC 2001o, Section 6.3.5.1).

Figure 4-63 shows contour plots of matrix liquid saturation at 1,000 years (a) without ventilation and (b) with ventilation (CRWMS M&O 2000c, Section 3.12.3.2). The plots show large decreases in matrix liquid saturation only near the drifts. Liquid saturation at a lateral distance of more than 50 m (160 ft) away from the potential repository is predicted to remain at near-ambient conditions. A zone of decreased saturation extends to about 50 m (160 ft) above and below the emplacement drifts. After 5,000 years, the matrix liquid saturation is almost fully recovered to ambient conditions. The fracture and matrix liquid saturation within the pillars remain at near-ambient levels and are controlled primarily by changes in the climate state.

Thermally Driven Changes in Percolation Flux—Although ventilation lasts only 50 years, it results in changes in flux patterns that persist for hundreds of years. This is because the heat removed by ventilation will delay the onset of boiling conditions. The model results show that ventilation leads to lower thermally driven liquid and gas fluxes because the effective thermal load is only 30 percent of the total heat output during the early period in which the heat output of the emplaced waste is greatest. Figure 4-64 shows the calculated liquid flux through the potential repository horizon with ventilation.

With ventilation, thermal evolution is delayed because the temperature of the rock must rise to boiling conditions to maximize evaporation and heat pipe activity. At 100 years (50 years after closure), the liquid flux is 10 mm/yr (0.4 in./yr). However, flux recovers to more than 150 mm/yr (6 in./yr) at 500 years as a result of higher temperatures. Because of the decay in thermal output, and the increased infiltration imposed by climate change, the liquid flux in the fracture continuum at the potential repository horizon recovers to the ambient percolation flux after approximately 5,000 years. With ventilation, the fracture flux through the pillars between the drifts remains at or above ambient conditions and temperatures remain below boiling throughout the thermal pulse. Beyond approximately 500 years, the fracture liquid flux in the pillars returns to the ambient percolation levels, though these levels change in response to changes in climate state. The liquid saturation within the pillars remains at or above the ambient liquid saturation because of vapor condensation and liquid flux being channeled through the pillars. This flux may be increased by condensate drainage for several hundred years.

Liquid flux through the pillars is the direct result of the counter-current vaporization and condensation (reflux) cycle associated with the development of heat pipe conditions predicted by the mountain-scale thermal-hydrologic model (CRWMS M&O 2000c, Section 3.12.3.3). At early times (less than 100 years), thermally induced liquid fluxes above the drift are up to two orders of magnitude larger than the ambient percolation flux. At around 1,000 years, model results show fracture liquid flux at several locations to be 20 to 50 mm/yr (0.8 to 2.0 in./yr). At around this time, mountain-scale thermal model results suggest liquid flux may occur downward into the drifts (CRWMS M&O 2000c, Section 3.12.3.3).

Mountain-scale thermal-hydrologic modeling shows that temperature increases at the base of the PTn unit and the top of the CHn unit are moderate (much less than boiling) and are not expected to cause mineralogical alteration of these units. Temperatures at a lateral distance of more than 100 m (330 ft) outside the potential repository are expected to remain near ambient conditions. The mountain-scale model predicts little impact on thermal-hydrologic conditions outside the potential repository layout, possibly because large-scale gas-phase buoyant convection promotes upward mass and heat transfer.

4.2.2.3.2 Multiscale Thermal-Hydrologic Model

The multiscale thermal-hydrologic model (
CRWMS M&O 2000cf) is used to model the effects of decay heat on water movement at the drift scale and within the emplacement drifts. Simulation results are used either directly or indirectly as input to other process models or abstractions for TSPA. The model uses the nonisothermal unsaturated-saturated flow and transport computer code to solve coupled heat and mass-transfer equations (CRWMS M&O 2000cf, Section 6).

This model estimates thermal-hydrologic conditions within the emplacement drifts and the surrounding rock as functions of time, waste package type, and location in the potential repository. The multiscale thermal-hydrologic model is based on drift-scale thermal-hydrologic models of the types used to analyze the field-scale thermal tests, combined with three-dimensional thermal models that incorporate thermal interactions between waste packages and mountain-scale heat transfer. The resulting multiscale model predicts fine-scale behavior within the emplacement drifts while including the effects of repository-scale heat transfer (CRWMS M&O 2000cf, Section 6).

The multiscale thermal-hydrologic model relates the results from different types of smaller models to capture the effects of key factors that can affect thermal-hydrologic conditions in the emplacement drifts and surrounding rock (CRWMS M&O 2000cf, Section 6):

  • Variability of the percolation flux on the scale of the potential repository

  • Temporal variability of percolation flux (as influenced by climate change)

  • Uncertainty in percolation flux (as represented by the mean, high, and low infiltration flux conditions described in Section 4.2.1.3.3)

  • Variability in hydrologic properties (e.g., those properties which control fracture–matrix interaction and capillarity of fractures) on the scale of the potential repository

  • Edge-cooling effect (cooling increases with proximity to the edge of the potential repository)

  • Dimensions and properties of the engineered barrier system components, such as the drip shield and invert

  • Waste-package-to-waste-package variability in heat generation rate

  • Variability in overburden thickness on the scale of the potential repository

  • Variability in rock thermal conductivity (emphasizing the host rock units) on the scale of the potential repository.

Table 4-16 lists the specific performance measures that are predicted using the multiscale thermal-hydrologic model. This model simulates time-varying thermodynamic conditions such as temperature, relative humidity, and evaporation and condensation. Thermal-hydrologic conditions are also predicted, such as liquid saturation, liquid-flux (percolation flux), gas-phase flux, and capillary pressure. These results are used in coupled models, in engineered barrier system performance analyses, and to describe the overall evolution of the thermodynamic and thermal-hydrologic environments in the emplacement drifts and surrounding rock for TSPA (CRWMS M&O 2000al, Sections 2.2.1 and 3.2; CRWMS M&O 2000cf, Sections 6.11 and 6.12).

The need for a multiscale modeling approach stems from the fact that the performance measures depend on thermal-hydrologic behavior within a few meters of the emplacement drifts and also on thermal and thermal-hydrologic behavior at the mountain scale. A single, explicit numerical model would require a large number (many millions) of grid blocks. The multiscale model is used to estimate results that would be obtained if such a single model were feasible. The multiscale thermal-hydrologic model also predicts the effects of different waste package types (e.g., different commercial spent nuclear fuel waste packages, codisposal of DOE high-level radioactive waste) on the various performance measures (CRWMS M&O 2000cf, Section 6.1).

The multiscale thermal-hydrologic model comprises four major models and includes multiple scales (mountain and drift), multiple dimensions (one-dimensional, two-dimensional, and three-dimensional), and different assumptions regarding the coupling of heat transfer to fluid flow (conduction-only and fully coupled thermal-hydrologic). The four types of models are (CRWMS M&O 2000cf, Section 6.1):

  • Line-averaged-heat-source, drift-scale, thermal-hydrologic model

  • Smeared-heat-source, mountain-scale, thermal-conduction model

  • Smeared-heat-source, drift-scale, thermal-conduction model

  • Discrete-heat-source, drift-scale thermal-conduction model.

It is useful to think of the line-averaged-heat-source, drift-scale, thermal-hydrologic model as the basis model. These two-dimensional drift-scale thermal-hydrologic models use hydrologic properties and other input data that are fully consistent with the unsaturated zone flow model. The models are run for 31 locations spaced throughout the potential repository area (Figure 4-65) for a range of thermal loading values that represents the influence of edge-cooling. Variability of the hydrologic properties at the scale of the potential repository is represented by the 31 locations (CRWMS M&O 2000cf, Section 5.1.1).

The other three types of models, which are thermal conduction-only models, are required to account for the influence of three-dimensional mountain-scale heat flow and three-dimensional drift-scale heat flow on drift-scale thermal-hydrologic behavior. Further details on these thermal-conduction-only models, the method used to modify the two-dimensional thermal-hydrologic model results to reflect the three-dimensional scale effects, and the representation of air spaces in the drifts are provided in supporting documentation (CRWMS M&O 2000cf, Sections 6 and 7.1).

In the multiscale thermal-hydrologic model, the waste package sequence is explicitly modeled in a drift segment and repeated hundreds of times throughout the footprint of the potential repository. Model geometry is consistent with the design basis described in this report. The emplaced waste packages fall into several categories, representing those that would contain (1) pressurized water reactor spent nuclear fuel, (2) boiling water reactor spent nuclear fuel, (3) DOE high-level radioactive waste, and (4) DOE (naval) spent nuclear fuel. All waste packages are emplaced at the same time and follow the same average thermal decay function (as a percentage of initial heat output). The 70 percent heat-removal efficiency and the 50-year ventilation period are applied uniformly throughout the potential repository footprint. The overall average areal mass loading of the potential repository for multiscale model applications is 60 MTHM/acre.

Figure 4-65 illustrates the potential repository footprint used in the model; this footprint closely approximates the actual plan view of the perimeter within which waste would be emplaced. Thirty-one locations are shown, which represent the lateral variability in hydrologic properties, stratigraphic thickness, and boundary conditions.

Major results of the multiscale thermal-hydrologic model summarized in this section are based on the higher-temperature thermal operating mode described in Section 2. During the preclosure period, host rock temperatures remain below the boiling point for the mean- and upper-infiltration flux cases, while boiling occurs in the host rock for the lower-flux case.

The expected duration of temperatures above the boiling point of water (96°C [205°F]) on the surfaces of the waste packages varies for three main reasons: (1) location within the repository layout, (2) spatial variation in the infiltration of recharge water at the ground surface, and (3) variability in the heat output of individual waste packages. The repository edges would cool first because they lose heat to the cooler rock outside the layout. The repository center would cool more slowly because heat flow would be limited mainly to the upward and downward directions. Water percolating downward through the host rock in response to infiltration at the ground surface would hasten cooling of the repository; locations with greater percolation will cool sooner. There will be relatively large variations in the heat output of individual waste packages depending on the type and age of the waste they contain. Each of these effects is represented explicitly in the multiscale thermal-hydrologic model, and the results of this model are used in TSPA.

During the preclosure period, peak waste package temperatures of 100°C (212°F) for the mean flux case, and 110°C (230°F) for the low flux case, are expected to occur at 10 to 15 years; peak drift-wall temperatures of 86°C (187°F) for the mean flux case, and 96°C (205°F) for the low flux case, are expected to occur at 20 to 25 years. Edge-cooling effects will not strongly affect preclosure temperatures.

During the postclosure period, peak waste package temperatures of 128° to 178°C (262° to 352°F) for the mean flux case, 127° to 189°C (261° to 372°F) for the low flux case, and 124° to 173°C (255° to 343°F) for the high flux case are expected to occur at 60 years. The difference in peak waste package temperature between the hottest and coldest waste packages for the mean flux case would be approximately 50 C° (90 F°). During the very early postclosure period, edge-cooling will have a small effect on temperatures. By 100 years, the influence of edge-cooling will be considerable, with waste package temperatures varying by 65 C° (117 F°) (98° to 163°C [208° to 325°F]) from the edge to the center of the potential repository for the mean flux case.

A typical waste package under nominal conditions would have an average surface temperature above the boiling point of water for about 1,000 years (CRWMS M&O 2000cf, Figure 6-50). For the mean infiltration case, the average temperature on the surfaces of all 21-PWR waste packages would cool to below the boiling point of water after about 1,400 years. Lower infiltration rates could increase the time until the waste packages cool to this temperature. Depending on the infiltration rate and the location in the repository, the time to cool could be less; for example, for the mean infiltration rate a 21-PWR waste package located on the edge of the repository would cool to below the boiling temperature of water within about 300 years. For brevity, these ranges are described elsewhere in this report as "from hundreds to thousands of years."

Liquid-phase flux in the host rock above the drift would be influenced by dryout and heat pipe activity. Heat pipe behavior can increase the liquid-phase flux in fractures to well above the ambient percolation flux. However, the duration of this effect would be greatly decreased in this design in comparison with the repository design used for the Viability Assessment (CRWMS M&O 1998g, Chapter 3, Section 3.5.4). For the higher-temperature operating mode, the increased liquid-phase flux is calculated to last for less than 600 years (the duration of the present-day climate period).

The maximum lateral extent of boiling temperatures (away from the drift wall) is a good indication of spatial extent of dryout around the emplacement drifts. The lateral extent of boiling would be greater for the low infiltration-flux case than for the mean or upper-flux cases. For the hottest waste package location and the lower flux distribution, the maximum lateral extent of boiling would be 18 m (59 ft); because the drifts would be 81 m (266 ft) apart, a maximum of approximately 44 percent of the potential repository area would exceed the boiling point. If the estimated infiltration increases, this percentage would decrease.

There is a much greater difference in dryout behavior (as evidenced by the maximum lateral extent of boiling and by relative humidity reduction) between the mean and low infiltration-flux cases than between the mean and upper flux cases. Therefore, if one considers a percolation threshold above which rock dryout becomes substantially limited by percolation, the threshold would be near the mean infiltration-flux case. Larger values of the percolation flux greatly limit the calculated extent of boiling temperatures and rock dryout.

4.2.2.3.3 Drift-Scale Thermal-Hydrologic-Chemical Processes and Models

Figure 4-40 shows schematically the relationships between thermal-hydrologic and geochemical processes in the zones of boiling, condensation, and water drainage in the rock surrounding the potential repository, particularly in the rock above the emplacement drifts. The emphasis in this section is on the changes in flow properties of the host rock due to these processes. Modeling of thermal-hydrologic-chemical effects on the aqueous chemistry of seepage water and gas composition in the potential repository host rock is described in Section 4.2.3.3.

Changes to hydrologic properties were evaluated using the thermal-hydrologic-chemical model. The thermal-hydrologic-chemical models account for two-dimensional heat and mass transfer within the drifts and in the surrounding rock, using separate continua in the rock to represent the connected network of fractures and the rock matrix in which the fractures reside (CRWMS M&O 2000al, Section 3.2.3.4.2). The thermal-chemical model component is implemented using TOUGHREACT, a heat transfer, mass transfer, and reactive-transport code. The model is used to calculate dissolution and precipitation of minerals that could change the porosity and permeability of the fracture system (CRWMS M&O 2000al, Section 3.2.3.4.3).

Thermal-hydrologic aspects of the model, such as the heating rate, ventilation, infiltration flux, and other boundary conditions, are identical to the thermal-hydrologic models discussed previously in this section. Discretization of the model domain is illustrated in Figure 4-66. Several cases were evaluated for different infiltration conditions (i.e., lower, mean, and upper) using the same climate-change scenarios used for the unsaturated zone flow model (BSC 2001o, Section 6.3).

The dual-permeability method was selected for modeling thermal-hydrologic-chemical processes. This is an important selection because a realistic representation of chemical interactions between fractures and the rock matrix depends on realistic representation of hydrologic interactions. The active fracture model is used (Liu et al. 1998; CRWMS M&O 2000bq, Section 6.4.5). Each matrix gridblock and each fracture gridblock has its own pressure, temperature, liquid saturation, water and gas chemistry, and mineralogy. Water–mineral reactions are considered to take place under either kinetic or equilibrium conditions, using simulation methods similar to those described by Reed (1982) and Steefel and Lasaga (1994). Because the dissolution rates of many mineral-water reactions are quite slow, most phases are treated using pseudo-order reaction kinetics.

As stated in Section 4.2.2.2.1.3, the initial water and gas chemistry selected for use in the thermal-hydrologic-chemical model is based on the chemical composition of matrix pore water collected from Alcove 5 (BSC 2001o, Sections 4.1.3 and 6.1.2). Although the rock permeability of the matrix is many orders of magnitude smaller in the matrix than in the fractures, the TSPA-SR thermal-hydrologic-chemical model assumes that infiltrating water in the fractures has the same composition as matrix pore water. This is justified in that the chloride-sulfate-type matrix-derived pore water composition is more concentrated in total dissolved minerals. The actual water composition within fractures tends to be more dilute (i.e., bicarbonate-type water). For all thermal-hydrologic-chemical modeling, the initial water composition is set to be the same in the fractures and matrix throughout the model domain (BSC 2001o, Section 4.1.3). Thermal-hydrologic-chemical model simulations were repeated with two sets of rock minerals to evaluate the sensitivity ofcalculated results to the mineral assemblage selected (BSC 2001o, Section 6.2). A full-simulation case included the major minerals found in the fractures and matrix of all rock units that are likely to be thermally perturbed based on mineral occur rence in deeper zeolitized units. A limited-simulation case included those minerals needed to represent basic aspects of Drift Scale Test data, such as pH and gas-phase carbon dioxide, while neglecting other species, such as silicates, ferric minerals, and fluorides. Details on derivation of model inputs, the numerical model and supporting sensitivity studies are provided in Drift-Scale Coupled Processes (DST and THC Seepage) Models (BSC 2001o, Sections 4.1 and 6.2).

Thermal-hydrologic-chemical model results consist of projections for the composition of water and gas that may enter the emplacement drifts for 100,000 years, including a 50-year preclosure period with ventilation. Figure 4-67 shows liquid saturation and temperature in the rock around the drifts at 600 years for the three infiltration conditions (i.e., lower, mean, and upper). This is the approximate time of the maximum extent of dryout for each infiltration case investigated. Note that these models predict conditions that would be encountered near the center of the potential repository, and that cooler conditions would be found near the edges.

Time histories for gas-phase carbon dioxide concentration, pH, chloride concentration, and total dissolved carbon concentration are predicted for the several locations in the host rock at the drift wall (CRWMS M&O 2000c, Figures 3.10-8 to 3.10-11). These results are summarized in Section 4.2.3.3, where they are used as boundary conditions for the physical and chemical environment in the emplacement drifts.

Comparison with data from the Drift Scale Test shows that the limited simulation (calcite, silica, and gypsum minerals only) matches observed chemical data more closely than the full simulation (including silicates, iron, and fluorides). However, for longer duration of reflux and boiling, as would be encountered in the potential repository, the system may trend toward the chemistry of the more complex full simulation.

Porosity changes in the rock matrix and fractures are directly related to volume changes from mineral precipitation and dissolution. Changes in fracture permeability are approximated using a parallel-plate model approach with fractures of uniform aperture (Steefel and Lasaga 1994, p. 556). Matrix permeability changes are calculated from changes in porosity using the Carmen-Kozeny relation (Bear 1988, p. 134). Capillary pressure in the matrix and fractures is modified using the Leverett scaling relation (Slider 1976, pp. 290 to 297), as previously mentioned in Section 4.2.1.3.2.

The calculated changes in fracture porosity for rock near the emplacement drifts, for the full simulation, at a simulation time of 10,000 years, are shown in Figure 4-68 for the three infiltration conditions (i.e., lower, mean, and upper). The fracture porosity change is expressed as a percentage of the initial porosity. Maximum porosity decrease is predicted for the high-infiltration case, predominantly above the drift. For all cases, the porosity change is relatively small (less than 1 percent of the initial porosity). For the limited simulation, porosity decrease results mainly from calcite precipitation, as was interpreted from the Drift Scale Test simulations. For the full simulation, the fracture porosity change is dominated by zeolite reactions. Because the fracture porosity changes are small compared to total fracture porosity, permeability changes are negligible and thermal-hydrologic processes will not be significantly affected by mineral precipitation or dissolution (CRWMS M&O 2000al, Section 3.2.3.4.3).

