li THERMAL HYDRAULICS OF CORE/CONCRETE INTERACTION by

advertisement
THERMAL HYDRAULICS OF CORE/CONCRETE INTERACTION
IN SEVERE LWR ACCIDENTS
by
L.S. Kao and M.S. Kazimi
li
A
i5l
MITNE-276
Department of Nuclear Engineering
Massachusetts Institute of Technology
June 1987
THERMAL HYDRAULICS OF CORE/CONCRETE INTERACTION
IN SEVERE LWR ACCIDENTS
by
L.S. Kao and M.S. Kazimi
Work Supported by
Electric Power Research Institute
Project Manager: Dr. B.R. Sehgal
THERMAL HYDRAULICS OF CORE/CONCRETE INTERACTION
IN SEVERE LWR ACCIDENTS
ABSTRACT
Several physical processes, including melt freezing, liquid/liquid interfacial heat
transfer, layer mixing, and droplet entrainment, involved in the analysis of the
core/concrete interaction were investigated by scoping simulant experiments. Water and cyclohexane were used to form a multi-layer pool in a test unit designed
with cooling capability and air injection through a porous plate. In the freezing
tests with gas velocities in the range of 5 to 126 mm/s, a stable solidified layer
was formed across the bubble agitated horizontal liquid/solid interface, while no
boundary crust was found at either the top surface or liquid/liquid interface. The
interfacial heat transfer between the water and cyclohexane layers was measured
under different air injection rates. The experimental data showed good agreement
with both the modified Szekely model and the Lee and Kazimi model. In the
layer mixing tests, it was found that the water and cyclohexane with density ratio
of 0.78 were entirely mixed under a modest superficial gas velocity of 50 mm/s.
The amount of liquid droplet entrainment measured in the simulant experiments
agreed qualitatively with the Kataoka and Ishii model.
The heat transfer from the corium to concrete will affect the cooling rate of the
corium and the amount of gas generated by concrete decomposition, which will,
in turn, affect the pressurization rate of the containment building and the degree
to which fission products could be released from the melt. A semiempirical correlation was developed to describe the heat transfer at the horizontal core/concrete
interface. The model assumes periodic contact between corium and concrete at
low gas evolution rates, and separation by a stable gas film at high gas generation rates. The proposed model has been incorporated into an existing computer
code, CORCON/MOD2, for integral analysis of the corium/concrete interactions.
Good agreement was found when using this model to analyze the German BETA
experiments involving several hundred kg oxidic and metallic melt with sustained
internal heating by induction. The proposed model is capable of producing erosion
results of the BETA experiments with a mean error of 5% and a standard deviation
of 27%. Less accuracy was found in the calculation of transient, one-dimensional
experiment results obtained at the Sandia National Laboratory.
The impact of the downward heat. transfer model on the calculations of concrete
erosion, gas generation, and ex-vessel aerosol release was studied by using COR-
CON/MIT (revised version of CORCON/MOD2) and VANESA. It was found that
with high initial melt temperatures, the fission product release calculated by the
periodic contact model was three to five times higher than the original gas film
model used in the CORCON code. At low initial melt temperatures, with the
ii
formation of an initial bottom crust, the various heat transfer models did not lead
to significant differences in the fission product release. Sensitivity studies involving variations in several parameters, such as initial melt temperature, concrete
properties, amount of unoxidized zirconium, amount of melt, decay heat, and layering potential of melt constituents, were also performed to identify the important
sources of uncertainties in calculation of the ex-vessel aerosol release. The initial
debris temperature was found to be the most significant parameter in the calculation. The release fraction of the non-volatile fission products such as lanthanum
was the most sensitive result of the parameter variations.
iii
ACKNOWLEDGMENTS
This report is based on the thesis submitted by the first author to M.I.T. in
fulfillment of the requirements for the degree of Doctor of Philosophy in Nuclear
Engineering.
The authors would like to thank Professor J. Meyer for the time he devoted to
comment on this work. Dr. B.R. Sehgal of EPRI is especially appreciated for the
continuous guidance he provided. This work would not have been possible without
the financial support of EPRI.
The first author would also like to extend special acknowledgment to his wife,
Ann-Tinn Shen, for the support and encouragement she provided.
iv
TABLE OF CONTENTS
Page
ABSTRACT
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii
Iv
. . .. .. . ....
ACKNOWLEDGMENTS
.. . .. .......
..........
. .. V
..........
TABLE OF CONTENTS
. . . .. . . . . .. . . . .. .. . ...
ix
. . .. .. .....
. . . .........
LIST OF FIGURES ....
..........
xviii
.. ........
LIST OF TABLES ......
.....................
Chapter 1
INTRODUCTION AND BACKGROUND
1.1 Introduction ...........................
1.1.1 The Molten Core/Concrete Interaction
1.1.1.1 Severe Accident Sequence.....
1.1.1.2 Physical Phenomena of MCCI
1.1.1.3 Consequences of MCCI .......
1.1.2 Scope of This Work .................
1.2 Background
...........................
1.2.1 G eneral ...........................
1.2.2 Melt/Concrete Interaction Experiments
1.2.2.1 Simulant Experiments
.......
1.2.2.2 Real Material Experiments
1.2.3 Analytical Modeling of MCCI
1
2
2
4
5
6
6
8
8
.. .
....
.......
1.2.4 Ex-Vessel Source Term Assessment ...
1.2.4.1 Integrated Approach
.........
1.2.4.2 Estimates of Uncertainties
1.3 Structure of This Work
Chapter 2
...
.................
10
....
13
....
....
....
....
22
24
28
29
SIMULANT EXPERIMENTS
2.1 Objective
......................
2.2 Introduction ....................
2.2.1 Freezing Phenomena ........
2.2.2 Interfacial Heat Transfer
.....................
.....................
.....................
.....................
.....................
.....................
.....................
.....................
.....................
.....................
.....................
.....................
.....................
....
..............
2.2.3 Layer Mixing
2.2.4 Droplet Entrainment ........
2.3 Experiment Descriptions ..........
2.3.1 General Features ............
2.3.2 Apparatus ..................
2.3.3 Simulant Materials ..........
2.3.4 Test Procedures ............
2.4 Experimental Results ............
2.4.1 Freezing Phenomena ........
v
31
31
31
32
37
42
45
45
45
49
49
52
52
TABLE OF CONTENTS (Continued)
Page
2.4.2 Interfacial Heat Transfer
2.4.3 Layer M ixing
2.4.4 Droplet Entrainment
2.5 Summary and Conclusions
Chapter 3
.....................
56
61
63
63
...............................
.........................
.........................
DOWNWARD HEAT TRANSFER MODEL FOR THE
MELT/CONCRETE INTERACTION
68
71
71
76
82
85
85
. 86
. 88
. 88
. 90
. 92
3.3.2 Transition Criteria of the Film Collapse Model ....
. 95
3.3.3 Post-Freezing Heat Transfer Model ..............
. 97
............
3.3.4 Sum m ary ....................................
. 97
.
.
.
.
.
.
.
.
..
.
.
3.4 Downward Heat Flux Calculation ....................
. 97
3.4.1 Cases Studied ............................................
3.4.1.1 Molten Core Configuration in Concrete Cavity .......... .104
.105
....................
3.4.1.2 Concrete Type of Reactor Cavity
.105
3.4.1.3 Solidus Temperature of Molten Core ..................
.106
3.4.2 Results and Discussions ....................................
.106
3.4.2.1 Downward Heat Transfer Model ....................
.110
3.4.2.2 M olten M aterial ....................................
.117
3.4.2.3 Concrete Type .....................................
.119
..............
3.4.2.4 Solidus Temperature of Molten Material
. . .. .. . . . . . . .. . . . . . . . . . . .. . . . . . . . . .. . . . . . . . . . . . . . . . . .124
3.5 Summary
3.1 Introduction ......................................
3.2 Review of the Downward Heat Transfer Models ........
3.2.1 The Gas Film Model ..........................
3.2.2 The Periodic Contact Model ....................
3.2.3 The Film Collapse Model ......................
3.3 Model Development and Implementation in CORCON ..
3.3.1 Revised Periodic Contact Model ................
3.3.3.1 Basic Definition ........................
3.3.1.2 Transient Heat Conduction ..............
3.3.1.3 Bubble Dynamics ......................
3.3.1.4 Interface Temperature ..................
Chapter 4
.
.
.
.
.
.
HEAT TRANSFER MODEL VALIDATION BY
COMPARISON TO INTEGRAL EXPERIMENTS
4.1 Introduction ....................................................
4.2 Review of Real Material Experiments ..............................
vi
125
126
TABLE OF CONTENTS (Continued)
Page
4.2-1 BETA Experim ents ..........................................
4.2.1.1 Descriptions of BETA Facility ..........................
4.2.1.2 Experimental Results and Discussions
..................
4.2.2 Sandia Experim ents
........................................
4.2.2.1 Descriptions of Sandia Experiments
....................
4.2.2.2 Results and Conclusions
..............................
4.2.3 Summary ................................................
4.3 Experiment Analysis and Model Validation ..........................
4.3.1 Experiments Analyzed and Input Parameters Used ..............
4.3.2 Results and Discussions ......................................
4.3.2.1 BETA Experiments
..................................
4.3.2.2 Sandia Experim ents ..................................
4.3.3 Model Validation ...........................................
4.4 C onclusions
....................................................
Chapter 5
126
126
130
137
137
142
145
145
145
150
150
179
187
194
SENSITIVITY STUDY OF THE EX-VESSEL SOURCE
TERM
5.1 Objective
....................................................
195
5.2 Introduction ....................................................
196
5.2.1 Characteristics of Ex-Vessel Source Term .....................
196
5.2.2 Background ..............................................
197
5.3 Formulation of This Study .......................................
200
5.3.1 Analysis Methods and Computer Codes ......................
200
5.3.2 Input Param eters
..........................................
201
5.3.2.1 Specifications of the Base Case
........................
201
5.3.2.2 Cases Analyzed ......................................
204
5.4 Results and Discussions ..........................................
204
5.4.1 Impact of the Downward Heat Transfer Models
.......... ..... 207
5.4.2 Effect of Concrete Decomposition Temperature ...........
.. 221
5.4.3 Effect of Concrete Type ...................................
226
5.4.4 Effect of Zirconium M etal ....................................
231
5.4.5 Effect of Initial Debris Temperature
.........................
.243
5.4.6 Effect of Amount of M elts ....................................
262
5.4.7 Effect of Ferrous Oxide ......................................
262
5.4.8 Effect of Decay Heat ........................................
269
5.4.9 Effect of Layer Configuration
................................
281
5.4.10 Effect of CORCON/MOD2 Version 2.01 ......................
288
5.4.11 Revised Periodic Contact Model .............................
288
5.4.12 Sum m ary
................................................
303
5.5 Conclusions
.............
...................................
311
Vii
TABLE OF CONTENTS (Continued)
Page
Chapter 6
SUMMARY AND CONCLUSIONS
6.1 Summary of This Work
.............................
6.1.1 Experimental Observations
.....................
6.1.2 Development and Validation of Heat Transfer Models
6.1.3 Impact of Heat Transfer Models
.................
6.1.4 Sensitivity Study on Ex-Vessel Aerosol Release .....
............
.............
...........
............
............
REFEREN CES .......................................
.............
VIll
. 313
313
. 315
. 318
. 319
321
LIST OF FIGURES
Page
Figure
17
..
..............
1.1
Schematic Diagram of CORCON System (Ref.[C3])
1.2
BM I-2104 Codes Suite (Ref.[S2])
1.3
Source Term Code Package (Ref.[S2])
2.1
Interfacial Heat Transfer Coefficient of the Metallic/Oxidic
Corium Pool Predicted by Different Models ....................
38
Interfacial Heat Transfer Coefficient of the Metallic/Oxidic
Corium Pool Based on Different Temperature Differences ........
39
2.2
2.3
2.4
26
..............................
27
..........................
Interfacial Heat Transfer Coefficient of the Metallic/Oxidic
Corium Pool Based on Different Bubble Diameters ..............
.40
Interfacial Heat Transfer Coefficient of the Metallic/Oxidic
Corium Pool Based on Different Layer Configurations ............
.41
46
....
2.5
Schematic Diagram of the Simulant Experimental Apparatus
2.6
Illustration of the Temperature Measurement Locations
2.7
Water Pool Temperature Histories
.
55
2.8
Interfacial Heat Transfer between Water and Cyclohexane Layers . .
60
2.9
Water Droplet Entrainment from the Bubbling Pool
3.1
Illustration of the Downward Heat Transfer of the Core/Concrete
................................................
Interaction
.48
..........
............................
66
............
70
.72
......................
3.2
Analytical Picture of the Gas Film Model
3.3
Analytical Picture of the Periodic Contact Model ................
3.4
Downward Ablation Distance of BETA Test VO.2
. . ....
...... . .
83
3.5
Analytical Picture of the Revised Periodic Contact Model
........ .
87
3.6
Downward Ablation Distances of Early BETA Tests
3.7
Descriptive Downward Heat Flux of the Film Collapse Model
3.8
Comparison between the Predicted and Measured Downward
.....................
Erosion Distances of BETA Test V0.2.
3.9
.79
..........
Comparison between the Predicted and Measured Downward
Erosion Distances of BETA Test VO.3 ..........................
ix
. . .94
99
....
..
100
101
LIST OF FIGURES (Continued)
Page
Figure
3.10
3.11
3.12
Comparison between the Predicted and Measured Downward
Erosion Distances of BETA Test V1.2 ..........................
102
Comparison between the Predicted and Measured Downward
Erosion Distances of BETA Test V1.3 ..........................
103
Downward Heat Fluxes of the Various Models for Core Oxide
Interacting with Limestone/Common Sand Concrete
3.13
3.14
3.15
3.16
3.17
3.18
3.19
3.20
3.21
3.22
............
109
Downward Heat Fluxes of the Various Models for Core Oxide
Interacting with Limestone Concrete ..........................
111
Downward Heat Fluxes of the Various Models for Core Oxide
Interacting with Basaltic Concrete ............................
112
Downward Heat Fluxes of the Various Models for Core Oxide
Interacting with KfK Concrete ................................
113
Downward Heat Fluxes of the Gas Film Model for Various Melts
Interacting with Limestone/Common Sand Concrete ............
114
Downward Heat Fluxes of the Periodic Contact Model for Various
......
Melts Interacting with Limestone/Common Sand Concrete
115
Downward Heat Fluxes of the Revised Periodic Contact Model for
Various Melts Interacting with Limestone/CS Concrete ..........
116
Downward Heat Fluxes of the Gas Film Model for Steel Melt
Interacting with Various Concretes ............................
120
Downward Heat Fluxes of the Periodic Contact Model for Steel
Melt Interacting with Various Concretes ........................
121
Downward Heat Fluxes of the Revised Periodic Contact Model for
Steel Melt Interacting with Various Concretes ..................
122
Downward Heat Fluxes of the Revised Periodic Contact Model
for Core Oxide Interacting with Various Concretes ..............
123
............
127
4.1
Schematic Diagram of BETA Experiment (Ref.[A5])
4.2
Dimensions of Concrete Cavity of BETA Experiment (Ref.[A5])
4.3
Measured Concrete Erosion Distances of BETA Test V1.8
........
132
4.4
Measured Concrete Erosion Distances of BETA Test V2.3 ........
133
4.5
SWISS Experimental Apparatus (Ref.[G12])
x
....................
128
139
LIST OF FIGURES (Continued)
Page
Figure
4.6
TIURC-ISS Experiment Facility (Ref.[G4])
....................
141
4.7
Comparison of the TURC-1SS Test Data and the Predictions of
the VANESA Code (Ref.[P13]) ................................
144
4.8
Power Input History of BETA Test V1.3
......................
148
4.9
Power Input History of BETA Test V2.3
......................
149
4.10
Power Input History of Sandia SWISS-1 Test
4.11
Comparison between the Predicted and Measured Erosion
Distances of BETA Test V1.5 ................................
154
Comparison between the Predicted and Measured Downward
Erosion Distances of BETA Test V1.6 ..........................
155
Comparison between the Predicted and Measured Downward
Erosion Distances of BETA Test V1.7 ..........................
156
Comparison between the Predicted and Measured Downward
Erosion Distances of BETA Test V1.8 ..........................
157
Comparison between the Predicted and Measured Downward
Erosion Distances of BETA Test V1.9 ..........................
158
Comparison between the Predicted and Measured Downward
Erosion Distances of BETA Test V2.1 ........................
159
Comparison between the Predicted and Measured Downward
Erosion Distances of BETA Test V2.3 ..........................
160
Comparison between the Predicted and Measured Downward
Erosion Distances of BETA Test V3.3 ..........................
161
Comparison between the Predicted and Measured Metallic
Layer Temperatures of BETA Test VO.2 ........................
162
Comparison between the Predicted and Measured Metallic
Layer Temperatures of BETA Test V 1.3 ........................
163
Comparison between the Predicted and Measured Metallic
Layer Temperatures of BETA Test V1.5 .....................
164
Comparison between the Predicted and Measured Metallic
Layer Temperatures of BETA Test V 1.6 ........................
165
Comparison between the Predicted and Measured Metallic
Layer Temperatures of BETA Test V 1.7 ........................
