Вы находитесь на странице: 1из 338

FRESCO-II

Verification and Validation

by
Karl Verfondern and Heinz Nabielek
IEK-6, Research Center Jülich

Report FZJ-IEK6-D-CC-02-Rev3

September 2012
ii
China Report 2: FRESCO-II Verification and Validation

Document Approval

name date signed

Prepared by Karl Verfondern

Approved by Hans-Josef Allelein

iii
iv
ABSTRACT

One of the prime characteristics of High Temperature Gas-Cooled Reactors is that fission
products are kept at their place of origin inside the fuel kernel, both in operation and in
accidents. To quantify this retention, diffusion models are considered embedded in computer
codes which have been or are being developed in various countries. The temperature
dependent diffusion coefficients play a central role in these models, because they – together
with operational parameters – determine the speed of fission product transport. Since they are
so important, recommended diffusion coefficients are under an international review process
and new recommendations may be published when more measurements and more analyses
become available.

The FZJ diffusion model FRESCO in combination with the particle failure model PANAMA,
both developed in the early 1980s, was chosen to be the German reference method for
calculating fission product release from HTGR fuel. In the course of testing and application of
the FRESCO model, numerous studies were conducted in the past including both
postcalculations for V&V purposes and predictions for either planned experiments or reactor
concepts as part of respective safety analyses. This report provides a detailed presentation of
the FRESCO-II code and an overview of what experience with this diffusion model has been
achieved up to now. The confidence in the applicability of these codes is strengthened by
validation along the experimental programs that include both irradiation and heating tests.

The diffusion coefficient is the essential input parameter to a diffusion code. The
recommended diffusion coefficients are a useful basis for scoping calculations. In most cases
of postcalculations of heating tests with the FRESCO code, the results prove to lead to a
conservative coverage of the experimental data. A drawback, however, is given in that it
appears from the comparison with the measurements that some important reference diffusion
coefficients are still varying over a broader range. Some of the currently recommended
diffusion coefficients have been derived by FRESCO calculations representing average data
from the heating experiments considered. This seems to be an appropriate approach to be used
in postcalculations of experiments.

The most comprehensive verification and validation effort with FRESCO-II was conducted
recently as part of the IAEA directed CRP-6 benchmark for accident conditions. The code-to-
code and code-to-experiment exercises have shown that the FRESCO performance compares
well with the performance of the other international codes. Main uncertainties are most likely
introduced less by shortcomings of the model, but rather by input data strongly depending on
material properties and exhibiting a broader variability. However, an appropriate set of input
data can be chosen which will allow a sufficiently conservative source term prediction for
HTGR reactor concepts.

In order to ascertain that the fuel is consistent with the internationally established standards,
additional irradiation and post-irradiation testing and examination will be required for newly
manufactured fuel, in order to expand and improve the existing statistical data base for fuel
performance analysis. The additional data will demonstrate successful fuel designs for future
reactors, and strengthen the statistical viability of the performance base.

v
vi
CONTENTS

ABSTRACT ............................................................................................................................. III

CONTENTS ............................................................................................................................VII

1. INTRODUCTION ............................................................................................................ 1
1.1. Efforts on Modeling Fission Product Release Behavior .................................... 1
1.2. Objectives of the Report ..................................................................................... 3

2. FRESCO-II THEORETICAL BASIS .............................................................................. 4


2.1. Source Term ....................................................................................................... 4
2.2. Safety Relevance ................................................................................................ 5
2.2.1. Radiologically relevant Radionuclides ................................................... 5
2.2.2. Radionuclides Considered Relevant in HTGR Analyses ....................... 6
2.3. HTGR Fuel Performance under Accident Conditions...................................... 12
2.4. Types of Particles ............................................................................................. 13
2.5. Fission Product Transport Phenomena ............................................................. 14
2.5.1. Inventory Buildup ................................................................................. 14
2.5.2. Recoil .................................................................................................... 16
2.5.3. Chemi-Sorption Effect of Metallic Fission Products............................ 17
2.5.4. Influence of Chemical Reactions on Release Behavior........................ 19
2.6. Simulation of Fission Product Transport and Release...................................... 20
2.6.1. Diffusive Transport of Metallic Fission Products................................. 20
2.6.2. Release of Long-Lived or Stable Fission Products from the
Particle Kernel ...................................................................................... 23
2.6.3. Retention by a Single Coating Layer.................................................... 24

3. THE FRESCO-II CODE................................................................................................. 27


3.1. Modeling........................................................................................................... 27
3.1.1. General Properties of FRESCO ............................................................ 27
3.1.2. Inventory Buildup ................................................................................. 27
3.1.3. Diffusion Model for Simulation of Fission Product Transport and
Release Behavior .................................................................................. 27
3.2. Key Parameters................................................................................................. 29
3.2.1. Free Uranium and Heavy Metal Contamination................................... 29
3.2.2. Initial Distribution in Accidents ........................................................... 29
3.2.3. Particle Failure Function....................................................................... 30
3.2.4. Diffusion Coefficients........................................................................... 31
3.2.4.1. Reference Data .................................................................................31
3.2.4.2. History of Reference Data ................................................................35
3.2.5. Sorption Isotherms................................................................................ 39
3.2.6. Fuel Temperature.................................................................................. 43
3.3. Special Treatment of Fission Gas and Iodine Release under Accident
Conditions......................................................................................................... 43
3.4. Limitations of the FRESCO-II Code ................................................................ 44

4. CRP-6 ACCIDENT CONDITION BENCHMARK (2010)........................................... 45


4.1. Background....................................................................................................... 45
4.2. General Conditions ........................................................................................... 46

vii
4.3. Results .............................................................................................................. 46
4.3.1. Sensitivity Study ................................................................................... 46
4.3.1.1. Case 1 – Fuel Kernel ........................................................................46
4.3.1.2. Case 2 – Fuel Kernel + Buffer + PyC ..............................................48
4.3.1.3. Case 3 – TRISO Coated Particle ......................................................48
4.3.1.4. Case 4 – TRISO Coated Particle + Irradiation Phase.......................50
4.3.1.5. Case 5 – TRISO Coated Particle + Irradiation Phase with
Temperature Cycles..........................................................................52
4.3.2. Postcalculations of Heating Experiments ............................................. 54
4.3.2.1. Case 6 – HFR-P4..............................................................................54
4.3.2.2. Case 7 – HRB-22..............................................................................59
4.3.2.3. Case 8 – HFR-K3 .............................................................................61
4.3.2.4. Case 9 – HFR-K6/3 ..........................................................................65
4.3.3. Predictive Calculations of Heating Experiments .................................. 67
4.3.3.1. Case 10 – HFR-EU1bis/1 .................................................................67
4.3.3.2. Case 11 – HTR-PM Fuel Sphere ......................................................70

5. VALIDATION STUDY FOR RECENT HEATING EXPERIMENTS (2011)............. 71


5.1. Transient Heating Tests .................................................................................... 71
5.2. FRESCO-II Results for New Heating Tests in KÜFA-II ................................. 93

6. VALIDATION STUDIES IN THE PAST (1983-88) .................................................. 115


6.1. First Validation Calculations .......................................................................... 115
6.2. Postcalculation of Heating Tests at FZJ ......................................................... 120
6.2.1. Objective of Postcalculations.............................................................. 120
6.2.2. Core Heatup Simulation Experiments ................................................ 120
6.2.3. Conduction of Calculations ................................................................ 123
6.2.4. FRESCO-II Results for Metallic Fission Product Release ................. 125
6.2.4.1. Heating Tests with HEU (Th,U)O2 TRISO Fuel............................125
6.2.4.2. Heating Tests with LEU UO2 TRISO Fuel ....................................140
6.2.4.3. Cesium and Strontium Concentration Profiles in Heated Fuel
Spheres ...........................................................................................204
6.2.4.4. Derivation of New Diffusion Coefficients in SiC and Matrix
Graphite ..........................................................................................210
6.2.5. FRESCO-II Results for Fission Gas and Iodine Release from UO2
Kernels and TRISO Coated Particles.................................................. 217
6.2.5.1. Heating Tests..................................................................................217
6.2.5.2. Derivation of New Diffusion Coefficients in UO2 .........................227
6.2.6. Fission Product Release from Contamination .................................... 229
6.2.6.1. Objective ........................................................................................229
6.2.6.2. Tests ...............................................................................................230
6.2.6.3. Results ............................................................................................236
6.2.6.4. Derivation of Diffusion Coefficient ...............................................236

7. PREDICTIVE CALCULATIONS FOR HTGR CONCEPTS OR EXPERIMENTS .. 239


7.1. HTR-100 (1988) ............................................................................................. 239
7.1.1. HTR-100 Reactor Design ................................................................... 239
7.1.2. Boundary Conditions .......................................................................... 239
7.1.2.1. Computer Codes .............................................................................239
7.1.2.2. Particle Failure Fraction .................................................................240
7.1.2.3. Thermodynamic Data .....................................................................240
7.1.3. Activity Release during Normal Operation of the HTR-100.............. 240

viii
7.1.4. Activity Release during a Core Heatup Accident of the HTR-100 .... 242
7.2. U.S. MHTGR (1988) ...................................................................................... 244
7.2.1. Definition of Benchmark Problem...................................................... 245
7.2.2. Initial Fuel Particle Conditions ........................................................... 246
7.2.3. Transport Data .................................................................................... 247
7.2.4. Results................................................................................................. 247
7.2.4.1. Exposed Fuel Kernel Release Model .............................................247
7.2.4.2. Diffusive Release from Standard (Intact) Particles ........................251
7.2.4.3. Comparison of Fission Product Release Models............................252
7.3. Modern German Fuel Spheres Irradiated in FRJ2-K15 (1990) ...................... 253
7.3.1. Boundary Conditions .......................................................................... 253
7.3.2. Results for the Fractional Release of Cesium..................................... 254
7.3.3. Final Remark....................................................................................... 256
7.4. HTTR (1999) .................................................................................................. 256
7.4.1. Description of HTTR Design.............................................................. 256
7.4.2. Boundary Conditions .......................................................................... 257
7.4.2.1. Operational Conditions...................................................................257
7.4.2.2. Fuel and Fuel Quality .....................................................................258
7.4.2.3. Free Uranium and Initial Distribution ............................................259
7.4.2.4. Transport Data................................................................................260
7.4.3. Metallic Fission Product Release during Normal Operation .............. 261
7.4.4. Comparison between FZJ and JAERI calculations............................. 263
7.4.4.1. JAEA Calculation Tools and Boundary Conditions.......................263
7.4.4.2. Cesium Release Behavior during HTTR Normal Operation..........263
7.5. Fuji Study (2003)............................................................................................ 264
7.5.1. FAPIG-HTGR Plant Concept ............................................................. 265
7.5.2. Boundary Conditions .......................................................................... 265
7.5.3. Results................................................................................................. 269
7.5.3.1. Release during Normal Operation ..................................................269
7.5.3.2. Release during Core Heatup Accident ...........................................270
7.6. German and Chinese Fuel Spheres Irradiated in HFR-EU1 (2006) ............... 272
7.6.1. Boundary Conditions .......................................................................... 272
7.6.2. Radionuclide Release.......................................................................... 275
7.6.2.1. Cesium Release ..............................................................................275
7.6.2.2. Strontium Release...........................................................................278
7.6.2.3. Silver Release .................................................................................280

8. CONCLUSIONS .......................................................................................................... 285

9. REFERENCES ............................................................................................................. 289

APPENDIX I. FRESCO-II INPUT DESCRIPTION.................................................. 295


I.1. Description Input Data Set ............................................................................. 295
I.2. BLOCK DATA............................................................................................... 324
I.3. FUNCTION PBRUCH ................................................................................... 324
I.4. Plot Routines................................................................................................... 325
I.5. Example .......................................................................................................... 325
I.5.1. Input File............................................................................................. 325
I.5.2. BLOCK DATA................................................................................... 326
I.5.3. FUNCTION PBRUCH ....................................................................... 328

ix
x
1. INTRODUCTION

It is the ultimate goal in nuclear systems to retain all radioactivity in all potential states of
normal and abnormal operation. In most reactors, active safety systems inside the primary
circuit accomplish this goal typically by means of redundant shut-down systems and decay
heat removal systems. With the introduction of small, modular-type high temperature gas
cooled reactors (HTGR), this classical approach has been modified in that it now relies
entirely on the high quality of coated particle fuel within the fuel element. The HTR-Modul
proposed by Siemens Interatom [Reutler 1984] was designed such that a high degree of safety
was derived from the ability of the all-ceramic fuel to retain their fission products under
normal and accident conditions, the safe neutron physics behavior of the core, the chemical
stability of the core and a design to dissipate decay heat by natural heat transfer mechanisms –
convection, radiation and conduction - without reaching excessive temperatures. The
maximum fuel temperatures reached within the HTGR core would not exceed a value that
results in significant fission product release or irreparable damage to coated particle fuel.

The philosophy of all modern HTGRs is to design the plant such that the radionuclides would
be essentially retained in the core during normal operation and postulated accidents. The
principal events to be considered when deriving fuel quality criteria for steam-cycle HTGRs
are the steady-state full-power operation and normal operating transients as well as postulated
accidents such as p rimary coolant leak (rapid depressurization), large water/steam ingress
plus pressure relief, or depressurized core conduction cooldown.

The key to achieving this safety goal is the reliance on ceramic-coated fuel particles for
primary fission product containment at their source, along with passive cooling to assure that
the integrity of the coated particles is maintained even if the normal cooling systems were
permanently disrupted. Therefore, the reactor design mandates the development and
qualification of coated particle fuels that meet stringent requirements for as-manufactured
quality and in-service coating integrity for both normal operation and accident conditions.

Demonstrating the capability of spherical fuel elements to withstand a severe depressurization


accident without any measurable loss of fission product was a primary objective of the
German Fuel Development Program. Since the 1970s, irradiated fuel elements from the
German high temperature pebble bed reactor AVR (“Arbeitsgemeinschaft Versuchsreaktor”)
and from material test reactor (MTR) tests have been investigated in accident simulation tests
[Schenk 1988, Freis 2010]. Similar efforts were also starting in the United States and Japan
with regard to their respective HTGR fuel designs.

1.1. Efforts on Modeling Fission Product Release Behavior

The experimental programs were accompanied by the development of computer codes to help
understanding the physical phenomena in fission product behavior and to provide the
appropriate tools for its prediction in specific HTGR designs. In parallel, efforts were made
on code verification, a quality control process to evaluate whether or not the code and its
submodels are in compliance with the initially given specifications or conditions, and the
physical laws (“correct programming”), and code validation, a quality assurance process to
prove evidence that the code accomplishes its intended requirements (“agreement with
physical reality”) by code-to-experiment comparisons.

1
At the Research Center Jülich, calculation codes have been developed in the past to describe
the transport behavior of radiologically relevant fission product nuclides in coated particles
and spherical fuel elements. According to the German HTGR designs that were considered in
the 1970s and 1980s, the codes should be valid for small and medium sized HTGR's covering
a temperature range up to about 2500°C. Most models were based upon the numerical
solution on Fick's first law of diffusion. The diffusion calculation model is the “classical”
approach for determining the fission product concentration in the particle kernel, the coating
layers, and the matrix graphite of the fuel element dependent on the temperature time history
during the postulated core heatup accident.

The FZJ diffusion model FRESCO combined with the particle failure model PANAMA was
chosen to be the German reference method for calculating fission product release from HTGR
fuel [IAEA 1997]. The computer model FRESCO exists in two versions. The “core” version
[Krohn 1982] was developed first to describe the fission product release behavior in a pebble
bed core during a core heatup accident phase taking account of redistribution of fission
products by the coolant flow and sorption effects at the graphite surfaces of both the fuel
elements and the top and bottom reflectors.

With the focus on small, modular-type HTGRs characterized by lower accident temperatures
with expected maximum values below 1600°C, the fission product release behavior during
normal operating conditions gained importance. A follow-up version, FRESCO-II [Krohn
1983], concentrates on a single, representative spherical fuel element and includes, in addition
to the heating/accident phase, also the phase of irradiation/normal operation. Main purpose of
this version was to simulate a complete irradiation and heating experiment for a spherical fuel
element. Furthermore, with the inclusion of the irradiation/normal operation phase, the code
allowed a more realistic determination of the source term (i.e., the quantity and duration of
radionuclide release) of small modular HTGRs.

In the course of testing and application of the FRESCO model, numerous studies were
conducted in the past including both postcalculations for V&V purposes and predictions for
either planned experiments or reactor concepts as part of respective safety analyses. This
report provides an overview of what experience with the FRESCO model has been achieved
up to now to verify and validate the code. It include also some examples of studies with
predictive calculations.

The major V&V studies in the past were:

1. Code verification and validation with the initial FRESCO (-I and -II) versions based on
some experimental data existing at that time and on the input data (mainly transport
data) available;

2. Systematic postcalculation of the core heatup simulation experiments conducted at FZJ


in the A.-Test furnace and the KÜFA furnace, and fixation of the final HBK data set of
recommended diffusion coefficients at the end of the HBK project [Verfondern 1989];

3. Accident condition benchmark exercise as part of the IAEA Coordinated Research


Project (CRP-6) on advances in HTGR fuel technology with code-to-code and code-to-
experiment comparisons for eight codes from five countries [IAEA 2011];

4. As the most recent activity, postcalculation of the new heating test series conducted at
JRC-ITU, Karlsruhe, in the KÜFA-II furnace.

2
In parallel to the applications of the code, the write-up and review of methodology to be
applied and the data to be used in safety analyses represented important steps:

 HBK standard data set;

 Compilation of the German methodology of modeling fission product release behavior


to be used in safety and risk analyses [Moormann 1987];

 Comparison of methods and data for HTR fuel performance and radionuclide release
modeling during normal operation and accidents for safety analysis as used in
Germany and the USA [Verfondern 1993];

 Comparison of methodologies used worldwide as part of the IAEA Coordinated


Research Project (CRP-2) on fuel performance and fission product behavior in gas
cooled reactors [IAEA 1997];

 Benchmark studies within CRP-6 [IAEA 2011].

1.2. Objectives of the Report

This report is part of a conceptual agreement between the Institute of Nuclear and New
Energy Technology (INET) of the Tsinghua University in Beijing and the Institute of Energy
and Climate Research (IEK-6) of the Research Center Jülich. It represents a verification and
validation report with the goals

- to describe in detail the theoretical basis of the FRESCO model;

- to present a user’s manual of the computer code including an input example;

- to present important results achieved with the code including the application to the
latest heating experiments and a comparison with other international codes’ results;

- to provide – as a conclusion from all V&V efforts, a justification for the FRESCO
model to be applied for the determination of the source term as part of safety and risk
analyses for pebble-bed HTGRs.

3
2. FRESCO-II THEORETICAL BASIS

2.1. Source Term

The release of radionuclides, especially cesium and silver isotopes, from the core during
normal operation is of particular importance to HTGRs because of their impact on operation
and maintenance dose rates. In addition, radionuclides that accumulate in the primary coolant
circuit during normal operation can become important source terms for postulated accidents.
Furthermore, for HTGR concepts dedicated to process heat production for chemical processes
and, therefore, designed for core outlet temperatures of 950°C and above, fission product
release from the core may become a critical issue.

Coated particle fuel for HTGRs has been under development for the past four decades. The
various HTGR fuel development programs have performed hundreds of irradiation tests with
a large variety of coated particle designs and conducted reactor surveillance programs to
assess fuel performance in operating HTGRs. Initial fuel development focused on
pyrocarbon-coated (BISO) fuel particles, and such particles were mass produced and used in
the early prototype HTGRs (Dragon, AVR, and Peach Bottom 1) and in the THTR. However,
the superior fission product retention capability of the TRISO particle - because of the
addition of an SiC layer - was recognized early (first serving as the driver fuel in Dragon
reload cores and then mass produced for Fort St. Vrain and for AVR reloads), and its
development and qualification became the emphasis by the late 1970s.

TRISO particles consist of multiple ceramic coatings of silicon carbide and pyrolytic carbon
surrounding dense kernels containing the fissile and fertile materials which constitute the
nuclear fuel. The TRISO coating system functions as a multi-shell pressure vessel and as a
diffusion barrier to retain fission products. The coatings are capable of maintaining their
integrity and retaining fission products at temperatures higher than those imposed during
normal operation and postulated accident conditions. The fuel particles are bonded together
by a carbonaceous matrix and formed into a spherical fuel zone encapsulated by a spherical
shell of unfueled matrix material for use in pebble-bed cores.

The ultimate measure for judging fuel performance is the ability of the coated particle to
retain fission products. Typically, the dominant sources of fission product release from the
core are as-manufactured, heavy-metal contamination (i.e., leachable heavy metal outside of
coated particles) and particles whose coatings fail in service. In addition, the volatile metals
(e.g., cesium, silver) can, at sufficiently high temperatures for sufficiently long times, diffuse
through the SiC coating and be released from intact TRISO particles. However, only for silver
is diffusive release from intact particles of any practical relevance. The important mechanisms
that can cause coating failure under irradiation have been identified as a result of irradiation
testing and postirradiation examination. Both structural/mechanical mechanisms and
thermochemical mechanisms have been observed, and models have been developed to predict
them for reactor design and safety analysis. Analytical modeling, including structural analysis
and in combination with experimental testing, has identified practical methods for minimizing
or even eliminating these particle failure mechanisms.

While the particle coatings are the most effective release barrier in the containment system,
credit is also taken for fission product retention in the fuel kernels, the fuel matrix, and core
graphite structures. Consequently, fission product transport in each of these core materials
must be characterized experimentally and modeled as part of the core design. Under the
auspices of the IAEA, an extensive international data base has been assembled from

4
laboratory measurements, in-pile experiments, and reactor surveillance programs to provide
the basis for deriving the material properties and models necessary to describe fission product
transport in core materials [IAEA 1997, IAEA 2011].

2.2. Safety Relevance

2.2.1. Radiologically relevant Radionuclides

The principal way of preventing radioactive release from the nuclear plant is based on the
defense-in-depth philosophy applied, a balance between accident prevention and mitigation.
General design criteria have been developed for LWR with a source term definition based on
credible fuel damage scenarios.

From the broad range of radionuclides generated in the course of reactor operation, not all are
equally important with regard to safety concerns. Main criterion is their impact on human
health. The significance of radionuclides is strongly depending on parameters such as

 rate of production,

 rate of radioactive decay,

 fraction released to the environment,

 pathway effects (mode of spreading in environment, uptake and concentration by food


chain/ human body).

The radiological impact on humans ascribes importance to the biological significance of the
radiation emitted which is influenced by the radionuclides’ decay energy, biogeochemical
availability, energy transfer to biological systems, and ubiquitous source during accidents.
The biologically significant radiation energy ranges from 18.6 keV (tritium) over 512 keV
(137Cs) to 5.16 MeV (239Pu). Dominant pathways are inhalation, ingestion, and groundshine.

Principal health problems arise from volatile radionuclides (e.g., 137Cs, 131I). Others represent
less a problem, because they are either not volatile (Sr, transuranium elements) or biologically
not (very) active (132Te, 133Xe). Some nuclides are available in the form of particulates (dust)
or can become attached to them. Released nuclides can fall out (rain), leached into the
soil/groundwater/ vegetation, and become part of the food chain.

Due to the large number of radionuclides and lack of data for individual nuclides, the
radionuclides are typically grouped according to their chemical and physical properties where
each group is represented by a lead nuclide. Thermodynamic properties of some relevant
fission products indicate the relative volatility of various core materials.

But classification is also influenced by the kind of safety analysis considered: For source term
determination, it is the behavior in the fuel and in the reactor building, with varying degrees
of importance for normal operation, or design basis accidents, or hypothetical accidents. For
the head end, it is the decommissioning phase and the spent fuel. Also of importance is the
type and purpose of the reactor itself. From the perspective of the target to be protected, i.e.,
humans (infant/adult) and environment, the biological effect of radionuclides may be the
criterion for grouping.

5
It is important that the source term/release categories (sometimes referred to as release
groups, release bins or source terms bins) are defined on the basis of appropriate attributes
that affect fission product releases and accident consequences. These attributes are specific to
the NPP and containment type, and there is no unique way to perform this task [IAEA 1995].

Various group structures have been proposed to date, ranging from very coarse to very fine
groupings. A selection of 54 radionuclides grouped in 7 classes was considered in the WASH-
1400 report of 1975 for severe accidents in LWRs [NRG 1975], later expanded to 9 classes. A
revision of 1995 has eventually resulted in a grouping into 8 classes [Socher 1995], which
was also recommended by IAEA to provide a reasonable group structure for LWR severe
accidents (Table 2-1) [IAEA 1995].

Table 2-1. Reasonable group structure for fission products [IAEA 1995].
Group Species
Xenon Xe, Kr
Iodine I, Br
Cesium Cs, Rb
Tellurium Te, Sb, Se
Barium Ba, Sr
Ruthenium Ru, Rh, Pd, Mo, Tc
Lanthanium La, Zr, Nd, Eu, Nb, Pm, Pr, Sm, Y
Cerium Ce, Pu, Np
Species chemical forms
Iodine (gas) I2, CH3I, HI
Iodine (aerosol) CsI
Cesium CsOH, CsI

2.2.2. Radionuclides Considered Relevant in HTGR Analyses

Due to the dominating role of LWR plants up to the present, the risk in the case of nuclear
accidents is principally conceived as a risk of severe core damage and large containment
releases. This approach, however, does not take credit of the particular safety features specific
to HTGRs. It is therefore necessary to define quantitative safety goals specially tailored to
modular HTGR technology [Carretero 2002].

Essential differences between an LWR and a modular HTGR do not only refer to the fuel
design, but also to the course of abnormal events. An important aspect is the subdivision of
the fuel in the HTGR core into ~109 – 1010 tiny fuel particles encapsulated by a ceramic
coating and their structural integrity which is retained even under accidental conditions with
maximum possible temperatures not exceeding 1600°C. Events that may lead to severe core
damage are highly unlikely. Therefore, by far most of the generated radioactivity remains
immobilized inside the particle coating and is not readily available for leakage from the fuel

6
into the primary circuit and the reactor building, respectively. This results in a very limited
amount of radioactivity that can be released.

In contrast, radionuclide release in an LWR can generally be subdivided into several


phenomenological phases associated with the degree of damage and degradation to core,
reactor pressure vessel, or concrete: (i) initial release of coolant activity associated with a
break or leak in the reactor coolant system; (ii) after fuel cladding failure, gap release of
mostly volatile nuclides; (iii) upon core degradation and loss of fuel geometry, release of
significant quantities of the volatile nuclides in the core inventory as well as small fractions of
the less volatile nuclides into containment; (iv) after contact and interactions of molten core
debris with the concrete structural materials, release of the left-over volatile radionuclides and
lesser quantities of non-volatile radionuclides as well as comprehensive amounts of concrete
decomposition gas products into the containment atmosphere.

A further significant difference between LWR and modular HTGR in the course of accident
scenarios is given in the time scale of the events and the radionuclide release connected with
it. Even in the case of a loss of active core cooling, the developing temperature transients are
very slow due to the small power density and the high thermal inertia of the graphite in the
core.

Another aspect is the different long-term behavior of the coolant as the carrier medium for
activity. Different from steam, helium remains inert and in a single phase, also under accident
conditions. A breach of the primary circuit will not result in a complete loss of the coolant,
but rather retain helium at atmospheric pressure, which might be – in the case of decreasing
temperatures – in a slow diffusion process gradually replaced with air.

The transport of any metallic fission or activation product could be calculated if the transport
parameters of the nuclide were available. But it is impractical (and unnecessary) to evaluate
production, transport and release of all species generated in a nuclear reactor. Therefore, only
the radiologically most significant nuclides are analyzed. Key radionuclides are selected
based on the combination of their fission yield, their transport and release properties and their
radiological hazard level.

According to an INL White Paper on HTGR mechanistic source terms [INL 2010], the
radionuclides of interest in HTGR design and safety analysis are classified according to
similar in-core and ex-core behavior with regard to radionuclide transport modeling
(Table 2-2). The definition is based on the facts that there is significant retention of the fission
products in the kernel itself, and that both fuel kernel and SiC-TRISO coating are the major
barriers to fission product release from the fuel particle.

Groups of fission products classified according to their transport behavior in the HTGR fuel
have been defined in [Brinkmann 2006]. Distinction is made between three principal groups
as outlined in Table 2-3.

For the release of radionuclides from the reactor building to the environment, the
classification is given their behavior in the reactor building determined by the deposition
behavior, formation of anorganic and organic compounds (e.g., methyl iodide), filter
effectiveness for given species, or attachment to aerosols. In the decommissiong phase and for
radioactive waste disposal, long-lived volatile/gaseous nuclides (e.g., tritium, 85Kr) and
activation products in the reactor components (14C, 60Co) are of highest importance.

7
Table 2-2. Classes of radionuclides of interest for HTGR design [INL 2010].
Radionuclide Key Form in fuel Principal Principal
class nucl in-core behavior ex-core behavior
3
Tritium H Element (gas) Permeates intact SiC; Permeates through
sorbs on core graphite heat exchangers
133
Noble gases Xe Element (gas) Retained by PyC/SiC removed by helium
purification system
131
Halogens I Element (gas) Retained by PyC/SiC Deposits on colder
metals
137
Alkali metals Cs Oxide-element Retained by SiC; Deposits on
some matrix/graphite metals/dust
retention
132
Tellurium group Te Complex Retained by PyC/SiC Deposits on
metals/dust
90
Alkaline earths Sr Oxide-carbide High matrix/graphite Deposits on
retention metals/dust
110m
Noble metals Ag Element Permeates intact SiC Deposits on metals
140
Lanthanides La Oxide High matrix/graphite Deposits on
retention metals/dust
239
Actinides Pu Oxide-carbide Quantitative Retained in core
matrix/graphite
retention

Table 2-3. Classification of radionuclides according to transport behavior in fuel.


No. Class Representative Characterized by (or assumed)
131
1 Non-metals I, 85Kr, 133Xe Completely retained by any intact coating layer
No holdup in graphite
If HMC, slow diffusion from graphite grain
No sorption on graphite surfaces
134
2 Reactive Cs, 137Cs, 90Sr Slow diffusion through SiC coating
metals Moderate diffusion through matrix
Certain holdup in fuel kernel at NO temperatures
High release from kernels at accident temperatures
High sorption capability on graphite surfaces
110m
3 Noble Ag Fast diffusion through SiC layer
metals Fast diffusion through matrix
No sorption on graphite surfaces

8
The radiologically most relevant radioisotopes monitored during accident simulation testing
are listed in Table 2-4. These include the long lived strontium (90Sr), silver (110mAg), cesium
(134Cs, 137Cs), and krypton (85Kr) activation and fission products, and the short lived iodine
(131I) and xenon (133Xe) fission products. Isotopes of the same chemical species are assumed
to result in same release fractions. More nuclides such as fission gases or other iodine isotopes
may be added to the above list depending on the expected fuel conditions.

Cesium and strontium are radioactive for a long period of time. The materials can contaminate
property and, therefore, would require extensive cleanup. Cesium and strontium can be
absorbed into the food chain and are potential cancer-causing elements. Cesium is a long-term
contaminant of pastures/crops, which may preclude food production from affected land.

For atmospheric releases from a reactor building, 131I is likely to be important because it is
produced in quantity, it is volatile and therefore readily released, and because it is absorbed in
the grass-cow-milk food chain and tends to concentrate in the thyroid gland with the potential
for causing thyroid cancer. Due to its short halflife, 131I is more or less a “first-month
problem. Tellurium becomes iodine as it undergoes radioactive decay; its health impacts are
similar to iodine.

Type and purpose of the nuclear plant will also have a more or less influence on the selection
of important nuclides. Safety analyses for HTGRs have shown that in direct cycle concepts,
the analysis of the silver isotope 110mAg is required with regard to maintenance of the helium
turbine. For nuclear process heat reactors connected to a (conventional) chemical factory, it is
of utmost importance to guarantee a clean product that is delivered to the consumer. The
decisive nuclide here is tritium produced during normal operation and easily permeating
through metal walls.

The relative characteristics of the metallic fission products of radiological interest for LEU
UO2 TRISO fuel under HTGR service conditions are as follows:

 Silver is known to begin to be released from TRISO fuel at temperatures above


1000°C of normal operating temperatures, and to be released from intact particles in
significant fractions after several days at 1600°C. It deposits in graphite at 900°C and
plates out on metallic surfaces at temperatures of 800°C and diffuses into the bulk
metal where it is effectively captured. As a result, silver does not present an important
concern with regard to offsite dose, but can be a dominant contributor to occupational
dose depending on component maintenance requirements.

 Strontium is retained to a large extent in oxide kernels during normal operation even
when coatings are defect and is slowly released at temperatures beyond modular
HTGR accident conditions (i.e., approaching 1800°C for several days). Additionally,
it is strongly adsorbed in matrix material and little is released from spherical fuel
elements. Different from other species, it is difficult to measure and the results shown
have an uncertainty of an order of magnitude compared to around 10% for others.

 Cesium is also released from particles with defective silicon carbide layers in normal
operation and can be released from the fuel spheres depending on the local conditions.
However, there is a significant delay in release from the sphere due to holdup in the
matrix. In the AVR, the LEU UO2 TRISO spheres were typically net absorbers of
cesium due to the low silicon carbide defect fractions.

9
Table 2-4. Selection of radiologically relevant fission products considered in HTGR safety analyses.
Radio Half life Relevance
nuclide
85
Kr 10.783 years Normal gaseous effluent, indicator of particle coating failure and conservative upper limit for iodine release, important for
waste management
88
Kr 2.84 hours Precursor of 88Rb, normal gaseous effluent
89
Kr 32.32 seconds Precursor of 89Sr, normal gaseous effluent
90
Kr 32.32 seconds Precursor of 90Sr, normal gaseous effluent
133
Xe 5.247 days Normal gaseous effluent, important for normal operation and waste management
137
Xe 3.818 minutes Precursor of 137Cs, normal gaseous effluent
138
Xe 3.818 minutes Precursor of 138Cs, normal gaseous effluent, important for waste management
131
I 8.021 days Similar behavior to noble gases, mobile, greatest significance for design and licensing, important for heatup accident
conditions, off-site dose, concentration in thyroid, absorption in food chain
133
I 20.81 hours Similar behavior to noble gases, mobile, greatest significance for design and licensing, important for heatup accident
conditions, off-site dose, concentration in thyroid, absorption in food chain
134
Cs 2.066 years Mobile at high temperatures, sorption on metallic and graphitic surfaces, activation product, may be available in the coolant
at low temperatures in the vapor phase as Cs, Cs2O, CsOH, long term behavior after extreme accidents and in risk analyses,
off-site whole body dose, absorption in food chain
137
Cs 30.07 years Mobile at high temperatures, sorption on metallic and graphitic surfaces, may be available in the coolant at low temperatures
in the vapor phase as Cs, Cs2O, CsOH, long term behavior after extreme accidents and in risk analyses, off-site whole body
dose, absorption in food chain
131
Te 25.0 minutes Precursor of 131I, relatively volatile, similar behavior to iodine
133
Te 25.0 minutes Precursor of 133I, relatively volatile, similar behavior to iodine
89
Sr 50.55 days Formation of very stable (carbides and) oxides, mobile at high temperatures, efficiently retained in UO2 kernels, strong
sorption surfaces
90
Sr 28.78 years Formation of very stable (carbides and) oxides, mobile at high temperatures, efficiently retained in UO2 kernels, strong

10
sorption surfaces, long term behavior after extreme accidents and in risk analyses, off-site dose
110m
Ag 249.8 days Small inventory, highly mobile, release from intact coated particles, important for maintenance, operation and maintenance
dose, high public inhalation dose coefficient
111
Ag 7.454 days Highly mobile, release from intact coated particles, important for normal operation and accident conditions
109
Pd 13.44 hours Precursor of stable 109Ag, major contributor to corrosive attack on silicon carbide
3
H 12.323 years Ternary fission and activation product, ubiquitous production, complete retention in intact particles, leakage from primary
circuit in the forms CH3T, HT, HTO, tendency to follow water cycle in nature, ability to become tissue-bound in humans
and biotic environment, important for waste management
14
C 5730 years Activation product from 14N, 17O, 13C with the first being the dominant source, leakage from primary circuit in the forms
CO2, CO, CH4, important for waste management
41
Ar 1.83 hours Activation product in air of rooms near reactor pressure vessel, high-energy gamma emitter
60
Co 5.272 years Activated metallic erosion product, important for waste management, high adult public inhalation dose coefficient

11
The primary isotopes of cesium of interest are 134Cs and 137Cs, whose inventories developed
differently as a function of burnup, but whose release fractions under accident conditions are
very similar. The combination of 85Kr and 137Cs data in the early phase of the heating tests
allow determination of both through-coating failure fractions and failure fractions of particles
with silicon carbide defects and an intact pyrocarbon layer.

This HTGR-typical source term vector is also in agreement with [Moormann 1992] to be used
in combination with German regulations for design basis accidents [GRS 1983] and
emergency planning [GMBL 1989, ICRP 1984] or with more realistic accident consequence
models [Ehrhardt 1988].

2.3. HTGR Fuel Performance under Accident Conditions

Retention of fission products in the advent of an accident is the overriding goal in nuclear
systems. In most reactors, active safety systems are located inside the primary circuit to
accomplish this goal and in many designs these are in the form of redundant shut-down
systems and redundant decay heat removal systems. In Germany in the 1980s, the
introduction of the 200 MW HTR-Modul design by Siemens [Lohnert 1988] employed an
approach which relied entirely on the ability of coated particle fuel within the spherical fuel
element to retain all key radionuclides during design-basis accidents. The HTR Modul was
designed such that a high degree of safety was derived from the ability of the all-ceramic fuel
to retain their fission products under normal and accident conditions, the safe neutron physics
behavior of the core, the chemical stability of the core and the ability of the design to dissipate
decay heat by natural heat transfer mechanisms – conduction, radiation and conduction –
without reaching excessive temperatures. The maximum fuel temperature reached within the
HTR core would not exceed a value that results in irreparable damage to coated particle fuel.

Demonstrating the capability of spherical fuel elements to withstand a severe depressurization


accident without any measurable loss of fission product was a primary objective of the
German Fuel Development Program. Since the seventies, irradiated fuel elements from AVR
and from MTR tests have been investigated in accident simulation tests. Out-of-reactor
heating of irradiated fuel elements is the method employed to simulate the depressurization
accident in small modular HTGRs [Lohnert 1988]. In response to the reactor designer’s call
for a passively safe HTGR, a unique set of specialized remote equipment was designed,
assembled and installed in the FZJ hot cells.The depressurization accident consequences are
assessed by measuring the released fission gases Xe, Kr and solid fission products Cs, Sr, Ag,
and Eu during controlled heatup of previously irradiated fuel elements. A dedicated hot cell
furnace with the capability of quantitatively measuring gaseous and solid fission product
inventories released during heating was developed and operated at the FZJ for this purpose.

The maximum expected fuel element temperature evolution in a depressurized loss of forced
coolant accident in a small modular HTGR is shown in Fig. 2-1. This “so-called” maximum
fuel temperature limit is set at 1600°C based upon the estimated maximum core temperature
of ~1500°C plus a reasonable estimate of the effect thermal property uncertainties have on the
maximum temperature estimate. For testing purposes the temperature range, 1600-1700-
1800°C are of interest for the core heatup simulation testing program as illustrated in the
figure. A wide range of test results have been reported and are the basis of licensing
applications for modern retentive HTR fuel (Schenk 1988, 1989, 1990, 1991b, 1997, Nabielek
1989, 1990, Kostecka 2004, Freis 2008, 2009).

12
Figure 2-1. Temperature evolution during the depressurization accident in small modular
HTGRs and temperature profiles in KÜFA heating tests at FZJ.

Through this dedicated effort, it was possible to demonstrate that modern LEU UO2 TRISO
coated fuel particles will retain all safety relevant fission products up to 1600°C at a level not
exceeding release under normal operating conditions, which in itself is very low. If the
maximum fuel burnup is kept strictly below 11% FIMA, as is typical for present HTGR
designs, the allowable temperature limit may be higher than 1600°C, but this must be
established with additional investigations. If LEU UO2 TRISO coated fuel particles are
irradiated up to15% FIMA, then the maximum temperature under all circumstances has to be
rigorously limited to 1600°C.

2.4. Types of Particles

In fully intact TRISO coated particles, fission products formed have to be transported through
all the coating layers before diffusing through the matrix material and being released into the
coolant, respectively. Different from metallic fission products, the diffusion parameters for
fission gases in PyC are very small. Therefore, the transport rates of gaseous fission products
are extremely slow, and release fractions from intact coated particles are negligible. Iodine is
considered to show a similar transport behavior as the fission gases.

Even under the best manufacturing conditions, a small fraction of coated fuel particles will be
defective. Furthermore, under operating conditions and also under abnormally high
temperatures, there is a certain chance for coated fuel particles to fail. In most models, a
defective/failed particle is represented conservatively by a bare fuel kernel. It is assumed that
fission products from failed and defective particles do not have to travel through coating
layers and only have to escape the kernel to be released to the matrix material.

Particles with defective or missing coating layers, particularly SiC, can be modeled by a
modification of the diffusion coefficients.

13
2.5. Fission Product Transport Phenomena

The schematic in Fig. 2-2 summarizes the most important physical phenomena that occur in
an HTGR fuel element regarding transport and release of fission products.

Figure 2-2. Fission product transport and release phenomena.

2.5.1. Inventory Buildup

A wide spectrum of fission product nuclides is created by the fission of uranium and
plutonium resulting from successive β-particle emission from the fission fragments. Most
have very short half-lives or form stable oxide compounds in the fuel kernel. The inventory
buildup of fission products during reactor operation is more or less linear for long-lived
isotopes, whereas short-lived isotopes soon reach an equilibrium state.

Neglecting several details, the generation of fission products is described by the differential
equation
n
dN
  Yi f i  N
dt i 1

where N is the number of fission product atoms, t is time [s], Yi is the fission yield and fi is
the fission rate of fissile isotope i [s-1], n is the number of fissile isotopes, and λ is the decay
constant [s-1] of the isotope with halflife T1/2 [s]:

ln 2
T12 

Neglected items are contributions from precursors, neutron absorption etc. When fission rates
are constant, the fission product activity A [Bq] is given by

14
 
n
A  N   Yi f i 1  e  t
i 1

Shorter-lived fission products reach the equilibrium activity faster


n
A  N   Yi fi
i 1

Long-lived isotopes like 90Sr and 137Cs show an approximately linear increase with irradiation
time and with burnup as shown in Fig. 2-3. Codes like ORIGEN [Vondy 1962, Bell 1973] are
used to predict fission product inventories in reactor systems.
Inventory [Bq] per spherical fuel element

Burnup [%FIMA]
Figure 2-3. Measured inventory buildup in a fuel sphere with burnup.

Activation products like 134Cs and 110mAg are created when a fission product captures a
neutron to form a new nuclide. Some of these activation products, whose inventory buildup is
different from fission products are very important for HTGR maintenance and safety. For
example, the 110mAg isotope is the product of an (n,γ)-reaction of the stable 109Ag whose
decay chain by β emissions is shown below:

15
       ( n , ) 
109
40 Zr  109
41 Nb  109
42 Mo  Tc 
109
43
109
44 Ru  109
45 Rh  109
46 Pd  109
47 Ag  110 m
47 Ag  110
48 Cd

Due to decay effects on fission product transport and the ability to diffuse through the layers
of a coated fuel particle, radiologically important fission products released from fuel elements
are divided into relatively short-lived gaseous fission products and long-lived metallic fission
products. With regard to the silver transport through a TRISO coating, which is comparatively
fast at temperatures above 1000°C, a part of the silver may be released from the fuel in form
of the stable (inactive) 109Ag and no longer be subjected to an activation to 110mAg. Different
theoretical models and sets of transport parameters (e.g., diffusion coefficients, correlations)
describe both fission products groups’ behavior.

2.5.2. Recoil

Due to their high kinetic energy, fission products travel a finite distance after birth before
being stopped by the surrounding material. Consequently, some of the fission products
formed near the surface of a fuel kernel are released directly into the buffer region. This
phenomenon is called recoil effect. Also a part of the fission products present in the oPyC
layer (due to uranium contamination during the fabrication process) may escape directly into
the adjacent matrix material. Similarly, some of the fission products formed in the grains of
the graphite matrix (heavy metal contamination is the source of such fission products) escape
the grains directly without diffusion onto the grain boundaries.

Direct recoil after fissioning from the kernel into the buffer layer is an additional release
mechanism. It is independent of irradiation parameters and, particularly, of temperature and
can be derived from purely geometrical considerations. At lower temperatures (below normal
operating temperatures) diffusion transport is very slow and releases are dominated by recoil.

The recoil contribution can be calculated as follows [Flügge 1938, Sasaki 2007]:

R  2
2 Rr  r0  R 2  r 2 
r0  R 4 Rr
 4r 2 dr
3 R  1 R 
 
2

Frecoil    1  
4r0
3
4 r0  12  r0  
 
3

where Frecoil is the release fraction into the buffer (or the matrix material or the grain
boundaries), ro is the radius [m] of the kernel (or the coated particle or the graphite grain), and
R is the mean recoil distance [m].

The additional 137Cs constant release term of 2.1% has also been established experimentally.
This is consistent with 7 μm recoil distance of Cs atoms from the kernel into the buffer layer
and a 250 μm UO2 kernel radius. Respective release fractions from differently sized particle
kernels are listed in Table 2-5.

16
Table 2-5. Recoil fraction as a function of kernel radius.
Kernel Radius Cs-recoil release fraction
[µm] [%]
100 5.25
125 4.20
150 3.50
200 2.62
250 2.10
300 1.75
350 1.50
400 1.31
450 1.17
500 1.05

2.5.3. Chemi-Sorption Effect of Metallic Fission Products

At the boundary between fuel element and coolant, transition of diffusing atoms occurs due to
adsorption and desorption (evaporation) processes, respectively. In most practical cases, the
processes are sufficiently fast that a local equilibrium may be assumed between the
concentration adsorbed at the solid surface and that in the adjacent boundary layer in the gas
(“vapor pressure”). Both are connected by a steep concentration difference of several orders
of magnitude as is shown in Fig. 2-4. These concentrations are put into relation by the so-
called sorption isotherms, which are a function of the temperature and at high surface
concentrations (in the so-called Freundlich regime, additionally of the surface concentration
itself). In practice, adsorption is mainly of importance in the case of porous solids which have
a large specific surface. The vapor pressure is expressed as an exponential function of
temperature and sorbate concentration.

The sorption effect is strongly depending on the concentrations of the nuclide considered in
bulk and coolant, on the type of radionuclide and the thermodynamic conditions, but also on
the type of graphitic material. The sorption effect is furthermore influenced by the coolant
velocities. It is less pronounced during normal operation with high speed coolant flows along
the graphite surfaces, but will become more significant at low gas velocities like in the
scenario of an unrestricted core heatup in the depressurized reactor.

Coked phenolic resin binder, a component of the fuel element matrix graphite, has a high
sorption capacity for cesium and strontium. The high values of sorption enthalpies show that
Cs and Sr are strongly bound by chemi-sorption. The large quantity of matrix material in the
core indicates a high potential for cesium and strontium retention during normal operation and
in the case of a core heatup accident. This potential increases with fast neutron irradiation at
temperatures below ~1100°C and decreases, if irradiation temperatures reach ~1400°C
[Hilpert 1988]. In contrast, the sorption capacity of graphite for silver and iodine is much
smaller.

17
Figure 2-4. Concentration profiles for adsorption and desorption processes.

Matrix material was found to have a significantly higher sorption capacity for cesium and
strontium than other nuclear graphites. This can be attributed to the coked phenolic resin
binder component, which has a low density compared to graphite, and where the carbon
layers are not parallel nor planar and have many defects. The sorption potential increases on
irradiation with fast neutrons and decreases with increasing irradiation temperatures. In
comparison to cesium and strontium, the sorption potential of silver and iodine is much
smaller and therefore usually not taken into consideration meaning that an unrestricted
transition is assumed from the fuel element to the coolant. The fact that the sorption capability
of a graphitic material is strongly influenced by its amount of binder coke can be seen from
Fig. 2-5. It shows for cesium the equilibrium concentration at 1100°C in four graphitic
material samples with different binder coke contents [Moormann 2006].

Sorption investigations were usually conducted for single species. For HTGR release
predictions, however, multiple species sorption or co-sorption should be taken into account.
The competition of different species for sorption sites in the graphite can lead to a reduction
of the overall sorption capability of a single species. This effect was examined in more detail
for the Cs-Rb system [Myers 1979], the Cs-Ba system [Kazi 1980], and the Cs-Sr system
[Kazi 1980b]. One approach for the conservative treatment of this effect as was used in
German safety analyses was to consider the sum of all inventories of chemically similar
species as the “inventory” input of a species relevant for calculating the sorption effect.

18
Figure 2-5. Comparison of sorption capability of different graphitic materials for cesium
(partial pressure of Cs: 0.03 Pa, temperature: 1100°C).

2.5.4. Influence of Chemical Reactions on Release Behavior

Chemical reactions of the impurities generally present in the primary (and secondary) coolant
such as H2O, CO2, CO, H2, CH4, or N2 can occur over the total temperature range of an
HTGR, from roughly 300 to 1200°C [Nieder 1981]. Corrosive reactions with the core
graphite are principally of importance because they may degrade the strength of core
structures and may also lead to the formation of flammable gas mixtures. The reactions with
the graphite are of either thermal or radiolytic or sorptive nature. While in colder core areas,
radiolytic reactions are dominating, thermal reactions are essential at higher temperatures.

The fission products bonded to the graphite surface may interact directly with the water or
steam or oxygen, or may be released with the gasified graphite. The latter process would even
liberate fission products trapped inside graphite grains. Porosity or cracks in the graphite
developing through the oxidation processes may offer a faster way for the air or water to
reach the fuel particles. It was found that matrix material is more reactive than nuclear-grade
structural graphite.

The water-carbon reaction is an endothermal process which consumes nuclear decay heat.
Therefore high temperatures and high material loss are not expected. Reaction rates are
dependent on the temperature. Changes in the chemical form of fission products or changes of
the properties of the graphite material could change the transport rates through the matrix.
Liquid water in contact with graphite is expected to mobilize a great part of the chemisorbed
metallic fission products which form stable compounds with the water. Although the
penetration by water is a slow process, in safety analyses, the assumption is made that all
metallic fission products in graphite regions, which get into contact with liquid water, will be
released [Moormann 1992]. The kinetics of most reactions is determined by an activation
energy, typically described in form of an Arrhenius approach.

19
Calculation models have been applied to assess air ingress accidents. The contributing
transport processes are convection, conduction, and radiation with their shares depending on
the conditions of flow rate, oxygen concentration, temperature, thermal gradient, local
distribution of chemical heating, and local and global burning.

The release behavior of fission products is influenced by their chemical state [Moormann
1985]. Experience acquired from the accident in the Three Mile Island (TMI) nuclear power
station, for example, has shown that the release of iodine was by several orders of magnitude
less than expected, explained by chemical reactions within the aqueous phase. Distinguishing
between metals and non-metals, the former are representing the dominant part of the fission
products. For the HTR-Modul, the estimated ratio is

(Cs + Rb + Sr + Ba) / (I + Br + Te + Se) ≈ 7.5

2.6. Simulation of Fission Product Transport and Release

2.6.1. Diffusive Transport of Metallic Fission Products

The mathematical modeling of fission product transport uses two basic laws:

 the conservation of mass;

 the formulation for the transport speed.

The conservation of mass can be expressed by

 c   
 p  c  t  dV
V
  j  dS
S

where V [m3] are volume and S [m2] surface under consideration, p [m-3·s-1] and c [m-3] are
volume specific generation rate and concentration, λ [s-1] is the decay constant, and j [m-2·s-1]
is the transport flux through all outer surfaces. λ and p describe removal and production in a
general way and may involve several additional processes.

The transport of metallic fission products through the fuel materials is modeled as a transient
diffusion process. The transient diffusion equation is typically solved numerically with
appropriate boundary and interface conditions.

The rate of migration of a species through a homogeneous medium is defined by its mass flux
and a concentration gradient as the driving force. According to Fick’s first law as described in
[Crank 1975], the flux of atoms diffusing through a medium is proportional to the
concentration gradient:

c
Jx   D
x

where J is the diffusion flux [atoms/(m2·s)], D is the diffusion coefficient [m2/s], c is the
concentration [atoms/m3], and x is position or length [m].

20
The second Fickian law describes the time dependent change of the concentration field by
diffusion. Assuming the diffusion coefficient D to be a constant and including source (e.g.,
production from nuclear fission) and sink (e.g., radioactive decay) terms, Fick’s equation
reads as follows:

c  2c
 D  c  S
t x2

or in spherical coordinates:

c   2 c 1 c 
 D  2     c  S
t   r r  r 

where λ is the decay constant [s-1], and S is the fission product production rate [atoms/(m3·s)].
The following boundary conditions apply:

 Concentration gradient is, for symmetry reason, 0 in the particle center at radius r = 0

c
=0
r r 0

 Continuity of flux and concentration is given at the interface between two adjacent
materials with diffusion coefficients D1, and D2:

c c
c1  c 2 and  D1   D2
r 1
r 2

At the surface of the particle, the concentration is assumed to be the average


concentration in the matrix graphite:

c r  rp   c gr

 For fission product transport in the fuel element matrix graphite, a third boundary
condition applies describing the mass transfer at the fuel surface which is given by

c
D    cbl c gas  with cbl  f c w ,T    c w
r r  rs

where β is the mass transfer coefficient from the surface to the coolant, cbl is the concentration
in the coolant immediately above the graphite surface [atoms/m3], cgas is the mean
concentration in the coolant [atoms/m3], cw is the concentration in the graphite near the
surface [atoms/m3]. α is the ratio between boundary layer and wall concentration and is
determined by means of sorption isotherms.

This third boundary condition defines the release rate from the surface of the fuel sphere into
the coolant. In most cases, free evaporation from the surface can be assumed which is
expressed by

21
   or cw  0

The release rate R [s-1] from the surface at r = rs [m] is then

c
Rt   4rs2 D
r r  r s

and the cumulative fractional release F after time t

 R  e
t
  t  
d
F t   0
Total Inventory

The transport speed is determined by the diffusion coefficient D [m2/s] which is typically
depending on the temperature. For gas-in-gas binary diffusion, the diffusion coefficient is
approximately proportional to T1.5. For transport in solids, an Arrhenius-type temperature
dependence [Jost 1960] of the form

 Q 
D  Do exp  
 RT 

is assumed where Do is a pre-exponential factor [m2/s], Q is the activation energy [J/mol], R is


the universal gas constant, R= 8.3145 J/(mol·K), and T is the absolute temperature [K].

The transport of mobile fission metals is certainly structure-sensitive and more complex than
classical Fickian diffusion and likely a combination of lattice diffusion, grain boundary
diffusion, pore diffusion, etc., further complicated by effects like irradiation-enhanced
trapping and adsorption. Consequently, any quoted diffusion coefficient should be called an
“effective” diffusion coefficient which implies that the overall migration process is
approximately described by Fick’s laws [Nabielek 1974].