4.2.2.3.4 Drift-Scale Thermal-Hydrologic-Mechanical Processes and Models

The drift-scale thermal-hydrologic-mechanical modeling in support of the TSPA accomplishes two objectives. This section emphasizes the potential effect on hydrologic properties in the surrounding rock resulting from thermal-mechanical changes, specifically potential effects on permeability. The drift degradation analysis presented in
Section 4.2.3.3.5 assesses thermally caused movement of blocks on fractures intersecting the drift and the potential for rockfall to affect the engineered barrier system.

Most prior thermal-mechanical modeling had the objective of determining the evolution of stresses in the near-field rock, in order to estimate the requirements for rock support in the emplacement drifts. These models treated the rock as a continuum and conservatively assumed that the mid-pillar locations were symmetry boundaries. This assumption is conservative (produces higher calculated stresses) because the overall repository footprint can expand due to the heating. The movement of rock blocks at fractures was captured in the continuum models by the rock mass properties, such as the coefficient of thermal expansion (the fractional expansion of the rock per degree of temperature rise). Measurements of this coefficient depends on the size of the sample, with the coefficient decreasing as the scale moves from core to small blocks to small field tests and then to large field tests (CRWMS M&O 1999m, Table 9-3). The decrease in expansion coefficient can be attributed to the increased number of fractures which can accommodate expansion of the rock blocks. The results of the modeling indicate that horizontal compressive stresses increase more than vertical compressive stresses during the thermal pulse, due to the stiff boundary conditions at the mid-pillar locations.

In this analysis supporting the TSPA-SR model, the distinct-element code 3DEC is used to simulate normal and shear displacement and other behavior on discrete fractures in the rock mass surrounding the drift (CRWMS M&O 2000al, Section 3.5.2). Fracture orientations and fracture densities are represented and discretized (gridded) in three dimensions. Fracture orientations are based on field observations (Albin et al. 1997). Joint and rock-mass properties used in the calculation are based on field and laboratory studies of rock and fracture behavior, such as those by Barton et al. (1985) and Olsson and Brown (1994). Fracture densities are based on the assumption that only a few well-connected fractures are mechanically and hydrologically active (CRWMS M&O 2000al, Section 3.5.2).

Calculated joint deformations are used to compute permeability deformation values over a period of 1,000 years to capture the effects of heating and cooling. The mathematical formulation is described in the Near-Field Environment Process Model Report (CRWMS M&O 2000al, Section 3.5.2.3). Using this formulation, shear deformation always produces an increase in permeability, while normal deformation will increase permeability if the fracture opens and decrease permeability if the fracture closes. In general, fracture closing is expected during the heating phase in response to thermal expansion, while fracture opening occurs during cooling, the effects of shear displacements notwithstanding.

The results of the calculation (CRWMS M&O 2000al, Section 3.5) are that the major thermal-mechanical effect on fracture permeability occurs during cooldown due to both shear and normal deformation. Shear deformation of fractures during the cooldown causes permeability of the fractures in a region within two drift diameters of a drift wall to increase in permeability as much as an order of magnitude. Specifically, shear deformation on vertical fractures during cooldown produces the maximum amount of permeability change. Farther away from the drift wall, smaller increases in permeability (a factor of five) may occur on vertical fractures (CRWMS M&O 2000al, Section 3.5.3).

Results also indicate that normal deformation of fractures causes permeability to increase but to a lesser degree than shear deformation. Normal deformation during heating causes permeability to decrease significantly within one drift diameter of the drift wall. During cooldown, some vertical fractures above the drift open, thus increasing the permeability by a factor of two from the ambient values (CRWMS M&O 2000al, Section 3.5.3).

Ambient fracture permeability at the repository horizon is high, greater than 10-13 m2 (100 millidarcy) (CRWMS M&O 2000ch, Table 5). Near-Field Environment Process Model Report (CRWMS M&O 2000al, Section 5.5) concludes that the potential order of magnitude increase in permeability due to shear movement is not likely to significantly affect seepage.

4.2.2.3.5 Limitations and Uncertainties

As discussed in
Section 4.1.1.2, uncertainties are an inherent component of the TSPA method. Uncertainty is introduced through the conceptual model selected to characterize a process, as well as the mathematical, numerical, and computational approaches used to implement the model. Uncertainty is also introduced from imperfect knowledge of important parameters used for input to the models (e.g., physical properties).

The DOE has performed several supplemental activities to address uncertainties and limitations in the TSPA-SR model. Additionally, as noted in Section 4.1.4, the DOE is evaluating the possibility for mitigating uncertainties in modeling long-term repository performance by operating the design described in this report at lower temperatures. Consequently, some of the models describing the effect of decay heat on water movement have been updated since the TSPA-SR model. Some alternative conceptual models have been implemented, and sensitivity analyses conducted to address parameter and model uncertainties. These supplemental model analyses are summarized in FY01 Supplemental Science and Performance Analyses (BSC 2001a, Section 1; BSC 2001b, Section 1).

Mountain-Scale Thermal-Hydrologic Model—The unsaturated zone flow model is the basis for the mountain-scale thermal-hydrologic model. Therefore, uncertainties associated with the unsaturated zone flow model also pertain to the mountain-scale model (see Section 4.2.1.3). The spatial resolution of the numerical mountain-scale thermal-hydrologic model is large enough that it limits the interpretation of calculated temperature, saturation, and fluxes within the emplacement drifts and in the host rock near the drift openings. Consequently, the multiscale thermal-hydrologic model, with a finer spatial scale than the mountain-scale model, is used for these purposes.

Multiscale Thermal-Hydrologic Model—Two categories of model uncertainties are defined for the multiscale thermal-hydrologic model: (1) uncertainties related to thermal-hydrologic modeling and (2) uncertainties related to the multi-scale estimation methodology.

The unsaturated zone flow model provides an important basis for the multiscale thermal-hydrologic model. Consequently, uncertainties associated with the flow model are propagated to the multiscale thermal-hydrologic model. Uncertainties related to the special features of the unsaturated zone seepage process model are not incorporated in the TSPA-SR model. As discussed in this section, the TSPA-SR model uses a conservative approach for calculating seepage during the thermal pulse, based on the percolation flux (5 m [16 ft] above the drift openings) and calculated by the multiscale thermal-hydrologic model. Although this is considered to be a conservative approach (leading to overestimated thermal seepage into drift openings), model improvements have been suggested such as discrete representation of flow focusing along faults and fractures and representation of episodically increased flow in the host rock (BSC 2001a, Section 4.3.5.6).

The percolation flux contains uncertainties related to mean, high, and low infiltration flux conditions, temporal variability (e.g., changes in climate), and spatial variability (e.g., repository cooling at the edges). The principal effects of these uncertainties on TSPA are related to the timing for cooling and return of moisture to the in-drift environment. Differences in timing of hundreds to a few thousands of years (as discussed in Section 4.2.2.3.2) will have a minor impact on the estimated longevity of the drip shield and waste package (with expected lifetimes greater than 10,000 years calculated in the TSPA-SR model). Another uncertainty is the arrangement and heat output of different types of waste packages, either in a lower- or higher-temperature operating mode. An arrangement of waste types with different heat output is used in the multiscale model, but the extent to which this is representative of repository conditions is uncertain, pending final decisions on repository design and operations parameters.

Uncertainties related to the multiscale estimation methodology include the effects of mountain-scale gas-phase convective circulation and the movement of water vapor along the axis of emplacement drifts from warmer to cooler regions. The TSPA-SR modeling approach did not include heat transfer by these mechanisms, probably resulting in overestimation of predicted temperatures and the duration of the thermal pulse.

Supplemental studies have added insight to some uncertainties and limitations identified with the TSPA-SR model. Approaches included alternative thermal property sets representing lithophysal tuff, Monte Carlo simulations of spatially heterogeneous fracture properties, simulation of a through-going vertical fracture zone intersecting the drift opening, and decreased thermal loading representing a lower-temperature operating mode (BSC 2001a, Sections 4.3.5.3, 4.3.5.4, and 4.3.5.5). Specific studies included:

  • Alternative thermal seepage models incorporating the effects of flow focusing and episodicity (BSC 2001a, Section 4.3.5). The results support previous TSPA-SR analyses that found thermal seepage to be negligible for the relatively small values of seepage that may occur.

  • Representation of fractures by spatially heterogeneous properties (BSC 2001a, Section 5.3.1.4.2).

  • Evaluation of the bulk permeability of the host rock, the thermal conductivity of the lithophysal Tptpll unit and the invert, and the effects of lithophysal porosity of the Tptpll unit on vapor storage and heat capacity (BSC 2001a, Sections 5.3.1.4.1, 5.3.1.4.7, 5.3.1.4.8, 5.3.1.4.9, and 5.3.1.4.10).

  • Evaluation of mountain-scale gas-phase convective process and the movement of vapor along the axis of the emplacement drifts (BSC 2001a, Sections 5.3.1.4.4 and 5.3.2.4.6).

Supplemental studies substantiate the overall model and analytical results of the TSPA-SR model, providing quantification of uncertainty either quantitatively or qualitatively. Thermal-hydrologic-chemical and thermal-hydrologic-mechanical processes will not significantly affect temperature or relative humidity within the emplacement drifts (BSC 2001a, Sections 5.3.1.4.5 and 5.3.1.4.6). In evaluating a lower-temperature operating mode, supplemental studies show thermal perturbations to be less than in the higher-temperature operating mode (BSC 2001b, Sections 3.2.2.6 and 4.2.2).

Drift-Scale Thermal Hydrologic-Chemical Processes Model—Uncertainties exist in the chemical parameters used to describe mineral precipitation and dissolution. Temperatures and flow rates are better constrained than other parameters of these models. Geochemical reactions are strongly influenced by temperature, the presence of water, and mass transport; so while the spatial distribution of mineral precipitation and dissolution is considered to be representative, the quantities minerals formed or dissolved at a given time and location are more uncertain. Furthermore, the potential for rapid boiling in the rock to cause mineral behavior outside the range of the models is recognized.

The assumed initial water and gas compositions as well as the geochemical conceptual model may also introduce uncertainty. Uncertainty is also recognized in the relationship used to determine fracture permeability based on changes in fracture porosity. These model uncertainties affect predictions of host rock pore water chemistry and changes in permeability in the host rock caused by thermal-hydrologic-chemical processes.

Model parameters, such as the effective reaction rates, are calibrated to results from field thermal tests, including the Drift Scale Test. Comparison of model predictions to geochemical data is important for confidence building, and such comparisons have shown that the model is reasonable. Results from the full simulations yield higher pH values than measured in water samples from the Drift Scale Test, which results from a greater calculated reaction rate for feldspars. Therefore, porosity changes as a result of feldspar alteration in the potential repository host rock will probably be slower than predicted, suggesting that the TSPA-SR model gives an upper bound on such changes.

Supplemental studies have added insight to some uncertainties and limitations identified with the TSPA-SR thermal-hydrologic-chemical process and abstractions. These activities included a range of different input data and assumptions, such as host rock mineralogy and thermodynamic input data (BSC 2001a, Sections 4.3.6.3.1 and 6.3.1.6). Studies included:

  • Supplemental model validation activities of water and gas compositions conducted for Drift Scale Test results (BSC 2001a, Sections 4.3.6.3.1 and 4.3.6.9)

  • Supplemental sensitivity studies of different initial water and gas boundary conditions (BSC 2001a, Section 4.3.6.5)

  • Simulated evolution of water and gas compositions in the lower lithophysal as well as the middle nonlithophysal tuff unit (BSC 2001a, Section 6.3.1.4.3).

These additional drift-scale thermal-hydrologic-chemical model simulations support the TSPA-SR evaluation, concluding only negligible changes in fracture permeability resulting from thermal-chemical rock and water interactions (BSC 2001a, Sections 4.3.6.7.4 and 4.3.6.9).

Drift-Scale Thermal-Hydrologic-Mechanical Processes and Models—The thermal-hydrologic-mechanical model, used in the TSPA-SR model, used a simplified thermal history to calculate results (CRWMS M&O 2000al, Section 3.5). Coupling the model to the multiscale thermal-hydrologic model has been suggested as a model improvement. Uncertainties in the TSPA-SR thermal-hydrologic-mechanical model also include mechanical boundary conditions, joint and block properties, block geometry, and the calculation of permeability change due to aperture change.

The model domain for drift-scale near-field thermal-mechanical models is bounded by the mid- pillar locations between the emplacement drifts. For stress calculations, using a zero lateral displacement boundary condition at this location would be conservative (i.e., producing greater horizontal compressive stress) because in reality the entire repository layout can expand, allowing somedisplacement of the mid-pillar locations during the heating period.

The thermal-hydrologic-mechanical model used in the TSPA-SR model imposed boundary conditions equal to the ambient in situ stress (CRWMS M&O 2000al, Section 3.5). This is nonconservative (producing smaller stresses) because it is equivalent to assuming that more lateral displacement of the repository layout would occur. An alternative method would calculate the overall large-scale repository response in a coarsely gridded model and use the results to develop time-dependent displacement and stress boundary conditions for a drift-scale model. An additional advantage of this method would be the ability to consider the variability of the thermal-hydrologic-mechanical response due to proximity to an edge of the repository.

The joint and rock mass properties are based on field and laboratory studies (CRWMS M&O 2000al, Section 3.5.2). As discussed in Section 4.2.2.3.4, the effective thermal expansion coefficient depends on the size of the sample because of the tendency for fractures to deform in response to thermal stress changes. A similar situation exists for discrete fracture models because the rock between the discretely modeled joint may also include fractures. Also, the input values used for joint friction and cohesion are both variable and uncertain. Possible model improvements include adjusting the model grid for more gradual transitions in block sizes, also using site-specific fracture mapping data to develop the block sizes and shapes.

Supplemental studies have added insight to some uncertainties and limitations identified with the TSPA-SR thermal-hydrologic-mechanical process evaluation. Two models were used to assess selected uncertainties related to thermal-hydrologic-mechanical processes: a revised and extended discrete fracture (distinct element) analysis and a fully coupled thermal-hydrologic-mechanical continuum model (BSC 2001a, Section 4.3.7). These studies evaluated different boundary conceptual models, a simplified cubic-block conceptual model, and alternative empirical nonlinear relationships to calculate permeability change from porosity change (BSC 2001a, Section 4.3.7). Sensitive parameters were identified (e.g., residual permeability and rock stiffness) and conclusions are similar to those reached for the TSPA-SR model: permeability changes are within the range of the ambient seepage model, and thus uncertainty is already captured in the TSPA-SR model (BSC 2001a, Section 4.3.7.5).

4.2.2.3.6 Alternative Conceptual Processes

As with limitations and uncertainties, some of the following alternative conceptual models have been implemented or addressed in supplemental uncertainty analyses. These results are summarized in FY01 Supplemental Science and Performance Analyses (
BSC 2001a, Section 1; BSC 2001b, Section 1).

Mountain-Scale Thermal-Hydrologic Model and Multiscale Thermal-Hydrologic Model—Alternative conceptual models can be organized in several categories: representation of fractured rock in numerical models, selection of representative property values, the potential for permanent changes in those properties from the effects of heating, and alternative models implemented to quantify uncertainties.

The host rock is represented in the mountain-scale and multiscale models as a continuous porous medium, although the rock contains discontinuities such as fractures and fracture zones. An alternative model represents the fractures discretely, and the resulting discrete fracture model approach has been applied to example problems (Hardin 1998, Section 3.3.3). The approach is very computationally intensive and probably not practical for drift-scale and mountain-scale calculations. In addition, the number of interconnected fractures present in the host rock is so large that features of the network can be represented by a continuous medium (Section 4.2.1.3.1.1). Use of the discrete fracture model approach has been limited to modeling studies that support understanding of thermal-hydrologic processes.

Another category of alternative models involves the manner in which the network of fractures in the host rock is represented by a continuous medium. The thermal-hydrologic models described here represent fracturing using a dual-permeability continuum approach, based on the active fracture concept, which is also used in the unsaturated zone flow model (Section 4.2.1.3.1.1). The dual-permeability approach controls the movement of liquid and gas between fractures and the adjacent intact rock. Other available approaches include dual-porosity models and the equivalent single-continuum model (Hardin 1998, Section 3.3). The need for dual-permeability has been demonstrated by comparison to field thermal test data (Hardin 1998, Section 3.4); other approaches have been determined to provide less realism than the dual-permeability approach.

Alternative models for potential permanent changes in thermal and hydrologic properties of the host rock may be summarized as follows:

  • Heating, cooling, and resulting water movement occur in a system with fixed thermal and hydrologic properties (such as porosity, permeability, and thermal conductivity). Properties of the rock may vary with temperature and water saturation but return to pre-repository values after the temperature returns to ambient levels.

  • The same processes occur, but thermal effects permanently alter certain properties of the host rock through the action of coupled processes. For example, thermal-hydrologic-chemical coupling may change the hydrologic properties because of dissolution and precipitation of minerals in different regions.

The understanding of thermal-hydrologic-chemical effects on flow is summarized in Section 4.2.2.3.3. The model results indicate that changes in fracture porosity and permeability caused by chemical dissolution and precipitation will be minor compared to total porosity and permeability; hence, the first conceptual model (i.e., stationary properties) is selected as the most credible. This is the conceptual basis for both the mountain-scale thermal-hydrologic model and the multiscale thermal-hydrologic model.

As noted in the previous Section 4.2.2.3.5, supplemental studies have addressed additional alternative models:

  • Supplemental studies implemented an alternative seepage method. Instead of using percolation flux as input to the seepage abstraction model, the unsaturated zone seepage model is used, incorporating new models for focusing seepage flow along discontinuities (e.g., faults) and episodic flow (BSC 2001a, Section 4.3.5).

  • Supplemental studies implemented effects of mountain-scale gas-phase convective circulation and the movement of water vapor along the axis of emplacement drifts from warmer to cooler regions (BSC 2001a, Sections 5.3.1.4.4 and 5.3.2.4.6).

  • Supplemental studies implemented effects of vapor storage and altered heat capacity within lithophysal cavities (porosity) of the Tptpll unit (BSC 2001a, Sections 5.3.1.4.1 and 5.3.1.4.9).

  • Sensitivity studies incorporated drift-scale heterogeneity such as the influence of drift-scale heterogeneity of fracture properties, including permeability, porosity, and capillary properties (BSC 2001a, Section 5.3.1.4.2).

Supplemental analyses employing alternative models substantiate the overall model and analytical results of the TSPA-SR model (BSC 2001b, Sections 3.2.2 and 4.2.2).

Drift-Scale Thermal-Hydrologic-Chemical Processes and Models—A model proposed by Matyskiela (1997) suggests that silica precipitation in the rock matrix adjoining fractures will strongly reduce the permeability of the matrix, resulting in significantly decreased imbibition of percolating waters. The time required for strong sealing by silica was estimated for volcanic glasses under saturated conditions. Matyskiela (1997) observed complete filling of pore spaces with silica at fracture–matrix interfaces around a basaltic magma intrusion, the 50-m (160-ft) wide Papoose Lake sill, in the Paiute Ridge area of the Nevada Test Site. He estimated that fracture flow could be enhanced five times in magnitude with the sealing of the matrix pores (Matyskiela 1997, pp. 1117 to 1118). Formation of a silica cap by plugging of fractures with siliceous minerals, as predicted by recent simulations conducted for the potential repository at Yucca Mountain (Hardin 1998, Section 8.5.1), is the opposite behavior. More recent simulation modeling (CRWMS M&O 2000al, Section 3.3.3.5) has shown that fracture plugging will be of limited importance, given estimated fracture porosity of 1 percent.