166
4.12
4.13
4.14
4.15
4.16
4.17
4.18
4.19
4.20
4.21
4.22
4.23
xi
..................
152
LIST OF FIGURES (Continued)
Figure
4.24
4.25
4.26
4.27
4.28
Pag
Comparison between the Predicted and Measured Metallic
Layer Temperatures of BETA Test V1.8 ........................
167
Comparison between the Predicted and Measured Metallic
Layer Temperatures of BETA Test V1.9 ........................
168
Comparison between the Predicted and Measured Metallic
Layer Temperatures of BETA Test V2.3 ........................
169
Comparison between the Predicted and Measured Metallic
Layer Temperatures of BETA Test V3.3 ........................
170
Gas Generation Rates of BETA Test V1.3 Calculated by the
Film Collapse M odel
4.29
4.30
4.31
4.32
........................................
Gas Generation Rates of BETA Test V1.9 Calculated by the
Film Collapse M odel ........................................
173
Gas Generation Rates of BETA Test V2.3 Calculated by the
Film Collapse M odel ........................................
174
Gas Generation Rates of BETA Test V3.3 Calculated by the
Film Collapse M odel ........................................
175
Comparison between the Predicted and Measured Downward
Erosion Distances of SW ISS-1 Test
4.33
172
............................
180
Comparison between the Predicted and Measured Downward
Erosion Distances of SWISS-2 Test ............................
4.34
Comparison between the Predicted and Measured Downward
Erosion Distances of TURC-1T Test
4.35
181
..........................
182
Comparison between the Predicted and Measured Downward
Erosion Distances of TURC-1SS Test ..........................
4.36
Comparison between the Predicted and Measured Downward
Erosion Distances of TURC-2 Test
4.37
............................
4.39
4.40
184
Comparison between the Predicted and Measured Melt
Temperature Histories of SWISS-1 Test
4.38
183
........................
186
Downward Erosion Distances of BETA Tests (Prediction
versus Experiment) .........................................
.189
Least Square Fits of the Various Heat Transfer Models on the
Predictions of the Downward Erosion of BETA Tests ............
191
Relative Downward Erosion Distances of BETA Tests
192
xii
............
LIST OF FIGURES (Continued)
Page
Figure
5.1
5.2
5.3
5.4
5.5
5.6
5.7
5.8
5.9
5.10
5.11
5.12
5.13
5.14
5.15
5.16
5.17
Concrete Ablation Distances Predicted by Different Heat
Transfer Models ...........................................
209
Melt Temperature Histories Predicted by Different Heat
Transfer Models ...........................................
210
Gas Generation Rates Predicted by Different Heat
Transfer Models ...........................................
212
Aerosol Generation Rates Predicted by Different Heat
Transfer Models ...........................................
213
Accumulated Aerosol Releases Predicted by Different Heat
Transfer Models ...........................................
215
Lanthanum Release Rates Predicted by Different Heat
Transfer Models ...........................................
216
Lanthanum Release Fractions Predicted by Different Heat
Transfer Models ...........................................
217
Tellurium Release Fractions Predicted by Different Heat
Transfer Models ...........................................
218
Antimony Release Fractions Predicted by Different Heat
Transfer Models ...........................................
219
Strontium Release Fractions Predicted by Different Heat
Transfer M odels ...... ......................................
220
Downward Ablation Distances for Different Concrete
...............................
Decomposition Temperatures
222
Melt Temperature Histories for Different Concrete
................................
Decomposition Temperatures
223
Gas Generation Rates for Different Concrete
...............................
Decomposition Temperatures
224
Aerosol Generation Rates for Different Concrete
................................
Decomposition Temperatures
225
Accumulated Aerosol Releases for Different Concrete
................................
Decomposition Temperatures
227
Lanthanum Release Fractions for Different Concrete
................................
Decomposition Temperatures
228
Downward Ablation Distances for Different Types of Concrete
XHi'
....
229
LIST OF FIGURES (Continued)
Figure
Page
5.18
Melt Temperature Histories for Different Types of Concrete
5.19
Gas Generation Rates for Different Types of Concrete
5.20
Aerosol Generation Rates for Different Types of Concrete
5.21
Accumulated Aerosol Releases for Different Types of Concrete
5.22
Tellurium Release Fractions for Different Types of Concrete
5.23
Downward Ablation Distances for Different Amounts of
Zirconium .................................................
......
..........
230
..
........
232
233
....
234
......
235
237
5.24
Melt Temperature Histories for Different Amounts of Zirconium
5.25
Gas Generation Rates for Different Amounts of Zirconium
5.26
Aerosol Generation Rates for Different Amounts of Zirconium
5.27
Accumulated Aerosol Releases for Different Amounts of
Zirconium ...................................................
241
Lanthanum Release Fractions for Different Amounts of
Zirconium ..................................................
242
5.29
Tellurium Release Fractions for Different Amounts of Zirconium
244
5.30
Downward Ablation Distances Predicted by the Periodic Contact
Model with Different Initial Debris Temperatures ................
246
5.28
5.31
5.32
5.33
5.34
5.35
5.36
5.37
..
........
238
239
....
240
Downward Ablation Distances Predicted by the Gas Film
Model with Different Initial Debris Temperatures ................
247
Melt Temperature Histories Predicted by the Periodic Contact
Model with Different Initial Debris Temperatures ................
248
Melt Temperature Histories Predicted by the Gas Film
Model with Different Initial Debris Temperatures ................
249
Gas Generation Rates Predicted by the Periodic Contact
Model with Different Initial Debris Temperatures ................
251
Gas Generation Rates Predicted by the Gas Film
Model with Different Initial Debris Temperatures ................
252
Aerosol Generation Rates Predicted by the Periodic Contact
Model with Different Initial Debris Temperatures ................
253
Aerosol Generation Rates Predicted by the Gas Film
Model with Different Initial Debris Temperatures ................
xiv
254
LIST OF FIGURES (Continued)
Page
Figure
5.38
5.39
5.40
5.41
5.42
5.43
5.44
Accumulated Aerosol Releases Predicted by the Periodic Contact
Model with Different Initial Debris Temperatures ................
255
Accumulated Aerosol Releases Predicted by the Gas Film
Model with Different Initial Debris Temperatures ................
256
Fission Products Release Rates Predicted by the Periodic
Contact Model with Different Initial Debris Temperatures
........
257
Lanthanum Release Fractions Predicted by the Periodic Contact
Model with Different Initial Debris Temperatures ................
258
Lanthanum Release Fractions Predicted by the Gas Film
Model with Different Initial Debris Temperatures ................
259
Tellurium Release Fractions Predicted by the Periodic Contact
Model with Different Initial Debris Temperatures ................
260
Tellurium Release Fractions Predicted by the Gas Film
Model with Different Initial Debris Temperatures ................
261
Downward Ablation Distances for Different Amounts of Melt
5.46
Melt Temperature Histories for Different Amounts of Melt ........
5.47
Gas Generation Rates for Different Amounts of Melt
5.48
Aerosol Generation Rates for Different Amounts of Melt .
5.49
Accumulated Aerosol Releases for Different Amounts of Melt
5.50
Lanthanum Release Fractions for Different Amounts of Melt
5.51
Downward Ablation Distances Predicted by the Periodic Contact
........................
Model with Different Amounts of FeO
5.52
....
266
....
267
......
268
..
.271
........................
........................
........................
272
273
274
Melt Temperature Histories of the High Initial Debris Temperature
Cases with Different Amounts of Decay Heat
5.56
265
Lanthanum Release Fractions Predicted by the Gas Film
Model with Different Amounts of FeO
5.55
..
............
Lanthanum Release Fractions Predicted by the Periodic Contact
Model with Different Amounts of FeO
5.54
264
Accumulated Aerosol Releases Predicted by the Periodic Contact
Model with Different Amounts of FeO
5.53
263
....
5.45
...................
276
Accumulated Aerosol Releases of the High Initial Debris Temperature
Cases with Different Amounts of Decay Heat
xv
..................
277
LIST OF FIGURES (Continued)
Page
Figure
5.57
5.58
5.59
Accumulated Aerosol Releases of the Low Initial Debris Temperature
..................
Cases with Different Amounts of Decay Heat
278
Lanthanum Release Fractions of the High Initial Debris Temperature
..................
Cases with Different Amounts of Decay Heat
279
Lanthanum Release Fractions of the Low Initial Debris Temperature
..................
Cases with Different Amounts of Decay Heat
280
. .
282
5.60
Downward Ablation Distances for Different Layer Configurations
5.61
Melt Temperature Histories for Different Layer Configurations
5.62
Aerosol Generation Rates for Different Layer Configurations
5.63
Accumulated Aerosol Releases for Different Layer Configurations
286
5.64
Lanthanum Release Fractions for Different Layer Configurations
287
5.65
..................
........................
5.70
5.71
5.72
5.73
290
291
292
Effect of the Corrected Version of CORCON/MOD2 on the
Prediction of the Accumulated Aerosol Release ..................
293
Effect of the Corrected Version of CORCON/MOD2 on the
Prediction of the Lanthanum Release Fraction ..................
294
Effect of the Corrected Version of CORCON/MOD2 on the
....................
Prediction of the Radial Ablation Distance
295
Effect of the Corrected Version of CORCON/MOD2 on the
Prediction of the Released Gas ................................
296
Effect of the Revised Periodic Contact Model on the
Prediction of the Downward Ablation Distance ..................
5.74
289
Effect of the Corrected Version of CORCON/MOD2 on the
Prediction of the Aerosol Generation Rate ......................
5.69
285
Effect of the Corrected Version of CORCON/MOD2 on the
Prediction of the Gas Generation Rate
5.68
......
Effect of the Corrected Version of CORCON/MOD2 on the
Prediction of the Melt Temperature Histories
5.67
283
Effect of the Corrected Version of CORCON/MOD2 on the
Prediction of the Downward Ablation Distance ..................
5.66
....
297
Effect of the Revised Periodic Contact Model on the
Prediction of the Melt Temperature Histories
xvi
..................
298
LIST OF FIGURES (Continued)
Page
Figure
5.75
5.76
5.77
5.78
Effect of the Revised Periodic Contact Model on the
........................
Prediction of the Gas Generation Rate
299
Effect of the Revised Periodic Contact Model on the
Prediction of the Aerosol Generation Rate ......................
300
Effect of the Revised Periodic Contact Model on the
Prediction of the Accumulated Aerosol Release ..................
301
Effect of the Revised Periodic Contact Model on the
Prediction of the Lanthanum Release Fraction ..................
302
xvii
LIST OF TABLES
Table
Page
1.I
Analytical Models of Corium/Concrete Interactions ..............
14
1.2
Elements and Vapor Species Considered in the VANESA Model
20
1.3
Output Data of the VANESA Model
23
2.1
Corium M aterials Properties
2.2
Simulant M aterials Properties
2.3
Freezing Phenomena Test
2.4
Interfacial Heat Transfer Test
2.5
Layer M ixing Test
2.6
Counts per Minute of Different Particle Sizes
2.7
Filter Collection Test
3.1
Downward Heat Transfer Calculated by the Gas Film Model ......
77
3.2
Test Matrix of the Early BETA Experiments
93
3.3
Empirical Constants of the Film Collapse Model
3.4
Compositions and Physical Properties of Various Melts
3.5
Compositions and Physical Properties of Concretes
3.6
Typical Values of the Parameters Used in the Periodic Contact
M od el
..........................
.
..................................
36
50
................................
....................................
.
54
59
................................
..........................................
62
..................
........................................
..................
................
..........
..............
....................................................
...
64
.
65
98
107
108
118
4.1
Test Matrix of the BETA Experiments
........................
131
4.2
Test Matrix of the SWISS Experiments
........................
140
4.3
Test Matrix of the TURC Experiments
........................
143
4.4
Test Conditions of the BETA Experiments
4.5
Melt Compositions of the BETA Experiments
4.6
Test Conditions of the Sandia Experiments
4.7
Input Parameters Used in CORCON/MIT for Experiment
..................................................
A n aly sis
Xvii
....................
..................
....................
146
147
151
153
LIST OF TABLES (Continued)
Page
Table
........
.188
4.8
AxiaTConcrete Erosion Rates of the BETA Experiments
4.9
Statistics for the Various Heat Transfer Models in the
............................
Calculations of the BETA Results
193
......................
202
5.1
Parameters Used in the Base Case Study
5.2
Melt Inventory of the Base Case (Ml) at the Start of MCCI
5.3
Phenomena and Parameters Range Used in the CORCON/MIT-
VANESA Sensitivity Study
5.4
......
203
205
..................................
Compositions of Various Melts Used in the CORCON/MIT-VANESA
Sensitivity Study ............................................
206
5.5
Phenomena and Timing of Events of the Base Case ..............
208
5.6
Timing of Events of the Cases with Different Initial Melt
Temperatures and Different Amounts of FeO ....................
5.7
Integral Results of the Base Case at 3 Hours after the Start
of M C C I
5.8
Radial and Axial Concrete Erosion Distances Relative to PCM1
5.11
305
306
307
Accumulated Releases of Decomposition Gases Relative to GFM1
..................
308
Fission Products and Total Aerosol Releases Relative to PCM1
Base Case at 3 Hours after the Start of MCCI
5.13
..................
Accumulated Releases of Decomposition Gases Relative to PCM1
Base Case at 3 Hours after the Start of MCCI ..................
Base Case at 3 Hours after the Start of MCCI
5.12
..................
Radial and Axial Concrete Erosion Distances Relative to GFM1
Base Case at 3 Hours after the Start of MCCI
5.10
304
..................................................
Base Case at 3 Hours after the Start of MCCI
5.9
270
..................
Fission Products and Total Aerosol Releases Relative to GFM1
Base Case at 3 Hours after the Start of MCCI ..................
xix
309
310
CHAPTER 1
INTRODUCTION AND BACKGROUND
1.1 Introduction
In the current design of nuclear power plants, reactor containment systems
must withstand a set of design basis accidents, for example, the large loss-ofcoolant accident (LOCA), without the release of excessive amounts of radioactive
material. The March 28, 1979, accident at Three Mile Island (TMI) has prompted
new initiatives regarding nuclear safety. This accident, which involved a degraded
core, reached conditions more severe than those of design basis accidents. Since
then, concern about the potential for accidents beyond the design basis, namely,
core meltdown accidents, has correspondingly increased.
Before the TMI acci-
dent, the Reactor Safety Study (RSS or WASH-1400) [N1],
published by the
U.S. Nuclear Regulatory Commission (NRC) in 1975, had indicated that the core
meltdown accidents were the dominant contributors to the risk. This provided
motivation for further investigation of physical phenomena that may influence the
consequences of postulated core meltdown accident sequences.
The Nuclear Regulatory Commission (NRC) issued, on October 2, 1980, an
"advance notice of long-term rulemaking to consider to what extent. if any, nuclear power plant should be designed to deal effectively with degraded core and
core melt accidents" [N2).
This advance notice of rulemaking proposed to ad-
dress the-objectives and content of a degraded core regulation, the related design
and operational improvements, and their costs and benefits. NRC subsequently
issued a proposed Commission Policy Statement
N3,N4] which would implement
the Advance Notice of Rulemaking with severe accident regulatory determination
on specific standard plant designs and regulatory decisions on classes of existing
plants. The Commission decision. which could have certain economic impact. on
the indtistry, will slowly evolve diring the next several years.
1
1.1.1 The Molten Core/Concrete Interaction
1.1.1.1 Severe Accident Sequence
In the event of a LWR degraded core accident with complete failure of normal
and emergency coolant flow, the decay heat would cause fuel rods to heat up to
templ~eratures above the design limit. If the cooling failure persisted for extended
time periods, the combination of decay heat and the exothermic zirconium/water
reaction would cause melting of the reactor core. This could lead to slumping of
the molten core material (corium) down into the vessel lower plenum. Mechanical
and thermal loads imposed on the reactor vessel by the corium could lead to
vessel failure and deposition of these materials into the concrete reactor cavity.
C'orium could then attack the concrete of the cavity floor, releasing gas and vapor
(CO2
and H 2 0) and ablating the solid concrete, a phenomenon known as Molten
Core/C'oncrete Interaction (MCCI).
1.1.1.2 Physical Phenomena of MCCI
The MCCI is a long term endothermic erosion of concrete by high temperature
corium, and results in decomposition and melting of the concrete with production
of very large quantities of carbon dioxide and steam.
Radiodecay power and
chemical reaction heat are the sources which sustain long term core-concrete interactions. While the sensible heat content of the corium pool will be lost by heat
transfer either to the containment atmosphere and containment structure directly
or to the concrete. Both the heat transfer and gas release into the containment
atmosphere would pressurize and threaten the integrity of the containment building.
During the concrete decomposition, weight loss in the concrete consist of three
distinct events P11: loss of evaporable water (30 to 250 'C). loss of chemicallyconstituted water (400 to 550 C), and loss of carbon dioxide (550 to 800 C). Loss
of evaporable water is due to vaporization of molecular water from species such as
Tobermorite. Ettringite. and 3CaO - 25i0 2 - 3H 2 0.