An effective diffusion coefficient can be deduced experimentally for specific temperatures or


temperature ranges, since the diffusion of fission product atoms increases as the temperature
is raised according to the Arrhenius equation. Sometimes experimental findings suggest the
incorporation of additional dependencies like neutron fluence, Γ, or concentration of the
considered species, c, or the use of two temperature ranges with different activation energies
like one for the normal operation phase and another one for the accident temperature range:

 Q 
Deff  D T , , c,...   Do ,i exp  i 
i 1, 2  RT 

Burnup dependence may be taken into account by the following equation:

D0  C1[1  (1  n)C2 F n ]

where C1, is C2 and n are derived from experiments and F the burnup expressed in FIMA
(fissions per initial metal atom, 1 % FIMA ≈ 10 MWd/t).

22
2.6.2. Release of Long-Lived or Stable Fission Products from the Particle Kernel

In the Equivalent Sphere Model [Nabielek 1974], the primary fission product retaining object
is a UO2 particle kernel (or a grain) that is modeled as a sphere with radius a. The
accumulated fractional diffusive release F of long-lived or stable fission products (decay
constant ) is given by the transient solution of the diffusion equation to yield

6 

1  exp  n 2 2 D't 
F 1
D't

n 1 n 4 4

where D’ is the reduced diffusion coefficient [s-1] and t is the irradiation time [s].

F is shown as the lower curve in Fig. 2-6 as a function of diffusion time. F can be efficiently
approximated by

D t
3
F 4  D t for D t  0.35 and
 2

1
F  1 for D t  0.35
15 D t

During accident condition heating tests, the release of long-lived or stable fission products
from the particle kernel or a spherical grain can be approximated by solving the diffusion
equation, but without source/sink terms, with the boundary condition

c 1  2 rc 
D  0, cr  a, t  = 0 (a is kernel/grain radius).
r r r 2 r 0

The initial condition of a uniform concentration in the sphere approximates the case of prior
cold irradiation. During an isothermal heating, it is assumed that gas which diffuses to its
surface is released from the sphere. As shown in [Booth 1957, Nabielek 1974, Förthmann
1977], the out-of-pile fractional release F after a time t is then given by


Fout of  pile  1  6 

exp  n  D t
2
2

2
2

n 1 n

where D   D a 2 is the reduced diffusion coefficient [s-1].

Fout  of  reactor is efficiently approximated by

Dt
F 6  3Dt for Dt  0.15 and

F  1

6 exp   D t
2
2
 for Dt  0.15

23
The release curve, applicable to a heating test of a bare kernel or broken particle, is shown as
the upper term in Fig. 2-6.

1 E+0
Fractional Release

1 E-1

1 E-2
1 E-4 1 E-3 1 E-2 1 E-1 1 E+0
tau = D't

Figure 2-6. Fractional release of long-lived fission products during irradiation (lower curves)
and during heating (upper curves) as a function of diffusion time t = D’t
according to the equivalent sphere model.

2.6.3. Retention by a Single Coating Layer

In the case where fission product retention is mainly effected by one coating layer, frequently
the concept of a breakthrough time [Crank 1956]

1 D
tbreakthrough  with D* 
6 D* d2

is used whereby d is the layer thickness and D is the diffusion coefficient in the retaining
layer. This concept is somewhat misleading in that the breakthrough is defined arbitrarily and
fission products are released through the layer at all times before the “breakthrough”.

In a spherical coating layer of inner radius ri and outer radius rs, it is

ri  r  rs ; d  rs  ri

characterized by diffusion coefficient D, the diffusion equation describing concentration c as a


function of time t and radius r is given in spherical geometry by

c 1 2
D r c 
t r  r2

In the coating layer, the initial condition is c  0 and   0 for long-lived fission products.
Boundary condition at the outer surface r = rs is c  0 at all times, i.e., infinitely fast
evaporation from the surface.

24
Inside the spherical shell, we have initially N freely movable atoms, i.e., an initial inner
concentration

N
ci t  0  
4
 ri3
3

that can be depleted only by diffusion into the shell. Rewriting the mass balance equation
across this inner volume and the inner surface of the shell gives

4 ri3 d ci c
 4 ri 2 D
3 dt r r  ri

We assume the partition coefficient between inner volume and shell to be 1, i.e., ci  c r  r at
i

all times. Of interest is the cumulative fractional release defined by


t
c
  4r dt 
2
D
s
r r r
F t   0 s

The solution (Minato 1995) for the fractional release is

 

F    1  T m exp  ym2
m0

Dt
where   D*t  is the dimensionless diffusion time.
d2

d
  ; d  rs  ri are the geometric parameters. The coefficients in the infinite sum terms are
ri
given by

3 ym
tan ym  , ym  0
ym2  3 2

and

1  1
Tm  
cos ym y 12
3 2  4.5 3
   2
m
6 2 ym  3 2

For short diffusion times, the fractional release can be approximated by

24 1   
Fsmall _ times 
 1 
exp    1.5 1  6 1     12 5  6  6  
2 2

  4 

The large diffusion time approximation is

25
Fbig _ times  1  1    exp 3 
 2

This correlation is more efficient and has better precision than the breakthrough
approximation in diffusion text books [Crank 1956]:

F  3
rs d  1
    2


 1n exp  n 2 2  
ri ri  6 n 1 n 2 2 

Single layer approximations have been useful in estimating the diffusion coefficient of silver
in SiC [Nabielek 1977] and in estimating diffusion coefficients of Cs, Ag, Ru, Ba, Sr in ZrC
[Minato 1995].

26
3. THE FRESCO-II CODE

3.1. Modeling

3.1.1. General Properties of FRESCO

The FRESCO code is flexible in the particle design by allowing treatment of both BISO and
TRISO-coated particles. The diffusion model distinguishes between two types of particles,
intact particles and defective/failed particles. A defective or failed particle is represented by a
bare kernel only meaning that fission products are released immediately into the matrix
graphite. At the moment of failure of a particle, the fission product content inside the coating
layers is released into the matrix graphite. For the inventory still left in the kernel, this
moment is “time zero”, from which an independent diffusive release calculation will be
starting. The particle failure function, which is input to the code, is actually a step function
with a maximum of 10 steps.

The procedure of calculation within a time step is performed in two consecutive parts:

1. Calculation of diffusive release from intact and defective particles and from the
graphite grains is conducted. The sum of those single releases is the overall source
term for the fuel element matrix graphite (grain boundaries).

2. Calculation of diffusive transport on graphite grain boundaries (quick phase) through


the matrix and release by desorption from fuel sphere surface into the coolant is
calculated.

3.1.2. Inventory Buildup

The diffusion calculation is conducted based on relative values. The inventory of a fission
product is built up during the normal operation phase according to its decay constant until
reaching 100% at the end of irradiation and beginning of the accident phase, respectively. For
long-lived species, the buildup is (almost) linear with irradiation time, while short-lived
species quickly reach the equilibrium state.

The simple assumption is that each radionuclide to be calculated is treated like a fission
product – characterized by its halflife – in a constant neutron flux, while ignoring additional
sources and sinks of that nuclide. This procedure is, e.g., not quite correct for activation
products and may be replaced with a more accurate one, or – ideally – with the results directly
taken over from a dedicated inventory calculation (ORIGEN, TNT).

3.1.3. Diffusion Model for Simulation of Fission Product Transport and Release
Behavior

Most of the currently used models are based upon the numerical solution of Fick's laws of
diffusion representing the "classical" approach for studying the fission product transport in the
particle kernel, the coating layers, and the matrix graphite of the fuel element as a function of
the temperature-time history during normal operation and under core heatup accident
conditions.

27
In general, the diffusion equation cannot be solved analytically. Therefore, numerical
approximations have to be used. For the spatial resolution, there are the Finite Difference and
the Finite Volume approximations and the Finite Element method. For digitizing the
evolution in time, the Crank-Nicolson approximation is frequently used that centers on the
mid-point between present and future step. Back-integration is also a possibility that centers
on the future time step; both methods require the inversion of a square positive definite
matrix. Forward integration – although computationally much simpler and numerically more
precise – is not possible with diffusion problems, because this leads to unacceptable
oscillations.

Whether or not a numerical diffusion approximation is acceptable, must be demonstrated by


verification and validation of the code.

The numerical solution starts from the diffusion differential equation in spherical coordinates,
which applies to both particle kernels, intact particles, graphite grains, and the fuel element.
The sphere considered is subdivided into N spherical shells of volume Vi and diffusion
coefficient Di, i=1,…N, and a constant source term Q. Concentration curve is assumed linear
in all sections except for the central one (i = N) where a square-shaped curve is assumed
(Fig. 3-1).

ci 1  ci
c  ci   r  ri 
ri 1  ri

and for the central element:

c N 1  c N c c
 r  rN   N 1 2 N  r  rN 
2
c  cN  2
rN rN

Figure 3-1. Partitioning of sphere in sub-shells (left), concentration curve in the sphere
(right).

28
The fission product mass flows ji are:

c N 1  c N
ji   2 DN 
 rN

ci 1  ci
jN   Di 
ri 1  ri

The mass balance for a volume element Vi is:

d .

dt Vi
c dV  Q Vi    c dV  ji Fi  ji 1 Fi 1
Vi

3.2. Key Parameters

3.2.1. Free Uranium and Heavy Metal Contamination

In a study conducted by Interatom [Ragoss 1989, Nabielek 2006] evaluating burn-leach tests
on as-manufactured spherical fuel elements (AVR-19, proof-test fuel) with TRISO-coated
LEU particles, the fraction of free uranium, Ufree/Utot, i.e., uranium not covered by an intact
SiC coating, was generally found to be either below or near the detection limit (typically
1×10-6), or a multiple of the uranium inventory of a single coated particle. These
measurements indicate that in most cases the measured free uranium fraction is due to
fabrication-induced defective particles, whereas the quasi-homogeneously distributed
contamination in the matrix material is comparatively small. (For more details see China
Report 3 on HTR-Modul.)

From an analysis of burn-leach measurements and also from the irradiation tests of the FRJ2-
KA series, it was concluded that the U contamination is at a very low level corresponding to
the natural contamination with heavy metals, while the Ufree is dominated by the defective
particles [Ragoss 1989]. The natural contamination of the matrix material with uranium is
existing in traces only close to the detection limit of 10-6; the fraction of fissile material
assuming a 10% enrichment is thus in the order of 10-7.

Irradiated AVR spheres compared to spheres irradiated in MTRs reveal a relatively larger
release fraction for cesium and iodine. This is due to cross-contamination in the AVR primary
circuit which was transported by the coolant and deposited and penetrated the fuel spheres
from the outside.

3.2.2. Initial Distribution in Accidents

The initial distribution of the specific fission product nuclide considered is equivalent to the
initial fissile uranium distribution. By far most of the fissile material is concentrated in the
particle kernels surrounded by an intact coating. The particles are homogeneously dispersed in
the fuel zone of the spherical fuel element. However, there are small fractions of uranium
outside the kernel in the coating and matrix graphite due to the manufacturing process or
natural contamination, respectively. The data given in Table 3-1 are typical to the German

29
reference fuel element and have been used as standard input in the calculations. The
complement to 100% of the sum of the above fractions is set as the initial kernel inventory.

Table 3-1. Uranium inventories outside the fuel kernels.


Uranium outside fuel kernel Inventory Fraction
U in buffer 1.0×10-3
U in iPyC 1.0×10-4
U in SiC 1.0×10-6
U in oPyC 1.0×10-6
U in matrix graphite 1.0×10-7

The natural contamination of the matrix material with uranium is existing in traces only close
to the detection limit of 10-6; the fraction of fissile material assuming a 10% enrichment is
thus in the order of 10-7. All inventories are assumed to be homogeneously distributed within
the respective material zone.

Uranium distribution is assumed to be equivalent to the distribution of the fission product


isotope considered. Furthermore, no chemical reactions are assumed among fission products.

3.2.3. Particle Failure Function

As already mentioned above, defective or failed particles consist of a particle kernel only so
that fission products are released from the kernel into the matrix graphite immediately. At the
moment of failure of a particle, fission product contents inside the coating layers is considered
to be released into the matrix graphite, while for the inventory still remaining in the kernel, it
is “time zero”, from which an independent diffusive release calculation is starting.

The assumption of a defective or failed particle in the diffusion code FRESCO-II is equivalent
to simulating an exposed particle kernel. This means that fission products released from the
particle kernel by diffusive transport are immediately released to the surrounding graphite
structure.

The failure of a particle coating, however, is not necessarily equivalent to a non-existing


coating; such an assumption is a conservative approach. There is experimental evidence for a
better retention of fission products in particles with a broken coating compared to exposed
kernels. The simulation of a partly existing coating as a kind of a 3rd type of particle (besides
intact particles and exposed kernels) has been proposed by JAEA [Sawa 1991]. A possibility
of modeling a broken but still existent coating is the choice of a smaller diffusion coefficient
in the kernel as an "effective" diffusion coefficient for this 3rd type of particle. An alternative
option is the increase of the SiC diffusion coefficient to simulate a cracked SiC layer while
PyC layers are still intact. The former approach was once pursued with the FRESCO-II code
by postcalculating the 1800°C isothermal heating test HFR-K3/3 trying to explain the high-
release peak of the observed IMGA bimodal distribution of single particle inventories
[Verfondern 1992]. But this procedure is arbitrarily imprecise, since no decent diffusion
coefficient could be derived.

30
3.2.4. Diffusion Coefficients

3.2.4.1. Reference Data

Fission product migration in coated particle and fuel element materials has been studied for
many years in many different laboratories, but the results are still not completely clear yet.
Main reason is the fact that any in-pile situation is too complex for a derivation of basic
material data and a basic understanding of the phenomena, while any simple out-of-pile
investigation of individual aspects produces as a result only one little piece of the whole
picture. Therefore, the use of conservative assumptions in order to bridge some existing gaps
in the knowledge of fission product behavior in HTGRs is necessary to conduct a successful
safety analysis [Nabielek 1974].

In a porous medium, the migration is occurring partly by diffusion and permeation in the gas
phase (pores) and partly by surface diffusion of adsorbed molecules along the pore walls with
an evaporation-adsorption exchange flux between the two phases. A strong adsorption has the
effect of depleting the pore gas of fission products, so that the effects of gas diffusion and gas
permeation become less important compared to surface diffusion and substantial amounts of
fission products can be stored in the solid phase. Diffusion is visualized as a series of random
jumps in a periodic lattice due to a number of simple mechanisms involving vacancies and
interstitials and also some more complex, combined mechanisms. Because of the large variety
of microscopic local geometries, it is not reasonable to specify this geometry in detail; any
(average) geometry assumed would produce similar results. A reasonable approximation is
the assumption of a spherical geometry for the crystallite grain. The significant scatter
observed between various experiments suggests the existence of some variability of the
grains’ sizes which is influenced by the fuel operating conditions [Nabielek 1974].

With regard to radionuclide transportation, certain key fission products are being considered
which can be classified into three groups of typical behavior [Flowers 1973, Nastri 1973]:

(a) Short-Lived Rare Gases (Kr, Xe)


Fission gases have very low diffusion coefficients in particle coating materials
and are therefore only released from broken fuel particles. Rare gas release
measurements in-pile and out-of-pile PIE annealing experiments in the case of
defective/failed fuel particles is provided by delay and decay in the UO2 kernel.
No adsorption or plate-out effects are observed for gases.

(b) Iodine
At normal operation temperatures, iodine behaves like a rare gas of equivalent
half-life meaning that a release can only occur from "as-manufactured" and "in-
service" broken fuel particles. There is only negligible delay and retention in
graphite, since noble gases and iodine are not adsorbed. However, there is a very
considerable attenuation of circulating iodine activity due to plate-out in the
primary circuit.

(c) Metallic Fission Products (Sr, Cs, Ag)


Some diffusional release of metallic fission products is observed to occur
through plain PyC (BISO) coatings, especially if derived from propylene instead
of methane, through defective TRISO particle coatings, and even to some extend
through intact particle coatings. The ability of these fission products to form
chemical compounds suggests a higher rate of migration especially in the grain

31
boundary volume but at the same time also an additional retention effect due to
chemical gettering. Furthermore, adsorption phenomena, particularly of cesium
and strontium, in the graphite components are significant.

In most cases, the diffusive transport of fission products is calculated assuming that the fuel
materials are homogeneous described by effective diffusion coefficients. The use of effective
diffusion coefficients, i.e., the assumption of the Fickian diffusion model, is a simplified
simulation of the transport mechanism. Their application will result in the same fractional
release of fission products from the fuel material as was observed in the measurements from
which the diffusion coefficients were derived. A comprehensive set of effective diffusion
coefficients has been established as a reference for key radionuclides in all fuel materials,
which is recommended to be used in safety analyses [IAEA 1997]. The data as currently
being used in German safety analyses are given in Table 3-2 and plotted in Fig. 3-2.

Table 3-2. Effective diffusion coefficients used in the calculations.


Diffusion Coefficient (a) in
UO2 PyC SiC Matrix
Cesium
Do,1 [m2/s] 5.6×10-8 6.3×10-8 5.5×10-14 × eΓ/5 3.6×10-4
(b)

Q1 [kJ/mol] 209 222 125 189


2 -4 -2
Do,2 [m /s] 5.2×10 0. 1.6×10 0.
Q2 [kJ/mol] 362 - 514 -
Strontium
Do,1 [m2/s] 2.2×10-3 2.3×10-6 1.2×10-9 1.0×10-2
Q1 [kJ/mol] 488 197 205 303
Do,2 [m2/s] 0. 0. 1.8×106 0.
Q2 [kJ/mol] - - 791 -
Silver
Do,1 [m2/s] 6.7×10-9 5.3×10-9 3.6×10-9 6.8×101
Q1 [kJ/mol] 165 154 215 262
Do,2 [m2/s] 0. 0. 0. 0.
Q2 [kJ/mol] - - - -
Iodine, fission gases
Do,1 [m2/s] 1.3×10-12 2.9×10-8 - 6×10-6 6×10-18
Q1 [kJ/mol] 126 291 - 0. 71
2 -15
Do,2 [m /s] 8.8×10 0. - in in
pores grains
Q2 [kJ/mol] 54 - -
(a) Effective diffusion coefficient: D = Do,1 · exp[-Q1/(R·T)] + Do,2 · exp[-Q2/(R·T)]
(b) Γ is the fast neutron fluence in 1025 n/m2, E>0.1MeV

32
1.0E-04

cesium
1.0E-06
strontium
1.0E-08
silver
Diffusion Coefficient [m2/s]

1.0E-10
fission gases,
iodine AC
1.0E-12
fission gases,
iodine NO
1.0E-14

1.0E-16

1.0E-18

1.0E-20

1.0E-22
4.00 4.50 5.00 5.50 6.00 6.50 7.00 7.50 8.00 8.50 9.00 9.50 10.00 10.50

1.0E+04/T

Figure 3-2a. Reference diffusion coefficients in UO2.


1.0E-04

cesium
1.0E-06

strontium
1.0E-08
Diffusion Coefficient [m2/s]

1.0E-10 silver

1.0E-12 fission gases,


iodine

1.0E-14

1.0E-16

1.0E-18

1.0E-20

1.0E-22
4.00 4.50 5.00 5.50 6.00 6.50 7.00 7.50 8.00 8.50 9.00 9.50 10.00 10.50

1.0E+04/T

Figure 3-2b. Reference diffusion coefficients in pyrocarbon.

33
1.0E-04

cesium
1.0E-06

strontium
1.0E-08
Diffusion Coefficient [m2/s]

1.0E-10 silver

1.0E-12 fission gases,


iodine

1.0E-14

1.0E-16

1.0E-18

1.0E-20

1.0E-22
4.00 4.50 5.00 5.50 6.00 6.50 7.00 7.50 8.00 8.50 9.00 9.50 10.00 10.50

1.0E+04/T

Figure 3-2c. Reference diffusion coefficients in silicon carbide


(cesium diffusion coefficient is based on Γ=2).
1.0E-04
cesium

1.0E-06 strontium

silver
1.0E-08
fission gases,
Diffusion Coefficient [m2/s]

iodine in pores
1.0E-10
fission gases,
iodine in grains
1.0E-12

1.0E-14

1.0E-16

1.0E-18

1.0E-20

1.0E-22
4.00 4.50 5.00 5.50 6.00 6.50 7.00 7.50 8.00 8.50 9.00 9.50 10.00 10.50

1.0E+04/T

Figure 3-2d. Reference diffusion coefficients in matrix graphite.

34
3.2.4.2. History of Reference Data

Diffusion coefficients have been measured by means of various experimental methods, but
have also been derived from the calculational evaluation of numerous irradiation and heating
experiments with complete spherical fuel elements, fuel compacts, single fuel particles, or
graphite samples.

In a joint effort of the Research Center Jülich and German industries within the project
“Hochtemperaturreaktor-Brennstoff'kreislauf” (HBK), a set of data to be used for predicting
fission product transport during HTGR normal operation has been established. The data were
annually revised and published in the HBK project reports from 1978, until the project was
terminated in 1985 [HBK 1986]. Purpose of this data set was to summarize the data base for
HTGR design calculations for licensing procedure. The comprehensive data collection created
in the 1970s by UKAEA AERE Harwell with the OECD Dragon Project and summarized in
the so-called "Red Book" [UKAEA 1979] served as a basis. The HBK data set was extended
by additional recommendations to apply to core heatup accident conditions. Later, additional
or new recommendations derived from experiments were given. The last and most
comprehensive collection comprises the actual data as are being used in the various countries
participating in the IAEA CRP-2; it is published in the respective technical document [IAEA
1997].

Data for UO2

It was early found that significant kernel retention exists in oxide kernels for Sr because of
SrO and/or SrZrO3 formation, but no significant kernel retention was found for Sr in carbidic
fuel kernels. The same observations show also clearly that no significant kernel retention
exists for Ag in either oxide or carbide kernels [Nabielek 1974].

Diffusion coefficients in UO2 fissile particle kernels which were used in German safety
analyses were derived from irradiation and heating experiments with designed-to-fail UO2
particles (kernel + buffer layer). According to an evaluation of the German irradiation and
heating experiment FRJ2-P28, cesium (see Figs. 3-2a and 6-64) and iodine release (see Figs.
3-2a and 6-65) under accident conditions is considered at FZJ to be best approximated by a
two-branch diffusion coefficient (see chapter 6.2.5) while those for strontium and silver were
modified by a factor of 20 and 10, respectively, compared to the recommendation of the last
HBK data set. A different recommendation is given for iodine at lower temperatures
(< 1000°C) which was derived from irradiation experiments in the R2 reactor at Studsvik
[Müller 1976].

Data for Pyrocarbon

Pyrocarbon is an extremely complex and flexible material with regard to its microstructure. It
consists of a practically infinite range of molecular structures made of carbon and hydrogen
atoms. In addition to crystallites and more amorphous phases, the material contains
microcracks and micropores. Several modes of diffusion (grain, grain boundary, surface and
gaseous diffusion) as well as trapping and adsorption processes therefore occur
simultaneously. Any diffusion coefficient values should therefore be called "effective" and
imply that the overall migration process can be phenomenologically described by Fick' s laws.

35
This holds particularly for cesium. Migration in terms of diffusion and adsorption shows a
strong correlation with PyC characteristic variables [Nabielek 1974].

Diffusion coefficients in pyrocarbon were mostly derived from early experimental data with
LTI- and HTI-PyC from BISO particles. A rapid diffusion through PyC is assumed at
temperatures above 1900°C. Recommended diffusion coefficients are those from the HBK
data set (see Fig. 3-2b). The diffusion coefficient relation given for fission gases/iodine was
derived from GA heating data at very high temperatures; it is, however, not of importance in
the as-is FRESCO version since at least the SiC layer is considered gas-tight not allowing any
gases/iodine to be released from intact coated particles.

The carbonaceous buffer layer is assumed in the diffusion model to have no significant
retention capability for fission products. In calculations, a constant diffusion coefficient of
10-8 m2/s is assumed which allows for a rapid diffusion.

Data for Silicon Carbide

SiC is by comparison to PyC a material of much simpler microstructure and also of much less
microstructural variation and scatter. Especially as its density (as fabricated) approaches the
theoretical maximum of 3.210 Mg/m3 there is much less possibility for porosity, internal
surface areas and microcracks than in PyC and the range of possible migration phenomena is
mainly limited to lattice and grain boundary diffusion. Accordingly, much lower diffusion
coefficients and adsorption coefficients are expected than for PyC. However, similar to PyC,
we still see for SiC the same complications due to material variations from producer to
producer, between discs and particles, and with deposition conditions may occur [Nabielek
1974].

The permeability of the SiC coating layer for fission products, particularly cesium, was
investigated more than any other relevant transport data. The Arrhenius relation for cesium
proposed by Allelein [Allelein 1980] and derived from heating tests on particles became part
of the HBK data set as recommended data for use in HTGR core release predictions under
normal operating conditions and was still the recommendation to be used in German safety
calculations [Moormann 1987].

For irradiation experiments with modern HTGR fuel, the Allelein diffusion coefficient was
found to overestimate in some cases the cesium release fraction from fuel particles. An
evaluation of the HFR-K3 experimental data in [Christ 1985] included the fast neutron
fluence as a parameter for the Arrhenius relation. This new recommendation for the low-
temperature branch of the diffusion coefficient is – for a fast fluence of 2×1025 n/m2,
E>0.1MeV – at a temperature of, e.g., 1600°C lower by a factor of ~8 compared to the
Allelein data. The HRB relation was fixed as a new recommendation in [Verfondern 1993]
and remained unchanged in [IAEA 1997].

In the accident temperature range, a theoretical evaluation of several researchers' experimental


cesium release data from coated particles was made by Myers in the early 1980s. The result
was a diffusion coefficient subdivided into a low temperature branch and a high temperature
branch dominant at > 1600°C. The data at high temperatures were likewise subdivided and
identified to characterize two different types of silicon carbide material [Myers 1984]. The
Myers upper curve was chosen in 1986 as the reference data for accident temperature
conditions.

36
The experimental results from the KÜFA heating tests with modern spherical fuel elements
starting in 1984 showed a distinct discrepancy in the release behavior from loose coated
particles and from complete fuel spheres. Due to an optimized fuel manufacturing process
after 1985, the considerable retentivity of the matrix graphite for metallic fission products,
and the much better statistics of about 104 particles per fuel element, experiments with loose
particles were no longer regarded as representative of modern high quality fuel.

This experience coincides with the fact that postcalculations of cesium release from heated
fuel elements using the Myers diffusion coefficient have resulted in an overestimation by up
to several orders of magnitude. The evaluation of the new KÜFA data using the FRESCO-II
diffusion code (see also chapter 6.2.4) has led to a new KFA recommendation for a diffusion
coefficient in silicon carbide [Verfondern 1989, Verfondern 1991] which remained part of
[IAEA 1997].

Use of the wrong diffusion coefficient in FRESCO for cesium in SiC as was discovered in
CRP-6 calculations originated from the historic coding, at a time when – for accident
conditions – the reference diffusion coefficient was “Myers” with a low-temperature and
high-temperature branch. The “Verfondern”-1989 was meant to give a new recommendation
for the high-temperature branch. The mistake was to still leave the low-temperature “Myers”
curve in the code, while there was the “HRB” curve for normal operation. As a consequence,
FRESCO coding of diffusion data should be modified such that – supposed there is a two-
branch relation – it should be with regard to temperature; there should be no longer an
additional distinction between normal operation and accident.

In analogy to cesium, an evaluation was made for strontium on the basis of 11 KÜFA heating
tests in the temperature range between 1600 and 1800°C resulting in a new diffusion
coefficient [Verfondern 1991] which is at 1600°C lower by a factor of about 20 compared to
the HBK recommendation. However, the combination of this new high-temperature branch
with the old HBK relation for the lower temperatures to a single two-branch diffusion
coefficient does not benefit from this potential for reduced release (see chapter 6.2.4.4). This
is why postcalculations of Sr release still overestimate the experimental data [IAEA 2011].
Therefore more strontium release data at lower temperatures should be made available.

The case that really needs more experimental study and better understanding is the large
release of silver (110mAg) through apparently intact SiC coating layers which has been
observed above 1250°C. It is suspected that this is a structure sensitive property which is
shown up particularly by silver since this element is also not retained by pyrocarbon.
Recommended silver transport data in [IAEA 1997] taken from the HBK data set are based on
a 1500°C heating test at Harwell. An evaluation of available Ag release data from KÜFA tests
rendered difficult since no consistent release behavior was observed and is in many cases not
reproducible by any existing model. The recent evaluation presented in [Van der Merwe
2009] for the lower temperature range appears to be a decent diffusion coefficient valid for
modern HTGR fuel and may become part of a new recommended data set.

Data for Matrix Material

The choice of material and design for the graphite components of a fuel element like the
matrix, a relatively porous and permeable graphite structure with gaps and vents and possibly
also cracks developing upon irradiation, causes a complex release situation.

37
Effective diffusion data are relatively well known from measurements at the HMI, Berlin, for
cesium [Hoinkis 1983], strontium [Hensel 1991] and silver [Hoinkis 1983, Hoinkis 1994] in
German A3-3 matrix graphite in the Henrian concentration regime as well as the influence of
irradiation and corrosion on the diffusive process. The Germany data in [IAEA 1997] refer to
the more conservative irradiated A3-3 matrix graphite. According to the experience from
measurements, it is further assumed that cesium and strontium proceed in matrix rapidly at
temperatures beyond 2000°C, for silver already at temperatures above 1000°C.

For strontium, no significant deviations from Fick's laws have been observed and no
discrepancy between laboratory measurements and PIE of irradiated fuel elements. Its strong
adsorption on matrix is so effective that practically the only release of 90Sr from the core is
due to gaseous precursor decay (90Kr). In contrast, for cesium, significant discrepancies
between in-pile experimental data and those from laboratory experiments have been observed.
Also significant deviations from Fick's laws like profiles consisting of several components
and contradictions between profile shapes and relased or transmitted amounts have been
observed. The data suggest that Cs migrates in graphite via all sorts of transport mechanisms
[Nabielek 1974].

A trapping mechanism was found by the HMI to describe more realistically the cesium
transport process in graphite [Hensel 1995]. The diffusion/trapping/re-emission model
assumes that migrating particles are captured in traps for some time, before they are re-
emitted and become mobile again. Traps are uniformly distributed, predominantly in the
polymeric binder carbon (due to its numerous adsorption sites), and do not lose their trapping
character if occupied. It was observed, however, that the traps for cesium disappear, if the
matrix material was slightly oxidized (up to a weight loss of 0.3%) due to opening of pores. A
calculation study has shown that the cesium release from the core cavern is by a factor of
maximum 8 larger, if diffusion trapping is taken into account, but still remains with << 10-8
insignificant from the radiological point of view. Therefore the use of effective diffusion
coefficients may be justified.

The systematic evaluation of heated spheres, for which cesium inventory measurements in the
matrix were made available, has demonstrated that the diffusion coefficient in A3-27 is at
least one order of magnitude lower compared to A3-3 in the accident temperature range of
1600 to 1800°C. But besides the diffusion coefficient in matrix, there is also the heavy metal
contamination fraction in the matrix which is an important input for the model calculations
with a major effect on the release from the fuel element, especially at lower heating
temperatures when the release from the coated particles is comparably small, and especially
for strontium which is more strongly bound in graphite than is cesium. An adequate
recommendation for cesium and strontium transport in A3-27 matrix graphite, of which the
spheres of most heating tests were made, may be the reduction of the Hoinkis diffusion
coefficient curves (for A3-3) by one order of magnitude which would still envelope the
calculated results [Verfondern 1991] (see chapter 6.2.4.4). But reference remains the A3-3
matrix data as given in [IAEA 1997].

The rare fission gases (Kr, Xe) and also iodine are not adsorbed on graphite at normal
operation temperatures, therefore, they migrate mainly by in-pore gaseous diffusion with no
significant delay within the graphite components and by the various helium convection
processes. With regard to iodine, certain retention in graphite materials has been identified in
the lower temperature range. Based on the perception that the heavy metal contamination is
buried in the graphite grains, the iodine release is connected with a comparatively slow
transport process outwards to the grain boundaries (or pores), from where there will be a fast

38
transport to the fuel element surface and the coolant, respectively. The fast transport
mechanism for iodine and fission gases is assumed to apply also to all iodine that is being
released from the defective/failed fuel particles into the matrix. Other experimental findings
such as the influence of fast neutron fluence, fission product concentration, porosity, the
degree of oxidation, or materials grades were often tried to be incorporated into empirical
relationships [IAEA 1997, GA 1987] but are currently not part of the recommended transport
data.

The transport of iodine and fission gases in graphite was the subject of a special series of
experiments at FZJ [Schenk 1990] where spherical fuel elements were activated to low
burnup and then heated to investigate the release behavior of short-lived isotopes originating
from the heavy metal contamination. Iodine release curves were found to be similar to those
for xenon (see chapter 5.2.7).

3.2.5. Sorption Isotherms

The transition of metallic fission products from the fuel element surface into the coolant
(desorption) or vice versa (adsorption) is determined by sorption isotherms giving the partial
pressures of a species above a graphite (or metal) surface as a function of temperature and the
concentration in the graphite near the surface.

At low radionuclide concentrations, the so-called “Henry regime”, there is a direct


proportionality between vapor pressure and concentration of the adsorbed species with a
constant heat of adsorption, whereas at higher concentrations, the so-called “Freundlich
regime”, the heat of adsorption is decreasing with increasing concentration of the adsorbed
species. Both regimes are separated at the transition concentration ct. The fission product
concentration adsorbed on the surface is in equilibrium with the partial vapor pressure, pH or
pF, above the surface:

 B  E
ln p F   A     D    ln c
 T  T

 B  E
ln p H   A     D  1    ln ct  ln c
 T  T

p  pF  pH

ln ct  d1  d 2  T

where pF is the vapor pressure in the Freundlich regime [Pa], pH is the vapor pressure in the
Henry regime [Pa], T is the temperature [K], c is the concentration of sorbed species
[mmol/kg of graphite], and A, D, d1 are constants as well as B, E [K] and d2 [K-1].

Sorption isotherms have been measured at the Research Center Jülich for both cesium,
strontium, silver, and iodine on matrix graphites. Experiments were carried out applying two
different methods:

 high temperature mass spectrometry with a Knudsen cell

39
 isopiestic method

Two systems based on mass spectrometry were used, a single-focussing sector-type


instrument and a double Knudsen cell. The Knudsen cells were made of tungsten and
molybdenum which do not interact chemically with the species examined or the graphite
during the measurements. Pressures down to 10-7 could be determined in the cell [Hilpert
1985].

The isopiestic method employs a high-vacuum furnace having inside a cylindrical Mo cell
sized 40 mm both in length and inner diameter. The cell containing various samples is closed
at high temperature and high vacuum. Loading of the graphite samples with a species takes
place via the gaseous phase until equilibrium is reached. This method is suitable for studying
the influence of, e.g., graphitization or BET surface in comparison to a reference sample
[Hilpert 1983].

Sorption isotherms for cesium, strontium, silver, and iodine have been measured for a variety
of nuclear graphites and matrix materials, and also metals. The data are summarized in [IAEA
1997]. Unirradiated A3-3 matrix powder (size between 40-80 μm) loaded with a certain
concentration of Cs or Sr was given into the Knudsen cell, and the partial pressure of Cs or Sr
was measured as a function of temperature. The concentrations selected in a test were in the
range of both Henry and Freundlich regimes; temperatures were ranging up to ~1500°C for
cesium and ~1700 for strontium.

The sorption isotherms derived from these experiments were slightly modified to become
applicable to the higher temperature range up to 2300°C and given an additional uncertainty
range to be then applied in safety analyses [IAEA 1997]. Plots of these sorption isotherms are
presented in Fig. 3-3, left, for cesium and in Fig. 3-3, right, for strontium. Each vertical line
represents a test and indicates the concentration selected and temperature range examined.
The results exhibit for strontium a much stronger sorption capability on A3-3 matrix
compared to cesium.

Only few data are available for silver and iodine giving evidence of a much weaker sorption
on graphite for both species (Fig. 3-4). This is why a sorption effect of silver and iodine was
never considered in German safety analyses.

An important observation resulted from the sorption experiments with the isopiestic method
when examining the single constituents of A3 matrix. Comparing the sorption capabilities of
natural graphite, graphitized coke, and the coked resin binder, it could be shown that by far
most strontium or cesium was bound on the resin binder (96%) which is a highly porous, non-
graphitized material [Hilpert 1985]. Similar results were obtained for cesium.

Structural graphites, such as the German reference reflector graphites ATR-2E and ASR-1RS
are more graphitized due to high temperature treatment compared to matrix material and
therefore have a reduced sorption capability, which was demonstrated in an isopiestic test
[Hilpert 1985].

40
Figure 3-3. Sorption isotherms for cesium (left) and strontium (right) over A3 matrix.

Sorption investigations were usually conducted for single species. For HTGR release
predictions, however, multiple species sorption or co-sorption should be taken into account.
The competition of different species for sorption sites in the graphite can lead to a reduction
of the overall sorption capability of a single species. This effect was examined in more detail,
e.g., for the systems Cs-Rb, Cs-Ba, or Cs-Sr. One approach for the conservative treatment of
this effect as was used in German safety analyses was to consider the sum of all inventories of
chemically similar species as the “inventory” input of a species relevant for calculating the
sorption effect.

Applying the FRESCO-II code means that only one fuel sphere is considered. The sorption
effect here is just desorption leading to a minor reduction of the fuel element release, but
would not take into account the adsorption of cesium from the coolant that might have come
from other fuel spheres.

41
Figure 3-4. Comparison of sorption isotherms for cesium, strontium, and iodine
over A3 matrix.

The effect of the sorption process was demonstrated in a calculation study considering a core
heatup accident scenario for the THTR-300 for two different accident temperature histories
with maximum fuel temperatures of 2080°C and 2360°C, respectively. Figure 3-5 shows the
comparison of FRESCO-I calculational results for cesium release from the coated particles
(blue), from the core cavern (red), and the inventory in the top reflector graphite (green). In
the case of lower core temperatures (dashed lines), the matrix graphite provides a very high
potential of cesium sorption supported by the top reflector whose inventory gradually
increases. Therefore the release from the core cavern remains on a low level over 280 h,
before it eventually increases. In contrast, in the case of higher core temperatures (solid lines),
the Cs release from the coated particles is so high that concentrations in the fuel and reflector

42
graphite soon reach its capacity limits such that all additional Cs released from the fuel
directly contributes to the release from the core cavern.

For practical calculations within safety analyses for small modular HTGRs, the sorption effect
is typically neglected for conservative reasons. If it were assumed for the pebble-bed (i.e., in a
FRESCO-I calculation), it would result in a practically zero core release because of the
enormous sorption potential from the large available surfaces of relatively cold graphite in
most parts of the pebble-bed.

Figure 3-5. Prediction of cesium release behavior in the THTR-300 core for two different
accident temperature distributions.

3.2.6. Fuel Temperature

In FRESCO calculations, it is assumed that the temperature at a given moment in time is


spatially constant throughout the sphere and the coated particles, respectively. This
assumption is justified under core heatupaccident conditions (reactor shut-down), when a
constant temperature level inside a fuel sphere after approximately 20 minutes after shut-
down is reached. The assumption, however, is not quite correct for the normal operation/
irradiation phase. Power producing fuel spheres develop a temperature profile inside the
sphere with a steadily increasing temperature to the center. This effect should be taken into
account in a future improved version of the FRESCO model.

3.3. Special Treatment of Fission Gas and Iodine Release under Accident Conditions

The transport behavior of iodine in graphite under accident conditions is considered in


German model calculations to be similar to that of the noble gases krypton and xenon. This is
in agreement with observations from heating experiments where both iodine and fission gas

43
release was found to be in direct correlation with the presence of defective/failed particles
[Schenk 1990]. Rather than simulating an effective, one-phase diffusion, the iodine transport
is conservatively treated in the KFA reference modeling to consist of a slow diffusion phase
out of the graphite grain [Müller 1976] at temperatures ≤ 1250°C and of a rapid diffusion
phase via the pores and the graphite grain boundaries. The iodine from the heavy metal
contamination is assumed to be buried deep inside the grains. In contrast, iodine released from
defective particles into the graphite is assumed to be immediately transported in the graphite
pores. The quick phase transport is assumed independent of temperature and corresponds to
the diffusive behavior for xenon in helium, and does not allow for any realistic retention in the
graphite.

Approach for assessing 131I release:

(a) during normal operation


Booth formula applied to both particle kernel and graphite grain; then multiply with
particle failure fraction and heavy metal contamination, respectively;

(b) during core heatup accident:


Two separate FRESCO calculations, one simulating the “normal” coated fuel particle
with given failure function and assuming zero heavy metal contamination, the second
one simulating a graphite grain, then combining both contributions to an overall
release.

3.4. Limitations of the FRESCO-II Code

In the course of FRESCO code applications in the past, a couple of deficiencies have been
encountered which are recommended to be considered in the development of future updates or
new code systems. They will be shortly summarized in the conclusions, and are explained in
more detail in China Report 7.

44
4. CRP-6 ACCIDENT CONDITION BENCHMARK (2010)

4.1. Background

In the IAEA Coordinated Research Project CRP-6 on “Advanced in HTGR Fuel


Technology”, important parts were two benchmark exercises, a first one on fuel performance
modeling under normal operating conditions (see China Report 1 on PANAMA V&V) and a
second one on fission product release under accident conditions. These benchmarks which
included both code-to-code and code-to-experiment comparisons have demonstrated the
enormous progress on computer model development in various countries in recent years
[Verfondern 2005, Phelip 2006].

Most of these more recent modeling approaches to describe fuel performance during normal
operation and under accident conditions are more sophisticated (than PANAMA) in that they
simulate the complete TRISO coated particle and include the analysis of mechanical and
thermal stresses acting inside the particle coating. The result of the codes’ calculations is the
quantification of the particle failure probability with most codes including the subsequent
simulation of radionuclide transport and release.

The accident condition benchmark exercise was composed of three parts:

(i) a sensitivity study to examine fission product release from a fuel particle starting
with a bare kernel and ending with an irradiated TRISO particle for verification
purposes;

(ii) the postcalculation of well documented irradiation and heating experiments for
validation purposes; and

(iii) predictive calculations for heating tests which were planned in future.

A total of eight codes from France, Germany, Korea, South Africa, and the USA have been
applied to all or a part of the proposed benchmark cases. From the historic perspective, the
comparison is between three “old” (Germany, South Africa, USA/GA) and five newly
developed models. The codes were:

ATLAS France
FRESCO-II Germany
COPA Republic of Korea
GETTER South Africa
PARFUME USA (Idaho National Laboratory)
SORS USA (General Atomics)
NRCDIF USA (Nuclear Regulatory Commission)
MELDIF USA (Sandia National Laboratory)

Depending on the code development stage and case definition, codes were applied to the
various cases as appropriate.

45
4.2. General Conditions

The general characteristics of the fuel spheres/compacts and the coated particles, respectively,
considered in the different calculation cases are described in detail in [IAEA 2011]. For the
cases of the sensitivity study, basically the nominal data of the German reference fuel element
were assumed.

Main input data for the codes are, apart from the fuel design, the initial fuel distribution, the
transport data, the amount of failed particles, and the temperature-time history of the fuel. All
models all principally based on the diffusive transport of radionuclides through the coated
particles and the fuel element determined by solving the simple Fickian diffusion equation in
discrete steps of time and location. Differences are mainly given in the numerical solution
methods and the choice for number of nodes and time step widths. Effective diffusion
coefficients in form of an Arrhenius relationship for all materials are applied. While the
assumption of particle failure for the sensitivity part of the benchmark is obsolete, the
transient failure fractions for the real heating experiments are considered according to the
observed fission gas release during the tests.

The diffusive transport of fission products (cesium, strontium, silver, fission gases) is
calculated assuming that the fuel materials are homogeneous. Therefore, effective diffusion
coefficients are used in code calculations. The set of data to be applied corresponds to the
diffusion coefficients for “Germany” as listed in [IAEA 1997]. No adjustment was made in
these benchmark calculations to obtain optimal agreement with the measurements.

Since the heated spheres considered here were not in a reactor environment, the assumption
was that no credit should be taken from the sorption effect and that rather transition of the
fission products from the fuel element surface into the coolant is unhindered.

4.3. Results

4.3.1. Sensitivity Study

The cases 1a, 1b refer to cesium release from a fuel kernel at two different temperatures,
while the cases 2a, 2b consider the same heating conditions, but for a particle consisting of
kernel, buffer, and dense PyC layer. All other cases of the sensitivity study are based on a
complete TRISO coated particle.

In the five cases 3a-e, the TRISO particle is exposed to heating temperatures of 1600°C and
1800°C. Distinction is made with regard to a broken SiC layer (3d) and a through-coating
failure (3e). The four cases 4a-d correspond to the previous cases 3a-c, 3e with the difference
of assuming now a realistic irradiation phase of 500 efpd at 1000°C preceding the heating
phase. The two cases 5a, 5b finally are focusing on a modified irradiation phase assuming 10
temperature cycles to simulate the multi-pass feature of fuel in a pebble-bed reactor.

4.3.1.1. Case 1 – Fuel Kernel

The simplest case is the release from a spherical particle kernel with homogeneously
distributed fission products. This case can also easily be calculated analytically by applying

46
the fractional release term derived from the “equivalent sphere model” [Nabielek 1974] (see
chapter 2.0.0). The calculated fractional release values for 137Cs at the end of the 200-hour
heating phase at 1200°C (case 1a) and 1600°C (case 1b), respectively, are listed in Table 4-1.

Table 4-1. Fractional release of 137Cs after 200 h for case 1.


Participant Fractional release of 137Cs from a bare kernel
Case 1a (1200°C) Case 1b (1600°C)
France 0.472 1.000
Germany 0.456 1.000
Korea 0.473 1.000
South Africa 0.498 1.000
US/GA 0.453 0.970
US/INL 0.467 1.000
US/NRC 0.463 0.998
US/SNL 0.465 1.000
Analytical solution 0.4673 0.99999959

The comparison of the fractional cesium release results shows that most codes come close to
the analytical solution. The three “old” codes (FRESCO-II, GETTER, GA code) are a little bit
off in the 1200°C case, although it is only about 2% deviation.

1.E+00

8.E-01
Fractional Release Cesium

6.E-01

4.E-01

2.E-01

0.E+00
0 20 40 60 80 100 120 140 160 180 200
Heating Time [h]

Figure 4-1. FRESCO-II transient fractional release of 137Cs for case 1 (red, black, white
curves) and case 2 (blue, green curves) in linear scale [Verfondern 2008].

47
One problem in FRESCO-II has been identified to be the limited number of spatial
discretization steps for the diffusion calculation. Figure 4-1 shows the transient cesium release
calculated with FRESCO-II for 1200°C (lower red curve) and 1600°C (upper red curve). Both
red curves are sandwiched between a black curve above representing the analytical solution
and a white curve below, where a smaller number of nodes in the kernel was chosen (35 as
standard vs. 150 for case 1 here). There is also a deviation in the 1600°C case, which is
clearly visible especially in the first 50 hours of heating.

4.3.1.2. Case 2 – Fuel Kernel + Buffer + PyC

The results of the calculations for case 2, cesium release from a particle kernel plus buffer
plus dense PyC layer after 200 h heating at 1200°C and 1600°C, respectively, are listed in
Table 4-2.

Table 4-2. Fractional release of 137Cs after 200 h for case 2.


Participant Fractional release of 137Cs from a
bare kernel + buffer + pyrocarbon layer
Case 2a (1200°C) Case 2b (1600°C)
France 0.028 0.995
Germany 0.026 0.991
Korea 0.029 0.995
South Africa 0.030 0.993
US/GA 0.006 0.968
US/INL 0.026 0.996
US/NRC 0.026 0.989
US/SNL 0.026 0.995

The comparison shows that most codes arrive at a fractional release of 2.6-2.9% in the
1200°C case and more than 99% for the 1600°C case. The transient release curves as
calculated with FRESCO-II were already included in Fig. 4-1, the green curve for 1200°C and
the blue curve for 1600°C. Their respective differences to the red curves represent the
retention potential of the buffer and PyC layers during heating showing a certain delay in the
release.

4.3.1.3. Case 3 – TRISO Coated Particle

The results of the calculations for case 3 are listed in Table 4-3 considering cesium release
from a TRISO coated particle after 200 h heating at 1600°C (3a) and 1800°C (3b), heating at
1600°C following by 1800°C (3c), in addition assuming SiC crack at 1800°C (3d), and finally
assuming SiC failure from the beginning plus through coating failure at 1800°C (3e). The
through coating failure in the last case allows also the calculation of fission gas release.

48
Table 4-3. Fractional release after 200/400 h for case 3.
Participant Fractional release from a TRISO coated particle
Case 3a Case 3b Case 3c Case 3d Case 3e
137 137 137 137 137
Cs Cs Cs Cs Cs I, gases
after 200 h after 400 h
-5
France 6.59×10 0.207 0.222 0.999 0.97 0.98
-4
Germany 2.06×10 0.218 0.239 1.000 1.00 1.00
-4
Korea 4.72×10 0.210 0.224 1.000 1.00 1.00
South Africa 1.14×10-4 0.203 0.230 1.000 1.00 1.00
US/INL 1.32×10-4 0.208 0.222 1.000 1.00 1.00
US/NRC 1.25×10-4 0.207 0.22
US/SNL 1.00×10-4 0.208

The fractional release curves as calculated with FRESCO-II are given in Fig. 4-2 for all five
cases 3a-e. For the first 200 h, the red and orange curves (identical) represent heating at
1600°C (3a, 3c, 3d), the blue curve heating at 1800°C (3b), and the green curve heating at
1600°C, but with a failed SiC layer. This latter case (3e) is similar to the previous case 2b,
both heating at 1600°C, but with the difference that the cesium has to pass now two dense
PyC layers.

1.E+00

1.E-01
Fractional Release Cesium

1.E-02

1.E-03

1.E-04

1.E-05
TRISO Cs 1800
1.E-06 TRISO Cs 1600 + 1800
TRISO Cs 16-800 failed SiC at 1800
TRISO Cs failed SiC at 1600 total at 1800
1.E-07 TRISO I failed SiC at 1600 total at 1800

1.E-08
0 50 100 150 200 250 300 350 400
Heating Time [h]
Figure 4-2. FRESCO-II transient fractional release of 137Cs for case 3 [Verfondern 2008].

49
For the second 200 h, which is heating at 1800°C, the red curve represents the intact TRISO
particle (3c), while the orange curve is based on the assumption of a failed SiC (3d), and the
through coating failure (3e, green curve) is not visible since release fraction has reached
already 100%.

Case 3a represents the “classical” 1600°C accident case for a TRISO coated particle.
According to Minato [Minato 1995], the fractional release of cesium after 200 h at 1600°C
should be 2.1×10-4. The calculations here exhibited in first approach surprisingly diverging
fractional release values after 200 h for the different models. In particular, the result of the
(first) FRESCO calculation was 1.15×10-3, the highest release among all codes. This raised
discussion about whether or not the same diffusion coefficient in SiC was applied. The
recommended diffusion coefficient of cesium in SiC as generally understood from [IAEA
1997] is composed of a low-temperature branch for irradiation/normal operation derived by
HRB [Christ 1985] and a high-temperature branch for heating/ accident conditions derived by
FZJ [Verfondern 1989]. The combined diffusion coefficient is:

14    125,000  2  514,000  m2


D  5.5  10 exp  exp    1.6  10 exp  
5  RT   RT  s

The diffusion coefficient at 1600°C (and Γ = 2) should then be:

D = 1.011×10-16 m/s2

A re-check has revealed that the original FRESCO coding, although utilizing the “HRB”
curve for normal operation conditions, included for accident conditions still the low-
temperature branch of the previous recommendation by Myers combined with the today’s
recommendation by FZJ for the high-temperature branch. The combination of “FZJ” plus
“Myers” at 1600°C leads to a higher value of

D = 1.477×10-16 m/s2

A new FRESCO calculation with the corrected diffusion coefficient eventually resulted in a
fractional release of 2.01×10-4, much closer to the other results. Most codes reach the Minato
value within a factor of 2 or less. For the 1800°C heating over 200 h, all codes predict a
release fraction of ~21%.

In case 3c, the combined 1600/1800°C heating, the predicted result should be 0.236. It is also
pretty well met by all codes. Cases 3d and 3e assuming failed SiC and through coating failure,
respectively, will certainly result in a practically complete release of cesium, which is
predicted by all codes. In case 3e assuming a through coating failure with beginning of the
1800°C heating phase, fission gases will be released also. The black curve in Fig. 4-2
indicates a spontaneous, steep increase of the fractional release reaching very soon 100%.

4.3.1.4. Case 4 – TRISO Coated Particle + Irradiation Phase

Case 4 also treats a complete TRISO particle, but now includes a preceding irradiation
history. Fractional release results at the end of the heating phase are listed in Table 4-4 for
both 137Cs and 110mAg. In Fig. 4-3, the transient release curves as calculated with FRESCO-II
are given for cesium, silver, and fission gases/iodine.

50
Assuming a crack of the SiC layer at 1600°C, i.e., with beginning of the heating phase,
cesium release will immediately steeply rise, since the retention in the still existing PyC
layers is small at 1600°C. The additional assumption of cracked PyC layers at 1800°C merely
has a further influence on the cesium release from the particle. It does, however, decisively
influence the release of fission gases, since they will escape the particle only upon a through-
coating failure as is shown with the black curve in Fig. 4-3 starting at time 200 h.

Table 4-4. Fractional release after 200/400 h for case 4.


Participant Fractional release from an irradiated TRISO coated particle
Case 4a Case 4b Case 4c Case 4d
after 200 h after 400 h
137
Cs
France 2.55×10-4 0.201 0.21 1.00
Germany 5.00×10-4 0.220 0.24 1.00
Korea 8.25×10-4 0.215 0.23 1.00
-4
South Africa 1.64×10 0.206 0.23 1.00
-4
US/INL 4.10×10 0.226 0.23 1.00
110m
Ag
France 0.27 0.58 0.98 0.98
Germany 0.41 0.87 0.92 1.00
Korea 0.55 0.95 0.98 1.00
South Africa 0.42 0.88 0.93 1.00
US/INL 0.43 0.89 0.93 1.00

Unlike the case 3 with no significant irradiation phase, the diffusion process from the kernel
into the coating over 500 days at 1000°C has further proceeded into the coating with fission
products being sooner released during the heating phase. This effect can be seen from the
difference of the two lower red curves in Fig. 4-3; it is, however, less pronounced with
progressing time and increasing heating temperature.

Comparing the 1600°C case 4a with the corresponding case 3a without irradiation (the lower
two red curves), the increase of fractional release after 200 h is for all codes approximately
the same, the difference being ~3×10-4. It shows the effect of a preceding 500-days irradiation
phase at 1000°C, which allows already some part of the cesium to penetrate the coating and
then to appear sooner during the 1600°C heating. Calculated cesium release values are
ranging between 2.55×10-4 and 8.25×10-4 with the – corrected – FRESCO result well in
between these limits. Silver release data are as expected much higher compared to cesium and
are ranging between 27% and 55% among the codes.

For the 1800°C case (4b), calculation results for cesium are with 20-23% in a narrow range,
for silver with 58-95% in a somewhat wider range. Similar results were obtained for the
combined 1600/1800°C heating case 4c.

51
1.E+00

1.E-01

1.E-02
Fractional Release

1.E-03

1.E-04

1.E-05

1.E-06 Cesium
Cesium no irradiation
Silver
1.E-07 Strontium
Iodine
Cesium SiC failure at 1600
1.E-08
0 50 100 150 200 250 300 350 400
Heating Time [h]

Figure 4-3. FRESCO-II fractional release of 137Cs for cases 4a, 3a, and 4d (red), of 110mAg
and 90Sr for case 4a (blue and green), and of fission gases/iodine for case 4d (black)
[Verfondern 2008].