Lichtner et al. (1999) showed that for a given matrix porosity, fracture plugging depends on the fracture porosity and the particular silica mineral that precipitates. The two-phase numerical simulation results suggest that at distances of tens of meters from the larger Paiute Ridge intrusion in their study, prolonged boiling conditions were established for times on the order of several thousands of years. Amorphous silica, with its higher solubility, is more readily transported by water and therefore produces the largest decrease in porosity, followed by chalcedony and quartz. For substantial sealing of fractures, a very small value of the fracture porosity is necessary. Lichtner et al. (1999) questioned the conclusions of Matyskiela (1997).

Comparison of the geochemical environment around a potential repository at Yucca Mountain with the geochemical environment around a basaltic magma intrusion is provided in supporting documentation (CRWMS M&O 2000c, Section 3.10.9). As discussed in Section 4.2.2.3.3, models of permeability changes due to mineral precipitation indicate that any such changes will be minimal. The sealing effects of silica deposition will probably be less developed at Yucca Mountain because (1) devitrified tuff reacts slowly compared with volcanic glass, (2) unsaturated fractures have less wetted surface area, and (3) the silica concentration in condensate draining through fractures will probably be limited by reaction rate processes.

As noted in Section 4.2.2.3.5, supplemental studies have addressed additional alternative models, including:

  • Alternative initial water and gas compositions boundary conditions (BSC 2001a, Section 4.3.6.5)

  • Alternative representation of the host rock, including the explicit representation of the Tptpll lithophysal unit mineralogy (BSC 2001a, Section 6.3.1.4.3).

Supplemental analyses employing alternative models substantiate the overall model and analytical results of the TSPA-SR model (BSC 2001b, Section 3.2.4.2).

Drift-Scale Thermal-Hydrologic-Mechanical Processes and Models—Alternative approaches fall into two categories: continuum versus discrete fracture models and method of coupling thermal-mechanical results to hydrologic flow. Both continuum and discrete fracture models have been used on the project. The continuum approach is satisfactory for calculating spatially averaged stress fields but is unable to resolve fracture displacements that affect permeability. The discrete fracture method can calculate movement of a significant number of representative fractures, which can then be related to permeability change.

Fracture displacement through normal or shear movement results in aperture change. The aperture change can be used to calculate both fracture porosity and fracture permeability, based on assumptions about fracture geometry. The approach used in Section 4.2.2.3.4 was to calculate fracture permeability change directly from fracture aperture change, using an empirical relationship based on laboratory studies. An alternative approach, used for thermal-hydrologic-chemical modeling in Section 4.2.2.3.3, assumes a fracture geometry (parallel plates) and calculates permeability change from theoretical considerations.

As noted in Section 4.2.2.3.5, supplemental studies have addressed additional alternative models, including:

  • A revised and extended distinct element analysis and a fully coupled thermal-hydrologic-mechanical continuum model (BSC 2001a, Section 4.3.7)

  • A simplified cubic block conceptual model and alternative empirical nonlinear relation ships to calculate permeability from porosity (BSC 2001a, Section 4.3.7).

Conclusions from the supplemental studies are similar to those reached for the TSPA-SR model: permeability changes are within the range of the ambient seepage model, and thus uncertainty is already captured in the TSPA-SR model (BSC 2001a, Section 4.3.7.5).

4.2.2.3.7 Model Calibration and Validation

Mountain-Scale Thermal-Hydrologic Model—There are no directly applicable data for validation of the mountain-scale response to thermal loading associated with the potential repository. However, numerical models of geothermal and petroleum systems can be validated from a wealth of field-scale testing and geothermal production data. The validity of mountain-scale model predictions is demonstrated by corroborative results from the modeling of analogue systems, from previously published unsaturated zone modeling studies, and from field-scale thermal tests in the Exploratory Studies Facility.

Table 4-15 in Section 4.2.2.2.3.3 lists selected geothermal systems (and, where available, analyses of those systems) that are comparable to the mountain-scale model. Applications for thermal-hydrologic modeling include detailed studies of the genesis, production history, and future performance of geothermal fields. Justification for the modeling approaches used in the mountain-scale thermal-hydrologic model is found in the successful modeling of fluid and heat transport in large natural subsurface systems for which extensive field data are available. The magma intrusion analogues for thermal-hydrologic-chemical processes are discussed in Section 4.2.2.3.5. In addition, models for the recently completed Single Heater Test (Tsang and Birkholzer 1999) and the ongoing Drift Scale Test (CRWMS M&O 2000c, Section 2.2.4) use the same approach and input data as the mountain-scale model. In summary, the mountain-scale thermal-hydrologic model is considered valid because of its similarity to the models developed for field tests and the demonstrated validity of the geothermal analogue models.

Multiscale Thermal-Hydrologic Model—The multiscale thermal-hydrologic model uses a method based on industry-standard finite-difference software that includes both mass and energy balances. Model documentation addresses input data, assumptions, initial and boundary conditions, software, uncertainties, and other information required to replicate the model results.

Several validation approaches are used for the multiscale thermal-hydrologic model, including comparison of thermal-hydrologic modeling with results from the Large Block Test and the Drift Scale Test and comparison of multiscale thermal-hydrologic model results with mountain-scale thermal-hydrologic simulation, as described below. These comparisons are discussed in more detail in supporting documentation (CRWMS M&O 2000cf, Section 6.13).

Thermal-Hydrologic Models of the Large Block Test—A similar modeling approach was used to simulate the entire history of the Large Block Test (CRWMS M&O 2000cf, Section 6.13.1). As an example of model comparison with data, Figure 4-69 shows simulated borehole temperature profiles compared to observed temperatures. Evaluation of goodness-of-fit to measured temperatures shows accuracy of a few degrees Celsius.

Figure 4-70 shows the simulated and measured liquid-phase saturation profiles along another borehole in the Large Block Test. The simulated dryout zone develops more slowly than observed, but the difference resolves with time. At later times, the model is in close agreement.

Thermal-Hydrologic Models of the Drift Scale Test—Thermal-hydrologic modeling of the Drift Scale Test heating period, from startup to the present, was compared to observations (CRWMS M&O 2000cf, Section 6.13.2). As an example of model comparison with data, Figure 4-71 compares the simulated and measured temperatures along an observation borehole. The model results are in close agreement with measured temperatures, only slightly overpredicting temperatures in the dryout zone and slightly underpredicting temperatures in the sub-boiling zone.

In general, close agreement with observed temperature in the sub-boiling zone indicates that heat flow there is dominated by conduction and that the value of thermal conductivity is reasonable. Close agreement in the region close to the heated drift indicates that (1) thermal radiation is adequately represented inside the heated drift, (2) heat flow in the boiling and above-boiling zones is dominated by conduction, and (3) the value of thermal conductivity in this region is reasonable.

Comparison of the Multiscale Thermal-Hydrologic Model with the Mountain-Scale Numerical ModelFigure 4-72 compares the drift-wall temperature predicted by the multiscale thermal-hydrologic model with temperatures predicted by east–west cross-sectional mountain-scale thermal-hydrologic models (for details, see CRWMS M&O 2000cf, Section 6.13.3). The mountain-scale thermal-hydrologic model is coarsely gridded, so the comparison is limited to drift-wall temperature from the multiscale thermal-hydrologic model vs. drift temperature from the mountain-scale thermal-hydrologic model.

Before comparing the two approaches (Figure 4-72), it is important to discuss other differences in the models. Differences between the multiscale model and mountain-scale modeling approaches include:

  • The temperature predicted by the mountain-scale model is for a grid block that occupies an entire drift, so it produces a lumped representation of drift temperature, whereas the multiscale model resolves temperature differences within the drift.

  • The mountain-scale thermal-hydrologic model uses a line-averaged heat source that axially smooths the differences between hotter and cooler waste package locations.

  • The initial areal power density (at emplacement) in the multiscale thermal-hydrologic model is 92.3 kW/acre, compared to 99.4 kW/acre in the mountain-scale thermal-hydrologic model.

  • The mountain-scale thermal-hydrologic model representation of the heated footprint of the potential repository extends slightly further to the west than in the multiscale thermal-hydrologic model.

Near the center of the potential repository, the approaches predict nearly the same duration of boiling (Figure 4-72, left). Near the edge, the mountain-scale model predicts a longer duration of boiling (Figure 4-72, right). During the post-boiling period, the temperatures predicted by the approaches are in close agreement. During the early heating period, the coarse gridding of the mountain-scale model cannot capture the more rapid changes that the multiscale model predicts. Also because of the coarse gridding, the mountain-scale model tends to overpredict heat pipe behavior. Given the differences in technical approach, the models are in reasonable agreement throughout much of the thermal evolution of the potential repository.

Drift-Scale Thermal-Hydrologic-Chemical Processes and Models—Comparison of model predictions with data from the Drift Scale Test involves (1) modeled patterns of fracture drainage compared to locations where water has been collected during the test, (2) comparison of carbon dioxide concentrations from gas samples, and (3) comparison of the evolution of water composition in boreholes sampled over time.

Simulated distributions of temperature and carbon dioxide concentration are shown in Figure 4-73 for the limited simulation approach (Case 2). The comparison shows that the simulations follow general trends in measured carbon dioxide concentration (CRWMS M&O 2000c, Figure 3.10-5). Two exceptions were when heater power loss occurred temporarily and when the gas samples were acquired at boiling temperatures and condensation occurred during the sampling. Detailed comparison of the modeling results and the measured carbon dioxide concentration data is discussed in Drift-Scale Coupled Processes (DST and THC Seepage) Models (BSC 2001o, Section 6.2.7.2).

The simulated pH of water in fractures is shifted to pH 6.5 from an initial pH of approximately 8.3, with the lowest pH values predicted where carbon dioxide concentration is greatest. The predicted shift in pH is similar to that observed in water samples collected from the Drift Scale Test. Chloride concentrations in waters collected from hydrology boreholes are considerably more dilute (a factor of 5 to 10) than the matrix pore water predicted by the modeling. Other species, such as calcite and silica, show similar trends in the modeled fracture water compositions compared to measured water compositions.

In the model simulations, calcite is the major phase forming in the zone above the heaters, although the quantity of calcite is small. Amorphous silica also precipitates but is less abundant than calcite. Direct observation of calcite precipitation or dissolution has yet to be observed in the Drift Scale Test, which is ongoing. However, other indications, such as the composition of water samples, provide indirect evidence for calcite precipitation. Fracture porosity changes predicted for the Drift Scale Test after 20 months are very small (on the order of 0.01 percent of the initial fracture porosity). Such small changes would likely have no measurable effect on the hydrologic properties of the rock.

Drift-Scale Thermal-Hydrologic-Mechanical Processes and Models—Model calibration and validation of fracture displacements due to heating and cooling is done using laboratory and field test results. Calibration includes normalizing models to test results and using observations to determine physical phenomena needed in the models. An example of the latter is the observation of sharp movements in multipoint-borehole-extensometer data at specific times; this observation has resulted in the adoption of a discrete fracture conceptual model in Section 4.2.2.3.4. Validation is the comparison of calculated results to test data, with the calculations being done independent of the test data themselves. Both continuum models and discrete fracture models have been compared to test data from the Large Block Test, the Single Heater Test, and the Drift Scale Test; these tests are described in Section 4.2.2.2.3.2.

4.2.2.4 Total System Performance Assessment Abstraction

Because of the limited thermal effect on water movement at large distances, the mountain-scale thermal-hydrologic model results are not directly included in the TSPA. Abstraction of thermal-hydrologic model results for the TSPA was therefore based on the near-field behavior predicted by the multiscale model.

The mountain-scale thermal-hydrologic model shows that the impacts of repository heating on temperature, saturation, and liquid flux in the unsaturated zone will have limited duration and will be limited to the repository area. Some effects, such as elevated liquid flux associated with heat pipe activity, will be limited to the vicinity of the emplacement drifts. Also, mineralogical alteration of the overlying and underlying hydrogeologic units will be minimal with preclosure ventilation. Accordingly, the thermal-hydrologic effects on far-field flow and transport are not currently considered in the TSPA (
CRWMS M&O 2000c, Table 3.13-2).

The purpose of the report Abstraction of NFE Drift Thermodynamic Environment and Percolation Flux (CRWMS M&O 2000cc) is to abstract the multiscale, process-level thermal-hydrologic model results (CRWMS M&O 2000cf) so that they can be implemented in the TSPA model. The purpose of the abstraction is to simplify the detailed thermal-hydrologic description of the potential repository that is produced by the multiscale model. An averaging process ("binning") is used to compute these quantities, based on a subdivision of the repository footprint that preserves a wide range of thermal-hydrologic variability. Multiscale model results used directly in support of the TSPA-SR model include waste package temperature, relative humidity at the waste package surface, and the percolation flux in the host rock 5 m (16 ft) above the emplacement drift. Temperature and relative humidity are used for the corrosion model, and percolation flux is used for the seepage model. Time-histories of waste package temperature, percolation flux, evaporation rates, and maximum and minimum waste package surface temperatures are also provided (CRWMS M&O 2000cc, Section 6.3). The abstraction of thermal-hydrologic data represents the potential variability and uncertainty in thermal-hydrologic conditions. It provides a quantitative description of thermal-hydrologic variability (i.e., from variability in the host rock unit, edge proximity, waste package type, infiltration rate, and climate state) and also incorporates uncertainty associated with the infiltration (i.e., lower, mean, and upper).

Abstraction of predicted water and gas compositions for the mean infiltration rate (with climate change), including both limited and full mineral suite simulations, is summarized in Section 4.2.3.4. Also, since the predicted thermal-hydrologic-chemical coupled effects on flow properties are relatively small, the effects on seepage are not included in the TSPA-SR.

The abstracted seepage model used in TSPA-SR performance assessment calculations did not include changes in permeability due to thermal-mechanical effects. This approach was based on Near-Field Environment Process Model Report (CRWMS M&O 2000al, Section 5.5) and was confirmed by supplemental TSPA analyses (BSC 2001a, Section 4.3.7.4.4).

4.2.3 Physical and Chemical Environment

The lifetimes of the drip shield and waste package will depend on the environmental conditions to which they are exposed: the in-drift physical and chemical environment (
CRWMS M&O 2000a, Section 3.3). Once a waste package is breached, the transport of radionuclides released from the waste form also depends on the environment in the emplacement drifts.

This section describes estimates of how the physical and chemical conditions in the drifts are expected to evolve with time, based on the thermal operating mode described in Section 2. The description is based on the estimated response of the host rock to heating and on data concerning behavior of the engineered materials used to construct the potential repository. The estimates are based primarily on results from laboratory and field-scale testing, supplemented by observations from natural and man-made analogues.

As noted in Section 4.1.4, the DOE is evaluating operation of the repository at lower temperatures. Operating the repository at lower temperatures may change the evolution of the physical and chemical conditions in the drifts described in this section. The data and analytical results presented in this section reflect the effects of higher-temperature operating mode conditions, specifically the process models and abstractions employed in the TSPA-SR model (CRWMS M&O 2000a). Alternative thermal operating modes and supplemental uncertainty evaluations related to the in-drift physical and chemical environment models are described and summarized in FY01 Supplemental Science and Performance Analyses, (BSC 2001a, Sections 6.3.3., 6.3.4, 7.3.1, 7.2.4, and 10.3.4; BSC 2001b, Sections 3.2.4.2 and 4.2.4).

Results of the in-drift models used directly in the TSPA-SR model include time-dependent estimates of the infiltration rate, temperature, and relative humidity at the drift wall, as well as the evolution of the chemical conditions at the drift wall over four discrete time periods: (1) preclosure, (2) boiling, (3) transitional cooldown, and (4) extended cooldown (CRWMS M&O 2000a, Section 3.3.3.4.2).

Physical Environment—The physical environment is described by the evolution, with time, of thermal-hydrologic conditions in the emplacement drifts. Estimation of the temperature, relative humidity, and rate of evaporation at locations throughout the potential repository is described in Section 4.2.2. The results show that every location in the potential repository could evolve from very dry conditions at temperatures greater than boiling to cooler conditions and increasing humidity. Differences between locations are limited mainly to the timing of these changes—for example, the duration of boiling temperatures on the waste package will depend on its location in the repository layout, local infiltration flux, and the heat output of individual waste packages (see Section 4.2.2.3.2). Cooling, and return of moisture to the emplacement drifts, would occur hundreds to a few thousands of years sooner at the edges of the potential repository, compared with the center. Cooling also would occur sooner at locations where there is greater recharge of water from the ground surface.

The potential for liquid water seepage into the emplacement drifts is described in Section 4.2.1. Seepage is combined with temperature, relative humidity, and evaporation rate to represent the physical environment for the engineered barriers in the TSPA-SR. Diversion of seepage by the drip shield and waste package is described in Section 4.2.5. The potential for condensation under the drip shield during the thermal period is also discussed in that section. A model of the flow of liquid water through breaches in the drip shield and waste package is used to assess advective releases of radionuclides in TSPA-SR. Thermal-hydrology, seepage, and water diversion model results that were developed in Sections 4.2.1, 4.2.2, and 4.2.5 are implicit in the following description of the physical and chemical environment and are not discussed further in this section.

The physical environment also includes the potential for rockfall, which could damage the drip shields or waste packages. The effects of rockfall are estimated based on observations from site characterization and use approaches that represent the effects of heating and seismic loading. Estimates of block size and rockfall frequency have been used to design the drip shield, which is designed to withstand rockfall over its design lifetime and thereby protect the waste package. The approach to estimating rockfall events is also described in this section.

Chemical Environment—Important processes affecting the chemical environment include evaporation and condensation of water, the formation of salts, and the effects of gas composition. During the thermal period, relative humidity will likely control the equilibrium solution chemistry and is therefore a principal descriptor of the chemical environment. The approach to analyzing the chemical environment involves several types of predictions:

  • Composition of water and gas in the host rock around the drifts that can enter drift openings

  • Composition of waters within the drifts that can further evaporate and form precipitates and salts

  • The effect of microbial activity on the chemical environment

  • The effects of engineered materials such as steel and cement

  • The chemical environment at the surfaces of the drip shield and waste package.

These analyses are complementary and together describe the in-drift chemical environment as it is represented for performance assessment. Each is either incorporated explicitly in TSPA-SR or has been considered to have minor consequences to system performance and is excluded from consideration. The approach for each analysis is described in the following sections.

4.2.3.1 Conceptual Basis

This section describes the conceptual models that form the basis for analytical treatment of processes in TSPA. Although the descriptions may contain statements that appear to be definitive, it is important to recognize that there are uncertainties associated with the selection of appropriate conceptual models. Alternative conceptual models are discussed in
Section 4.2.2.3.7. Model results based on the selected conceptual models are generally considered to be best estimates, incorporating uncertainty, such that the models are suitable for use in TSPA.

4.2.3.1.1 Conceptual Basis for the Composition of Liquid and Gas Entering the Drifts

Composition of Liquid Seepage—The chemistry of waters in the host rock will act as a boundary condition on the in-drift chemical environment (
CRWMS M&O 2000al, Section 3.4.2). During the thermal pulse, water vapor will move away from the heated drifts while liquid water percolates downward and replaces the water that evaporates in a thermal refluxing process (Section 4.2.2). The percolating waters will contain dissolved chemical species, such as sodium, calcium, sulfate, chloride, carbonate, and silica (CRWMS M&O 2000a, Section 3.3.3.4.2). When evaporation occurs, the chemical species will be left behind in the rock as precipitated minerals and salts.

The areal extent of the dryout zone produced by the higher-temperature operating mode would shrink as the heat output from the waste packages decreased with time. This will cause the region of boiling conditions to slowly converge on the drift openings. Liquid water will tend to sweep through formerly boiling regions, redissolve precipitates and salts, and move them closer to the openings. Soluble salts will tend to be concentrated near the openings. Depending on local hydrologic conditions, this process could cause seepage to be concentrated in soluble salts relative to the ambient (preheating) water composition.