The weight loss assigned to
chemically-constituted water is caused mainly by dehydration of Ca(OH)2. The
decarboxylation weight loss in basaltic concrete is due to thermal decomposition
of CaCO3 formed in the cementituous phase during concrete fabrication. Decarboxylation weight loss in calcareous concrete is, of course, principally the result of
thermal decomposition of the aggregate.
The corium melt contains both oxidic (U0
2,
ZrO2 , FeO, and fission product
oxides) and metallic (Fe, Cr, Ni, Zr, and metallic fission products) materials,
which may differ in density from each other. The experimental evidence shows
that the various oxides in the corium are highly miscible, as are the metallic
species, but that the two groups are mutually immiscible [P1]. Buoyancy forces
may be sufficient to separate the molten debris into two layers. However, the layer
formation may be destroyed by the large turbulence produced by the gases and
concrete constituents entering the pool from the bottom interface with concrete.
A homogeneously mixed pool could be formed if the gas generation rate is high
and the density difference between the oxidic and metallic materials is small.
The concrete decomposition gases, initally CO2 and H 2 0, may percolate
through the pool unless they can escape at the pool periphery, e.g. the gases
generated from the sideward erosion. The gases which pass through the pool may
encounter the metallic elements and be reduced as the metal is oxidized. These
chemical reactions will change the composition of the pool, add energy to the
pool and generate flammable gases H 2 and CO. The composition of the pool is
also changed by the addition of slag (molten concrete oxide) to the oxidic phase
of coriui.
The slag will dilute the oxidic layer, decrease its power density and
reduce the freezing point. Thermal properties of the melt mixtures will also be
changed.
The presence of the gases in the pool will elevate the pool surface and increase
the layer thickness, therefore changing the geometry of the corium pool.
The
bubbling of gases through the pool tends to enhance the heat transfer between
3
layers. The heat transfer process between corium and concrete is also complicated
by this gas percolation.
Furthermore, the gases, the fission products and other materials in the melt
form various chemical compounds, which may be vaporized and carried away with
the flowing gases. These vaporized materials, after emerging from the corium pool,
will form an aerosol source as they condense in the containment atmosphere. In
addition to the vapor source the flowing gases may entrain some melt material,
which could contain fission products. The releases of radionuclides and production
of aerosols from core/concrete interactions into the containment atmosphere are
identified as ex-vessel source terms.
As time progresses, the pool grows, its surface area increases and decay heat
decreases. Therefore, the pool temperature will decrease and eventually the possibility of freezing arises. In general, concrete melts between 1250 and 1775 K,
while corium melts between 1800 and 2700 K (depending upon compositions).
These data imply that even solidified core debris could melt the concrete. After
the pool solidification, the attack on the concrete shifts from the molten pool to
partially solidified debris. The relatively slow attack on concrete by solidified or
partially solidified debris may persist for a few days.
1.1.1.3 Consequences of MCCI
The accident sequences developed by the Reactor Safety Study indicated that
the interaction of molten core materials with concrete was important because it
affects two primary modes of release of radioactive materials from the containment building. First, the overpressurization release occurs when the containment
pressure is increased to containment failure pressure by the added heat and water
vapor during blowdown phase and MCCI period. This release mode can be directly affected by the gas release and heat flow from the core/concrete interaction.
Also, if the concentrations of the combustible gases (H
2
and CO) generated trom
core concrete interaction are high enough inside the containment atmosphere, de-
I
flagration and detonation of these gases may occur and lead to containment failure
by a pressure spike. Second, the molten core materials interacting with the concrete base erode the concrete which can lead to fission product release to the soil.
The detailed phenomena of the core,/concrete interaction determine if and how
fast these release modes can occur.
The core/concrete interaction also affects the characteristics (such as the nagnitude, the content, the physical and chemical properties) of the radioactive release
source term. In addition, the timing of the ex-vessel aerosol release, relative to
that of the fission product aerosol release from the primary system, and to that of
the containment failure, is largely determined by the physical phenomena of the
core/concrete interaction. All of these parameters are extremely important in the
estimation of the severe accident source term.
Recently. results from the Containment Loads Working Group [S1], Containment Performance Working Group [N5], the NRC Accident Source Term Reassess-
ment Program [G1.K1,S2,D2],
IDCOR program (I1], and other studies [G2,L1]
have demonstrated that the core/concrete interactions present the greatest threat
to the containment integrity. Most importantly. it becomes clear that refractory
fission products (such as Te, Sr, Ba, La and Ce) released during the core/ concrete
interaction, dominate the environmental source terms if containment fails
S21.
[D2.K2,
Especially in some of the accident. scenarios for the BWR design, the im-
portance of the ex-vessel releases stems from the fact that they take place into a
failed drywell and are thus available for release to the environment without passing
throughThe suppresion pool.
1.1.2 Scope of This Work
This work will focus on thermal hydraulics of the core, concrete interaction.
Some physical processes of the MCCI which are difficult to interpret from real
material experimeit will be investigated by siimulant experimnents in this study.
Tie experimental apparattus used is desigied with gas agitation aid coldiig ca-
pabilities of both a single-layer and a multi-layer simulant liquid pool to obtain
an understanding of the fundamental processes of the MCCI, such as (1) freezing
phenomena of a corium pool; (2) mixing phenomena of two immiscible layers; (3)
liquid/liquid interfacial heat transfer; and (4) liquid droplet entrainment due to
gas sparging.
Major contributions of this work are development and validation of computational models describing the heat transfer across the horizontal corium/concrete
interface. The proposed heat transfer model will be developed based on transient
heat conduction theory and bubble dynamics. and it will be incorporated into an
existing computer code for integral analysis of the MCCI. Validation of various
downward heat transfer models will be performed by comparison of the calculated
downward concrete erosion to results of real material experiments.
A study of the uncertainty caused by the downward heat transfer modeling
in the calculation of the ex-vessel source term will be made.
Impacts of other
parameters on the calculation of the ex-vessel release will be examined as well.
1.2 Background
1.2.1 General
The Reactor Safety Study provided estimates of the radioactive source term
that might result from a severe reactor accident. At that time, detailed knowledge
of the phenomena that might occur in the severe accident was not available. The
analyses were based, therefore, on simple bounding models, and the estimates
given in the RSS were intended to be conservative.
The Reactor Safety Study assumed that the mechanism for concrete erosion
by the core melt was rapid spallation (i.e. mechanical disruption) of the first half
meter depth of concrete in about 20 minutes, followed by concrete decomposition
at a rate of 0.04 mm /s.
This assumed mechanism resulted in rapid concrete erosion
and ignored t he physical fact that the core mielt cools aid would eventually begin to
6
solidify; therefore this could be considered an upper bound on the rate of erosion.
For the reference reactors in WASH-1400. concrete basemat melt-through was
predicted to occur in about 18 hours. Because of this rapid concrete erosion rate,
the rate of gas generation was also high and the containment was predicted to fail
by overpressure well within the first day of the accident. This early containment
failure time may have caused an overestimate of the health consequences of the
accident because of the large airborne radioactive inventory early in the accident
and the lack of time for population evacuation.
Since then, considerable progress in developing both a scientific basis and
computational ability have been achieved. Simulant experiments as well as large
scale real material experiments have been conducted in recent years providing
useful data, which can lead to better understanding of the phenomena that might
take place in a severe reactor accident. Analytical models for core-concrete interactions have been developed to calculate erosion rate, melt temperatures, gas
evolution and other parameters. The chemical reactions that occur in the debris
and the related aerosol production can also be calculated.
Based on the new developments, calculation of the volatile radioactive material that could be released to the environment is substantially smaller than those
reported in the RSS [A1,A2).
This finding resulted from better understanding of
containment integrity, natural retention potential of reactor systems and chemistry
of cesium iodine (CsI). However, one mechanism that might, for some sequences.
increase the radionuclide releases above those calculated in the Reactor Safety
Study is the release of nonvolatile radionuclides in the core-concrete interaction.
The magnitude of the contribution from the nonvolatile radioniclide which could
be available for long-term release is still open to question. primarily because of
the modeling unmcertainty of the MCCL. While the fundamental concepts of current models have been generally accepted, heat transfer modeling is not fully
developed and validation of chemical effects is inconplete. The understanding of
core/concrete interaction is far from complete, and calculations are also known to
depend strongly on the details of core melt progression for which many uncertainties still exist. It is important to spend more efforts on both experimental and
analytical research to improve our knowledge of the physics and chemistry in this
crucial area.
1.2.2 Melt/Concrete Interaction Experiments
The interaction between molten corium and concrete consists of many physical
and chemical processes, such as heat and mass transfer. gas genention, melting of
concrete, oxidization of metal, and aerosol release. To understand these processes,
small scale simulant material tests as well as large scale real material experiments
have been conducted. These experimental programs have provided valuable qualitative and quantitative information, enhancing the understanding of core/concrete
interactions.
1.2.2.1 Simulant Experiments
Due to the extremely high temperature involved in the melt/concrete interaction experiments, the physical details of the interaction processes are difficult
to interpret.
Various simulant experiments have been conducted to investigate
separate physical phenomena that might occur in the core/concrete interaction.
Among those simulant experiments, a conceptually simple experiment of gas
agitated ablation can be realized by using dry ice as the eroding (sublimating)
material due to heat transfer from an overlying pool of water or other low temperature liquids. Indeed, such experiments were conducted by Dhir et a].
[Di1
and Alsmeyer et a]. [A3]. Such experiments address only the non-radiative heat
transfer phenomena without accounting for mass transfer effects resulting from
the inclusion of the molten concrete slag. Based on observations from these experiments, a gas film was expected to be formed at the -olid-liqiid interface, which
would control the downward rate of energ' transfer. However, in an experiment by
Felde et a).
F1 usiig gas injection thiroighi a porus plate ito a volunetricallv
heated liquid pool, no continuous gas film was identified at modest superficial gas
velocities-(-
10 mm/s).
A simulant experiment using adipic acid (T,
151 *C) to simulate corium
and azelaic acid, sodium bicarbonate plus polyethylene glycol mixture to simulate
substrate concrete was performed by Plys [P2.
This experiment preserved most
of the major phenomena of melt/concrete interaction. On decomposition, gas was
generated and the remainder of the melt substrate became miscible with the pool
material. The decay heat of the corium was simulated by an immersed heater. In
this experiment, the measured pool temperatures indicated rapid initial cooling,
followed by slow heatup, and then renewed attack. An early aggressive phase of
the interaction showed limited effects of crusting. After crust buildup, erosion was
severely reduced due to the increase in heat transfer resistance. Pool heatup was
observed at this phase. Finally, the interface temperature rose, the crust broke
up, and a quasi-steady attack was established.
With regard to the heat transfer rate after pool solidification, an experiment
was conducted by Ahmed and Dhir [A4] in which a solid copper block embedded
with electric heaters penetrated an underlying dry ice slab. It was concluded that
the heat transfer coefficient depends on the temperature difference between the
solid and substrate.
An experiment was performed at M.I.T. [L21 to study the phenomena of
crust stability. In this experiment, air was injected through a porous plate into a
single-layer water pooland heat was removed from the bottom of the pool by a
condensing unit. The superficial gas velocity achieved in the experiment ranged
between 6.5 and 130 mm/s. No gas film was observed and a stable ice crust was
always formed across the liquid/solid interface under bubble agitation condition.
The thickness of the ice crust increased gradually. Gas bubbles penetrated the
ice crust through several locations. Air bubbles trapped in the crust layer were
observed as well.
9
Recently, an experiment conducted at the University of Wisconsin [G3] was
intended to investigate the effect of gas injection on liquid-liquid entrainment in
an isothermal system. In this scoping experiment, air was injected into a layered
pool of two immiscible liquids of different densities.
Different refrigerants. oil,
mercury, and water were used in various comiblinations to scope out the effects
of density ratio, surface tension, and viscosity on the onset of entrainment. The
experiments performed to date covered a range of density ratios of 1.5 to 15. At
the lower end of this density ratio for water and RI 13, the entrainment threshold
was observed to be at a superficial gas velocity of 20-30 mm/s. At the higher
end of this density ratio for water and mercury, the entrainment threshold was
reached at a velocity of 150-200 mm/s.
Simulant experiments have the advantage that they are relatively easy to
perform to allow detailed observations of the basic processes. However, questions
may be raised about their direct applicability to a real case situation, in view of
the potential for unforeseen scaling effects.
In addition, radiative heat transfer,
which is the dominant heat transfer mechanism in a real case, was neglected in
the sinuilant experiments.
1.2.2.2 Real Material Experiments
Many experimental programs using molten core material and various concretes have been performed to study the phenomena and provide data bases for
corium/ concrete interactions.
Thefirst real material experiment was reported by Baker et al.
(BI .
consisted of a small scale apparatus with resistive heating to melt 1 kg of
It
02 .
A penetration rate of 0.003--0.03 mm/s was measured and some spallation of
concrete was observed.
In a number of large scale experiments conducted by
Perinic et al. P3) no large spallat ion was noted and an erosion rate of about 0. 17
mm a was measured when the iron-alumina therniite melts dropped into concrete
cricibles. Vigorous gas stirring by decomposed coicrete gases was observed.
10
Peehs et al.
[P4] at Kraftwerk Union conducted a number of small scale
experiments in which molten steel was deposited into a concrete crucible.
The
experiments have been both transient and steady state. The experiments were
mainly scoping experiments in which concrete erosion was observed.
Concrete
thermophysical properties and enthalpy of decomposition were also measured.
Separate-effect experiments were conducted at Sandia National Laboratory
[C1,M1,P1] to investigate the response of concrete exposed to a high heat flux.
In these experiments, cylindrical concrete samples were exposed to nominal heat
fluxes of 280 to 2800 kW/m
2
provided by a plasma jet or radiant sources.
At
each experiment, after a brief initial transient period (30~60 seconds), erosion
took place at a constant rate. Tests at various applied heat fluxes showed that the
rate of erosion was approximately linear to the heat flux actually deposited in the
concrete.
A large variety of scoping experiments, both small scale and large scale, were
performed at Sandia by Powers et al. [M2,P5,S3].
These experiments were tran-
sient tests without internal sustained heating of melts. Various melts were generated either by thermite reaction or inside an induction furnace, and poured
on a basaltic or calcareous concrete crucible.
The erosion rate was traced by
the response of thermocouples embedded in the concrete crucibles. In these experiments, the interaction was marked by vigorous evolution of gases. Chemical
analysis showed the composition of the gases to be predominantly a mixture of
C0, C'02, H 2 , and H20. The decomposition products (slag) were largely immiscible with steel melt.
Density driven stratification of the melt into slag and
metal phases occurred quickly and was not greatly disrupted by the gas evolution
process. The erosion rates in the radial and axial directions were approximately
the same and were proportional to the absolute temperature of the melt.
Several experiments [P6,S4] have been also made at Sandia to provide internal
heating within the melt to simulate the decay heat generation of corium. It was
I1
found that the erosion was dominantly downward and radial erosion of the concrete
cavity was fairly insignificant. There are limited available quantitative results for
these tests.
An experimental program, sponsored by EPRI, has been undertaken at Argonne National Laboratory [S5] to study corium/concrete interactions, with particular emphasis on measurements of the magnitude and chemical species present
in the aerosol releases.
In addition, other aspects of the interaction to be ex-
amined include the downward heat transfer and concrete ablation rates, mixing
of corium melt, and gas release rate. Experiments have been performed inside
rectangular corium containments with base dimensions measuring 100 mm square
for small-scale tests and 200 mm square for intermediate-scale tests. Two types
of the corium containments have been constructed for different tests. One had
a water-cooled brass base with gas sparge tubes in which hydrogen and carbon
monoxide were injected into the corium to represent the concrete decomposition
gases. This containment was designed for the gas sparge test. Another was used
for a corium/concrete interaction test in which a concrete block with 305 mm
thickness was placed within the containment as a concrete basemat. Both containments consisted of water-cooled brass sidewalls where heat losses through
these walls were determined by measuring the water flow rate and temperature
rise. Various core melt mixtures (5
-
27 kg) consisting of UO 2 , ZrO2 , Zr, Fe,
Cr, Ni, CaO, SiO2 , and fission product mockup were melted by direct electrical
heating inside the corium containment.
The direct heating technique has been
successfully used in generating an internal heating rate of 1 kW/kg and achieving
melt temperatures of 2000 0C. Downward erosion was measured by the response
of embedded thermocouples. An aerosol and gas sampling system was used to collect aerosol samples. In the corium-concrete interaction tests, downward erosion
rate of about 0.02 mm/s of concrete was measured. Some quantitative results of
tl)e released aerosol and gas composition were reported. A new facility is being
developed to preform
12
larger scale (300 kg corium inventory inside 50 cm square concrete base cavity),
integral tests for addressing issues important to the modeling of the MCCI.
Recently. two major programs designed to investigate the integral behavior
of the MCCI were conducted at the Sandia Large Melt Facility and the German BETA Facility. The Sandia experiments included molten corium/concrete
[G4,G51, melt/concrete with overlying water [B2], and hot solid/concrete interactions [C2]. The BETA experiments
[A5,A6,A7]
were performed at various tem-
perature levels using sustained induction heating of the metallic melt to reach a
quasi-steady state condition. Concrete erosion rates, melt temperatures, aerosol
production, and evolved decomposition gases were measured during the tests. Detailed descriptions and results of these tests will be discussed in Chapter 4.