4.3.1.5. Case 5 – TRISO Coated Particle + Irradiation Phase with Temperature Cycles

Case 5 describes heating at 1600°C, but with a preceding irradiation phase consisting of 10
temperature cycles in the range 600-1000°C with a total irradiation time of 1000 days.
Fractional release data after irradiation (5a) and after the subsequent heating phase (5b) are
given in Table 4-5.

Table 4-5. Fractional release after 200/400 h for case 5.


Participant Fractional release from a cycles-irradiated TRISO coated particle
Case 5a (after irradiation) Case 5b (after 200 h heating)
137 110m 137 110m
Cs Ag Cs Ag
France 3.78×10-12 1.57×10-5 6.44×10-4 0.14
Germany 2.19×10-19 5.55×10-6 1.22×10-3 0.39
Korea 1.92×10-11 1.25×10-5 6.63×10-4 0.54
US/INL 6.45×10-14 5.06×10-5 3.07×10-4 0.42

Due to the low irradiation temperature, the fractional release data for cesium at end-of-life are
very low and should not be taken too seriously, since they may be dominated by effects of the
numerical calculation method rather than the physical model. Only silver release reaches a
level which is within the typically displayed release range. Calculated values differ by one
order of magnitude between 5.6×10-6 (FRESCO-II) and 5.1×10-5 (PARFUME).

52
1.E+00

1.E-01

1.E-02
Cesium Strontium Silver
Fractional Release

1.E-03 Silver 1000 Temperature

1.E-04

1.E-05

1.E-06

1.E-07

1.E-08
0 5000 10000 15000 20000
Irradiation Time [h]

Figure 4-4. Fractional release of metallic fission products for case 5, irradiation phase
[Verfondern 2008].

1.E+00

1.E-01

1.E-02
Fractional Release

1.E-03

1.E-04

1.E-05

1.E-06
Cesium Silver
1.E-07 Silver 1000 Cesium case 4a
Cesium case 3a
1.E-08
0 50 100 150 200
Irradiation Time [h]

Figure 4-5. Fractional release of metallic fission products for case 5, heating phase
[Verfondern 2008].

53
To see the difference compared to higher operating temperatures, the calculation was repeated
with FRESCO-II for a maximum cycle temperature of 1250°C. Figure 4-4 shows the release
of the metallic fission products 137Cs and 110mAg. An additional curve (thin blue line) for the
silver radionuclide illustrates the difference to the original case 5a of cycles with a maximum
temperature of 1000°C.

In case 5b, for the subsequent heating phase over 200 h at 1600°C as shown in Fig. 4-5 for
FRESCO-II, transient silver release for the two different temperature curves (blue lines) is
compared showing that for 1000°C maximum temperature, silver release is starting at a much
lower level, but is quickly approaching the 1250°C release curve. For cesium (red lines), the
transient release for 1250°C maximum temperature (upper red line) is compared with case 4a
with shorter irradiation at lower temperatue (middle red line), and also with case 3a (lower red
line) where no irradiation was assumed. All three cesium curves here are based on the
uncorrected, too high diffusion coefficient. The FRESCO-II calculation for case 5b was not
repeated; the corresponding release value in Table 4-5 should be lower, if it were corrected.

4.3.2. Postcalculations of Heating Experiments

For this postcalculation part of the accident benchmark study, a total of seven heating
experiments with fuel samples from four irradiation experiments have been selected. For the
purpose of easy comparison, the heating temperature/time history was precisely defined
including the heatup phase. Furthermore, the krypton release records from the heating tests
were translated into failures of coated particles as a pre-defined boundary condition for the
fission product release calculations.

One aspect that should be mentioned here was not treated the same in all codes: for a
fractional release in a heating test, the reference inventory is the inventory at the beginning of
the test. Therefore, in a calculation, the released amount of a fission product species at the end
of irradiation should be subtracted from the calculated fractional release during heating.

Another note again on FRESCO-II postcalculations: all 1600°C cases were conducted with
the uncorrected, too high diffusion coefficient. The mistake is insignificant at higher heating
temperatures.

4.3.2.1. Case 6 – HFR-P4

The irradiation test HFR-P4 was conducted to explore the potential limits of the German high-
quality UO2 LTI TRISO fuel beyond the target limits of the HTR-Modul. Particles were
embedded in cylindrical compacts machined from spherical fuel elements with a reduced
spherical fuel zone. Low in-pile R/B values of fission gases indicated no particle failure
during irradiation. Two of the irradiated compacts considered here, HFR-P4/1-12 and /3-7,
have achieved burnups of 11.1 and 13.9% FIMA and fast neutron fluences of 5.5 and
7.5×1025 n/m2, E>0.1 MeV, respectively. These two compacts were heated in the KÜFA
furnace at 1600°C over 300 h.

The curves of the postcalculations shown in the figures represent release from the coated
particles neglecting retention of the compact matrix material. Due to the small number of
coated particles per compact, the level corresponding to one failed particle is 6×10-4.

54
HFR-P4/1-12

Figures 4-6, 4-7, 4-8 show the fractional release of cesium, strontium, and silver, respectively,
as a function of heating time for the different codes and the comparison with the
corresponding measurements indicated by the red symbols. No experimental data are
available for silver release.

Cesium release measurements (Fig. 4-6) start at a relatively high level > 10-5, which may be
an indication of the presence of a coated particle with a defective/failed SiC layer. The
calculated release curves remain below the measurements in the first half of the test, but are
rising at a higher rate and eventually exceeding the measurements. Calculations of three codes
are almost identical in the second half of the test, while there are some larger differences in
the first phase. The plateau section in the curves results from the quick release of
contamination outside the SiC.
1.E+00 5000

1.E-01

1.E-02

1.E-03
Cs Release Fraction

1.E-04

1.E-05 US/INL
US/GA
France
1.E-06 Korea
Germany
South Africa
1.E-07 Expt Data
temperature

1.E-08 0
0 50 100 150 200 250 300 350

Time (Hours)

Figure 4-6. Fractional release of 137Cs from HFR-P4/1-12.

With regard to strontium (Fig. 4-7), all codes predict nearly the same transient fractional
release within a range of a factor of about 5, but are by more than three orders of magnitude
above the measurements. Silver release predictions shown in Fig. 4-8 in a linear scale vary
between 40% and 90% at the end of the heating test.

55
1.E+00 500

1.E-01

1.E-02

1.E-03
Strontium Release Fraction

1.E-04

1.E-05

1.E-06 US/INL France

Korea Germany
1.E-07 South Africa Expt Data

temperature
1.E-08 0
0 50 100 150 200 250 300 350

Time (Hours)

Figure 4-7. Fractional release of 90Sr from HFR-P4/1-12.

1.00 500

US/INL
US/GA
0.80 France
Korea
Germany
South Africa
0.60 temperature
Silver Release Fraction

0.40

0.20

0.00 0
0 50 100 150 200 250 300 350

Time (Hours)

Figure 4-8. Fractional release of 110mAg from HFR-P4/1-12.

56
HFR-P4/3-7

In this 1600°C test, three particle failures have been observed. The assumption to be
considered in the postcalculation was that the 1st coated particle failed after 49 h @ 1600°C,
the 2nd after 115 h, and the 3rd after 200 h.

Figures 4-9, 4-10, 4-11 show the fractional release of cesium, strontium, and silver,
respectively, as a function of heating time for the different codes and the comparison with the
corresponding measurements indicated by the red symbols.

The observed release of cesium (Fig. 4-9), which is significantly higher compared to 1600°C
heating tests with fuel of lower burnup < 10% FIMA, is in much better agreement with the
calculations than compact HFR-P4/1-12. The particle failures can be clearly identified in the
cesium release curve of FRESCO-II from the sudden puff releases (Fig. 4-11).
1.E+00 50

1.E-01

1.E-02

1.E-03
Cesium Release Fraction

1.E-04

US/INL
1.E-05 US/GA
France

1.E-06 Korea
Germany
South Africa
1.E-07 Expt Data
temperature

1.E-08 0
0 75 150 225 300

Time (Hours)

Figure 4-9. Fractional release of 137Cs from HFR-P4/3-7.

For strontium (Fig. 4-10), there is again a good agreement among the codes and a large
discrepancy to the measurements. The fact that PARFUME and ATLAS start at a high level is
probably because the release value at the end of irradiation was not subtracted from the
calculated release during heating. For silver (Fig. 4-11), calculations overestimate the
measurements. They are matched well at the beginning, but the measured silver release
flattens abruptly after 50 h and remains almost constant, a behavior that can hardly be
reproduced by a diffusion calculation.

57
1.E+00 5000

1.E-01

1.E-02

1.E-03
Strontium Release Fraction

1.E-04

US/INL
1.E-05
France

Korea
1.E-06 Germany
South Africa
1.E-07 Expt Data
temperature

1.E-08 0
0 50 100 150 200 250 300 350

Time (Hours)

Figure 4-10. Fractional release of 90Sr from HFR-P4/3-7.

1.00 500
US/INL
US/GA
France
0.80 Korea
Germany
South Africa
Expt Data
0.60 temperature
Silver Release Fraction

0.40

0.20

0.00 0
0 50 100 150 200 250 300 350

Time (Hours)

Figure 4-11. Fractional release of 110mAg from HFR-P4/3-7.

58
4.3.2.2. Case 7 – HRB-22

HRB-22, Test 3

The HRB-22 was an irradiation experiment at the HFIR in Oak Ridge dedicated to advanced
Japanese fuel [Minato 1998]. Four heating tests with unbonded coated particles, which were
extracted from selected compacts, were conducted in the Oak Ridge CCCTF furnace. Two of
these tests have been selected as benchmark examples. Only three codes have been applied for
postcalculating these two heating tests.

For the accident condition test 3 (ACT 3), compact 10 was chosen to extract 25 coated
particles to be heated for 270 h at 1700°C. There was a little amount of fission gas release
observed during the test which, however, never reached a level to justify the assumption of a
particle failure. Postcalculations for cesium are in very good agreement with the experimental
data (Fig. 4-12). The results for silver, here shown at a linear scale, conservatively cover the
experimental data with a difference of less than a factor of 2 (Fig. 4-13).

1.E+00 5000

1.E-01

1.E-02

1.E-03 US/INL

US/GA
Cesium Release Fraction

1.E-04 France

Korea

1.E-05 Germany

Expt Data

temperature
1.E-06

1.E-07

1.E-08 0
0 50 100 150 200 250 300

Time (Hours)

Figure 4-12. Fractional release of 137Cs from HRB-22, Test3.

59
1.00 5000

0.80

US/INL
0.60 US/GA
France
Silver Release Fraction

Korea
Germany
0.40
Expt. Data
temperature

0.20

0.00 0
0 50 100 150 200 250 300

Time (Hours)

Figure 4-13. Fractional release of 110mAg from HRB-22, Test3.

HRB-22, Test 4

For test ACT 4, again 25 individual coated particles were taken from compact 10 for heating
in CCCTF. Heating time was 220 h at 1800°C. The results given in Figs. 4-14 and 4-15 show
again a conservative coverage of the experimental data by the postcalculations.

1.E+00 5000

1.E-01

1.E-02
US/INL

US/GA
Cesium Release Fraction

1.E-03 France

Korea

Germany
1.E-04 Expt Data

temperature

1.E-05

1.E-06 0
0 50 100 150 200 250

Time (Hours)

Figure 4-14. Fractional release of 137Cs from HRB-22, Test4.

60
1.00 5000

0.80

US/INL
0.60 US/GA
France
Silver Release Fraction

Korea
Germany
0.40
Expt. Data
temperature

0.20

0.00 0
0 50 100 150 200 250

Time (Hours)

Figure 4-15. Fractional release of 110mAg from HRB-22, Test4.

4.3.2.3. Case 8 – HFR-K3

HFR-K3/1

The HFR-K3 irradiation experiment was considered a reference test for a steam-cycle HTGR
using fuel of the AVR 19 reload charge. This test with four spherical fuel elements, of which
two were heated after irradiation (Schenk 1989), belongs to the most well documented of the
German HTGR fuel program.

Sphere 1 was heated over 500 h at 1600°C. Despite severe irradiation conditions with a 7.5%
FIMA burnup and a fast neutron fluence of 4.0×1025 n/m2, E>0.1 MeV, the observed krypton
gas release remained below the inventory of one particle, therefore no particle failure was
assumed.

The measured transient release curves for the metallic fission products (Figs. 4-16, 4-17,
4-18) show that, after an initial increase, the release remains more or less constant, before – in
the case of cesium – the release fraction starts to increase and reaches at the end of the heating
test a value of about 1×10-4. Strontium remains below the 10-5 throughout the whole test,
while silver, after rapid increase to a level of ~3%, remains practically constant.

Postcalculation of this test with the various models reveals a tremendous overestimation by all
codes. Measurements are hardly reproducible with a diffusion model using a simple effective
diffusion coefficient.

61
1.E+00 50

Cesium Release Fraction 1.E-01

1.E-02

1.E-03

1.E-04

1.E-05 US/INL
US/GA
France
1.E-06
Korea
Germany
1.E-07 South Africa
Expt Data
temperature
1.E-08 0
0 100 200 300 400 500

Time (Hours)

Figure 4-16. Fractional release of 137Cs from HFR-K3/1.

1.E+00 50

1.E-01

1.E-02

1.E-03
Strontium Release Fraction

1.E-04

1.E-05

1.E-06 US/INL France

Korea Germany

1.E-07 South Africa Expt Data

temperature

1.E-08 0
0 100 200 300 400 500

Time (Hours)

Figure 4-17. Fractional release of 90Sr from HFR-K3/1.

62
1.00 50

0.80

0.60
Silver Release Fraction

0.40

US/INL US/GA

France Korea
0.20 Germany South Africa

Expt Data temperature

0.00 0
0 100 200 300 400 500

Time (Hours)

Figure 4-18. Fractional release of 110mAg from HFR-K3/1.

It should be mentioned here that both heating tests were part of the analysis study conducted
in 1989 where cesium release from 44 heating tests and strontium release from 10 heating
tests were taken to derive new recommendations for a diffusion coefficient in silicon carbide
(see chapter 5.0.0.0).

HFR-K3/3

Sphere 3 of the same irradiation test was heated at 1800°C over 100 h, where after 25 h, the
test was accidentally interrupted and later resumed. In contrast to sphere 1, heating of sphere 3
has shown a much more diffusion-like transient behavior as can be seen from the Figs. 4-19,
4-20, and 4-21, respectively. A steadily increasing Kr release was observed after 10 h heating
at 1800°C exceeding the level of one failed particle (6×10-5) after about 50 h at 1800°C. For
the postcalculation, it was assumed that 10 particle failures occurred during heating at 1800°C
after 50 h, 55 h, 65 h, 70 h, 75 h, 80 h, 85 h, 89 h, 92 h, and 97 h.

Corresponding postcalculations with diffusion codes – consequently – results in better


qualitative agreement with the measurements. But also the quantitative agreement is much
better in this case, even for silver, for which a more or less diffusion-like behavior was
recorded (which is not always the case). Largest discrepancy is again with the strontium
overestimated in the calculation by more than an order of magnitude.

63
1.E+00 5000

1.E-01

1.E-02

1.E-03
Cesium Release Fraction

1.E-04

US/INL
1.E-05 US/GA
France

1.E-06 Korea
Germany
South Africa
1.E-07 Expt Data
temperature

1.E-08 0
0 50 100 150 200

Time (Hours)

Figure 4-19. Fractional release of 137Cs from HFR-K3/3.

1.E+00 500

1.E-01

1.E-02

1.E-03
Stronlum Release Fraction

1.E-04

US/INL
1.E-05
France

Korea
1.E-06 Germany
South Africa

1.E-07 Expt Data

temperature

1.E-08 0
0 50 100 150 200

Time (Hours)

Figure 4-20. Fractional release of 90Sr from HFR-K3/3.

64
1.E+00 500

1.E-01

1.E-02

1.E-03
Silver Release Fraction

1.E-04
US/INL

1.E-05 US/GA
France
Korea
1.E-06
Germany
South Africa
1.E-07 Expt Data
temperature

1.E-08 0
0 50 100 150 200

Time (Hours)

Figure 4-21. Fractional release of 110mAg from HFR-K3/3.

The massive overprediction of strontium release in the HFR-K3 heating tests suggests to
possibly adjusting downwards the Sr diffusion coefficient in silicon carbide. This, however,
can only be done by taking all experimental evidence together.

4.3.2.4. Case 9 – HFR-K6/3

The heating test with an HFR-K6 irradiated sphere was one of the first in the new KÜFA-II
furnace operated at the ITU in Karlsruhe. The test was conducted with sphere 3 which had
reached a burnup of 10.9% FIMA (after revision: 9.7% FIMA) and a fast neutron fluence of
4.8×1025 n/m2, E>0.1 MeV. Heating temperature levels chosen were 1600, 1700, and 1800°C
over periods of 100 h each followed by a fourth heating phase again at 1800°C over additional
300 h [Freis 2008, Freis 2010].

The measured krypton release remained surprisingly low, even below the level of 10-5 during
the three heating phases. Only with beginning of the 4th heating phase, the release increased
significantly. The assumption with regard to the postcalculation was the failure of 5 particles:
after 119 h, 174 h, 214 h, 258 h, and 288 h, respectively, of the final 1800°C heating phase.

Also the observed cesium behavior (Fig. 4-22) has shown extremely low fractional release
data which remain during the 1600°C and 1700°C heating phases near the 10-6 level, before
the release starts to increase with the 1800°C phase and eventually reaches ~4%. The 110mAg
isotope was, due to the long time after the irradiation, no longer present in the sphere, while
strontium has not been measured so far.

65
1.E+00 500

1.E-01

1.E-02

1.E-03
Cesium Release Fraction

1.E-04
US/INL
US/GA
1.E-05
France
Korea
1.E-06 Germany
South Africa
1.E-07 Expt Data
temperature

1.E-08 0
0 100 200 300 400 500 600 700

Time (Hours)

Figure 4-22. Fractional release of 137Cs from HFR-K6/3.

1.E+00 5000

1.E-01

1.E-02

1.E-03
Strontium Release Fraction

1.E-04

US/INL
1.E-05
US/GA
France
1.E-06 Korea
Germany
South Africa
1.E-07 temperature

1.E-08 0
0 100 200 300 400 500 600 700

Time (Hours)

Figure 4-23. Fractional release of 90Sr from HFR-K6/3.

66
The postcalculations with the various codes exhibit good agreement among each other, but a
strong overestimation of the cesium release with a difference of several orders of magnitude
during the first few hundred hours of heating, and reduced to about a factor of 10 at the end of
the test. The effect of the few failed particles on the cesium release is not visible due to the
much larger total release (both measured and calculated). The calculated silver release
(Fig. 4-24), again shown for a linear scale, is ranging between 60% and 100% fractional
release at test end.

1.00 500

0.80

0.60
Silver Release Fraction

US/INL
0.40 US/GA

France

Korea

0.20 Germany

South Africa

temperature

0.00 0
0 100 200 300 400 500 600 700

Time (Hours)

Figure 4-24. Fractional release of 110mAg from HFR-K6/3.

4.3.3. Predictive Calculations of Heating Experiments

4.3.3.1. Case 10 – HFR-EU1bis/1

As part of the revitalized irradiation program for HTGR fuel irradiation testing in the
European Union, the so-called HFR-EU1 irradiation experiment was initiated with the goal to
explore the performance limits of the presently existing German and Chinese high-quality fuel
[Laurie 2010]. This test with a focus on high burnup (towards 20% FIMA) was terminated in
February 2010. A parallel, congenial test, called HFR-EU1bis, with a focus on high operation
temperature was conducted under simplified conditions and was terminated in October 2005
with postirradiation examination work having started soon afterwards [Fütterer 2006].

An example suggested for the prediction section was the heating experiment with sphere 1 of
the irradiation test HFR-EU1bis. This test was, at least at the time of definition of the accident
benchmark, in the planning stage and therefore considered a prediction. Experimental cesium
and silver release data have been made available in the meantime, and the boundary
conditions for both irradiation and heating have been defined for the calculations accordingly.

67
Sphere 1 out of the five fuel spheres inserted in the EU1bis test, which reached a burnup of
9.3% FIMA and a fast neutron fluence of 3×1025 n/m2, E>0.1 MeV, was planned to be heated
at temperature levels of 1250, 1600, 1700, and 1800°C over 200 h each. The first, relatively
low temperature level was chosen to account for the analysis of the release behavior of silver,
of which still plenty is available. In the real heating test, there was no 1800°C heating phase.
Another difference was the assumption in the calculations to proceed to the next temperature
level immediately, whereas in the real heating test, the fuel sphere was cooled down to room
temperature after each heating phase. These cooldown and heatup interim phases took a
longer time than the 2 hours assumed in the calculations.

Furthermore, in the release calculations, no particle failure was assumed. For the heating test
itself, this was found to be in agreement with the experiment, since measured krypton release
during heating remained below the level of the inventory of a single coated particle. What
may not be in agreement with the real experiment, is the high chance of presence of
irradiation-induced failures of an SiC layer or even the whole coating that could be concluded
from the high R/B gas release measured particularly toward the end of irradiation. To what
extend sphere EU1/1 had contributed to the overall gas release could not be distinguished
during the irradiation test. The transient fractional release curves for the metallic fission
products are plotted in Figs. 4-25, 4-26, and 4-27.

The measured 137Cs activity (see symbols in Fig. 4-25) is on a relatively high level already
during the 1250°C heating. This is presumably the result of the high temperature during
irradiation with enhanced diffusion through the particle coatings and release from particles
with an SiC layer that failed during irradiation.

1.E+00 5

1.E-01

1.E-02

1.E-03
Cesium Release Fraction

1.E-04

US/INL
1.E-05 US/GA
France

1.E-06 Korea
Germany
South Africa
1.E-07
Expt Data
temperature

1.E-08 0
0 100 200 300 400 500 600 700 800

Time (Hours)

Figure 4-25. Fractional release of 137Cs from HFR-EU1bis/1.

68
1.E+00 500

1.E-01

1.E-02

1.E-03
Strontium Release Fraction

1.E-04

1.E-05 US/INL
France

1.E-06 Korea
Germany
South Africa
1.E-07
temperature

1.E-08 0
0 100 200 300 400 500 600 700 800

Time (Hours)

Figure 4-26. Fractional release of 90Sr from HFR-EU1bis/1.

1.00 500

0.80

0.60
Silver Release Fraction

US/INL
0.40
US/GA
France
Korea
Germany
0.20
South Africa
Expt Data
temperature

0.00 0
0 100 200 300 400 500 600 700 800

Time (Hours)

Figure 4-27. Fractional release of 110mAg from HFR-EU1bis/1.

69
In agreement with the experience with high silver release at temperatures above 1000°C, the
calculated fractional release of 110mAg (Fig. 4-27) is reaching already the percentage range
towards the end of the 1250°C heating phase, followed by the expected steep increase at
higher temperatures. Also 137Cs and 90Sr are predicted to escape significantly from the sphere
with beginning of the sphere.

4.3.3.2. Case 11 – HTR-PM Fuel Sphere

The “HTR-PM” called test has actually no direct relation to the Chinese HTR-PM design. It is
a simple code-to-code comparison where fuel sphere data assumed are those of the German
reference fuel exposed to a fictive irradiation history.

The calculations for the heating experiment with an HTR-PM fuel sphere case were made
based on the assumption of no particle failure during the heating. Results are shown in
Fig. 4-28 with all codes being in fairly good agreement with each other.

1.E+00 5000

1.E-01

1.E-02

1.E-03
Cesium Release Fraction

1.E-04

US/INL
1.E-05
US/GA

France
1.E-06
Korea
Germany
1.E-07 South Africa
temperature

1.E-08 0
0 100 200 300 400 500 600 700 800

Time (Hours)

Figure 4-28. Prediction of fractional release of 137Cs for HTR-PM fuel sphere.

70
5. VALIDATION STUDY FOR RECENT HEATING EXPERIMENTS (2011)

5.1. Transient Heating Tests

Six spherical fuel elements have been heated in the KÜFA facility in the Hot Cells at the
Research Center Jülich under conditions of the maximum temperature transient calculated for
a core heatup accident in the depressurized HTR-Modul. In one of these tests, the heating
temperature was raised by a constant margin such that a maximum temperature of 1700°C
was obtained. The fuel elements for these tests are listed in Table 5-1.

Table 5-1. FZJ heating tests with UO2 TRISO coated fuel particles exposed to a transient
heating temperature representative of the HTR-Modul core heatup accident.
Test Irradiation Burnup FZJ Transient heating
facility [% FIMA] heating facility temperature [°C]
AVR 85/18 AVR 9.2 KÜFA, Jülich max. 1620
AVR 89/13 AVR 9.1 KÜFA, Jülich max. 1620
AVR 90/2 AVR 9.3 KÜFA, Jülich max. 1620
AVR 90/5 AVR 9.2 KÜFA, Jülich max. 1620
AVR 90/20 AVR 9.8 KÜFA, Jülich max. 1620
AVR 91/31 AVR 9.0 KÜFA, Jülich max. 1700

The transient heating temperature which the fuel elements were exposed to in the KÜFA
furnace is shown in Fig. 5-1. It corresponds to the case of a core heatup in the depressurized
reactor predicted for the HTR-Modul at the hottest position in the core plus an uncertainty
addition (about 100 degrees). The temperature curve starting at the normal operation
temperature level of 857°C increases at the beginning, passes through a maximum value of
1620°C reached after about 30 h into the accident, and then gradually decreases. The duration
of the heating tests was nominally 300 h when the temperature is below 1200°C.

The test with sphere AVR 90/2 was shut down after 120 h. The spheres AVR 89/13 and AVR
90/5 were exposed twice to the same heating transient. For sphere AVR 91/31 heated to
higher temperatures, the transient started at the normal operation temperature level of 857°C,
rising to a maximum temperature of 1695°C reached after 30 h, before cooling down to
1200°C after 300 h.

The characteristic data of the six transient-heated AVR GLE-3 fuel elements are summarized
in Table 5-2.

Table 5-2. Characteristic data of the UO2 TRISO particles used in the heating tests.
AVR GLE-3
Fuel LEU UO2
Particle batch HT232-245
No. of cp per sphere 16,350
Fraction corresponding to one cp 6.1×10-5

71
Figure 5-1. Temperature transient during the heating experiment of fuel spheres exposed to
the HTR-Modul accident temperature curve plus an uncertainty margin.

Special assumptions:

 The fraction of heavy metal contamination in the fuel spheres is assumed to be


4.9×10-5 which is the typical cross contamination level for AVR fuel. A sorption of
fission products at the fuel element surfaces is not taken into account meaning the
unhindered transition of fission products from the fuel element into the coolant.

 All six fuel elements had seen very similar irradiation conditions with narrow ranges
for heavy metal burnup of 9.0-9.8% FIMA and for fast neutron fluence of
2.6-2.9×1025 n/m2. The assumption was made that these two parameters are in a linear
relationship with the irradition time in the AVR reactor (1311-1427 efpd).

 An accurate value for an averaged irradiation temperature in the AVR during the fuel
life time cannot be given. For the calculations, the typically selected range was
assumed with a lower limit at 700°C and an upper limit at 1000°C.

 An initial value of 857°C for the heating temperature was chosen to simulate the
operating conditions.

 The matrix type of GLE-3 spheres is A3-27. Therefore, according to the


recommendation given in [IAEA 1997], the diffusion coefficients for both cesium and
strontium in the matrix material was divided by 10 compared to the reference data
(valid for A3-3).

 In these calculations, the diffusion coefficient for cesium in SiC was corrected
according to the finding from CRP-6 of a wrong low-temperature branch under
accident/heating conditions.

 As is usually done, the EOL fractional release has been subtracted from the fractional
release data in the heating phase.

72
AVR 85/18

Characteristic data of experiment AVR 85/18

Burnup: 9.2% FIMA Fast neutron fluence: 2.6×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1340 efpd
Heating temperature: Transient, Heating time: 300 h
max. 1620°C

The heating test AVR 85/18 was conducted in the KÜFA furnace at FZJ with transient
temperatures simulating the HTR-Modul core heatup accident with a maximum temperature
of 1620°C reached after 30 h into the transient test. The measured krypton release (1.4×10-7)
indicates that no particle failed during the heating period.

Cesium

The measured cesium release fraction at the end of the 300 h heating test was 1.3×10-5
showing – in connection with the low release from the coated particles and the low krypton
release – that the cesium originates from contamination only.

Figure 5-2a shows the calculation results of cesium release from the particles (blue curve) and
from the fuel element (red curves) as function of heating time as well as their comparison
with the cesium measurements (symbols). For the release from the fuel element, it is
distinguished between the irradiation temperature of 700°C (solid red curve) and 1000°C
(dashed red curve). The release curves reveal a cesium fraction released from the particles to
remain below the measured release from the fuel elemcnt during the whole experiment. The
amount of cesium that escaped the particles after 300 h corresponds to less than the inventory
of one particle which means that the cesium that escaped the fuel element originates from
heavy metal contamination being uninfluenced by the release from the coated particles.

73
1 E+0

AVR 85/18 Cesium


1 E-1

1 E-2
Release from CP (700) Release from FE (700)
Cs-137 measurements Temperature
1 E-3 Release from FE (1000)
Failure fraction

1 E-4

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350

Heating time (h)

Figure 5-2a. FRESCO-II postcalculation of 137Cs fractional release for AVR 85/18.

Strontium

Strontium could not be precisely measured due to the high silver activity on the plates. Figure
5-2b shows the reference postcalculations for strontium with the expected low release from
the fuel element. The calculated release from the coated particles rises above the 10-3 level
which may be very conservative if taking account of the experience from other heating tests.

Silver

The measured silver release fraction strongly increased within increasing temperatures to
1.2×10-3 at the maximum temperature and increased further to 6.5×10-3 by the end of the
300 h heating test. These experimental data are still by a factor of about 30 lower compared to
the predicted release (Fig. 5-2c).

74
1 E+0

AVR 85/18 Strontium

1 E-1

1 E-2

1 E-3
Failure fraction

Release from CP (700) Release from FE (700)


1 E-4
Sr-90 measurements Temperature
Release from FE (1000)

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350

Heating time (h)

Figure 5-2b. FRESCO-II postcalculation of 90Sr fractional release for AVR 85/18.
1 E+0

AVR 85/18
Silver
1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4
Release from CP (700) Release from FE (700)
Ag-110m measurements Temperature
1 E-5 Release from FE (1000)

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350

Heating time (h)

Figure 5-2c. FRESCO-II postcalculation of 110mAg fractional release for AVR 85/18.

75
AVR 89/13

Characteristic data of experiment AVR 89/13

Burnup: 9.1% FIMA Fast neutron fluence: 2.6×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1325 efpd
Heating temperature: 2 x transient, Heating time: 2 x 300 h
1620°C

The heating test AVR 89/13 was conducted in the KÜFA furnace at FZJ where the sphere was
heated twice (similar to heating test AVR 90/5) according to the transient temperatures
simulating the HTR-Modul core heatup accident with a maximum temperature 1620°C
reached after ~30 h. The reason for the two consecutive heating phases was to examine
whether or not it is possible to resume reactor operation after a core heatup depressurization
accident without a change of the fuel. The measured krypton release fraction which was
2.0×10-7 after the first heating phase and practically zero during the second heating phase
indicates that no particle failed during the two heating periods.

Cesium

The cesium fractional release remains very low throughout both heating periods. Similar to
the krypton behavior, the 137Cs release during the second heating period was with 1.4×10-6
much lower than during the first heating period with 1.1×10-5. This shows that the cesium
originates exclusively from contamination of the matrix material.

The experimental data are consistent with the postcalculation based on the 700°C irradiation
temperature. Despite the predicted significant release from the coated particle, its influence on
the release from the fuel sphere is only marginal remaining below the assumed level of heavy
metal contamination. Only in the case of Tirr = 1000°C, has proceeded inside the TRISO
coating such that the cesium released from the coated particles (not shown in the figure)
contributes significantly to the release from the fuel element.

76
1 E+0

AVR 89/13 Cesium


1 E-1

1 E-2 Release from CP (700) Release from FE (700)


Cs-137 measurements Temperature
Release from FE (1000)
1 E-3
Failure fraction

1 E-4

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350 400 450 500 550 600 650 700

Heating time (h)

Figure 5-3a. FRESCO-II postcalculation of 137Cs fractional release for AVR 89/13.

Strontium

There are no strontium measurements available from this test.

The calculated strontium release from the fuel element shows some further increase during the
second transient heating, but remains below 10-5 for the total investigation time due to the
strong holdup in the matrix material.

Silver

Different from cesium (and krypton), the release of silver has become much stronger in the
second heating period. While the total 110mAg release fraction was measured to be 8.3×10-4
during the first heating, it was 1.46% during the second heating accumulating to a total of
1.54×10-2 after 600 h.

The postcalculation of silver release is rapidly increasing above the 10% level reaching a
fractional release of 32% after the first heating transient, further increasing to 57% after the
second phase.

77
1 E+0

AVR 89/13 Strontium

1 E-1

1 E-2

1 E-3
Failure fraction

Release from CP (700) Release from FE (700)


Sr-90 measurements Temperature
1 E-4
Release from FE (1000)

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350 400 450 500 550 600 650 700

Heating time (h)

Figure 5-3b. FRESCO-II postcalculation of 90Sr fractional release for AVR 89/13.

1 E+0

AVR 89/13
Silver
1 E-1

1 E-2

1 E-3
Failure fraction

Release from CP (700) Release from FE (700)


Ag-110m measurements Temperature
1 E-4 Release from FE (1000)

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350 400 450 500 550 600 650 700

Heating time (h)

Figure 5-3c. FRESCO-II postcalculation of 110mAg fractional release for AVR 89/13.

78
AVR 90/2

Characteristic data of experiment AVR 90/2

Burnup: 9.3% FIMA Fast neutron fluence: 2.7×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1354 efpd
Heating temperature: Transient, 1620°C Heating time: 120 h

The heating test AVR 90/2 was conducted in the KÜFA furnace at FZJ where the sphere was
heated according to the transient temperatures simulating the HTR-Modul core heatup
accident with a maximum temperature of 1620°C reached after ~30 h. Due to interruption of
power supply, the test was discontinued at 120 h into the test during the cooldown phase at
1435°C. In this heating test, three pressure vessel failed particles were observed with the
overall krypton release fraction totaling to 1.2×10-4. For the postcalculation, a corresponding
particle failure function with three steps (black curve in the figures) has been assumed.

Cesium

The release fraction was measured to be 2.4×10-5 at the maximum temperature and then
further increased by the end of the (short) test to 4.6×10-5, which is still below the assumed
level of heavy metal contamination in the matrix graphite, but higher compared to other
transient heating tests. The cesium measurements of test AVR 90/2 shown in Fig. 5-4a give
rise to a certain but small influence by the failure of the three particles, the reason of which is
obviously a buffer effect in the fuel element graphite.

The reference postcalculation shown in Fig. 5-4a is in pretty good agreement with the
experimental data. Differences are not large and well within the uncertainty range of the
cesium transport in graphite and of the scattering of the beavy metal contamination of AVR
fuel clements. The release from the fuel element (red curve) appears to come mainly from
contamination and, compared with the release from the particles (blue curve), has no obvious
influence of the three particle failures on the red curve. Also the conservative assumption in
the model that a particle failure directly results in an exposed particle kernel could affect the
release behavior due to the fact that only 57%, and 46%, and 31%, respectively, of the
krypton inventory have escaped the particles in consideration after their failure.

The dashed red curve, representing the calculation with a 1000°C irradiation temperature,
shows in this example a larger difference (compared to other examples) to the 700°C case.
The later rise of the 1000°C curve could be explained by a stronger “cleaning” of the sphere
from contamination during irradiation with less cesium left for escaping during the heating
phase.

79
1 E+0
Cesium
AVR 90/2
1 E-1

1 E-2 Release from CP (700)


Release from FE (700)
Cs-137 measurements
1 E-3 Failure fraction
Failure fraction

Temperature
1 E-4 Release from FE (1000)

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150

Heating time (h)

Figure 5-4a. FRESCO-II postcalculation of 137Cs fractional release for AVR 90/2.

In Fig. 5-4b, an older postcalculation, the effect of a particle which is defective since the
beginning of the heating phase is being examined. From the viewpoint of the calculation,
there is no difference, at least at an irradiation temperature of 700°C, whether the particle has
become defective during the fabrication process or not until the operation in the AVR. The
assumption of a defective particle leads to a rapid increase of the release from the particles
during the heating test up to the level of the inventory of one particle (blue curve) and remains
then almost constant. There is only a negligible influence of the release from the particles on
the release behavior from the fuel element if assuming a reference value of 5×10-5 for the
contamination fraction in the graphite. However, due to the fact that the measured cesium
release from the fuel elements of other transient tests appears to stabilize at a lower level of
~2×10-5, this lower contamination value was assumed in a separate calculation together with
the presence of a defective particle. The additional assumption of a defective particle raises
the release curve up to the level of the heating test data. These calculations show that, in the
case of low cesium release, assumption on both presence of defective/failed particles and
heavy metal contamination may certainly influence the release behavior and help explaining
differences to the measurements.

80
Figure 5-4b. FRESCO-II postcalculation of 137Cs fractional release for AVR 90/2.

Strontium

There are no strontium measurements available for this test.

The postcalculation of strontium release in Fig. 5-4c shows the expected behavior with a high
release from the coated particles, also from intact particles, and a release from the fuel
element that remains at a low level due to the strong holdup in the matrix graphite.

Silver

Completely unaffected by the particle failures is the silver release behavior. With reaching the
maximum temperature, its fractional release has reached already a value of 3.5×10-3 which is
above the krypton failure fraction. In the subsequent 90 h cooldown phase, silver release has
further increased to 3.7×10-2.

The postcalculation of silver release (Fig. 5-4d) is, in this case, in surprisingly good
agreement with the measurements, well covering the experimental data.

81
1 E+0
Strontium
AVR 90/2
1 E-1

Release from CP (700) Release from FE (700)


1 E-2 Sr-90 measurements Failure fraction
Temperature Release from FE (1000)
1 E-3
Failure fraction

1 E-4

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150

Heating time (h)

Figure 5-4c. FRESCO-II postcalculation of 90Sr fractional release for AVR 90/2.

1 E+0

AVR 90/2 Silver


1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4

1 E-5 Release from CP (700)


Release from FE (700)
1 E-6 Ag-110m measurements
Failure fraction
1 E-7 Temperature
Release from FE (1000)
1 E-8
0 50 100 150

Heating time (h)

Figure 5-4d. FRESCO-II postcalculation of 110mAg fractional release for AVR 90/2.

82
AVR 90/5

Characteristic data of experiment AVR 90/5

Burnup: 9.2% FIMA Fast neutron fluence: 2.7×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1340 efpd
Heating temperature: 2 x transient, Heating time: 300+300 h
1620°C

The heating test AVR 90/5 was conducted in the KÜFA furnace at FZJ where the sphere was
heated twice (similar to heating test AVR 89/13) according to the transient temperatures
simulating the HTR-Modul core heatup accident with a maximum temperature of 1620°C
reached after ~30 h. The reason for the two consecutive heating phases was to examine
whether or not it is possible to resume reactor operation after a core heatup depressurization
accident without a change of the fuel. The overall measured krypton release fraction at the
end of both heating phases was 1.9×10-7.

Cesium

The cesium release calculated for sphere AVR 90/5 is shown in Fig. 5-5a. The measurements
(symbols) stabilize in the first heating period at a constant level determined by the heavy
metal contamination. This level practically persists throughout the whole second heating
period by taking into account an eventually lower contamination level in this fuel ball.
Experimental results for cesium release are very similar to those of heating test AVR 89/13.

While release from the coated particles (blue curve) remains at a low level during the first
heating phase, there is a significant increase in the subsequent heating period with reaching
the maximum temperature and, after that, also a slight increase of the release from the fuel
clement (red solid curve), a stronger one for the 1000°C irradiation temperature (red dashed
curve) – again in contrast to the measurements. Postcalculation of the cesium release is very
similar to that of transient test AVR 89/13.

83
1 E+0

AVR 90/5 Cesium


1 E-1

1 E-2 Release from CP (700) Release from FE (700)


Cs-137 measurements Failure fraction
Temperature Release from FE (1000)
1 E-3
Failure fraction

1 E-4

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350 400 450 500 550 600 650 700

Heating time (h)

Figure 5-5a. FRESCO-II postcalculation of 137Cs fractional release for AVR 90/5.

An older calculation the results of which are given as solid curves in Fig. 5-5b demonstrates
that, in case of a reduction of the diffusion coefficient for cesium in SiC by a factor of 2 and
also a variation in the heavy metal contamination level, the release from the fuel clement
remains low, in good agreement with the experimental data.

However, the same conclusion could also be drawn if, instead of the diffusion coefficient in
SiC, the one in graphite would be varied. A corresponding postcalculation with a different
diffusion coefficient in matrix graphite reduced by one order of magnitude is successful; here
the release from the particles (dash-dotted curve) remains practically unchanged compared to
the reference calculation. The variation of the diffusion coefficients in SiC and in graphite,
however, would reveal only an upper limit for the release.

84
Figure 5-5b. FRESCO-II postcalculation of 137Cs fractional release for AVR 90/5.

Strontium

There are no experimental data for strontium available for this heating test.

Also the calculated strontium release behavior from the coated particles (blue curve) and from
the fuel element (red curve) is, due to similar operating and heating conditions, very similar
from that of heating test AVR 89/13 (Fig. 5-5c).

Silver

The amount of silver released during the second heating period (9.0×10-4) was approximately
the same as in the first heating period (1.1×10-3) accumulating to a total of 2.0×10-3 after
600 h. Measurements soon reach the level of approximately 10-3 where the release slightly,
but steadily increases. The release behavior during the first heating period is practically the
same as that in test AVR 89/13. Only during the second heating period, the release level
remains almost constant here, whereas is it significantly increasing to above the 10-2 level in
the other test.

The calculation results (Fig. 5-5d) are, again similar to test AVR 89/13, soon reaching high
release values far above the measurements.

85
1 E+0

AVR 90/5 Strontium

1 E-1

1 E-2

1 E-3
Failure fraction

Release from CP (700)

1 E-4 Release from FE (700)

Sr-90 measurements

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350 400 450 500 550 600 650 700

Heating time (h)

Figure 5-5c. FRESCO-II postcalculation of 90Sr fractional release for AVR 90/5.
1 E+0

AVR 90/5
Silver
1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4 Release from CP (700) Release from FE (700)


Ag-110m measurements Failure fraction
Temperature Release from FE (1000)
1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350 400 450 500 550 600 650 700

Heating time (h)

Figure 5-5d. FRESCO-II postcalculation of 110mAg fractional release for AVR 90/5.

86
AVR 90/20

Characteristic data of experiment AVR 90/20

Burnup: 9.8% FIMA Fast neutron fluence: 2.9×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1427 efpd
Heating temperature: Transient, Heating time: 300 h
max. 1620°C

The heating test with the GLE-3 sphere AVR 90/20 was conducted in the KÜFA furnace at
FZJ with transient temperatures simulating the HTR-Modul core heatup This was the first
heating test with a high-burnup (9.8% FIMA) fuel sphere to be exposed to the transient
temperature curve. Shortly before reaching the maximum temperature, a burst of krypton
release was observed presumably from up to 5 defective/failed particles. Krypton
measurements corresponded to the inventory of 3-5 particles, two of which may have been
initially defective coated particles. In the postcalculations, the assumption of five particle
failures was made, two after 22 h into the transient test, one more after 27 h, plus two more
after 36 h (see black curve in following figures).

Cesium

A gradual increase in cesium release during the test was observed reaching a final release
fraction of 6.5×10-6 after 300 h. This is most probably cesium originating from the free
uranium in the fuel sphere (estimated 5×10-5), still with a strong holdup by the matrix
graphite.

In the postcalculation (Fig. 5-6a), the release from the coated particles (blue curve) is seen to
closely follow the assumed particle failure function reaching 3×10-4, which corresponds to the
inventory of five particles. The release fraction from the fuel element reaches 1.7×10-4, well
above the experimental data.

87
1 E+0

AVR 90/20 Cesium


1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4

1 E-5

1 E-6

1 E-7
Release from CP (700) Release from FE (700)
Cs-137 measurements Failure fraction
Temperature Release from FE (1000)
1 E-8
0 50 100 150 200 250 300 350 400

Heating time (h)

Figure 5-6a. FRESCO-II postcalculation of 137Cs fractional release for AVR 90/20.

Strontium

Experimental data for strontium release are not available for this test.

The postcalculation (Fig. 5-6b), similar to the others, exhibits a strong release from the coated
particles reaching soon a level above 10-3 and fairly unaffected by the presence of the five
failed particles.

Silver

Silver release increases with the fuel sphere heatup to a value of around 7% and then remains
almost constant. Release fraction after 300 h is 7.6%. The corresponding calculation curve
(Fig. 5-6c) envelops pretty well the experimental data.

88
1 E+0

AVR 90/20 Strontium

1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4

1 E-5

1 E-6

1 E-7 Release from CP (700) Release from FE (700)


Sr-90 measurements Failure fraction
Temperature Release from FE (1000)
1 E-8
0 50 100 150 200 250 300 350 400

Heating time (h)

Figure 5-6b. FRESCO-II postcalculation of 90Sr fractional release for AVR 90/20.

1 E+0

AVR 90/20
1 E-1

Silver
1 E-2

1 E-3
Failure fraction

1 E-4

1 E-5

1 E-6

1 E-7 Release from CP (700) Release from FE (700)


Ag-110m measurements Failure fraction
Temperature Release from FE (1000)
1 E-8
0 50 100 150 200 250 300 350 400

Heating time (h)

Figure 5-6c. FRESCO-II postcalculation of 110mAg fractional release for AVR 90/20.

89
AVR 91/31

Characteristic data of experiment AVR 91/31

Burnup: 9.0% FIMA Fast neutron fluence: 2.6×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1311 efpd
Heating temperature: Transient, 1700°C Heating time: 300 h

The heating test with the GLE-3 sphere AVR 91/31 was conducted in the KÜFA furnace at
FZJ with transient temperatures simulating the HTR-Modul core heatup accident and
extrapolated such that a nominal maximum value of 1700°C was reached. According to the
krypton release measurements, single particle failures were registered in the temperature
range 1585 to 1700°C. The estimated number of particle failures by the end of the test is
18-20. The first two krypton release bursts are presumed to have been caused by the presence
of two initially defective coated particles, where the left-over (from the irradiation phase)
fission gas inventory escaped not until the heating at higher temperatures in a burst-like
release. In the postcalculation, the assumption was made that 18 coated particles failed in
eight steps between 21 h and 50 h into the transient.

Cesium

After the observed failure of a few particles, a somewhat delayed cesium release can be seen
which can obviously be traced back to the defective/failed coated particles. After about 100 h
into the transient, the cesium fractional release is exceeding the krypton fractional release.
Reasons are presumed to be release coming from some intact coated particles, or a somewhat
larger number of failed coated particles, or also the presence of some particles with a failed
SiC layer, which would explain the lower Kr release. Also there is only minor holdup in the
A3 matrix material due to the higher heating temperature history compared to the 1620°C-
cases.

90
1 E+0

AVR 91/31 Cesium


1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4

1 E-5

1 E-6

1 E-7
Release from CP (700) Release from FE (700)
Cs-137 measurements Failure fraction
Temperature Release from FE (1000)
1 E-8
0 50 100 150 200 250 300 350 400

Heating time (h)

Figure 5-7a. FRESCO-II postcalculation of 137Cs fractional release for AVR 91/31.

Strontium

There are no experimental data for strontium release available for this test.

Due to the higher heating temperatures, the calculated fractional release of strontium
(Fig. 5-7b) reaches a final value of 1.8%. The release from the fuel element is somewhat
higher compared to the 1620°C cases, but remains below 10-5 throughout the investigation
time.

Silver

The release of silver increases to 12.4% with reaching the maximum temperature of 1700°C
and continues increasing. Final release fraction is with 62% much higher compared to the
1620°C-transient tests.

The calculated silver release from the fuel element is in surprisingly good agreement with the
experimental data (Fig. 5-7c).

91
1 E+0

AVR 91/31 Strontium

1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4

1 E-5

1 E-6

1 E-7 Release from CP (700) Release from FE (700)


Sr-90 measurements Failure fraction
Temperature Release from FE (1000)
1 E-8
0 50 100 150 200 250 300 350 400

Heating time (h)

Figure 5-7b. FRESCO-II postcalculation of 90Sr fractional release for AVR 91/31.
1 E+0

AVR 91/31
Silver
1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4

1 E-5

1 E-6

1 E-7
Release from CP (700) Release from FE (700)
Ag-110m measurements Failure fraction
Temperature Release from FE (1000)
1 E-8
0 50 100 150 200 250 300 350 400

Heating time (h)

Figure 5-7c. FRESCO-II postcalculation of 110mAg fractional release for AVR 91/31.

92
5.2. FRESCO-II Results for New Heating Tests in KÜFA-II

The new heating furnace KÜFA-II at JRC-ITU in Karlsruhe has been used up to now for eight
heating experiments. Of the spheres heated, two were of GLE-3 type manufactured in 1981
and irradiated in the AVR, of which the first one (AVR 73/21) was just a short test to
demonstrate the ability of the new facility. The other six have been fuel spheres irradiated in
the HFR Petten, two in the experiment HFR-K6, which was the proof test for HTR-Modul
fuel, and four in the experiment HFR-EU1bis, which was a high (irradiation) temperature test
representing VHTR conditions.

The tests are listed in Table 5-3, all of which except for the last one are well documented in
[Freis 2010]. The heating experiment HFR-EU1bis/5 was conducted under South Africa
direction. Some characteristic data of the different types of fuel used in these tests are given in
Table 5-4.

Table 5-3. Heating tests in new KÜFA-II furnace at Karlsruhe.


Test Irradiation Burnup Heating facility Heating
facility [% FIMA] temperature [°C]
AVR 73/21 AVR, Jülich 2.5 KÜFA-II, Karlsruhe 1600, 1800
AVR 74/18 AVR, Jülich 4.8 KÜFA-II, Karlsruhe 1800
HFR-K6/2 HFR, Petten 9.3 KÜFA-II, Karlsruhe 1600, 1800
HFR-K6/3 HFR, Petten 9.7 KÜFA-II, Karlsruhe 1600, 1700, 1800
HFR-EU1bis/1 HFR, Petten 9.3 KÜFA-II, Karlsruhe 1250, 1600, 1700
HFR-EU1bis/3 HFR, Petten 11.1 KÜFA-II, Karlsruhe 1250, 1600
HFR-EU1bis/4 HFR, Petten 11.1 KÜFA-II, Karlsruhe Transient, 1720
HFR-EU1bis/5 HFR, Petten 9.7 KÜFA-II, Karlsruhe 1000, transient,
1500, decrease

Table 5-4. Characteristic data of the fuel used in the heating tests in KÜFA-II.
AVR HFR-K6 HFR-EU1bis
Type GLE-3 Proof test GLE-4
Fuel LEU UO2 LEU UO2 LEU UO2
Particle batch HT232-245 EUO2358-2365 HT354-383
No. of cp per sphere 16,350 14,580 9560
Fraction corresponding to one cp ~6×10-5 ~7×10-5 ~1×10-4
Mean SiC strength [MPa] 834 834 834
Weibull modulus 8.02 8.02 8.02

93
Special assumptions:

 For all MTR-irradiated fuel spheres, the fraction of heavy metal contamination in the
fuel spheres was assumed to be 1.0×10-6. Due to the typical cross contamination level
in the AVR reactor, all AVR-irradiated fuel elements have a much higher heavy metal
contamination fraction, which was (probably) assumed to be 5.0×10-5. A sorption of
fission products at the fuel element surfaces is not taken into account meaning the
unhindered transition of fission products from the fuel element into the coolant.

 For all spheres with A3-27 matrix material, the diffusion coefficient for cesium and
strontium was reduced by a factor of 10.

 For the AVR spheres, an averaged irradiation temperature of 820°C over the total
lifetime was assumed [Freis 2010].

 In all AVR and HFR-K6 spheres heated in KÜFA-II, the irradiation has been long
time ago. Therefore all 110mAg inventory has decayed at the time of the heating and
could not be measured any more.

 EU1bis experimental data must be considered with caution, since a high uncertainty
was introduced due to impact by the temperature excursion at the beginning of the test.
Tendency of enhanced gas release towards the end of the test leads to presumption that
between several and a hundred coated particles inside the five spheres may have
failed. Assumptions in the postcalculations were made with 0 and with 10 failed
particles with beginning of the heating.

94
AVR 73/21

Characteristic data of experiment AVR 73/21

Burnup: 2.5% FIMA Fast neutron fluence: 0.4×1025, E>0.1MeV


Irradiation temp.: ~700°C Irradiation time: 235 efpd
Heating temperature: 1600, 1800°C Heating time: 5, 5 h

The heating test AVR 73/21 served the sole purpose to demonstrate the proper functioning of
the new KÜFA furnace and its measurement devices at the ITU, Karlsruhe. This low-burnup
sphere was heated up to 1600°C and remained there for 5 hours before cooled down to 300°C.
The same procedure was repeated for 1800°C. During the test, the krypton release remained
below the detection limit. Apart from the typical expected contamination of AVR spheres, the
cesium measurements on the condensing plates could also be traced back to some additional
contamination of the plates and the aluminum bins, in which they were given. Therefore, the
measurements were of no use.

No need for postcalculation.

95
AVR 74/18

Characteristic data of experiment AVR 74/18

Burnup: 4.8% FIMA Fast neutron fluence: 0.8×1025, E>0.1MeV


Irradiation temp.: ~820°C Irradiation time: 480 efpd
Heating temperature: 1600, 1800°C Heating time: 100, 100 h

The heating test AVR 74/18 was conducted in the KÜFA-II furnace at ITU in a heating phase
at 1600°C over 100 h followed by a second heating phase at 1800°C over another 100 h. It
was the first real heating experiment in the new KÜFA-II furnace. With reaching the 1600°C
heating temperature, a krypton fractional release was measured at ~1×10-6, which increased in
the course of the test to 6×10-6, still below the level of the inventory of one coated particle.

Cesium

The experimental data of cesium release fraction from the fuel element remain below 10-5
indicating that its origin is mainly from the cesium contamination of this AVR sphere.

In agreement with this observation, the calculated release during the 1600°C heating phase
also appears to be a cleaning of the sphere from cesium contamination (Fig. 5-8a). But in
contrast to the observation, the predicted cesium release during the 1800°C heating phase
strongly increases to fractional release values of 9.3% from the coated particles (blue curve)
and 5.2% from the fuel sphere (red curve), respectively.

96
1 E+0

AVR 74/18 Cesium


1 E-1

1 E-2

1 E-3 Release from CP (820)


Failure fraction

Release from FE (820)


Cs-137 measurements
1 E-4 Temperature

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300

Heating time (h)

Figure 5-8a. FRESCO-II postcalculation of 137Cs fractional release for AVR 74/18.

Strontium

No strontium release measurements are available for this test.

The calculated strontium release (Fig. 5-8b) from the coated particles is significantly higher
than cesium release reaching ~0.6% at the end of the 1600°C phase (2.5×10-6 for Cs) and 17%
(9% for Cs) at the end of the test. The Sr release from the fuel sphere, however, remains as
expected in the contamination range during the 1600°C heating and only significantly
increases at 1800°C reaching a fractional release value of 3×10-3 at the end of the test.