With seepage, salts such as calcium carbonate and sodium chloride can form in the drifts (for example, from dripping and evaporation) directly on the drip shield or waste package. In the TSPA-SR model, only a fraction of the waste package locations in the repository would be affected by seepage (Section 4.2.1), especially during the thermal pulse when the conditions are relatively dry (Section 4.2.2). Without seepage, the effects of chemical processes in the host rock on the in-drift chemical environment will be limited to the gas composition. Seepage will be more likely in the future as the climate changes to the cooler, wetter, glacial-transition conditions discussed in Section 4.2.1. However, by the time the effects of this climate change propagate down to the host rock, cooldown will have progressed so that the drip shield temperature will be below boiling throughout the potential repository (CRWMS M&O 2000cf, Section 6.11.4).

After cooldown, and after soluble salts precipitated during the thermal pulse are redissolved and remobilized, the composition of seepage water will become increasingly similar to the ambient percolation in the host rock units. Some minerals precipitated during the thermal pulse may be stable, or slow to dissolve, but effects from such minerals are incorporated into the thermal-hydrologic-chemical model (BSC 2001o, Section 6.1).

Composition of the Gas Phase—The gas-phase composition in the host rock will also act as a boundary condition on the in-drift chemical environment (CRWMS M&O 2000a, Sections 3.3.3.2.3 and 3.3.3.4.2). The gas composition will initially be similar to atmospheric air, but during the thermal pulse, the gas phase will be strongly modified by evaporation of water and by interaction with carbon dioxide in waters and carbonate minerals (BSC 2001o, Section 6.2.7.2). Change in the in-drift gas flux and composition will affect water pH, including water that may occur on the surface of the drip shield or waste package CRWMS M&O 2000ck, Section 6.2.4; CRWMS M&O 2000a, Section 3.3). Relationships among thermal, hydrologic, and chemical processes in the host rock around the drift openings, and within the drifts, are depicted schematically in Figure 4-74.

Evaporation of water from heating of the host rock will cause much of the dissolved carbon dioxide to be released as gas (the remainder will be precipitated as carbonate minerals). The gaseous carbon dioxide will form a broad halo around the drift openings that encompasses the cooler region where water vapor condenses. Condensate will be enriched in carbon dioxide and slightly acidified (CRWMS M&O 2000al, Section 3.6.4.2). In the zone of evaporation closer to the drift openings, calcite and other carbonate minerals will be precipitated but may be redissolved later during cooldown when liquid water returns. Oxygen will exhibit simpler behavior because it is less soluble in water and forms different kinds of minerals. Analyses of uncertainty and the thermal sensitivity of chemical conditions within the drifts are described in FY01 Supplemental Science and Performance Analyses (BSC 2001a, Section 6.3).

4.2.3.1.2 Conceptual Basis for Evolution of the Chemical Environment for the Engineered Barriers

At low relative humidity, any minerals or salts that exist in the emplacement drifts will be dry. These minerals and salts may be introduced as ventilation dust, from evaporation of construction water, or from the occurrence of seepage when the drifts are hot and dry. As the relative humidity increases during cooldown, salts will condense to trap water vapor from the air (deliquescence) and form brine. Minerals and salts may occur on a small scale, for example, an evaporated droplet on the surface of the drip shield. Eventually, increasing humidity will cause the brines to capture more water vapor and become diluted. Seepage (where it occurs) will flush soluble salts from the drifts and redissolve the less soluble minerals. After the thermal period, the composition of waters and gases in the emplacement drifts will return to ambient (preheating) conditions (
CRWMS M&O 2000as, Section 3.1.2.5.1).

Minerals and Salts Formed by Evaporation—Temperatures well above the boiling point of pure water (96°C [205°F] at the potential repository elevation), and associated low relative humidity, will persist hundreds to thousands of years after emplacement, depending on the location within the repository layout, the local infiltration flux, and the heat output from individual waste packages, as discussed in Section 4.2.2.3.2. These conditions are conducive to evaporation. Seepage into the drifts during this period, if it occurs, may be transient, but the nonvolatile dissolved constituents will accumulate in the drifts as salts and minerals. These solids will include soluble salts with the potential to form brines when relative humidity increases to approximately 50 percent and greater (CRWMS M&O 2000as, Section 3.1.2.5.1). Another source of minerals and salts is dust on the drip shield or waste package surfaces, which could be derived from the rock or from aerosols introduced from the atmosphere by ventilation during the preclosure operational period (CRWMS M&O 2000ck, Section 6.2).

Composition of Waters in the Emplacement Drifts—Temperature and humidity would be slowly varying in the repository, so equilibrium relationships will apply between brine concentration and relative humidity. Thus the relative humidity, which is readily predicted from thermal-hydrologic calculations (Section 4.2.2), is a good estimator of brine composition (CRWMS M&O 2000ck, Section 6.4). As noted previously, all waste package locations will evolve to high relative humidity, but at different rates, so the effect on TSPA is limited mainly to the timing of changes in equilibrium brine composition and seepage.

Where seepage occurs, the rate of seepage entering the drift during cooldown will eventually exceed the rate of evaporation in the drift as the thermal output decays (CRWMS M&O 2000cl, Section 6.2). When this happens, liquid water will begin to flow through the drift. Any brines present will tend to be diluted and flow out of the drifts with the seepage. If seepage is flowing into the drift, then considerable dilution of brines has probably occurred already from the associated humidity. Therefore, the details of seepage mixing with pockets of brine are not critical to predicting the chemical environment.

Microbial Activity in the Emplacement Drifts—Microbial activity is important primarily because of the potential for microbially influenced corrosion of the waste package. Microbes may also increase the rates of degradation for other engineered materials and can contribute to radionuclide transport. Bacteria and fungi, including molds, occur naturally in the host rock and would also be introduced by repository construction and operation. Because of dryness and elevated temperature, heating the rock at the potential repository will arrest microbial activity for time periods on the order of hundreds of years. However, heating will redistribute water so that cooler locations are wetter, which will locally increase the microbial activity. Factors that will limit microbial activity include elevated temperature, low humidity, and availability of nutrients and energy from engineered materials. Engineered materials such as steel may provide metabolic energy sources and limiting nutrients, such as phosphate. Measures to control the use of organic materials during construction and operation will also limit microbial activity (CRWMS M&O 2000as, Section 3.1.2.4).

4.2.3.1.3 Conceptual Basis for the Effects of Engineered Materials on the Chemical Environment

Engineered materials have the potential to affect the chemical environment as they degrade. Drip shield and waste package materials will degrade slowly, and the effects on oxygen availability will be minor. Structural steel in the emplacement drifts will corrode and consume more oxygen, depending on the relative humidity. Cementitious grout will be used to anchor rock bolts in a portion of the potential repository (
CRWMS M&O 2000cg) and may contribute cement leachate to the chemical environment.

Degradation of Steel and Alloys—Measured penetration rates for the titanium drip shield and the Alloy 22 waste package outer barrier are small (see Section 4.2.4). These materials obtain corrosion resistance from a passive layer of oxides on the exposed surfaces. The rate of oxygen consumption from maintaining the oxide layer is directly related to the penetration rate.

Steel will be used in roof supports, rails, and beams in the invert that support the drip shields and waste packages (see Section 2.4). These are preclosure structural applications; the steel will readily degrade during the postclosure period when humidity returns. Steel corrosion will begin when the temperature is near boiling and much of the air is displaced by water vapor. The steel is likely to corrode relatively quickly, within a few hundreds to thousands of years. While it is active, steel corrosion may affect the oxygen budget in the emplacement drifts (CRWMS M&O 2000cm, Section 6.2.2).

Cementitious Materials—Cement that will be used in rock bolt installation (like all Portland-based cements) is an assemblage of minerals and other phases, some of which dissolve to produce highly alkaline leachate. The composition of leachate will be determined by the solubilities of cement mineral phases. The phases present in "fresh" cement are more alkaline than those in aged cement because of carbonation and other processes (CRWMS M&O 2000cg, Section 5.3.1). Therefore, the leachate composition can be bounded using readily available information on cement composition. Several factors will act to limit the quantity and quality of leachate produced, including cement carbonation, low grout permeability, limited exposure to seepage flux, and neutralization of leachate by carbon dioxide in the drift environment (CRWMS M&O 2000cg, Sections 6.3.1 and 6.7.5).

Colloidal Particles Produced by Degradation of Engineered Materials—Colloidal particles are important as carriers for radionuclides, particularly isotopes of relatively insoluble elements, such as americium and plutonium. These radionuclides are transported very slowly (or not at all) in groundwater, except for colloidal modes of transport. Colloid generation and radionuclide transport are discussed at length in Sections 4.2.6, 4.2.7, 4.2.8, and 4.2.9. Colloids derived from the host rock will be present in seepage water. Additional colloids will be generated from degradation of engineered materials in the potential repository, including waste forms and other materials within the waste package (CRWMS M&O 2000cn, Section 6.1; CRWMS M&O 2000cg, Section 6.6).

Engineered materials in the emplacement drifts will consist mainly of corrosion-resistant alloys, cement, crushed rock, and steel. Corrosion products of titanium and Alloy 22 are mechanically stable (hence the corrosion resistance of these materials) and are unlikely to form significant colloids. Degradation of cementitious materials can form colloids (Hardin 1998, Chapter 6), but the usage of cement and its exposure to seepage will be limited. The invert ballast material will be crushed tuff derived from the host rock (see Section 2.4.1); the resulting colloids will be similar to colloids introduced with seepage. By contrast, steel in the emplacement drifts will be an abundant source of ferric-oxide colloids that are potentially important for radionuclide transport.

It is anticipated that quantities of colloids will be mobilized as a result of alteration of both the high-level radioactive waste and spent nuclear fuel waste forms. Colloid abundance within a breached waste package will depend on the extent of waste form alteration and the alteration products formed from in-package steel components. Colloid abundance and stability also depend on many environmental factors, including the ionic strength, pH, cation concentrations, colloid content of groundwater entering the waste package from the drift, presence of fulvic and humic acids, and microbe fragments (CRWMS M&O 2000co, Section 1). The colloid source term and transport models are described in Sections 4.2.6 and 4.2.7 of this report, respectively.

Contribution of Engineered Materials to Microbial Activity—As stated previously, microbial activity is important primarily because of the potential for microbially influenced corrosion of the waste package; microbial activity may also increase the rates of degradation for other engineered materials and contribute to radionuclide transport. Microbes exploit chemical reactions (oxidation-reduction) that are rate-limited under abiotic conditions by providing faster alternative reaction pathways that also support cell-building and energy production (CRWMS M&O 2000cp, Section 6.3.1.1). Engineered materials include metals, which are important sources of reactants for these chemical reactions.

Engineered materials in the emplacement drifts will consist mainly of corrosion-resistant alloys, steel, cement, and crushed rock. Each of these can interact with microbes in particular ways, but steel will probably be the most important contributor. Steel will oxidize completely in the first few hundred years after sufficient humidity returns to the in-drift environment; after that, it will contribute little to microbial activity.

4.2.3.1.4 Conceptual Basis for the Environment on the Surfaces of the Drip Shield and Waste Package

Behavior of Water on the Barrier Surfaces—The environments on the surfaces of the drip shield and the waste package will determine the potential for corrosion processes and the penetration rates. Surface chemical conditions will be controlled by temperature, humidity, gas-phase composition (especially oxygen and carbon dioxide), the composition of dripping water, and the minerals and salts that may be deposited.

During the thermal pulse, the drip shield will be warmer than the surrounding in-drift environment in some instances, and the local relative humidity at the surface will be lower than the average humidity in the drift. Under these conditions, any water on the drip shield would tend to evaporate, resulting in concentration of aqueous solutions and precipitation of minerals and salts (
CRWMS M&O 2000ck, Section 6.2.3).

The titanium and Alloy 22 surfaces will react with atmospheric oxygen to form thin, resistant layers of metal-oxides (corrosion processes and rates are discussed in Section 4.2.4). These oxide layers will be mechanically stable and chemically unreactive, which confers corrosion resistance to these materials. Because the oxides are chemically inert compared to other species in the environment, they are assumed to not contribute to the chemical evolution of aqueous solutions on the barrier surfaces (CRWMS M&O 2000ck, Section 5).

Waste packages under intact drip shields will be exposed to moisture and chemical species. Humidity will penetrate the air gap between the waste package and drip shield, although the increase of relative humidity at the waste package surface will be delayed because of the warmer temperature there (CRWMS M&O 2000ck, Section 6.2.3). Moisture will form thin films on the surfaces by adsorption or capillary condensation (CRWMS M&O 2000ck, Section 6.3). Minerals and salts will be present in small quantities from dust and aerosols transported in the drift air (CRWMS M&O 2000ck, Section 6.1). If a drip shield is breached, seepage can contact the underlying waste package, which can lead to additional precipitation of minerals or salts on the waste package surface. As the waste package cools, less evaporation will occur on the surface, and accumulation of minerals or salts will be increasingly unlikely.

Potential for Acidic Conditions—There are two mechanisms by which acidic conditions could occur on the surfaces of the drip shield or waste package: radiolysis and localized corrosion. Neither mechanism will be important in the potential repository.

Radiolysis (outside the waste package) will be caused by gamma radiation that penetrates the waste package wall and interacts with air and moisture in the environment to produce small amounts of hydrogen peroxide (CRWMS M&O 2000n, Section 3.1.6.6) and possibly other species, such as nitric acid. Radiolysis inside the waste package is discussed in Section 4.2.4. Several factors will limit the effect of radiolysis on the environment at the waste package and drip shield surfaces. The rate of gamma radiation from spent fuel and other waste forms will decline steeply within the first 1,000 years from the decay of relatively short-lived fission products. Only a small portion of the gamma radiation from the waste package will interact with the air space between the waste package and drip shield; much of this radiation will penetrate the drip shield or be absorbed within it. Acidic compounds formed in the drift environment during the thermal period will likely condense on cooler surfaces such as the drift wall and not on the drip shield or waste package, which will be warmer than their surroundings. If acidic compounds tend to precipitate on surfaces, then the drip shield will afford some protection to the waste package. Finally, the drip shield and waste package materials are resistant to attack by products of radiolysis (Section 4.2.4).

Localized corrosion can, in principle, cause acidic conditions to develop in cracks, crevices, or interfaces where exposure to the bulk chemical environment in the drifts is limited (CRWMS M&O 2000ck, Sections 6.5 and 6.6). Corrosion modes are discussed at length in Section 4.2.4, and localized corrosion is found to be of minor importance for titanium and Alloy 22.

4.2.3.1.5 Conceptual Basis for Rockfall on the Drip Shield

Fractures intersecting emplacement drifts can form "key blocks" that may become dislodged and fall directly onto the drip shields. Key blocks typically form at the crown of the existing excavations, are of minor size, and fall immediately after excavation, prior to ground support installation (
CRWMS M&O 2000a, Section 3.3.1). In the design described in Section 2.4.4, the drip shield segments will be pinned together to prevent movement (with allowance for longitudinal thermal expansion and seismic strain). Structural bracing will provide capacity to resist permanent deformation from rockfall. The drip shield segments will have overlapping and interlocking joints to impede water leakage, even with small displacements between segments. The connections between segments will tend to stiffen the structure, so that loads will be shared by adjacent segments. Determination of the size distribution for rock blocks that may fall on the drip shield and analysis of the structural response of the drip shield to rockfall are ongoing activities for which preliminary results are presented in Section 4.2.3.3.5.

4.2.3.2 Summary State of Knowledge

4.2.3.2.1 Composition of Liquid and Gas Entering the Drifts

Ambient Water Composition in the Unsaturated Zone—As discussed in
Section 4.2.2.1.3, infiltrating water chemistry could be chosen from either the pore water chemistry in the unsaturated zone at or above the repository horizon or from a more dilute composition found in the perched water or saturated zone. These are referred to as a chloride-sulfate-type water composition and a bicarbonate-type water composition, respectively. As discussed in Section 4.2.2.3.3, the thermal-hydrologic-chemical model assumes the chloride-sulfate-type water as the basis for the initial ambient water composition in fractures in the unsaturated zone.

Composition of Evaporatively Concentrated Waters—Laboratory evaporation tests have been performed using both types of waters to investigate the effects of partial evaporation on solution chemistry and to determine which minerals and salts form as the waters are evaporated completely (CRWMS M&O 2000cg, Section 6.5.2). The bicarbonate-type water has been shown to evolve by evaporative concentration to a brine with high pH, whereas the chloride-sulfate-type water evolves to a brine with a nearly neutral pH. The chloride, sulfate, and nitrate concentrations in these waters tend to increase linearly with evaporation, and precipitate only in the later stages, because they are highly soluble. The major differences in behavior are attributed to the relative abundance of bicarbonate and carbonate among the anions present.

Studies of saline lakes in the western United States show that alkaline sodium-carbonate brines of the type that were produced from bicarbonate-type water in laboratory tests occur in nature (CRWMS M&O 2000cl, Section 6.1.2). Many of these waters occur in volcanic geology similar to Yucca Mountain and have high silica content. These waters also are typically enriched in chloride and sulfate. Similarly, carbonate-poor brines of the type that were produced from matrix pore water in laboratory tests also occur in nature, such as those resulting from evaporation of sea water.

Water samples collected from the rock, in field-scale thermal tests performed at Yucca Mountain, are analogous to waters that would form during the heating of the host rock around the potential repository. The waters were collected where fractures intersected boreholes; therefore, they represent fracture waters that could potentially seep into repository drifts. The composition of such waters has varied with location, temperature, and other conditions but is dominated by condensation and interaction with fracture minerals (CRWMS M&O 2000cg, Section 6.5). The water compositions show that calcite present in the fractures is more readily dissolved than clays, feldspars, quartz, or other forms of silica. The noncalcite constituents of the waters are present in relative amounts that are similar to perched water and water found in well J-13, referred to hereafter as J-13 water (CRWMS M&O 2000cg, Section 6.5). The overall rate of condensation in these tests greatly exceeds the rate of water input from natural percolation, so the waters are dilute. Mixing of condensate with matrix pore waters was limited over the duration of the field tests but may be important over hundreds or thousands of years. From this discussion, it is appropriate to consider a range of water compositions, including bicarbonate-type and chloride-sulfate-type waters, for estimating the effects of evaporatively concentrated water and projecting the evolution of the in-drift chemical environment (CRWMS M&O 2000cg, Section 6.7.4.6).

Ambient Gas Composition—The natural composition of the gas phase in the unsaturated zone at Yucca Mountain is similar to atmospheric air, except that the carbon dioxide concentration is elevated by a factor of three. Elevated carbon dioxide is associated with soil processes near the ground surface and is commonly observed in nature. At other locations where there is more plant activity, the carbon dioxide activity in the subsurface may be an order of magnitude greater than at Yucca Mountain (CRWMS M&O 2000cg, Section 6.2). The oxygen concentration throughout the unsaturated zone is apparently close to atmospheric, indicating there are no natural processes that consume oxygen at rates nearing the rate of potential supply from the ground surface.

Analysis of radiocarbon data from the site, from sampling of pore gas in surface-based boreholes and analysis of core samples, has been used to investigate the natural processes that deliver carbon dioxide to the unsaturated zone (CRWMS M&O 2000cg, Section 6.2). The influx of carbon dioxide dissolved in infiltrating waters, plus that transported in the gas-phase, has been estimated from isotopic mass balance to be in excess of 500 mg carbon dioxide per square meter per year. The results indicate that carbon dioxide is transported in the unsaturated zone by percolating waters and by gas-phase processes, such as barometric pumping.