1.2.3 Analytical Modeling of MCCI
The analytical studies of the MCCI include three general areas:
(1) The
proportion of core-melt heat transferred downward or sideward into the concrete
causing erosion versus upward into the containment or an overlying water layer;
(2) The rate of condensible and noncondensible gas generation due to the concrete
erosion; (3) Characterization of the aerosol source term and transport during the
MCCI. Various models have been developed to calculate the concrete erosion,
decomposition gas generation, upward heat flux, aerosol production, and other
parameters (see Table 1.1).
Among those analytical models, the one to describe the heat transfer across
t he horizontal corium/concrete interface has the most direct impact on the calculations of the important parameters specified above. Two fundamental assumptions
about the existence of a stable gas filn at the horizontal corium/concrete interface
laid the foundation for development of different downward heat transfer models.
Some of the downward heat transfer models have been extensively used in the
analysis of MCCI.
i3
Table 1.1
Analytical Models of Corium/Concrete Interactions
Reference
Subject
Heat Transfer across
Dhir [D1]; Alsmeyer [A3];
Corium/Concrete Interface
Felde [F1]; Benjamin [B3];
Blottner [B4]; Murfin [M3];
Muir et al. [M4]; Henry [H1];
Ahmed and Dhir [A4];
Lee and Kazimi [L3,L4];
Muir and Benjamin [M5];
Reimann and Murfin [R1];
Kutateladze and Malenkov [K3]
Heat Transfer across
Szekely [S6]; Grief [G6];
Liquid/Liquid Interface
Konsetov [K4]; Werle [Wi];
Blottner [B4]; Greene [G7];
Muir et al. [M4];
Lee and Kazimi [L4];
Reimann and Murfin [Ri]
Heat Transfer across
R.Cole,Jr. [C3,C4];
Solid Crust
Plys [P2]; Henry [H1]
Heat Transfer to
R.Cole,Jr. et al. [C3]; Henry [Hi];
Overlying Water Pool
Ginsberg and Greene [G8]
Aerosol Generation
Powers et al. [P7]; Ginsberg [G9];
Plys et al. (P8,PIO,P11];
Butland et al. [B51;
Clough et al. [C5]
Layer Mixing
Gonzalez and Corradini [G3]
14
Since the physical processes involved in the core/concrete interactions are
quite complex and interrelated, computer codes have been developed to analyze
the integral behavior of the MCCI. They are INTER [M3], CORCON [C3,M4],
DECOMP [H1], WECHSL [R11 and VANESA [P7]. The first such program, INTER, was developed by W.B. Murfin at Sandia in 1977. This model was developed
based on very limited experimental data and many untested assumptions. INTER
was incorporated into the MARCH
[W2]
code as a qualitative tool for sensitivity
analysis. A more detailed modeling effort was undertaken and produced the CORCON series of codes, which replaced INTER in the calculation of the core/concrete
interaction. DECOMP was developed for the IDCOR (Industrial Degraded Core
Rulemaking) [I1].
WECHSL was developed at KfK, West Germany, and is in
many respects similar to CORCON. VANESA was developed at Sandia to calculate aerosol generation during the MCCI. Among these codes, CORCON and
VANESA have the most advanced modeling capability and will be described in
detail here.
(i) CORCON
The code predicts the behavior of the core melt-concrete interaction, including heat transfer, concrete ablation, cavity shape change, melt temperature history,
and gas generation. The first version, CORCON/MOD1, was released in 1981. An
improved model called CORCON/MOD2 based on insights from additional experimental data became available in 1984. The major changes are the inclusion of
models for solidification of the melt and for its interactions with coolant water.
The CORCON/MOD2 code is now widely used in the ex-vessel source term calculations.
CORCON is a stand-alone computer code which models most of the physical
and chemical processes which may occur during the thermal interaction between
molten core debris and concrete. The concrete is assumed to have a cavity initially
of one of several axisymmetric geometries. A mixture of molten oxidic and metallic
15
melts is deposited into the cavity, consisting primarily of molten fuel, U0 2 and
oxidized iron and zirconium for the oxide phases, and steel and zirconium for the
metallic phases.
These oxides and metals, being immiscible. are assumed to separate into distinct layers with no mixing between oxides and metals.
The pool structure is
assumed to consist of up to six layers with their spatial orientation determined
by their respective densities (see Fig. 1.1). Several of the layers, i.e.. oxide-metal
mixture layers are not currently modeled in the code although present in the numerical solution scheme, and as such are assumed to have zero mass. In a typical
calculation, the oxidic layer is calculated to be more dense initially than the metallic layer. Later, when molten concrete slag dilutes the heavy oxide layer, the oxidic
layer becomes lighter than the metal layer and rises to the top.
Energy transfer in the system is modeled by empirical and analytical correlations from the literature based upon available empirical data wherever possible. The heat transfer modes that are modeled include heat transfer across the
melt/concrete interfaces, between layers in the pool, and from the pool surface to
the surrounding atmosphere and structures.
The gas film model [A3,D1] is employed to describe heat transfer across the
melt /'concrete interface. It assumes the existence of a stabilized gas film on the
pool bottom and a flowing gas film along steeper portions of the sides of the
pool. The gas film is assumed to be composed of concrete decomposition gases.
Without-water coolant layer, upward heat transfer to the containment atmosphere
is described by a combination of radiative and convective processes. With coolant
layer, boiling heat transfer is also included to describe the upward heat loss of
the molten pool. The model in CORCON/MOD2 includes the full boiling curve,
based on standard pool boiling correlations. No correction is made for the effects
of gas injection at the melt/coolant interface.
16
SURROUNDINGS
MELT ATMOSPHERE
(REACTING GAS MIXTURE)
COOLANT/
CONCRETE
INTERFACE
REGION
J
L- VENT TO
IF
CONTAINMENT
SCOOLANT LAYER
UGHT OXIC LAYER
METALLIC LAYER
CONCRETE
HEAVY OXDIC LAYER
(PRINCIPALLY UO2 )
MELT/CONCRETE
INTERFACE REGION
Figure 1.1
CONCRETE
Schematic Diagram of CORCON System (Ref.'C3)
17
The solidification model in the CORCON/MOD2 assumes that a crust forms
on any surface whose temperature falls below the solidification temperature. The
mechanical stability of the crust is not considered. It is believed that other regimes
may exist, and that both the mechanical strength of a crust and the loads imposed
on it by concrete decomposition gases are important in determining the true behavior in any given case. Heat transfer across the crust is modeled as heat conduction.
At the solid core/concrete interface, existence of a stable gas film is assumed and
the heat transfer is again described by the gas film model.
Mass transfer between layers and the surroundings is assuned to be instantaneous between time steps of the calculation. Light concrete oxides entering the
pool in the oxidic layer remain there and reduce the average density of the oxidic
layer. Those entering the metallic layer rise through the layer to form a light oxide
layer above. Oxidized metallic components also leave the metallic layer to join the
light oxide layer. Eventually, when concrete slag has diluted the heavy oxide layer
such that its density is less than that of the metallic layer, the layers flip and the
heavy oxide and light oxide layers, being soluble combine to form one.
Concrete decomposition is treated as a one-dimensional, quasi-steady state
ablation process, dependent on the local heat transfer rates at several hundred
integration points around the periphery of the melt/concrete interface. Based on
the assumption that the concrete erosion is a quasi-steady process, all the sensil;ie
heat, reaction heat and latent heat of the concrete are lumped together to form an
effective decomposition enthalpy. The concrete ablation rate is then determined
by the energy balance at the decomposition interface.
Two phase hydrodynamics is accounted for by a bubbly/churn-turbulent flow
drift flux model based upon the gas flow rates resulting from the concrete decomposition rate. The void fraction of each layer is calculated to determine pool swell
and layer geometry.
Comparisons have been made between calculations of CORCON/MOD2 and
results of the Sandia and KfK BETA experimental programs. Despite good agreement with some tests (B6,K51, the gas film model is not adequate for describing
the downward heat transfer of the core/concrete interaction for most of the tests
"K6".
(ii) VANESA
VANESA is a mechanistic description of the aerosol generation and fission
product release during core debris interaction with concrete. The model predicts
the mass, composition, and mean particle size of radioactive and non-radioactive
materials liberated as vapors or particles during the interaction. Thus, mass release
by vaporization and mechanical processes are included.
A substantial portion of the VANESA model is devoted to the analysis of
vaporization.
It contains a library of thermodynamic properties (free energies
from which vapor pressures are calculated) for about 125 chemical species (mostly
elements, oxides and hydroxides) that might be formed by fission products and
other melt constituents. The gas phase chemical species recognized by the model
are shown in Table 1.2.
This model considers not only the detailed thermochemistry of vaporization
but also the kinetic factors which might. prevent the vaporization process from
reaching the equilibrium limit defined by the thermochemistry. Equilibrium partial pressures of the melt constituents are first calculated to be driving forces for
vaporization of these constituents into gas bubbles. Secondly, inhibition of vaporization due to kinetic factors is considered to be caused by: (1) available surface
area, (2) mass transport in the condensed phase, (3) surface vaporization rates,
and (4) transport of vapors within the bubble. The behavior of gas bubbles. such
as bubble shape, trajectory, rise velocity and bubble size. must be counted in a vaporization kinetics model. Standard equationii
for these calculations.
19(
taken from the literature are used
Table 1.2
Elements and Vapor Species Considered in the VANESA Model
Element
Hydrogen
Oxygen
Carbon
Iron
Chromium
Nickel
Molybdenum
Ruthenium
Tin
Antimony
Tellurium
Silver
Manganese
Calcium
Aluminum
Sodium
Potassium
Silicon
Uranium
Zirconium
Barium
Strontium
Cesium
Lanthanum
Cerium
Niobium
Iodine
Vapor Species
H, H2, OH, H2 0
0, 02, OH, H20, CO, CO 2
CO, CO 2
Fe, FeO, FeOH, Fe(OH)2
Cr, CrO, Cr0 2 , Cr03 , H 2 CrO4
Ni, NiO, NiOH, Ni(OH)2
Mo, MoO, MoO 2 , MoO 3 , H 2 MoO4 , (MoO 3 )2 ,
(MoO 3 )3
Ru, RuO, RuO 2 , RuO 3 , RuO4
Sn, SnO, SnOH, Sn(OH)2 , SnTe
Sb, SbOH, Sb(OH)2 , Sb 2 , Sb 4 , SbTe
Te, TcO, TeO 2 , Te 2 0 2 , H 2 TeO 4 , Te 2
H 2 Te, SnTe, SbTe, AgTe
Ag, AgOH, Ag(OH) 2 , AgTe
Mn, MnOH, Mn(OH) 2
Ca, CaO, CaOH, Ca(OH)2
Al, AIO, AlOH, Al 2 0, A10 2 , Al 2 02
Al(OH) 2 , AIO(OH)
Na, Na 2 , NaOH, (NaOH)2 , NaO, NaH
K, K2, KOH, (KOH)2 , KO, KH
Si, SiO, SiO2 , SiOH, Si(OH)2 , Si(OH)4
U, UO, U0 2 , U0 3 , H 2 UO4
Zr, ZrO, ZrO2 , ZrOH, Zr(OH)2
Ba, BaO, BaOH, Ba(OH)2
Sr, SrO, SrOH, Sr(OH)2
Cs, Cs 2 , CsOH, Cs 2 (OH) 2 , CS 2 O
CsO, CsI
La, LaO, LaOH, La(OH)2
Ce, CeO, CeOH, Ce(OH)2
Nb, NbO, NbO 2 , NbOH, Nb(OH) 2
Cs!, HI, 12, 1
20
The VANESA model also accounts for aerosol production by mechanical processes.
Mechanical aerosols have the bulk composition of the melt from which
they are formed rather than being enriched in volatile species as are aerosols
formed by vaporization. Within the context of the VANESA model only the uppermost portion of the core debris (oxidic layer) participates in the mechanical
aerosol production process.
The contribution of mechanical action is relatively
small compared to the vaporization process, however, the amount of mechanically
generated aerosol is not negligible. Especially, later in the course of core/concrete
interactions, mechanical process may be the dominant source of aerosols.
The layer configuration assumed in the VANESA model is rather simple, a
stratified configuration with the oxidic layer on top of the metallic layer.
The
actual density of each layer and the possibility of mixing are not considered in
the determination of the layer configuration.
Fission products and other melt
constituents (steel, zircaloy, and concrete) are apportioned initially to the metallic
and oxidic layers based on prior experimental observations and calculations.
In case of an overlying water pool, the VANESA code will account, for aerosol
scrubbing by gravitational settling, random diffusion, and inertial impaction. A
decontamination factor for each particle size is calculated to give the mass within
the size range that emerges from the water pool.
The key input requirements of VANESA are:
" chemical composition of the core debris including fission product inventories
at the start of core/concrete interaction:
" chemical composition of the concrete;
" temperature histories of the melt during the core/Iconcrete interaction;
" the rate at which molten concrete is incorporated into the melt;
* gas (C'02 and H 2 0) generation rates by melt attack on tihe concrete: and
" the geometric top surface area of the molten pool during core/concrete interact ion.
21
The first of these input quantities is obtained from the results of accident analyses
with the MARCH and CORSOR [K7] models. The second of the input quantities
is obtained from specifications of the reactor plant under analysis. The remaining
input data are obtained from the CORCON code.
A detailed listing of the specific output quantities from the VANESA model
is provided in Table 1.3. The most important of these output quantities for subsequent use in the analysis of a severe accident source term are:
" aerosol mass generation rate;
" chemical composition of the aerosolized mass in terms of fission products,
concrete constituents, and strurctural materials;
e particle size of the aerosol;
" material density of the aerosol;
" gas flux during core/concrete interaction; and
" chemical composition of released gases.
These quantities specify the ex-vessel source term for calculations of containment
response with the NAUA [B7] or CONTAIN [B81 code.
At this time, few experimental data exist for overall validation of the VANESA
code. Additional validation is being carried out as more data becomes available.
Uncertainties certainly exist (e.g., the assumptions of a layered melt, of unity
activity coefficients, of initial chemical form, and continued gas permeability below
the solidus temperature) in the calculation of aerosol release.
1.2.4 Ex-Vessel Source Term Assessment
A source term is defined as the quantity, timing, and characteristics of the
release of radioactive materials to the environment following a postulated severe
reactor accident. Source term assessment is employed for a variety of regulatory
applications [S2], including plant siting evaluation, emergency planning, evaluation
of engineering safety features such as containment isolation and containment spray
99
Table 1.3
Output Data of the VANESA Model
1. Aerosol Properties
* Density of aerosol material (g/cm3 )
" Mean aerosol particle size (ym)
" Aerosol generation rate (g/s)
" Aerosol concentration at STP (9/m 3 )
" Aerosol concentration in cavity (g/m 3 )
2. Aerosol Composition
" Fission products (mass percent CsI, Cs 2 0, Te, Ru, Sb, Mo, SrO, BaO,
CeO2 , La 2 0 3 , Nb 2 Os)
" Concrete constituents (mass percent Na 2 0, K 2 0, Al 2 03, SiO 2 , CaO, FeO,
Cr
2
03)
* Fuel and structural materials (mass percent Fe, Ni, Cr, Mn, Sn, Ag, ZrO2 ,
U0 2 )
3. Melt Composition
" Change caused by aerosol formation
" Change caused by metal oxidation
" Change caused by concrete melting
4. Released Gas Characteristics
" Composition (volume percent CO, CO 2 , H 2 , H20,
" Gas flow rate (moles/s)
" Superficial velocity (m/s)
23
02,
OH, 0, H)
additives, qualification of safety-related electrical equipment for performance under accident conditions, environmental impact statements, post-accident monitoring requirements, and criteria for re-entry of a plant after an accident. Furthermore. an understanding and quantitative assessment of source terms is necessary
for conducting probabilistic risk assessments, which are becoming a significant
part of the regulatory decision process.
The core/concrete interactions could persist for many days, and the ex-vessel
fission products and aerosols released from core/concrete interactions into the containment atmosphere, denoted as the ex-vessel source term, could be crucial to
the total aerosol concentration in the containment. The resulting higher aerosol
concentration in containment could lead to increases in agglomeration rates which
could produce a significant reduction in source terms. In the cases where containment failure is substantially delayed, the fission products escaping from the
core/concrete interaction might become the principal source of radioactivity transported to the environment. The less volatile fission product species would be the
principal concern in this event, since the more volatile species would have escaped
from the fuel and be largely removed from the containment atmosphere by condensing on surfaces or agglomerating and settling earlier in the sequence. The overall
effect of the core/concrete interaction on source terms depends on the relative
quantities of materials and the timing of the peak release rate. These parameters, in turn, are strong functions of the modeling of the core/concrete interaction
process, the specifics of the accident sequence, and the plant parameters.