Silver

The silver isotope 110mAg could not be identified any more due to the long time passed since
the end of irradiation.

The postcalculation (Fig. 5-8c) shows that the fractional release of silver rises rapidly eaching
values of 22% after 100 h heating at 1600°C and 72%, respectively, after 100 h more hours
heating at 1800°C.

97
1 E+0

AVR 74/18 Strontium

1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4

Release from CP (820)


Release from FE (820)
1 E-5
Sr-90 measurements
Temperature

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300

Heating time (h)

Figure 5-8b. FRESCO-II postcalculation of 90Sr fractional release for AVR 74/18.
1 E+0

AVR 74/18
Silver
1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4
Release from CP (820)
Release from FE (820)
1 E-5 Ag-110m measurements
Temperature

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300

Heating time (h)

Figure 5-8c. FRESCO-II postcalculation of 110mAg fractional release for AVR 74/18.

98
HFR-K6/2

Characteristic data of experiment HFR-K6/2

Burnup: 10.6% FIMA* Fast neutron fluence: 4.6×1025, E>0.1MeV


Irradiation temp.: ~940°C, max 1130°C Irradiation time: 634 efpd
Heating temperature: 1600, 1800°C Heating time: 100, 200 h
* corrected [Kühnlein 2003] (Petten value was 9.3% FIMA).

The heating test HFR-K6/2 was conducted in the KÜFA-II furnace at ITU in a heating phase
at 1600°C over 100 h (in three periods) followed by a second heating phase over another
200 h at 1800°C. Krypton fractional release remained below the detection limit during the
first 240 h of the test well into the 1800°C heating phase. An increase to 2×10-7 at 275 h was
measured, finally reaching 1×10-5 at the end, a value which is unexpectedly low for the given
heating history.

Although the final gas release does represent much less than the krypton inventory of one
coated particle, the assumption of one failed particle was made starting approximately
halfway into the 1800°C heating phase.

Cesium

The experimental data of cesium release fraction from the fuel element remain in the range of
1-4×10-5 during the 1600°C heating phases. For an MTR-irradiated fuel sphere and also
compared to the other heated HFR-K6 sphere (see next section), this is a relatively high
cesium release and might be explained with the presence of one or perhaps a few coated
particles with a failed SiC layer which easily release cesium but not krypton. Also the
diffusion-like release behavior of the krypton gives rise to the presumption of the presence of
one or several particles with a failed SiC layer (and still intact oPyC). Cesium fractional
release only starts rising significantly after about 100 h into the 1800°C phase with a final
value of 2.1×10-3.

The reference calculation shows that the Cs release during the 1600°C heating phases remains
at a low level, much lower than the measurements, and strongly increases with beginning of
the 1800°C heating phase far beyond the measurements (Fig. 5-9a). Final cesium fractional
release from the fuel sphere is 17% compared to measured 0.2%. To explain the large
discrepancy at the beginning, an additional calculation was made assuming a failed particle
from the beginning of the heating. Results (red dashed curve) show that the calculated release
level now is pretty close to the experimental data, while this assumption does not affect the
release behavior at 1800°C.

99
1 E+0

HFR-K6/2 Cesium
1 E-1

Release from CP (940)


1 E-2
Release from FE (940)
Cs-137 measurements
Temperature
1 E-3 Failure fraction
Release from CP (1 fail)
Failure fraction

Release from FE (1 fail)

1 E-4

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350 400

Heating time (h)

Figure 5-9a. FRESCO-II postcalculation of 137Cs fractional release for HFR-K6/2.

Strontium

No strontium release measurements are available for this test.

The calculated strontium release from the coated particles (blue curve in Fig. 5-8b) steeply
rises from the beginning reaching ~0.5% at the end of the 1600°C phases and 27% at the end
of the test. The Sr release from the fuel sphere, however, remains as expected in the
contamination range during the 1600°C heating and only significantly increases at 1800°C
reaching a fractional release value of ~2% at the end of the test.

Silver

The silver isotope 110mAg could not be identified any more due to the long time passed since
the end of irradiation.

The postcalculation shows a steeply increasing fractional release of silver with the fuel sphere
practically depleted of silver at the end of the test (Fig. 5-8c).

100
1 E+0

HFR-K6/2 Strontium

1 E-1

1 E-2

Release from CP (940)


1 E-3 Release from FE (940)
Failure fraction

Sr-90 measurements
Temperature
1 E-4 Failure fraction

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350 400

Heating time (h)

Figure 5-9b. FRESCO-II postcalculation of 90Sr fractional release for HFR-K6/2.


1 E+0

HFR-K6/2
Silver
1 E-1

1 E-2 Release from CP (940)


Release from FE (940)
Ag-110m measurements
1 E-3
Temperature
Failure fraction

Failure fraction

1 E-4

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350 400

Heating time (h)

Figure 5-9c. FRESCO-II postcalculation of 110mAg fractional release for HFR-K6/2.

101
HFR-K6/3

Characteristic data of experiment HFR-K6/3

Burnup: 10.9% FIMA* Fast neutron fluence: 4.8×1025, E>0.1MeV


Irradiation temp.: ~940°C, max 1140°C Irradiation time: 634 efpd
Heating temperature: 1600, 1700, 1800°C Heating time: 100, 100, 100+300 h
* corrected [Kühnlein 2003] (Petten value was 9.7% FIMA).

The heating test HFR-K6/3 was conducted in the KÜFA-II furnace at ITU in four heating
phases, the first at 1600°C over 100 h followed by a second heating phase at 1700°C over
another 100 h. The final two phases were conducted at 1800°C over 100 h and 300 h,
respectively. Krypton fractional release remained very low during the first three heating
phases, slightly increasing to 8.1×10-6. It was not until the fourth heating phase that a
significant Kr release occurred reaching a final value of 5.5×10-4, still extremely low
compared to what would habe been expected after 400 h heating at 1800°C. The observed
krypton release has been translated into a 5-step particle failure function with a total of 8
particles assumed to have failed.

Cesium

As can be seen in Fig. 5-10a, the cesium release measurements remain in the contamination
range during the first two heating phases and only start rising significantly with beginning of
the third, 1800°C heating phase. Fractional release values are 0.1% after 100 h at 1800°C, 1%
after about 210 h, and 4.3% at the end of the test after 400 h at 1800°C. The measurements are
well above the measured krypton release data indicating the presence of more coated particles
with a broken SiC layer.

The respective postcalculation is widely covering all experimental data showing for this test a
conservative character of the reference SiC diffusion coefficient. The eight coated particles
assumed to have failed have no influence on the cesium release behavior, since both
measurements and calculation are above that level.

102
1 E+0

HFR-K6/3
Cesium
1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4

1 E-5

1 E-6
Release from CP (940)
Release from FE (940)
1 E-7 Cs-137 measurements
Failure fraction
Temperature
1 E-8
0 50 100 150 200 250 300 350 400 450 500 550 600 650 700 750

Heating time (h)

Figure 5-10a. FRESCO-II postcalculation of 137Cs fractional release for HFR-K6/3.

Strontium

No strontium release measurements are available for this test.

The calculated strontium release (Fig. 5-10b) from the coated particles is steeply increasing
from the beginning reaching ~0.5% at the end of the 1600°C phase, 5% at the end of the
1700°C phase, 20% after 100 h at 1800°C, and 41% at the end of the test. The Sr release from
the fuel sphere, however, remains as expected at a low 1.4×10-7 during the 1600°C heating
and then increases to 1.1×10-4 after the 1700°C phase, 6.4×10-3 after 100 h at 1800°C, and
8.2% at the end of the test.

Silver

The silver isotope 110mAg could not be identified any more due to the long time passed since
the end of irradiation.

The postcalculation shows (Fig. 5-10c) that the fractional release of silver rises rapidly
reaching values of 21% after 100 h heating at 1600°C, 54% after 100 h more hours heating at
1700°C, and 99% after 400 h more hours heating at 1800°C.

103
1 E+0

HFR-K6/3
1 E-1

Strontium
1 E-2

1 E-3
Failure fraction

1 E-4

1 E-5

1 E-6

Release from CP (940)


Release from FE (940)
1 E-7
Sr-90 measurements
Failure fraction
Temperature
1 E-8
0 50 100 150 200 250 300 350 400 450 500 550 600 650 700 750

Heating time (h)

Figure 5-10b. FRESCO-II postcalculation of 90Sr fractional release for HFR-K6/3.


1 E+0

Silver
1 E-1

HFR-K6/3
1 E-2

1 E-3
Failure fraction

1 E-4

1 E-5

1 E-6

Release from CP (940)


Release from FE (940)
1 E-7 Ag-110m measurements
Failure fraction
Temperature
1 E-8
0 50 100 150 200 250 300 350 400 450 500 550 600 650 700 750

Heating time (h)


Figure 5-10c. FRESCO-II postcalculation of 110mAg fractional release for HFR-K6/3.

104
HFR-EU1bis/1

Characteristic data of experiment HFR-EU1bis/1

Burnup: 9.3% FIMA Fast neutron fluence: 2.4×1025, E>0.1MeV


Irradiation temp.: ~1250°C Irradiation time: 249 efpd
Heating temperature: 1250, 1600, 1700°C Heating time: 210, 200, 150 h

The heating test HFR-EU1bis/1 was conducted in the KÜFA furnace at ITU in three heating
phases, the first one at 1250°C over 210 h followed by a heating phase at 1600°C over 200 h,
and a final one at 1700°C over 150 h. The measured krypton release fraction was increasing
during the first heating phase to 1.4×10-6, further increasing in the second phase to 1×10-5, and
further increasing in the final phase to 2.5×10-5.

Cesium

Cesium release measurements are at a relatively high level (for a sphere with very low
contamination) above 10-5 and increase to 7.2×10-5 during the first heating phase despite the
low temperature of only 1250°C. A further increase of about one order of magnitude was
observed during the second, 200 h heating phase at 1600°C reaching the 10-3 level. Only a
few experimental data from the initial cooling plates of the third, 1700°C heating phase are
available (indicating a further increase) with the measurements discontinued due to high silver
activity (but are planned to be resumed later).

Looking at the reference postcalculation (solid curves in Fig. 5-11a), the predicted release
remains below the measurements and only coincide at the end of the 1600°C phase.

For the irradiation experiment HFR-EU1bis, it should be recalled that this was a high-
temperature test with a 1250°C irradiation temperature and that a temperature excursion
accidentally occurred at the beginning of the test, both effects of which may have resulted in a
breaching of the TRISO coating (and/or at least the SiC) in several to several tens of particles
(over all five spheres irradiated) during irradiation indicated by a significant increase in
fission gas release towards the end of the test. Therefore more calculations were conducted
varying the number of initially failed particles of 0 (reference), 1, 10, and 100. Results show
that the calculations approach the measurements with increasing number of failed particles,
but don’t quite reproduce the slope of the release curve.

The slope of the release curve, however would match with the measurements much better, if
assuming a larger diffusion coefficient for the cesium in the matrix graphite. This
presumption is derived from the result that the qualitative release curve from the particles (all
bluish curves) agrees pretty well with the slope of the measurements. An increase of the
diffusion coefficient in the graphite reduces the buffer effect and has the curve for release
from the fuel element approach that for release from the particles.

105
1 E+0

HFR-EU1bis/1 Cesium
1 E-1

1 E-2

1 E-3
Failure fraction

Release from CP (1250)


Release from FE (1250)
Cs-137 measurements
1 E-4
Temperature
Release from CP (1 fail)
Release from FE (1 fail)
1 E-5 Release from CP (10 fail)
Release from FE (10 fail)
Release from CP (100 fail)
Release from FE (100 fail)
1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350 400 450 500 550 600 650

Heating time (h)

Figure 5-11a. FRESCO-II postcalculation of 137Cs fractional release for HFR-EU1bis/1.

Strontium

No strontium release measurements are available for this test.

The calculated strontium release (Fig. 5-11b) shows again the large difference between the
high release from the coated particles (blue curve) and the delayed release from the fuel
sphere (red curve) due to the strong buffer effect of the matrix graphite.

Silver

The silver (110mAg) release measurement start at about the 10-3 level showing a moderate
increase to 2.3×10-3 at the end of the 1250°C heating and further increasing to 6.8×10-3 during
the 1600°C.

The observed release behavior is less pronounced than the calculated curves based on the
reference diffusion coefficients. The calculated values of 1.8% at the end of the 1250°C phase
and 36% at the end of the 1600°C are a factor of 8 and 53, respectively, larger than the
measurements.

106
1 E+0
Strontium
HFR-EU1bis/1
1 E-1

1 E-2

Release from CP (1250)


1 E-3
Release from FE (1250)
Failure fraction

Sr-90 measurements
1 E-4 Temperature

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350 400 450 500 550 600 650

Heating time (h)

Figure 5-11b. FRESCO-II postcalculation of 90Sr fractional release for HFR-EU1bis/1.


1 E+0

HFR-EU1bis/1
Silver
1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4
Release from CP (1250)
Release from FE (1250)
1 E-5 Ag-110m measurements
Temperature

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350 400 450 500 550 600 650

Heating time (h)

Figure 5-11c. FRESCO-II postcalculation of 110mAg fractional release for HFR-EU1bis/1.

107
HFR-EU1bis/3

Characteristic data of experiment HFR-EU1bis/3

Burnup: 11.1% FIMA Fast neutron fluence: 2.9×1025, E>0.1MeV


Irradiation temp.: ~1250°C Irradiation time: 249 efpd
Heating temperature: 1250, 1600°C Heating time: 100, 200 h

The heating test HFR-EU1bis/3 was conducted in the KÜFA-II furnace at ITU in two heating
phases, the first one at 1250°C over 90 h followed by a second heating phase at 1600°C over
another 200 h. Krypton fractional release remained at a low level throughout the test,
gradually increasing to a final value of 2.3×10-6.

Cesium

Regarding cesium release measurements, there is already a significant increase during the
1250°C heating which is probably again due to the early presence of particles with failed
TRISO/SiC. The release value after 90 h at 1250°C is 2.4×10-5. Enhanced release is observed
during the 1600°C heating with a final value of 2.1×10-3 after 200 more hours at this
temperature.

Also in this EU1bis sphere, the calculation curves remain below the measurements, but can be
approached pretty well with the (realistic) assumption of 10 particles that failed right from the
start. There is, however, a discrepancy remaining at the very beginning where release
measurements are already at a higher level which cannot be reproduced by the calculation.

108
1 E+0
Cesium
HFR-EU1bis/3
1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4

1 E-5
Release from CP (1250)
Release from FE (1250)
1 E-6
Cs-137 measurements
Temperature
1 E-7
Release from CP (10 fail)
Release from FE (10 fail)
1 E-8
0 50 100 150 200 250 300 350

Heating time (h)

Figure 5-12a. FRESCO-II postcalculation of 137Cs fractional release for HFR-EU1bis/3.

Strontium

No strontium release measurements are available for this test.

The calculated strontium release (Fig. 5-12b) shows again the large difference between the
high release from the coated particles (blue curve) and the delayed release from the fuel
sphere (red curve) due to the strong buffer effect of the matrix graphite. Strontium release
from this sphere is predicted to remain below 10-5.

Silver

The silver (110mAg) release measurements exhibit a comparatively flat release profile where
even the heating temperature increase from 1250 to 1600°C does not cause a perceivable
change. The experimental data are 1×10-3 at the end of the 1250°C heating and 3.6×10-3 at the
end of the 1600°C phase.

The calculated release are broadly enveloping the experimental data with values of ~1% at the
end of the 1250°C phase and ~35% at the end of the 1600°C.

109
1 E+0
Strontium
HFR-EU1bis/3
1 E-1

1 E-2

Release from CP (1250)


1 E-3
Failure fraction

Release from FE (1250)


Sr-90 measurements
1 E-4 Temperature

1 E-5

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350

Heating time (h)

Figure 5-12b. FRESCO-II postcalculation of 90Sr fractional release for HFR-EU1bis/3.

1 E+0

HFR-EU1bis/3
1 E-1

Silver
1 E-2

1 E-3
Failure fraction

Release from CP (1250)


1 E-4
Release from FE (1250)
Ag-110m measurements
1 E-5 Temperature

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300 350

Heating time (h)

Figure 5-12c. FRESCO-II postcalculation of 110mAg fractional release for HFR-EU1bis/3.

110
HFR-EU1bis/4

Characteristic data of experiment HFR-EU1bis/4

Burnup: 11.1% FIMA Fast neutron fluence: 2.9×1025, E>0.1MeV


Irradiation temp.: ~1250°C Irradiation time: 249 efpd
Heating temperature: 800, transient, 1720°C Heating time: 50, 90, 140 h

The heating test HFR-EU1bis/4 was conducted in the KÜFA-II furnace at ITU in a heating
phase at 800°C over 50 h followed by a transient temperature phase simulating the core
heatup accident scenario for a 600 MW block-type reactor until the maximum temperature of
1720°C was reached. Isothermal heating at that level was continued over another 140 h.
Krypton fractional release was expectedly low during the 800°C phase, then increased during
the transient to 4.1×10-6 and further increased during the isothermal 1720°C phase to a final
value of 1.7×10-5.

Cesium

For this fuel sphere initially heated at not more than 800°C over 50 h, a flat release profile
was measured in this phase, but already at the 10-5 level meaning that this amount of cesium
still must have been stored in the sphere at the end of irradiation and escaped at an early stage
into the heating test. With beginning of the temperature transient, the fractional release of
cesium strongly increases reaching a final value of ~1% at the end of the test.

Like in the preceding two examples, the calculated cesium release from the fuel sphere
remains below the measurements except for the final phase of the transient. And again, the
difference between observation and prediction is getting smaller when assuming 10 coated
particles to have failed from the beginning.

111
1 E+0

HFR-EU1bis/4 Cesium
1 E-1

1 E-2

1 E-3
Failure fraction

Release from CP (1250)

1 E-4
Release from FE (1250)
Cs-137 measurements
Temperature
1 E-5 Release from CP (10 fail)
Release from FE (10 fail)

1 E-6

1 E-7

1 E-8
0 50 100 150 200 250 300

Heating time (h)

Figure 5-13a. FRESCO-II postcalculation of 137Cs fractional release for HFR-EU1bis/4.

Strontium

No strontium release measurements are available for this test.

The calculated strontium release (Fig. 5-13b) from the coated particles (blue curve) is steeply
increasing with reaching the temperature of 1250°C and above. As expected, the release of
strontium from the fuel sphere is largely delayed and only with reaching the maximum
temperature of 1720°C, the release from the fuel sphere (red curve) exceeds the contamination
level. Predicted strontium release from this sphere at the end of the test is ~10-3.

Silver

The silver (110mAg) release measurements exhibit during the heatup transient an increase to
the percentage level and then, during the constant 1720°C phase, continues with a
comparatively flat release profile with only a small further increase to 2.5% at the end of the
heating.

The calculated release of silver is, apart from the initial stage before the heatup actually
started, in good agreement with the measurements, conservatively covering the experimental
data. The final value of ~53% silver release at the end of the test is roughly by a factor of 20
higher than the corresponding measurement.

112
1 E+0

HFR-EU1bis/4 Strontium

1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4

1 E-5

1 E-6

Release from CP (1250)


1 E-7 Release from FE (1250)
Sr-90 measurements
Temperature
1 E-8
0 50 100 150 200 250 300

Heating time (h)

Figure 5-13b. FRESCO-II postcalculation of 90Sr fractional release for HFR-EU1bis/4.


1 E+0

HFR-EU1bis/4 Silver
1 E-1

1 E-2

1 E-3
Failure fraction

1 E-4

1 E-5

1 E-6

Release from CP (1250)


1 E-7 Release from FE (1250)
Ag-110m measurements
Temperature
1 E-8
0 50 100 150 200 250 300

Heating time (h)

Figure 5-13c. FRESCO-II postcalculation of 110mAg fractional release for HFR-EU1bis/4.

113
HFR-EU1bis/5

Characteristic data of experiment HFR-EU1bis/5

Burnup: 9.7% FIMA Fast neutron fluence: 2.5×1025, E>0.1MeV


Irradiation temp.: ~1250°C Irradiation time: 249 efpd
Heating temperature: between 950°C and Heating time: 980 h (incl. heatup
1800°C and cooldown)
SiC strength: 834 MPa Weibull modulus: 8.02

The heating test HFR-EU1bis/5 was conducted in the KÜFA-II furnace at ITU. The sphere
was exposed to a rather complex heating temperature history. It is characterized by
consecutive periods of constant heating temperatures at various levels plus, towards the end, a
period of 8 rapid cycles ramping up and down between room temperature and 1800°C. The
total test extended over a time of almost 1000 hours, of which 164 h were at 1800°C (not
included the cycles). Krypton release remained very low throughout the test.

Cesium

Cesium release measurements for this test are still pending.

Strontium

No strontium release measurements are available for this test.

Silver

Silver release measurements for this test are still pending.

114
6. VALIDATION STUDIES IN THE PAST (1983-88)

6.1. First Validation Calculations

With the development of the FRESCO-II code [Krohn 1983], it became now possible to
consider the preceding irradiation/normal operation phase to be more precise in the
calculation of fission product release from heated fuel spheres. First calculations were – again
– dedicated to heating tests on AVR-irradiated spheres conducted in the A.-Test furnace. This
equipment allowed (not more than) the comparison of the cesium inventory measurements
before and after a heating phase.

The essential input parameters to be checked by the postcalculations were the Arrhenius-type
diffusion coefficients, particularly the important one of cesium in silicon carbide. The
reference data for cesium at that time are listed in the following Table 6-1:

Table 6-1. Reference diffusion coefficients for cesium in the fuel element materials as used for
the first FRESCO calculations.
Material Pre-exponential factor Activation energy Remark
[m2/s] [kJ/mol]
(Th,U)O2 4.5×10-12 117 Irradiation
3.0×10-6 300 Accident
LTI 6.3×10-8 222
-1
HTI 1.8×10 493
-11
SiC 1.8×10 181 Allelein
-4
Matrix graphite 2.0×10 181 Hoinkis

In Fig. 6-1, the postcalculations are shown for a heating test with an AVR GO-3 sphere
(14.4% FIMA burnup) with BISO particles which was heated within 16 h from 1000°C up to
2000°C and then staying at this level for 5 more hours. For some reason, the heating ramp up
to 2000°C was much shorter (3.6 hours) in the calculation; for both figures, the beginning of
the isothermal 2000°C phase is at t = 5.6 h. The measured 137Cs release at the end of the test
was 74% (Harwell: 72%). The calculated 137Cs release from the fuel element using reference
diffusion coefficients remained with 42% somewhat below the experimental data. The results
for strontium differed even more with regard to release from the sphere with calculated 3%
versus measured 92%. In this case of a high heating temperature, the assumed irradiation
temperature (730°C) did not really matter with regard to the data at the end of the test; a
higher value (1050°C) resulted only in a higher starting level at the beginning.

Figure 6-2 presents the calculated transient inventories for kernel, buffer, and HTI layer for
the case with Tirr = 730°C. The kernel inventory of ~25% (or ~18% for Tirr = 1050°C) is much
higher than the measured inventory of < 1%. Problem here is obviously the diffusion model
applied for the whole coated particle. While the author explained the difference with a
potential adsorption effect for cesium in the buffer layer, it might also be wise to take into
consideration the replacement of Fickian diffusion in the kernel with the Booth model (which
would predict an inventory of only ~20% left just after 5 h of heating at 2000°C and not
included yet the 16-hours heating ramp and) which would be closer to the measurement.

115
Figure 6-1. Prediction of cesium fractional release from an AVR GO-3 fuel sphere at 2000°C
heating temperature (Tirr = 730°C).

Figure 6-2. Prediction of cesium distribution in the particles of an AVR GO-3 fuel sphere
at 2000°C heating temperature (Tirr = 730°C).

The second example described here refers to a postcalculation of the heating test
FRJ2-K11/03 with TRISO coated particles, which was heated at 1400°C over 100 h, 1500°C
over 100 h, and three times at 1600°C over 50 h, 50 h, and 38 h, respectively. Cesium release
measurements above the detection limit (1%) are available for the end of the second and third
1600°C phase, which were 4.5% and 14.5% fractional release, surprisingly high according to
today’s standards. Figure 6-3 shows the reference postcalculation based on the reference
diffusion coefficient for cesium in SiC [Allelein 1980] and an irradiation temperature of
1050°C with the measurement data way above the calculation.

116
Figure 6-3. Prediction of cesium fractional release from FRJ2-K11/03 at 1400, 1500, 1600°C
heating temperature (Tirr = 1050°C) with reference diffusion coefficient.

Figure 6-4. Prediction of cesium fractional release from FRJ2-K11/03


with modified diffusion coefficient.

An effort to come closer to the measurements (Fig. 6-4) was only successful after
modification of the cesium diffusion coefficient in SiC to the upper confidence limit of all
respective diffusion data from different authors existing at that time:

Cesium in SiC: Do = 6.76×10-3 m2/s Q = 459 kJ/mol

The third example treats the postcalculation of the heating test with sphere 4 of the irradiation
experiment R2-K12. This sphere contained US fuel composed of fissile particles with 200 μm
UC2 kernels and fertile particles with 500 μm ThO2 kernels (the latter type of particle was

117
ignored in the calculation). The heating test was conducted over 50 h at 1600°C followed by a
100 h heating at 1600°C and a 50 h at 1700°C. The diffusion coefficient chosen for UC2 was
the same as for the oxide kernel due to lack of data, but also due to the fact that the kernel
diffusion coefficient was found to be not very sensitive to the calculation results.

Similar to the previous example, the postcalculation has revealed an underestimation of the
experimental data when the reference diffusion coefficient for cesium in SiC was assumed,
and only came close to the measurements,, when a modified diffusion coefficient, the same
used in the previous example was applied (Fig. 6-5).

Figure 6-5. Prediction of cesium fractional release from R2-K12/4 sphere heated at 1600 and
1700°C (Tirr = 1050°C) applying a modified diffusion coefficient.

Efforts on checking consistency of the FRESCO model and comparison with available
experimental data have been starting actually with the – older – FRESCO-I code [Krohn
1982]. The principles of these two consecutively developed codes are the same despite some
modifications in the code programming. As an example, the release behavior of iodine is
shown in the following two figures.

In Fig. 6-6, the integral fractional release of 131I from the core of the PNP-500 during an
unrestricted core heatup accident is given. The maximum temperature reached after about
30 h is ~2200°C. Iodine is expected to escape only from failed coated particles. Calculations
were done for two different particle failure functions which, according to the temperature
distribution in the core as a function of time, result in total integral failure fractions of ~2.2%
and ~2.8×10-4, respectively.

Figure 6-7 shows the transient integral release of iodine activity from the fuel for several
isotopes of iodine and the overall iodine release. It can be noted from a comparison that the
131
I activity release in relation to the total 131I activity is slightly higher than the particle
failure fraction which is due to the fact that the inventory of the shorter-lived 131I correlates
with the higher power production and higher fuel temperatures, respectively, in the upper part
of the OTTO core, which therefore exhibits the locally higher failure fractions.

118
Figure 6-6. Prediction with FRESCO-I of the integral iodine fractional release from the
PNP-500 core using two different particle failure functions (upper and lower curve) and
varying the diffusion coefficient for iodine in the kernel by a factor of 10 (dashed curves).

Figure 6-7. Prediction with FRESCO-I of the integral activity release from the PNP-500 core
for the different iodine isotopes assuming the lower particle failure function
(from previous figure).

119
6.2. Postcalculation of Heating Tests at FZJ

This section describes the results of the systematic validation of the FRESCO-II model made
against numerous experiments with spherical fuel elements containing TRISO-coated fuel
particles and heated at accident temperatures in the range of 1600-2500°C [Verfondern 1989].
The heating tests were conducted between end of the 1970s and beginning of the 1990s in the
A.-Test facility [Schenk 1983] and KÜFA furnace at FZJ [Schenk 1988, Schenk 1989,
Schenk 1997].

6.2.1. Objective of Postcalculations

The main objectives of the study were

 to apply the FRESCO model (selected in FZJ as the reference code for future studies)
to all available heating test data in order to qualify it as a proper tool for use in safety
and risk analyses for the HTGR concepts under discussion

 to check the existing set of recommended diffusion coefficients developed within the
HBK project against the new available results from heating tests and derive, if
necessary, new recommendations of diffusion coefficients for accident conditions.

6.2.2. Core Heatup Simulation Experiments

Over many years, heating experiments in the two furnaces operated in the Hot Cells at FZJ
have been one of the most important sources for transport data of fission products in HTGR
fuel elements.

The older A.-Test apparatus was a resistance furnace (Fig. 6-8) which could reach heating
temperatures up to 2500°C. The released fission gases were swept away with a helium flow
from the sphere, captured in cold traps with the 85Kr gas continuously monitored by means of
NaI detectors. All metallic fission products were collected in a trap. The heatup procedure
was defined such that the initial phase was simulating the normal operation conditions before
the sphere was further heated up to the desired accident temperature.

The release of the cesium isotopes 137Cs and 134Cs was determined by comparing the
inventories before and after the heating test. Postcalculations could therefore only refer to a
comparison with a single measured release value at the end of the heating test. To obtain more
than one measuring point, the experiments in this furnace were split into several phases, to
allow cesium measurements before and after each heating phase. Disadvantage of this method
was the temperature cycling down to room temperature with a potential effect on fuel
performance.

120
Figure 6-8. A.-Test apparatus for core heatup accident simulation testing of fuel spheres
up to 2500°C [Schenk 1983].

The cold finger apparatus (KÜFA) (Fig. 6-9), starting operation in 1983 at FZJ, allowed the
recording of the transient behavior of metallic fission product release from the fuel elements.
The possible heating temperature range was up to 1800 °C.

The helium purge gas from the KÜFA furnace is routed through activated charcoal cold traps
to capture the krypton and xenon fission gases released from the fuel. The traps were sitting
above Na(Tl) scintillation detectors that continuously monitored the 85Kr activity.

To permit a quasi on-line measurement of the release of metallic fission products, a water-
cooled steel condensation plate intruded into the heating furnace and was positioned directly
above the high-temperature zone. Water-cooling within the probe allows the surface
temperature of the condensate plate to be at ~100°C compared to the fuel specimen at
~1600°C. Thus, volatile fission or activation products released from the irradiated fuel
specimen will preferentially deposit on the low temperature condensate plate. Periodically, the
probe with the condensation plate in place was withdrawn from the furnace and replaced with
a new condensate plate. The time required to change and reinstall a new condensate plate was
on the order of 20 minutes. Using a calibration source, the plate-out efficiency of the
condensing plates was found to be ~70% for cesium, silver, iodine, whereas it was not more
than ~20% for strontium.

121
Figure 6-9. KÜFA apparatus for core heatup accident simulation testing used at FZJ and ITU
(top) and measurement of solid fission products from fuel spheres up to 1800°C
[Schenk 1988, Freis 2010].

122
The capability to repeatedly replace this condensation plate during KÜFA operation allows
the generation of time-dependent release curves for important gamma emitters isotopes like
137
Cs, 134Cs, 110mAg, 106Ru and 154Eu and the beta emitting 90Sr isotope. For the measurement
of 131I, because of its short half-life, the irradiated fuel element had to be reactivated and then
quickly transported to the hot-cell KÜFA facility for measurement. The inventory of beta-
emitting isotopes could be measured either by performing beta spectrometry on the
condensate plate or by removing the deposited fission products from the plate and measuring
the specific species using standard wet-chemistry techniques.

For some of the KÜFA experiments, the cooling plates were re-examined by AERE Harwell.
There are arge uncertainties for the strontium and silver measurements. With regard to
strontium, the Harwell data were considered to be more accurate compared to the FZJ data.

Experimental results from heating tests with modern HTGR spherical fuel elements beginning
in 1984 [Schenk 1988] however, showed a discrepancy in the high temperature release
behavior from single coated particles and from complete fuel balls, with the latter exhibiting
much lower release fractions. Due to a considerable retentivity of the matrix graphite for
metallic fission products, due to the optimized fuel manufacturing process which may have
influenced the fission product transport characteristics in the particle coating, and due to the
better statistics of about 104 particles per fuel element, experiments with loose particles were
no longer regarded as representative of modern high quality fuel.

6.2.3. Conduction of Calculations

For the postcalculation of the release of the isotopes 137Cs, 90Sr, and 110mAg, only heating tests
with TRISO fuel were considered. The study was actually concentrating on the diffusion
coefficient in the silicon carbide as the most important coating layer dominating the release
behavior. Those heated spheres which were deconsolidated afterwards and post-examined for
local fission product distribution would also allow conclusions with regard to the diffusion
coefficients in the matrix graphite. Finally, there were some heating experiments dedicated to
the study of release behavior from the particle kernel, which allowed checking of the diffusion
coefficients in the kernel material UO2. Derivation of new diffusion coefficients for accident
conditions

A typical way of evaluating a heating experiment with the diffusion model is a calculation
based on the recommended data set of diffusion coefficients to be compared with another one
where transport data have been varied leading to a good agreement with the measurements.

The diffusion coefficient in the silicon carbide layer as the most important and effective
barrier against particle release was generally taken as parameter for variation, as it does not
make much sense to change several transport data at a time especially if only a single release
value is available from the (A.-Test) experiment.

In the study described in this chapter, the reference diffusion coefficient, here “upper Myers”,
was modified by multiplying with a factor to adjust the calculated results to the
measurements. The factor, in practically all cases a reduction factor, was a pure fitting factor
obtained by a trial-and-error procedure. In a heating ramp test, it applied to all temperatures of
the ramp in the same way. In isothermal heating tests with distinct phases of a constant
temperature at different levels, a factor could be derived for each separate heating phase.

123
For the AVR spheres, an averaged irradiation temperature of constant 700°C over total
irradiation time was assumed. Furthermore, the AVR cross contamination level was assumed
as the initial heavy metal contamination.

The results of the postcalculations are presented in several series:

a. Heating tests with TRISO mixed oxide particles (chapter 6.2.4.1) and TRISO UO2
particles (chapter 6.2.4.2) to describe for both the calculation of transient release,
radial profiles in the sphere (chapter 6.2.4.3), derivation of new diffusion coefficients
for cesium (and strontium) in SiC at accident temperatures (chapter 6.2.4.4);

b. Heating tests with designed-to-fail (DTF) particles to describe calculation of transient


release (chapter 6.2.5.1), derivation of new diffusion coefficients for cesium (and
strontium) in UO2 (chapter 6.2.5.2);

c. Heating tests with shortly activated fuel spheres to describe calculation of transient
release from the uranium contamination of the matrix graphite (chapter 6.2.6).

Explanation of plots:

The diagrams for visualization of the transient release curves are given in different formats.
All curves in a plot are given as function of the heating time with the ordinate axis ranging
from 10-8 to 100. Red curves describe the release from the fuel element, blue curves the
release from the coated particles. The solid red line represents the optimized postcalculation
which is to be compared with the measurements (o: FZJ data, x; AERE Harwell data), while
the dashed red curves marked with ”Ref.” represent the fuel element release based on the
Myers reference diffusion coefficients. The optimization consists of the use of modified
diffusion coefficients for cesium in silicon carbide DSiC and in matrix graphite DG – chosen
reduction factors are written in the bottom line of a plot.

The dashed black curve describes the particle failure curve as a step function (as required by
the FRESCO input). The dotted (sometimes long-dashed) black curve in the bottom part of
the figure describes the heating temperature in a linear scale from 0°C (@ log fraction = -8) to
2500°C (@ log fraction = -5). The long-dashed, green line is a relict from the past and can be
ignored. It represents the so-called matrix retention factor (MRF) defined as the ratio of fuel
element release over coated particle release whose scale on the ordinate is going linearly from
0 (bottom) to 1 (top). Its value at the end, of the heating test is also written in the bottom line.

The box below the figure caption contains the factors, with which the respective diffusion
coefficients were modified to achieve the optimized curves.

124
6.2.4. FRESCO-II Results for Metallic Fission Product Release

Eight heating tests with irradiated fuel spheres containing HEU (Th,U)O2 TRISO coated
particles are considered here. Fuel batch is HT150-160 162-167. Seven spheres were of GO-2
type irradiated in the AVR since February 1981 (AVR-15). One sphere was irradiated in the
R2 MTR in Studsvik and contained the particle batch EO 1674.

6.2.4.1. Heating Tests with HEU (Th,U)O2 TRISO Fuel

Eight heating tests with irradiated fuel spheres containing HEU (Th,U)O2 TRISO coated
particles are considered here and listed in Table 6-2. Fuel batch is HT150-160 162-167. Seven
spheres were of GO-2 type irradiated in the AVR since February 1981 (AVR-15). One sphere
was irradiated in the R2 MTR in Studsvik and contained the particle batch EO 1674.

Table 6-2. FZJ heating tests of fuel spheres with HEU (Th,U)O2 TRISO coated particles.
Test Irradiation Burnup FZJ Heating
facility [% FIMA] heating facility temperature [°C]
AVR 69/13 AVR, Jülich 8.6 A.-Test 1800
AVR 69/28 AVR, Jülich 6.8 A.-Test 2150 ramp
AVR 70/18 AVR, Jülich 7.1 A.-Test 2400
AVR 70/26 AVR, Jülich 8.2 KÜFA 1600
AVR 74/17 AVR, Jülich 10.3 A.-Test 2500 ramp
AVR 74/20 AVR, Jülich 11.9 A.-Test 1900
AVR 74/24 AVR, Jülich 11.2 A.-Test 2100
R2-K13/1 R2, Studsvik 10.2 KÜFA 1600

125
AVR 69/13

Characteristic data of experiment AVR 69/13

Burnup: 8.6% FIMA Fast neutron fluence: 2.1×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1100 efpd
Heating temperature: 1800°C Heating time: 10+10+22+50 h

The heating test AVR 69/13 was conducted in the A.-Test furnace in four separate heating
phases at 1800°C. From the gas release measurements, a massive coated particle failure was
observed with a failure fraction (krypton release) of 9×10-3 reached by the end of the test. A
corresponding particle failure function was assumed in the FRESCO postcalculation.

Cesium

No cesium loss from the sphere could be measured after the first three 1800°C heating phases,
meaning < 1% due to the high gamma spectrometry detection limit of 1% for the A.-Test
furnace. At the test end, a high cesium release of 48% was measured. Besides, before heating
and after each of the first three heatings, Cs profiles in the fuel-free zone were measured.
While the initial profile was flat, later profiles were rising toward the fuel zone and were
increasing with heating time.

The postcalculation (Fig. 6-10) was trying to reproduce this final measurement point of 48%
which could be achieved only with an increase of the reference diffusion coefficient by a
factor of 10. Calculated cesium release from the almost 1% failed particles did not
significantly influence the overall cesium release. The experiment might have seen more SiC
failures with still intact oPyC which would raise the cesium release but keep the gas release
lower.

126
Figure 6-10. FRESCO-II postcalculation of 137Cs fractional release for AVR 69/13.

Cesium in SiC: 1.23×10-17 m2/s


Cesium in Matrix: 1987 reference

127
AVR 69/28

Characteristic data of experiment AVR 69/28

Burnup: 6.8% FIMA Fast neutron fluence: 1.7×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 870 efpd
Heating temperature: 1250-2150, 2150°C Heating time: 52, 6 h

The heating test AVR 69/28 was conducted in the A.-Test furnace with a slow heatup phase
within 52 h to 2150°C and a holding time of 6 h. Kr release measured was 4.4×10-5. This level
corresponds to less than the gas inventory of a single coated particle and is presumably mainly
due to release from contamination in the matrix.

Cesium

The diagram shows the single 2150°C cesium fractional release measurement point of 22%.
After the heating test, the sphere was deconsolidated and the Cs profile measured. The
relatively flat profile through the fuel zone at a high activity level shows that the cesium was
principally released from the coated particles. Three samples with 10 particles each were
taken from three different sphere regions and the particles crushed to measure the cesium
distribution resulting in inventories of 24% in the kernel, 51% in the coating and 3% in the
matrix, with 22% being released from the sphere. Ceramography exhibited clear sign of SiC
corrosive attack.

A reduction of the reference diffusion coefficient of cesium in SiC (Myers) by a factor of 25


leads to a reproduction of the measurement (Fig. 6-11). From the information of 3% cesium
inventory in the fuel element, adjustment could be extended to the A3 matrix diffusion
coefficient resulting in reduction factor of 0.8. The reference calculation predicts a cesium
release fraction from the sphere of more than 90%.

128
Figure 6-11. FRESCO-II postcalculation of 137Cs fractional release for AVR 69/28.

Cesium in SiC: 3.72×10-14 m2/s


Cesium in Matrix: 1987 reference × 0.8

129
AVR 70/18

Characteristic data of experiment AVR 70/18

Burnup: 7.1% FIMA Fast neutron fluence: 1.7×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 910 efpd
Heating temperature: 1250-2400°C Heating time: 28 h

The heating test AVR 70/18 was conducted in the A.-Test furnace in a 28 h heating ramp to
2400°C. Measured krypton release at the end of the test was 1.2%.

Cesium

Measured cesium fractional release was 82% providing no specific additional information.
From the comparatively low measured krypton release, it is concluded that the cesium release
behavior was not affected by the fraction of failed particles. In post-heating examinations, the
SiC was found mainly damaged by thermal decomposition.

The adjustment made in the cesium release calculation leads to a reduction of the 1987
reference diffusion coefficient by a factor of 5 (Fig. 6-12).

Figure 6-12. FRESCO-II postcalculation of 137Cs fractional release for AVR 70/18.

Cesium in SiC: 1.82×10-12 m2/s


Cesium in Matrix: 1987 reference

130
AVR 70/26

Characteristic data of experiment AVR 70/26

Burnup: 8.2% FIMA Fast neutron fluence: 2.0×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1050 efpd
Heating temperature: 1600°C Heating time: 312 h

The heating test AVR 70/26 was conducted in the KÜFA furnace in one heating phase at
1600°C over 312 h. No particle failure occurred during the heating test, since the measured
krypton fractional release remained for the whole time below 10-5.

Cesium

Cesium release from the fuel sphere begins at a level of below 10-5, rises then relatively
quickly and saturates at a level of 3×10-5. With regard to the release rate, there is a steady
decrease over the total heating time. The observed high starting level is due to the high
concentration in the surface layer resulting from the well-known cross contamination in the
AVR core. Decreasing release rates indicate no active cesium release from the coated
particles, and no particle failed during heating.

Figure 6-13a. FRESCO-II postcalculation of 137Cs fractional release for AVR 70/26.

Cesium in SiC: 2.82×10-17 m2/s


Cesium in Matrix: 1987 reference

131
For the postcalculation (Fig. 6-13a), the cesium diffusion coefficient was reduced by a factor
of ~33 to avoid significant release from the particles and to keep the release from the fuel
sphere at the contamination level over the whole testing range.

Strontium

Strontium release (1.7×10-5 at the end of the heating test) remains below the cesium release
curve. It does not shown the initial strong increase because of the slower transport through the
graphite. Release rate of Sr was found to be similar to Cs but with a considerable delay.

In the postcalculation (Fig. 6-13b), the diffusion coefficients for both SiC and A3 matrix were
adjusted to best reproduce the fuel element release curve. The calculated strontium release
from the coated particles, which significantly rises after ~40 hours into the 1600°C heating,
represents a conservative upper limit that allows reproducing the fuel element release curve. It
is presumably much lower.

90
Figure 6-13b. FRESCO-II postcalculation of Sr fractional release for AVR 70/26.

Strontium in SiC: Reference × 0.07


Strontium in Matrix: Reference × 0.4

Silver

Only one measurement for Ag release was made with a fractional release of 1.9% at the
beginning of the 1600°C heating phase.

132
AVR 74/17

Characteristic data of experiment AVR 74/17

Burnup: 10.3% FIMA Fast neutron fluence: 2.5×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1310 efpd
Heating temperature: 1250-2500°C Heating time: 27 h

The heating test AVR 74/17 was conducted in the A.-Test furnace in a 27 h heating ramp to
2500°C followed by a 9-hours cooldown phase to room temperature. This sphere showed
enhanced gas release with a measured Kr release fraction of 12%.

Cesium

Measured Cs fractional release was 83%. In order to meet this release value in the
postcalculation, the diffusion coefficient of cesium in SiC was reduced by a factor of 25
compared to the 1987 reference (Fig. 6-14). No credit was taken from the enhanced failure
rate of coated particles due to the observed gas release fraction of 12%. The assumption of a
particle failure fraction in the calculation would have further increased the reduction factor
(but this effect was not quantified).

Figure 6-14. FRESCO-II postcalculation of 137Cs fractional release for AVR 74/17.

Cesium in SiC: 7.94×10-13 m2/s


Cesium in Matrix: 1987 reference

133
AVR 74/20

Characteristic data of experiment AVR 74/20

Burnup: 11.9% FIMA Fast neutron fluence: 2.9×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1520 efpd
Heating temperature: 1250-1900, 1900°C Heating time: 15, 50 h

The heating test AVR 74/20 was conducted in the A.-Test furnace in a heating ramp from
1250°C to 1900°C within 15 hours followed by a 50-hours holding time at 1900°C.

Cesium

Measured fractional release of cesium was 43%. In order to meet this release value in the
postcalculation, the diffusion coefficient of cesium in SiC was reduced by a factor of ~3
compared to the 1987 reference (Fig.6-15).

Figure 6-15. FRESCO-II postcalculation of 137Cs fractional release for AVR 74/20.

Cesium in SiC: 1.86×10-14 m2/s


Cesium in Matrix: Reference

134
AVR 74/24

Characteristic data of experiment AVR 74/24

Burnup: 11.2% FIMA Fast neutron fluence: 2.7×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1430 efpd
Heating temperature: 1250,2100, 2100°C Heating time: 18, 30 h

The heating test AVR 74/24 was conducted in the A.-Test furnace in an 18 hours heatup
phase to 2100°C, a holding time of 30 hours at this temperature, and then a cooldown within
7 h. From the gas release measurements, a massive coated particle failure was observed
during the 2100°C heating phase with a failure fraction (krypton release) of 10% reached by
the end of the test. A corresponding particle failure function was assumed in the FRESCO
postcalculation.

Cesium

Measurement of Cs fractional release was 50%. In the postcalculation, the reference diffusion
coefficient of cesium in SiC was reduced by a factor of ~20 to meet the cesium measurement
of 50% (Fig. 6-16).

Figure 6-16. FRESCO-II postcalculation of 137Cs fractional release for AVR 74/24.

Cesium in SiC: 1.23×10-12 m2/s


Cesium in Matrix: 1987 reference

135
R2-K13/1

Characteristic data of experiment R2-K13/1

Burnup: 10.2% FIMA Fast neutron fluence: 8.3×1025, E>0.1MeV


Irradiation temp.: 1000-1200°C Irradiation time: 517 efpd
Heating temperature: 1600°C Heating time: 1000 h

The heating test R2-K13 was conducted in the KÜFA furnace in one heating phase at 1600°C
over 1000 hours, the longest ever heating test. After heating, the sphere (manufactured in
1979) was deconsolidated for further detailed analyses. The fuel element R2-K13/1 (also /4)
was made of A3-3. Krypton gas release remained at a very low level of < 10-6 for more than
200 h before gradually increasing to a final value of 3.5×10-4.

Cesium

Due to the high irradiation temperature and the fact that no particle failed, a part of the heavy
metal contamination was already released during irradiation, such that the fractional release
level of fission products in the heating test started from below 10-6. Cesium release was
steadily increasing reaching a final value after 1000 h of 1.5% from the KÜFA plates (it was
an estimated 2.1% from the – less accurate – difference of gamma measurements before and
after heating). Looking at the cesium release rate in this test, is was also rising for at least the
first 300 hours indicating that cesium was released from the particles.

Besides, before and after heating, Cs profiles in the fuel-free zone were measured. While the
initial profile was flat, the profile after heating was at a two orders of magnitude higher level
and slightly rising toward the fuel zone. Postheating gammaspectrometry measurements on
1925 singulized particles have revealed that 62 have released more than 15% of their 137Cs
inventory. From these, seven particles have failed early in the test with release fractions >
50%. About 30 particles had releases between 15 and 28% having failed at a late stage of the
test. Failed SiC still provide a certain degree of cesium retention.

A lot of effort has been done to postcalculate this 1000 h test at 1600 °C and the FRESCO
diffusion model still has problems with it as can be seen from Fig. 6-17a. The transient
cesium release behavior is not well reproduced. At least it was successful to get agreement
with the final conditions of fuel element release and (small) cesium inventory in the matrix
graphite. The different attempts have also shown an influence of the averaged irradiation
temperature and/or the number of particles with defective or failed coating layers (but no
through-coating failure). Disagreement with the measured transient release behavior might be
improved with the assumption of a concentration dependence of the diffusion coefficient
resulting in a slower transport at the beginning and a quicker transport later on.

136
Figure 6-17a. FRESCO-II postcalculation of 137Cs fractional release for R2-K13/1.

Cesium in SiC: 2.82×10-17 m2/s


Cesium in Matrix: 1987 reference × 0.8

Strontium

Different from cesium, the postcalculation of the strontium release with a final value after
1000 h of 1.2×10-3 has shown a relatively good agreement of the transient release curve with
the diffusion model (Fig. 6-17b).

Figure 6-17c shows corresponding results from a parameter study with modification of the
diffusion coefficients of strontium both in SiC and in matrix. The figure also includes a
second set of measurements on the same KÜFA plates in Harwell, UK, with supposedly more
accurate strontium measurements.

137
90
Figure 6-17b. FRESCO-II postcalculation of Sr fractional release for R2-K13/1.

Strontium in SiC: 1987 reference × 0.09


Strontium in Matrix: 1987 reference

90
Figure 6-17c. FRESCO-II postcalculation of Sr fractional release for R2-K13/1.

138
Silver

Massive silver release began already during the standard heatup phase. At the end of the 16
hours heating at 1250°C, from the initially existing 110mAg, a fractional release of 55% was
measured. When reaching the 1600°C, fractional release was 94%. With the first
measurement after 18 h at 1600°C, silver release was 100%. Similar to strontium, these FZJ
measured release data were considered conservative due to the, at that time, existing
inaccurate measurement methods for Ag and Sr. New Ag measurements of the same KÜFA
plates at Harwell has led to lower silver release data reaching after 1000 h a total fractional
release of 10%. In Fig. 6-17d, only the first 500 hours of the heating test are displayed,
revealing again that the qualitative silver release behavior cannot be reproduced by a simple
diffusion calculation.

Figure 6-17d. FRESCO-II postcalculation of 110mAg fractional release for the first 500 h of
R2-K13/1.

139
6.2.4.2. Heating Tests with LEU UO2 TRISO Fuel

This part of the study includes 21 heating tests with spherical fuel elements irradiated in the
AVR. The spheres were of GLE-3 type manufactured in 1981 and irradiated since July 1982
as the first set of high quality TRISO fuels in AVR (AVR-19). The spheres contained LEU
UO2 LTI TRISO coated particles of the fuel batch HT232-245. Twelve more fuel samples
heated after irradiation in MTRs including two spheres from the FRJ2 Jülich, two spheres and
five compacts from the HFR Petten, and three compacts from the Siloe reactor Grenoble. The
heating tests conducted are listed in Table 6-3; some characteristic data of the fuel are given
in Table 6-4.

The HFR-K3 irradiation experiment was considered a reference test for a steam-cycle HTGR
using fuel of the AVR-19 reload charge. This test with four spherical fuel elements, of which
two were heated after irradiation [Schenk 1989], belongs to the most well documented of the
German HTGR fuel program.

The irradiation test HFR-P4 was conducted to explore the potential limits of the German high-
quality TRISO fuel beyond the target limits of the HTR-Modul, while the SL-P1 tests were
focusing on the conditions up to the potential reactor burnup and neutron fluence. Particles
were embedded in cylindrical compacts machined from spherical fuel elements with a
reduced (~20 mm diameter) spherical fuel zone. Particle batch in the fuel samples of HFR-P4,
capsules 1 and 3, and in Sl-P1 was EUO2308. The capsule 2 of HFR-P4 contained coated
particles with a 51 μm thick SiC layer (batch EUO2309).

Table 6-3. FZJ heating tests with UO2 TRISO coated fuel particles.
Test Irradiation Burnup Heating facility Heating
facility [% FIMA] temperature [°C]
AVR 70/19 AVR, Jülich 2.2 A.-Test 2400 ramp
AVR 70/33 AVR, Jülich 1.6 A.-Test 1800
AVR 71/7 AVR, Jülich 1.8 A.-Test 2000
AVR 71/22 AVR, Jülich 3.5 KÜFA, Jülich 1600
AVR 73/12 AVR, Jülich 3.1 A.-Test 1900
AVR 74/6 AVR, Jülich 5.6 A.-Test 2100
AVR 74/8 AVR, Jülich 2.9 A.-Test 2500 ramp
AVR 74/10 AVR, Jülich 5.5 A.-Test 1800
AVR 74/11 AVR, Jülich 6.2 KÜFA, Jülich 1700
AVR 76/18 AVR, Jülich 7.1 KÜFA, Jülich 1800
AVR 76/19 AVR, Jülich 7.3 A.-Test 1900
AVR 76/27 AVR, Jülich 7.4 A.-Test 2100
AVR 76/28 AVR, Jülich 6.9 A.-Test 2100
AVR 80/14 AVR, Jülich 8.4 A.-Test 2500 ramp
AVR 80/16 AVR, Jülich 7.8 A.-Test 2000
AVR 80/22 AVR, Jülich 9.1 A.-Test 1900

140
AVR 82/9 AVR, Jülich 8.9 KÜFA, Jülich 1600
AVR 82/20 AVR, Jülich 8.6 KÜFA, Jülich 1600
AVR 88/15 AVR, Jülich 8.7 KÜFA, Jülich 1600, 1800
AVR 88/33 AVR, Jülich 8.5 KÜFA, Jülich 1600, 1800
AVR 88/41 AVR, Jülich 7.6 KÜFA, Jülich 1800
FRJ2-K13/2 DIDO, Jülich 8.1 KÜFA, Jülich 1600
FRJ2-K13/4 DIDO, Jülich 7.6 KÜFA, Jülich 1600/1800
HFR-K3/1 HFR, Petten 7.7 KÜFA, Jülich 1600
HFR-K3/3 HFR, Petten 10.2 KÜFA, Jülich 1800
HFR-P4/1-8 HFR, Petten 13.8 KÜFA, Jülich 1600
HFR-P4/1-12 HFR, Petten 11.1 KÜFA, Jülich 1600
HFR-P4/2-8 HFR, Petten 13.8 KÜFA, Jülich 1600
HFR-P4/3-7 HFR, Petten 13.9 KÜFA, Jülich 1600
HFR-P4/3-12 HFR, Petten 9.9 KÜFA, Jülich 1800
SL-P1/C6 Siloe, Grenoble 10.7 KÜFA, Jülich 1600
SL-P1/C9 Siloe, Grenoble 10.7 KÜFA, Jülich 1700
SL-P1/C10 Siloe, Grenoble 10.3 KÜFA, Jülich 1700

Table 6-4. Characteristic data of the UO2 TRISO particles used in the heating tests.
AVR GLE-3 FRJ2-K13 HFR-P4/1+3 HFR-P4/2 SL-P1
HFR-K3
Fuel LEU UO2 LEU UO2 LEU UO2 LEU UO2 LEU UO2
Particle batch HT232-245 EUO2308 EUO2308 EUO2309 EUO2308
No. of cp per 16,350 16,350 1631 1631 1666
sphere
Fraction 6.1×10-5 6.1×10-5 6.1×10-4 6.1×10-4 6.0×10-4
corresponding
to one cp

141
AVR 70/19

Characteristic data of experiment AVR 70/19

Burnup: 2.2% FIMA Fast neutron fluence: 0.6×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 303 efpd
Heating temperature: 1250-2400°C Heating time: 27 h

The heating test AVR 70/19 was conducted in the A.-Test furnace in a heatup ramp test to
2400°C within 27 hours. Measured krypton release was still at ~10-5 when passing 2200°C,
but then rapidly increased to 1%.