Effects of Heating on Gas Composition—With heating of the potential repository, carbon dioxide will be released to the gas phase from evaporating matrix pore water (CRWMS M&O 2000al, Sections 3.3.1.2 and 3.3.3). At the same time, the humidity will increase with temperature, and water vapor will displace the air. This will initially cause increased carbon dioxide activity, as observed in field thermal tests, relative to the ambient level for the unsaturated zone. Years later, the carbon dioxide activity will decrease as the temperature approaches 96°C (205°F) (the boiling point at the potential repository elevation) and air is displaced by water vapor. Eventually, during cooldown, the carbon dioxide concentration in the emplacement drifts will approach the ambient level for the unsaturated zone.

Gas-phase convection during the thermal pulse will have an important impact on the fluxes of carbon dioxide and oxygen to the emplacement drifts and on the resulting gas-phase concentrations available for chemical reactions. As the host rock heats up, and the humidity increases, the density of the gas phase will decrease, and buoyant convection may occur. Buoyant convection can occur in porous media, and has been interpreted as the cause for thermal effects associated with hot springs and igneous intrusions (Turcotte and Schubert 1982, pp. 367 to 370). Convection could circulate gas in the unsaturated zone, and thereby move air into the emplacement drifts, decreasing the humidity and increasing the availability of carbon dioxide and oxygen during the thermal pulse (CRWMS M&O 2000cg, Section 6.2.2.2). Increased availability of carbon dioxide will have the effect of buffering evaporatively concentrated, alkaline solutions, particularly above pH 10 (CRWMS M&O 2000cg, Section 6.7.5.3). Increased availability of oxygen will lessen the impact of steel corrosion on the in-drift chemical environment (CRWMS M&O 2000cg, Section 6.3.2.2).

The gas-phase concentration of carbon dioxide in the emplacement drifts during the thermal pulse will also be augmented by liquid water percolation in the host rock. As waters percolate downward toward the potential repository, carbon dioxide will exsolve because its solubility decreases with increasing temperature. If the waters are evaporatively concentrated, more carbon dioxide may be produced. The concentration of carbon dioxide species in such waters can be inferred from the composition of perched waters sampled from field thermal tests (CRWMS M&O 2000cg, Section 6.5) and is predicted using the approach described below.

Approach to Modeling the Composition of Liquid and Gas Entering the Drifts—The modeling approach for seepage and gas-phase composition during the thermal period couples thermal-hydrologic processes with both liquid and gas-phase chemistry. Temperature and evaporative concentration are determined from thermal-hydrology, while dissolution and precipitation reactions are modeled simultaneously (CRWMS M&O 2000c, Section 3.10.5). The thermal-hydrologic-chemical modeling approach is integrated with the unsaturated zone flow model (Section 4.2.1) because the same rock properties and boundary conditions are used. The same coupled model is used in Section 4.2.2 to evaluate the potential for coupled effects on hydrologic properties.

As noted previously, every location in the potential repository will be subject to similar evolution of thermal-hydrologic and chemical processes, but the timing of these processes will depend on local conditions. The principal factors that will control timing are edge versus center locations in the potential repository layout and the local percolation flux (CRWMS M&O 2000cg, Section 6.1). Based on similarity of the process evolution at different locations, it is possible to represent thermal-hydrologic-chemical behavior using a limited set of chemical computational models. The major drawback of such an approach is uncertainty of hundreds to thousands of years in the timing of thermally driven changes in chemical conditions. This uncertainty corresponds to the predicted variability in the duration of elevated temperature on individual waste packages, as discussed in Section 4.2.2.3.2.

For developing and validating this model, results from the Drift Scale Test constrain the thermal-hydrologic response, gas composition, and composition of liquid water in fractures for the first few years of repository evolution. Longer-term evolution is predicted by extrapolation. The concentrations of carbon dioxide and oxygen depend on thermally driven gas-phase circulation, and a range of models and properties are considered and compared (CRWMS M&O 2000cg, Section 6.2).

Water and gas compositions are predicted at the drift wall to represent effects of processes in the host rock and used as boundary conditions for processes in the drift (CRWMS M&O 2000c, Section 3.10.5). These predictions are then used as boundary conditions on a different model for in-drift processes, which uses an approach formulated to accommodate evaporatively concentrated waters with high ionic strength (CRWMS M&O 2000cg, Section 6.7.4.6; CRWMS M&O 2001b, Section 1). Evaluation of the carbon dioxide and oxygen budgets for chemical processes occurring in the drifts and the surrounding host rock shows that chemical reactions will probably not strongly perturb conditions in the surrounding rock (CRWMS M&O 2000cg, Section 6.7). The concentrations of carbon dioxide and oxygen will be uniform within the drift air-space (although relative humidity will vary with temperature) because gas-phase diffusion and convective mixing are rapid compared to the potential rates of consumption of these reactants. Finally, alternative chemical boundary conditions for water composition (i.e., bicarbonate-type and chloride-sulfate-type waters) are used to reflect the present state of knowledge of mobile waters in the host rock (CRWMS M&O 2000cg, Section 6.7.4).

4.2.3.2.2 Evolution of the Chemical Environment

The following discussion describes the summary state of knowledge for processes that will affect the bulk chemical environment in the emplacement drifts. This environment is distinguished from local conditions associated with cracks and crevices, microbial colonies, and within layers of engineered material degradation products.

Minerals and Salts Formed by Evaporation—As Yucca Mountain waters are evaporated to dryness, various minerals and salts are formed in sequence as the solution conditions exceed their solubility constraints. The formation of brines by evaporative concentration of natural waters can be conceptualized as a series of chemical divides, which are caused when salts with limited solubility drop out of solution (
CRWMS M&O 2000ck, Section 6.5). The concept of chemical divides is straightforward. When a salt such as calcium sulfate is precipitated from solution during continual evaporation, one or the other of the component species (i.e., calcium or sulfate) will be effectively removed from the water. The remaining one will jointly determine what precipitates next, and so on, in a series of chemical divides. (The situation is more complex for solutions with multiple soluble species, as discussed below.) The complete precipitation of all limited-solubility species is observed in natural waters that have been concentrated by evaporation, such as Owens Lake in southeastern California (CRWMS M&O 2000ck, Section 6.5).

As discussed previously, for particular conditions of temperature and relative humidity, there is a specific extent of evaporative concentration for Yucca Mountain waters, such that the solution is in moisture equilibrium with the gas phase. In other words, the solution will either evaporate or absorb moisture from the air until equilibrium is reached. This equilibrium behavior is well known and has been observed in laboratory tests (CRWMS M&O 2000ck, Section 6.6). Below a critical value of the relative humidity, or deliquescence point, the solution evaporates completely and the resulting salts remain dry. Coligative behavior is observed when solutions are concentrated by boiling; the boiling temperature increases with solution concentration until a limit is reached and the salt precipitates. The deliquescence point and the boiling point have been measured for a number of solutions of pure salts, as shown in Figure 4-75. Various salts exhibit different behavior, but solutions of salts with lower deliquescence points have consistently higher boiling points.

During cooldown of the potential repository, dissolution of salts will occur rapidly when the relative humidity exceeds the critical value for each salt present. Rapid dissolution is consistent with the observation that puddles of dissolved salt (primarily sodium chloride) occur overnight on salt flats when the relative humidity exceeds the deliquescence point for sodium chloride, but the temperature remains above the dew point (CRWMS M&O 2000cl, Section 6.1.4.2). These puddles can then dry up during the day when the relative humidity decreases.

Laboratory evaporation tests have been used to identify the minerals and salts that could form in the emplacement drifts. Tests were conducted using two different water compositions: a bicarbonate-type water and a chloride-sulfate-type water (CRWMS M&O 2000cg, Section 6.5). Evaporation was performed at below-boiling conditions (85°C [185°F]) to represent the behavior of slowly migrating waters in the engineered barrier system. At this temperature, the assemblage of mineral phases resulting from evaporation is controlled by precipitation kinetics. It is thought that the assemblages are representative of potential repository conditions.

The first set of tests using bicarbonate-type water (similar to water from well J-13) showed that the salts formed during complete evaporation included niter, halite, thermonatrite, calcite, and silica (CRWMS M&O 2000cg, Section 6.5). Calcium and magnesium precipitated as carbonates early in the evaporation process, while halite and niter, being more soluble, precipitated when evaporation was nearly complete. The precipitates formed are sensitive to the concentration of carbon dioxide in the environment, which affects the pH environment. The tests provided no evidence of hydroxide precipitates, which would indicate very high pH. When the tests were repeated with tuff particles mixed with the water, similar results were obtained.

Test results for the chloride-sulfate-type water (similar to matrix pore water or sea water) concluded that chloride and sulfate are the major anionic brine constituents, and that carbonate species are substantially less important (CRWMS M&O 2000cg, Section 6.5). As the water was evaporatively concentrated, the pH decreased to approximately 6 or lower. On complete evaporation, chloride and sulfate salts (such as halite and gypsum) predominated. In summary, test results show that calcium and magnesium carbonates, if present, will precipitate early in the evaporation process, while halite and niter are among the last salts to form. Other, additional chloride and sulfate salts will precipitate from chloride-sulfate-type waters.

Composition of Waters in the Emplacement Drifts—Throughout most of the repository performance period, composition of water in the drifts will be similar to that of ambient percolation in the host rock. During the thermal pulse, development of brine compositions that could potentially accelerate corrosion of the drip shield and waste package will depend on the chemical composition in late stages of evaporation.

In the laboratory tests described previously, evaporation of the bicarbonate-type water (similar to J-13 well water) produced a carbonate-rich brine that was highly concentrated in sodium, chloride, sulfate, carbonate, and silica. The pH increased to at least 10 and possibly higher as evaporation progressed. Evaporation of chloride-sulfate-type water (similar to matrix pore water or sea water) produced a carbonate-poor brine that was concentrated in sodium, chloride, sulfate, calcium, and magnesium (CRWMS M&O 2000cg, Section 6.5). Other dissolved components that can be enriched in natural brines include fluoride, bromide, strontium, phosphate, and boron (CRWMS M&O 2000ck, Section 6.1). Laboratory test results have shown that fluoride is not concentrated in brines representing Yucca Mountain waters, probably because these waters contain sufficient calcium and sodium to precipitate fluoride minerals (CRWMS M&O 2000cg, Section 6.5). Other components of natural brines (bromide, strontium, phosphate, and boron) are trace constituents in Yucca Mountain waters and are unlikely to achieve significant concentrations. Thus, the major chemical constituents of concentrated brines in the potential repository drifts will be sodium, calcium, magnesium, chloride, sulfate, and nitrate. For high-pH carbonate-rich brines, silica species and carbonate will also be present.

Brine compositions are outside the range that can be calculated exactly with widely used chemical activity models (i.e., much greater than 1 molal ionic strength). Accordingly, descriptions of brine behavior are more empirical, particularly with multiple dissolved species (not a single salt). The Pitzer approach (CRWMS M&O 2000cl, Section 6.4.2; CRWMS M&O 2000cg, Section 6.7.3) is based on observations of analogous ion interactions and for simulating brine behavior at ionic strengths as high as 10 molal.

Microbial Activity in the Emplacement Drifts—Laboratory testing, combined with microbial test results in the scientific literature, show that threshold conditions for microbial activity and growth are relative humidity above 90 percent, and temperature below 120°C (248°F) (CRWMS M&O 2000cg, Section 6.4.5.1). Field investigations have been performed at Yucca Mountain and at nearby Rainier Mesa to characterize microbial populations in situ (CRWMS M&O 2000cg, Section 6.4). Additional microbial observations from the exploratory tunnels at Yucca Mountain are underway.

Site characterization investigations have shown that in situ microbial growth and activity are limited by availability of water and nutrients, particularly phosphate and organic carbon. In the potential repository environment, microbial activity will also be limited by temperature and radiation (CRWMS M&O 2000cg, Section 6.4). Water will be locally available during the repository thermal evolution, as discussed previously. Phosphate is a trace constituent of J-13 water, probably because the Topopah Spring Tuff contains apatite, a phosphate mineral. Organic carbon may be limiting for some classes of organisms, such as molds, but microorganisms that fix carbon from carbon dioxide are common in rock samples obtained underground at Yucca Mountain; therefore, a source of organic carbon is not required for bacterial growth and activity. Temperature will decrease over time, so microbial activity will be possible within a few hundred years after permanent closure. Radiation from the waste package may sterilize the surrounding environment, but the shielding effect of the host rock will ensure that nearby microbes can recolonize the emplacement drifts. From this discussion, it is evident that microbial activity will occur in the emplacement drifts. The effects of microbial activity on the degradation of the waste package outer barrier material are discussed in Section 4.2.4.3.3. The potential effects on radionuclide transport are discussed in Section 4.2.7.

Approach to Modeling the Evolution of the In-Drift Chemical Environment—In this approach, relative humidity is used as a "master variable" to control the evolution of brines. Initially, the formation of minerals and salts, as seepage waters are evaporated to dryness, is modeled using a normative approach based on the laboratory tests described previously (CRWMS M&O 2000cg, Section 6.5). As relative humidity increases in the emplacement drifts, the salts formed by evaporation gradually become brines and are diluted. For the salts formed from evaporation of bicarbonate-type water (similar to J-13 well water), and considering the behavior of the salts separately and without interaction, deliquescence begins at relative humidity of 50 percent (CRWMS M&O 2000ck, p. 75 and Figure 8). At a greater value of relative humidity, all the salts are considered to be dissolved, and further changes in solution composition are calculated using the Pitzer modeling approach (or other more widely used modeling approaches) as the environment approaches ambient (preheating) conditions. If seepage occurs, it either evaporates completely, forming minerals and salts, or it is partially evaporated and the evaporative concentration effect is accommodated in the chemical modeling approach (CRWMS M&O 2000cg, Section 6.5).

To evaluate the effects of microbial activity on the bulk chemical environment while taking into account the importance of engineered material degradation, an approach is used that incorporates energy balance and mass balance. Threshold conditions determine when microbial growth and activity can resume, then degradation rates for the materials present in the in-drift environment determine the biomass that can be supported. The result of this calculation can be compared with other descriptors of the in-drift environment to assess whether microbial effects are significant (CRWMS M&O 2000cg, Section 6.4).

The same threshold conditions apply to the onset of microbially influenced corrosion of the waste package. While the drip shield and waste package materials are included in the energy- and mass-balance approach, the corrosion rates are developed directly from laboratory tests (see Section 4.2.4).

4.2.3.2.3 Effects of Engineered Materials on the Chemical Environment

Degradation of Steel and Alloys—The product of drip shield corrosion is predominantly titanium dioxide. Alloy 22 corrosion produces oxides of major components nickel, chromium, iron, and molybdenum. These oxides tend to be chemically inert and do not react with other chemical species in the drift environment. They will accumulate slowly, corresponding to general corrosion rates on the order of 10 to 1,000 nm/yr (0.0000004 to 0.00004 in./yr) (see
Section 4.2.4). This will consume oxygen at a rate that is negligible compared with the rate of gas flux through the drift openings (CRWMS M&O 2000cg, Section 6.3).

Steel corrosion products are insoluble ferric-oxides or oxyhydroxides. The concentration of ferric iron for aqueous solutions in equilibrium with hematite, goethite, and other iron oxides is very small, comparable to the concentration associated with iron-bearing nontronite clays (CRWMS M&O 2000cg, Section 6.7).

The rate of degradation for structural carbon steel, for vapor-phase conditions representative of the drift environment, has been measured for many samples, at different temperatures, in close proximity to different solutions conditions, such as pH and chloride concentration (CRWMS M&O 2000cg, Section 6.3). The measured penetration rates are on the order of tens to hundreds of microns per year. The corresponding consumption rates for oxygen are comparable to the convective flux of oxygen through the drift openings during the time period when this corrosion will occur.

Corrosion products can increase in volume, and particles can move in the drifts, potentially changing flow characteristics. These effects are neglected because the liquid flux in the drifts will be orders of magnitude smaller than the flow capacity, and redistribution of particulate matter can therefore have only a minor impact on the flow.

Effects of Cementitious Materials—A modified Type-K (Portland-based) expansive cement would be used for rock bolt anchorage in the potential repository. The mix will contain silica fume, plasticizer, and a low water-cement ratio to promote strength, durability, and low permeability (10-19 m2).

The alkaline composition of leachate from rock bolt cement grout will be bounded by equilibrium with portlandite. This approximation is conservative because portlandite will be carbonated over tens or hundreds of years from exposure to carbon dioxide in the gas phase. This behavior is known to occur with long-term exposure of concrete to air (CRWMS M&O 2000cg, Section 6.3).

Colloidal Particles Produced by Degradation of Engineered Materials—The following paragraphs describe laboratory data and field analogue data that are used to estimate the importance of ferric colloids for radionuclide transport.

Naturally percolating, mobile fracture waters have not been intercepted or sampled in the host rock, so groundwater analogues have been used to estimate colloid concentrations in seepage water. Colloids have been sampled by pumping of groundwater from 18 well intervals, from the saturated zone, at or near Yucca Mountain. The concentrations of particles in different size ranges were determined by instrumental analysis (CRWMS M&O 2000cg, Section 6.6). It is assumed that these results are analogous to unsaturated zone fracture waters in the potential repository host rock because the waters are derived from broadly similar rock types and likely have similar compositions. Accordingly, colloids are probably of similar types, consisting mainly of clays and silica.

It is noted that for more concentrated waters with greater ionic strength, such as will be produced by evaporative concentration during the thermal pulse, colloid concentrations are suppressed (Section 4.2.6). As the solution concentration increases, colloid stability decreases (i.e., the maximum possible concentration of colloids, which may vary for different sizes).

Production of colloids from engineered materials will be dominated by steel corrosion products, as discussed previously. The maximum concentration of ferric colloids in seepage waters can be estimated using the groundwater analogue. The affinity of these colloids for radionuclides has been measured in the laboratory (CRWMS M&O 2000cg, Section 6.6). Ferric-oxide colloids were shown to have high affinity for plutonium and americium.

Application to Modeling the Effects of Engineered Materials—The effects of steel on oxygen consumption in the drifts, and radionuclide transport, are modeled using the available information. Consumption of oxygen by steel is modeled by applying the average of measured corrosion rates, subject to a condition that the onset of steel corrosion occurs when the relative humidity is 70 percent. Consumption of oxygen by titanium and Alloy 22 corrosion, and the effects of cementitious materials on the bulk chemical environment, are not included in performance analyses, based on the previous discussion.

Use of groundwater colloids as an analogue for the concentration of ferric-oxide colloids in the drifts is conservative because pumping wells are highly dynamic, and not all colloids in the drift will be ferric-oxide or derived from steel. Also, different types of colloids become unstable in certain pH ranges (point of zero charge), and ferric-oxide colloids are unstable near pH 8, which is very close to the predicted pH of seepage. Use of laboratory test data for radionuclide affinity is also a conservative approach because smaller distribution coefficients would probably be observed in chemical systems with more components, containing anions and cations that could compete with radionuclides for sorption sites.

It is noted that the majority of ferric-oxide corrosion products in the drifts will be immobile in the drift environment and yet exposed to seepage water so that released radionuclides can be immobilized by sorption. The potential for retardation by corrosion products in the drifts was not incorporated into the TSPA-SR model but has been considered in supplemental studies. This supplemental evaluation is described further in Section 4.2.7.