1.2.4.1~Integrated Approach
A 1981 review of source term technical bases, NUREG-0772
[N6),
pointed out
the need for an integrated approach to source term assessment and an improved
data base. To initiate this reassessment, the NRC funded a source term study
at Battelle Columbus Laboratories reported in BMI-2104 [G1]. This study was
the first integrated assessment of source terms, involving a number of computer
21
codes based on the recent severe accident reseaech results. These codes were
then coupled to form the BMI-2104 code suite (see Fig. 1.2) that provided the
appropriate feedback involved in realistic accident sequences. Subsequently, the
NRC made improvements to the BMI-2104 suite of codes as a result of extensive
reviews. The revised set of codes is referred to as the NRC's Source Term Code
Package (see Fig. 1.3) [G10]. Also, anticipating further research advances, NRC is
developing a new, fully integrated, code package called MELCOR.
The industry has also made substantial effort to provide integrated source
term assessment by developing its own integrated code package called MAAP (a
proprietary code) [F2]. MAAP simulates an accident transient, and specifically
accounts for system events which occur during the transient including operator interventions, until a permanently coolable state is achieved or until the containment
has failed and depressurized. MAAP includes models for the important phenomena which might occur during accident sequences leading to degraded cores. The
code is highly modularized so that it can incorporate alternate physical models
as they are developed and so that it can be adapted for different reactor plant
configurations.
When the effort of BMI-2104 was initiated, there was an expectation among
many in the nuclear community that a correct treatment of the physical and chemical behavior of fission product release and transport would show a reduction of
several orders of magnitude in calculated source terms compared with the Reactor
Safety Study. The American Nuclear Society concluded in the technical summary
of its report [Al], under the heading "Major Findings" that current knowledge
is sufficient "to warrant the reduction of calculated source terms from estimates
in WASH-1400 by more than an order of magnitude to several orders of magnitude, except for noble gases".
However, the results presented in NUREG-0956
based on the analyses of the Source Term Code Package do not substantiate such
generalization for certain severe accident sequences. For some sequences, it was
25
Fission Product
Transport
Thermal Hydraulic
ORIGEN
MARCH
-Behavor--7
Fission Product
Inventory in Fuel
Overall
I
I
Behavior of
Reactor Coolant System,
Molten Core, and
Containment
CORSOR
I
Retained
Release
in Fuel
fromFe
TRAPMELT
MERGE
3
Reactor Coolant System
m mm
Transport and
Reactor Coolant System
Retention
Release from
Core-Concrete Melt
I
Detailed Temperature,
Pressure, and Flow in
<*
H
ConcN
Detaild CoreConcrete Temperature
and Interactions
~1
I
I
I
I
I
I
I
I
I
I
I
I
-J
1
I
NAUA, SPARC, ICEDF
Containment Transport
and Retention
Release of fission products to the
environment: Source Term
Figure 1.2 BMI-2104 Codes Suite (Ref.(S21)
26
Release of Fission Products to the
Environment: Source Term
Figure 1.3 Source Term Code Package (Ref.['S21)
27
found that the calculated release fractions of refractory, nonvolatile fission products
(such as Sr, Ba, La and Ce), which are critically depend upon the modeling of
the core/concrete interactions, are even higher than those of the Reactor Safety
Study.
1.2.4.2 Estimates of Uncertainties
Analytical models for corium/concrete interactions have been developed to
calculate erosion rate, debris temperatures, gas evolution, aerosol production, fission product release, and other parameters. However, it is recognized [A2] that
uncertainties inherent in the modeling efforts are still large. The calculation of
the release of refractory fission products critically depends the modeling of the
core/concrete interaction. A considerable effort is being made by several groups
in the US and abroad to estimate the uncertainty of the results that come from
various calculational methods.
A study aimed on estimation of the uncertainty of radiological source term,
referred to as Quantitative Uncertainty Estimation for the Source Term (QUEST)
[L1], has been performed at Sandia National Laboratories. In this study, several
computer codes were incorporated to model the whole possible sequences during
the course of a severe reactor accident. It was pointed out that the uncertainty can
develop in two ways. First, because the models of a particular stage of the accident are not accurate, uncertainty in the calculations can be introduced. Second,
uncertainty in the previous stage of the accident will be propagated through the
present stage, and can be either amplified or attenuated during the propagation
througirthe series of codes.
The QUEST study provided the first attempt to evaluate the changes in the
calculated source terms resulting from reasonable changes in the input parameters
and the models.
Only three seqiences were evaluated.
Based on this study,
however, it can be concluded for those sequences that if the mode and timing of
containment failure is considered to be fixed. the release fractions would likely
2S
show a span of two decades because of input and model changes. An important
development in the QUEST study was a method of scanning the entire calculation
to determine the important parameters for a particular output of interest. Based
on this method, future studies can focus on the estimation of the uncertainty
factors important to the purpose at hand.
Recently, a program called QUASAR (Quantification and Uncertainty Analysis of Source Terms for Severe Accidents in LWRs) was initiated at Brookhaven
National Laboratory [P9). The QUASAR study is, in a way, an extension of the
earlier QUEST study. Several important improvements are being incorporated in
QUASAR. The major improvement is related to data interpretation; QUASAR
will provide probability distribution functions for the ranges of parameter (input)
variations, and will determine a probability distribution for source term values
(output). QUEST used ranges of input parameter variations that were considered reasonable, and provided high and low source term estimates based on a
judgmental selection of compatible input values.
Several industry efforts to analyze uncertainties and sensitivities in their results have recently been reported [F3,F4). These efforts identified major factors
and parameters controlling uncertainties, but they did not result in quantified
ranges of source term values.
1.3 Structure of This Work
Simulant experiments using water and cyclohexane were conducted to simulate several physical phenomena related to the core/concrete interaction. In these
experiments, water and/or cyclohexane were poured into a test cell to form a
single- or a multi-layer liquid pool. The test cell was constructed with porous
plates where air was injected through the bottom of the pool. Heat content of the
pool was removed by a condensing unit. Various phenomena of the gas agitated
pool were observed. Results of these experiments are presented in Chapter 2.
29
In Chapter 3, a downward heat transfer model was formulated from transient
heat conduction theory and bubble dynamics to describe the heat transfer across
the horizontal corium/concrete interface. The proposed model is a revision of the
Lee and Kazini approach [L4] which assumes that no gas film exists at the bottom
interface. Based on the early BETA experimental observations, a combined gas
film-periodic contact model was also developed involving hydrodynamic stability
limits of the gas film to determine the actual mode of the downward heat transfer.
In addition, a simple model was developed to calculate the heat transfer between
a solid pool and concrete. These proposed models were incorporated into CORCON/MOD2 to form a new version CORCON/MIT for integral analysis of the
MCCI. A review of existing models and their differences in the calculation of the
downward heat flux is also presented.
In Chapter 4, several large scale real material experiments were reviewed.
Analyses of these experiments using CORCON/MIT were performed. Validation
of various downward heat transfer models by comparisons of their calculations
with experimental results was made.
Since significant differences among various models of the downward heat
transfer were found, a sensitivity study was made to investigate the impact of the
downward heat transfer model on the ex-vessel aerosol release. CORCON/MIT
and VANESA codes were used in this study. Uncertainties of the calculated aerosol
release caused by other parameters were studied as well. The results and discussions are presented in Chapter 5.
Finally, conclusions are drawn in Chapter 6, and some recomendations for
future wortare made.
30
CHAPTER 2
SIMULANT EXPERIMENTS
2.1 Objective
Several physical processes, such as melt freezing, liquid/liquid interfacial heat
transfer, layer mixing, and droplet entrainment due to gas sparging, are involved
in the analysis of the core-concrete interaction. Since these processes are hard to
observe in a real material experiment, simulant experiments are designed with gas
agitation and cooling capabilities of both a single-layer and a multi-layer liquid
pool to investigate such phenomena in qualitative and quantitative ways.
2.2 Introduction
2.2.1 Freezing Phenomena
During early stages of the MCCI, the corium pool can be fully molten and
result in a vigorous thermal attack on the concrete. Since the MCCI is an endothermic reaction, and internal heating by fission products decreases as time
progresses, the temperature of the corium will fall below its solidification point
and freezing begins after some period of interaction. Initially, before freezing, the
dominant heat transfer process within the pool is convection. After solidification
begins, the heat transfer mode at the pool boundaries has to be determined by
the freezing characteristics of the corium pool. There are two possible situations
for melt freezing. First, a thin crust may form on the boundaries separating the
molten material and its surroundings. Second, solid crystals may precipitate out
of solution creating a two phase solid-liquid slurry pool.
These two models of freezing lead to different heat transfer modeling of the
core/concrete interaction.
If a slurry pool forms, it is expected that the heat
transfer rates across the pool boundaries is governed by the convective process
before the pool is fully solidified. On the other hand, if the solid exists as a stable
31
growing crust surrounding the molten pool, then the crust provides an additional
thermal resistance between the interior of the pool and its boundaries, and the
heat transfer rate is limited by a conductively controlled process which is far less
effective than the convective process.
In CORCON/MOD2, a crust freezing model was developed assuming that a
stable growing crust forms on any surface (such as the corium/concrete, metallic
layer/oxidic layer, and corium/water interfaces) whose temperature falls below the
solidification temperature. This model also assumes that the mechanical strength
of the crust is enough to hold it coherent in spite of the loads imposed by the
concrete decomposition gases. However, there is no experimental evidence to substantiate these assumptions.
In this study, a simulant experiment will be conducted to investigate the
freezing phenomena of molten materials under the gas agitation condition.
2.2.2 Interfacial Heat Transfer
In MCCI, the immiscible metallic and oxidic phases will be separated into
two layers due to the density difference, provided that the layer mixing due to gas
agitation is limited. The temperature responses of the metallic and oxidic materials
depend not only on the amount of heat generated (decay heat in the oxidic layer
and chemical reaction heat in the metallic layer) but also the heat transfer rates
along various heat transfer paths. The interfacial heat transfer between these two
immiscible layers is characterized by natural convection agitated by transverse gas
flow. As the interfacial heat transfer increases, the heat transferred downward
into the concrete from the initially heavy oxidic layer is reduced and the upward
heat flux into the overlying metallic layer is increased. Therefore, the amount of
decomposed concrete, the melt temperatures, and the gas generation rate are all
affected to some extent by the magnitude of the interfacial heat transfer. Actually,
it has been found by a sensitivity study [G7] that the interfacial heat transfer is
an important factor in predicting the melt behavior of MCCI.
32
Several correlations have been proposed to calculate the interfacial heat transfer (see Table 1.1). The one used in CORCON/MOD1 was developed by Konsetov
[K4] and later modified by Blottner [B4], which is expressed in terms of heat transfer coefficients for the upper, U, and lower, L, layers as:
hL,U -
k (Pr
)
1/3
(2.1)
0.00274A.42]1/3
where ATia,,ye, is the temperature difference between the layer bulk and its interface
boundary, and a is the void fraction. Applying the definition of the heat transfer
coefficient to each layer separately and then together yields the following expression
for the liquid/liquid interface temperature, Tint:
(2.2)
hLTL + hUTU
hL + hU
where TL and TU are the lower and upper layer temperatures, respectively. The
overall interface heat transfer coefficient, hj, can be expressed as:
hy =
hUhL
hLT+ hL
ATBulk
where q'j is the heat flux across the liquid/liquid interface, and the temperature
difference ATBik is equal to (TU7 - TL). The modified Konsetov correlation has
taken into account the effect of natural convection in terms of
A Ti yer, and also
the effect of superficial gas velocity in terms of the void fraction a.
The void
fraction a is obtained by [M4]:
a =
,
u
1.53
gA
(2.4)
where tit is the terminal velocity of the rising bubble, and the Laplace constant A
is defined as:
-g0.5
A
g(e-P)(2.5)
33
The constants appearing in the modified Konsetov correlation were obtained by
fitting limited experimental data (only one data point) of a slag-metal system [S6].
It was found later that this correlation significantly underestimates the oil/water
experimental data of Werle [W1], by as much as two orders of magnitude.
Based on the assumption of transient heat conduction between the arrival of
consecutive bubbles, Szekely [S6] derived an interfacial heat transfer coefficient
for bubble stirred interface of two immiscible liquid layers. It was given in the
following form:
hU,L =
2
(2.6)
(pkc)
where t, is the time interval between the arrival of consecutive bubbles, and Szekely
suggested that te can be calaulated as:
A
Ab No
te = A
(2.7)
where
A = total cross section area of interface
Ab = surface area swept by single bubble
No = number of bubbles produced per unit time
The Szekely's model was later modified by Blottner [B4] based on the derivation
of t, as:
t,
AVb
AbNoV
(2.8)
0.445-
Vb
Abig
j(
where V is bubble volume and rb is equivalent bubble radius. Combining equations
(2.6) and (2.8), gives:
NuUL = 1.69 (Re - Pr)0 '5
(2.9)
where
NUU,L =
k
L;
'bh
Re =34
p
-
Pr =
k
(2.10)
The modified Szekely model was then incorporated in the WECHSL code for
interfacial heat transfer calculation.
In CORCON/MOD2, an empirical correlation developed by Greene [C3] is
Greene modified the Szekely correlation based on his own experimental
used.
data (oil/water and water/mercury) as:
0 8
0
NuU,L = 5.05 Re .5Pr -
(2.11
In this correlation, the interfacial heat transfer increases with the increasing of the
liquid viscosity.
Lee and Kazimi
of
te,
[L4]
have considered a different approach in the derivation
and the contribution from the carried liquid along with the rising bubble.
They proposed an interfacial heat transfer correlation based on the Szekely model
as:
0 5
5
Nuv,L = 8.84 (J *)1' (Re- Pr)
(2.12)
where the dimensionless superficial gas velocity J* was defined as:
jg
J*L59(2.13)
=
P
-r(0.25
L
.
and PL is the density of the lower liquid layer.
The correlation constant was
obtained by fitting the water/mercury [G7] and oil/Wood's metal [W1] data with
an assumption of constant bubble radius of 1 mm.
This model increases the
dependency of the interfacial heat transfer on the superficial gas velocity from the
power of 0.5 (equation 2.9) to the power of 1.0 (equation 2.12).
All these correlations have )een applied to a real case (metal-over-oxide liquid
pair) based on the physical properties of the coriumn materials (Table 2.1) and
assumed values of bubble radius and ATBulk.
35
It is found that there are order
Table 2.1
Corium Materials Properties
Oxidic Layer
Metallic Layer
p (kg/m 3 )
7000
5750
c, (J/kg K)
600
740
k (W/m K)
3.3
40.0
(1/K)
1.0x10- 4
0.6 x 10-4
y. (pPa - s)
4000
4000
a (N/m)
0.45
1.5
Tsol (K)
2673
1673
#
36
of magnitude differences among the interfacial heat transfer coefficients predicted
by the various models (see Fig. 2.1). The effects of the parameters (rb,
\TBaLk,
and layer configuration) on the prediction of the 'interfacial heat transfer are also
shown from Figs. 2.2 through 2.4.
In this experimental study, the interfacial heat transfer between two immiscible simulant liquid materials under different gas velocities will be measured, and
compared with these analytical models.
2.2.3 Layer Mixing
In the BETA experiments [A6], it was observed that as the input power
increased, the amount of the metallic materials entrained into the oxidic layer increased significantly due to the increased gas generation rate. This entrainment
was recognized by the observation of a sharp decrease in the input power because
of the induction method of heating used in the BETA test.
This observation
supported the assumption that there may be times during the core/concrete interaction when the gas sparging rate is high enough to lead to mixing of the two
immiscible layers, oxidic and metallic materials. At other times, when the gas generation rates are reduced, the two layers will separate. This physical phenomenon
has not been properly modeled in the integral analysis codes, and it may affect
the core/concrete interaction. First, the heat transfer rate between the melt and
concrete could be affected not only because the physical properties of the reaction
materials will be altered but also the freezing phenomena of the mixed layer could
be changed. Second, the vaporization rate and chemical composition of the melt
may be changed due to the change of the oxygen potential and zirconium oxidation rate. A sensitivity study [B5] has indicated that the degree of mixing of the
pool is important in determining the magnitudes of fission product release during
core/concrete interaction. However, the mixing criterion of the gas flow limit is
not well-developed and very difficult to be identified in a real material experiment.
A study group at the University of Wisconsin [G3] is now conducting simulant
37
6
LJ
10
z
105
E0
z
'N
2
10
1[_
00
10-4
SUPERFICIAL GAS VELOCITY [m/s]
Figure 2.1
Interfacial Heat Transfer Coefficient of the Metallic/Oxidic Corium
Pool Predicted by Different Models
38
MODIFIED KONSETOV CORRELATION
z
1.
E-a
z
I=
10
2
L
10-4
100
SUPERFICIAL GAS VELOCITY [m/s]
Figure 2.2
Interfacial Heat Transfer Coefficient of the Metallic/Oxidic Corium
Pool Based on Different Temperature Differences
39
108
-
-
I
I I
1 1 11111
1 111111
Modified Szekely (rb=0. 001
I
m)
Greene (rb=0.001 m)
Lee & Kazimi (rb=0.001 r m)
Modified Szekely (rb=0.01 m)
Greene (r '=0.01 m)
Lee & Ka zimi (rb= 0.01 m
E"
Q
I
1 1 1 1 111i
106
raw
C2
z
r
................