Cesium

Cesium release at the end of the test with this low-burnup sphere (2.2% FIMA) was 3.0%
which is relatively low for this temperature.

In the postcalculation, the red curve shows predicted Cs particle release fraction and the blue
curve release from the sphere (Fig. 6-18). In order to hit the final release value, the diffusion
coefficient of Cs in SiC had to be reduced by a factor of 200. Both curves for particle release
and sphere release are very close to each other in the final part of the ramp showing that there
is practically no holdup in the matrix graphite at temperatures above 2000°C.

Figure 6-18. FRESCO-II postcalculation of 137Cs fractional release for AVR 70/19.

Cesium in SiC: 4.57×10-14 m2/s


Cesium in Matrix: 1987 reference

142
AVR 70/33

Characteristic data of experiment AVR 70/33

Burnup: 1.6% FIMA Fast neutron fluence: 0.4×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 220 efpd
Heating temperature: 1800°C Heating time: 50+24.5+100 h

The heating test with the very low-burnup sphere AVR 70/33 was conducted in the A.-Test
furnace at 1800°C over 174.5 h in three phases of 50 h + 24.5 h + 100 h. Measured krypton
release values during the three heating phases were 5.1×10-4 , 9.9×10-4 , and 2.2×10-4 totaling
to 1.7×10-3. The number of particles that failed during the heating test was estimated to be
~40.

Cesium

Cesium release remained below the detection limit of 1% during the first two phases, while
during the third heating period, an increase to 2.2×10-2 release was measured. Increase was
slower compared to the MTR test spheres HFR-K3/3 (10.3% FIMA) and FRJ2-K13/4 (7.6%
FIMA) due to smaller burnup (1.6% FIMA) and fast neutron fluence (Fig. 6-19).

Figure 6-19. FRESCO-II postcalculation of 137Cs fractional release for AVR 70/33.

Cesium in SiC: 3.47×10-16 m2/s


Cesium in Matrix: 1987 reference

143
AVR 71/7

Characteristic data of experiment AVR 71/7

Burnup: 1.8% FIMA Fast neutron fluence: 0.5×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 248 efpd
Heating temperature: 1250-2000, 2000°C Heating time: 18, 50 + 18, 50 h

The heating test AVR 71/7 was conducted in the A.-Test furnace in two phases over 50 h each
with an 18 h ramp to reach the 2000°C. Measured krypton release for this low-burnup sphere
remained low with 2.5×10-6 after the first heating phase and a total of 8.6×10-5 after both
heating phases. Some damage to the SiC passing through the whole layer could be observed
after heating, but the extent of the damage was comparatively small, probably due to the low
burnup.

Cesium

Measured cesium release fractions were 5.0% and 4.2%, a total fractional release from the
sphere of 9.2%.

The postcalculation with FRESCO-II was done by trying to reproduce the final cesium
measurement (Fig. 6-20). This could be achieved by reducing the reference diffusion
coefficient in SiC by a factor of > 100. With these data, the ffirst cesium measurement is
slightly underpredicted. The assumption of a two-step function for particle failure according
to the observed krypton release reaching 1×10-4 in the 2nd heating phase did not influence the
cesium release which is several orders of magnitude higher and therefore dominated by
release from intact particles.

144
Figure 6-20. FRESCO-II postcalculation of 137Cs fractional release for AVR 71/7.

Cesium in SiC: 5.89×10-14 m2/s


Cesium in Matrix: 1987 reference

145
AVR 71/22

Characteristic data of experiment AVR 71/22

Burnup: 3.5% FIMA Fast neutron fluence: 0.9×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 481 efpd
Heating temperature: 1600°C Heating time: 500 h

The heating test AVR 71/22 was conducted in the KÜFA furnace at 1600°C over 500 h.
Measured krypton fractional release after 500 h was 4.0×10-7. After the long heating, a
minimal attack on the SiC structure from the inside could be identified.

Cesium

The measured cesium release increased soon to the expected level of matrix contamination
typical for AVR and then practically remained constant for the total heating phase with a final
fractional release value after 500 h of 2.0×10-5. Due to low burnup and fluence data, the
particles remained unaffected indicated by the observed low krypton release level and did not
contribute to cesium release from the sphere. It was also confirmed by postheating
examinations of fission product distributions on single particles that there was no evidence of
Cs release from the particles.

Figure 6-21a. FRESCO-II postcalculation of 137Cs fractional release for AVR 71/22.

Cesium in SiC: 1.86×10-17 m2/s


Cesium in Matrix: 1987 reference × 0.1

146
The red short-dashed curve in Fig. 6-21a represents the prediction based on the 1987
recommended diffusion coefficient and shows clearly that this diffusion coefficient must be
reduced. Full red line is predicted sphere release and dashed blue line particle release that
shows startup of Cs after 400 hours, an effect that is unfortunately hidden by the high
contamination levels.

Strontium

Strontium also remained below 10-5 during the whole heating phase showing that the main
source of Sr was from the contamination of the AVR sphere. The strontium release fraction
measured at the end was 5.3×10-6. The measured concentration profiles of Sr in the matrix
graphite was above the 10-3 level indicating release from the coated particles, but good
retention in the matrix.

In the postcalculation, the diffusion coefficients in SiC and matrix were selected such that the
red curve of Sr release from the fuel sphere remained at the contamination level, close to the
measurements (Fig. 6-21b). This could be achieved by reducing the SiC diffusion coefficient
by a factor of 20 and the matrix diffusion coefficient by a factor of ~12. This leads to a release
fraction from the coated particles of ~2.0×10-3 which is to be considered an upper limit of
particle release, because a further increase (by choosing a smaller reduction factor) would also
result in an increase of the fuel sphere release. From the overall experience with
postcalculations of Sr release, it is obvious that the recommended diffusion coefficient in SiC
is too high and too conservative, respectively.

90
Figure 6-21b. FRESCO-II postcalculation of Sr fractional release for AVR 71/22.

Strontium in SiC: 1987 reference × 0.05


Strontium in Matrix: 1987 reference × 0.08

147
Silver

Silver release measured was 9.0×10-4 after 500 h at 1600°C. The release curve is slightly but
steadily increasing from ~1.0×10-4 at the beginning of the 1600°C heating phase to its final
value.

148
AVR 73/12

Characteristic data of experiment AVR 73/12

Burnup: 3.1% FIMA Fast neutron fluence: 0.8×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 426 efpd
Heating temperature: 1900°C Heating time: 50+50 h

The heating test AVR 73/12 was conducted in the A.-Test furnace at 1900°C in two phases
over 50 h each. Measured krypton release remained low with 1.6×10-5 after the first heating
phase indicating no coating failure, and 1.4×10-4 after both phases indicating single SiC layer
failures during the second phase.

Cesium

Measured cesium release after the first phase remained below the detection limit of 1%; it
increased during the second heating phase to 10%. The postcalculation with FRESCO-II was
done by trying to reproduce the final cesium measurement (Fig. 6-22). This could be achieved
by reducing the reference diffusion coefficient in SiC by a factor of 40. With these data, the
first cesium measurement is overpredicted. The assumption of a three-step function for
particle failure according to the observed krypton release did not influence the cesium release
which is several orders of magnitude higher and dominated by release from intact particles.

Figure 6-22. FRESCO-II postcalculation of 137Cs fractional release for AVR 73/12.

Cesium in SiC: 1.55×10-15 m2/s


Cesium in Matrix: 1987 reference

149
AVR 74/6

Characteristic data of experiment AVR 74/6

Burnup: 5.6% FIMA Fast neutron fluence: 1.4×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 770 efpd
Heating temperature: 1250-2100, 2100°C Heating time: 18, 30 h

The heating test AVR 74/6 was conducted in the A.-Test furnace in a heating ramp to 2100°C
over 18 h plus a holding time of 30 h at that temperature. The measured krypton release was
2.4% at the end of the test. Severest damage to the SiC layer on numerous particles could be
observed after heating.

Cesium

The measured cesium release was 47%. Measured krypton release is 2.4%. In the
postcalculation with FRESCO-II, a particle failure function was not assumed despite the
observed gas release fraction of 2.4%. The measured cesium release of 47% is certainly
dominated by release from intact particles as would be expected at such high heating
temperatures (Fig. 6-23).

Figure 6-23. FRESCO-II postcalculation of 137Cs fractional release for AVR 74/6.

Cesium in SiC: 9.77×10-13 m2/s


Cesium in Matrix: 1987 reference

150
AVR 74/8

Characteristic data of experiment AVR 74/8

Burnup: 2.9% FIMA Fast neutron fluence: 0.7×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1100 efpd
Heating temperature: 1250-2500°C Heating time: 27 h

The heating test AVR 74/8 was conducted in the A.-Test furnace in a heating ramp test to
2500°C within 27 hours. Krypton release from this sphere remains at a low level < 10-5 until a
steep increase occurs after exceeding ~2200°C. Final value measured was 4.6%.

Cesium

Cesium release at the end of the test with this low-burnup sphere (2.9% FIMA) was 25%.
Measured krypton release at the end was 4.6%. Also in this postcalculation, a particle failure
function was not assumed despite the observed gas release fraction of 4.6%. The measured
cesium release of 25% is mainly determined by release from intact particles as would be
expected at such high heating temperatures (Fig. 6-24).

Figure 6-24. FRESCO-II postcalculation of 137Cs fractional release for AVR 74/8.

Cesium in SiC: 1.58×10-13 m2/s


Cesium in Matrix: 1987 reference

151
AVR 74/10

Characteristic data of experiment AVR 74/10

Burnup: 5.5% FIMA Fast neutron fluence: 1.4×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 756 efpd
Heating temperature: 1800°C Heating time: 30+30+30 h

The heating test AVR 74/10 was conducted in the A.-Test furnace in three 1800°C phases of
30 h each. Measured krypton release was 1.3×10-4, 8.3×10-4, and again 8.3×10-4, respectively,
during the single phases totaling to 1.8×10-3 at the end.

Cesium

Cesium release fractions were < 1% (below detection limit), 3.7%, and 4.2%, a total fractional
release from the sphere of 7.9%. In the postcalculation, with the assumption of an SiC
diffusion coefficient reduced by a factor of 10, the two measurement data above the detection
limit could be well reproduced (Fig. 6-25).

Figure 6-25. FRESCO-II postcalculation of 137Cs fractional release for AVR 74/10.

Cesium in SiC: 1.74×10-15 m2/s


Cesium in Matrix: 1987 reference

152
AVR 74/11

Characteristic data of experiment AVR 74/11

Burnup: 6.2% FIMA Fast neutron fluence: 1.6×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 853 efpd
Heating temperature: 1700°C Heating time: 184.5 h

The heating test AVR 74/11 was conducted in the KÜFA furnace at 1700°C over 185 h. The
pressure vessel failure of a coated particle occurred after 84 hours into the 1700°C heating
phase. The measured krypton fractional release was 3.0×10-5, those of xenon and iodine were
lower.Investigation after the heating test has shown serious damage to the SiC layer partially
penetrating the whole layer, obviously the result of the higher (than 1600°C) heating
temperature.

Cesium

Cesium release measured at the end of the heating test was 7.6×10-5 and 4.3×10-5 in the
Harwell measurement. Fission product concentration profiles through the heated sphere have
shown that particle failure occurred in localized positions. The cesium inventory of estimated
5 to 10 particles has been released into the matrix graphite during the 1700°C heating test.

Figure 6-26a. FRESCO-II postcalculation of 137Cs fractional release for AVR 74/11.

Cesium in SiC: 8.51×10-17 m2/s


Cesium in Matrix: 1987 reference × 0.02

153
The postcalculation was made with the reference diffusion coefficient of Cs in SiC reduced by
a factor of 50 (Fig. 6-26a). With the assumption of a sudden through-coating failure of one
particle, the particle release curve (blue) immediately jumped to the level of cesium inventory
of one failed particle (black short-dashed curve). Since it was approximately the level already
reached by the sphere release (red), mainly due to contamination of the matrix, it had hardly
any influence on the latter curve. The cesium inventory in the matrix graphite after heating
was measured to be 2.8×10-4 suggesting a strong reduction of the diffusion coefficient in the
matrix by a factor of 50.

Strontium

Strontium release was steadily increasing to 7.2×10-6 (8.3×10-5 in the FZJ measurement),
starting delayed in comparison to cesium due to the slower transport in matrix graphite, but
eventually exceeding the level of cesium release.

The ideal postcalculation of strontium release from the fuel sphere could be managed by
assuming a slight reduction of the diffusion coefficient in SiC and a larger reduction (factor of
~30) for that in the matrix graphite (Fig. 6-26b). This choice was based on the additional
knowledge from the postheating deconsolidation of the sphere revealing that at the end of the
test, 5.1% of the strontium were found in the matrix material.

90
Figure 6-26b. FRESCO-II postcalculation of Sr fractional release for AVR 74/11.

Strontium in SiC: 1987 reference × 0.8


Strontium in Matrix: 1987 reference × 0.03

154
Silver

Silver release steadily increased from the beginning. Final value measured at FZJ was
3.2×10-2, the measurement at Harwell was 4.8×10-2.

The postcalculation for silver was made with the reference data showing a conservative
coverage of the measured release (Fig. 6-26c). The single failed particle did not influence the
release behavior. Despite a high silver release from the coated particles, the measured silver
inventory after heating remained small (4.3×10-4) as expected due to rapid transport through
the matrix.

Figure 6-26c. FRESCO-II postcalculation of 110mAg fractional release for AVR 74/11.

155
AVR 76/18

Characteristic data of experiment AVR 76/18

Burnup: 7.1% FIMA Fast neutron fluence: 1.9×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1034 efpd
Heating temperature: 1800°C Heating time: 200 h

The heating test AVR 76/18 was conducted in the KÜFA furnace in a 12 h heating ramp to
1800°C followed by a 200-hours heating at that temperature and a 12-hours cooldown phase
to room temperature. There was a steadily increasing krypton release during the 1800°C
heating from the 10-8 level a final value of 1.2×10-4.

Cesium

Measured 137Cs release at the end of the 1800°C heating phase was 4.5×10-2, in good
agreement with Harwell data, and also compares well with the 4×10-2 derived from inventory
measurements in post-heating examinations. From the profile measurements after sphere
disintegration, high cesium concentrations were found in the matrix showing that cesium was
significantly released from the coated particles, presumably from particles with a failed SiC
layer as could be concluded from the much lower and steadily increasing krypton release
curve.

Figure 6-27a. FRESCO-II postcalculation of 137Cs fractional release for AVR 76/18.

Cesium in SiC: 1.29×10-15 m2/s


Cesium in Matrix: 1987 reference × 0.02

156
Postcalculation of the cesium (Fig. 6-27a) measurement could be achieved by reducing the
reference diffusion coefficient for SiC by a factor of 2 and, due to the knowledge of Cs
release from the coated particles, by reducing the reference diffusion coefficient for matrix by
a factor of 20. This test is one of the “good” examples where the calculated curve matches
excellently the diffusive release profile over the whole duration of the heating. This could be
achieved by reducing the 1987 reference diffusion coefficient by a factor of ~13 and that in
matrix graphite by a factor of 50 (after knowing the cesium inventory in the matrix measured
to be 14%).

Strontium

Fractional release value for strontium was 6.6×10-2 measured by Harwell (no FZJ data).
Measurements (symbols) represent a good diffusive release profile as shown in the figure in
comparison with the postcalculation.

The calculated curve (Fig. 6-27b) was obtained by reducing the SiC diffusion coefficient by a
factor of 5. No change was made for the matrix diffusion coefficient since no Sr inventory
measurement was made after sphere disintegration.

90
Figure 6-27b. FRESCO-II postcalculation of Sr fractional release for AVR 76/18.

Strontium in SiC: 1987 reference × 0.2


Strontium in Matrix: 1987 reference

157
Silver

The fractional release value of 62% for silver measured by FZJ is in fairly good agreement
with Harwell data. It is one of the few tests where the transient release corresponds
surprisingly well with the release profile from a diffusion calculation.

The FRESCO calculation shown here was conducted with reference data (Fig. 6-27c).

Figure 6-27c. FRESCO-II postcalculation of 110mAg fractional release for AVR 76/18.

158
AVR 76/19

Characteristic data of experiment AVR 76/19

Burnup: 7.3% FIMA Fast neutron fluence: 1.9×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1063 efpd
Heating temperature: 1250-1900, 1900°C Heating time: 15, 30 h

The heating test AVR 76/19 was conducted in the A.-Test furnace in a 15 h heating ramp to
1900°C followed by a 30-hours heating at that temperature and a 4-hours cooldown phase to
1250°C temperature. Measured krypton fractional release was with 6.9×10-7 extremely low.

Cesium

Measured Cs release at the end of the 1900°C heating phase was 46%.

Postcalculation of the single cesium measurement point could be achieved by reducing the
reference diffusion coefficient for SiC by a factor of 2.

Figure 6-28. FRESCO-II postcalculation of 137Cs fractional release for AVR 76/19.

Cesium in SiC: 3.09×10-14 m2/s


Cesium in Matrix: 1987 reference

159
AVR 76/27

Characteristic data of experiment AVR 76/27

Burnup: 7.4% FIMA Fast neutron fluence: 1.9×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1078 efpd
Heating temperature: 1250-2100, 2100°C Heating time: 19, 30 h

The heating test AVR 76/27 was conducted in the A.-Test furnace in a 19 h heating ramp to
2100°C followed by a 30-hours heating at that temperature and a 6-hours cooldown phase to
room temperature. Measurements of krypton release increased from contamination level
(4.6×10-7) to 3.0×10-5 when reaching the 2100°C level, soon during the isothermal heating
phase reaching the percentage range, and ending at 14%.

Cesium

Measured Cs release at the end of the test was 69%.

Postcalculation of the cesium measurement could be achieved by reducing the reference


diffusion coefficient for SiC by a factor of ~15.

Figure 6-29. FRESCO-II postcalculation of 137Cs fractional release for AVR 76/27.

Cesium in SiC: 1.74×10-12 m2/s


Cesium in Matrix: 1987 reference × 0.8

160
AVR 76/28

Characteristic data of experiment AVR 76/28

Burnup: 6.9% FIMA Fast neutron fluence: 1.8×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1005 efpd
Heating temperature: 1250-2100, 2100°C Heating time: 19, 30 h

The heating test AVR 76/28 was conducted in the A.-Test furnace in a 19 h heating ramp to
2100°C followed by a 30-hours heating at that temperature and a 6.5-hours cooldown phase to
room temperature. Measured krypton fractional release increased from contamination level
(< 10-7) to 2.5×10-4 when reaching the 2100°C level. It further increased to 4.8% at test end.

Cesium

Measured Cs release at the end of the test was 55%.

Postcalculation of the cesium was made with the assumption of a particle failure function
where in several steps, the final krypton release measurement was reached. The single cesium
measurement of 55% could be reproduced by reducing the reference diffusion coefficient for
SiC by a factor of ~16. The ~5% fraction of failed particles contributed only little to the
overall cesium release dominated by release from intact particles.

Figure 6-30. FRESCO-II postcalculation of 137Cs fractional release for AVR 76/28.

Cesium in SiC: 1.48×10-12 m2/s


Cesium in Matrix: 1987 reference

161
AVR 80/14

Characteristic data of experiment AVR 80/14

Burnup: 8.4% FIMA Fast neutron fluence: 2.2×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1223 efpd
Heating temperature: 1250-2500°C Heating time: 27 h

The heating test AVR 80/14 was conducted in the A.-Test furnace in a 27 h heating ramp
from 1250°C to 2500°C followed by a 9-hours cooldown phase to room temperature.
Measured krypton fractional release increased to 14% at test end.

Cesium

Measured Cs release at the end of the test was 99%. Measured krypton fractional release
increased to 14% at test end.

Postcalculation of the cesium measurement could be achieved by reducing the reference


diffusion coefficient for SiC by a factor of ~25. A particle failure function was not assumed.

Figure 6-31. FRESCO-II postcalculation of 137Cs fractional release for AVR 80/14.

Cesium in SiC: 7.94×10-13 m2/s


Cesium in Matrix: 1987 reference × 0.8

162
AVR 80/16

Characteristic data of experiment AVR 80/16

Burnup: 7.8% FIMA Fast neutron fluence: 2.0×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1136 efpd
Heating temperature: 1250-2000, 2000°C Heating time: 16, 30 h

The heating test AVR 80/16 was conducted in the A.-Test furnace in a 16 h heating ramp to
2000°C followed by a 30-hours heating at that temperature and a 6-hours cooldown phase to
room temperature. Measured krypton fractional release was 1.3×10-4 when reaching the
2000°C level. It further increased to 1.9% at test end.

Cesium

Measured Cs release at the end of the test was 22%. Measured krypton fractional release was
1.9% at test end.

Postcalculation of the cesium measurement could be achieved by reducing the reference


diffusion coefficient for SiC by a factor of ~16. A particle failure step function was assumed
as indicated in the figure, but it only marginally contribute to the overall cesium release.

Figure 6-32. FRESCO-II postcalculation of 137Cs fractional release for AVR 80/16.

Cesium in SiC: 5.01×10-13 m2/s


Cesium in Matrix: 1987 reference

163
AVR 80/22

Characteristic data of experiment AVR 80/22

Burnup: 9.1% FIMA Fast neutron fluence: 2.4×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1325 efpd
Heating temperature: 1250-1900, 1900°C Heating time: 15, 30 h

The heating test AVR 80/22 was conducted in the A.-Test furnace in a 15 h heating ramp to
1900°C followed by a 30-hours heating at that temperature and a 6-hours cooldown phase to
room temperature. First krypton release measurement was 4.0×10-4 when passing the 1600°C
level during the ramp. It further increased to 1.5% at test end.

Cesium

Measured Cs release at the end of the test was 4.8%.

Postcalculation of the cesium measurement could be achieved by reducing the reference


diffusion coefficient for SiC by a factor of ~16. The particle failure function was assumed to
be one step to a level of ~2×10-3. (However, I cannot recall, why there are no further steps up
to 1.5%. At least, it is conservative with regard to the reduction factor for the SiC diffusion
coefficient.)

Figure 6-33. FRESCO-II postcalculation of 137Cs fractional release for AVR 80/22.

Cesium in SiC: 3.72×10-15 m2/s


Cesium in Matrix: 1987 reference × 0.8

164
AVR 82/9

Characteristic data of experiment AVR 82/9

Burnup: 8.9% FIMA Fast neutron fluence: 2.5×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1296 efpd
Heating temperature: 1600°C Heating time: 500 h

The heating test AVR 82/9 was conducted in the KÜFA furnace in one heating phase at
1600°C over 500 h. The burnup of this sphere was 8.9% FIMA representative of the HTR-
Modul target burnup. Measured krypton release remained below 10-6 during the whole heating
test indicating zero particle failure with the measured krypton gas coming from contamination
of the matrix graphite only.

Cesium

Cesium release from the fuel sphere begins already at a level of 4.5×10-5, rises then to
3.4×10-4 after 108 h into the 1600°C heating phase and eventually reaches 7.6×10-4 at the end
of the test.

Figure 6-34. FRESCO-II postcalculation of 137Cs fractional release for AVR 82/9.

Cesium in SiC: 4.68×10-17 m2/s


Cesium in Matrix: 1987 reference

165
The high release level is explained only partly with the typical AVR fuel sphere
contamination. Another, unquantifiable contribution is given by a contamination of
components in the KÜFA furnace from the preceding heating test, which unfortunately could
not be replaced prior to this test for technical reasons. Therefore cesium release data are not
reliable.

For the postcalculation, the cesium diffusion coefficient was reduced by a factor of ~20 to
avoid significant release from the particles and to keep the release from the fuel sphere at the
(typical AVR) contamination level over the whole testing range. The KÜFA furnace
contamination was not taken account of.

Strontium

The first strontium measurement available at the beginning of the 1600°C heating phase is
4.6×10-6, increases further to 5.7×10-5 after 108 h into the 1600°C heating phase and
eventually reaches 8.3×10-5 at the end of the test.

No calculation

Silver

The silver measurements started with a value of 4.5×10-5 after 12 hours at 1050°C during the
heatup phase, reach 2.1×10-4 at the beginning of the 1600°C heating phase, increases further
to 6.8×10-4 after 108 h into the 1600°C heating phase, and eventually reach 1.9×10-2 at the
end of the test.

No calculation

166
AVR 82/20

Characteristic data of experiment AVR 82/20

Burnup: 8.6% FIMA Fast neutron fluence: 2.2×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1253 efpd
Heating temperature: 1600°C Heating time: 100 h

The heating test AVR 82/20 was conducted in the KÜFA furnace in one heating phase at
1600°C over 100 h. The 8.9% FIMA burnup of this sphere and estimated 2.2×1025 n/m2 fast
neutron fluence represent HTR-Modul target data. Measured krypton fractional release at the
end of the test was as low as 1.5×10-7 indicating no particle failure.

Cesium

Cesium release from the fuel sphere begins at a level of 5.1×10-6 and rises to 6.2×10-5 at the
end of the test.

For the postcalculation, the cesium diffusion coefficient in SiC was reduced by a factor of ~5
combined with a reduction of that in matrix by a factor of ~3 to follow as close as possible the
measured cesium release from the sphere. Reduction of the diffusion coefficients is
comparatively small. The experiment should have continued further to allow for a better
postcalculation.

Figure 6-35a. FRESCO-II postcalculation of 137Cs fractional release for AVR 82/20.

Cesium in SiC: 1.86×10-16 m2/s


Cesium in Matrix: 1987 reference × 0.3

167
Strontium

Strontium release from the fuel sphere gradually rises to 3.8×10-6 at the end of the test and
remains more than one order of magnitude lower than cesium release.

For the postcalculation, the Sr diffusion coefficents in SiC and in matrix were reduced by a
factor of 10 to arrive at the release curves as shown in the figure.

90
Figure 6-35b. FRESCO-II postcalculation of Sr fractional release for AVR 82/20.

Strontium in SiC: 1987 reference × 0.1


Strontium in Matrix: 1987 reference × 0.1

Silver

Silver release from the fuel sphere are available only for three moments during the initial
heatup process with a final value of 4.4×10-3 at the beginning of the 1600°C heating phase.

No calculation

168
AVR 88/15

Characteristic data of experiment AVR 88/15

Burnup: 8.7% FIMA Fast neutron fluence: 2.3×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1267 efpd
Heating temperature: 1600, 1800°C Heating time: 50, 50 h

The heating test AVR 88/15 was conducted in the KÜFA furnace in a heating phase at
1600°C over 50 h followed by a second heating phase at 1800°C over 21 h. Purpose of this
relatively short heating test was the measurement of iodine release. The 8.7% FIMA burnup
of this sphere and estimated 2.3×1025 n/m2 fast neutron fluence represent approximately HTR-
Modul target data. According to the measured krypton release, no particle failed during the
1600°C phase, whereas a couple of particles obviously have failed during the 1800°C heating
represented by the assumed particle failure step function.

Cesium

Cesium release from the fuel sphere begins at contamination level and rises significantly with
beginning of the 1800°C heating reaching a value slightly above 1% at the end of the test.

In the experiment AVR 88/15, no deposition of 131I on the condensing plate was discovered.
Since in this test, due to insufficient connection between condensing plate and cooling finger,
the temperature of the plate was relatively high, namely at ~200°C, it is considered proven
that iodine was released from the fuel sphere as element (boiling point: 184°C) and not as CsI
(boiling point: 1280°C). In contrast, microprobe examinations have shown that below
1800°C, iodine in the particle coating is available as CsI.

For the postcalculation, the cesium diffusion coefficient in SiC was reduced by a factor of ~30
at 1600°C and by a factor of ~10 at 1800°C to follow as close as possible the measured
cesium release from the sphere. The fact that measured cesium release from this high-burnup
fuel sphere is fairly high compared to the short heating time might be due to the presence of a
number of particles with a failed SiC layer (to easily release cesium) and a still intact oPyC
layer (to retain fission gases).

169
Figure 6-36. FRESCO-II postcalculation of 137Cs fractional release for AVR 88/15.

Cesium in SiC: 2.82×10-17 m2/s (1600°C)


1.38×10-15 m2/s (1800°C)
Cesium in Matrix: 1987 reference

170
AVR 88/33

Characteristic data of experiment AVR 88/33

Burnup: 8.5% FIMA Fast neutron fluence: 2.2×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1238 efpd
Heating temperature: 1600, 1800°C Heating time: 50, 21 h

The heating test AVR 88/33 was conducted in the KÜFA furnace in a heating phase at
1600°C over 50 h followed by a second heating phase at 1800°C over 21 h. Purpose of this
relatively short heating test was the measurement of iodine release. The 8.5% FIMA burnup
of this sphere and estimated 2.2×1025 n/m2 fast neutron fluence represent approximately HTR-
Modul target data. According to the measured krypton release, no particle failed during the
1600°C phase, whereas a couple of particles obviously have failed during the 1800°C heating
represented by the assumed particle failure step function.

Cesium

Cesium release from the fuel sphere shows from the beginning a significant increase reaching
a value of 1.2×10-4 after 50 h heating at 1600°C. Cesium release continues to rise at 1800°C.
Final value after the additional 21 hours heating at 1800°C is 4.4×10-4.

For the postcalculation, the cesium diffusion coefficient in SiC was reduced by a factor of ~30
at 1600°C and by a factor of ~10 at 1800°C to follow as close as possible the measured
cesium release from the sphere. The fact that measured cesium release from this high-burnup
fuel sphere is fairly high even during the 1600°C phase might be due to the presence of a few
particles with a failed SiC and still intact oPyC layer.

171
Figure 6-37. FRESCO-II postcalculation of 137Cs fractional release for AVR 88/33.

Cesium in SiC: 2.82×10-17 m2/s (1600°C)


1.74×10-15 m2/s (1800°C)
Cesium in Matrix: 1987 reference × 0.5

Strontium

The strontium release fraction increased during the 1600°C heating phase from 2.3×10-7 to
8.4×10-6. It further increased at 1800°C to 2.3×10-4.

No calculation

Silver

The silver release fraction increased during the 1600°C heating phase from 4.3×10-4 to
1.2×10-3. It further increased at 1800°C to 20%.

No calculation

172
AVR 88/41

Characteristic data of experiment AVR 88/41

Burnup: 7.6% FIMA Fast neutron fluence: 2.0×1025, E>0.1MeV


Irradiation temp.: ~700-1000°C Irradiation time: 1106 efpd
Heating temperature: 1800°C Heating time: 24 h

The heating test AVR 88/41 was conducted in the KÜFA furnace in one heating phase at
1800°C over 24 h. Purpose of this relatively short heating test was the measurement of iodine
release, but at the time of the heating test, six 131I half-lifes later, no iodine could be detected
any more. This, and also the measured krypton release, indicated that no particle failed during
the short heating period.

Cesium

Cesium release from the fuel sphere shows an increase from contamination level to slightly
above the contamination level reaching a value of 9.4×10-5 after the heating at 1800°C and
cooldown. The higher cesium release could be explained by a few particles with failed SiC
and still intact oPyC.

Figure 6-38. FRESCO-II postcalculation of 137Cs fractional release for AVR 88/41.

Cesium in SiC: 1.74×10-15 m2/s


Cesium in Matrix: 1987 reference

173
In the postcalculation (Fig. 6-38), the reference cesium diffusion coefficient in SiC was
reduced by a factor of 10. The test was actually too short to make a decent comparison.

Strontium

The strontium release fraction increased during the 1800°C heating phase from 3.9×10-6 to
1.2×10-4.

No calculation

Silver

The silver release fraction increased during the 1600°C heating phase from 5.4×10-4 to 7.7%.

No calculation

174
FRJ2-K13/2

Characteristic data of experiment FRJ2-K13/2

Burnup: 8.0% FIMA Fast neutron fluence: 0.2×1025, E>0.1MeV


Irradiation temp.: 990-1150°C Irradiation time: 396 efpd
Heating temperature: 1600°C Heating time: 138 h

The heating test FRJ2-K13/2 was conducted in the KÜFA furnace at 1600°C over 138 h. The
krypton release fraction remained at a very low level with 6.4×10-7 measured at the end
indicating no failed particle. The structure of the SiC after heating was found to be unchanged
compared to unirradiated material.

Cesium

Cesium release during the 1600°C heating phase increased from 1×10-5 to 3.9×10-5. The level
of early cesium release indicates the presence of one particle with a defective SiC layer.

With the assumption of a manufacture-induced particle defect, an improved poslcalculation


could be made with respect to fuel element release.

Figure 6-39a. FRESCO-II postcalculation of 137Cs fractional release for FRJ2-K13/2.

Cesium in SiC: 2.82×10-17 m2/s


Cesium in Matrix: 1987 reference × 0.1

175
Strontium

Strontium release measurements conducted at Harwell are only available for six condensing
plates applied during the 1600°C heating phase. The Harwell data yield a fractional release of
3.1×10-9 (not visible in the figure) increasing to 3.3×10-7. at the end.

90
Figure 6-39b. FRESCO-II postcalculation of Sr fractional release for FRJ2-K13/2.

Strontium in SiC: 1987 reference × 0.06


Strontium in Matrix: 1987 reference × 0.01

176
Silver

Only five condensing plates were measured at Harwell with regard to silver release (crosses in
figure). Fractional release was 2.8×10-3 at the end of the test.

Figure 6-39c. FRESCO-II postcalculation of 110mAg fractional release for FRJ2-K13/2.

177
FRJ2-K13/4

Characteristic data of experiment FRJ2-K13/4

Burnup: 7.6% FIMA Fast neutron fluence: 0.2×1025, E>0.1MeV


Irradiation temp.: 980-1120°C Irradiation time: 396 efpd
Heating temperature: 1600, 1800°C Heating time: 138, 100 h

The heating test FRJ2-K13/4 was conducted in the KÜFA furnace at constant 1600°C over
138 h and subsequently at constant 1800°C over 100 more hours indicated by the dotted line.
In this test, the krypton release fraction remained at a very low level with 3.0×10-7 measured
at the end of the 1600°C phase. One particle failed after 8 hours into the 1800°C phase and a
second one after 94.5 h and then a third one. Measured fractional release was 7.2×10-5.

Cesium

At the end of the 1600°C heating, cesium release was measured to be 2.5×10-6, while it
rapidly increased to ~1.0% during the 1800°C phase.

Figure 6-40a. FRESCO-II postcalculation of 137Cs fractional release for FRJ2-K13/4.

Cesium in SiC: 2.82×10-17 m2/s (1600°C)


1.74×10-15 m2/s (1800°C)
Cesium in Matrix: 1987 reference × 0.01

178
This 1600 / 1800°C test is another example for the excellent reproduction of the transient
cesium release behavior and the relation with coated particle and fuel element release with the
diffusion model. A reduction of the SiC diffusion coefficient by a factor of ~30 at 1600°C at
10 at 1800°C, respectively, plus a strong reduction of the matrix diffusion coefficient by a
factor of 100 compared to reference data was necessary to arrive at the measurements. No
influence of the observed particle failures during the 1800°C heating phase on the release
from the fuel element could be seen.

Strontium

Harwell measured 14 out of the 20 condensing plates used during the test, while the final six
plates were dropped because of too low activities. Strontium release was increasing during the
1600°C heating phase from 5.7×10-9 after 42 h (not visible in the figure) to 2.0×10-8 after
138 h, and then more rapidly increasing to 1.4×10-3 during the additional 100 h heating at
1800°C.

90
Figure 6-40b. FRESCO-II postcalculation of Sr fractional release for FRJ2-K13/4.

Strontium in SiC: 1987 reference × 0.02 (1600°C)


1987 reference × 0.3 (1800°C)
Strontium in Matrix: 1987 reference × 0.01 (1600°C
1987 reference × 0.09 (1800°C)

179
Silver

According to the Harwell measurements, silver fractional release was 4.5×10-4 (KFA: 1×10-3)
at the end of the 1600°C heating phase and 53% at the end of the additional 1800°C test. No
silver was identified at KFA on the final two plates; therefore, a 100% fractional release was
assumed.

Figure 6-40c. FRESCO-II postcalculation of 110mAg fractional release for FRJ2-K13/4.

180
HFR-K3/1

Characteristic data of experiment HFR-K3/1

Burnup: 7.7% FIMA Fast neutron fluence: 3.9×1025, E>0.1MeV


Irradiation temp.: 1045-1150°C Irradiation time: 359 efpd
Heating temperature: 1600°C Heating time: 500 h

See also chapter 8.4.2

The heating test HFR-K3/1 was conducted in the KÜFA furnace at 1600°C over 500 h.
Observed krypton release remained over more than 400 hours below 10-6 showing a gradual
increase to 1.8×10-6 after 500 h. Postheating examination has revealed an enhanced damage of
the SiC structure due to the long heating time in combination with the severe impact during
irradiation.

Cesium

Cesium, krypton and strontium releases remain extremely low during the first 200 h
demonstrating that this fuel element had no initial coating defects, a low manufacturing
contamination level, no in-pile coating failure and no particle failure during the initial phases
of the heating test. Cesium fractional release increases after 200 h reaching a final value of
1.2×10-4 (FZJ: 1.1×10-4) measured after 500 h.

Figure 6-41a. FRESCO-II postcalculation of 137Cs fractional release for HFR-K3/1.

Cesium in SiC: 3.80×10-17 m2/s


Cesium in Matrix: 1987 reference × 0.03

181
The deconsolidation of the sphere after heating has shows that by the end of the 500 h
heating, certain amounts of Cs and Sr had been released from the particles, but had been
strongly retained in the sphere. Cesium release from the particles was 1.2×10-3, while release
from the sphere was 1.1×10-4. The ceramographic sections through a number of particles
make a small degradation of the SiC layer visible.

The reduction of the 1987 reference diffusion coefficient by a factor of 20 is in agreement


with other postcalculations of 1600°C heating tests. In combination with a rather strong
reduction of the diffusion coefficient in matrix graphite by a factor of ~30, the measured ratio
of release from the fuel element over release from the coated particle could be reproduced
well. The overprediction by the reference calculation is up to 4 orders of magnitude.

Strontium

Strontium release remains at a very low level throughout the whole heating phase with
1.2×10-8 at the beginning of the 1600°C phase and reaching 1.8×10-7 after 500 h. The data
refer to the more accurate Harwell measurements.

90
Figure 6-41b. FRESCO-II postcalculation of Sr fractional release for HFR-K3/1.

Strontium in SiC: 1987 reference × 0.1


Strontium in Matrix: 1987 reference × 0.025
Contamination: 1987 reference × 0.1

182
Silver

Silver release is already above 1% when reaching the 1600°C heating temperature (1.64%).
The final fractional release value after 500 h is 2.4×10-2 (all Harwell data).

Figure 6-41c. FRESCO-II postcalculation of 110mAg fractional release for HFR-K3/1.

183
HFR-K3/3

Characteristic data of experiment HFR-K3/3

Burnup: 10.2% FIMA Fast neutron fluence: 6.0×1025, E>0.1MeV


Irradiation temp.: 740-1045°C Irradiation time: 359 efpd
Heating temperature: 1800°C Heating time: 25.5+74.5 h

(See also chapter 4.3.2.3)

The heating experiment HFR-K3/3 was conducted in the KÜFA furnace at 1800°C over 100 h
in two tests due to an unintentional interruption after ~25 h into the 1800°C heating. The
interruption was taken into account in the postcalculations. Measured krypton release after
100 h at 1800°C was 6.5×10-4.

Cesium

Cesium release strongly increases with heating at 1800°C to a value of 1.0×10-3 after 25 h.
With 5.9% for cesium after 100 h, the fractional release remains significantly below expected
levels under these conditions.

Figure 6-42a. FRESCO-II postcalculation of 137Cs fractional release for HFR-K3/3.

Cesium in SiC: 2.57×10-15 m2/s


Cesium in Matrix: 1987 reference × 0.1

184
At 1800° C, the influence of irradiation conditions is less pronounced than at 1600°C.
Postheating gammaspectrometry examination of particles has shown that the observed 137Cs
loss in these particles was ranging between 10 and 81%, underlining the stochastic nature of
the release process (which is not reproducible by a diffusion model).

With the reduction of the reference diffusion coefficients by a factor of ~7 for SiC and 10 for
matrix, the calculation is in very good agreement with the measured data. As the release level
was already in the order of percent, the small amount of failed particles did not influence the
calculated diffusive release behavior.

Strontium

Strontium release strongly increases with heating at 1800°C reaching 3.6×10-5 after 25 h (end
of heating part 1) and further increases to 1.8×10-3 after a total of 100 h at 1800°C (all
Harwell data).

Here is the rare case that the SiC diffusion coefficient was raised by 20% (whereas that in
matrix graphite was reduced by a factor of 10) to fit the measurement data.

90
Figure 6-42b. FRESCO-II postcalculation of Sr fractional release for HFR-K3/3.

Strontium in SiC: 1987 reference × 1.2


Strontium in Matrix: 1987 reference × 0.1

185
Silver

Silver release begins at a much lower level compared to HFR-K3/1 due to the lower (by
~200°) irradiation temperature, but strongly increases with the higher heating temperature of
1800°C to 8.8% after 25 h and 67% after 100 h (all KFA data).

The postcalculation shows a curve being above the measurements during the first part of the
experiment, but remaining below the measurements during the longer second part of the test.

Figure 6-42c. FRESCO-II postcalculation of 110mAg fractional release for HFR-K3/3.

186
HFR-P4/1-8

Characteristic data of experiment HFR-P4/1-8

Burnup: 13.8% FIMA Fast neutron fluence: 7.2×1025, E>0.1MeV


Irradiation temp.: 915-940°C Irradiation time: 351 efpd
Heating temperature: 1600°C Heating time: 304 h

Low in-pile R/B values of fission gases indicated no particle failure during irradiation. The
compact (small sphere) was heated within 7.5 h from 1250 to 1600°C and then remained at
1600°C over 304 h [Schenk 1997]. Measured krypoton release began to increase after about
80 h at 1600°C and reached a value of 5.4×10-5 at the test end.

Cesium

The measured fractional release of cesium was steadily increasing and reached a final value of
2.0×10-3. After deconsolidation of the compact, the cesium inventory in the matrix graphite
was measured to be 1.3×10-3.

Figure 6-43a. FRESCO-II postcalculation of 137Cs fractional release for HFR-P4/1-8.

Cesium in SiC: 1.17×10-16 m2/s


Cesium in Matrix: 1987 reference × 0.04

187
Strontium

The measured fractional release of strontium was gradually increasing and reached a final
value of 1.5×10-4. After deconsolidation of the compact, the strontium inventory in the matrix
graphite was measured to be 2.9×10-3.

90
Figure 6-43b. FRESCO-II postcalculation of Sr fractional release for HFR-P4/1-8.

Strontium in SiC: 1987 reference × 0.06


Strontium in Matrix: 1987 reference

Silver

Silver was below the detection limit and could not be measured.

188
HFR-P4/1-12

Characteristic data of experiment HFR-P4/1-12

Burnup: 11.1% FIMA Fast neutron fluence: 5.5×1025, E>0.1MeV


Irradiation temp.: 915-940°C Irradiation time: 351 efpd
Heating temperature: 1600°C Heating time: 304 h

See also chapter 4.3.2.1.

Low in-pile R/B values of fission gases indicated no particle failure during irradiation. The
compact (small sphere) was heated within 7.5 h from 1250 to 1600°C and then remained at
1600°C over 304 h [Schenk 1997]. Krypton release remained at a low level during the whole
test with a final fractional release of 5.4×10-7.

Cesium

Cesium release measurements start at a relatively high level > 10-5, which may be an
indication of the presence of a coated particle with a defective/failed SiC layer. The fractional
release was gradually increasing and reached a final value of 2.6×10-4. After deconsolidation
of the compact, the cesium inventory in the matrix graphite was measured to be 5.8×10-4.

Figure 6-44a. FRESCO-II postcalculation of 137Cs fractional release for HFR-P4/1-12.

Cesium in SiC: 7.41×10-17 m2/s


Cesium in Matrix: 1987 reference × 0.04

189
The calculated release curves remain below the measurements in the first half of the test, but
are rising at a higher rate and eventually exceeding the measurements.

Strontium

The measured fractional release of strontium was steadily increasing and reached a final value
of 6.0×10-6. After deconsolidation of the compact, the strontium inventory in the matrix
graphite was measured to be 6.7×10-4.

90
Figure 6-44b. FRESCO-II postcalculation of Sr fractional release for HFR-P4/1-12.

Strontium in SiC: 1987 reference × 0.01


Strontium in Matrix: 1987 reference × 0.1

Silver

Silver was below the detection limit and could not be measured.

190
HFR-P4/2-8

Characteristic data of experiment HFR-P4/2-8

Burnup: 13.8% FIMA Fast neutron fluence: 7.2×1025, E>0.1MeV


Irradiation temp.: 920-945°C Irradiation time: 351 efpd
Heating temperature: 1600°C Heating time: 304 h

Low in-pile R/B values of fission gases indicated no particle failure during irradiation. The
compact (small sphere) was heated within 7.5 h from 1250 to 1600°C and then remained at
1600°C over 304 h [Schenk 1997]. Krypton release began to increase after ~140 h at 1600°C
and reached a value of 5.4×10-5 at the test end.

Cesium

The measured fractional release of cesium was steadily increasing and reached a final value of
1.4×10-3. After deconsolidation of the compact, the cesium inventory in the matrix graphite
was measured to be 8.0×10-4.

Figure 6-45a. FRESCO-II postcalculation of 137Cs fractional release for HFR-P4/2-8.

Cesium in SiC: 1.91×10-16 m2/s


Cesium in Matrix: 1987 reference × 0.04

191
Strontium

The measured fractional release of cesium was steadily increasing and reached a final value of
1.1×10-4. After deconsolidation of the compact, the strontium inventory in the matrix graphite
was measured to be 2.2×10-3.

90
Figure 6-45b. FRESCO-II postcalculation of Sr fractional release for HFR-P4/2-8.

Strontium in SiC: 1987 reference × 0.09


Strontium in Matrix: 1987 reference

192
Silver

Silver was below the detection limit and could not be measured.

Since this fuel specimen contained particles of batch EUO2309 with a 50 μm thick SiC layer,
calculations were made to quantify the effect of the thicker SiC layer versus the reference
thickness. The difference in the transient release curves (during isothermal heating at 1600°C)
is small such that a significant reduction of the silver source term ist not expected to justify an
increase of the reference layer thickness.

Figure 6-45c. FRESCO-II postcalculation of 110mAg fractional release for HFR-P4/2-8.

193
HFR-P4/3-7

Characteristic data of experiment HFR-P4/3-7

Burnup: 13.9% FIMA Fast neutron fluence: 7.5×1025, E>0.1MeV


Irradiation temp.: 1050-1075°C Irradiation time: 351 efpd
Heating temperature: 1600°C Heating time: 304 h

Low in-pile R/B values of fission gases indicated no particle failure during irradiation. The
compact (small sphere) was heated within 7.5 h from 1250 to 1600°C and then remained at
1600°C over 304 h [Schenk 1997].

In this test, three particle failures have been observed. after 47 h, an obvious pressure vessel
failure of a particle occurred. Two more increases in gas release could be identified which
may be connected with particle failures, after 115 h, and after 200 h. A total of 9.9×10-4 or
about 1.6 times the inventory of a particle was released by the end of the test.The assumption
to be considered in the postcalculation was that the 1st coated particle failed after 49 h @
1600°C, the 2nd after 115 h, and the 3rd after 200 h.

Cesium

The measured fractional release of cesium was steadily increasing and reached a final value of
3.9×10-3. After deconsolidation of the compact, the cesium inventory in the matrix graphite
was measured to be 1.4×10-3. The observed release is significantly higher compared to
1600°C heating tests with fuel of lower burnup < 10% FIMA.

The postcalculation is in much better agreement with the calculations than compact HFR-
P4/1-12. The particle failures can be clearly identified in the cesium release curve of
FRESCO-II from the sudden puff releases.

194
Figure 6-46. FRESCO-II postcalculation of 137Cs fractional release for HFR-P4/3-7.

Cesium in SiC: 1.10×10-16 m2/s


Cesium in Matrix: 1987 reference × 0.04

Strontium

The measured fractional release of strontium was steadily increasing and reached a final value
of 2.4×10-4. After deconsolidation of the compact, the strontium inventory in the matrix
graphite was measured to be 1.9×10-3.

Silver

The measured fractional release of silver reached a final value of 3.9×10-4.

195
HFR-P4/3-12

Characteristic data of experiment HFR-P4/3-12

Burnup: 9.9% FIMA Fast neutron fluence: 5.5×1025, E>0.1MeV


Irradiation temp.: 1050-1075°C Irradiation time: 351 efpd
Heating temperature: 1800°C Heating time: 279 h

See also chapter 4.3.2.1.

Low in-pile R/B values of fission gases indicated no particle failure during irradiation. The
compact (small sphere) was heated within 12 h from 1250 to 1800°C and then remained at
1800°C over 279 h [Schenk 1997]. Krypton release began to increase continuously from the
beginning and reached a value of 1.0×10-3 at the test end. Metallic fission product release
during heating from this compact was highest among all heated HFR-P4 compacts. It was the
only test where silver still could be detected, whereas this was not possible in all other heated
compacts, where most silver-110m was decayed four years after irradiation.

Cesium

The measured fractional release of cesium was steadily increasing and reached a final value as
high as 52%. After deconsolidation of the compact, the cesium inventory in the matrix
graphite was measured to be 5.2%.

Figure 6-47. FRESCO-II postcalculation of 137Cs fractional release for HFR-P4/3-12.

Cesium in SiC: 1987 reference × 0.13


Cesium in Matrix: 1987 reference

196
This compact was postcalculated with a variation of the diffusion coefficient in SiC only. Its
modification by a factor of 0.13 leads to a good agreement with the release level reached at
the end of the test, but to a poor agreement with the transient release behavior.

Strontium

Strontium release from the compact for this test could not be measured with sufficient
accuracy. After deconsolidation of the compact, the strontium inventory in the matrix graphite
was measured to be 2.3%.

Silver

Silver was below the detection limit and could not be measured.

197
SL-P1/C6

Characteristic data of experiment SL-P1/C6

Burnup: 10.7% FIMA Fast neutron fluence: 6.7×1025, E>0.1MeV


Irradiation temp.: 790°C Irradiation time: 330 efpd
Heating temperature: 1600°C Heating time: 304 h

Low in-pile R/B values of fission gases indicated no particle failure during irradiation. The
compact (small sphere) was heated within 7.5 h from 1250 to 1600°C and then remained at
1600°C over 304 h [Schenk 1997].

Cesium

The measured fractional release of cesium was increasing and then leveled out near the
inventory of one particle reaching a final value of 3.9×10-4. After deconsolidation of the
compact, the cesium inventory in the matrix graphite was measured to be 9.7×10-4.

Figure 6-48a. FRESCO-II postcalculation of 137Cs fractional release for SL-P1/C6.

Cesium in SiC: 8.91×10-17 m2/s


Cesium in Matrix: 1987 reference × 0.04

198
Strontium

The measured fractional release of strontium, qualitatively similar to the cesium release, was
increasing to a final value of 4.3×10-5. After deconsolidation of the compact, the strontium
inventory in the matrix graphite was measured to be 1.2×10-3.

Figure 6-48b. FRESCO-II postcalculation of 90Sr fractional release for SL-P1/C6.

Strontium in SiC: 1987 reference × 0.05


Strontium in Matrix: 1987 reference × 0.2
HMC: 1×10-4

Silver

Silver was below the detection limit and could not be measured.

199
SL-P1/C9

Characteristic data of experiment SL-P1/C9

Burnup: 10.7% FIMA Fast neutron fluence: 6.3×1025, E>0.1MeV


Irradiation temp.: 790°C Irradiation time: 330 efpd
Heating temperature: 1700°C Heating time: 304 h

Low in-pile R/B values of fission gases indicated no particle failure during irradiation. The
compact (small sphere) was heated within 9 h from 1250 to 1700°C and then remained at
1700°C over 304 h [Schenk 1997]. After approximately 110 h at 1700°C, krypton release
began to increase and reached a value of 3.7×10-5 at the test end. With regard to the presence
of 1631 coated particles in this compact, the released krypton corresponds to less than 10% of
the inventory of one particle.

Cesium

The measured fractional release of cesium was steadily increasing and reached a final value of
9.8%. After deconsolidation of the compact, the cesium inventory in the matrix graphite was
measured to be 3.6%.

Figure 6-49a. FRESCO-II postcalculation of 137Cs fractional release for SL-P1/C9.

Cesium in SiC: 6.17×10-16 m2/s


Cesium in Matrix: 1987 reference × 0.04

200
Strontium

The measured fractional release of strontium reached a final value of 5.8%. After
deconsolidation of the compact, the strontium inventory in the matrix graphite was measured
to be 4.5%.

Figure 6-49b. FRESCO-II postcalculation of 90Sr fractional release for SL-P1/C9.

Strontium in SiC: 1987 reference × 0.2


Strontium in Matrix: 1987 reference

Silver

Silver was below the detection limit and could not be measured.

201
SL-P1/C10

Characteristic data of experiment SL-P1/C10

Burnup: 10.3% FIMA Fast neutron fluence: 6.0×1025, E>0.1MeV


Irradiation temp.: 790°C Irradiation time: 330 efpd
Heating temperature: 1700°C Heating time: 304 h

Low in-pile R/B values of fission gases indicated no particle failure during irradiation. The
compact (small sphere) was heated within 9 h from 1250 to 1700°C and then remained at
1700°C over 304 h [Schenk 1997]. Krypton release began to increase after about 40 h at
1700°C and reached a value of 9.1×10-5 at the test end corresponding to less than the
inventory of one particle.

Cesium

The measured fractional release of cesium was steadily increasing and reached a final value of
10.4%. After deconsolidation of the compact, the cesium inventory in the matrix graphite was
measured to be 2.3%.

Figure 6-50a. FRESCO-II postcalculation of 137Cs fractional release for SL-P1/C10.

Cesium in SiC: 7.41×10-16 m2/s


Cesium in Matrix: 1987 reference

202
For this compact, a good agreement with the measurements was obtained for the second half
of the release curve while the first part is underpredicted by the calculations. This result is
similar to the previously mentioned compact HFR-P4/3-12.

Strontium

The measured fractional release of strontium was steadily increasing and reached a final value
of 2.1%. After deconsolidation of the compact, the strontium inventory in the matrix graphite
was measured to be 9.0%.

90
Figure 6-50b. FRESCO-II postcalculation of Sr fractional release for SL-P1/C10.

Strontium in SiC: 1987 reference × 0.1


Strontium in Matrix: 1987 reference

Silver

Silver was below the detection limit and could not be measured.

203
6.2.4.3. Cesium and Strontium Concentration Profiles in Heated Fuel Spheres

From the AVR operation experience, it is known that activity concentrations in the coolant are
several orders of magnitude larger for solid fission products than they are for gaseous species.
The activity mostly originated from the heavy metal contamination. As shown in the lower
curve in Fig. 6-51, cesium profiles in fuel elements with “normal” particle performance, i.e.,
no failure, reveal an increase in concentration near the surface due to adsorption of cesium
from the gaseous phase (“cross-contamination”).