4.2.3.2.4 Environment on the Surfaces of the Drip Shield and Waste Package

Behavior of Water on the Barrier Surfaces—As discussed previously, the deliquescence point defines the minimum relative humidity at which a salt will absorb water from the atmosphere and form a concentrated brine.
Figure 4-75 shows deliquescence points for a number of pure salts as functions of temperature. Among the salts which have been identified from evaporation of representative waters, sodium nitrate has the lowest deliquescence point (50 percent). Humidity corrosion of the drip shield and waste package can therefore begin at a relative humidity of 50 percent. This is conservative because deliquescence apparently occurs at greater humidity for mixtures of salts. For example, the deliquescence point for a mixture of salts evaporated from J-13 well water can be inferred from the observed boiling point of a concentrated brine representing the result of evaporating a bicarbonate/carbonate solution. The boiling point for such a brine is approximately 112°C (234°F) (CRWMS M&O 2000ck, Section 6.12.5), which is less than the boiling point of a saturated sodium nitrate solution (Figure 4-75). This means that humidity corrosion of the drip shield or waste package is not actually likely to occur until relative humidity reaches a value greater than 50 percent.

Potential for Acidic Conditions—Inorganic acids could form in small quantities from chemical evolution of brines, or they could be produced in conjunction with localized corrosion processes. These acids, such as hydrochloric acid, have known vapor pressures and will tend to evaporate over time, especially at elevated temperature (CRWMS M&O 2000ck, Section 6.10).

Application to Modeling the Environment on Barrier Surfaces—Consideration of the environment on the drip shield and waste package surfaces is limited mainly to identifying the value of relative humidity at which corrosion will begin. Humidity corrosion begins at 50 percent relative humidity, which is conservative, as discussed previously. The relative humidity is determined from thermal-hydrologic predictions that represent the metallic barriers explicitly.

Once corrosion begins, the water composition is estimated using the approach described previously for the in-drift chemical environment. Possible water compositions range from a concentrated sodium nitrate brine to dilute bicarbonate-type or chloride-sulfate-type waters. The corrosion rates used with these water compositions in the TSPA-SR model are based on laboratory-measured corrosion data for several different water compositions ranging from dilute bicarbonate-type water to concentrated brine, including elevated temperature conditions and both acidified and basified compositions (CRWMS M&O 2000ck, Section 6.12). The selection of these water compositions for corrosion test conditions is central to the approach for representing environmental conditions on the surface of the drip shield and waste package.

Considering the factors limiting production and deposition of inorganic acids and hydrogen peroxide, radiolysis outside the waste package is considered to be of minor importance to corrosion and is not considered in the performance analyses.

4.2.3.2.5 Rockfall on the Drip Shield

The geotechnical parameters required to predict rockfall include data and information collected either by field mapping or by laboratory testing. Joint mapping data for the subunits that constitute the emplacement horizon of the potential repository have been collected from the Exploratory Studies Facility, including the ECRB Cross-Drift (
CRWMS M&O 2000e, Section 4.1). Joint strength parameters, including cohesion and friction angle, have been developed from laboratory shear strength test data from core samples and are used to predict both the size and number of fallen rock blocks (CRWMS M&O 2000e, Section 4.1). Rock density data and intact rock elastic properties are used to assess seismic effects and to determine the load applied to the drip shield; these data have also been obtained from laboratory tests performed on core samples (CRWMS M&O 2000e, Section 4.1).

Key block analysis in underground excavations located in jointed rock masses has been considered for a number of design situations. Deterministic methods of block theory in rock engineering were advanced by Warburton (1981) and Goodman and Shi (1985). The literature provides examples for deterministic analysis of the maximum block size, given the spacing and orientation of three joint sets, and the excavation size and orientation. Subsequent work by other authors has been oriented toward probabilistic risk assessment of key block failure (CRWMS M&O 2000e, Section 6.3). These more recent methods are considered suitable for the analysis of densely jointed and faulted rock masses (i.e., greater than three joint sets) where planar joint surfaces can reasonably be assumed. Such conditions typically exist in the potential host rock units at Yucca Mountain. The probabilistic approach used in the TSPA-SR model is distinguished from traditional key block analysis because it not only assesses the maximum size of key blocks, but it also predicts the number of potential key blocks that will be formed in a certain length of tunnel for any tunnel orientation.

4.2.3.3 Process Model Development and Integration

The discussion of the physical and chemical environment is divided into five parts: (1) the chemical composition of seepage water and gas flux into the emplacement drifts (these are near-field conditions and are bounding conditions to the chemical and physical environment for the engineered barriers); (2) the chemical environment in emplacement drifts; (3) the effects of engineered materials on the chemical environment; (4) the chemical environment on the surfaces of the drip shield and waste package; and (5) the model for the rockfall on the drip shield.

4.2.3.3.1 Modeling the Composition of Liquid and Gas Entering the Drifts

Thermal-Hydrologic-Chemical Seepage Model Approach—This section discusses implementation of the thermal-hydrologic-chemical seepage model. The unsaturated zone flow model and the drift seepage model, which are the bases for thermal-hydrologic modeling, are discussed in
Section 4.2.1. This model predicts, at the drift scale, the composition of seepage and the associated gas-phase chemistry for 100,000 years, including the effects of heating.

The effects of heating on gas composition are important for predicting the composition of incoming seepage. The concentration of carbon dioxide in the immediate vicinity of the drift openings will decrease because of displacement and dilution by water vapor. This will be accompanied by decreased carbon dioxide activity in any associated seepage, which will cause pH to increase. In the zone of condensation further from the drift openings, carbon dioxide enrichment will occur, causing pH to decrease. Diffusivity for gaseous species is much greater than for aqueous species, and transport is more rapid. The result is that the region affected by changes in gas composition will be larger than that affected by changes in liquid-phase transport.

Flow of information from various models and data sources to the thermal-hydrologic-chemical seepage model is shown in Figure 4-76. The model uses input from modeling of the Drift Scale Test, the unsaturated zone flow model (CRWMS M&O 2000c), and other sources of geochemical data (CRWMS M&O 2000al, Section 3.3). These inputs ensure consistency between the thermal-hydrologic-chemical seepage model and the other models and data used to calculate drift seepage and the movement of water in response to heating. Other model inputs, including reactive surface areas of minerals, fracture–matrix interaction area, and the mineral volume fractions, are estimated from observed data (CRWMS M&O 2000al, Section 3.2.2).

Comparison with observations from the Drift Scale Test (CRWMS M&O 2000ch) and sensitivity studies on mineral assemblages and water compositions were used to guide development of the model. An example of the comparison of model results and observed data is shown in Figure 4-77. Evolution of the concentration of carbon dioxide in the gas-phase, over time, is compared with model results at four sampling intervals in the Drift Scale Test. For some data, calculated results are shown for nearby points in the model grid. For example, locations labeled "above" or "below" are calculated somewhat above or below the sampled location in the test. Similarly, locations labeled "center" or "end" were calculated near the center or the far end, respectively, of the sampled location. There are comparable trends between the test data and the calculations, and agreement is obtained in an average sense throughout much of the time period modeled.

Thermal-Hydrologic-Chemical Seepage Model Results—The model incorporates elements of the repository thermal operating mode nominally described in Section 2 to represent waste package heating over time, changes in heating from ventilation, heat transfer within the drift, and thermal-hydrologic-chemical processes (CRWMS M&O 2000al). As an example of model results, time profiles for gas-phase carbon dioxide concentrations at three locations around the drift, for one simulation using chloride-sulfate-type water, are shown in Figure 4-78. Carbon dioxide concentrations in fractures decrease significantly during dryout and increase again during rewetting. The increase during rewetting is caused by the dissolution of calcite deposited during dryout and by heating of ambient percolation and condensate waters as they approach the drift opening (the solubility of carbon dioxide in water decreases as temperature increases).

Using the chloride-sulfate-type water as the initial condition, the chloride concentration in fracture water immediately above the drifts is estimated to increase approximately fourfold from evaporative concentration. Higher concentrations could occur below the drifts. Upon rewetting, chloride concentrations are estimated to decrease to near-ambient values. The results indicate that seepage water should not be concentrated more than approximately one order of magnitude in chloride, compared to the ambient pore water. For the same simulation, the estimated pH for waters reaching the drift wall ranges from near-neutral (pH 7.2 to 8.3) to sub-alkaline (pH 8.6 to 9.0), depending on the chemical system properties used in the calculation, such as mineral assemblage and gas-phase carbon dioxide (CRWMS M&O 2000al, Section 3.3; CRWMS M&O 2000cl).

4.2.3.3.2 Modeling the Evolution of the In-Drift Chemical Environment

Three complementary modeling approaches are used to describe the formation and behavior of salts under different relative humidity conditions:

  • Normative Precipitates and Salts Model—Based on empirical data, this model is used to estimate which minerals and salts will form when waters are completely evaporated.

  • Low-Relative-Humidity Salts Model—The low-relative-humidity salts model describes the behavior of salts when relative humidity is less than 85 percent.

  • High-Relative-Humidity Salts Model—The high-relative-humidity salts model calculates water composition and dissolution/precipitation reactions for solutions that can occur when relative humidity is greater than 85 percent.

These models are used to determine the timing of in-drift environmental conditions that permit aqueous corrosion and to estimate the composition of waters that can transport released radionuclides.

Normative Precipitates and Salts Model—For waters having compositions similar to either bicarbonate-type water or chloride-sulfate-type water, a set of precipitates is identified that is consistent with the laboratory test data discussed previously. This model describes an assemblage of precipitates formed by complete evaporation. It is approximate in that the laboratory tests may not exactly duplicate evaporation conditions in the potential repository. It is supported by arguments based on thermodynamic data, which indicate that:

  • Thermonatrite and calcite are favored over hydroxides such as portlandite because of the presence of sufficient carbon dioxide in the environment.

  • Anhydrite is favored to form over gypsum (both forms of calcium sulfate) for the humidity conditions that will be present in the drifts.

  • Thenardite (sodium sulfate) and calcite are a more stable assemblage than thermonatrite and anhydrite.

The final species considered in the assemblage is amorphous silica (silicon dioxide), which was detected in all the laboratory samples derived from the bicarbonate-type water. The presence of the tuff would allow for more geochemical processes, such as formation of clays, cation exchange, and silicate buffering associated with tuff dissolution (
CRWMS M&O 2000cg, Section 6.5).

Low-Relative-Humidity Salts Model—This model begins at a point in time during the thermal period when the emplacement drifts are dry and at low humidity. Incoming seepage (if it occurs) is completely evaporated, and the salts and minerals that form are determined from the normative approach. As relative humidity continues to increase over time, the salts are allowed to dissolve according to their deliquescent behavior. The amount and composition of brine produced are controlled by the solubilities of the salts and the fraction of each salt that is allowed to dissolve. For the nitrate salts, the entire amount is allowed to dissolve when the relative humidity reaches 50 percent. For the remaining salts, the fraction dissolved is abstracted from the individual salt properties. For simplicity in performance assessment analyses, all of the remaining salts are modeled to dissolve exponentially from zero to unity as relative humidity increases from 50 percent to 85 percent. The timing of relative humidity evolution at different locations in the potential repository is obtained from thermal-hydrologic calculations (Section 4.2.2). At 85 percent relative humidity, all the accumulated soluble salts are considered to be completely dissolved.

High-Relative-Humidity Salts Model—The high-relative-humidity salts model is used for relative humidity greater than 85 percent, conditions for which soluble salts are fully dissolved and the relative rates of evaporation and seepage control the aqueous chemistry. Modeling the behavior of concentrated aqueous solutions (greater than 1 molal ionic strength) is performed using a Pitzer approach, implemented using the EQ3/6 chemical modeling code (CRWMS M&O 2000cl, Section 6.4.2). A modified Pitzer database was developed from existing published data, adding chemical species such as nitrate and silica, and extending the temperature range of applicability to 95°C (203°F). Details are provided in supporting documentation (CRWMS M&O 2000cl, Section 5.3). It is a conservative, approximate model that is used for predicting the composition of water in the emplacement drifts for a time interval during the thermal period when seepage is strongly concentrated by evaporation.

Three sources of experimental data were used for validation of the model (but not for model development or calibration; CRWMS M&O 2000cl, Section 4.1.2). All three studies involved evaporation of synthetic J-13 well water in a beaker open to the atmosphere and maintained at constant elevated temperature. In one study, 30 L (7.9 gal) of synthetic average J-13 water were evaporated to 30 mL and the precipitated solids analyzed. In a second study, the pH of the water was monitored during evaporation, and, in a third, 100-times concentrated average J-13 well water was dripped through a column of heated tuff and final solution composition compared to the initial composition (CRWMS M&O 2000cl, Section 4.1.2).

Results from the high-relative-humidity salts model show that sodium, potassium, chloride, sulfate, and carbonate species simply concentrate without much precipitation. As J-13 water is concentrated to a total ionic content of 10 molal, the pH approaches 10. Species containing silicon, aluminum, calcium, magnesium, iron, and fluoride tend to precipitate in significant amounts relative to their aqueous concentrations. Among the precipitates included in the model, calcite and chalcedony are produced in greatest quantity; magnesium-bearing clay (represented in the model by sepiolite) and fluorite are the next most abundant.

Model for Microbial Activity in the Emplacement Drifts—The approach to assessing the potential effects from microbial activity has two parts: (1) a threshold model for environmental conditions that permit microbial growth and activity and (2) a quantitative model to bound the quantity of biomass, including microbes, that could develop in the emplacement drifts.

Relative humidity, temperature, and radiation dose conditions are combined to formulate threshold environmental conditions for microbial activity. Published results based on extreme behaviors of known organisms are used as a guide where site-specific data are sparse. More precise estimates based on characterization of organisms at Yucca Mountain are unwarranted because of the uncertainty about which types of organisms will be present and the potential for biological adaptation.

Microbial growth and activity does not occur until the local temperature decreases to less than 120°C (248°F) and relative humidity increases to 90 percent. Microbially influenced corrosion of the waste package is conservatively assumed to begin when the relative humidity reaches 90 percent.

The microbial communities model quantifies the abundance and metabolic activity of microorganisms in the engineered barrier system environment. It is based on models used in the Swiss and Canadian nuclear waste programs (CRWMS M&O 2000cp, Section 6). An idealized elemental composition for microbial biomass is used, consisting of carbon, nitrogen, sulfur, and phosphorous in fixed proportions, plus water. The rates of supply for these constituents are input as constant release rates for each natural and engineered material present in the drifts. The other major constraint is the chemical energy available for microbes to grow, which is maximized from possible oxidation/reduction reactions.

The presence of water, nutrients, and energy sources is required for microbial growth and activity. There are three main categories of introduced materials that could furnish nutrients and energy: steels/alloys, cementitious materials, and organic substances (CRWMS M&O 2000as, Section 3.1.2.4.2.1).

Three basic approaches to modeling microbial nutrient and energy balances are possible, based on nutrient balance, thermodynamic energy balance, and chemical kinetics. The microbial communities model (CRWMS M&O 2000cp, Section 6) combines the nutrient and thermodynamic approaches. The available materials are decomposed into their basic elements and combined with the constituents available from groundwater and gas fluxes through the repository. Release of chemical constituents is controlled by estimating the degradation lifetime of each material. The model assumes that all available nutrients and redox energy sources are used for microbial processes and is therefore bounding.

Application of the microbial communities model to the potential repository design shows that approximately 10 grams of biomass would be produced per lineal meter of emplacement drift per year, during the first 10,000 years. Based on this small generation rate, effects on the bulk chemical environment are considered negligible for the TSPA-SR analyses (CRWMS M&O 2000as, Section 5.3.2.4). Localized effects of microbial activity could nevertheless alter the longevity of materials and the transport of radionuclides. For that reason, microbially influenced corrosion of the waste packages was included in TSPA-SR calculations (CRWMS M&O 2000a, Section 3.4.1.6).

4.2.3.3.3 Modeling the Effects of Engineered Materials on the Chemical Environment

Modeling the Effects from Corrosion of Steel and Alloys—Steels and alloys introduced to the potential repository as ground support and other structural materials will degrade with time. The important effects of metal corrosion on the bulk chemical environment, and the approaches used to represent those effects, are summarized as follows.

Oxygen in the drift environment will be consumed, decreasing the partial pressure of oxygen while corrosion is active. This effect is modeled by converting rates of metal corrosion to rates of oxygen consumption and comparing the results with estimates of the available oxygen flux (
CRWMS M&O 2000as, Section 3.1.2.3.4).

A corrosion-rate model for structural steel was used to develop estimates of oxygen availability. The potential consumption of oxygen from corrosion of other alloys is also discussed briefly. Corrosion rates for A516 carbon steel have been measured for representative vapor-phase conditions (not immersion), elevated temperature, and proximity to synthetic groundwater with a composition similar to evaporatively concentrated J-13 water. These data are used to represent corrosion of structural steel in the drifts, as there is a paucity of equivalent data for other steels. The rate can vary with temperature, pH, and water composition. In addition, laboratory data for a similar carbon steel composition (CRWMS M&O 2000cg, Section 6.4) suggest an approximate sixfold increase in the corrosion rate once moisture returns to the drifts because of microbial activity. Steel corrosion will be insignificant until the relative humidity exceeds 70 percent (CRWMS M&O 2000cg, Section 6.3.2.2).

Applying measured corrosion rates for structural steel and using environmental conditions representing the potential repository environment yields the following results (CRWMS M&O 2000cg, Section 6.3):

  • Steel present in the drifts will completely corrode within a few hundred years, starting at times from approximately 300 to 2,000 years after waste emplacement. (Timing will depend on the duration of preclosure operations and location in the repository layout; steel corrosion can begin at 300 years after emplacement if the preclosure period is 50 years, and corrosion first occurs near the repository edge, which cools fastest.)

  • The calculated rate of steady-state oxygen consumption will be replenished by the flux of oxygen that is transported to the drifts by buoyant gas-phase convection.

Results indicate that oxygen partial pressure in the drifts would be decreased but that the flux of oxygen would probably be sufficient to maintain oxic (i.e., corrosive) conditions. In the potential repository, the corrosion rate for structural steel will tend to decrease as the oxygen partial pressure decreases, prolonging the degradation process but moderating the impact on oxygen partial pressure. This model is therefore conservative because the laboratory corrosion data on which the rate model is based were acquired for oxic conditions (CRWMS M&O 2000as, Section 3.1.2.3.4).

The drip shield, waste package, and waste package supports in the design described in this report would be made from titanium, Alloy 22, and stainless steel. Corrosion of these materials will be much slower than for carbon steel and therefore slower to consume oxygen. Consumption of oxygen by corrosion of Grade 7 titanium or Alloy 22 could contribute slightly to depletion of oxygen in the engineered barrier system. However, the potential rate of consumption is a small fraction of the oxygen availability (CRWMS M&O 2000cg, Section 6.3).

Modeling the Effects from Cementitious Materials—The model for the effects of cement grout used in rock bolts is based on dissolution of cement mineral phases and subsequent interaction of cement leachate with carbon dioxide in the drift environment. The results indicate that the potential contribution of cement leachate to the bulk chemical environment will be minor (CRWMS M&O 2000cg, Sections 6.3 and 6.7).

Chemical equilibrium calculations for cement are performed using a mineral assemblage that includes constituents of young cement as a bound on the potential to produce alkaline leachate (CRWMS M&O 2000cg, Section 6.3). Mineral assemblages representing alteration of Portland cement by aging and carbonation are available but tend to produce less alkaline leachate.

The initial composition of water that interacts with cement is assumed to be that of J-13 water, but the potential contribution of influent water chemistry to the leachate is minor because leachate composition (before reaction with carbon dioxide) will be dominated by the cement. Reaction of groundwater with the grout is assumed to be closed (i.e., it does not include reactions with gas-phase carbon dioxide), which maximizes the leachate pH. Biotic processes are also assumed to have a negligible effect on leachate composition because the organic content of the plasticizer admixture probably has very low biological reaction rates (CRWMS M&O 2000cg, Section 6.3).