44
wei
10 2
101
I
I
1 I II
i
1
I
I
1 1 1 11
10 -4
I
I
1 1 1 1 11
I I I I II
10-1
100
SUPERFICIAL GAS VELOCITY [m/s]
Figure 2.3
Interfacial Heat Transfer Coefficient of the Metallic/Oxidic Corium
Pool Based on Different Bubble Diameters
40
LEE & KAZIMI CORRELATION
107
Bottom Layer: Oxide
Bottom Layer: Metal
- -.
10
6
E---
z
105
4
z
E--
E- 103-
2
Si
10-
1I
iii
10-4
lmol
10-3
i
ii
i
ml
10-2
I
iiiii
ii
10-1
100
SUPERFICIAL GAS VELOCITY [m/s]
Figure 2.4
Interfacial Heat Transfer Coefficient of the Metallic/Oxidic Corium
Pool Based on Different Layer Configurations
41
experiments, involving various combinations of liquid pairs (refrigerants, oil, mercury, and water) to scope out the effects of density ratio, surface tension, and
viscosity on the layer mixing.
In this study, the intermixing phenomena of an immiscible liquid pair will be
examined with and without the influence of the freezing process.
2.2.4 Droplet Entrainment
An aerosol can be generated from a pool of molten corium due to mechanical
breakup of the liquid melt by the flowing gases. The mechanical aerosol generation
process can occur in two ways -
bursting of bubbles at the melt surface and melt
splashing at a liquid/gas interface. At low gas velocities (bubbly flow), the liquid
droplets are expected to generate from the bursting of individual bubbles as they
break through the liquid surface.
If the gas generation rate is high enough to
achieve a churn-turbulent flow regime, the gas velocities would be sufficiently high
to entrain droplets of liquid (caused by liquid filament and sheet instabilities at the
liquid/gas interface) within the bulk of the two phase medium. The entrainment
of liquid by sparging gases has been reviewed by Kataoka and Ishii [K9] and by
Ginsberg [G9].
Kataoka and Ishii suggested that gas flow through liquids in the churnturbulent regime can cause noticeable entrainment when:
jg ;> 0.325 [got(Pl 2 Pg)]
0 .25
(2.14)
They also found that correlations of the amount of material entrained had to be
categorized in terms of distance from the liquid surface. Their analysis revealed
three regions of entrainment in the axial direction from a pool surface.
In the
near surface region, entrainment is independent of height and gas velocity. In a
momentum controlled region, the amount of entrainment decreases with increasing
height from the free surface and increases with increasing gas velocity.
In the
deposition controlled (far field) region, the entrainment increases with increasing
42
gas velocity and decreases with increasing height due to deposition of droplets. The
boundaries between the regions are also dependent on the gas velocity. Kataoka
and Ishii developed single correlations for the entrainments present in the near
surface and far field regions, however, they found that the correlations for the
momentum controlled region had to be categorized in terms of the magnitude of
the gas velocity, defined as low gas flux, intermediate gas flux, and high gas flux
flow regimes. Several dimensionless parameters have been used by Kataoka and
Ishii in the development of entrainment correlations; defined as follows:
E
Gas Velocity:
J* = .
5g
0.25
a9(pl - pg
9)
Height:
H*
2
H
=
05
[g(pt - pg)
Gas Viscosity:
Ni
=
0.5
t
[pgO
D* =
Vessel Diameter:
[
D
]0.5
g(pt -
E*
Entrainment:
where
jg
pg)
Pg.g
is the superficial gas velocity, H is the height above the pool surface,
and DH is the equivalent diameter of the pool. The boundaries and entrainment
correletions were then given by:
(i) Near Surface Region
This region is limited to the viscinity of the pool surface given by:
0
< H*
0-5 (D*)04 2 (
< 1038J*N;
9
_
H~
)03
(2.15)
A~P
In this region, the entrainment is given as:
E*
=
4.84 x 10-
3
)
(
pg
43
(2.16)
(ii) Momentum Controlled Region
This region is limited to the intermediate height range given by:
lo.42(
1038JNo45(D*
3
P )-
< H* < 1970N, .3(D* )o.42
o-2
(2.17)
This region is subdivided into three regimes, depending on the gas velocity:
Low Gas Flux Regime
(
)
< 6.39 x 10-4
(2.18)
with
E= 2.213N
5
(D* )1
25
0.3 (
)
(2.19)
Intermediate Gas Flux Regime
( )
6.39 x 10~4
withH
with
5.7 x 10-4N-5(D* )-o.42(g o)0
E* =54 7
E
0 N " D )H
g
x
A.1
1N7(D)1(P
5.417 x
(2.20)
)31
2.1
(2.21)
Jg
High Gas Flux Regime
(i )
>
5.7
x
10-4 N
5(D*)
0.42
0.
(2.22)
with
)7~20
E*c
(2.23)
(iii) Far Field Region
This region is above the height given by:
H* > 1970N,.33(D* )o.42(g o. 2 3
(2.24)
In this region, the entrainment when considering the deposition is given by:
J* ) 3 exp[-0.205H/DH]
E* = 7.13 x 10-4N,.5
(2.25)
Pg
Without the deposition effect, the entrainment in the far field region is given by:
E
1.99 x 10- 3 N
44
5
Ag
(
Pg
)(J)3
(2.26)
For most reactor accident analyses, the correlation suggested by Kataoka
and Ishii in the far field region, without considering the deposition effect, is the
most proper model for mechanical aerosol generation from core melt into the
containment atmosphere.
In this experiment, however, an attempt is made to
develop a procedure which can be employed to characterize the liquid droplets
entrainment in the momentum controlled region under low gas flux conditions.
2.3 Experiment Description
2.3.1 General Features
The apparatus was designed by M. Lee [L2] to simulate two fundamental
physical processes of the corium/concrete interaction, namely, gas evolution at
the pool boundary and freezing of the pool material. However, the physical processes of internal heat generation of the corium and melting of the concrete are not
included in this experiment. The experiment apparatus consisted of four major
parts: a test cell, a cooling unit, an air supply system, and measurement instruments. A schematic diagram of the experimental apparatus is shown in Fig. 2.5,
and a detailed description of the apparatus can be seen in Ref. [L2].
2.3.2 Apparatus
(i) Test Cell
The test cell is a rectangular pool bounded from the bottom and two sides
by bronze porous plates. The front and back walls of the test cell are made of
transparent plexiglass plates from which the behavior of the pool can be observed.
The dimensions of the test unit are 225.5 mm by 244.5 mm in cross section and
250 mm in depth. Simulation of gas evolution from the pool boundary is accomplished by injecting air through the porous plate into the liquid pool. The cooling
capability of the test cell is provided by a freon 12 refrigeration cycle. An evaporator is placed behind each porouis plate for removal of the heat content of the
pool.
45
TEST CELL
F
B
COOLING UNIT
E
A
AIR SUPPLY UNIT
I ~~~;&,
Figure 2.5
rgauge
Schematic Diagram of the Simulant Experimental Apparatus
(ii) Cooling Unit
The major part of the cooling unit is a Tecumseh AH7514AC condensing unit
with a maximum capacity of 16800 BTU/hr (4.9 kW). The cooling unit contains
three independent cooling loops which are controlled by refregeration shut off
valves to activate the heat extraction in the sideward or downward directions. A
back pressure regulator was installed between the suction and discharge lines of
the freon compressor to control the operating pressure and temperature of the
condensing unit.
(iii) Air Supply Sysytem
The injected air is supplied by the air compressor of the laboratory building.
The volumetric air flow rate is measured by a Fisher & Porter 10A3557 series
(tube size 12.7 mm) rotameter. The maximum air flow reading of the rotameter
is 6.43 SCFM (3.03 £/s).
A pressure gauge was installed at the upstream of
the rotameter to measure the air system pressure.
Accounting for a pressure
correction factor and the flow area, the maximum superficial gas velocity achieved
in this experiment was 126 mm/s.
(iv) Temperature Measurement
All temperatures are measured with Type E thermocouples. A KAYE ramp
processor and scanner system are used to record the temperature data at every
prespecified time interval. The positions of the temperature measurements are
shown in Fig. 2.6.
(v) Entrainment Measurement
An optical particle counter is used to investigate the size distribution of the
entrained liquid droplets generated from the bubbling pool. A millipore membrane
(0.2 pm pore size FG filter) contained within a filter holder is looped together
with a conventional carbon-vane pump (with flow capacity of 2
e/s)
to collect
the liquid droplets. During the collection procedure, the liquid droplets above the
pool surface together with air flow are driven into the filter holder by the pump,
47
AIR INLET
Figure 2.6
Illustration of the Temperature Measurement Locations
48
and deposited on the millipore membrane. A microbalance with accuracy to 1 pg
is used to find the weight difference of the millipore membrane before and after
the collection procedure.
2.3.3 Simulant Materials
The simulant materials selected for the experimental study were water (H 2 0),
cyclohexane (C6 H
2
), and air. Physical properties of these simulant materials are
listed in Table 2.2. There are advantages for using the water and cyclohexane as
simulant materials because they are transparent, nontoxic, noncorrosive, inexpensive, chemically inert, easy to handle, and immiscible. In addition, the density
ratio of water and cyclohexane (0.78) is very close to the density ratio of the
metallic and oxidic materials (0.82) at early stages of the MCCI. This preserves
an important factor in the simulation of the intermixing phenomena of the MCCI.
In the simulation of the freezing phenomena, the solidification temperature of the
cyclohexane is higher, therefore, the upper cyclohexane layer will be frozen earlier
than the lower water layer during the cooling down process. This is similar to the
situation at later stages of the MCCI, in which freezing of the upper layer (oxidic
material) occurs earlier than that of the lower layer (metallic material).
With
this feature, the freezing phenomena of the cyclohexane can be used to indicate
whether the oxidic material in the MCCI will be frozen in the form of a slurry or
a crusting layer.
2.3.4 Test Procedures
Various experiments were conducted in this study to investigate different
physical phenomena, divided into the following categories:
" FP Test: Freezing Phenomena
" HT Test: Interfacial Heat Transfer
" LM Test: Layer Mixing
" DE Test: Droplet Entrainment
Detailed description of these tests will be given in what follows.
49
Table 2.2
Simulant Materials Properties
Water
Cyclohexane
Air
p (kg/m 3 )
1000
778
1.28
c, (J/kg K)
4180
1770
1000
k (W/m K)
0.55
0.10
0.025
/ (1/K)
p (pPa- s)
a (N/m)
Tso
0 (K)
4
1.8x10-
2.0x10-4
1000
960
0.0728
0.0253
273
279.6
50
18
(i) FP Tests
These tests were performed to observe the freezing phenomena of a bubble agitated liquid pool. Both single-layer (water dr cyclohexane) and two-layer
(cyclohexane-over-water) pools were tested under different gas velocities to identify the existence and stabilities of the boundary crusts. During the experiment,
the air flow rate was kept constant, and the pool temperature was continuously
monitored. Heat content of the liquid pool was removed in the downward direction only. Sideward cooling was not activated in these tests. The experiment was
started with liquids at room temperature, and ended at the time when significant
solidification occurred. The freezing phenomena of the liquid pool were visually
observed.
(ii) HT Tests
In this test series, fixed amounts of water (3.0 kg) and cyclohexane (1.56 kg)
were poured together into the test unit. The highest gas flow rate employed in
this test was based on the limitation of keeping the water and cyclohexane layers
well-separated without significant mixing. Heat subtraction was activated in the
downward direction.
The liquid layers temperatures as well as the surrounding
(above the pool) temperature were recorded every two minutes. Each experiment
was tested for more than 30 minutes before the cyclohexane started freezing.
(iii) LM Tests
The mixing phenomena were examined in the tests with different gas injection
rates. The water was colored congo red in order to assist the visual identification
of the mixing phenomena. The tests were performed with and without cooling the
pool to scope out the effect of solidification on the layer mixing process.
(iv) DE Tests
These tests were conducted to investigate the characteristics of the liquid particles generated from an aqueous solution pool under bubble agitation conditions.
Two types of experiments were performed: (1) optical particle counter experiments
51
for droplet size measurement; and (2) filter collection experiments for quantifying
the amount of liquid entrainment.
In the optical counter experiments, two tests were performed, one with a
pure water pool for background test, another with 1% by weight aqueous solution
of potassium sulfate (K 2 SO4 ). Both tests were conducted at a gas velocity of
jg
= 8.0 mm/s. Under this condition, the entrained liquid droplets accompanied
with the flowing air were sucked into the optical particle counter. A dryer using silica gel was set up on the collection loop in front of the counter to ensure that only
solid K 2 SO
4
particles passed through, and the numbers of particles of different
sizes in the flowing gas stream were counted by the counter. In the filter collection
experiment, several tests with the aqueous solution pool were performed to quantify the droplet entrainment subjected to different gas injection rates. During the
experiment, the bubble-induced liquid droplets suspended in the air stream were
forced by the carbon-vane pump into the filter holder which was located at certain
positions above the pool surface. After several tens of minutes, the filter was carefully removed and placed into an oven to dry the water. Residual aerosol particles
(K
2
SO4 ) remained on the filter. The amount of collected potassium sulfate was
weighed by a microbalance. The total amount of liquid entrainment could then
be calculated from the known weight fraction of potassium sulfate in the solution
pool.
2.4 Experimental Results
2.4.1 Freezing Phenomena
During the initial, transient stages of the core/concrete interaction, the debris
may be relatively hot and gas generation rates can be quite high. Superficial gas
velocities over 1 m/s have been observed in a real material experiment [P12].
However, as the melt cools down, the gas velocities can be reduced to 100 mm/s
during the freezing stages. In this experiment, a superficial gas velocity of up to
52
126 mm/s was used to investigate the freezing phenomena of the simulant materials. A list of the freezing experiments is shown in Table 2.3. Observed phenomena
are described as follows.
(i) Single Water Layer
In the water experiments, two different phenomena were observed at the moment of freezing. One is defined as bulk freezing (FP-1,2,3 tests), in which pieces
of ice appear and float around in the middle of the pool accompanied with a thin
layer of ice crust across the bottom liquid/solid interface simultaneously. Significant supercooling (liquid below solidification temperature before freezing actually
starts) was found in this type of freezing. Upon freezing, the pool bulk temperature jumped up to the solidification temperature of water. Another phenomenon
is defined as layer freezing (FP-4,5,6 tests), in which only a bottom ice crust was
formed upon freezing, and slight supercooling was observed. The bulk temperature histories of the water pool are shown in Fig. 2.7. In any event, the bottom
ice crust was always stable and the thickness of the crust increased gradually. No
top crust at the water/atmosphere interface was observed during these tests.
Even though some of the bubble generation sites were completely blocked
by the bottom crust, the permeation of the gas through the ice crust was always
observed. The porous nature of the ice crust layer could easily be seen by post-test
examination. Air bubbles trapped in the crust layer were observed as well.
(ii) Single Cyclohexane Layer
In the cyclohexane experiments, the pool was always frozen in the same fashion of the bulk freezing of the water pool. However, in the test of low gas velocity
(FP-7,8 tests), a stable growing crust was formed at the the top surface at a later
time than the formation of the bottom crust. In the test with higher gas velocity
(FP-9,10,11 tests), no top crust was observed even though the bottom crust was
still stably formed.
53
Table 2.3
Freezing Phenomena Tests
Water
Cyclohexane
ig
Test
(kg)
(kg)
(mm/s)
FP-1
1.00
6.3
FP-2
1.00
13.1
FP-3
1.00
83.3
FP-4
1.00
9.2
FP-5
4.00
51.7
FP-6
2.00
FP-7
-
1.56
6.3
FP-8
-
1.56
13.1
FP-9
-
2.34
45.0
FP-10
-
1.56
66.3
FP-11
-
1.56
126.0
-
126.0
FP-12
2.00
1.56
6.3
FP-13
2.00
1.56
13.1
FP-14
2.00
2.34
20.1
FP-15
2.00
1.56
40.0
FP-16
3.00
2.34
51.7
FP-17
3.00
2.34
104.1
54
15
o
-A
FP-5 Test (Layer Freezing)
10
0-
- 5I
0
I
I
I
I
20
40
60
80
TIME [min]
Figure 2.7 Water Pool Temperature Histories
'A,4
100
120
(iii) Water/Cyclohexane Layers
In the low gas flow tests (FP-12,13,14 tests), in which the two liquids stayed
separate into distinct layers, small pieces of solidified cyclohexane appeared in the
upper layer at the moment when the pool temperature reached the melting point
of the cyclohexane.
After a while, the cyclohexane layer was frozen as a single
piece floating atop the water layer. During this time period, no cyclohexane crust
was actually formed on any of its boundaries. As the cooling process progressed,
a stable ice crust was eventually formed at the bottom interface.
In the high gas flow rate tests (FP-15,16,17 tests), in which these two liquids
were mixed into one homogenous layer, cyclohexane was frozen as small pieces
in the homogenized layer at the beginning.
After a while, bigger pieces of the
solidified cyclohexane appeared attached to the sidewall of the test unit. Some of
the peripheral areas of the bottom plate were covered by the solidified cyclohexane.
When the pool temperature dropped further, a pure ice crust was formed at the
center region of the bottom plate. No other boundary crust was observed during
this test.
The important observation of the water/cyclohexane test is that the cyclohexane was always frozen in the slurry form without any formation of a boundary
crust.