1 E+0
Relative Concentration Cs 137

1 E-1
GLE-1, 6.3% FIMA, with
failed particles
1 E-2

1 E-3

1 E-4 GLE-1, 4.3% FIMA, without


failed particles
1 E-5

1 E-6
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Radius Fuel Element [cm]

Figure 6-51. Measured cesium concentration profiles in the outer fuel-free zone comparing
AVR GLE-1 fuel spheres with and without failed particles.

Figure 6-52. Measured cesium concentration profiles in the outer fuel-free zone for an AVR
fuel sphere (variant THTR) and a moderator sphere.

204
The level of contamination in AVR spheres is much higher compared to profiles in an MTR
irradiated fuel element of the same type. However, the reduction in this adsorbed activity
observed over the years reflects the gradual replacement of BISO by high-quality TRISO fuel
in the AVR core. The effect of cross-contamination could be identified even in moderator
spheres as shown in Fig. 6-52.

Figures 6-53 show the normalized profiles in the fuel-free zone of spherical fuel elements. For
the example AVR GK 17/6, it can be seen from Fig. 6-53a that even without particle failure,
strontium release from carbide fuels is very high and. Strontium and Europium move in an
identical manner in (Th,U)C2 fuel, through a BISO coating and along the A3 matrix of the
spherical fuel element. Note that 90Sr and 154Eu profiles are identical when normalized to their
respective inventories. Due to these similarities, the strontium beta measurements for AVR
spheres were dropped and Eu-154 gamma measurements were used instead.

1 E+0

1 E-1
Relative Concentration

Sr 90 AVR GK 17.6% FIMA

1 E-2

1 E-3 Eu 154 GK 17.6% FIMA

1 E-4
Cs 137 GK 17.6% FIMA
1 E-5

1 E-6
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Radius Fuel Element [cm]

Figure 6-53a. Normalized profiles in the fuel free zone of spherical fuel elements for
AVR sphere GK 17/6 after end of irradiation.

The fuel element FRJ2-K13/4 (Fig. 6-53b) had zero manufacturing defects, zero in-pile
failures and no detectable fission product release. The measured profiles are characteristic of
detection limits in hot cell work and represent handling contamination values. In Fig. 6-53c,
the example of an AVR GLE-3 fuel sphere with 8% FIMA is given showing fission product
deposition profiles from the outside. Release from the fuel zone cannot be detected, because
normalized profiles for 134Cs/137Cs and 90Sr/154Eu would then be identical.

205
1 E+0

1 E-1
Relative Concentration Sr 90 FRJ2-K13/4

1 E-2

1 E-3 Eu 154 FRJ2-K13/4

1 E-4
Cs 137 FRJ2-K13/4
1 E-5

1 E-6
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Radius Fuel Element [cm]

Figure 6-53b. Normalized profiles in the fuel free zone


of spherical fuel element FRJ2-K13/4 after end of irradiation.

1 E+0

1 E-1 Sr-90
Relative Concentraion

1 E-2
Cs-134
1 E-3
Cs-137
1 E-4

1 E-5 Eu-154

1 E-6
0 0.5 1 1.5 2 2.5 3
Radius Fuel Element [cm]

Figure 6-53c. Normalized fission product profiles in the fuel free zone
of an 8% FIMA AVR GLE-3 element.

206
Diffusion calculations which were optimized with respect to fuel element and, whenever
possible, coated particle release should principally allow the reproduction of the graphite
inventory at the end of the heating test. Calculated cesium profiles in the fuel element were
indeed found to be in very good agreement with the measured profiles.

The assumption that the classical Fickian diffusion of cesium in matrix graphite be
sufficiently correct and conservative, could be demonstrated by the postcalculation of heating
tests with fuel spheres, for which a post-heating deconsolidation was made (Figs. 6-54, 6-55,
6-56). A similarly good agreement was achieved for some heated spheres with respect to
strontium (Figs. 6-57, 6-58).

Figure 6-54 shows the calculated concentration profile of 137Cs in the matrix over the sphere
radius in comparison with respective measurements for the example of an AVR-irradiated
sphere heated at 1600°C and two MTR-irradiated spheres heated at 1600 and 1800°C,
respectively. Curves and measurements are describing the state at the end of the heating tests.
For sphere AVR 71/22 (lower curve) with moderate irradiation conditions, low concentrations
were measured indicating no significant release from the coated particles during the 1600°C
heating. In contrast, in the two HFR-K3 spheres with much more severe irradiation conditions
(higher burnup and neutron fluence), the particles are at the end of the heating test in a stage
of highly releasing cesium. Respective profiles in the matrix graphite calculated with the
FRESCO code (where the cesium release from the coated particles into the matrix was
adjusted to the measurements which allows to isolate the diffusion transport in the matrix)
show a good agreement with the measurements. It means that the “effective” diffusion
approach is sufficiently well describing cesium transport behavior in matrix material.

Figure 6-54. Comparison of cesium profile measurements for heated fuel spheres with
FRESCO postcalculations (with adjusted release from coated particles).

207
Figure 6-55. Comparison of cesium profile measurements for heated fuel spheres with
FRESCO postcalculations (with adjusted release from coated particles).

Figure 6-56. Comparison of cesium profile measurements for heated fuel spheres with
FRESCO postcalculations (with adjusted release from coated particles).

208
Figure 6-57. Comparison of strontium profile measurements for heated fuel spheres with
FRESCO postcalculations (with adjusted release from coated particles).

Figure 6-58. Comparison of strontium profile measurements for heated fuel spheres with
FRESCO postcalculations (with adjusted release from coated particles).

209
Since from the deconsolidation process [Schenk 1988], activity inventory information became
available over the diameter of the sphere, typically two data points per radial position
examined were generated. The example AVR 74/11 is shown in Fig. 6-55 with four data
points per radial position including both FZJ and Harwell measurements.

6.2.4.4. Derivation of New Diffusion Coefficients in SiC and Matrix Graphite

Silicon Carbide

Application of the Myers diffusion coefficient (curve 1 in Fig. 6-59) to the heating tests with
modern German HTGR spherical fuel elements at FZJ have shown a strong and steady
tendency to overestimate the cesium release as can be seen from many of the above described
postcalculations. A total of 44 heating tests was evaluated in terms of cesium release. It was
considered a sufficiently solid data base for recommending a new high-temperature branch
(curve 2 in Fig. 6-59), which is the fit through all symbols, as the new reference diffusion
coefficient [Verfondern 1989]. Table 6-5 summarizes for each calculated heating test the
adjusted diffusion coefficient of cesium in SiC which resulted in good agreement with the
measurements of the heating tests.

Table 6-5. Adjustment constants for diffusion coefficient of cesium in SiC.


Experiment Heating Adjusted
temperature T diffusion coefficient
[°C] [m2/s]
AVR 70/26 1600 2.8×10-17
AVR 71/22 1600 1.9×10-17
AVR 82/9 1600 4.7×10-17
AVR 82/20 1600 1.9×10-16
AVR 88/15 1600 2.8×10-17
AVR 88/33 1600 2.8×10-17
FRJ2-K13/2 1600 2.8×10-17
FRJ2-K13/4 1600 2.8×10-17
HFR-K3/1 1600 3.8×10-17
HFR-P4/1-8 1600 1.2×10-16
HFR-P4/1-12 1600 7.4×10-17
HFR-P4/2-8 1600 1.9×10-16
HFR-P4/3-7 1600 1.1×10-16
SL-P1/6 1600 8.9×10-17
R2-K13/1 1600 2.8×10-17
AVR 74/11 1700 8.5×10-17
SL-P1/9 1700 6.2×10-16

210
SL-P1/10 1700 7.4×10-16
AVR 69/13 1800 1.2×10-17
AVR 70/33 1800 3.5×10-16
AVR 74/10 1800 1.7×10-15
AVR 76/18 1800 1.3×10-15
AVR 88/15 1800 1.4×10-15
AVR 88/33 1800 1.7×10-15
AVR 88/41 1800 1.7×10-15
FRJ2-K13/4 1800 1.7×10-15
HFR-K3/3 1800 2.6×10-15
HFR-P4/3-12 1800 3.7×10-15
AVR 73/12 1900 1.6×10-15
AVR 74/20 1900 1.9×10-14
AVR 76/19 1900 3.1×10-14
AVR 80/22 1900 3.7×10-15
AVR 71/7 2000 5. 9×10-14
AVR 80/16 2000 5.0×10-13
AVR 74/6 2100 9.8×10-13
AVR 74/24 2100 1.2×10-12
AVR 76/27 2100 1.7×10-12
AVR 76/28 2100 1.5×10-12
AVR 69/28 2150 3.7×10-14
AVR 70/18 2400 1.8×10-12
AVR 70/19 2400 4.6×10-14
AVR 74/8 2500 1.6×10-13
AVR 74/17 2500 7.9×10-13
AVR 80/14 2500 7.9×10-13

Curve fitting procedure:

Y = -1.80 - 2.68 X, r2 = 0.84

Do = 1.59×10-2, Q = 514 kJ/mol

211
Figure 6-59. Derivation of a new diffusion coefficient for cesium in SiC from heating tests
with modern HTGR spherical fuel elements.

An analogous evaluation has been made for strontium on the basis of 11 KÜFA heating tests
with modem HTGR fuel in the temperature range between 1600 and 1800°C. The new
diffusion coefficient [Verfondern 1991] shown as curve 2 in Fig. 6-60 has a higher activation
energy resulting in a diffusion coefficient lower by a factor of about 20 at 1600°C compared
to the old recommendation. Unfortunately, the combination of this new high-temperature
Arrhenius relation with the old HBK relation for normal operation temperatures into a single
two-branch diffusion coefficient significantly underestimates the experimental retention
quality of SiC for strontium at 1600°C, which is obviously the reason for the overestimation
of experimental strontium release data in ostcalculations. In order to benefit from this
potential for reduced release, more strontium release data at lower temperatures should be
made available. All reduction factors derived for strontium are listed in Table 6-6, the
respective new high-temperature branch of the diffusion coefficient in Fig. 6-60.

212
Table 6-6. Adjusted diffusion coefficients of strontium in SiC.
Experiment Heating Adjusted
temperature diffusion coefficient
[°C] [m2/s]
AVR 70/26 1600 1.6×10-16
AVR 71/22 1600 1.2×10-16
FRJ2-K13/2 1600 1.4×10-16
FRJ2-K13/4 1600 4.6×10-17
HFR-K3/1 1600 2.3×10-16
R2-K13/1 1600 4.6×10-17
AVR 74/11 1700 3.6×10-15
AVR 76/18 1800 8.2×10-15
FRJ2-K13/4 1800 2.5×10-15
HFR-K3/3 1800 9.8×10-15

Curve fitting procedure:

Y = 6.25 - 4.13 X, r2 = 0.91

Do = 1.78×106, Q = 791 kJ/mol

As can be seen from Fig. 5-2, in the case of strontium, the discovered reduction potential of
the diffusion coefficient at 1600°C does not really become effective, when the overall
diffusion coefficient is used, i.e., the sum of the Arrhenius relations for normal operation and
for accident conditions.

Silver

Silver transport data have been taken from the HBK data set. No evaluation of available silver
release data from KÜFA heating tests has been made so far because no consistent release
behavior was observed and is in many cases not reproducible by any existing model. A
possible explanation as proposed by Myers could be the silver to be trapped at neutron-
induced defects in the silicon carbide structure at nominal operation temperatures resulting in
a burst release of the trapped silver at elevated temperatures.

213
Figure 6-60. Derivation of a new diffusion coefficient for strontium in SiC from heating tests
with modern HTGR spherical fuel elements.

Matrix Graphite

Comparative measurements with A3-27 type material at 1000°C showed that the diffusion
coefficient for cesium was lower by a factor of 20 in relation to A3-3 material [Hoinkis 1983].
In addition, FRESCO postcalculations of heating tests indicated that for A3-27 materials a
reduction of the A3-3 diffusion data for cesium could be reduced by at least a factor of 10 in
the accident temperature range 1600-1800°C (Fig. 6-61, left). All presented fuel elements
except R2-K13/ L are made of A3-27 matrix graphite.

A respective study for strontium (Fig. 6-61, right) has also resulted in lower diffusion
coefficients compared to the reference data. A new recommendation, however, was here not
given, since the reference data appear to represent a conservative coverage of the
experimental data [Verfondern 1991].

214
The input data for heavy metal contamination fraction and for the diffusion coefficients in SiC
and matrix graphite are usually varied until the calculated release data fit the measurements.
The adjustment of fuel element release to a lower level can - if the release from the coated
particles is correctly reproduced - be achieved by reducing either the diffusion coefficient in
graphite or the contamination fraction. The latter is not explicitly known for every single fuel
element, only average values can be estimated.

Table 6-7. Adjustment constants for diffusion coefficient of cesium and strontium
in matrix graphite.
Experiment Heating Adjustment factor to reference
temperature diffusion coefficient (Hoinkis)
[°C] Cesium Strontium
AVR 70/26 1600 0.4
AVR 71/22 1600 0.1 0.08
AVR 82/20 1600 0.3 0.1
AVR 88/33 1600 0.5 -
FRJ2-K13/2 1600 0.1 0.01
FRJ2-K13/4 1600 0.01 0.01
HFR-K3/1 1600 0.03 0.025
R2-K13/1 1600 0.8 0.2
AVR 74/11 1700 0.02 0.03
AVR 76/18 1800 0.02
AVR 88/33 1800 0.5
FRJ2-K13/4 1800 0.01 0.09
HFR-K3/3 1800 0.1 0.1
AVR 80/22 1900 0.8
AVR 76/27 2100 0.8
AVR 69/28 2150 0.8
AVR 80/14 2500 0.8

215
Figure 6-61. Derivation of a new diffusion coefficient for cesium (left) and strontium (right)
in matrix graphite from heating tests with modern HTGR spherical fuel elements.

The uncertainty for the recommended transport data of metallic radionuclides in graphitic
materials is considered to be fairly high. This is essentially due to strong dependencies on
graphite type and nature, structure, temperature, fission product concentration, state of
oxidation, irradiation damage, coolant pressure, or interference with other fission product
species. The range of uncertainty was estimated in [IAEA 1997] to be at least one order of
magnitude, for silver even up to three orders of magnitude; and no further diffusion data
measurements are known of to have been made since. For safety analysis purposes, however,
the simple Fickian diffusion model can be applied to graphite to obtain concervative release
data.

216
6.2.5. FRESCO-II Results for Fission Gas and Iodine Release from UO2 Kernels and
TRISO Coated Particles

Nine heating tests using fuel specimens with UO2 kernels and with UO2 TRISO coated
particles were conducted with the objective to measure fission product release with a focus on
fission gas and iodine release. The tests also provided valuable knowledge on the release of
metallic fission products from UO2 kernels. Each coupon of FRJ2-P28 contained 5 UO2
kernels surrounded by a porous buffer layer. The investigated coupon and compact of FRJ2-
P27 and also sphere 3 of the FRJ2-K14 test contained TRISO coated particles with UO2
kernels. Sphere 3 of FRJ2-K14 contained mixed oxide TRISO coated particles. All heating
tests were conducted in the KÜFA furnace.

6.2.5.1. Heating Tests

The heating tests conducted are listed in Table 6-8; some characteristic data of the fuel are
given in Table 6-9.

Table 6-8. FZJ heating tests with focus of fission gas and iodine release.
Test Irradiation Burnup FZJ Heating
facility [% FIMA] heating facility temperature [°C]
FRJ2-K14/2 Dido, Jülich 5.2 KÜFA, Jülich 1400, 1600
FRJ2-K14/3 Dido, Jülich 4.3 KÜFA, Jülich 1400, 1600
FRJ2-P27/C15 Dido, Jülich 7.6 KÜFA, Jülich 1100, 1400
FRJ2-P27/C8-2 Dido, Jülich 9.0 KÜFA, Jülich 960, 1400, 960
FRJ2-P28/C1 Dido, Jülich 8.7 KÜFA, Jülich 800, 1400, 800
FRJ2-P28/C2 Dido, Jülich 8.6 KÜFA, Jülich 800, 1800
FRJ2-P28/C6 Dido, Jülich 8.7 KÜFA, Jülich 1000, 1600
FRJ2-P28/C7 Dido, Jülich 0.4 KÜFA, Jülich 900, 1700
FRJ2-P28/C8 Dido, Jülich 0.4 KÜFA, Jülich 1400

Table 6-9. FZJ heating tests with UO2 TRISO coated fuel particles.
FRJ2-K14 FRJ2-P27 FRJ2-P28
Specimen Sphere 2 Sphere 3 Compact 15 Coupon Coupon
Fuel HEU LEU UO2 LEU UO2 LEU UO2 LEU UO2
(Th,U)O2 1 defect kernels +
buffer
Particle batch HT150-160, EUO2308 EUO 2308 EUO 2308
162-167 EUO 2309 EUO 2309
No. particles 10,700 1850 2424 34 5
per specimen

217
FRJ2-K14/2

Characteristic data of experiment FRJ2-K14/2

Burnup: 5.2% FIMA Fast neutron fluence:


Irradiation temp.: 1000 - 1200 Irradiation time: 139 efpd
Heating temperature: 500-1400, 1400, Heating time: 30, 20, 20 h
1600°C

The heating test with sphere FRJ2-K14/2 was conducted in the KÜFA furnace with a 30-h
heating ramp to 1400°C and remain at that level for 20 h, followed by a 4 h temperature ramp
to 1600°C with an isothermal heating phase over another 22.5 h before cooling down to room
temperature within 20 h.

Release data for the investigated fission product nuclides measured were:

Kr-85 Xe-133 I-131 Cs-137 Ag-110m


-4 -5 -5 -5
when reaching 1400°C 2.0×10 4.9×10 4.4×10 1.5×10 2.0×10-2
end of 1400°C phase 2.2×10-4 5.5×10-5 7.8×10-5 2.8×10-5 6.4×10-2
end of 1600°C phase 8.6×10-4 3.0×10-4 7.5×10-4 7.9×10-5 2.3×10-1

No calculation

218
FRJ2-K14/3

Characteristic data of experiment FRJ2-K14/3

Burnup: 4.3% FIMA Fast neutron fluence:


Irradiation temp.: ~300 Irradiation time: 139 efpd
Heating temperature: 500-1400, 1400, Heating time: 30, 20, 43 h
1600°C

The heating test with sphere FRJ2-K14/3 was conducted in the KÜFA furnace with a 30-h
heating ramp to 1400°C and remain at that level for 20 h, followed by a 4-h temperature ramp
to 1600°C plus an isothermal heating phase over 43 h.

The measured 133Xe fractional release remained low with 4.6×10-9 when reaching 1400°C,
increased to 2.2×10-8 at the end of the 1400°C phase, and 2.0×10-7 at the end of the 1600°C
phase. Neither iodine nor silver could be measured. Cesium release data were discarded due
to overlap with contamination.

No calculation

219
FRJ2-P27/15

Characteristic data of experiment FRJ2-P27/15

Burnup: 7.6% FIMA Fast neutron fluence: Very low


Irradiation temp.: 1080 - 1130°C Irradiation time: 232 efpd
Heating temperature: 1100, 1400°C Heating time: 18.5, 24 h

The heating test with compact 15 of FRJ2-P27 was conducted in the KÜFA furnace first at
1100°C to remain at that level for 18.5 h, followed by a 30-h heating ramp to 1400°C to
remain there for 24 h.

No calculation

220
FRJ2-P27/8-2

Characteristic data of experiment FRJ2-P27/8-2

Burnup: 7.6% FIMA Fast neutron fluence: Very low


Irradiation temp.: 1080 - 1130°C Irradiation time: 232 efpd
Heating temperature: 960, 960-1400, Heating time: 19, 30, 10, 30 h
1400, 1400-960°C

The heating test with coupon 8/2 of FRJ2-P27 containing 34 UO2 TRISO particles was
conducted in the KÜFA furnace first at 960°C to remain at that level for 19 h, followed by a
30-h heating ramp to 1400°C to remain there for 10 h, and then followed by 30-h cooling
ramp down again to 960°C.

After the 19-h holding time at 930°C, the fractional release data measured for 133Xe and 131I
were 1.3×10-4 and 1.8×10-4, respectively. They increased after heatup to 1400°C to 4.5×10-4
and 3.9×10-4, respectively. After the 1400°C isothermal phase and cooldown to 930°C plus
2-h holding time, the measurements were 1.0×10-3 for 133Xe and 1.2×10-3 for 131I.

No calculation

221
FRJ2-P28/C1

Characteristic data of experiment FRJ2-P28/C1

Burnup: 8.1 + 0.6% FIMA Fast neutron fluence: Very low


Irradiation temp.: 940°C Irradiation time: 251 + 25 efpd
Heating temperature: 800, 800-1400, Heating time: 12.5, 30, 10, 30, 10 h
1400, 1400-800,
800°C

The heating test with 5 UO2 kernels in coupon C1 of FRJ2-P28 was conducted in the KÜFA
furnace first at 800°C to remain at that level for 12.5 h, followed by a 30-h heating ramp to
1400°C to remain there for 10 h, and then followed by 30-h cooling ramp down again to
800°C to remain there for another 10 h.

No calculation

222
FRJ2-P28/C2

Characteristic data of experiment FRJ2-P28/C2

Burnup: 8.2 + 0.4% FIMA Fast neutron fluence: Very low


Irradiation temp.: 940°C Irradiation time: 251 +25 efpd
Heating temperature: 800, 800-1800°C, Heating time: 17, 30, 20 h,
300-1800, 1800°C 18, 50 h

The heating experiment with 5 UO2 kernels in coupon C2 of FRJ2-P28 was conducted in the
KÜFA furnace in two separate tests. The first test included a heating phase at 800°C to
remain at that level for 17 h, followed by a 30-h heating ramp to 1800°C to remain there for
20 h. The second test consisted of an 18-h heating ramp to 1800°C to remain there for another
50 h.

Figure 6-62. FRESCO-II postcalculation of 137Cs fractional release for FRJ2-P28/C2.

Reference diffusion coefficients:


Cesium in SiC Myers
Cesium in Matrix
Modification diffusion coefficients:
Cesium in SiC Reference * 1.
Cesium in Matrix

223
FRJ2-P28/C6

Characteristic data of experiment FRJ2-P28/C6

Burnup: 8.1 + 0.6% FIMA Fast neutron fluence: Very low


Irradiation temp.: 1020°C Irradiation time: 251 + 25 efpd
Heating temperature: 1000, 1000-1600, Heating time: 20, 30, 200 h
1600°C

The heating test with 5 UO2 kernels in coupon C6 of FRJ2-P28 was conducted in the KÜFA
furnace first at 1000°C to remain at that level for 20 h, followed by a 30-h heating ramp to
1600°C to remain there for 10 h, and then followed by 30-h cooling ramp down again to
800°C to remain there for 200 h.

Figure 6-63. FRESCO-II postcalculation of 137Cs fractional release for FRJ2-P28/C6.

Reference diffusion coefficients:


Cesium in SiC Myers
Cesium in Matrix
Modification diffusion coefficients:
Cesium in SiC Reference * 1.
Cesium in Matrix

224
FRJ2-P28/C7

Characteristic data of experiment FRJ2-P28/C7

Burnup: 0.4% FIMA Fast neutron fluence: Very low


Irradiation temp.: Irradiation time: 23 efpd
Heating temperature: 1600°C Heating time: 500 h

The heating test with 5 UO2 kernels of coupon C7 of FRJ2-P28, which was only activated,
was conducted in the KÜFA furnace first at 900°C to remain at that level for 17 h, followed
by a 30-h heating ramp to 1700°C to remain there for 22.5 h.

No calculation

225
FRJ2-P28/C8

Characteristic data of experiment FRJ2-P28/C8

Burnup: 0.4% FIMA Fast neutron fluence: Very low


Irradiation temp.: Irradiation time: 23 efpd
Heating temperature: 1000-1400°C Heating time: 500 h

The heating test with 5 UO2 kernels of coupon C8 of FRJ2-P28, which was only activated,
was conducted in the KÜFA furnace in a 30-h heating ramp from 1000°C to 1400°C and
remained at that level for 10 h, before cooled down in a 30-h ramp back to 1000°C.

No calculation

226
6.2.5.2. Derivation of New Diffusion Coefficients in UO2

Release data from heating and irradiation of FRJ2-P28 were taken to compare with reference
calculations. Cesium (Fig. 6-64) and iodine (Fig. 6-65) release under accident conditions were
found to be best approximated by a two-branch diffusion coefficient while those for strontium
and silver were modified by a factor of 20 and 10, respectively, compared to the
recommendation of the last HBK data set.

Figure 6-64. Derivation of a new diffusion coefficient cesium in UO2 from heating tests with
modern HTGR spherical fuel elements.

The new recommendation for cesium in UO2 written as a reduced diffusion coefficient is:

 in UO2
DCs  0.90e  209, 000 RT  8320e 362, 000 RT s 
1

227
Figure 6-65. Derivation of a new diffusion coefficient iodine in UO2 from heating tests with
modern HTGR spherical fuel elements.

228
6.2.6. Fission Product Release from Contamination

6.2.6.1. Objective

The level of heavy metal contamination and the transport of iodine and fission gases in
graphite in the temperature range of 1000-1800°C were subject to two heating test series,
FRJ2-KA1 and FRJ2-KA2, at the Research Center Jülich [Schenk 1989b, Schenk 1991b],
where spherical fuel elements were activated to low burnup and then heated to investigate the
release behavior of short-lived isotopes originating from the heavy metal contamination. In
each test, three AVR spherical fuel elements of type GLE-4 (AVR 21-2) with UO2 TRISO
coated fuel particles were short-term irradiated at a temperature of about 420°C to 0.6 and 2%
FIMA, respectively, to produce short-lived isotopes. Shortly afterwards the spheres were
moved to the KÜFA furnace and heated at 1000, 1200, or 1400°C over 50 h followed by a
1600 or 1800°C over another 50 h, or even at 2000°C for 2 hours in order to investigate the
release behavior of the short-lived isotopes from the heavy metal contamination. Heating in
the high temperature range was expected to release all contamination-related fission gases.

Table 6-10. FRJ2-KA heating tests with UO2 TRISO coated fuel particles.
Test Irradiation Burnup FZJ Heating
facility [% FIMA] heating facility temperature [°C]
FRJ2-KA1/1 Dido, Jülich 0.73 KÜFA, Jülich 1000, 1600
FRJ2-KA1/2 Dido, Jülich 0.85 KÜFA, Jülich 1200, 1600, 1800,
2000
FRJ2-KA1/3 Dido, Jülich 0.79 KÜFA, Jülich 1400, 1600
FRJ2-KA2/1 Dido, Jülich 1.84 KÜFA, Jülich 1000, 1600
FRJ2-KA2/2 Dido, Jülich 2.00 KÜFA, Jülich 1400, 1600
FRJ2-KA2/3 Dido, Jülich 2.02 KÜFA, Jülich 1200, 1600, 1800

Table 6-11. Characteristic data of the UO2 TRISO particles and fuel spheres used in the
heating tests.
FRJ2-KA1 /KA2
Fuel element GLE-4/2
Fuel LEU UO2
Particle batch HT 385-393, 395-404, 406-423
No. particles per sphere 9500
Matrix grade A3-3

229
6.2.6.2. Tests

FRJ2-KA1/1

Characteristic data of experiment FRJ2-KA1/1

Burnup: 0.73% FIMA Fast neutron fluence: Very low


Irradiation temp.: ~420(s)/~550(c) °C Irradiation time: 22 efpd
Heating temperature: 1000, 1600°C Heating time: 50, 50 h

The heating test FRJ2-KA1/1 was conducted in the KÜFA furnace first at 1000°C over 50 h,
before heated up within 13 h to 1600°C to remain at this level for another 50 h. The fractional
release of 133Xe at 1600°C is by a factor of about 10 higher compared to the level at 1000°C.

Figure 6-67. FRESCO-II postcalculation of fission gas/iodine fractional release


for FRJ2-KA1/1.

230
FRJ2-KA1/2

Characteristic data of experiment FRJ2-KA1/2

Burnup: 0.85% FIMA Fast neutron fluence: Very low


Irradiation temp.: ~420(s)/~550(c) °C Irradiation time: 22 efpd
Heating temperature: 1200, 1600, 1800, Heating time: 50, 36.5, 2.5, 0.5 h
2000°C

The heating test FRJ2-KA1/2 was conducted in the KÜFA furnace in four steps with
increasing temperatures, first at 1200°C over 50 h, then at 1600°C over 36.5 h, then at 1800°C
for 2.5 h, and finally at 2000°C for 0.5 h. The fractional release of 133Xe at 1600°C is by a
factor of about 6 higher compared to the level at 1200°C.

Figure 6-68. FRESCO-II postcalculation of fission gas/iodine fractional release


for FRJ2-KA1/2.

231
FRJ2-KA1/3

Characteristic data of experiment FRJ2-KA1/3

Burnup: 0.79% FIMA Fast neutron fluence: Very low


Irradiation temp.: ~420(s)/~550(c) °C Irradiation time: 22 efpd
Heating temperature: 1400, 1600°C Heating time: 51, 50 h

The heating test FRJ2-KA1/3 was conducted in the KÜFA furnace first at 1400°C over 51 h,
before heated up to 1600°C to remain at this level for another 50 h. The fractional release of
133Xe at 1600°C is by a factor of less than 2 higher compared to the level at 1400°C.

Figure 6-69. FRESCO-II postcalculation of fission gas/iodine fractional release


for FRJ2-KA1/3.

232
FRJ2-KA2/1

Characteristic data of experiment FRJ2-KA2/1

Burnup: 1.84% FIMA Fast neutron fluence: Very low


Irradiation temp.: 420(s)-550°C Irradiation time: 68 efpd
Heating temperature: 1400, 1600°C Heating time: 51, 50 h

The heating test FRJ2-KA2/1 was conducted in the KÜFA furnace first at 1000°C over 50 h,
before heated up within 13 h to 1600°C to remain at this level for another 50 h.

Figure 6-71. FRESCO-II postcalculation of fission gas/iodine fractional release


for FRJ2-KA2/1.

233
FRJ2-KA2/2

Characteristic data of experiment FRJ2-KA2/2

Burnup: 2.00% FIMA Fast neutron fluence: Very low


Irradiation temp.: 420(s)-550°C Irradiation time: 68 efpd
Heating temperature: 1200, 1600°C Heating time: 50, 50 h

The heating test FRJ2-KA2/2 was conducted in the KÜFA furnace first at 1200°C over 50 h,
before heated up within 8.5 h to 1600°C to remain at this level for another 50 h.

Figure 6-72. FRESCO-II postcalculation of fission gas/iodine fractional release


for FRJ2-KA2/2.

234
FRJ2-KA2/3

Characteristic data of experiment FRJ2-KA2/3

Burnup: 2.02% FIMA Fast neutron fluence: Very low


Irradiation temp.: 420(s)-550°C Irradiation time: 68 efpd
Heating temperature: 1400, 1600, 1800°C Heating time: 50, 50, 50 h

The heating test FRJ2-KA2/3 was conducted in the KÜFA furnace in three steps at different
temperature levels, first at 1400°C over 50 h, then at 1600°C at 50 h, and finally at 1800°C for
another 50 h.

Figure 6-73. FRESCO-II postcalculation of fission gas/iodine fractional release


for FRJ2-KA2/3.

235
6.2.6.3. Results

As expected and in agreement with respective PANAMA calculations, in none of the


FRJ2-KA heating tests did a particle failure occur.

While 133Xe could be measured well in all FRJ2-KA/1 tests, the 131I activities deposited on the
condensing plates were often close to the detection limit and have therefore some uncertainty
range. Despite the given uncertainty in the iodine data, a good agreement was found between
iodine and fission gas release behavior. The similar transient release curves seem to support
the interpretation of a release from traps in the matrix graphite. The fact that a burst-like
release occurred between 1000 and 1600°C, supports the interpretation of a release from traps
within the matrix [Schenk 1989b].

In contrast, the extremely small amounts of iodine could well be measured in FRJ2-KA2 by
reducing the furnace contamination and improving measuring methods. The released iodine
could be traced back to contamination of the matrix with natural uranium of < 0.1 ppm. Also
here, the release behavior of 131I and 133Xe was almost identical. A burst-like release of
iodine between 1600 and 1800°C shows that still at this high temperature level, iodine is
being retained to a certain degree. Overall release fractions were very low with an average
value of 6×10-9 (FRJ2-KA1) and 8×10-9 (FRJ2-KA2). Other fission products such as 137Cs or
85
Kr could not be measured accurately due to low inventories and low releases. The cesium
identified was originating from contamination of the furnace from previous heating tests
rather than from the fuel spheres under investigation [Schenk 1991b].

6.2.6.4. Derivation of Diffusion Coefficient

The effective diffusion coefficients derived from these experimental data using the FRESCO
model [Verfondern 1991] revealed that the reference diffusion coefficient was too small at
lower temperatures and too large at higher temperatures > 1250°C. But from the arrangement
of the fitted data, the activation energy appears to be similar to that for xenon transport in
irradiated AUF graphite which was impregnated with uranium carbide [Cubiciotti 1952]. That
Canadian investigation has shown that relatively high fractions of xenon were retained in the
graphite even at temperatures up to 2000°C.

The evaluation of the FRJ2-KA heating experiments with the FRESCO-II code was
conducted by assuming for the contamination fraction in the spheres a single value of 1×10-7
(Fig. 6-74, symbols in top figure = red symbols in bottom figure, also Table 6-12) rather than the
release value measured at the end of each series (Fig. 6-74, black-and-white symbols in
bottom figure), and adjusting the diffusion coefficient for the matrix graphite to fit/envelop
the experimental data. The calculation results are shown in Fig. 6-74a as circles (red circles in
Fig. 6-74b which is just a section of Fig. 6-74a marking all circles with the respective KA-
test, from which they were derived). Results are up to two orders of magnitude lower than the
earlier results, and the low temperature data fall close to the Müller slow phase diffusion
coefficient [Müller 1976]. However, all things considered, these calculational results did not
seem to represent a good basis for recommending a new diffusivity. The old one remains
sufficiently conservative for safety analyses.

236
(a)

(b)

Figure 6-74. FRESCO-II postcalculation of 137Cs fractional release for FRJ2-KA2/3.

237
Table 6-12. Adjusted diffusion coefficients of iodine in matrix to a
heavy metal contamination level of 1×10-7.
Experiment Heating Adjusted
temperature diffusion coefficient
[°C] [m2/s]
FRJ2-KA1/1 1000 1.1×10-13
FRJ2-KA2/1 1000 3.0×10-13
FRJ2-KA1/2 1200 1.5×10-13
FRJ2-KA2/2 1200 3.0×10-13
FRJ2-KA1/3 1400 1.5×10-12
FRJ2-KA2/3 1400 3.0×10-13
FRJ2-KA1/1 1600 5.0×10-12
FRJ2-KA1/2 1600 5.0×10-12
FRJ2-KA1/3 1600 1.6×10-11
FRJ2-KA2/1 1600 1.6×10-11
FRJ2-KA2/2 1600 5.0×10-12
FRJ2-KA2/3 1600 1.8×10-12
FRJ2-KA2/3 1800 1.6×10-10

238
7. PREDICTIVE CALCULATIONS FOR HTGR CONCEPTS OR EXPERIMENTS

In the course of testing and application of the FRESCO model, numerous studies were
conducted in the past including both postcalculations for V&V purposes and predictions for
either planned experiments or reactor concepts as part of respective safety analyses. This
chapter provides an overview of what experience with the FRESCO model has been achieved
up to now to verify and validate the code. Some examples of studies with predictive
calculations will be presented in the next chapter.

7.1. HTR-100 (1988)

7.1.1. HTR-100 Reactor Design

The HTR-100 was designed as a two-reactor plant with a thermal power of 2 × 258 MW.
Similar to the AVR reactor, the steam generator is arranged above the pebble-bed core. The
helium coolant flows upwards at a rate of 103 kg/s entering the core at a temperature of
255°C and reaching an average outlet temperature of 740°C. The system pressure is 7.0 MPa.

7.1.2. Boundary Conditions

7.1.2.1. Computer Codes

The analysis of fission product release behavior in the HTR-100 was made according to the
methodology and data described in [Moormann 1987]. For normal operation, the FRESCO-II
was used to predict the release of the metallic fission products 137Cs, 90Sr, and 110mAg, while
the release of iodine and fission gases was assessed by applying the Booth model.

Due to the homogeneous mixing of fuel spheres with different age resulting from multiple
core passages, the predicted release was expected to be representative of the core release.

The core version, FRESCO-I was used to calculate fission product release during a postulated
core heatup accident with the THERMIX code to provide the necessary thermohydraulic input
data and the PANAMA code to provide the fractions of accident-induced particle failure. The
core cavern of the HTR-100 with a radius of 3.5 m and an average height of 11.75 m was
represented in a 6 (r) times 22 (z) cylindrical calculation grid comprising pebble-be, top and
bottom reflectors, and neglecting the four graphite noses intruding in the core.

The fuel element design assumed corresponds to the German reference spherical fuel element
with LEU UO2 TRISO coated particles. Most of the radioactivity inventory is initially in the
particle kernels.

The FRESCO diffusion model was applied using effective diffusion coefficients. With regard
to iodine transport in graphite, it was distinguished between a rapid transport through the
graphite pores and slow transport in the graphite grain assumed spherical with a 6 μm
diameter. The iodine release was determined in two separate calculations:

239
 release from the defective/failed coated particles only with rapid transport through the
graphite, ignoring heavy metal contamination in the graphite,

 release from the heavy metal contamination only (assuming 100% intact particles) by
slow transport out of the graphite grain,

with the results of both calculations added up to the overall iodine release during the core
heatup accident.

7.1.2.2. Particle Failure Fraction

For normal operation, it was assumed that the irradiation-induced failure fraction corresponds
to half of the end-of-life value (to represent an average fuel element) reached in five linear
steps after 1000 efpd and added to the constant level of the manufacture-induced defect
fraction.

For the core heatup accident, the PANAMA model integrated into the FRESCO-I code was
used to determine the particle failure fractions for each mesh of the calculation grid depending
on its temperature history. The final values were increased by a factor of 5 and then being
used to construct the 5-step failure functions individually for each mesh. In such a way, the
accident-induced failure fractions should be determined sufficiently conservative.

7.1.2.3. Thermodynamic Data

During normal operation, a fuel element was assumed to pass 5 times the active core, each
time being exposed to a temperature of 1000°C (expected) and 1250°C (design) at the top,
which decreases linearly to 400°C at the core bottom. Temperatures were assumed constant
throughout the pebble and the coated particle, respectively.

Fuel temperatures as well as the radial and axial velocity components of the coolant during
the core heatup accident were taken from the respective THERMIX results file and transferred
into discretized time and space structures of the FRESCO model.

7.1.3. Activity Release during Normal Operation of the HTR-100

The FRESCO-II results for the fractional release of the metallic fission products 137Cs, 90Sr,
and 110mAg as a function of the fuel element life time are plotted in Fig. 7-1 for expected
values and in Fig. 7-2 for design values. In all curves, the five cycles of the fuel element can
be well identified. The figures are completed by additional curves for the transient irradiation
temperature (dash-dotted) and the assumed 5-step particle failure function (dashed).

240
Figure 7-1. Prediction of fractional release of metallic fission products from the HTR-100
pebble-bed during normal operation (based on expected values).

Figure 7-2. Prediction of fractional release of metallic fission products from the HTR-100
pebble-bed during normal operation (based on design values).

241
The slope of the release curves is characterized by radioactive decay, inventory buildup, and
diffusional release. A decrease in the fractional release occurs at lower temperatures, when
diffusional release becomes negligible, while the actual inventory is continuously increasing.
The sequential order of fission product species corresponds to the experimentally observed.
Silver shows the highest fractional release from the fuel element with 3.2×10-5 (expected) and
3.5×10-2 (design). It is followed by cesium and strontium, the latter exhibiting strong holdup
in the UO2 kernels and the matrix material. Final values after a 1000-efpd life time are for
cesium 8.0×10-6 (expected) and 1.5×10-4 (design), and for strontium 1.4×10-7 (expected) and
3.0×10-6 (design), respectively.

Reducing the diffusion coefficient for cesium and strontium by a factor of 10 as


recommended for A3-27 matrix material, fractional release data at end-of-life would have
decreased by a factor of 3. An evaluation applying the core version FRESCO-I has shown that
20-30% of the cesium and strontium would be bound in the bore holes of the top reflector.
These fractions would tremendously increase in a depressurized reactor, when coolant
velocities were small. Consideration of the sorption effect would even result in a complete
binding of Cs and Sr activities due to the vast amount of relatively cold graphite surfaces in
the pebble-bed.

For the further source term assessment, it is assumed that 80% of the metallic fission product
activity and 100% of the iodine activity released from the pebble-bed will be deposited on the
metallic surfaces of the steam generator.

7.1.4. Activity Release during a Core Heatup Accident of the HTR-100

The results of PANAMA used here as input to the FRESCO-I calculations are described in
China Report 1. The transient fractional release of metallic fission products from the coated
particles during a core heatup accident in the depressurized reactor is shown in Fig. 7-3.
Release from the fuel particles begins for all nuclides at the same level corresponding to the
heavy metal contamination increasing significantly in the time interval of 20-40 h when
maximum temperatures are being reached. As expected silver is liberated most reaching a
release fraction of 1.2% of its total core inventory after 200 h. It is followed by strontium with
1.7×10-3 and cesium with 1.5×10-4. The somewhat higher strontium release compared to
cesium is due to a less favorable diffusion coefficient in SiC in the temperature range
1000-1700°C.

With regard to the release from the fuel elements (Fig. 7-4), i.e., the core, the silver fraction
released remains with 1.2% unchanged, since there is no holdup in the matrix graphite. This is
followed – now in the expected order – by cesium with 1.1×10-4 and strontium with 1.2×10-5,
due to the strong retention capability of matrix graphite for strontium. If the sorption effect of
the fuel sphere surfaces were to be taken into account, the core release for both during the
heatup accident would be negligible (< 10-8).

242
Figure 7-3. Prediction of fractional release of metallic fission products from the coated
particles during a core heatup accident in the depressurized HTR-100 reactor.

Figure 7-4. Prediction of fractional release of metallic fission products from the fuel spheres
(no sorption) during a core heatup accident in the depressurized HTR-100 reactor.

243
The fractional release of iodine from the coated particles and the core, respectively, shown in
Fig. 7-5 is directly coupled to the particle failure fraction. About half of the release iodine
originates from heavy metal contamination in the matrix graphite. The other half of the iodine
is released by diffusion from the kernels of defective/failed particles as shown in the lower
curve of the figure. Integral release fraction reached after 200 h is 5.3×10-6 corresponding to
about 10% of the inventory in the defective/failed particles. This result would be the same,
even if the sorption effect would be limited to temperatures < 1350°C and a further
uncertainty factor of 20 on the sorption isotherms were assumed.

Figure 7-4. Prediction of fractional release of iodine from the fuel during a core heatup
accident in the depressurized HTR-100 reactor distinguished between release from the
defective/failed particles and from heavy metal contamination.

7.2. U.S. MHTGR (1988)

This study describes cooperative work on HTGR safety research as agreed to in the "US/FRG
Umbrella Agreement for Cooperation in GCR Development: Safety Research Subprogram
Plan". The objective was to compare calculation models used in safety analyses in the US and
FRG which describe fission product release behavior of TRISO coated fuel particles under
core heatup accident conditions. The study was conducted in 1988 and published in 1991
[Verfondern 1991b].

A quantitative comparison of the US and FRG reference calculation models was made by
assessing a benchmark problem predicting particle failure and radionuclide release under
MHTGR conduction cooldown accident conditions.

244
7.2.1. Definition of Benchmark Problem

The benchmark problem for the fuel particle to which the US and German fuel failure and
release models were applied and the results compared, is based upon a depressurized
conduction cooldown accident (loss of coolant and loss of forced circulation accident in
connection with a core heatup) in the MHTGR. The maximum fuel temperature is limited to
around 1600°C through passive heat removal by conduction and radiation to a reactor cavity
cooling system surrounding the reactor vessel.

The temperature histories used in the benchmark problem for comparing US and FRG models
were based on the predicted peak and average temperatures during a depressurized conduction
cooldown in the MHTGR. They are shown in Fig. 7-6 for the first 1000 h of the transient. The
location for the "maximum temperature" history is where the absolute peak temperature
occurs during the transient with a maximum of 1621°C reached after 81 h and then gradually
decreasing to 1372°C at 250 h and down to 782°C after 1000 h. For the "average temperature"
conditions, a particular location in the core was selected which has a temperature history
similar to the average core temperature. The average temperature history peaks at 1224°C
after 90 h and then gradually decreases to 1053°C at 250 h and down to 636°C after 1000 h.
The time-averaged temperature values in the first 250 hours are 1508°C for the maximurn
temperature history and 1148°C for the average temperature history. A time of 250 hours has
been selected for the comparisons in this study since beyond that time any additional fuel
failure and radionuclide release is insignificant.

Figure 7-6. Maximum and average temperature histories für the benchmark problem.

245
7.2.2. Initial Fuel Particle Conditions

The integrity of the US and FRG fuel at the start of an accident to be used in caleulations for
risk analyses is presented in Table 7-1. By far the majority of the particles are expected to
remain intact and contain all radionuclides.

US models for accident conditions assume that particles with missing buffer layer fail
completely during normal operation producing exposed kernels. For the other possible
defective particles and for the standard particles, pressure vessel failure is insignifieant for the
operating conditions that the MHTGR or the HTR-Modul experience. The US value for
exposed kernels at the start of an accident is 5×10-5. The US also assumes an additional
6.1×10-5 of the particles to have a failed SiC layer.

In FZJ calculations (at that time), a particle fraction of 1.5×10-5 with a defective coating due
to manufacture (= half of the "free" uranium) is conservatively treated as exposed kernels. An
additional particle fraction of 1×10-4 (= expected value, = half of the target specification for
the HTR-Modul) is assumed to fail during normal operation by the end of life [Moormann
1987]. The total failed fraction in FZJ calculations at the beginning of an accident is 6.5×10-5
(= defect fraction in fresh fuel plus half of the end-of-life value as average for the total core)
being homogeneously distributed over the whole active core.

These FZJ assumptions for initial fuel particle conditions were a conservative choice made in
1986 [Moormann 1987]. (According to new experience within the test program of modern
German HTGR fuel, the assumptions for the "free" uranium fraction in fresh fuel were later
changed.

Table 7-1. Comparison of US and German fuel integrity data at the start
of an accident sequence.
Defect/failure USA Germany
Heavy metal contamination 1.0×10-5 1.5×10-5
Exposed kernels 5.0×10-5 6.5×10-5
Missing buffer 0. -
-5
Defective SiC, intact oPyC 6.1×10 -
-5
Defective iPyC, intact SiC 3.0×10 -
-2
Defective oPyC, intact SiC 3.0×10 -

The focus of the benchmark problem is on the fissile fuel particle of the US reference fuel
particle design. The nominal specification of the fuel particle and the initial fuel conditions
which have an influence on fuel particle failure and radionuclide release are presented in
Table 7-2.

For the equilibrium cycle and assuming a uniform loading of the core, the number of particles
at the specific locations is estimated to be 2.3×108 fissile and 1.3×108 fertile particles at the
hot node and 1.7×108 fissile and 1.4×108 fertile particles, respectively, at the location with
average conditions.

246
Table 7-2. Assumed initial fuel conditions for the benchmark case.
Parameter Value
Kernel
Composition UCO
Enrichment [%] 19.8
Heavy metal loading [g] 2.1×10-4
Diameter [μm] 350
Coating layer thicknesses [μm]
Buffer 100
IPyC 50
SiC 35
OPyC 40
Irradiation temperature
for “average temperature history” 692°C
for “maximum temperature history” 612°C
Irradiation time 1100 efpd
Fast neutron fluence 3.6×1025 n/m2
Burnup 15.3% FIMA
Fission Density 5.8×1026 m-3
SiC layer strength after irradiation 480 MPa
Weibull modulus after irradiation 5

7.2.3. Transport Data

The set of diffusion coefficients selected for the German calculations was taken from
[Moormann 1987], which means in particular, the “upper Myers” relation as reference for
cesium in SiC.

7.2.4. Results

The results of both GA model SORS and FZJ model FRESCO-II were compared for the
release of cesium, which illustrates metallic radionuclide release, and iodine, which illustrates
gaseous radionuclide release.

7.2.4.1. Exposed Fuel Kernel Release Model

Cesium Release

The FZJ (KFA/ISF) modeling of the release behavior from defective or failed particles during
the accident transient assumes that the transport is governed by diffusion from the exposed
particle kernel into the matrix graphite directly. The calculation for the maximum temperature
history (Fig. 7-7) indicates that the exposed kernels have almost lost their total cesium

247
inventory within 100 hours into the accident. The cesium release for the average temperature
history (Fig. 7-8) increases more slowly. After 250 hours, the fraction of cesium outside the
kernel is 2.1×10-5 which corresponds to approximately 40% of the initial cesium inventory
within the exposed fuel kernel.

Figure 7-7. Cesium release from exposed particle kernels


for the maximum temperature history.

Figure 7-8. Cesium release from exposed particle kernels


for the average temperature history.

248
The corresponding GA cesium release curve for the maximum temperature history rises more
rapidly during the initial stage of the transient, but then slows down and eventually remains
below the FZJ curve. After 250 hours, the cesium release fraction is 2.5×10-5 which is 50% of
the initial cesium inventory of the exposed kernels and is a factor of 2 lower than the 100%
release predicted with the German model. The release fraction after 250 hours for the average
temperature history is 1.4×10-5 (or about 30%) which is a factor of 1.5 lower than the
FRESCO prediction.

Iodine Release

A similar conclusion as for cesium can be drawn for iodine release from the initially exposed
particle kernels. Aecording to the KFA/ISF diffusion model, iodine transport should be slower
than cesium in the kernel material since the diffusion coefficient is smaller. FRESCO results
for the maximum temperature history (Fig. 7-9) show that the release fraction reaches
4.5×10-5 after about 150 hours and remains practically constant through the remainder of the
250 hours. The release corresponds to 90% of the initial iodine inventory. For the average
temperature history (Fig. 7-10), the iodine release increases slowly reaching 5.4×10-6 (or
11%) after 250 hours.

In the GA calculation for the maximum temperature history, a steep rise in the release is seen
within the first 10 hours, followed by a gradual rise. After 250 hours, a release fraction of
2.8×10-5 is expected, which corresponds to 44% of the iodine inventory being retained in the
kernel. For the average temperature history, the GA calculation indicates that most of the
iodine (98%) is retained within the kernel.

GA release estimates were found much lower compared to FZJ results due to either a different
modeling approach and/or to different temperature interpolation schemes. The KFA/ISF
model is based on bare kernel release data while the GA model is based on data für
constrained failed particles. Experience has shown that the fission product release from
exposed kernels is higher compared to that from constrained failed particles. Further study
would be required in order to resolve this difference.

249
Figure 7-9. Iodine release from exposed particle kernels
for the maximum temperature history.

Figure 7-10. Iodine release from exposed particle kernels


for the average temperature history.

250
7.2.4.2. Diffusive Release from Standard (Intact) Particles

According to GA modeling philosophy (1985 statistical model and its revision in 1988),
standard particles are of non-releasing character and therefore not expected to release any
fission products, except for silver at high temperatures.

In the FZJ modeling, the FRESCO code is applied to both intact (standard) and
defective/failed particles to determine the diffusive release of fission products. Transport data
in the SiC layer have been subject to many studies. Experience has shown already that the
diffusion coefficient of cesium in silicon carbide recommended by Myers for accident
conditions leads to overestimating the experimental results for modern German HTGR fuel
and definitely requires its revision. A better approach which was found to better describe the
actual experimental data is a diffusion coefficient for SiC with a different structure also
derived by Myers. The calculations of the cesium release from standard particles shown in
Fig. 7-11 for maximum temperature conditions were conducted for both approaches. The
release fraction after 250 hours with the more realistic approach is 5.0×10-4 being two orders
of magnitude lower compared to a calculation based on the reference data. For average
temperature conditions, the cesium release from standard particles remains below 10-8 in both
cases.

No iodine is expected to escape from particles with an intact coating layer in both GA and
FZJ modeling.

Figure 7-11. Cesium release from standard fuel particles according to the diffusion model
for the maximum temperature history.

251
7.2.4.3. Comparison of Fission Product Release Models

Regarding the comparison of the prediction of fission product release using respective
reference methods, it should be noted that the philosophy the modeling is based upon is
different among the codes. Figure 7-12 shows the GA results obtained with the 1985
statistical model compared with FZJ results obtained with the diffusion model for cesium
release at maximum temperature condition. The qualitative difference between the two curves
– later and steeper rise of the FZJ curve compared to GA's – is due to the different
assumptions for the modeled thermally activated processes and the corresponding activation
energies (FZJ: diffusion in SiC; GA: SiC corrosion). The cesium release fraction after 250 h is
calculated by FZJ to be 5.0×10-4 which is higher than the GA result of 1.2×10-4. A reason for
it could be given by the fact that the diffusion model takes account of a diffusive transport in
the matrix graphite. On the other hand, the 1985 statistical model has been derived by
correlating the SiC failure directly with the measured cesium release from the fuel element
(neglecting a cesium holdup in the graphite), thus resulting in a smaller effective permeability
of the SiC layer, i.e., a smaller cesium release from the coated particles.

Figure 7-12. Comparison of cesium release for the maximum temperature history
using the GA and KFA/ISF reference models.

252
7.3. Modern German Fuel Spheres Irradiated in FRJ2-K15 (1990)

Three spherical fuel elements have been irradiated at FZJ in the experiment FRJ2-K15 with
the objective to provide a maximum heavy metal burnup of about 16% FIMA at different
irradiation temperatures. For these fuel elements, each placed in a separate capsule, predictive
calculations have been made (actually before the end of the irradiation test) with respect to
fuel performance and fission product release in postirradiatin accident simulation experiments
in the KÜFA test facility.

7.3.1. Boundary Conditions

The spherical fuel elements inserted in the irradiation test were part of the AVR reload
production charge 21 of GLE-4 type consisting of A3-27 matrix graphite and each containing
9500 UO2 LEU TRISO fuel particles of the batch HT354-383. The geometric data of a
particle were: kernel diameter 501 μm, thickness of buffer 94 μm, iPyC 41 μm, SiC 35 μm,
and oPyC 40 μm.

The content of heavy metal contamination in the matrix graphite was assumed to be 1.2×10-5.
A sorption of fission products at the fuel element surface was not taken into account
corresponding to an unrestricted transition of fission products from the surface into the
coolant.

The irradiation boundary conditions are summarized in the following Table 7-3:

Table 7-3. Irradiation Boundary Conditions in FRJ2-K15.


Parameter FRJ2-K15
Assumed in Real test data
calculations
Capsule 1 + 3 Capsule 2 Capsule 1 + 3 Capsule 2
Irradiation temperature 800 - 900 950 - 1050 800 - 9901280 980 - 11501280
[°C]
Irradiation time [d] 550 550 533 533
Fast fluence 0.2 0.2 < 0.2 < 0.2
[1025 m-2, E>0.1 MeV]
Heavy metal burnup 16.0 16.0 14.1 / 14.8 15.3
[% FIMA]

253
The assumed heating program for the irradiated fuel spheres was closely following that of the
experiment FRJ2-K13 (Table 7-4).