Once the leachate flows into the drift opening or the surrounding rock, it will be exposed to carbon dioxide in the emplacement drifts, which will moderate the pH. Because of the small amount of leachate compared to the total percolation through the host rock, open-system conditions are used for evaluating this interaction. In addition, contact with silica, such as the cristobalite present in the host rock, will also moderate leachate pH.

The grout permeability is small, which limits chemical interaction with the drift environment while increasing the longevity of the grout to dissolution. Very small flow rates (a few milliliters per year per rock bolt) are obtained using the saturated hydraulic conductivity of the grout. A more conservative bounding approach is used to account for the possibility of grout cracks, allowing higher water contact rates. This approach is based on the ratio of the rock bolt grout cross-sectional area to the drift diameter. This method produces flow rates on the order of a few percent of the total seepage inflow to the drifts.

Results from this model show that prior to contact with carbon dioxide or siliceous minerals, the leachate will be highly alkaline, with a pH value of 11 or greater. This is caused primarily by dissolution of portlandite, which has retrograde solubility (higher solubility at lower temperatures). Buffering on contact with carbon dioxide and near-field rock in the drift environment will result in substantial decrease in leachate pH. Accordingly, elevated pH values from the leachate are unlikely to be present at the drip shield or waste package.

The model does not consider the possibility that grout could break up into fragments, some of which may fall and come to rest on the drip shield. A transient pulse of alkaline leachate could result from such failure, and the current model does not explicitly address rapid leaching. However, the analysis shows that if such leaching occurs during the thermal pulse, the chemical composition of the leachate will be similar to evaporatively concentrated waters that will form in the drift. During and after cooldown, the concentration of carbon dioxide will increase in the drift environment, and cement fragments will be exposed to increased carbon dioxide levels. The potential for strongly alkaline conditions will then be moderated by reaction of leachate with the drift environment. Also, carbon dioxide gas will diffuse into the grout and react directly with the cement. This process is not considered in the current model but will be more important for fragmented grout. Finally, the current model indicates that the quantities of leachate produced are likely to be small, whether or not the grout is fragmented.

Modeling the Effects of Colloids Produced from Engineered Materials—Ferric-oxide colloids with strong affinity for plutonium and americium will be produced from corrosion of steel in the emplacement drifts. The size and concentration distributions of natural colloids observed in local groundwaters are used to represent the behavior of ferric-oxide colloids. This is justified because these waters are similar in composition to seepage waters that are likely in the potential repository (CRWMS M&O 2000al, Section 3.3.2.4). It is conservative because the effect of ionic strength in evaporatively concentrated waters, and the behavior of ferric-oxide at near-neutral pH, will decrease the stability of ferric-oxide colloids. Colloids are discussed further in Section 4.2.7 in the context of the radionuclide transport model.

4.2.3.3.4 Modeling the Environment on the Surfaces of the Drip Shield and Waste Package

Conditions for Aqueous Corrosion and Humidity Corrosion—At permanent closure, the surfaces of the drip shield and waste package will be dry, with relative humidity less than 50 percent, and no humidity corrosion or aqueous corrosion processes active. This condition will continue through the period of peak temperatures. During cooldown, the relative humidity on the barrier surfaces will gradually increase (
CRWMS M&O 2000n, Section 3.1.3).

Figure 4-75 shows that aqueous salt solutions on the drip shield and waste package can exist for relative humidity less than 100 percent but greater than the deliquescence point. Analysis of salts formed from laboratory evaporation tests shows that the salt with the lowest deliquescence point is sodium nitrate (CRWMS M&O 2000n, Section 3.1.3.1).

For modeling purposes, it is assumed that sodium nitrate—the salt with the lowest deliquescence point among all the salts likely to be present—will determine the minimum relative humidity at which aqueous conditions can occur (50 percent). Another value of relative humidity is selected to represent the condition at which all salts, including chlorides, sulfates, and carbonates, are dissolved as brine (85 percent). This value is greater than the deliquescence points for all salts that are likely to be present. The low-relative-humidity salts model is used to predict the composition of brines present on the drip shield and waste package surfaces as the relative humidity increases from 50 to 85 percent. This is a conservative approach because waste packages under intact drip shields will be affected only by dust and aerosols, which may be more benign than the brine conditions considered in the model.

The high-relative-humidity salts model is used to calculate the composition of brines as they undergo further dilution at relative humidity greater than 85 percent. If seepage occurs during the thermal pulse, the waters that contact the drip shield and waste package surfaces will be evaporatively concentrated because the relative humidity on the surfaces will be less than in the surroundings. The high-relative-humidity salts model is used to calculate the composition of evaporatively concentrated seepage waters. The degree of evaporative concentration will decrease with time as the relative humidity increases.

The high-relative-humidity submodel of the in-drift precipitates and salts analysis was implemented for relative humidity greater than 85 percent. In this regime, the steady-state water composition is controlled by the ratio of the evaporation rate to the seepage rate; this ratio is always less than one. The submodel calculates water composition, pH, chloride molality, and ionic strength in the repository for several temperatures, relative humidities, relative evaporation rates, and carbon dioxide gas fugacity values (CRWMS M&O 2000cl, Section 6.4.2).

Composition of Aqueous Solutions Used for Corrosion Testing—The exact chemistry of the water that contacts the drip shield and waste package surfaces cannot be known precisely. However, several test solutions have been developed for laboratory corrosion testing of titanium, Alloy 22, and other materials. These solutions were selected to represent a range of dilute and concentrated conditions, pH, and temperatures that could result from evaporative concentration in the repository. The test solutions are described in Section 4.2.4.2.

The chemistry of the waters that contact drip shield and waste package surfaces would vary in a repository, and the test solutions described previously represent a range of possible conditions. Results from corrosion testing with these solutions are discussed in Section 4.2.4.

4.2.3.3.5 Model for Rockfall onto the Drip Shield

The effect of rockfall on the emplaced drip shield was screened from the TSPA-SR model based on the following analysis. Drift Degradation Analysiss the analysis of key blocks, including block failure due to seismic and thermal effects (
CRWMS M&O 2000e, Section 6), using the probabilistic Discrete Region Key Block Analysis method (CRWMS M&O 2000a, Section 3.3.1; CRWMS M&O 2000e, Attachments VIII through XI).

Based on mapping data and results from key block analysis, the orientation of the emplacement drifts, as discussed in Section 2.1.2, was selected to maximize drift stability (CRWMS M&O 2000e, Sections 1.4 and 6.4). Considering static plus seismic loads, rockfalls were estimated using the probabilistic key-block analysis. The results include probability distributions for the size of fallen key blocks and the number of such blocks per length of emplacement drifts.

Heating and subsequent cooling of the repository host rock would impose stresses, deposit minerals, and shift rock joints. This was represented in the analysis as reduction in joint cohesion from the time of closure until about 2,000 years after closure and thereafter. The resulting changes in estimated rockfall are provided in supporting documentation (CRWMS M&O 2000e, Section 6). Results for combined thermal and static loads are similar to those for combined seismic and static loads.

Given the distribution of key block size, a structural calculation was performed to determine the stresses in the titanium drip shield (BSC 2001l). The analysis considered a range of rock sizes falling onto a 3-m (10-ft) length of drip shield. This length is half that of a drip shield segment and was chosen to take advantage of symmetry and thereby reduce computational effort. A range of rock sizes up to 10 metric tons (approximately 4.15 m3 [147 ft3]) is the design basis for the drip shield. For rock sizes up to 4 metric tons, the entire volume of a rock can be located above a 3-m (10-ft) partial-length drip shield. An analysis of the joint geometry suggests that an increase in rock size must occur by an increase in length of the rock block along the drift, rather than an increase in block height (CRWMS M&O 2000e, Attachment IX). The probabilistic key-block analysis showed that the maximum block size could be greater than 10 metric tons, but such large blocks would be relatively unlikely. Probabilistic key-block analysis was performed for the middle nonlithophysal, lower lithophysal, and lower nonlithophysal host rock units. For these units, the proportion of rock blocks simulated that exceeded 10 metric tons ranged from 0 to 5.8 percent (CRWMS M&O 2000e, Figures 27 to 29 and Attachment II). Also, because of their large dimensions, each block would be distributed over more than one 6-m (20-ft) drip shield segment, so the loads would be shared. Using the concept of effective rock mass over a 3-m (10-ft) partial-length drip shield, the maximum rock mass is determined to be 10 metric tons per 3-m (10-ft) partial-length drip shield. In other words, any rock mass greater than 10 metric tons will load a 3-m (10-ft) partial-length drip shield the same as a 10-metric-ton rock. Calculations for the effective rock mass for different size blocks were calculated using finite element techniques and are documented in Rock Fall on Drip Shield (BSC 2001l, Section 5.2 and Table 7-1). Preliminary results from the analysis indicated that the impact would not dent the drip shield in such a way that it would contact the waste package. (Design criteria are presented in Emplacement Drift System Description Document [CRWMS M&O 2000ab].) The impact from a 10-metric-ton block was estimated to produce one 13-cm (5-in.) long crack in a 3-m (10-ft) long portion of a drip shield segment because of stress corrosion cracking that could occur after the rockfall (BSC 2001l, Table 6-1). Smaller rocks, such as a 2-metric-ton rock, were estimated not to produce enough residual stress to initiate stress corrosion cracking. It is expected that such stress corrosion cracks would be narrow and would not conduct much water flow (see Section 4.2.4). The calculation is conservative because it does not take credit for energy expended by fracturing of the rock or dislocation of the drip shield.

Effects from rockfall events involving multiple rock blocks have not been analyzed explicitly. However, because the dimensions of the maximum expected block size are of the order of the dimensions of the waste packages, it is reasonable to assume that multiple falls of blocks of the maximum size would have to act independently. The effects from multiple, smaller-size block falls would be bounded by the analyses of the effects of the maximum block size.

4.2.3.3.6 Limitations and Uncertainties

As discussed in
Section 4.4.1.2 of this report, uncertainties are an inherent component of the TSPA method. Uncertainty is introduced through the conceptual model selected to characterize a process, as well as the mathematical, numerical, and computational approaches used to implement the model. Uncertainty is also introduced from imperfect knowledge of important parameters used for input to the models (e.g., physical properties). This section emphasizes limitations and uncertainties relative to the in-drift physical and chemical environment: (1) the thermal-hydrologic-chemical model used to characterize seepage water chemistry that enters the drift, (2) the in-drift chemical environment model and supporting analyses, and (3) the rockfall model and supporting analyses.

Since the TSPA-SR model, the DOE has performed several supplemental activities to address uncertainties and limitations in the TSPA-SR model. Additionally, as noted in Section 4.1.4, the DOE is evaluating the possibility for mitigating uncertainties in modeling long-term repository performance by operating the design described in this report at lower temperatures. Consequently, some of the models describing the thermal-hydrologic-chemical model, precipitates salts model, and the drift degradation model have undergone further evaluation since the TSPA-SR model. Some alternative models have been implemented and sensitivity analyses conducted to address parameter and model uncertainties. These supplemental analyses are summarized in FY01 Supplemental Science and Performance Analyses (BSC 2001a, Sections 4, 5 and 6). The sensitivity of TSPA model results to alternative process models is discussed in Volume 2, Sections 3.2 and 4.2 of FY01 Supplemental Science and Performance Analyses (BSC 2001b) and is summarized in Section 4.4.5.5 of this report.

Thermal-Hydrologic-Chemical Seepage Model Limitations and Uncertainties—Uncertainties exist in the thermodynamic and kinetic input data used in the model. For example, as the concentrations of dissolved solutes increase during evaporation, the theoretical limitations of the chemical activity model may be exceeded. Rapid boiling can lead to mineral precipitation that is controlled by nucleation kinetics and other surface-related phenomena. The specific conditions for which the model becomes invalid, however, may not be important for the overall dynamics of the system. Geochemical reactions are a strong function of temperature and the presence of water, both of which are better constrained than the rates of reaction and may control the spatial distribution even if the exact quantities of the phases at a given time are uncertain (CRWMS M&O 2000al, Section 3.3.4).

The model is based on information obtained from thermal testing in the middle nonlithophysal unit of the Topopah Spring Tuff, which may exhibit somewhat different thermal-hydrologic behavior or mineralogy than lower lithophysal units, in which most of the repository emplacement drifts would be located. Although this could change the system response, it is noted that the bulk chemistry is similar for all the welded host rock units (CRWMS M&O 2000al, Section 3.3.4). Thermal testing that is planned for the lower lithophysal unit will evaluate the applicability of the current model over the repository footprint.

Another aspect of thermal-hydrologic response that will affect aqueous and gas-phase chemistry in the host rock is buoyant gas-phase convection (CRWMS M&O 2000cm), controlled by the large-scale permeability. As discussed previously, gas-phase convection will be an important source of carbon dioxide and oxygen as chemical reactants to the in-drift chemical environment. Buoyant gas-phase convection is not evident from the Drift Scale Test, probably because the rock bulk permeability at this location does not support it. However, results from the unsaturated zone flow model (Section 4.2.1) indicate that permeability of the host rock increases with scale, and current thermal-hydrologic models do not incorporate mass transfer along the axes of the drift openings. Therefore, it is likely that further investigation will show that gas-phase convection in the potential repository will exceed that observed in the Drift Scale Test. It is anticipated that the magnitude of gas phase convection will be great enough to supply sufficient amounts of gas through drift openings to preclude depletion of gas components through chemical reactions with drift components (CRWMS M&O 2000cg, Sections 6.3.2 and 6.3.3).

Section 4.2.2.3.5 summarized supplemental studies to the TSPA-SR model related to physical changes in permeability and seepage resulting from thermal-hydrologic-chemical processes. These studies also addressed thermal-hydrologic-chemical model uncertainties related to the chemistry of potential seepage water. Specific studies included:

  • Supplemental sensitivity studies of different initial water and gas boundary conditions (BSC 2001a, Section 4.3.6.5)

  • Simulated evolution of water and gas compositions in the Tptpmn and Tptpll units (BSC 2001a, Section 6.3.1.4.3)

  • Modified base-case and extended case geochemical systems (BSC 2001a, Sections 4.3.5.3.2, 4.3.6.4, and 6.3.1.5).

These activities included a range of different input data and assumptions, including thermodynamic and kinetic data input data (BSC 2001a, Sections 4.3.6.4 and 6.3.1.4).

Precipitates and Salts Model Limitations and Uncertainties—This modeling approach is based on literature data that describe the minerals and salts produced by evaporative concentration and the relations between solution composition and relative humidity. In certain aspects, the thermal-chemical data support is incomplete. Therefore, the accuracy of this model was tested using independent collected laboratory data for the evolution of solids and water composition during evaporation. As described in In-Drift Precipitates/Salts Analysis (CRWMS M&O 2000cl, Section 7.3), the precipitates and salts model is expected to provide results that are within an order of magnitude for chloride concentrations and ionic strength and within a pH unit for pH predictions. This degree of accuracy is acceptable because it greatly reduces the potential ranges of these variables, thereby considerably reducing uncertainty.

Data were not available to directly evaluate the accuracy of the low-relative-humidity salts model. The model did, however, produce reasonable trends and results in chloride and ionic strength outputs in its applied relative humidity range and produces consistent results at 85 percent relative humidity, where the high-relative-humidity salts model takes over (CRWMS M&O 2000cl, Section 7.3).

Simplifying assumptions were required to reduce the complexity of the precipitates and salts analysis and to avoid sophisticated approaches where data were lacking, although these assumptions tended to be conservative. The greatest uncertainties in the analysis are likely the thermal-hydrologic and thermal-hydrologic-chemical predictions and other predicted inputs to the analysis (CRWMS M&O 2000cl, Section 7.3). Therefore, the models were applied for a variety of these conditions.

The precipitates/salts model has been modified in several ways since the TSPA-SR model analyses. Each modification involved the high relative humidity submodel. The Pitzer database was improved by the addition of several minerals and thermodynamic data, and the amount of output data reported in lookup tables was increased (BSC 2001a, Section 6.3.3.4).

Additional discussion of uncertainties related to the precipitate and salts model is described in supplemental studies (BSC 2001a, Sections 6.3.3.3, 6.3.3.4, and 6.3.3.5). Supplemental studies emphasized selected uncertainties, including the sensitivity of starting water composition on evaporative chemical evolution and effects of mineral suppression (BSC 2001a, Section 6.3.3.5.1). Sensitivity studies show in-drift chemistry to be sensitive to starting water composition and that thermodynamic and kinetic data are secondary in importance or negligible (BSC 2001a, Section 6.3.3.5). These studies also show that using the water composition obtained from the thermal-hydrologic-chemical model is reasonable, as implemented in the TSPA-SR model and supplemental TSPA model analyses.

Microbial Communities Model Limitations and Uncertainties—This model is not intended to quantify localized microbial activity or its consequences but bounds overall microbial growth and activity in the engineered barrier system. It has been validated by testing against laboratory data, natural system observations, and other modeling efforts, as described in Section 6.6 of In-Drift Microbial Communities (CRWMS M&O 2000cp). Comparison with these independent data sets indicates that the estimates are within an order of magnitude of the actual values.

The approach uses a well-mixed reaction system to represent the emplacement drift and provides estimates of total biomass production that can be used to bound the extent of microbially influenced corrosion of the waste package. In this approach, one of the primary uncertainties is in the degradation rates that supply nutrients and energy for microbial growth. This is evaluated by varying the rates over large ranges and using bounding results (CRWMS M&O 2000cp, Section 6).

Engineered Materials Effects Modeling Limitations and Uncertainties—The rate of steel corrosion for sub-oxic conditions, including the effects of microbial activity, is an important factor in predicting the oxygen budget in the emplacement drifts. Current estimates for steel corrosion are based on (1) abiotic test results obtained for similar steel under oxic conditions and (2) preliminary data for microbially influenced corrosion of similar steel at water-saturated and nutrient-augmented culture conditions. The results for oxic conditions overestimate the rate of oxygen consumption for conditions of decreased oxygen availability, which could occur for a few hundred years in the repository during peak thermal conditions. Applicability of the available microbial testing data to environmental conditions in the repository is more uncertain (CRWMS M&O 2000cg, Section 6.3).

Boiling of water during the thermal event will produce water vapor that will displace much of the air from the drift, with the result that oxygen will be depleted. This may lead to decreased corrosion rates, but the extent of such decreases is uncertain for steel and corrosion-resistant materials such as titanium and Alloy 22. In any case, the effect of decreased oxygen availability will be of limited duration.

Supporting analyses to the TSPA model has previously demonstrated that including these uncertainties would increase the potential for reduced corrosion and solubility, which would have a beneficial impact on performance (BSC 2001a, Section 6.3.2.3.1).

Conceptualization of leachate that could be contributed from degradation of cement grout is adapted from simplifying assumptions and studies of concrete reported in the literature (CRWMS M&O 2000cg, Section 6.3.1; CRWMS M&O 2000as, Section 3.1.2.3.5). Confirmatory testing of rock bolt longevity, carbonation of the grout, microbial attack of organic superplasticizers, and other aspects of grouted rock bolt performance has not been undertaken. The final design of ground support in the emplacement drifts is under development, and analysis of ground support function is preliminary. Accordingly, the longevity of grouted rock bolts in response to thermal loading and environmental conditions remains somewhat uncertain. Rock-mass deformation could lead to failure of the rock bolts (not necessarily associated with failure of the drift openings).