2.4.2 Interfacial Heat Transfer
The interfacial heat transfer can be analyzed according to a macroscopic energy balance of the cyclohexane layer. Heat flow into and out of the cyclohexane
layer are calculated from measured temperatures and some estimations, given as
follows:
Qdown
[McP dT]
+ QP + Q.ide - Qar
(2.27)
where Qdown is the heat extraction rate in the downward direction through the
water/cyclohexane interface.
M and c, are the mass and heat capacity of the
56
cyclohexane, respectively.
and Qzd, are the heat addition rates from the
Q2,
pool upper surface and side walls, respectively. Qai, is the heat reduction rate
due to cool air injection. These heat flow rates are calculated as:
Qup = hup(Tatm
Qside = haide(Tv
Qair
Tu)At,
(2.28)
Tu)A,
(2.29)
-
(ThCp)air(TU - TL)
(2.30)
and the overall interfacial heat transfer coefficient is calculated by:
h
=
dow
Atu(TU - TL)
(2.31)
where
Tatm = upper surrounding temperature
Tw = sidewall surface temperature
Tu = upper layer (cyclohexane) temperature
TL = lower layer (water) temperature
Atu = cross section area of the test unit
A,
= sidewall contact area
mair = mass flow rate of the injected air
(cp)air = heat capacity of air
Based on the natural convection over a plate with gas injection, the heat transfer
coefficient hu, can be estimated as [M4]:
hu, = 10.0 W/m 2 K
Without gas injection on the sidewall, a relatively small sideward heat transfer
coefficient was assumed as:
hsIde
1.0 W/m
57
2
K
In the experiments, the sideward heat transfer area A,
upward heat transfer area At,
is comparable to the
but the temperature drop in the sideward direction
((T. - TLJ) < 1.0 K) is far less than that in the up vard direction ((Ttrn - TT) ~_
10.0 K). Therefore, the sideward heat addition rate, Qside, can be neglected in
the energy balance calculation. In equation (2.30), it is assumed that the injected
air is thermally equilibriated with each layer.
A major error in the experimental data reduction can be caused by the uncertainty of the measured temperature difference (TU - TL). The accuracy of the
temperature measurement is ±0.05 K while the temperature difference measured
at the high gas flux test can reach as low as 0.1 K.
However, the results of the
experiments shown in Table 2.4 are calculated from the measured temperatures
averaged over a time interval of 20 minutes (one measurement every two minutes).
Comparisons of the various interfacial heat transfer models to the cyclohexanewater experiment data are shown in Fig. 2.8. It is seen that both the modified
Szekely and Lee and Kazimi models agree reasonably well with the experimental
data. The Greene's model overpredicts and the modified Konsetov correlation
underestimates the experimental data.
At low gas velocity (Jg < 10 mm/s), the experimental data disperse over a
wider range and seem to deviate from the trend of the experimental data obtained
at higher gas fluxes.
This is caused by the non-uniform pattern of the rising
bubbles observed in the low gas flux test. Unfortunately, most of the bubbles were
generated at the peripheral region of the pool when the injected gas rate was low.
At the center region of the pool, where the thermocouples were located to measure
the layers temperatures, the bubble-induced internal circulation was quite limited
because of low bubble frequency. Therefore, the non-uniform bubble pattern will
result in a higher thermal gradient, i.e. a higher temperature difference (TLT - TL),
which gives a lower interfacial heat transfer coefficient than one can expect from
a pool with uniform bubble distribution.
Table 2.4
Interfacial Heat Transfer Tests
jg
-MC
UJj
Q,
Qair
Qdown (Tu - TL)
h
(W) (W)
(W)
(K)
(W/m 2 K)
5.5
0.31
10.4
0.733
258
4.97
5.5
0.27
10.2
0.626
296
6.67
6.25
5.5
0.22
11.5
0.459
455
HT-4
7.30
5.10
5.5
0.18
10.4
0.343
551
HT-5
7.93
5.10
5.5
0.10
10.5
0.177
1078
HT-6
9.19
6.89
5.5
0.42
12.0
0.636
341
HT-7
9.19
6.38
5.5
0.12
11.8
0.182
1174
HT-8
9.19
6.25
5.5
0.09
11.7
0.131
1611
HT-9
9.19
7.52
5.5
0.10
12.9
0.152
1548
HT-10
9.84
6.89
5.5
0.22
12.2
0.318
694
HT-11
11.2
8.54
5.5
0.31
13.7
0.394
632
HT-12
11.8
6.63
5.5
0.11
12.0
0.136
1598
HT-13
12.4
7.78
5.5
0.27
13.0
0.303
780
HT-14
12.4
10.2
5.5
0.35
15.4
0.394
707
HT-i5
12.4
7.65
5.5
0.11
13.0
0.126
1873
HT-16
12.4
11.0
5.5
0.18
16.3
0.202
1463
HT-17
13.8
9.31
5.5
0.12
14.7
0.121
2198
HT-18
14.4
9.56
5.5
0.23
14.8
0.229
1173
HT-19
15.7
9.44
5.5
0.21
14.7
0.192
1393
HT-20
17.2
11.6
5.5
0.22
16.9
0.182
1686
HT-21
17.2
8.80
5.5
0.24
14.1
0.192
1330
HT-22
19.5
11.3
5.5
0.19
16.6
0.139
2174
HT-23
19.5
8.54
5.5
0.14
13.9
0.101
2497
HT-24
21.0
7.14
5.5
0.20
12.4
0.136
1655
HT-25
21.0
11.4
5.5
0.22
16.6
0.146
2060
HT-26
21.8
8.16
5.5
0.16
13.5
0.106
2309
HT-27
22.5
10.8
5.5
0.23
16.1
0.146
1995
HT-28
25.0
12.9
5.5
0.24
18.1
0.136
2413
HT-29
27.4
16.9
5.5
0.26
22.2
0.132
3050
HT-30
27.4
14.8
5.5
0.20
20.1
0.101
3605
Test
(mm/s)
(W)
HT-1
6.04
5.24
HT-2
6.04
HT-3
59
%%
105
Modified Szekely
- --
E-
Z--
. Greene
Lee & Kazimi
Modified Konsetov
~
Experiment
--
10
C
z
10 3
0
5
10
15
20
SUPERFICIAL GAS VELOCITY
2z
Figure 2.8
25
(mm/s]
Interfacial Heat Transfer between Water and Cyclohexane Layers
30
2.4.3 Layer Mixing
The layer mixing tests are listed in Table 2.5. In these tests, four different
liquid patterns were observed in both LM-1 and LM-2 tests with different gas
injection rates: (1) At
j
< 15 nm/s, the two liquid materials stay separate.
Small amounts of water drops were entrained into the upper cyclohexane layer
when the gas bubbles penetrated through the layer interface. The entrained water
droplets then fell back into the water layer due to their higher density. Basically,
a clear water/cyclohexane interface was observed at this low gas velocity, and
both layers were in the regime of bubbly flow. (2) As J, ranged from 15 to 30
mm/s, the rate of water droplet entrainment was high enough to create a mixing
layer sandwiched between the water and cyclohexane layers. The thickness of the
mixing layer increased with increasing superficial gas velocity. (3) As
jg
reached
30 to 50 mm/s, the pool exhibited a two-layer pattern: a well-mixed layer (in
churn flow regime) formed on top of a pure water layer (in bubbly flow regime)
while distinguished cyclohexane layer no longer existed.
(4) At jg > 50 mm/s,
the two liquids were violently agitated and mixed into one homogenous layer. A
churn flow pattern was observed in this gas velocity range.
Based on the correlation (equation 2.14) proposed by Kataoka and Ishii to
characterize the transition between the bubbly flow and churn-turbulent flow, the
critical superficial gas velocities of the water-air and cyclohexane-air systems are
53.0 and 43.0 mm/s, respectively. The flow regimes observed in the experiment
are in good agreement with these predictions. The observed mixing phenomena
discussed above seem to indicate that the two immiscible layers are entirely mixed
when the superficial gas velocity is higher than the critical velocities of both layers.
In the cooling tests (LM-3,4,5,6 tests), it was found that the mixing patterns
were not affected by the solidification of the pool materials. Since no boundary
crust was formed at the liquid/liquid interface, the extent of layer mixing was not
disturbed by the freezing phenomena.
61
Table 2.5
Layer Mixing Tests
Water
Cyclohexane
jg
Test
(kg)
(kg)
(mm/s)
Cooling
LM-1
2.0
1.56
6.0-126.0
No
LM-2
2.0
2.34
6.0-126.0
No
LM-3
2.0
1.56
6.0
Yes
LM-4
2.0
1.56
21.0
Yes
LM-5
2.0
1.56
45.0
Yes
LM-6
2.0
1.56
65.0
Yes
62
2.4.4 Droplet Entrainment
The numbers of the K 2 SO4 particles with different particle sizes in the gas
stream were counted by the optical particle counter. Each size particle was counted
ten times with the counting period of two minutes. Averaged values (counts/min)
of different particle sizes are shown in Table 2.6. It is seen that the detected solid
K 2 SO
4
particles had a median size of 0.5 pmp and a maximum size of 5.0 pm. The
particle size of the entrained liquid droplet can be calculated by:
-11/3
DH
20
1
_
.f
where
f
20
PK 2 SO
-
PH
4
DK2SO 4
(2.32)
J
is the weight fraction of K 2 SO4 in the aqueous solution, and D is the
diameter of particle.
Therefore, the median and maximum sizes of the water
droplets entrained by the bubbly flow (j9 = 8.0 mm/s) were about 2.0 and 20.0
pm,, respectively.
The amounts of collected potassium sulfate in the filter collection tests are
listed in Table 2.7. The dimensionless parameters of these tests are also shown
in this table. Comparison of the data with the Kataoka and Ishii correlation is
presented in Fig. 2.9. It is seen that the differences between the experimental data
and-the prediction of entrainment are within a factor of four.
2.5 Summary and Conclusions
The freezing phenomena experiments conducted in this study are of scoping
nature. It was observed that a bottom crust could be formed across the bubble agitated horizontal liquid/solid interface, with gas velocities up to 126 mm/s. This
observation confirms a freezing model assumption used in the current MCCI integral analysis codes. However, the liquid/liquid interface crust also assumed in the
analysis codes was not formed in the simulant experiments. The stability of a top
crust is also called into question by the observations of this experiment. Therefore,
the freezing of the oxidic layer involved in the MCCI could be in the slurry form
rather than a crusting
boundarv. In addition,
63
the supercooling plhenonenon
Table 2.6
Counts per Minute of Different Particle Sizes
Particle Size (pim)
Test
DE-Al
0.3-0.5 0.5-0.7 0.7-1.0 1.0-3.0 3.0-5.0 5.0-7.0
7.0-
504703 109942
27455
9169
371
16
6
479706
73422
12357
3230
125
13
7
24997
36520
15098
5939
246
3
-1
(Aqueous Solution)
DE-A2
(Pure Water)
Difference
64
Table 2.7
Filter Collection Tests
H
Collected Water
J*/H*
E*
9.38 x 10-6
Test
(mm/s)
(mm)
(mg/s)
DE-B1
16.4
180
0.0108
5.44 x10-5
DE-B2
16.4
180
0.0102
5.44 x 10-
5
8.80x 10
DE-B3
19.1
180
0.0517
6.34 x 10-
5
3.84x10-5
DE-B4
32.9
180
0.177
1.09
10-
4
7.62x 10~ 5
DE-B5
46.9
180
0.248
1.56x 10-
4
7.51x 10-5
DE-B6
70.1
180
0.492
2.33x10-
4
9.96x10-5
DE-B7
82.6
320
0.0908
1.54x 10-
4
1.57 x 10-5
65
x
6
bA
bi
10-5_
10
6
10-5
I
10-4
I
11
1
J9*/H*
Figure 2.9
Water Droplet Entrainment from the Bubbling Pool
66
observed in the experiments contradicts the assumption used in predicting the
timing of freezing.
The interfacial heat transfer between the water and cyclohexane layers was
measured under various superficial gas velocities. Comparisons of the data with
the existing models were made. The modified Szekely model used in the WECHSL
code agrees well with the experimental data. While the Greene model incorporated
in the CORCON/MOD2 seems to overpredict the experimental results.
In the mixing test, it is found that two immiscible liquids with density ratio of
0.78 were entirely homogenized under a modest superficial gas velocity of 50 mm/s.
The transition patterns of the mixing phenomena with different gas velocities were
observed as well.
Liquid droplets entrained by the flowing gas were quantified by the experiment. The median and maximum-sizes of the water droplets entrained by a gas
flow of j 9 = 8.0 mm/s were found to be 2.0 and 20.0 pm, respectively. The amount
of entrainment was found in good agreement with the prediction by the Kataoka
and Ishii model.
67
CHAPTER 3
DOWNWARD HEAT TRANSFER MODEL FOR
THE MELT/CONCRETE INTERACTION
3.1 Introduction
The behavior of a pool of molten core materials in a concrete cavity is governed by a simple energy balance.
The decay heat and chemical reaction heat
generated in the pool may be lost through its top surface to the containment
atmosphere and containment structure or to the surrounding concrete. The partition of energy between concrete and the top surface is determined by the various
thermal resistances from the pool of molten core materials to the surroundings.
Heat transfer phenomena involved include: radiative heat loss from hot corium
boundaries, convective heat transfer within the internally heated pool, heat transfer between immiscible liquid layers with bubble agitation, heat transfer at eroding
interfaces with gas injection and conductive heat transfer through solidified crust.
All these heat transfer processes will, in fact, affect the corium temperature in a
coupled fashion.
Among various phenomenological heat transfer models, the one having the
most direct impact on the core-concrete interactions process is that describing
the heat transfer across the melt/concrete interface. The extent of concrete ablation, the melt temperature response, the amount of decomposition gas release and
therefore the amounts of chemical heat and aerosol releases, all directly depend
on the amount of heat that can be transferred across the melt/concrete interface.
The downward heat transfer is driven by the temperature difference between
the molten core materials and the concrete. Vigorous agitation of the melt by
concrete decomposition gases is expected to enhance the convective heat transfer
68
process. Besides the decomposition gases, melting concrete (slag) generated beneath the corium pool will be buoyed up due to its relatively low density, and will
also affect the downward heat transfer.
As the pool cools down, formation of a bottom crust provides an additional
thermal resistance to the downward heat flow path.
This conductive thermal
resistance will limit the amount of heat loss to concrete. For some accident scenarios, the core debris may initially be solid or partially solid. If the degree of
solidification is such that internally generated heat cannot be removed, the debris
temperature will rise and the core materials will melt until the convective heat
transfer is sufficient to allow a thermal balance to be achieved.
Both analytical and experimental efforts have been reported on the heat transfer between the molten core and the concrete. M. Plys [P2] and M. Lee [L2] presented excellent reviews of these efforts. In this chapter, only a brief review of the
previous downward heat transfer work will be presented.
In Fig. 3.1, schematic diagrams of the melt/concrete heat transfer in the downward direction are shown. The principal components involved in the process are:
molten core materials, an underlying concrete, a solidified bottom crust if present,
and an interface region which comprises the decomposed concrete materials. The
various analytical models that have been proposed to predict the downward heat
transfer rate can be divided into two types: a film boiling-like model and a nucleate boiling-like model. The major difference between these models concerning the
interface region is based on the observations of different simulant experiments.
The objectives of this study are to develop downward heat transfer models for
various stages of the interaction, and to incorporate them into CORCON/MOD2
so as to analyze the integral behavior of the melt/concrete. The new version of
CORCON/MOD2 with additional downward heat transfer models will be referred
to as CORCON/MIT. Important factors affecting the calculation of the downward
69
0
0
0
O0
MOLTEN POOL
0
O0
Figure 3.1
0
1*1j
0
0
0
0
O0 0
0
O0
0
0
0
0~ 0
0
0
Illustration of the Downward Heat Transfer of the Core/Concrete
Interaction
70
heat flux will be discussed and presented in the following sections. Validation of
various models based on real material experimental data will be presented in the
next chapter.
3.2 Review of the Downward Heat Transfer Models
3.2.1 The Gas Film Model
Prototypic melt/concrete interaction tests and concrete decomposition experiments showed that large amounts of gases are produced as the concrete dehydrates
and decomposes while it is exposed to a thermal attack [P1,P5]. Separate effects
experiments using simulant materials for the molten pool and decomposing concrete have been performed by placing a horizontal slab of dry ice beneath a pool
of warm water or benzene. These simulant experiments indicated the existence of
a gas film at the pool/solid interface [D1,A3].
In view of these observations, an
assumption was made that the melt/concrete interface consists of a gas film comprised of concrete decomposition gases during a MCCI. Based on this assumption,
a gas film model was developed and used in the CORCON and WECHSL codes
to describe the heat transfer between the molten core and concrete.
As shown in Fig. 3.2(a), interfacial heat transfer across the gas film is composed of radiative (q/ad) and convective (q'
,) processes. The radiative flux across
the interface is given by the relation:
qrad = FCTB(Tj - T6)
where Oc
(3.1)
is the Stefan-Boltzmann constant and T, and TD are the core melt
surface and ablating concrete wall temperatures, respectively. The radiation form
factor, F, for two infinite parallel plates is defined as:
F-
(3.2)
1
1
ep
6c
where e and c, are the emissivities of the molten pool and concrete, respectively.