Table 7-4. Heating boundary conditions for the FRJ2-KI5 fuel elements.
Temperature T [°C] Duration of heatup [h] Duration of heating at T [h]
1250 - 9
1600 7 300
1800 2 182

The diffusion calculations were based upon both “Reference 1986”, i.e., according to the
Myers recommendation for silicon carbide with a so-called "laminar" structure, and
“Reference 1990”, i.e., the newly derived diffusion coefficient for cesium in SiC [Verfondern
1989]. Furthermore, the diffusion coefficient of cesium in matrix graphite was reduced by a
factor of 10 due to the presence of A3-27 type matrix.

No particle failure during the heating phase was assumed in the FRESCO calculations.

7.3.2. Results for the Fractional Release of Cesium

The influence of the irradiation temperature Tirr on the diffusive release of cesium can be
recognized in different moments the release curves start to rise significantly. For Tirr = 800°C,
the release from the coated particles remains at a low level for the first about ~50 h at 1600°C
before increasing rapidly. At higher irradiation temperatures, this time interval becomes
shorter. At Tirr = 1050°C, there is almost no delay in the increase of the release. The
differences, however, are soon getting relatively small, therefore no distinction between
capsules 1 + 3 and 2, respectively, was furtheron made. The calculations shown in the Figs.
7-13 and 7-14 are related to Tirr = 900°C.

Figure 7-13 shows the comparison between the new reference calculation (solid curves) and
the calculation based on the original methodology report [Moormann 1987] (dashed curves).
The blue curves represent the cesium release from the coated particles, the red curves that
from the fuel sphere, the difference being stored in the matrix graphite. Due to all experience
from the past years, cesium release based on former reference data will be overestimated to a
large extend. Also the retention of cesium in matrix graphite is stronger than assumed before.
With the new recommendations of diffusion coefficients, most of the investigated heating
experiments could be well reproduced and could, at the same time, still conservatively cover
the measurements with the exception of few examples [Verfondern 1989]. Based on these
data, for the planned FRJ2-K15 heating tests, a cesium release fraction of 2×10-3 after 300 h at
1600°C and 1.5×10-1 after additional 182 h at 1800°C are predicted.

Figure 7-14 presents the relative concentration of cesium in the matrix graphite as a function
of the fuel element radius for different times during the heating phase. Concentrations in the
graphite increase with heating time. The only small increase between 100 h and 182 h at
1800°C points out that about the same amount of cesium is released from the fuel element as
is released from the coated particles into the fuel element graphite.

254
Figure 7-13. Predicted cesium release from FRJ2-K15 spheres
during heating at 1600/1800°C.

Figure 7-14. Predicted cesium concentration profiles in the FRJ2-K15 spheres


at different times during heating at 1600/1800°C.

255
In the presented reference calculation of the cesium release, no fraction of failed particles was
assumed. If a step function for the failure fraction were to be taken into account, a significant
influence on the cesium release from the fuel element would have been found only for high
failure fractions as being predicted for the sphere of capsule 2. In case of a low particle failure
fraction, no effect would be recognized.

7.3.3. Final Remark

Unfortunately, the FRJ2-K15 fuel elements were never heated. The irradiated spheres were
planned to be transported to Karlsruhe for heating in the new KÜFA-II furnace, but they were,
by mistake, confused with other three spheres. The FRJ2-K15 spheres having shown an
excellent irradiation behavior up to very high burnups have gone to the waste in the
meantime, since another transport was not possible.

7.4. HTTR (1999)

One of the objective of this study was the prediction of the fuel performance and fission
product release behavior during HTTR normal operation using the German computer codes
with the results to be compared with respective JAEA (then: JAERI) results [Verfondern
2001c]. The methodology applied was that typically used in German safety and risk analyses
of HTGR reactor designs [IAEA 1997].

7.4.1. Description of HTTR Design

The active core of the HTTR produces a thermal power of 30 MW at an average power
density of 2.5 MW/m3. During the normal operation of the HTTR core, the helium coolant
enters the block-type core from top with a nominal temperature of 395°C and flows
downwards through an annular gap between the outside of the fuel rod and the inner wall of
the bore hole of the fuel assembly. The average outlet temperature is 850°C and 950°C,
respectively. The system pressure is 3.9 MPa with a design limit of 4.7 MPa. The total coolant
mass flow is 12.4 kg/s at 850°C average gas outlet temperature and 10.2 kg/s, respectively, at
950°C.

The nominal irradiation time of the 1st-core fuel was set at 660 efpd. According to the
„standard HTTR operation plan“, is was assumed that the so-called „rated operation“
characterized by a mean gas outlet temperature of 850°C be interrupted by a 110 efpd period
between day 220 and day 330 of „high-temperature (HT) test operation“ with a mean outlet
temperature of 950°C.

The active core of the HTTR is composed of 150 hexagonal pin-in-block type fuel assemblies
arranged in 30 columns of five blocks each, plus seven columns to guide control rods. The
dimensions of the active core are 2.9 m of total height and 2.3 m of equivalent diameter. Each
fuel assembly contains 31 (or 33) fuel rods which are inserted in vertical bore holes. The fuel
rod consists of 14 annular fuel compacts in a graphite sleeve. A fuel compact contains
approximately 13,000 LEU UO2 TRISO particles, at a volume packing fraction of 30%. The
degree of 235U enrichment was fixed at 12 different levels ranging between 3.4 and 9.9%

256
depending on the position in the core to achieve a uniform fuel temperature distribution; the
average value is 6%.

The HTTR primary circuit is separated into two parallel circuits outside the reactor pressure
vessel containing the two main heat sinks. These are the Intermediate Heat Exchanger (IHX)
in one line with one gas circulator and the primary Pressurized Water Cooler (PWC) in the
other line with three circulator units. In the mode of simultaneous operation of the IHX and
primary PWC, 10 and 20 MW of thermal power, respectively, are being removed from the
core. The other operational mode possible is the removal of the total heat by the primary PWC
only. From the PWCs, the heat is discharged via air coolers into the atmosphere. In the IHX,
the heat is transferred to a secondary helium circuit, before it is discharged via the same air
coolers into the atmosphere.

7.4.2. Boundary Conditions

7.4.2.1. Operational Conditions

The vertical fuel temperature profiles along a column of the active core, a stack of five fuel
blocks, as predicted for the HTTR operation during the “HT phase” at the inside of the
annular fuel compact, are given in Fig. 7-15. The maximum fuel temperature is 1290°C
reached in layer 3 of the active core; another relative maximum is found in layer 5. The
respective figure during the „rated operation“ is 1150°C, also reached in layer 3.

Figure 7-15. Vertical temperature profiles at various positions during the “HT phase”. The
curve for the compact inner surface is assumed as the fuel temperature in the calculations.

257
In radial direction, the temperatures differ by not more than a few tens of centigrades. A
safety margin derived from a hot spot factor analysis taking account of both systematic and
random uncertainties was considered (dashed line in the figure) resulting in an upper
temperature limit of 1360°C at „rated operation“ and 1492°C at „high-temperature test
operation“, which is still below the maximum design temperature of 1495°C.

The average core fuel burnup at the end of its lifetime is 22 GWd/t (equivalent to 2.4%
FIMA). The maximum value is 31.5 GWd/t (3.5% FIMA) remaining below the design limit
specified at 33 GWd/t (3.6% FIMA). The maximum burnup will be attained in the fuel
assembly layer 2 of the active core. The maximum layer-averaged fast neutron fluence to
which the fuel is exposed is 1.3×1025 n/m2. Maximum fluence values will also be attained in
the fuel assembly layer 2 of the active core.

Table 7-5 lists the core physical data as used in the FRESCO calculations, representing mean
values for a fuel block (‘layer’). For the fuel temperature, two values are given, the lower one
for the operation periods 0-220 and 330-660 efpd with 850°C coolant outlet temperature, the
higher one for the period 220-300 efpd with 950°C coolant outlet temperature. The values in
parentheses behind burnup represent the share of power production, which determines the
fission product distribution; they are used as weighing factors when determining the fractional
release for a whole core column. In addition, the fuel temperature is distinguished between
expected and safety design, the latter includes uncertainties covering systematic errors.

Table 7-5. Physical parameters during normal operation as a function of the


vertical position (layer) in the active HTTR core.
Averaged core physical parameters in fuel block
Layer # Fuel temperatures [°C] Burnup [% FIMA] Fast neutron
0-220 and 330-660 / 220-330 efpd (Fraction of power fluence
production) [1025 n/m2]
Expected Safety design
1 760/790 880 / 910 2.3 (19.0%) 1.1
2 1040 / 1090 1230 / 1300 3.4 (28.1%) 1.3
3 1150 / 1290 1360 / 1500 3.0 (25.0%) 1.0
4 1100 / 1260 1290 / 1440 2.0 (16.5%) 0.6
5 1100 / 1290 1280 / 1460 1.4 (11.4%) 0.4

7.4.2.2. Fuel and Fuel Quality

The reference concept for the TRISO coated fuel particles of the 1st core is based on a 600 μm
diameter UO2 kernel surrounded by a typical TRISO coating with its sequence of layers of
buffer (60 μm), inner pyrocarbon (30 μm), silicon carbide (25 μm), and outer pyrocarbon
(45 μm). In the calculations, a distinction is made for the thickness of the important silicon
carbide layer thickness between the expected value, which is 30 μm, and the specification of
25 μm.

258
Free uranium is considered to exist as heavy metal contamination of the fuel graphite and as
particles with a through-coating failure and those with a defective SiC layer. No additional
failure of particles is expected during the service life. The total fraction at end-of-life (EOL) is
8.5×10-5. Design values in terms of regulatory requirements have been fixed such that besides
the initial total free uranium outside intact coated particles, an additional fraction of 2×10-3
during the service life shall not be exceeded during the full operation period. Furthermore a
safety design limit for the HTTR operation has been set at a total of 1×10-2. The
comparatively high regulatory requirement for the fraction of free uranium outside intact SiC
layers, heavy metal contamination of the graphite plus SiC defects, originates from the limit
fixed for off-site exposure during normal operation. Therefore a minimization of the coated
particles defects is required as well as the avoidance of significant additional failures during
the reactor operation.

From the fuel quality control, the total number of particles with a through-coating failure in
the active core has been estimated to be 2500. Based on a total number of approx. 9×108
particles in the core, this translates into a fraction of about 2.5×10-6. The contamination of the
fuel element matrix graphite has been estimated to be approximately 2.5 g of uranium.
According to the total loading of 900 kg of U in the core, the respective contamination
fraction is about 2.5×10-6.

Since the FRESCO calculation model is only able to treat a spherical geometry of the fuel
element, the annular HTTR fuel compact has been simulated by an equivalent sphere of the
same volume resulting in a diameter of 32 mm. To follow the typical FRESCO input
structure, a fuel-free zone around the sphere has been assumed, however, of minimal
thickness in order to avoid a significant influence on fission product transport and retention,
respectively; such a fuel-free zone is not existing for the „real“ fuel compacts.

A distinction is made between particles with as-fabricated SiC defects and those with
throughcoating failures to allow for a more precise assessment of the fission gas release which
is completely suppressed by a still intact oPyC layer. JAEA has made an assessment of the
performance of particles with as-fabricated SiC layer defects under irradiation conditions. In a
special simulation of the performance of such particles with the PANAMA code here, the
oPyC layer serves as the wall of the pressure vessel. Its position in the HTTR particle is
between 420 and 465 μm particle radius. Respective input data as used by JAEA are 160 MPa
for the layer strength and 4 for the Weibull modulus [Sawa 1999]. A weakening of the oPyC
layer under neutron irradiation is not assumed.

7.4.2.3. Free Uranium and Initial Distribution

Certain fractions as listed in Table 7-6 are assumed to be available as contamination of the
matrix material or the coating layers or in bare kernels, whereas 100% of the uranium minus
the above fractions is contained in the kernels of intact particles at the beginning. The
calculations were carried out for both expected values and safety design values.

No distinction is made in the FRESCO calculations between particles with a through-coating


failure and those with a defective SiC layer only. All defects are simulated as bare kernels
equivalent to through-coating failures. The difference between the these two types of defects
is of major importance particularly for the fission gas release, for which the reference model
applied is anyway the Booth model.

259
The particle failure fraction as input to the FRESCO model defined as a step function was
selected in the „safety design“ calculations to describe a linear increase during the operation
period from the level of initial failures to the end-of-life failure level. It means in this case that
the step length is 132 days (= 1/5 of 660 days). This does not apply to the „expected value“
calculations where the irradiation-induced failure fraction is zero.

Table 7-6. Free uranium fractions assumed in FRESCO calculations.


Fractions of free uranium
Expected value Safety design value
Distribution at beginning-of-life
U contamination in matrix 2.5×10-6
OPyC 1.0×10-6
SiC 1.0×10-6
IPyC 1.0×10-4
Buffer 1.0×10-3
Initial SiC defects plus 8.0×10-5 2.0×10-3
through-coating defects
Distribution at end-of-life
Irradiation-induced failure 0 8.0×10-3
Linear five-steps failure constant: 8.0×10-5 until 66 efpd: 2.0×10-3
function during operation from 66 efpd: 3.6×10-3
from 198 efpd: 5.2×10-3
from 330 efpd: 6.8×10-3
from 462 efpd: 8.4×10-3
from 594 efpd: 1.0×10-2

7.4.2.4. Transport Data

The recommended sets of diffusion coefficients for the radiologically relevant fission product
species in the different reactor materials have been taken from IAEA (1997) and adjusted to
the HTTR fuel design, i.e., the frequency factors of UO2 being based on a 600 μm diameter. It
should be noted that the German diffusion data of iodine (and gases) in UO2 and matrix
graphite are formally valid up to a temperature of 1250°C only, which is the specification
limit for the German reference fuel, but have been extrapolated to higher temperatures for this
study here.

All fission product calculations with the FRESCO code include the recoil effect in the fuel
kernel, i.e. the prompt release of radionuclides due to fission near the surface into the
graphite. This effect is dependent on the fuel geometry and the assumed recoil distance (here:
7.7 μm). However, the calculated release data by considering the recoil effect might be
somewhat overpronounced since it is not quite clear whether and to which extend the recoil
release is already incorporated in the diffusion data as derived from heating experiments.

260
7.4.3. Metallic Fission Product Release during Normal Operation

The results of the calculations with the diffusion code FRESCO for the fractional release of
the radionuclides 137Cs, 90Sr, and 110mAg from the HTTR core under normal operation
conditions are given in the Figs. 7-16 and 7-17, distinguished for the calculations based on
expected values and on safety design values. The average values take already account of the
(vertical) fission product inventory distribution dependent on the different power generation
in the various layers. The respective weighing factors were derived from the burnup
distribution (see Table 7-5).

The release behavior of cesium is at the beginning dominated by the release from the
defective fuel particles. With the start of the “HT phase”, an enhanced diffusive release can be
recognized, soon exceeding the level of the particle failure fraction. This means that cesium
diffusion through intact and – in the design case – only 25 μm thin SiC layers is contributing
significantly to the overall cesium release. The fractional release from the fuel compact into
the coolant at EOL is 5.6×10-4 (expected) and 2.9% (design), respectively.

Looking at the local release curves for the single core layers, the difference in the release
behavior is quite large when comparing, e.g., the curves of layer 1 with those of layer 3. At
the low temperatures, the cesium originates from defective particles only with the release
level, determined by the recoil effect in the bare kernels. The fractional release after 660 efpd
(EOL) corresponds to approximately 8% of the inventory of the defective particles based on
expected values and 23% (because of the higher temperatures), if based on safety design
values. In contrast, the release in layer 3 is soon dominated by the release from intact
particles, reaching final fractional release values of 0.15 and 5.3%, respectively.

The anticipated fractional release of strontium remains extremely small at the lower
temperatures due to its effective retention in the UO2 kernel and the relatively high buffer
effect of the graphite. The release from the particles is only given by recoil release from the
defective particles. Both diffusion coefficients, however, have a high activation energy
resulting in a rapid transport process with increasing temperatures. For the layers 3, 4, and 5
with the higher fuel temperatures, the differences between the calculated fractional release for
expected and for safety design values (which also includes a thinner SiC layer for the latter) is
about two orders of magnitude. The final strontium release fraction even exceeds that of
cesium from the particles by more than one order of magnitude and reaches 7.2×10-3. The
recoil effect has a significant influence on the release behavior of strontium. Although a
purely geometrical effect independent of the fission product species, the release from the
kernels by recoil counteracts its good retention capability for strontium in combination with a
comparatively high diffusive transport through the SiC layer.

A small release fraction also for silver is observed in layer 1 despite the small retention
capability of the SiC layer; the reason is the large activation energy of the silver transport
process through silicon carbide. At higher temperatures, the silver inventories will be rapidly
released from the particles. The average release fraction at EOL is 10% and about 50% for
expected and safety design values, respectively. The maximum local release fractions are 28%
in layer 2 (expected values) and 77% in layer 3 (safety design values). There is no holdup of
the silver by the compact graphite. A more detailed look into the silver diffusion behavior
shows that during the first operation phase, the silver inventory produced in the fuel kernels is
being released to a large extend into the particle coating and stored there. Only a small
fraction in the order of 1-3% of the actual inventory is passing the coating and the compact
matrix and released into the coolant.

261
Figure 7-16. Metallic fission product release from the TRISO particles during HTTR normal
operation, predicted with FRESCO based on expected values.

Figure 7-17. Metallic fission product release from the TRISO particles during HTTR normal
operation, predicted with FRESCO based on safety design values.

262
With beginning of the HT phase, the release from the fuel compact significantly increases
resulting from an enhanced diffusive transport through the SiC layer. As soon as the fuel
temperature is back to its lower level for “rated operation”, the continuously produced silver
is again predominantly stored in the coating, while the release into the coolant is significantly
reduced and its share of the actual total inventory is slightly decreasing.

7.4.4. Comparison between FZJ and JAERI calculations

7.4.4.1. JAEA Calculation Tools and Boundary Conditions

For the simulation of the release of metallic fission products from the particles and fuel
compact, respectively, JAEA uses the FORNAX code [Sawa 1992]. It is based upon the
diffusive transport through the various materials considering the spherical geometry of the
fuel particle and the cylindrical geometry of the fuel rod in a block reactor. The transport
through the gaseous phase in the gap between compact and sleeve is determined by
desorption/adsorption processes at the surfaces. The prompt release from the particle kernels
by recoil is also included. The FORNAX code handles three kinds of particles, intact
particles; bare kernels (through-coating failures); and-at a stage in between-particles with a
degraded SiC layer.

FORNAX usually treats a complete fuel column at a time with the vertical profiles of fuel
temperature, burnup, and neutron fluence as input data. For the purpose of comparison with
the FORNAX calculations of cesium release, FRESCO was separately applied to each of the
five fuel element assemblies of a column. The calculations are based on expected values. The
diffusion coefficient for cesium in SiC used by JAEA exhibits a somewhat higher activation
energy intersecting the FZJ curve between 1100 and 1200 °C. The data for degraded SiC are
shifted by a factor of 100 towards higher values. A major difference is given for the graphite
diffusion data with the JAEA values higher by about three orders of magnitude.

7.4.4.2. Cesium Release Behavior during HTTR Normal Operation

The FORNAX calculations of the fractional release of cesium from the coated particles into
the fuel compact are given in Fig. 7-18 for three of the five core layers. For comparison
purposes, the respective FRESCO results are plotted as dashed curves.

The calculations of cesium release from the particles with the two models show for the core
layer 1 (not displayed in the figure) the dominance of the recoil effect starting at a release
level of about 3% of the fraction of defective particles. While the FORNAX results remain at
that level, the FRESCO results show an additional diffusive release (FZJ diffusion
coefficients for both UO2 and SiC are higher than the Japanese data) reaching approximately
the same release level as from recoil.

For the other core layers, the enhanced rise of the release during the HT phase exceeding the
level of SiC-degraded (failed) particles can be recognized. (The curves referring to layer 5 not
shown in Fig. 7-18 would be roughly between those of layer 3 and layer 4.) The increase of
the release curves, however, is less pronounced in the FZJ calculations starting at a higher
level in the initial phase and remaining somewhat below the JAERI results during the second
half of the operation time. The differences in the final results with higher values achieved

263
with FORNAX in the higher temperature range compared with the FRESCO results are
attributable to the difference in the diffusion coefficients in silicon carbide. The lower
FORNAX results for the layers 1 and 2 might also be explained by the assumption of the non-
intact particle fraction of 8×10-5 to exist as particles with a degraded SiC layer, still keeping a
significant retention capability in the lower temperature range.

Unlike the release from the particles, the release from the fuel compact (FRESCO) and the
release into the sleeve (FORNAX) are not directly comparable. The difference is given by the
modeling of the cesium transport through the gap. FORNAX simulates a
desorption/adsorption process, whereas it is ignored in the FRESCO calculations. The
difference results in a reduction of the release from the compact with FORNAX by a factor of
around 3-10 depending on the temperature. This relation can well be recognized when
comparing the results of the two models.

Figure 7-18. Comparison of FORNAX and FRESCO calculations of the cesium release for the
HTTR core layers 2-4.

7.5. Fuji Study (2003)

The FAPIG-HTGR presented by Fuji Electric [Nakano 2003] is designed as a small modular
220 MW(th) direct-cycle pebble-bed high temperature reactor for electricity production of
100 MW. The objective of this study was the assessment of metallic fission product release
behavior in the core of the FAPIG-HTGR using the FZJ methodology [Verfondern 2004].

264
7.5.1. FAPIG-HTGR Plant Concept

In Japan, the “First Atomic Power Industry Group”, FAPIG, conducted a feasibility study for
a commercial, small modular High Temperature Reactor, the FAPIG-HTGR. Its concept aims
at an electric power generation of 100 MW per unit, low construction cost of < 1200 $/kW(e),
and sufficient inherent safety such that no evacuation is necessary even in the case of
hypothetical accidents. The reactor is designed as a small modular 220 MW(th) direct-cycle
gas turbine pebble-bed reactor with the standard plant consisting of four units. The high
thermal efficiency was estimated to be 46%.

The pebble-bed core of the FAPIG-HTGR has a diameter of 3 m and a height of 11 m. A


multi-pass fuel loading scheme has been selected distinguishing two core regions. Fresh fuel
with a higher thermal power will be loaded to the outer core, while more aged fuel with a
lower power production will be loaded to the inner core. The average power density is
2.6 MW/m3. The helium coolant flows at a rate of 106 kg/s through the active core heating up
from 500 to 900°C at a system pressure of 6 MPa. The nominal maximum burnup of the fuel
is 80,000 MWd/t.

7.5.2. Boundary Conditions

The computer codes FRESCO and PANAMA were applied to assess the release of the
radiologically relevant fission products 137Cs, 90Sr, and 110mAg during the fuel lifetime under
normal operation and core heatup accident conditions of the FAPIG-HTGR.

The design of the spherical fuel element has been selected according to the latest German
reference concept. It is an A3-3 matrix graphite sphere with 60 mm diameter which contains
approximately 11,000 fissile coated particles homogeneously dispersed in the fuel zone of
55 mm diameter with the outermost 5 mm shell remaining fuel-free. The coated fuel particles
consist of a 500 μm diameter oxide fuel kernel with 10.6% enriched uranium surrounded by
subsequent layers of buffer (thickness: 95 μm), inner pyrocarbon (40 μm), silicon carbide
(35 μm), and outer pyrocarbon (40 μm). The heavy metal loading of a fuel element is 7 g with
0.5 g of fissile 235U plus 6.5 g of fertile 238U material.

With regard to the PANAMA calculations to estimate the particle failure fraction, the data for
tensile strength and corresponding Weibull modulus of the SiC, respective values for the
German particle batch EO 1607 have been used. Its strength and modulus represent typical
data, for which the influence (decrease) due to fast neutron fluence was measured explicitly.
The SiC data for the unirradiated state are σo = 834 MPa for the median value of ultimate
tensile strength and mo = 8.02 for the Weibull modulus.

The distinction between an inner core and an outer core was made with a boundary line at a
radius of r = 0.75 m cutting the core radius in two equal distances such that the volume
fractions are 25% for the inner core and 75% for the outer core. These values were used as
weighing factors when reference is made to “total core” values. No mixed reshuffling of the
fuel has been implemented in the model, i.e., a fuel element of the inner (outer) core has spent
its entire lifetime in the inner (outer) core.

The fuel elements are expected to pass 12 times through the core of the FAPIG reactor. The
time duration for one cycle is 116.8 efpd winding up to a maximum life time of the fuel of
1401.6 efpd. It applies to both the inner and the outer core. This assumption seems to be

265
optimistic for fuel in the outer core, since fuel lifetime will presumably be increasing with
core radius. In the equilibrium core, not more than one twelfth or about 8% of the fuel will
have reached maximum lifetime before being discharged. All other fuel would have
experienced a respectively shorter operation time and thus lower burnup and neutron fluence.
Since calculations are done for the total life time of a fuel element, an integration of the
release results to the whole core would represent a conservative approach.

The temperature distribution assumed during normal operation and the core heatup accident
were the result of a prediction conducted by Fuji Electric (Fig. 7-19). Temperatures refer to
the “average channel” in inner core and outer core, respectively. The fuel temperature inside
the sphere and the coated particle is assumed to be constant in the model, which is realistically
correct for the accident phase (shutdown reactor), but not quite correct for the normal
operation phase.

1600
temperature outercore temperature innercore
1400
temperature average

1200
Temperature [°C]

1000

800

600

400

200

0
0 200 400 600 800 1000 1200 1400
Time of normal operation [d]
1600

1400

1200
Temperature [°C]

1000

800

600
temperature outercore temperature innercore

400 temperature average

200

0
0 50 100 150 200
Time into accident [h]
Figure 7-19. Fuel temperatures during normal operation (top) and accident conditions.

266
The temperature distribution for the core heatup accident scenario is again distinguished
between outer and inner core, referring to that position in the core which reaches the
maximum temperature during the core heatup accident. This is found to be the neighboring
grid meshes at the boundary line between inner and outer core in the vertical position
z = 4.125 m.

It should be stressed that, unlike the normal operation phase, fuel temperatures during the core
heatup accident phase are higher in the inner core than in the outer core.

During the normal operation phase, inventory is built up according to the decay constant of
the considered species until reaching 100% at the end of irradiation and beginning of the
accident phase, respectively.

By far most of the fissile material is concentrated in the particle kernels surrounded by an
intact coating. However, there are small fractions of free uranium present as particle kernels
without intact coating and simulated as bare kernels or as contamination of the coating layers
and the matrix graphite. The data given in Table 3-1 (see chapter 3.2.2) have been used in the
calculations.

There are usually three contributions to the fuel particle defect or failure fraction:

(a) Manufacture-induced defects which are existing from the beginning of fuel life and
which typically represent the first step of the failure function. The value
representative of German reference fuel and assumed in the calculations here is
3×10-5.

(b) Irradiation-induced failures which occur during normal operation. The end-of-life
value of 2×10-5 representative for German fuel is assumed here. This EOL value
used regardless of reactor operating conditions or fuel position in the core is added
to the BOL value of 3×10-5. For the study here, this addition is made by an increase
of the failure fraction in six steps starting at a level of 3×10-5 (BOL) and ending at
the level of 5×10-5 (EOL). The step size is 4×10-6, step length is 233.6 efpd (or two
cycles) except for the first and the final step which both are 116.8 efpd. This
simulates a (quasi) linear increase in the failure fraction, which is conservative since
typically an exponential growth of the failure fraction would be expected.

(c) Accident-induced failures that occur at elevated temperatures during heatup


accidents. These failure fractions as a function of accident time and temperature
were estimated by applying the PANAMA model to the accident condition
temperatures (Fig. 7-19). The obtained failure fraction curve (Fig. 7-20) was then
transformed into four more steps to complete the step function for the whole
calculation range.

The final step function for the particle failure fraction as used in the FRESCO calculations is
given in Table 7-7.

267
Table 7-7. Step function for the cp failure fraction as input to the FRESCO calculation.
Normal operation
Operation time [efpd] Particle failure fraction
Step 1 0 3.0×10-5
Step 2 116.8 3.4×10-5
Step 3 408.8 3.8×10-5
Step 4 700.8 4.2×10-5
Step 5 992.8 4.6×10-5
Step 6 1284.8 5.0×10-5
Core heatup accident
Accident time [h] Particle failure fraction
Inner core Outer core Inner core Outer core
Step 7 9 12 5.53×10-5 5.21×10-5
Step 8 23 29 6.1×10-5 5.42×10-5
Step 9 33 40 6.6×10-5 5.63×10-5
Step 10 110 113 7.1×10-5 5.84×10-5

Figure 7-20. Failure probability of EOL coated particles and corresponding step functions
during core heatup accident scenario.

The diffusion coefficients of the different fission product species in the various fuel materials
have been chosen according to the recommendations given in [IAEA 1997] for Germany. The
sorption effect on graphite surfaces has been neglected in these calculations, rather assuming
an unhindered release from the fuel element surface into the coolant.

268
7.5.3. Results

7.5.3.1. Release during Normal Operation

The calculations of the release fractions for 137Cs, 90Sr, and 110mAg during normal operation
were conducted separately for the inner and outer core; the results of both calculations were
then combined to “total core” data considering the above mentioned volume-related weighing
factors.

Figure 7-21 summarizes all curves of release from the coated particles (thin lines) and from
the fuel element (thick lines) of the three metallic radionuclides. Since for normal operation,
the “outer core” curve is always the higher one, above the average curve, because of the
higher temperatures, the dominant contribution to the total core release will come from the
outer core.

1 E-2

1 E-3 silver release from cp silver release from fe


strontium release from cp strontium release from fe

1 E-4 cesium release from cp cesium release from fe

1 E-5
Fraction

1 E-6

1 E-7

1 E-8

1 E-9

1 E-10
0 200 400 600 800 1000 1200 1400
Time of normal operation [d]

Figure 7-21. Fractional release of metallic fission products during normal operation.

The wavy shape of the release curves reflects the temperature cycles of the passes through the
core during the lifetime of the fuel. It is the result of the combined effects of decay, diffusion,
and inventory buildup. For the long-lived isotopes (137Cs, 90Sr), the buildup of the inventory is
nearly linear. Diffusive release is highest in the second half of either cycle and bottom half of
the core, respectively, when temperatures are highest.

The cesium release curves are steadily increasing, but stay below the fraction of failed
particles during the whole lifetime of the fuel. End-of-life release from the coated particles is
1.8×10-5, corresponding to 60% of the inventory of the failed particles. For the given “low”
fuel temperatures, particles with an intact coating are not expected to contribute to the cesium
release. The release of cesium is practically from defective/failed particles only corresponding
to 36% of their inventory (“total core”).

The release of strontium is smaller compared to cesium both from the coated particles and the
fuel element. Due to a low diffusion coefficient in the kernel material UO2 at normal

269
operation temperatures, there is hardly any strontium escaping the place of its origin, even for
defective or failed particles. In addition, a similarly slow diffusive transport of strontium
through matrix graphite reduces the release from the fuel element into the coolant down to
insignificant values.

Silver shows a similar release behavior as cesium. Both species are in the same range of
release from coated particles and from the fuel element. Towards the end of life of the fuel,
silver appears to show the higher release rates. Still there are some interesting differences.
The waves in the silver curves are somewhat more pronounced resulting from the stronger
decay of the 110mAg compared to the long-lived cesium isotope. The fractional release of
silver from the particles in the inner core, i.e., at lower temperatures, is even continuously
decreasing after 400 d showing that diffusive release towards the end of a cycle cannot
compensate for the decay. This is in contrast to the release from outer core particles. Particle
kernels have released more than 50% of their silver inventory at EOL of the fuel; the fraction
for cesium is about 30%, while for strontium, the release fraction is negligible.

Another difference are the sharp, step-like increases of release from the particles at the
moments, when also the number of failed particles has increased in a step according to the
particle failure step function defined before. These transitions are much more smoothed in the
respective cesium release curves. Reason for this difference is the diffusion coefficient in the
UO2 kernel, which is highest for silver over the whole temperature range considered here. It
means that more silver has migrated from the kernels into the coating layers, resulting in a
sharp release step when, according to an assumption in the FRESCO model, the inventory in
the total coating is liberated immediately upon the failure of a particle. It confirms again that
under the given temperature condition, the release behavior of cesium and silver is very
similar, the respective curves being in a fairly narrow range over the whole fuel life time.

7.5.3.2. Release during Core Heatup Accident

The integral fractional release curves for the total core are given in Fig. 7-22 for all three
fission product species investigated. The comparison exhibits the release sequence typically
expected under elevated temperature conditions and often observed with silver being released
most, followed by cesium and strontium. According to the given temperature boundary
condition, the release from the outer core is higher than from the inner core at the beginning
as a result of normal operation release behavior. Since accident temperatures are higher in the
inner core, the release curves for the inner core soon increase such that they eventually
intersect the respective curves for the outer core.

The two thick curves represent the total core and are mainly dominated by the release from
the outer core. The curves of release from the particles approach, but do not exceed the
fraction of failed particles showing that even under the conditions of elevated temperatures,
the cesium release is mainly from the failed particles and not from the intact particles.

The release of strontium from the fuel element during the core heatup accident remains again
on a low level due to a very efficient retention in the matrix graphite even for the accident
temperatures.

270
1 E-2

1 E-3

1 E-4

1 E-5
Fraction

1 E-6 silver release from cp silver release from fe


strontium release from cp strontium release from fe

1 E-7 cesium release from cp cesium release from fe

1 E-8

1 E-9

1 E-10
0 50 100 150 200
Time into accident [h]

Figure 7-22. Fractional release of metallic fission products during core heatup accident.

With respect to the fractional release curves for 110mAg during the accident conditions, silver
release soon exceeds the fraction of failed particles (thin green line) revealing that it is
significantly released from intact particles also. The release curves for the particles (blue) are
very close to the corresponding curves for the fuel element (red) showing that there is a rapid
transport of silver through matrix graphite, i.e., practically no buffer effect of the matrix, at
elevated temperatures.

In terms of release from the particle kernels, the silver fraction outside the kernels is 73%, for
cesium 58%. For strontium, in comparison to normal operation, the release fraction from the
kernels has increased by several orders of magnitude up to 0.1%, but still is comparatively
low.

Also the retention capability of the matrix graphite can be seen directly given by the
difference between release from the particles (thin lines) and release from the fuel element
(thick lines). The inventory in the matrix is lowest for silver (9% of the amount released from
the particles) and largest for strontium (~100%) with cesium somewhere in between (40%).
The percentage data are valid for the moment 180 h.

271
7.6. German and Chinese Fuel Spheres Irradiated in HFR-EU1 (2006)

Within the European irradiation testing program for HTGR fuel, the so-called HFR-EU1
irradiation experiment was initiated with the goal to explore the performance limits of the
presently existing German and Chinese high-quality fuel [Conrad 2002, Laurie 2010]. This
test with a focus on high burnup (towards 20% FIMA) was started in September 2006 and
terminated in February 2010 with an almost 2-years discontinuation of the test due to reactor
shutdown. The objective of this study [Mao 2006], which was conducted prior to the
irradiation test, was predicting the metallic fission product release behavior during the
irradiation test and an assumed postirradiation heating experiment using the FZJ computer
codes FRESCO-II and PANAMA.

7.6.1. Boundary Conditions

There were totally five spherical fuel element samples to be tested in this experiment, three
German GLE-4 fuel elements (AVR reload 21/2) and two China INET fuel elements. Fuel
characteristics are listed in Table 7-8. Test conditions as were anticipated at that time and on
which the predictive calculations were based, are given in Table 7-9.

Table 7-8. Characteristic data of AVR GLE-4 and INET spherical fuel elements with
LEU TRISO UO2 fuel particles as loading for the experiment HFR-EU1.
AVR GLE-4/2 spheres INET spheres
Diameter [mm] 60 60
Matrix graphite grade A3-3 A3-3
Heavy metal loading [g/FE] 6.0 5.02
U-235 contents [g/FE] 1.00 0.858
Volume packing fraction [%] 6.2 5.0
Defective SiC layers [U/Utot] 7.8×10-6 2.3×10-7
Coated particle batch HT 385-393, 395-404, 406-423 V000802
UO2 kernel diameter [μm] 502 490
Enrichment [U-235 wt.%] 16.76 and 16.67 17.08
Coating thickness [µm]
Buffer 95 98
Inner PyC 41 42
SiC 35 38
Outer PyC 40 41

272
Table 7-9. Nominal irradiation characteristics of fuel elements of HFR-EU1 test.
AVR GLE-4 spheres INET spheres
Id. No 3 2 1 I-2 I-1
Fuel temperature center [°C] 1100 1100 1100 1000 1000
Fuel temperature surface [°C] 950 930 930 870 890
Fission power per FE [kW] 1.95 2.15 2.3 1.75 1.45
Fission power per CP [mW] 204 225 241 206 171
Burnup [% FIMA] 18.6 20.8 21.9 17.0 13.7
Fast neutron fluence [n/m2, 5.1×1025 5.7×1025 6.0×1025 4.6×1025 3.8×1025
E>0.1 MeV]

Since calculations are conducted for constant temperatures in the fuel sphere, which is not
quite correct for the irradiation phase, FRESCO for both surface and center temperature. For
the postirradiation core heatup accident simulation test, a potential temperature-time history
has been assumed consisting of a 1600°C and a 1800°C heating phase (Fig. 7-23).

2100

1800

1500
Temperature [°C ]

1200

900

600

300

0
0 25 50 75 100 125 150 175 200 225 250
Time to accid ent co ndi tion heati ng [h]

Figure 7-23. Temperature history assumed for the heating test of HFR-EU1 spheres.

By far most of the fissile material is concentrated in the particle kernels surrounded by an
intact coating. However, there are small fractions of free uranium present as particle kernels
without intact coating and simulated as bare kernels or as contamination of the coating layers
and the matrix graphite. The data given in Table 3-1 (see chapter 3.2.2) have been used in the
calculations.

273
There are three contributions to the fuel particle defect/failure fraction:

(a) Manufacture-induced defective particles typically represent the first step of the
failure function. The value representative for the German reference fuel is 3×10-5.
For the Chinese spheres, a value of 5×10-5 has been taken representative for the
average free uranium fraction of the first loading for the HTR-10 [Tang, 2001].

(b) Irradiation-induced failure may occur during the irradiation. The end-of-life value
(EOL) of 2×10-5 is the expected value of German reference fuel for the HTR-Modul
(10% FIMA) [IAEA, 1997]. For the high burnup fuel elements of the experiment
HFR-EU1, respective PANAMA results were taken, if they exceeded the above
German reference value.

The PANAMA calculated failure fraction curve or the linear increase from 0 to
2×10-5 as the German reference curve were transformed into step functions to adjust
to the required FRESCO input. These functions with a step length of 120 efpd
except for the first and the final step are shown in Fig. 7-24 for the irradiation
phase. Please note that the values for the Chinese spheres start at a higher level due
to the higher fraction of free uranium.

0.00

-0.50
Particle Failure Fraction [lg]

-1.00 GLE-4 at 20%FIM A 950°C

-1.50 GLE-4 at 20%FIM A1100°C

-2.00 INET at 15%FIM A 880°C &1000°C

-2.50

-3.00

-3.50
-4.00

-4.50

-5.00
0 50 100 150 200 250 300 350 400 450 500 550 600

Time [efpd]

Figure 7-24. Step function for manufacture plus irradiation-induced particle failure
as input to the FRESCO calculations.

(c) Accident-induced failure fractions as a function of time and temperature have also
been estimated by PANAMA. The obtained failure curves translated into a step
function with four steps (since six steps were already taken for the irradiation
phase), are plotted in Fig. 7-25. The figure includes as dashed line the heating
temperature transient.

274
0.00

Particle Failure Fraction [lg]


-1.00

-2.00

-3.00

-4.00

-5.00

-6.00

-7.00

-8.00
14400 14425 14450 14475 14500 14525 14550 14575 14600 14625 14650

Time [hours]

GLE-4 at 20%FIM A 950°C GLE-4 at 20%FIM A 1100°C INET at 15%FIM A 880°C


INET at 15%FIM A 1000°C Temprature

Figure 7-25. Step function for manufacture plus irradiation-induced particle failure
as input to the FRESCO calculations.

The diffusion coefficients of the different fission product species in the various fuel materials
have been chosen according to the recommendations given in [IAEA 1997] for Germany. The
sorption effect on graphite surfaces has been neglected in these calculations, rather assuming
an unhindered release from the fuel element surface into the coolant.

7.6.2. Radionuclide Release

The calculations of the metallic fission product release with the diffusion code FRESCO have
been conducted for the radiologically relevant isotopes 137Cs, 90Sr and 110m Ag for both the
German and the Chinese sphere at final conditions of the HFR-EU1 irradiation test. Two
calculations were made for each of these six cases, assuming the irradiation temperature to be
either the surface or the center temperature. The fractional release values achieved at the end
of irradiation, the 1600°C heating phase and the 1800°C heating phase are listed in
Table 7-10.

7.6.2.1. Cesium Release

The release behavior of cesium for the German fuel element is shown in Fig. 7-26a for the
irradiation phase and in Fig. 7-26b for the heating phase. The blue curves representing the
release from the coated particles (@ 950 and 1100°C) are approaching the fraction of failed
particles and also following the steps of the failure function towards the end of the irradiation
time. This indicates that the cesium release is mainly coming from the failed particles. The
two red curves represent the corresponding release from the fuel sphere smoothed by the
“buffering” effect of the matrix graphite.

275
0

Release fro m P articles at 950°C


-1
Release fro m FE at 950°C
Release fro m P articles at 1100°C
-2 Release fro m FE at 1100°C
Release Fraction [Lg] P artical Failure Fractio n GLE-4 at 1100°C
-3 P artical Failure Fractio n GLE-4 at 950°C

-4

-5

-6

-7

-8

-9

-10
0 2400 4800 7200 9600 12000 14400

Time [Hours]

Figure 7-26a. Predicted 137Cs release from German fuel sphere in HFR-EU1
during irradiation test.

-1

-2
Release fro m P articles at 950°C
Release Fraction [Lg]

-3 Release fro m FE at 950°C


Release fro m P articles at 1100°C
-4 Release fro m FE at 1100°C
P artical Failure Fractio n GLE-4 at 1100°C
P artical Failure Fractio n GLE-4 at 950°C
-5 Temperature

-6

-7

-8

-9

-10
14401 14430 14459 14488 14517 14546 14575 14604 14633 14662

Time [Hours]

Figure 7-26b. Predicted 137Cs release from German fuel sphere in HFR-EU1
during heating test.

276
0

-1 Release fro m P articles at 880°C Release fro m FE at 880°C


Release fro m P articles at 1000°C Release fro m FE at 1000°C
-2 P artical Failure Fractio n INET at 1000°C P artical Failure Fractio n INET at 880°C
T t

-3
Release Fraction [Lg]

-4

-5

-6

-7

-8

-9

-10
0 2400 4800 7200 9600 12000 14400

Tim e [Hours]

Figure 7-27a. Predicted 137Cs release from Chinese fuel sphere in HFR-EU1
during irradiation test.

-1

-2

-3
Release Fraction [Lg]

-4

-5

-6

-7

-8
Release fro m P articles at 880°C Release fro m FE at 880°C
Release fro m P articles at 1000°C Release fro m FE at 1000°C
-9 P artical Failure Fractio n INET at 1000°C P artical Failure Fractio n INET at 880°C
Temperature
-10
14401 14430 14459 14488 14517 14546 14575 14604 14633 14662

Tim e [Hours]

Figure 7-27b. Predicted 137Cs release from Chinese fuel sphere in HFR-EU1
during heating test.

277
The cesium release under heating conditions shows a similar behavior. The release from the
coated particles again closely follows the failure function with a delayed release from the fuel
element. The results confirm that for the irradiation and the 1600°C phase, the release from
the particles is practically from failed particles only. The contribution from the intact
particles, i.e., diffusion through the intact SiC layer, becomes gradually larger during the
1800°C heating phase. The difference between the blue and the red curves is the inventory in
the matrix graphite showing that at the end of irradiation, between 2-9% of the cesium, which
was released from the particles, also left the fuel element. For the heating phase, this share
grows to 39-52 % @ 1600°C and around 90 % @ 1800°C.

The corresponding cesium release results for the Chinese fuel element are given in Figs. 7-27a
and 7-27b. The release behaviour is similar to that for the German sphere with the release
from the particles closely following the particle failure function. The release from the fuel
element during irradiation remains comparatively lower because of the lower temperatures
and despite the somewhat higher failure fractions. During the heating phase, release curves
remain lower because of the significantly lower failure fractions.

The red figures in the table, which are “negative releases” or “increasing inventories” seem to
indicate some numerical problem of the code handling the large steps of the failure function
increase, which are characterized by the sudden liberation of all inventory in the coatings of
the breaking particles.

For the specified operating conditions of the reactor design considered and for the typical
German reference spherical fuel element, the results show that under the given thermal
hydraulic boundary conditions, the release remains on a very low level for all radionu¬clides
investigated. For cesium, the dominant release is from defective/failed particles even for
accident temperatures. The release of strontium remains always insignificant due to its
enormous retention capability in matrix and kernel material. Silver release is low during
normal operation similar to cesium, but shows significant release, even from intact coated
particles under the conditions of elevated temperatures during the core heatup accident. From
the perspective of these low metallic fission product re¬lease results, the FAPIG-HTGR can
be judged to have a safe design that can be main¬tained easily.

7.6.2.2. Strontium Release

The calculation results for strontium for the cases studied here are shown in Figs. 7-28a and
7-28b for the German sphere and in Figs. 7-29a and 7-29b for the Chinese sphere. In general,
the strontium release behaviour is characterized by a very slow diffusive transport both
through the kernel material UO2 and the matrix graphite. The release from the particle kernels
remains very low, even at 1600°C only a few percent. This explains why there is no effect on
the release curves when there are steps in the failure function.

As can be seen from the plots, the strontium release from the fuel elements remains on a low
level during irradiation reaching a value of about 10-8. The difference to the level of release
from the particles of three orders of magnitude is due to the slow diffusive transport in the
graphite.

Under heating conditions, the fractional release of strontium from the fuel element increases
to about 2×10-4 after 100 h @ 1600°C and to 3-5% after additional 100 h @ 1800°C, being
lower than the corresponding values for cesium.

278
0

Release fro m P articles at 950°C


-1 Release fro m FE at 950°C
Release fro m P articles at 1100°C
-2 Release fro m P articles at 1100°C
P artical Failure Fractio n GLE-4 at 1100°C
-3 P artical Failure Fractio n GLE-4 at 950°C
Release Fraction [Lg]

-4

-5

-6

-7

-8

-9

-10
0 2400 4800 7200 9600 12000 14400

Tim e [Hours]

Figure 7-28a. Predicted Sr-90 release from German fuel sphere in HFR-EU1
during irradiation test.

-1

-2

-3
Release Fraction [Lg]

-4

-5

-6

Release fro m P articles at 950°C


-7
Release fro m FE at 950°C
Release fro m P articles at 1100°C
-8 Release fro m P articles at 1100°C
P artical Failure Fractio n GLE-4 at 1100°C
-9 P artical Failure Fractio n GLE-4 at 950°C
Temperature

-10
14401 14430 14459 14488 14517 14546 14575 14604 14633 14662

Tim e [Hours]

Figure 7-28b. Predicted Sr-90 release from German fuel sphere irradiated in HFR-EU1
during heating test.

279
0

-1 Release fro m P articles at 880°C Release fro m FE at 880°C


Release fro m P articles at 1000°C Release fro m FE at 1000°C
-2 P artical Failure Fractio n INET at 1000°C P artical Failure Fractio n INET at 880°C
T t

-3
Release Fraction [Lg]

-4

-5

-6

-7

-8

-9

-10
0 2400 4800 7200 9600 12000 14400

Tim e [Hours]

Figure 7-29a. Predicted Sr-90 release from Chinese fuel sphere in HFR-EU1
during irradiation test.

-1

-2

-3
Release Fraction [Lg]

-4

-5

-6

-7

-8
Release fro m P articles at 880°C Release fro m FE at 880°C
Release fro m P articles at 1000°C Release fro m FE at 1000°C
-9 P artical Failure Fractio n INET at 1000°C P artical Failure Fractio n INET at 880°C
Temperature
-10
14401 14430 14459 14488 14517 14546 14575 14604 14633 14662

Tim e [Hours]

Figure 7-29b. Predicted Sr-90 release from Chinese fuel sphere irradiated in HFR-EU1
during heating test.

7.6.2.3. Silver Release

The calculated release behaviour of 110mAg is shown in Figs. 7-30a and 7-30b for the German
sphere and in Figs. 7-31a and 7-31b for the Chinese sphere.

280
0

-1

-2

-3
Release Fraction [Lg]

-4

-5

-6
Release fro m P articles at 950°C
Release fro m FE at 950°C
-7
Release fro m P articles at 1100°C
Release fro m FE at 1100°C
-8 P artical Failure Fractio n GLE-4 at 1100°C
P artical Failure Fractio n GLE-4 at 950°C
-9 Temperature

-10
0 2400 4800 7200 9600 12000 14400

Tim e [Hours]

Figure 7-30a. Predicted 110mAg release from German fuel sphere in HFR-EU1
during irradiation test.

-1

-2

-3
Release Fraction [Lg]

-4

-5

-6

Release fro m P articles at 950°C


-7
Release fro m FE at 950°C
Release fro m P articles at 1100°C
-8 Release fro m FE at 1100°C
P artical Failure Fractio n GLE-4 at 1100°C
-9 P artical Failure Fractio n GLE-4 at 950°C
Temperature

-10
14401 14430 14459 14488 14517 14546 14575 14604 14633 14662

Tim e [Hours]

Figure 7-30b. Predicted 110mAg release from German fuel sphere irradiated in HFR-EU1
during heating test.

281
0

-1 Release fro m P articles at 880°C Release fro m FE at 880°C


Release fro m P articles at 1000°C Release fro m FE at 1000°C
-2 P artical Failure Fractio n INET at 1000°C P artical Failure Fractio n INET at 880°C
T t

-3
Release Fraction [Lg]

-4

-5

-6

-7

-8

-9

-10
0 2400 4800 7200 9600 12000 14400

Tim e [Hours]

Figure 7-31a. Predicted 110mAg release from Chinese fuel sphere in HFR-EU1
during irradiation test.

-1

-2

-3
Release Fraction [Lg]

-4

-5

-6

-7

-8 Release fro m P articles at 880°C Release fro m FE at 880°C


Release fro m P articles at 1000°C Release fro m FE at 1000°C
-9 P artical Failure Fractio n INET at 1000°C P artical Failure Fractio n INET at 880°C
Temperature
-10
14401 14430 14459 14488 14517 14546 14575 14604 14633 14662

Tim e [Hours]

Figure 7-31b. Predicted 110mAg release from Chinese fuel sphere irradiated in HFR-EU1
during heating test.

282
The silver release behavior is typically characterized by a high diffusive transport even
through intact coating at temperatures above 1000°C. Furthermore, there is a fast transport of
silver through the matrix graphite corresponding to a small retention capability of the
graphite. The latter effect can be observed by the fact that in all cases, the curves of Ag
release from the particles and from the fuel element are very close or almost identical. The
former effect (of a relative rapid transport through intact SiC) can be recognized in the release
curves being nearly independent of the particle failure function. Only in the cases of lower
temperatures (950°C in Fig. 7-30a or 880°C in Fig. 7-31a), a step-like release can be
identified. The corresponding curves for release from the particles even show partially a
decrease in the fractional release (only during irradiation) indicating that at these
temperatures, the decay of the silver isotope dominates the diffusive release.

Silver release reaches as expected high values already during the irradiation phase which is
0.2% for the Chinese sphere (Tirr = 1000°C) and as high as 7% for the German sphere
(Tirr = 1100°C). During the heating phases, the silver release is further strongly increasing and
reaches final values of more than 70%.

283
Table 7-10. Predicted fractional release of 137Cs, 90Sr, and 110mAg from the German and Chinese fuel spheres after irradiation in the HFR-EU1 experiment.
Fuel Irradiation Burnup Time Cesium-137 from Strontium-90 from Silver-110m from
element temperature [% FIMA] [h]
CP FE CP FE CP FE
[°C]
End of HFR-EU1 irradiation
950 20 14400 4.64×10-5 7.58×10-7 2.12×10-6 1.38×10-9 2.17×10-4 1.28×10-4
AVR
1100 20 14,400 1.45×10-2 1.34×10-3 1.49×10-4 1.08×10-8 6.73×10-2 6.61×10-2
880 15 14400 1.77×10-5 6.37×10-8 1.11×10-6 3.59×10-10 4.40×10-6 1.70×10-6
INET
1000 15 14400 4.28×10-5 4.44×10-6 8.67×10-6 2.84×10-9 2.29×10-3 1.92×10-3
End of first heating phase after 100 h at 1600°C
950 20 14534 1.19×10-1 4.66×10-2 7.31×10-3 2.25×10-4 2.77×10-1 2.77×10-1
AVR
1100 20 14,534 2.89×10-1 1.50×10-1 5.26×10-3 1.69×10-4 4.09×10-1 4.09×10-1
880 15 14534 1.25×10-3 3.85×10-4 5.10×10-3 9.44×10-5 1.88×10-1 1.87×10-1
INET
1000 15 14534 1.25×10-2 6.58×10-3 5.06×10-3 9.71×10-5 2.11×10-1 2.10×10-1
End of second heating phase after additional 100 h at 1800°C
950 20 14663 6.77×10-1 5.86×10-1 1.13×10-1 5.02×10-2 7.62×10-1 7.62×10-1
AVR
1100 20 14,663 6.80×10-1 6.22×10-1 7.94×10-2 3.28×10-2 7.49×10-1 7.49×10-1
880 15 14663 1.26×10-1 7.74×10-2 1.70×10-1 6.06×10-2 6.77×10-1 6.76×10-1
INET
1000 15 14663 4.64×10-1 3.45×10-1 1.56×10-1 6.05×10-2 7.43×10-1 7.43×10-1

284
8. CONCLUSIONS

The postcatculation of heating experiments is a compulsory part of the verification and


validation procedure for computer models which are to be used for fuel performance and
fission product release predictions in risk analyses. They provide information about either
confirmation or the necessity of changing basic input data. At least they are able to
demonstrate the sensitivity of various input parameters that might be useful to define the
direction for further efforts.

Much experience has been gained with computer models in the past predicting and
postcalculating fission product release in irradiation and heating experiments as well as for
normal operation and core heatup accidents for numerous designs of small and medium sized
HTGRs. Purpose of these calculations was first of all a model verification to ensure that the
theory and principles are clearly understood, and a model validation by comparing
computational results with well documented experiments and with the results of other codes.

The objective of this report was the detailed presentation of the FRESCO-II code developed
in the early 1980s and a description of the comprehensive verification and validation efforts
for this code up to date, which permit judgement on its applicability for safety analyses within
the frame of licensing procedures for pebble-bed HTGRs.

The systematic utilization of the computer model FRESCO was done in several phases:

(a) Initial application to a few German and US heating experiments

(b) Application to the FZJ heating experiments conducted in the KÜFA furnace that
provided many more experimental data on complete fuel spheres from a
comprehensive post-heating examination program and that was more concentrating on
the lower temperature range 1600-1800°C with the objective to fix relevant diffusion
coefficients (mainly SiC and matrix graphite);

(c) Application to various HTGR reactor concepts based on – initially – only scanty
experimental data with concentration on the high temperature range up to 2500°C,
later also including small modular-type HTGRs and gradually improving data base;

(d) Application as part of an international benchmark exercise directed by IAEA (CRP-6)


with comprehensive code-to-code and code-to-experiment comparisons;

(e) Application to two series of experiments, the late FZJ-KÜFA tests (early 1990s) based
on the HTR-Modul accident temperature transient and the first ITU-KÜFA tests
(2008-2009) resumed after a longer hibernation period.