Supporting the TSPA-SR model analyses, supplemental analyses have provided additional confirmation indicating cement leachate-influenced seepage water would have a negligible affect on in-drift chemistry (BSC 2001a, Section 6.3.2.3.2).

Limitations and Uncertainties on Conditions Used to Model Barrier Corrosion—Microbial activity on the waste package surface under an intact drip shield may be negligible, even above 90 percent relative humidity, as long as water has not leaked through the drip shield onto the waste package and the waste package has not directly contacted the invert. Thus, significant corrosion would not commence until the drip shield fails and water contacts the waste package. Similarly, if salt-tolerant organisms are present, the relative humidity threshold for microbially influenced corrosion may decrease to 75 percent, which would mean that the onset of microbially influenced corrosion could shift to an earlier time. Such a shift represents only a small fraction of the time that microbial processes are modeled to be present.

Furthermore, uncertainty as to the timing for the onset of general corrosion or microbially influenced corrosion is of low importance for performance assessment because the expected corrosion rate for the waste package, even with microbes present, is low enough that a shift in the timing of the onset of corrosion is insignificant to the overall performance of the potential repository. Further discussion of corrosion rates for the drip shield and waste package is provided in Section 4.2.4.

Common scale minerals that are produced from natural waters include calcite, gypsum, and silica (CRWMS M&O 2000ck, Section 6.1.1). If seepage occurs, the chemical components of these minerals will be present in waters that contact the drip shield and waste package. The tendency for scale formation is enhanced by evaporative concentration and changes in pH. Scale can protect the underlying material if it forms a dense adherent layer, but it may have a deleterious effect if it forms a porous layer or a crevice with the underlying material. Interfaces between scale-covered and bare regions of the surfaces might be subject to corrosion. Few data on the effects of scale are presently available, but related laboratory tests have been performed to evaluate the potential for crevice corrosion. The results show that crevice corrosion is insignificant for titanium and Alloy 22 at the expected repository temperature and humidity. By inference, the effects of scale are also thought to be insignificant.

Another potentially important aspect of scale formation is the production of decomposition products, such as lime, from scale minerals at elevated temperatures. Subsequent dissolution of such products by seepage water or deliquescence could produce highly alkaline solutions. Decomposition temperatures for potential scale minerals are known, and some are within the operating temperature limits for the waste package (CRWMS M&O 2000ck, Sections 4.1.11 and 6.11). However, as discussed previously, the presence of gaseous carbon dioxide in the drift environment implies that the formation of highly alkaline hydroxide species is unlikely.

Rockfall Model Limitations and Uncertainties—The usefulness of the rockfall model is affected by how well the data inputs describe the actual fracture conditions. The natural variability of fractures within a rock mass always represents uncertainty in the design of structures in rock. The extensive fracture data collected at Yucca Mountain provide a good representation of fracturing at the emplacement drift horizon. The range of fracture variability from tunnel mapping has been captured in the rockfall model through multiple Monte Carlo simulations of the rock mass. To account for uncertainties associated with seismic, thermal, and time-dependent effects on rockfall, a conservative reduction of joint strength parameters has been included in the approach (CRWMS M&O 2000e, Section 5).

Sensitivity analyses supplemental to the TSPA-SR model did not change the results of the rockfall model analyses (BSC 2001a, Section 6.3.4.9). Consequently, the process remains insignificant to performance and remains screened from TSPA analyses (BSC 2001a, Section 6.3.4.6).

4.2.3.3.7 Alternative Conceptual Processes

As with limitations and uncertainties, some of the following alternative conceptual models have been addressed in supplemental analyses and summarized in FY01 Supplemental Science and Performance Analyses (
BSC 2001a, Section 1; BSC 2001b, Section 1).

Alternative Concepts for Thermal-Hydrologic-Chemical Seepage Model—Alternative approaches for modeling the compositions of the gas-phase and liquid seepage have been evaluated in Engineered Barrier System: Physical and Chemical Environment Model (CRWMS M&O 2000cg, Sections 6.2 and 6.7). Both the bicarbonate-type and chloride-sulfate-type waters were evaluated for a range of carbon dioxide conditions. The results were used to evaluate the carbon dioxide budget, taking into account processes in the host rock and within the drifts, as discussed previously. The liquid and gas compositions obtained are comparable to results from the thermal-hydrologic-chemical seepage model, which were abstracted for the TSPA-SR model.

As noted in Section 4.2.2.3.5, supplemental studies have evaluated a number of additional alternative starting water compositions in sensitivity analyses. Additional water compositions included water perched on top of the Calico Hills formation, water collected from the Drift Scale and Single Heater Tests, and water predicted by the thermal-hydrologic-chemical model, as well as other seepage water compositions discussed in this section (BSC 2001a, Section 6.3.3.5). Alternative conceptual models described in Sections 4.2.2.3.5 and 4.2.3.3.6 include alternative initial water and gas compositions, alternative host rock assumptions, and modified base-case and extended case geochemical systems (BSC 2001a, Sections 4.3.5.3.2, 4.3.6.4, 4.3.6.5, 6.3.1.4.3, and 6.3.1.5). In general, the sensitivity studies suggest that the models incorporated in the TSPA-SR are reasonable or conservative.

Alternative Concepts for Precipitates and Salts Model—An alternative approach for representing the composition of waters contacting the drip shield and waste package would be a bounding concept. Considering that some dripping of water onto the drip shields may occur throughout the repository in small quantities because of condensation, small accumulations of minerals and salts, on the scale of individual droplets, could occur in locations where seepage will never occur. Microscopic quantities of salts would interact with changes in the relative humidity but would not be mobilized by flow or diluted by seepage. In this alternative bounding approach, the environment on the drip shield surface would be represented by the permanent presence of minerals and soluble salts or the solutions obtained when they equilibrate with water vapor in the air (CRWMS M&O 2000cg). Although simpler than the current precipitates and salts model used for TSPA-SR, this approach would not yield very different corrosion conditions. Current estimates of corrosion rates are small, with limited sensitivity to the presence of soluble salts (see Section 4.2.4).

An alternative model for seepage/invert interactions has been developed but has not yet been implemented in its entirety (BSC 2001a, Section 6.3.3.4.2.1). This alternative model abstracts in-drift mixed solutions using ionic strength, pH, and an acid neutralizing capacity parameter (which is an indication of the resistance of a solution to pH changes). The abstracted solutions are mixtures of seepage fluxes from the crown above the drift, the water wicked through the rock and corroded metals in the invert, and the diffusion film or flux from a waste package after failure.

Alternative Concepts for Microbial Communities Model—As stated previously, the implemented model bounds the overall production of biomass in the emplacement drifts. Two approaches, while not precisely alternatives to the microbial communities model, could be used to extend the model from estimation of biomass to estimation of bounding rates for microbially influenced corrosion. The first approach is an empirical approach based directly on microbial corrosion laboratory test data. The second approach, which is the principal alternative concept for predicting microbial effects on the bulk chemical environment, is the chemical kinetics approach mentioned previously. In this concept, biotic and abiotic chemical reactions are treated similarly, as quantitatively explicit reactions distinguished by different reaction rates. Reaction pathways would include organic species and compounds produced by microbial activity. This approach has long been recognized as an alternative but has not been used because reasonable data support is not available. The advantages of the approach would include quantitative predictions involving specified chemical systems. Pathways for microbially mediated reactions can be complex, and extensive laboratory testing could be required to represent conditions in the repository. Process models for these approaches have not been developed.

Alternative Concepts for Engineered Materials Effects Modeling—An alternative concept for the possible effect of steel on the performance of the titanium drip shield was analyzed. Specifically, the potential for hydrogen embrittlement and cracking of the titanium was evaluated. It is likely that steel ground support members will eventually fail from corrosion, possibly augmented by rock support loads, and fall onto the drip shields so that steel would come to rest on the titanium. This concept is neglected in the TSPA-SR model because (1) at the time of failure the steel will probably be coated with oxides that have no deleterious effect on titanium and (2) fine cracks in the drip shields will not transmit significant amounts of water (CRWMS M&O 2001c, Section 6.1), similar to stress corrosion cracks in the waste package (see Section 4.2.4).

Alternative Approaches for Modeling Rockfall—An alternative approach that was considered for modeling rockfall involved the use of multidimensional distinct-element analysis (CRWMS M&O 2000e). The approach has the benefit of directly applying dynamic loading to the rock mass. However, the approach is deterministic and cannot readily accommodate the available data on variability of fracturing in the potential host rock units. Accordingly, distinct element modeling was used only to confirm results obtained with the probabilistic key-block analysis and to assess thermal effects on host rock permeability (see Section 4.2.2.3.4).

4.2.3.3.8 Model Calibration and Validation

Thermal-Hydrologic-Chemical Seepage Model Calibration and Validation—The goal of model validation is to determine reasonable bounds on the system behavior over 10,000 years or longer, based on relatively short-term tests. The thermal-hydrologic-chemical Drift Scale Test model is an implementation of the repository-scale thermal-hydrologic-chemical seepage model for the situation of the Drift Scale Test. The results were compared to gas and water samples, representing potential seepage collected during the Drift Scale Test (
CRWMS M&O 2000al, Section 3.6.3.2) as a means to validate the repository scale model. These comparisons include gas-phase carbon dioxide concentrations as a function of time and space, the pH of waters collected in boreholes, and general observations on changes in concentrations of chloride and other aqueous species.

Gas-phase carbon dioxide concentrations in the model results and in the measured values showed a similar halo of strongly elevated values (approximately 2 orders magnitude greater than the air in the observation drift) around the Drift Scale Test heaters that grew outward over time. Modeled pH values of fracture waters in the drainage zones of around 6.5 to 7.5 (calcite-silica-gypsum system) are within one pH unit of waters collected from boreholes during the Drift Scale Test. Increases in the modeled pH of the waters as the rock around the boreholes heated further and began to dry out were also similar to the measured values where multiple samples were collected over time. This indicates that the model also captured the time-dependence of thermal-hydrologic-chemical processes. Simulations employing a more complex set of minerals and aqueous species estimate pH values about 0.5 to 1 pH unit higher than for the calcite-silica system. Such behavior may be more characteristic of longer time-scale water-rock interaction, which can be validated as the Drift Scale Test produces data that reflect conditions of more stable isotherms, as opposed to the data that were collected during the more transient first two years of the heating phase.

Concomitant increases (with pH and temperature) in measured silica concentrations and depletions in calcium suggest silicate mineral dissolution and calcite precipitation, trends that were also predicted by the model. Ongoing studies (by side-wall sampling and overcoring) of the actual minerals precipitated in fractures and their effect on hydrologic properties will allow for further comparisons to model results.

Measured chloride concentrations (a conservative species that shows the effects of dilution, boiling, and fracture–matrix interaction) are 5 to 10 times lower than in the pore water, a characteristic that was also captured by the model in the drainage zones. This validates that the model approximates fairly well the overall effects of dilution through condensate formation and fracture–matrix interaction (diffusive equilibration).

Precipitates and Salts Model Calibration and Validation—The normative precipitates and salts model approach (CRWMS M&O 2000cg, Section 6.5) was applied to the laboratory test results on a qualitative basis, and the results were in reasonable agreement for the bicarbonate-type water. There evidently is some sensitivity to the relative availability of calcium and sodium to form sulfate salts, which is not accounted for in the model. Minor species such as sylvite were identified by the normative model but were not detected in the laboratory tests because they were scarce and because of interference from ambient humidity. Additional testing with the chloride-sulfate-type water produced similar agreement between the normative assemblage and the minerals observed on complete evaporation. From these results, the normative model is determined to provide a valid approximation to the major minerals and salts formed on complete evaporation of waters with composition similar to either J-13 water or matrix pore water.

The low- and high-relative-humidity salts models were independently developed and validated with laboratory data (CRWMS M&O 2000cl, Section 4.1.2). The low-relative-humidity salts model approach is conservative because it tends to shorten the dry period by not allowing dry conditions for relative humidity greater than 50 percent in the presence of nitrate salts. Also, it predicts elevated chloride concentrations at low relative humidity (CRWMS M&O 2000cl, Section 7.1). The model is valid because it reproduces trends in the known behavior of salts and provides results that are within an order of magnitude of independently developed data. Such accuracy reduces uncertainty associated with the processes that control water composition at low relative humidity. The low-relative-humidity salts model results that are important for TSPA modeling are a decrease in ionic strength (due to the thicker water film) and an increase in chloride concentration as the relative humidity rises from 50 to 85 percent.

Validation of the high-relative-humidity salts model is approached using results from laboratory tests, including those described previously in which bicarbonate-type and chloride-sulfate-type waters were evaporated, and handbook solubility values for pure salts. Reasonable agreement is obtained between measured and modeled values for pH and the concentrations of sodium, carbonate species, fluoride, chloride, and sulfate (CRWMS M&O 2000cl, Section 6.5.1). Agreement to within an order of magnitude is obtained for other constituents, which is acceptable for use in abstracted models for the TSPA. As an additional check, the model is also used to calculate the solubilities of several sodium and potassium salts that are potentially important products of evaporating J-13 water. Calculated values are within a factor of 2 of the handbook values, up to 10 molal (the limit of the Pitzer model).

Microbial Communities Model Calibration and Validation—This is a bounding model that has been validated by comparison to laboratory and field data. Both model validation and code verification are described in Section 6.6 of In-Drift Microbial Communities (CRWMS M&O 2000cp). In these applications, agreement was obtained with observed microbial abundances within the order-of-magnitude tolerance level identified for the use of this model. Three comparisons were used:

  • Replication of model results originally calculated for the Swiss repository program demonstrates that the code used for the microbial model (MING) functions correctly where natural materials are combined with engineered materials.

  • Modeling of microbial conditions investigated underground at Yucca Mountain and at an analogue site at Rainier Mesa replicated the ambient microbial abundance to within an order of magnitude and confirmed that water and phosphorous availability are limiting factors to microbial growth.

  • Modeling of the laboratory tests shows that the numbers of organisms calculated by MING can agree to within an order of magnitude with independently measured data from both energy-limited and nutrient-limited tests.

Engineered Materials Effects Model Calibration and Validation—Models for oxygen consumption by steel corrosion, the effects of cementitious materials, and the impact of ferric-oxide colloids on radionuclide transport are based on bounding approaches, and the best available supporting data are used. Key uncertainties and approaches to confirmation of these bounding arguments have been identified (CRWMS M&O 2000cg, Sections 6.3 and 6.6).

Rockfall Model Calibration and Validation—The rockfall model involved the use of probabilistic key block theory, which is an accepted approach for analyzing this type of geotechnical problem. The static key block results are in agreement with observed key block occurrence in the Exploratory Studies Facility main drift and cross-drift (CRWMS M&O 2000e, Section 7.2). The results from the rockfall model have shown that key blocks are most predominant in the Tptpmn unit, which agrees with field observations. The size of key blocks observed in the field is generally less than one cubic meter, which agrees with the simulated distribution of block sizes.

The seismic component of the rockfall model involves a quasi-static method of reducing the joint strength parameters. This method was verified based on the test runs using the dynamic functions of the distinct element code UDEC. Comparison between results from the dynamic and quasi-static analyses shows a consistent prediction of block failure at the opening roof (CRWMS M&O 2000e, Attachment V).

4.2.3.4 Total System Performance Assessment Abstraction

This section describes abstraction of only those models selected for inclusion in the TSPA. The microbial communities model and the models used to bound the effects of engineered materials are determined to have minor impacts on calculated waste isolation performance (although it is acknowledged that microbially influenced corrosion may affect the longevity of Alloy 22 locally) and are excluded from the base case. Also, design analysis has shown that rockfall will not degrade the functionality of the drip shield or waste packages, so the rockfall model is also not included in TSPA-SR.

Abstraction of the Thermal-Hydrologic-Chemical Seepage Model for the Total System Performance Assessment—The thermal-hydrologic-chemical abstraction using the mean distribution of infiltration rate (
CRWMS M&O 2000a, Section 3.3.3.4.2), including climate change, is calculated for both geochemical systems discussed above. The abstracted results for both the less complex and more complex chemical systems are shown in Table 4-17 (CRWMS M&O 2000al, Section 3.3.1). The evolution of chemical conditions at the drift wall is presented as a series of four discrete time periods for the TSPA-SR calculations: (1) preclosure, (2) boiling, (3) transitional cooldown, and (4) extended cooldown. The log carbon dioxide value represents the composition of gas at the drift wall. These time periods are selected so that relatively constant concentrations can be defined for constituents of interest. After the extended cooling period, the system is considered to have returned to the ambient conditions before thermal perturbation. For both chemical systems considered, major differences in concentrations for key constituents (factor of 10) are limited to calcium, sodium, and bicarbonate ions.

Abstraction of the Precipitates and Salts Model for the Total System Performance Assessment—For the TSPA-SR model, the evolution of water in the repository drifts as temperature decreases and relative humidity increases over time is generalized as an evolution from brine to increasingly dilute water. This evolution is modeled as a succession of time intervals. In each interval, the incoming seepage flow and its composition, as well as temperature, are assigned constant values; the evaporation rate varies, and the in-drift chemical environment is determined from the evolution of thermal-hydrologic conditions at the particular location evaluated (CRWMS M&O 2000a, Section 3.3.4).

Below 50 percent relative humidity, any salts present in the drift environment will exist in crystalline form and will not form brines. At relative humidities between 50 and 85 percent, salts in the environment will deliquesce and form brines; compositions are provided by the low-relative-humidity salts model. Above 85 percent relative humidity, all of the salts are considered dissolved, and the composition is estimated by the high-relative-humidity salts model using a quasi-steady-state approximation of the degree of evaporative concentration (CRWMS M&O 2000a, Section 3.3.4.5.1).

For each time interval, values of the temperature, carbon dioxide partial pressure, seepage, and evaporation rate are obtained from other models. Lookup tables are then developed for the in-drift water composition to be used in the TSPA, based on a set of calculations using the low-relative-humidity salts model and the high-relative-humidity salts model. The lookup tables are developed for the following conditions: temperature at 95°, 75°, 45°, and 25°C (203°, 167°, 113°, and 77°F); carbon dioxide partial pressure at 10-1, 10-3, and 10-6 atmospheres; and evaporative concentration ranges up to a thousandfold (CRWMS M&O 2000a, Section 3.3.4.5.1).

Inputs to and outputs from the precipitates and salts model were abstracted for use in the TSPA-SR model (CRWMS M&O 2001b, Section 1; CRWMS M&O 2000cg, Section 6.7.4). The abstracted results used in the TSPA-SR model were based on the thermal-hydrologic-chemical model as follows:

  • Incoming seepage composition represented by the thermal-hydrologic-chemical abstraction

  • Carbon dioxide fugacity and temperature fixed at thermal-hydrologic-chemical abstracted values.

The supplemental TSPA model used a modified abstraction of the thermal-hydrologic-chemical model. Modifications included the effects of operating temperature, the effects of different carbon dioxide partial pressures, and the effects of different initial pore water (and infiltration) compositions (BSC 2001a, Sections 6.3.1.6.3 and 6.3.1.9).

Likewise, the precipitates and salts model abstraction has been updated in supplemental TSPA modeling to include the effect of the concentrations of a revised suite of elements and a select number of aqueous species used to estimate alkalinity (BSC 2001a, Section 6.3.3.6). Another improvement to the precipitates and salts model is consideration of condensation. The combined effect of improvements to the near-field geochemical model shows differences at the subsystem level composition (BSC 2001a). However, supplemental analyses do not show a significant impact at the TSPA level for either higher- or lower-temperature operating modes (BSC 2001b, Sections 3.2.4.2 and 4.2.4).

Previous Section | Next Section