From the various studies conducted, the following conclusions can be drawn:

 Main uncertainties are most likely introduced less by shortcomings of the model, but
rather by input data strongly depending on material properties and exhibiting a broader
variability.

 In the CRP-6 accident benchmark, eight codes have been applied to all or part of the
cases suggested. Of these, five codes represented developments from the recent years
to be compared with three“older” codes (including FRESCO) where development was
often constraint by requirements of computer technology at that time. This benchmark

285
represents up to now the best validation effort for FRESCO providing the opportunity
for confirmation/validation by code-to-code and code-to-experiment comparison, and
identifying the areas for further improvements. Most codes have shown good
agreement among each other. Some differences can be explained by different
assumptions for input data or boundary conditions.

 The diffusion coefficient is the essential input parameter to a diffusion code. The
recommended diffusion coefficients are a useful basis for scoping calculations. They
have, however, drawbacks. In most cases of postcalculations of heating tests with the
FRESCO code, the results prove to lead to a conservative coverage of the
experimental data. A drawback, however, is given in that it appears from the
comparison with the measurements that some important reference diffusion
coefficients are still varying over a broader range.

 It would be difficult to demonstrate conclusively that all fission product transport


mechanisms in coated particle and fuel element components are single solid state
diffusion processes. This is because the materials have a complex structure: the
diffusion model based on Fick's law serves rather as a practical interpolation
procedure to reduce all available experimental evidence into the diffusion coefficient
as a single constant that – in most cases – is only temperature dependent.

 Some of the currently recommended diffusion coefficients have been derived by


FRESCO calculations representing average data from the heating experiments
considered. This seems to be an appropriate approach to be used in postcalculations of
experiments.

 From the code-to-code comparison in the CRP-6 benchmark, it became clear that
FRESCO was obviously employing a different diffusion coefficient for Cs in SiC.
Although formally correct, it was inconsistent with the description given in [IAEA
1997]. The correct recommended formula is the sum of the low-temperature “HRB-
branch” with the high-temperature “FZJ-branch”:
   125,000   514,000  m 2
D  5.5  10 14 exp  exp    1 . 6  10 2
exp  
5  RT   RT  s
where Г – fast neutron fluence, 1025 n/m2, E>16fJ.

 In general, it is not possible at this stage to provide a range of uncertainty in the


internationally agreed recommended transport data. If this is required, e.g., for design
calculations in licensing, it is then necessary to re-analyze the specifically applicable
data base and derive the range of confidence.

 The above mentioned 1989 study [Verfondern 1989] where 44 heating tests with
regard to cesium release were evaluated, formed the basis for a new high-temperature
branch of the diffusion coefficient in SiC with the frequency factor Do,2 = 0.016 m2/s
and the activation energy Q2 = 514 kJ/mole [IAEA 1997]. A statistical analysis of the
same data allows to derive the upper 95% confidence level of that curve that may be
regarded as a "design basis" diffusion coefficient with Do,2,db = 0.19 m2/s being the
new frequency factor, while the activation energy remains the same, Q2 = Q2,db.

286
 A re-analysis was also done in the context of silver release in a direct-cycle gas turbine
HTGR [Van der Merwe 2009] with new relationships for a diffusion coefficient in SiC
for “best estimate” and “design limit”.

 It is obvious from postcalculations that in particular the strontium data are largely
overpredicted in most cases. While the retention capability of UO2 and the matrix
material of the fuel spheres is visible, it is most probably the diffusive transport
through the silicon carbide layer that appears very conservative and should undergo a
thorough review. It needs a revision particularly of the data for the lower temperature
range.

 Silver release measurements are often unusual and inconsistent, and therefore
extremely difficult for postcalculation. There are presumably other mechanisms for
transport and retention working than treated by a simple diffusion model, which would
then require at least a conservative coverage of the experimental data.

 Iodine release behavior was found in heating tests to be similar to that of fission gases
with the released quantity being proportional the number of defective/failed coated
particles. Despite the observation that the iodine fractional release is generally lower
than xenon or krypton release, the same diffusion coefficient for fission gases and
iodine in the particle kernels is used which is considered conservative.

 The experimental data on fission product release in heating tests simulating the core
heatup accident temperature history up to 1620°C confirm in principle the model
assumptions and model data used at FZJ.

 For the heating tests with the HTR-Modul accident transient temperature, all AVR-
irradiated spheres, to 1620°C and with no particle failures, the measured cesium
release was ~1×10-5 and originated all from contamination. Postcalculations typically
revealed a release of 2×10-5, the difference presumably being a question of assumption
for the level of cross-contamination with cesium in the AVR core. Silver release in
these tests, rapid and principally unaffected by the presence of failed particles, is
calculated to be around 20%, while the respective measurements were in a broad range
between 10-3 and 10-1.

 The new (KÜFA-II) heating tests HFR-K6/2 and HFR-K6/3 representing proof test
fuel for the HTR-Modul have shown excellent retention behavior for fission products.
Metallic fission product release remained well below the expected levels from
previous experience. Release is largely overpredicted by the model calculations. The
comparison of cesium and krypton release gives indication that during
irradiation/heating particles with a failed SiC layer, but still intact oPyC layer may
have developed.

 In contrast, the heating tests with the spheres irradiated in HFR-EU1bis revealed
worse retention behavior for cesium. Although obviously no pressure-induced particle
failure occurred, cesium release level was comparatively high already at heating
temperatures as low as 1250°C and reached the 10-3 range at the end of the tests. Main
reasons for this behavior are believed to be on the one hand the high irradiation
temperature of 1250°C and on the other hand the occurrence of a temperature
excursion during the irradiation experiment that presumably caused impairment of
quite a number of coated particles. In addition, individual information was lost, since

287
all five fuel spheres were placed in one capsule. Therefore, postcalculations may be
considered “good” ones, although they were slightly below the experimental data.

From the experience with FRESCO code applications, several areas for further model
improvement have been identified (some of which are explained in more detail in China
Report 7):

 A shortcoming of FRESCO early recognized when small modular reactors came into
the focus where the normal operation phase is gaining importance, is the disregard of
the temperature profile inside the fuel sphere during irradiation. For a more realistic
modeling, the fuel temperatures inside the sphere should be calculated correctly.

 Another subject of improvement is the inventory buildup of a considered fission


product isotope during the normal operation phase. Current model approach is to have
the isotope build up according to its decay constant. Goal is to have the full inventory
available in the particle kernel at the end of the irradiation phase which will go rapidly
for short-lived and linearly for long-lived isotopes. However, ignoring all other
nuclear reactions to serve as potential sources or sinks introduces a certain error when
handling, e.g., activation products such as 110mAg or 134Cs. The ideal solution of this
problem is to take inventory data directly from an autonomously operating neutronics
calculation module (as is planned with the HCP model described in China Report 7).

 The particle failure function is currently limited to 10 steps in FRESCO-II, a relict


from the computational limitations in earlier times. Today’s possibilities should allow
to have a more precise approach to taking account of the failure of particles by using
the PANAMA results directly in FRESCO-II rather than converting them first into a
step function.

 As a result from CRP-6, the comparatively poor performance of the FRESCO model
regarding release from the kernel due to the limited mesh number, it is recommended
to replace the original diffusion calculation inside the kernel (based on only a very
limited number of nodes) with the analytical Booth formula for gas release.

 As a conclusion from this mistake in FRESCO, for all diffusion coefficients composed
of two branches (and therefore implemented in the code) should have these as a low-
temperature plus a high-temperature branch. An additional distinction between
“normal operation” branch and “accident condition” branch should be removed.

As a module of the currently on-going development of the HTR Code Package (HCP)
[Allelein 2010], the new calculation tool STACY will allow to comprehensively describe the
fission product transport and release behavior in an HTGR core. STACY is actually based on
the principles of the (combined) PANAMA and FRESCO (and the plate-out code SPATRA)
models. With the consideration of the results from a thermohydraulics code in terms of
temperature and coolant flow distribution as input (a characteristic feature of the core-version
FRESCO-I), it includes the possibility of determining the (re-)distribution of fission products
within the primary circuit. Also the inventories of radionuclides will be taken as input from
separate burnup calculations. Apart from eliminating many of the weaknesses of the FRESCO
and PANAMA models, additional features such as the simulation of the pebble flow by
tracking individual fuel spheres and their reshuffling in a multiple passage reactor core, the
new code goes much beyond the capablilities of the sum of the existing models so far.
STACY will also be available as stand-alone version.

288
REFERENCES

Allelein H.J., Spaltproduktverhalten - speziell Cs-137 - in HTR-TRISO Brennstoffteilchen.


Report Jül-1695, Research Center Jülich, 1980.

Bell M.J., ORIGEN - The ORNL Isotope Generation and Depletion Code. Report
ORNL-4628, Oak Ridge National Laboratory, TN, 1973.

Booth A.H., A Method of Calculating Fission Gas Diffusion from UO2 Fuel and Its
Application to the X-2-f Loop Test. Report CRDC-721, AECL No. 496, Chalk River,
Ontario, 1957.

Brinkmann G., et al., Important Viewpoints Proposed for a Safety Approach of HTGR
Reactors in Europe, Final Results of the EC-Funded HTR-L Project. Nuclear Engineering and
Design 236:463-474, 2006.

Carretero J.A., Current Requirements for the Containment. EU Project HTR-L, Deliverable
HTR-L-02/06-0-3 .0.2., 2002.

Christ A., Nachrechnung von Ausheizexperimenten. Technical Note HBK-5125-BF-GHRA


001555, BBC/HRB Mannheim, 1985.

Conrad R., Bakker K., Nabielek H., Verfondern K., HFR-EU 1 Test Specification for the
Irradiation Experiment HFR-EU1 within HTR-TN. Technical Memorandum, HFR / 02 /
4681, Revision No 5, Petten, 2002.

Crank J., The Mathematics of Diffusion. Oxford University Press, London, 1956.

Crank J., The Mathematics of Diffusion. 2nd Edition, Oxford University Press, New York,
1975.

Cubiciotti D., The Diffusion of Xenon from Uranium Carbide Impregnated Graphite at High
Temperatures. Report NAA-SR-194, North American Aviation, Downey, CA, 1952.

Ehrhardt J., et al., The Program System UFOMOD to Assess the Consequences of Nuclear
Accidents. Report KfK-4330, Research Center Karlsruhe, 1988.

Flowers R.H., Fission Product Control in the HTR. Proc. BNES Conference on Nuclear Fuel
Performance, Paper 23, London, 1973.

Flügge S., Zimens K.E., Die Bestimmung von Korngrößen und von Diffusionskonstanten aus
dem Emaniervermögen. Zeitschrift für Physikalische Chemie B42:179-220, 1938.

Förthmann R., Grübmeier H., Stöver D., Metallic Fission Product Retention of Coated
Particles with Ceramic Kernel Additives. Nuclear Technology 35:548-556, 1977.

Freis D., et al., Post Irradiation Testing of High Temperature Reactor Spherical Fuel Elements
under Accident Conditions. Proc. 4th International Topical Meeting on High Temperature
Reactor Technology HTR-2008, Paper 58203, Washington D.C., 2008.

289
Freis D., Störfallsimulationen und Nachbestrahlungsuntersuchungen an kugelförmigen
Brennelementen für Hochtemperaturreaktoren. PhD Thesis, Technical University RWTH
Aachen, 2010.

Fütterer M., et al., Irradiation Results of AVR Fuel Pebbles at Increased Temperature and
Burn-Up in the HFR Petten. In: Proc. of the 3rd International Topical Meeting on High
Temperature Reactor Technology HTR2006, Paper B00000035, Johannesburg, 2006.

GA, Fuel Design Data Manual, Issue F. Report GA-901866, General Atomics, San Diego,
CA, 1987.

GMBL, Rahmenempfehlungen für den Katastrophenschutz in der Umgebung kerntechnischer


Anlagen. Gemeinsames Ministerialblatt Nr. 40, p. 71, 1989.

GRS, Incident Calculation Bases for the Guidelines Issued by the BMI for the Assessment of
the Design of PWR Nuclear Power Plants Pursuent to § 28.3 of the Radiological Protection
Ordinance (18.10.1983). Translation Safety Codes and Guides; Editor: GRS Quantitative
Examinations, Köln, 1983.

HBK, Projektbericht 1985. Internal Report KFA-HBK-IB-1/86, Research Center Jülich, 1986.

Hensel W., Hoinkis E., The Diffusion of Sr in the Graphite Matrix A3-3 in Vacuum and in the
Presence of Hydrogen. Journal of Nuclear Materials 184:88-96, 1991.

Hensel W., Hoinkis E., The Diffusion of Cesium in the Graphitic Matrix A3-3 in the Presence
of Helium at Pressures up to 107 Pa. Journal of Nuclear Materials 224:1-11, 1995.

Hilpert K., Determination of Adsorption Isotherms of Sr on Matrix and Graphites – Influence


of Corrosion, BET Surface, Iodine and Water. In: Hoinkis E. (Ed.), Transport of Fission
Products in Matrix and Graphite. Proc. Colloquium, Berlin, Germany, 1981, Report
HMI-B 372, pp. 139-146, Hahn-Meitner-Institut, Berlin, 1983.

Hilpert K., et al., Sorption of Strontium by Graphitic Materials. Berichte der Bunsengesell-
schaft für Physikalische Chemie 89:43-48, 1985.

Hilpert K., et al., Sorption of Cesium and its Vaporization from Graphitic Materials at High
Temperatures. High Temperatures High Pressures 20:157-164, 1988.

Hoinkis E., The Determination of Diffusion Coefficients of Cesium and Silver by the Release
Method in As-received, Oxidized and Neutron Irradiated Graphitic Matrix. In: Hoinkis E.
(Ed.), Transport of Fission Products in Matrix and Graphite. Proc. Colloquium, Berlin,
Germany, 1981, Report HMI-B 372, pp. 77-102, Hahn-Meitner-Institut, Berlin, 1983.

Hoinkis E., The Diffusion of Silver in the Graphite Matrices A3-3 and A3-27. Journal of
Nuclear Materials 209:132-147, 1994.

IAEA, Procedures for Conducting Probabilistic Safety Assessment of Nuclear Power Plants
(Level 2). Safety Series No. 50-P-8, International Atomic Energy Agency, Vienna, 1995.

IAEA, Fuel Performance and Fission Product Behavior in Gas Cooled Reactors. Report
IAEA-TECDOC-978, International Atomic Energy Agency, Vienna, 1997.

290
IAEA, Advances in HTGR Fuel Technology. Report IAEA-TECDOC, International Atomic
Energy Agency, Vienna, to be published in 2011.

ICRP, Protection of the Public in the Event of Major Radiation Accidents: Principles for
Planning. ICRP-40, International Commission on Radiological Protection, 1984.

INL, HTGR Mechanistic Source Terms, White Paper. Report INL/EXT-10-17997, Idaho
National Laboratory, Next Generation Nuclear Plant Project, Idaho Falls, ID, 2010.

Jost W., Diffusion in Solids, Liquids, Gases. 3rd Printing with Addendum, Academic Press,
New York, 1960.

Kazi N.I., Zumwalt L.R., Sorption of Cesium by Barium-Impregnated Nuclear Graphite.


Transactions of the American Nuclear Society 34:214-215, 1980.

Kazi N.I., et al., Sorption of Graphites at High Temperatures. Report ORO-4682-4. North
Caroline State University, Raleigh, NC, 1980b.

Krohn H., Freisetzung von Spaltprodukten aus dem Core eines Kugelhaufenreaktors bei
Störfällen mit Core-Aufheizung. Report Jül-1791, Research Center Jülich, 1982.

Krohn H., and Finken R., FRESCO II – Ein Rechenprogramm zur Berechnung der
Spaltproduktfreisetzung aus kugelförmigen HTR Brennelementen in Bestrahlungs- und
Ausheizexperimenten. Report Jül-Spez-212, Research Center Jülich, 1983.

Kühnlein W., HFR-K5 und HFR-K6 Spaltproduktinventare und Abbrand. B-Z Internal Note,
Research Center Jülich, 2003.

Laurie M., et al., Results of the HFR-EU1 Fuel Irradiation of INET and AVR Pebbles in the
HFR Petten. Proc. 5th International Topical Meeting on High Temperature Reactor
Technology HTR2010, Prague, 2010.

Lohnert G., Nabielek H., Schenk W., The Fuel Element of the HTR-Module, a Prerequisite of
an Inherently Safe Reactor. Nuclear Engineering and Design 109:257-263, 1988.

Mao Y., Prediction of Metallic Fission Product Release Behaviour in Spherical Fuel Elements
Irradiated in the Experiment HFR-EU1 during Normal Irradiation and under Heating
Conditions. Studienarbeit at the Fachhochschule Jülich, 2006.

Minato K., Ogawa T., Fukuda K., Nabielek H., Sekino H., Nozawa Y., Takahashi I., Fission
Product Release from ZrC-Coated Fuel Particles during Post Irradiation Heating at 1600°C.
Journal of Nuclear Materials 224:85-92, 1995.

Minato K., et al., HRB-22 Capsule Irradiation Test for HTGR Fuel (JAERI/USDOE
Collaborative Irradiation Test. Report JAERI-Research 98-021, Japan Atomic Energy
Research Institute, Oarai, 1998.

Moormann R., Untersuchungen zum chemischen Verhalten der Spaltprodukte bei HTR-
Kernaufheizstörfällen. Proc. Reaktortagung Kerntechnik, pp. 163-166, München, 1985.

291
Moormann R., Verfondern K., Methodik umfassender probabilistischer Sicheranalysen für
zukünftige HTR-Anlagenkonzepte – Ein Statusbericht (Stand 1986), Band 3:
Spaltproduktfreisetzung. Report Jül-Spez-388_3, Research Center Jülich, 1987.

Moormann R., Source Term Estimation for Small Sized HTRs. Report Jül-2669, Research
Center Jülich, 1992.

Moormann R. AVR Experiments Related to Fission Product Transport. Proc. International


Topical Meeting on High Temperature Reactor Technology HTR2006, Johannesburg, 2006.

Müller A., Freisetzung gasförmiger Spaltprodukte (Kr, Xe, J) aus Brennelementen für
gasgekühlte Hochtemperaturreaktoren. Report Jül-1295, Research Center Jülich, 1976.

Myers B.F., Bell W.E., Cesium Transport Data for HTGR Systems. Report GA-A13990,
General Atomics, San Diego, CA, 1979.

Myers B.F., Cesium Diffusion in Silicon Carbide During Post Irradiation Anneals. Technical
Note HBK-TN-01/84, Research Center Jülich, 1984.

Nabielek H., Hick H., Wagner-Löffler M., Voice E.H., Performance Limits of Coated Particle
Fuel; Part III: Fission Product Migration in HTR Fuel. Report DP-828 (Part 3), OECD
Dragon Project, Winfrith, 1974.

Nabielek H., Brown P.E., Offermann P., Silver Release from Coated Particle Fuel. Nuclear
Technology 35:483-493, 1977.

Nabielek H., et al., Burn-Leach: The Most Important Test in the Manufacture of HTGR Fuel.
International Congress on Advances in Nuclear Power Plants ICAPP’06, Reno, NV, 2006.

Nakano M., et.al., FAPIG-HTGR Plant Concept of Low Cost and Small Power. Presentation
at the GENES4/ANP2003 Conference, Kyoto, 2003.

Nastri G.G., Rowland P.R., Shepherd J., The Prediction of Fission Product Behaviour in
HTR’s. Proc. BNES Conference on Nuclear Fuel Performance, Paper 17, London, 1973.

Nieder R., Grundzüge der HTR-Chemie. Proc. Seminar Chemie und Entsorgung, Jülich,
Germany, 1981, Report Jül-Conf-43, pp. 19-29, Research Center Jülich, 1981.

NRC, Reactor Safety Study An Assessment of Accident Risks in U.S. Commercial Nuclear
Power Plants, Report WASH-1400 (NUREG-75/014), U.S. Nuclear Regulatory Commission,
Washington D.C., 1975.

Phélip M.; et al. The CRP-6 Benchmark on HTGR Fuel Behavior under Normal Operation.
International Congress on Advances in Nuclear Power Plants ICAPP’06, Reno, NV, 2006.

Ragoß H., Description of Fuel Behavior Performance Used in the German Licensing
Procedures for the HTR-Module. Technical Note 38.07247.9, Interatom, Bensberg, 1989.

Reutler H., Lohnert G.H., Advantages of Going Modular in HTRs. Nuclear Engineering and
Design 78:129-136, 1984.

292
Sasaki T., et al., Theoretical Estimation of Radon Emanation Coefficients for UO2 Particles
Deposited on Surfaces of Uranium-bearing Wastes. Journal of Nuclear Science and
Technology 44:774-778, 2007.

Sawa K., et. al., Analytical Method of Fractional Release of Fission Products from Fuel
Elements of HTTR. Proc. IAEA Specialists' Meeting on Behaviour of GCR Fuel under
Accident Conditions, Oak Ridge TN, 1990, IWGGCR/25, pp. 55-61, IAEA, Vienna, 1991.

Sawa K., et al., Validation of Fission Product Release from Fuel Element of HTTR. Journal
of Nuclear Science and Technology 29:842-850, 1992.

Sawa K., et al., Fabrication of the First-Loading Fuel of the High Temperature Engineering
Test Reactor. Journal of Nuclear Science and Technology 36:683-690, 1999.

Schenk W., Störfällsimulation an bestrahlten Kugelbrennelementen bei Temperaturen von


1400 bis 2500°C. Report Jül-1883, Research Center Jülich, 1983.

Schenk W., Pitzer D., Nabielek H., Fission Product Release Profiles from Spherical HTR Fuel
Elements at Accident Temperatures. Report Jül-2234, Research Center Jülich, 1988.

Schenk W., Nabielek H., Kugelbrennelemente mit TRISO-Partikeln bei Störfalltemperaturen.


Report Jül-Spez-487, Research Center Jülich, 1989.

Schenk W., et al., Spaltproduktfreisetzung aus der Schwermetallkontamination bei 1000 bis
1600°C (FRJ2-KA1). Technical Note IRW-TN-92/89, Research Center Jülich, 1989b.

Schenk W., Nabielek H., Iodine Retention in Spherical Fuel Elements under HTR MODUL
Accident Conditions. Proc. Jahrestagung Kerntechnik’90, Nürnberg, Inforum GmbH, pp. 363-
366, Bonn, 1990.

Schenk W., et al., Jod und Xenonfreisetzung aus kurz bestrahlten Brennelementen bei 1000
bis 1600°C (FRJ2-KA2). Technical Note IRW-TN-9/91, Research Center Jülich, 1991b.

Schenk W., Gontard R., Nabielek H., Performance of HTR Fuel Samples under High-
Irradiation and Accident Simulation Conditions with Emphasis on Test Capsules HFR-P4 and
SL-P1. Report Jül-3373, Research Center Jülich, 1997.

Soffer L., Accident Source Terms for Light-Water Nuclear Power Plants, Final Report
NUREG-1465, U.S. Nuclear Regulatory Commission, Washington D.C., 1995.

Tang C.H., et.al., Fabrication of the First Loading Fuel of 10 MW High Temperature Gas-
Cooled Reactor. HTR-TN International HTR Fuel Seminar, Brussels, 2001.

UKAEA, Red Book. Data Collection, United Kingdom Atomic Energy Agency, Risley,
Warrington, 1979.

Van der Merwe J.J., Evaluation of Silver Transport through SiC during the German HTR Fuel
Program. Journal of Nuclear Materials 305:99-111, 2009.

Verfondern K., Müller D., Code Development for Fission Product Behavior in HTR Fuel for
Safety Assessments. Internal Report KFA-ISF-IB-10/89, Research Center Jülich, 1989.

293
Verfondern K. Müller D., Modeling of Fission Product Release Behavior from HTR Spherical
Fuel Elements Under Accident Conditions, Proc. IAEA Specialists' Meeting on the Behaviour
of GCR Fuel under Accident Conditions (Oak Ridge, 1990), IWGGCRl25, pp. 45-54,
International Atomic Energy Agency, Vienna, 1991.

Verfondern K., Dunn T.D., Bolin J.M., Comparison of US/FRG Accident Condition Models
for HTGR Fuel Failure and Radionuclide Release. Report Jül-2458, Research Center Jülich,
1991b.

Verfondern K., Possible Explanation for HFR-K3/3 IMGA Results. Report ORNL/M-2248,
Oak Ridge National Laboratory, TN, 1992.

Verfondern K., Martin R.C., Moormann R., Methods and Data for HTR Fuel Performance
and Radionuclide Release Modeling during Normal Operation and Accidents for Safety
Analysis. Report Jül-2721, Research Center Jülich, 1993.

Verfondern K., Sumita J., Ueta S., Sawa K., Modeling of Fuel Performance and Fission
Product Release Behavior during HTTR Normal Operation. Report JAERI-Research 2000-
067, Japan Atomic Energy Research Institute, Oarai, 2001c.

Verfondern K., Mikami H., Okamoto F., Prediction of Metallic Fission Product Release
Behavior in the 220 MW(th) FAPIG-HTGR during Normal Operation and Core Heatup
Accident. Proc. IAEA Technical Meeting on Current Status and Future Prospects of Gas
Cooled Reactor Fuels, Vienna, 2004.

Verfondern K., Lee Y.-W., Advances in HTGR Fuel Technology – A New Coordinated
Research Program. International Congress on Advances in Nuclear Power Plants ICAPP’05,
Paper 5050, Seoul, 2005.

Verfondern K., Nabielek H., Fission Product Release from HTR Fuel under Core Heatup
Accident Conditions. Proc. 4th International Topical Meeting on High Temperature Reactor
Technology HTR-2008, Paper 58160, Washington D.C., 2008.

Vondy D.R., Development of a General Method of Explicit Solutions to the Nuclide Chain
Equations. Report ORNL-TM-361, Oak Ridge National Laboratory, TN, 1962.

294
APPENDIX I. FRESCO-II INPUT DESCRIPTION

The input for a FRESCO-II calculation is composed of

 an input file, and

 data to be entered directly into the source code

o BLOCK DATA

o FUNCTION PBRUCH

o Plot routines

I.1. Description Input Data Set

The input file for FRESCO-II consists of a total of 27 lines to contain formatted data as
described on the following pages

295
Line Function
1.1 Fission Product Species

Column Format Variable Unit Comment


1-4 I4 NGNR - integer*4 NGNR
Group number for the identification of fission
product species considered:
1 = Cesium-137
2 = Strontium-90
3 = Silver-110m
4 = Iodine-131

296
Line Function
2.1 Headline

Column Format Variable Unit Comment


1-60 A60 TEXT(1) - character*60 TEXT(4)
TEXT(l) is a charcter string of length 60.
It appears as headline in the output file.
1st of 4 headlines.

297
Line Function
2.2 Headline

Column Format Variable Unit Comment


1-60 A60 TEXT(2) - character*60 TEXT(4)
TEXT(2) is a charcter string of length 60.
It appears as headline in the output file.
2nd of 4 headlines.

298
Line Function
2.3 Headline

Column Format Variable Unit Comment


1-60 A60 TEXT(3) - character*60 TEXT(4)
TEXT(3) is a charcter string of length 60.
It appears as headline in the output file.
3rd of 4 headlines

299
Line Function
2.4 Headline

Column Format Variable Unit Comment


1-60 A60 TEXT(4) - character*60 TEXT(4)
TEXT(4) is a charcter string of length 60.
It appears as headline in the output file.
4th of 4 headlines

300
Line Function
3.1 Time Step Control

Column Format Variable Unit Comment


1-10 E10.0 TE(1) s real*8 TE(5)
. . . TE(i) is a moment in time, which terminates
. . . the ith time interval. Until that moment, DT(i)
41-50 E10.0 TE(5) and DTOUT(i) apply (see lines 3.2 und 3.3).

Up to 5 time intervalls in increasing order can


be given. Final calculation time is given by
the last value ≠ 0.

301
Line Function
3.2 Time Step Control

Column Format Variable Unit Comment


1-10 E10.0 DT(1) s real*8 DT(5)
. . . DT(i) are time steps. Until time moment
. . . TE(i), (see line 3.1), calculations are
41-50 E10.0 DT(5) conducted using time step DT(i). Last time
step within an interval is adjusted such that
TE will be exactly reached.

Only those DT(i) will be considered, for


which a respective TE(i) > 0 exists.

302
Line Function
3.3 Time Step Control

Column Format Variable Unit Comment


1-10 E10.0 DTOUT(1) s real*8 DTOUT(5)
. . . DTOUT(i) are time steps. Until time moment
. . . TE(i), (see line 3.1), results are printed out
41-50 E10.0 DTOUT(5) using time step DTOUT(i).

Only those DTOUT(i) will be considered, for


which a respective TE(i) > 0 exists.

303
Line Function
4.1 Output Control

Column Format Variable Unit Comment


1-4 I4 IOPT(1) integer*4 IOPT(20)
. . . IOPT(i) is a 4-digit combination which
. . . controls the output. Up to 20 different output
56-60 I4 IOPT(15) options can be selected.

The last digit determines whether or not


results are printed.
0: no
1: yes

With the first 3 digits (= IOPT(i)/10), the


printout frequency can be further reduced. If
these 3 digits represents the number N, the
results of the ith option will be printed every
N*DTOUT(i) time steps.

Definition of IOPT(i):
IOPT( 1): concentration profile in particle,
graphite grain, fuel sphere
IOPT( 2): free
.
.
IOPT( 8): free
IOPT( 9): SiC layer corrosion
IOPT(10): diffusion coefficient graphite
IOPT(11): diffusion coefficient particle
IOPT(12): diffusion coefficient graphite grain
IOPT(13): fractional release from fuel sphere,
particle kernel, particle, cp failure fraction
IOPT(14): fission product inventory in
particle, graphite grain, pores, outside fuel
sphere
IOPT(15): fission product inventory in
particle kernel, coatings

304
Line Function
4.2 Output Control

Column Format Variable Unit Comment


1-4 I4 IOPT(16) integer*4 IOPT(20)
. . . See also line 4.1.
. . .
16-20 I4 IOPT(20) Definition of IOPT(i):
IOPT(16): fission product release from intact,
failed, all particles, graphite grain, fuel sphere
IOPT(17): fission product release rate from
intact, failed, all particles, graphite grain, fuel
sphere
IOPT(18): mass transfer coefficient, factor at
graphite-helium boundary, fission product
concentration in coolant
IOPT(19) fission product release rate by
recoil
IOPT(20) accumulated fission product release
by recoil

305
Line Function
4.3 Output Control

Column Format Variable Unit Comment


1-4 I4 IOPTPL No meaning
5-8 I4 IFPLOT integer*4 IFPLOT
IFPLOT is a 2-digit combination which
controls graphics output (plot).
1st digit defines time phase to be plotted
0 = normal operation/irradiation (see line 5.3)
(0,ZEIT0D)
1 = accident/heating (see line 5.3)
(ZEIT0D,ZEIT0D+ZEITPR)
nd
2 digit controls plotting
0 = no plot
1 = transient plots
2 = concentration profile plot
9-12 I4 IFBISO integer*4 IFBISO
Type of fuel particle
0 = TRISO
1 = BISO
13-16 I4 IFSICK integer*4 IFSICK
Consideration of SiC layer corrosion
0 = no
1 = yes

Remark:

IFBISO is used to jump to the correct diffusion coefficient calculation which is controlled by
the layer number, and to ignore SiC layer corrosion controlled by IFSICK.

IFSICK controls consideration (or not) of a model extension that calculates thinning of the
SiC layer by corrosion. It was introduced at a time when it was thought helpful to explain
certain experimental results. This idea, however, was abandoned later, since diffusion
coefficients derived from the experimental data, most certainly will include this effect
already.

306
Line Function
5.1 Geometry

Column Format Variable Unit Comment


1-10 E10.0 RBE(1) cm real*8 RBE(5)
. . . The fuel sphere can be subdivided into NRBE
. . . (≤ 5) zones.
41-50 E10.0 RBE(5) RBE(i) is the outer radius of the ith zone.
RBE(1) is the most inner zone. RBE(NRBE)
is the outer fuel-free zone.

NRBE is determined by the last RBE value


> 0.

Remark:

At least two zones need to be defined: the inner zone containing the fuel particles and the
outer fuel-free zone.

307
Line Function
5.2 Geometry

Column Format Variable Unit Comment


1-4 I4 NRB(1) integer*4 NRB(5)
. . . NRB(i) is the number of knots in an
. . . equidistant structure of the ith fuel sphere zone
16-20 I4 NRB(5) for the numerical solution of the diffusion
equation.
NRB(1) refers to the most inner zone.

Requirement:
NRBE
1  NRBi   200
i 1

308
Line Function
5.3 Geometry

Column Format Variable Unit Comment


1-10 E10.0 PZAHL0 real*8 PZAHL0
Number of particles in fuel sphere
11-20 E10.0 ZEIT0D d real*8 ZEIT0D
Normal operation/irradiation time
21-30 E10.0 XN0 atoms real*8 XN0
Total inventory of the fission product species
considered.
31-40 E10.0 ZEITPR h real*8 ZEITPR
ZEITPR is the time after begin of the
accident/heating phase, for which the
following information will be given:
- number of intact and failed particles and
fission product release in the single particle
failure phases (≤ 10) (see chapter xx on
particle failure function);
- concentration profiles in particle and fuel
sphere
- so-called „Matrix Release Fraction“
41-50 E10.0 GAMMA 1025 m-2, real*8 GAMMA
E>0.1MeV Fast neutron fluence at begin of the
accident/heating phase.

Remark:

XN0 is relevant only for the metallic fission products (Cs, Sr) in case the non-linear range of
the sorption isotherms (Freundlich regime) is touched. XN0 should bet he sum of the
inventories of all (longer-lived) isotopes of the species considered plus those of chemically
similar species: Cs+ Rb or Sr+ Ba.

ZEITPR is typically the end of a heating test to allow comparison with experimental data such
as IMGA measurements or concentration profiles measured. The “Matrix Release Fraction”
or MRF factor is the ratio of fuel element to particle release.

GAMMA is required in a cesium calculation for the determination of the fluence-dependent


diffusion coefficient in SiC during normal operation/irradiation.

309
Line Function
5.4 Geometry

Column Format Variable Unit Comment


1-10 E10.0 RCP(1) cm real*8 RCP(5)
. . . The fuel particle can be subdivided into NRP
. . . (≤ 5) zones (kernel, layers).
. . . RCP(i) is the outer radius of the ith zone.
41-50 E10.0 RCP(5) RCP(1) is the kernel radius.

NRP is determined by the last RCP value > 0.

Remark:

310
Line Function
5.5 Geometry

Column Format Variable Unit Comment


1-10 E10.0 RKORN cm real*8 RKORN
Outer radius of graphite grain

311
Line Function
5.6 Geometry

Column Format Variable Unit Comment


1-4 I4 NRC(1) integer*4 NRC(5)
. . . NRC(i) is the number of knots in an
. . . equidistant structure of the ith fuel particle
16-20 I4 NRC(5) zone for the numerical solution of the
diffusion equation.
NRC(1) refers to the kernel zone.

Requirement:
NRP
1   NRC i   200
i 1

312
Line Function
5.7 Geometry

Column Format Variable Unit Comment


1-4 I4 NK0 integer*4 NK0
NK0 is the number of knots in an equidistant
structure of the graphite grain for the
numerical solution of the diffusion equation.

Requirement:
1  NK 0  200
5-8 I4 IFBE integer*4 NK0
IFBE controls selection of concentration-
dependent diffusion coefficient in the fuel
element matrix graphite
0 = no
1 = yes

Remark:

The concentration dependence, well known for the sorption effect, will be applied to the
diffusion coefficient of cesium or strontium in the fuel element matrix graphite. If fission
product concentration in the graphite is greater than the boundary concentration (CGRENC,
see line xx), the diffusion coefficient in matrix is calculated according to

 M , Freundlich
DM  DM
 M , Henry

313
Line Function
6.1 Heavy Metal Contamination

Column Format Variable Unit Comment


1-10 E10.0 UKONGK real*8 UKONGK
Fraction of heavy metal contamination in
graphite grain
11-20 E10.0 UKONGP real*8 UKONGP
Fraction of heavy metal contamination in
graphite pores
21-30 E10.0 UKONTP(2) real*8 UKONTP(5)
. . . Fraction of heavy metal contamination in
. . . the ith particle zone (only layers, i ≥ 2).
51-60 E10.0 UKONTP(5)

Remark:

UKONTP(1), which is the heavy metal content of the particle kernel, is then determined by

 NRP

UKONTP1  1.  UKONGK  UKONGP   UKONTPi 
 i 1 

314
Line Function
7.1 Transport Data

Column Format Variable Unit Comment


2
1-10 E10.0 D0G cm /s real*8 D0G
Frequency factor for the diffusion coefficient
in the graphite pores
11-20 E10.0 AKG J/mol real*8 AKG
Activation energy for the diffusion coefficient
in the graphite pores
21-30 E10.0 F0G real*8 F0G
Factor to modify the diffusion coefficient in
the graphite pores
Default = 1.
31-40 E10.0 F0PK real*8 F0PK
Factor to modify the diffusion coefficient in
the particle kernel
Default = 1.
41-50 E10.0 F0PSIC real*8 F0PSIC
Factor to modify the diffusion coefficient in
silican carbide
Default = 1.
51-60 E10.0 F0DPK real*8 F0DPK
Factor to modify the diffusion coefficient in
the particle kernel of failed particles
Default = 1.

Remark:

The factors to modify diffusion coefficients are typically used to adjust calculation results to
measurements in heating experiments.

315
Line Function
7.2 Transport Data

Column Format Variable Unit Comment


2
1-10 E10.0 D0K cm /s real*8 D0K
Frequency factor for the diffusion coefficient
in the graphite grain
11-20 E10.0 AKK J/mol real*8 AKK
Activation energy for the diffusion coefficient
in the graphite grain
21-30 E10.0 F0GK real*8 F0GK
Factor to modify the diffusion coefficient in
the graphite grain
Default = 1.

316
Line Function
7.3 Transport Data

Column Format Variable Unit Comment


2
1-10 E10.0 D0P(1) cm /s real*8 D0P(5)
. . . D0P(i) is the frequency factor for the
. . . diffusion coefficient in the ith zone of the fuel
41-50 E10.0 D0P(5) particle
D0P(1) corresponds to particle kernel.

317
Line Function
7-4 Transport Data

Column Format Variable Unit Comment


1-10 E10.0 AKP(1) J/mol real*8 AKP(5)
. . . AKP(i) is the activation energy for the
. . . diffusion coefficient in the ith zone of the fuel
41-50 E10.0 AKP(5) particle
AKP(1) corresponds to particle kernel.

Remark:

The actual input structure for diffusion coefficients is only appropriate for simple effective
diffusion coefficients and does not allow more complex structures like those consisting of the
sum of a low-temperature and high-temperature branch or those which are distinguished
between normal operation/irradiation and accident/heating phase. In these cases, diffusion
coefficients are implemented directly in the code, i.e., in the FUNCTION DIFPA. These more
complex diffusion coefficients are selected by setting the respective values for D0P(i) and
AKP(i) zero (both!). Directly coded diffusion coefficients are currently given for:

 Diffusion coefficients with low-temperature and hogh-temperatur branch:

o Cesium in UO2 kernel (NGNR = 1, i = 1)

o Iodine in UO2 kernel (NGNR = 4, i = 1)

o Cesium in silicon carbide (NGNR = 1, i = 4)

o Strontium in silicon carbide (NGNR = 2, i = 4)

 Diffusion coefficients distinguished by normal operation/irradiation and


accident/heating:

o Strontium in (Th,U)O2 kernel (NGNR = 2, I = 1)

o Cesium in silicon carbide (NGNR = 1, i = 4)

o Silver (NGNR = 3, i = 4)

The diffusion coefficients for particle kernel and silicon carbide can be modified by the
factors F0PK und F0PSIC (see line 7.1).

For PyC layers, a rapid diffusion is assumed at temperatures beyond 2000°C.

318
The FUNCTION DIFPAD refers to the diffusion coefficient in the particle kernel
representing a defective/failed particle. The FUNCTION contains that part of the FUNCTION
DIFPA that applies to the particle kernel. The diffusion coefficient for the defective/failed
particle can be modified with a separate factor, F0DPK, which, e.g., allows the simulation of a
defective, but still partially existing coating, i.e., a particle between an intact particle and a
bare kernel.

319
Line Function
8.1 Recoil

Column Format Variable Unit Comment


1-10 E10.0 RECKER cm real*8 RECKER
Recoil distance in particle kernel
11-20 E10.0 RECPYC cm real*8 RECPYC
Recoil distance in PyC layer
21-30 E10.0 RECGRA cm real*8 RECGRA
Recoil distance in outer zone of fuel sphere

Remark:

Recoil distances are usually not considered in the calculations if assuming that the recoil
effect is already included in the diffusion coefficients derived from experiments.

320
Line Function
9.1 Sorption Isotherms

Column Format Variable Unit Comment


1-10 E10.0 ACI real*8 ACI
Coefficient for the linear (Henry) regime of
the sorption isotherm for matrix graphite
11-20 E10.0 BCI real*8 BCI
Coefficient for the linear (Henry) regime of
the sorption isotherm for matrix graphite
21-30 E10.0 CGRENC mmol/kg real*8 CGRENC
graphite Boundary concentration between linear
(Henry) and non-linear (Freundlich) regime
31-40 E10.0 IFADC integer*4 IFADC
Selection of whether or not the sorption effect
is considered
0 = sorption is considered
-1 = sorption is ignored
41-50 E10.0 TGRENC °C real*8 TGRENC
Upper temperature limit, beyond which the
sorption effect is no longer considered.
Default: 4000°C, i.e., sorption considered at
all temperatures.

321
Line Function
9.2 Sorption Isotherms

Column Format Variable Unit Comment


1-10 E10.0 ANCI real*8 ANCI
Coefficient for the non-linear (Freundlich)
regime of the sorption isotherm for matrix
graphite
11-20 E10.0 BNCI real*8 BNCI
Coefficient for the non-linear (Freundlich)
regime of the sorption isotherm for matrix
graphite
21-30 E10.0 ENCI real*8 ENCI
Coefficient for the non-linear (Freundlich)
regime of the sorption isotherm for matrix
graphite
41-50 E10.0 FNCI real*8 FNCI
Coefficient for the non-linear (Freundlich)
regime of the sorption isotherm for matrix
graphite

Remark:

With the coefficients given in lines 9.1 and 9.2, the partition coefficient αM of metallic fission
products between coolant and matrix graphite is determined according to

1  BCI 
 M , Henry  exp  ACI   for cMatrix ≤ CGRENC
T  T 

1  BNCI  FNCI  
 M , Freundlich  exp  ANCI    ENCI    ln c Matrix 
T  T  T  

If the option “sorption effect ignored” is selected (IFADC = -1), the partition coefficient is set
to

M  1

322
Line Function
10.1 Nuclear Data

Column Format Variable Unit Comment


1-10 E10.0 AINV Initial fission product inventory in fuel sphere
Default: 1.0×10-6
11-20 E10.0 ZERFK s-1 Decay constant of radionuclide considered
Default: If ZERFK = 0., then
ZERFK = 0. for NGNR = 0
ZERFK = 7.290×10-10 for NGNR = 1 (137Cs)
ZERFK = 7.712×10-10 for NGNR = 2 (90Sr)
ZERFK = 3.210×10-8 for NGNR = 3 (110mAg)
ZERFK = 1.000×10-6 for NGNR = 4 (131I)

323
Before starting a FRESCO-II run, further input data need to be introduced in a few
subroutines. These data, which are not part of the input file, are related to the temperature-
time histories for normal operation and accident phase (BLOCK DATA), and to the particle
failure function (FUNCTION PBRUCH). In an earlier version, the input of experimental data
for comparison purposes in the respective plot subroutines was also possible. But most
probably, this plotting part of the code is usually outsourced and done separately by modern
software. The current, still not optimized procedure at FZJ is to copy/paste the output parts to
be considered into an excel-sheet. Further interference into the source code is possible and
necessary, e.g., if two-branch diffusion coefficients, other than those offered, were to be
applied.

I.2. BLOCK DATA

The temperature-time histories for the irradiation/normal operation phase and the
heating/accident phase are defined in separate DATA statements where either one consists of
three strings:

NI0 to define how many pairs of time/temperature numbers are to be expected,


maximum: 500;
AZEIT0 [days] the moments in time, for which a temperature will be given;
ATEMP0 [°C] the irradiation temperatures for the given times.

The same structure applies to the accident phase:

NI to define how many pairs of time/temperature numbers are to be expected,


maximum: 500;
AZEIT [hours] the moments in time, for which a temperature will be given;
ATEMP [°C] the heating temperatures for the given times.

Final statement in BLOCK DATA contains the definition of the unit number for output
(NOUT) and input (NIN). The other two unit number variables are no longer of interest.

An example for a BLOCK DATA is given in chapter I.5.2.

I.3. FUNCTION PBRUCH

The fraction of defective/failed particles is directly allocated to the variable PBRUCH as a


function of time. The failure fraction is defined as a step function with a maximum of 10
possible steps for the whole investigation period.

Is one of the given time limits exceeded, the next, higher failure fraction value will be
allocated. For the additionally failed particles, which are now regarded as a bare kernel only, a
new, independent diffusion calculation will begin from this moment resulting in its own
(partial) source term (to the matrix graphite). In the calling subroutine PADIFF, a respective
counter index is raised by one, in order to identify the corresponding concentration arrays.
The strong limitation in storage capacity in former times required a limitation in the number
of steps.

324
An example for a FUNCTION PBRUCH is given in chapter I.5.3.

I.4. Plot Routines

The plot routines as were created in the 1980s and implemented in the codes, are usually
ignored in today’s FRESCO calculations. The user is kindly asked to handle the output file
according to modernized plot software and his/her computer system.

I.5. Example

I.5.1. Input File

Line Input
1 1

2 CESIUM RELEASE FROM SPHERICAL FUEL ELEMENT

3 DURING NORMAL OPERATION AND CORE HEATUP ACCIDENT

4
5
6 121098240.121746240.

7 3.6 E+4 3.6 E+2

8 3.6 E+5 3.6 E+3

9 200 0 1 1 1

10 1 1

11 0 01 0 0

12 2.5 3.0

13 9 20

14 16350.0 1401.6 180. 3.0

15 250. E-4 345. E-4 385. E-4 420. E-4 460.0 E-4

16 5. E-4

17 39 39 39 39 39

18 5

19 1.0 E-10 1.0 E-07 1.0 E-03 1.0 E-04 1.0 E-06 1.0 E-06

20 3.6 E-00 189.E+03

21 1.0 E+00 0.

22 0. E-05 1. E-08 6.3 E-04 0. 6.3 E-04

23 0.E+03 0. 222. E+03 0.E+03 222.E+03

24 0.0 E-04 0.0 E-04 0.0 E-04

325
25 21.94 -44093. 1.13 -1

26 22.11 -44543. -1.3689 3683.

27 0. 0.

I.5.2. BLOCK DATA


C---------------
BLOCK DATA
C---------------
C
IMPLICIT REAL*8 (A-H,O-Z)
C
COMMON /ITEMP/ NI0,NI
COMMON /TTEMP/ AZEIT0(500),ATEMP0(500),AZEIT(100),ATEMP(100)
COMMON /RIO/ NOUT,NIN,NFRK,NFRKM
C=======================================================================
C***** IRRADIATION TEMPERATURE/TIME HISTORY
C***** CASE: EXAMPLE
C***** 12 PASSAGES THROUGH CORE
C
DATA NI0 /204/
C***** IRRADIATION TIMES IN DAYS
DATA AZEIT0 / 0.0, 7.3, 14.6, 21.9, 29.2,
$ 36.5, 43.8, 51.1, 58.4, 65.7,
$ 73.0, 80.3, 87.6, 94.9, 102.2,
$ 109.5, 116.8,
$ 116.9, 124.1, 131.4, 138.7, 146.0,
$ 153.3, 160.6, 167.9, 175.2, 182.5,
$ 189.8, 197.1, 204.4, 211.7, 219.0,
$ 226.3, 233.6,
$ 233.7, 240.9, 248.2, 255.5, 262.8,
$ 270.1, 277.4, 284.7, 292.0, 299.3,
$ 306.6, 313.9, 321.2, 328.5, 335.8,
$ 343.1, 350.4,
$ 350.5, 357.7, 365.0, 372.3, 379.6,
$ 386.9, 394.2, 401.5, 408.8, 416.1,
$ 423.4, 430.7, 438.0, 445.3, 452.6,
$ 459.9, 467.2,
$ 467.3, 474.5, 481.8, 489.1, 496.4,
$ 503.7, 511.0, 518.3, 525.6, 532.9,
$ 540.2, 547.5, 554.8, 562.1, 569.4,
$ 576.7, 584.0,
$ 584.1, 591.3, 598.6, 605.9, 613.2,
$ 620.5, 627.8, 635.1, 642.4, 649.7,
$ 657.0, 664.3, 671.6, 678.9, 686.2,
$ 693.5, 700.8,
$ 700.9, 708.1, 715.4, 722.7, 730.0,
$ 737.3, 744.6, 751.9, 759.2, 766.5,
$ 773.8, 781.1, 788.4, 795.7, 803.0,
$ 810.3, 817.6,
$ 817.7, 824.9, 832.2, 839.5, 846.8,
$ 854.1, 861.4, 868.7, 876.0, 883.3,
$ 890.6, 897.9, 905.2, 912.5, 919.8,
$ 927.1, 934.4,
$ 934.5, 941.7, 949.0, 956.3, 963.6,
$ 970.9, 978.2, 985.5, 992.8, 1000.1,
$ 1007.4, 1014.7, 1022.0, 1029.3, 1036.6,
$ 1043.9, 1051.2,

326
$ 1051.3, 1058.5, 1065.8, 1073.1, 1080.4,
$ 1087.7, 1095.0, 1102.3, 1109.6, 1116.9,
$ 1124.2, 1131.5, 1138.8, 1146.1, 1153.4,
$ 1160.7, 1168.0,
$ 1168.1, 1175.3, 1182.6, 1189.9, 1197.2,
$ 1204.5, 1211.8, 1219.1, 1226.4, 1233.7,
$ 1241.0, 1248.3, 1255.6, 1262.9, 1270.2,
$ 1277.5, 1284.8,
$ 1284.9, 1292.1, 1299.4, 1306.7, 1314.0,
$ 1321.3, 1328.6, 1335.9, 1343.2, 1350.5,
$ 1357.8, 1365.1, 1372.4, 1379.7, 1387.0,
$ 1394.3, 1401.6,
$ 296*0./
C***** IRRADIATION TEMPERATURES IN CENTIGRADES
DATA ATEMP0 / 550., 565., 606., 656., 694.,
$ 736., 764., 790., 810., 825.,
$ 838., 847., 855., 860., 865.,
$ 867., 873.,
$ 550., 565., 606., 656., 694.,
$ 736., 764., 790., 810., 825.,
$ 838., 847., 855., 860., 865.,
$ 867., 873.,
$ 550., 565., 606., 656., 694.,
$ 736., 764., 790., 810., 825.,
$ 838., 847., 855., 860., 865.,
$ 867., 873.,
$ 550., 565., 606., 656., 694.,
$ 736., 764., 790., 810., 825.,
$ 838., 847., 855., 860., 865.,
$ 867., 873.,
$ 550., 565., 606., 656., 694.,
$ 736., 764., 790., 810., 825.,
$ 838., 847., 855., 860., 865.,
$ 867., 873.,
$ 550., 565., 606., 656., 694.,
$ 736., 764., 790., 810., 825.,
$ 838., 847., 855., 860., 865.,
$ 867., 873.,
$ 550., 565., 606., 656., 694.,
$ 736., 764., 790., 810., 825.,
$ 838., 847., 855., 860., 865.,
$ 867., 873.,
$ 550., 565., 606., 656., 694.,
$ 736., 764., 790., 810., 825.,
$ 838., 847., 855., 860., 865.,
$ 867., 873.,
$ 550., 565., 606., 656., 694.,
$ 736., 764., 790., 810., 825.,
$ 838., 847., 855., 860., 865.,
$ 867., 873.,
$ 550., 565., 606., 656., 694.,
$ 736., 764., 790., 810., 825.,
$ 838., 847., 855., 860., 865.,
$ 867., 873.,
$ 550., 565., 606., 656., 694.,
$ 736., 764., 790., 810., 825.,
$ 838., 847., 855., 860., 865.,
$ 867., 873.,
$ 550., 565., 606., 656., 694.,
$ 736., 764., 790., 810., 825.,
$ 838., 847., 855., 860., 865.,
$ 867., 873.,

327
$ 296*0./
C-----------------------------------------------------------------------
C***** ACCIDENT TEMPERATURE/TIME HISTORY
C***** CASE: EXAMPLE
C-----------------------------------------------------------------------
DATA NI /12/
C***** ACCIDENT TIMES IN HOURS
DATA AZEIT/ 0.0,0.0271,0.2208, 1.0, 10.0,
$ 20.0, 30.0, 35.0, 60.0, 90.0,
$ 120.0, 180.0,
$ 88*0./
C***** ACCIDENT TEMPERATURES IN CENTIGRADES
DATA ATEMP/ 760., 760., 795., 887., 1298.,
$ 1455., 1479., 1476., 1417., 1332.,
$ 1253., 1122.,
$ 88*0./
C=======================================================================
C***** NIN = FRESCO-II INPUT NOUT = FRESCO-II OUTPUT (PRINTER)
C***** NFRK = FRESCO-II RESULTS (GDDM) NFRKM = EXPERIMENTAL DATA (GDDM)
DATA NOUT /60/, NIN /50/, NFRK /7/, NFRKM /77/
C=======================================================================
C
END

I.5.3. FUNCTION PBRUCH


REAL FUNCTION PBRUCH*8(ZEIT,TEMPER)
C-------------------------------
C***** PARTICLE FAILURE FUNCTION
C***** EXPECTED FAILURE FRACTION DURING NORMAL OPERATION
C-------------------------------
IMPLICIT REAL*8 (A-H,O-Z)
COMMON /QUELL/ ZEIT0,QGESAM,QAKT,Q1
C
ZEIT1H = ZEIT/3600.
ZEIT1D = ZEIT1H/24.
PBRUCH = 3.0D-05
IF (ZEIT1D .LT. 116.8) GO TO 100
PBRUCH = 3.4D-05
IF (ZEIT1D .LT. 408.8) GO TO 100
PBRUCH = 3.8D-05
IF (ZEIT1D .LT. 700.8) GO TO 100
PBRUCH = 4.2D-05
IF (ZEIT1D .LT. 992.8) GO TO 100
PBRUCH = 4.6D-05
IF (ZEIT1D .LT.1284.8) GO TO 100
PBRUCH = 5.0D-05
IF (ZEIT1H .LT. 33650.) GO TO 100
PBRUCH = 5.0D-05 + 2.1D-06
IF (ZEIT1H .LT. 33667.) GO TO 100
PBRUCH = 5.0D-05 + 4.2D-06
IF (ZEIT1H .LT. 33678.) GO TO 100
PBRUCH = 5.0D-05 + 6.3D-06
IF (ZEIT1H .LT. 33751.) GO TO 100
PBRUCH = 5.0D-05 + 8.4D-06
100 CONTINUE
RETURN
END

328

Вам также может понравиться