71
1
00
'~
o.
0
U
0
0
0
0
0
0
0
0
0
0
0
MOLTEN POOL
0
0
0
S0
oAL
o
CRUST
-mk 0
-1
6AS FILM
CONCRETE
(a)
(b)
Figure 3.2 Analytical Picture of the Gas Film Model
,i*
0
0
0
0
o
0
0
0
0
0
000
1lilllilllllIIIIIIIIIIIli'illilllll11111lIIIIIIIIII
0
The convective heat flux across the gas film is given by:
(3.3)
q"onv = heonv(T1 - TD)
where the interface heat transfer coefficient, hconv, is described using Taylor instability models for a horizontal surface developed by Dhir et al. [D1], or Alsmeyer
et al. [A3]. The dimensionless convective heat transfer coefficient for this gas film
model is given by:
NUB
= CoRe-"
3
(3.4)
where
NuB=
;L_-1/
= pcnL;Re
k9
y,
gp9(pt -
pg)
and the Laplace constant A is defined as:
A
[g
L
]1
2
(3.5)
g( pt - p, )
where j, is the superficial gas velocity, and subscripts f and g refer to liquid pool
materials and gas film properties, respectively. The constant Co in equation (3.4)
is 0.326 and 0.256 for the Alsmeyer and Dhir models, respectively.
The downward heat flux is then given by the combination of the radiative and
convective heat fluxes as:
=q#d
to~0
(3.6)
However, there are still two unknowns which cannot be determined explicitly in
the calculation of the downward heat flux. One is the superficial gas velocity
j,
another is the interface temperature TI. The superficial gas generation velocity in
the melt/concrete interaction is defined by:
qI#
j=
d**
P 9 HDecomp
73
(3.7)
where x is the weight fraction of the gas content in concrete and HDecomp
is
the decomposition enthalpy of concrete. The jg cannot be determined before the
downward heat flux is known, while the downward heat flux can be calculated
only if the j 9 is given. An iteration is needed to calculate both the downward heat
flux and superficial gas velocity.
As for the interface temperature TI, it can be found by applying continuity of
heat flow at the core-melt surface. Heat transfer from pool bulk to its periphery
is equal to that transfer across the gas film (see Fig. 3.2(a)), i.e.
don
qFrad
+
nv
it
= "oot
(3.8)
where the pool internal heat transfer is given by:
"o= h, 00I(Tp - TI)
(3.9)
and Tp is the pool bulk temperature. Models describing the heat transfer coefficient hpoo, were developed using available data and correlations from simulant
experiments [B4].
The one used in the CORCON/MOD1 code is a modified
Konsetov correlation [K4] which combines the effects of natural and gas driven
convection. It is given by:
hpool = k (Pr+)
1/3
[0.0003,3aT +
0.4a 2]
1/3
(3.10)
where the unsubscripted variables are liquid pool properties and AT is the temperature difference between the pool bulk and its periphery (Tp - TI). The gas
volume fraction a is given by:
a=
B
1 - B'
B =
74
__
g
1.53
g
(3.11)
(.1
where jg is the superficial velocity of the gas entering the pool and A is the Laplace
constant defined in equation (3.7).
Since the release of CORCON/MOD1, Ginsberg and Greene [G8] have compared simulant data for horizontal liquid/liquid interface subjected to a gas flux
with the predictions of several models and concluded that the Konsetov forms of
CORCON/MOD1 seriously underpredicted the data. In fact, for very thin layers,
CORCON/MOD1 would sometimes predict convective heat transfer rates smaller
Therefore, in the CORCON/MOD2, an
than would occur by conduction [C3].
empirical correlation developed by Greene, accompanied with a conduction limit,
was used to predict the pool internal heat transfer. The Greene's correlation is
given by:
hp
=
5.05
rb
Reo' 5 Pr"'8
(3.12)
where the Reynold number is defined as :
Re = pajrb
(3.13)
and rb is the effective radius (based on volume) of the bubbles in the pool. This
model is similar to a so-called surface renewal model developed by Szekely [S6].
Szekely's model was derived from the idea that heat is transferred by transient
conduction with bubbles periodically disrupting the developing thermal gradients.
This model was adopted in the WECHSL code, and is given by:
hpoo = 1.69-
rb
(Re - Pr)0 '5
(3.14)
Based on the pool side heat transfer correlation, coupled with the heat transfer
coefficient across the gas film, the interface temperature T 1 can be calculated
implicitly.
During the post-freezing stage, a bottom boundary crust or an entirely solid
layer sitting on a blanket of gas film is again assumed in the CORCON model.
For
a liquid pool with a solid crust, as shown in Fig. 3.2(b), heat transfer in the liquid
region is governed by convection (natural or bubble-enhanced) with conduction
as a limit. In the solid region, heat transfer is governed by conduction.
Heat
transfer across the gas film is again described by a combination of the convective
and radiative processes. A simplified procedure has been employed in the CORCON/MOD2 to construct a steady-state solution to the heat transfer equations in
a right circular cylinder whose average temperature, boundary temperature, crust
and layer thicknesses, and volume all match those of the actual layer. Detailed descriptions of the post-freezing model used in the CORCON/MOD2 can be found
in Ref. [C4].
In the gas film model, the effect of the slag has been neglected. Heat transfer
across the interface region depends only on the thermal properties of the released
gases. For the pre-freezing stage, the gas film provides a major thermal resistance
in the downward heat transfer. Typical values of the downward heat fiux based on
the gas film model for different melt temperatures are shown in Table 3.1. It can
be seen that when the melt temperature is high, the radiative process dominates
the downward heat transfer. At lower temperatures, the fraction of the convective
part will increase.
3.2.2 The Periodic Contact Model
In an experiment by Felde et al. [Fl] using gas injection through a porous
plate into a volumetrically heated liquid pool, no continuous gas film was identified
at modest superficial gas Velocities, 0.0 ~ 20.0 mm/s. In an experiment at M.I.T.
[L2], air was injected through a porous plate into a water or cyclohexane pool,
and heat was removed from the bottom of the pool by a condensing unit. The
superficial gas velocity of the experiment ranged between 0.0 and 130.0 mm/s. No
gas film was observed at the liquid/solid interface. Based on these observations,
it is believed that a stable gas film will exist only if the superficial gas velocity
exceeds a certain limit, and one could not expect to have a stable gas film under
76
Table 3.1
Downward Heat Transfer Calculated by the Gas Film Model
Metallic Pool/Limestone Concrete: (TD = 1750 K)
{,
Tp
T1
heon,
(K)
(K)
(W/m 2 K)
hpoi
(W/m 2 K)
2950 2941
2.878 x 102
2.566 x 105 3.428 x 105
1.936 x 106 2.279 x 106
2750 2742
3.175 x 102
2.225 x 105 3.151 x 105
1.396 x 106 1.711 x 106
2450 2444
3.765 x 102
1.744 x 105 2.613 x 105
7.781 x 105 1.039 x 106
2050 2046
5.223 x 102
1.098 x 105
1.548 x 105
2.413 x 105 3.961 x 105
1850
7.408 x 102
6.648 x 104
7.252 x 104
6.749 x 104
1848
q"
(W/m 2 )
q'r'ad
(W/m
own
2
)
(W/m
2
)
1.390 x 105
Oxidic Pool/Limestone Concrete: (TD = 1750 K)
Tp
TI
hconv
hpooi
qq','aod
(K)
(K)
(W/m 2 K)
(W/m 2 K)
(W/m 2 )
2950
2939
3.844 x 102
2.093 x 105 4.569 x 105
1.929 x 106 2.386 x 106
2750
2740
4.231 x 102
1.896 x 10 5 4.191 x 105
1.391 x 106 1.810 x 106
2450
2442
4.993 x 102
1.468 x 105 3.457 x 105
7.752 x 105 1.121 x 106
*2050 1855
9.405 x 102
-
9.843 x 104
7.256 x 104 1.710 x 105
1753
2.510 x 103
-
7.296 x 103
3
1.848 x 103 9.144 x 10
*1850
* Entirely solidified pool.
77
(W/m
2
)
(W/m
2
)
certain MCCI conditions.
The gas injection into the pool can be viewed as being analogous to the
Based on this analogy and experimental
nucleate boiling of a saturated pool.
data, Felde et al. developed purely empirical correlations for the downward and
sideward heat transfer. The one for the downward heat transfer is of the following
form:
kg
h = 5.69-
ptj 3
0
A (gyt
(3.15)
The direct application of this model to the analysis of melt/concrete interaction is
questionable. The correlation was obtained by empirical curve fitting based on the
data of an experiment without having the eroding phenomena -
a major physical
process of MCCI. Nevertheless, their work suggested that a nucleate boiling-like
process may occur at the horizontal melt/concrete interface.
A periodic contact model developed at M.I.T. [L3,L4] was proposed to govern
the heat transfer process when a gas film cannot be sustained at the interface. The
periodic contact model considers the heat transfer mechanism as a transient heat
conduction process between the hot pool and the relatively cold concrete surface.
As the decomposed gas and molten slag rise up and away from the surface due
to the buoyancy force, the interface will be stirred and some hot liquid will be
periodically (based on the bubble departure frequency) brought into contact with
concrete surface. This model includes the slag effect by treating the decomposed
material as a two phase rising fluid which is contained in the interface region.
Thermal properties of this rising fluid are calculated from the volume-averaged or
weight-averaged values of the slag and decomposed gases.
The physical picture at the melt/concrete interface during a completed bubbling cycle assumed by the periodic contact model is shown in Fig. 3.3. As indicated in the figure, the downward heat flux across the pool/rising fluid interface is
represented by q". It can be expected that the downward heat flux is characterized
78
0
0
O0
molten pool
0
O
0
0
Oo
o n 0I
I
Figure 3.3 Analytical Picture of the Periodic Contact Model
79
by a transient behavior during the growth of the rising fluid in each bubbling cycle.
In the periodic contact model, a time-averaged downward heat flux q' was defined
as:
q' = hPc(Tp
-
(3.16)
TD)
The dimensionless downward heat transfer coefficient of the h-c is given by:
[
,f c(Te
Nuf = Coo
TD)
HDecomnp
f(
p
TD
(3.17)
where:
Nuf = hpcA/kf;
A=
1/2(3.18)
- pp)
g(p
yo =
-
TP -
TI;
Of .TP - TD_
=
(3.19)
/pkc
and the effective decomposition enthalpy, HeCmp is defined as:
Hbecomp = HDecomp + cf - (TI2TD)
Subscripts p,
f
(3.20)
and s used in these equations refer to liquid pool materials, rising
fluid and solid concrete, respectively. In equation (3.19), the T, represents the
average temperature of the pool/rising fluid interface during each bubbling cycle,
and it is given by:
TP TP
I
[Tp-T Of(321
Op +Of
TD
1
1(3.21)
which leads to:
Yo=
I+ O
(3.22)
Thermal properties of the rising fluid are defined by:
Pf =Pgas - C
+ pslag (II - a)
80
(3.23)
+ kslag
k 1 = kgasCf = Cgas
-
+
Cslag
(3.24)
(-a)
(1
-
x)
(3.25)
where X is the weight fraction of gas content in concrete, and the void fraction a
of the two phase rising fluid is defined by:
=
1-x~
1+
X
(3.26)
1
pga.,
-S
Pslag
with slip ratio
S=
(PL''g)1/3
Pgas
(3.27)
During early development of the periodic contact model, without having real
material experimental data, the constants appearing in equation (3.17) have been
determined by a least square curve fitting of the simulant experiment data of Dhir
et aL.. This procedure leads to
Co = 767
n = 1.53
m = 0.32
(3.28)
As mentioned in Ref. [L3], a correction factor (k,/k,) was applied in equation
(3.17) without theoretical justification to fit both the Water/Dry Ice and Benzene/Dry Ice experimental data. This could be a questionable term, since in the
dry ice experiments only one material (dry ice) was used as an eroded substrate.
It is therefore impossible to obtain the proper dependence of the downward heat
transfer coefficient on the thermal conductivity of the eroded material (k,) based
on the dry ice experimental data. On the other hand, it seems unreasonable that
the downward heat transfer coefficient based on a conduction mechanism be inversely proportional to the thermal conductivity of the substrate. Most important
81
of all, it is unpersuasive that the periodic contact model is obtained by fitting the
data of those experiments which have observed phenomena in contradiction to the
assumption that has been made in the model development.
When data of the first four BETA experiments became available, the constant
Co was reduced by a factor of 0.6 in order to best fit into the observed downward
concrete penetration distances [L4]. The constants n and m were kept the same
as those in equation (3.28).
The periodic contact model has been incorporated into the CORCON/MOD1
to analyze the integral behavior of the melt/concrete interaction. It is found that
there are significant differences between the gas film and the periodic contact models in the predictions of the concrete ablation, gas generation and melt temperature
response [K5,L4].
The periodic contact model was developed based on the assumption of direct
periodic contact of the liquid melt with the concrete surface. While the debris
cools down, direct contact will be prohibited by the formation of a bottom crust.
Therefore, the applicability of the periodic contact model is limited to the early
stages of the MCCI.
3.2.3 The Film Collapse Model
Both the gas film and the periodic contact models are conceptualized based
on presumable physical phenomena which cannot be directly observed in the real
material experiments.
The actual heat transfer mode may depend on the melt
temperature and is not well understood.
In the BETA experiments (VO.2 and
V1.2), in the case of high initial melt temperature (2473 K), the downward erosion
rate was initially somewhat limited, and a significant increased erosion rate was
observed after a period of time (see Fig. 3.4). Based on this observation, a socalled film collapse model was proposed [K5]. It assumes that the downward heat
transfer may follow a combination of the gas film and periodic contact models. If
82
BETA EXPERIMENT VO.2
500
I
I
I
i
400
/1(0
- 300
0
o200
100
0
0
500
1000
1500
2000
2500
TIME [s]
Figure 3.4
Downward Ablation Distance of BETA Test VO.2
3000
the melt initial temperature is high enough, a stable film may exist, and the heat
flux presumably follows the gas film model until a minimum stable film limit is
reached. When the film collapses, the heat transfer mode undergoes a transition to
the periodic contact model. If the melt initial temperature is too low to generate
a stable film, the heat transfer will be governed by the periodic contact model all
the time until the solidus temperature is reached.
The transition criteria of a stable gas film used in the film collapse model
are based on hydrodynamic considerations. The film establishment criterion was
related to the Kutateladze's flooding limit which is given by [K3]:
with
(3.29)
o/pg A
(jg)ilm = K,
{
30.0 M 2 / 3
K 6.3 M 2 / 3 Ar "/6
Ar > 104
Ar < 104
M 2 =pggA/P
Ar = g A 3 1v
A = [o-/g(p1
2
pg)]0.5
-
where P is the system pressure and subscripts
e and
g refer to liquid and gas,
respectively. The film collapse limit was related to Berenson's minimum gas flux
to stabilize the film, and is given as [B9];
(jg)min =
- 0.25
0.09 [o!(Pe Pg)
. ( pt + pg ) _
(3.30)
For the case of melt/concrete interaction, those limits differ by two orders of
magnitude.
In the early development of the film collapse model, multiplication
factors (Ai
and MB) were applied to both limits in order to get the best fit of
the early BETA experimental data.
84
For a given initial melt temperature, the superficial gas velocity calculated
by the periodic contact model is to be compared to the Kutateladze's limit to
determine the existence of an initial stable gas film. If the following relation
(3g)c ;> MK*
(ig)
ilm
(3.31)
is true, then the downward heat transfer would be calculated by the gas film model.
Otherwise, the periodic contact model is used. The superficial gas velocity (j,),
appearing in equation (3.31) is calculated by:
(ig)
=
P9
H
Decomp
-gx
(3.32)
For an initially stable gas film, film collapse is assumed when the superficial gas
velocity of the gas film model falls below a certain limit:
(jg)p _<MB - (j,)m
(3.33)
as the melt cools down. If the relation (3.33) is true, then the downward heat
transfer process switches from the gas film model to the periodic contact model.
When the film collapse model was first implemented into the CORCON/
MOD1 to analyze the early BETA experiment, the value of MB was found to be
6.0, while the multiplier MK had to be varied among different experiments (0.85
for VO.3 and 0.5 for VO.2 and V1.2) in order to obtain a predicted stable gas film
at the beginning whenever experimental data showed such a trend.
3.3 Model Development and Implementation in CORCON
3.3.1 Revised Periodic Contact Model
When more data became available from the BETA facility, it was found that,
in general, the periodic contact model somewhat overestimates the downward-penetration distance of the concrete [K6,K10]. A revised periodic contact model based
85
on theoretical considerations is developed to overcome a deficiency found in its
original derivation and improve its accuracy.
3.3.1.1 Basic Definition
In the revised periodic contact model, the downward heat transfer coefficient
is redefined as:
hPc =
"""n)
c
(3.34)
(Tp - TD)
Variables appearing in this equation are indicated in Fig. 3.5(a).
Compared to
the original definition in equation (3.16), it can be seen that the downward heat
flux across the pool/rising fluid interface, qg, is replaced by the one across the
rising fluid/concrete interface, (q'jow),,, in equation (3.34). The downward heat
transfer coefficient ho, as shown in Fig. 3.5(a), is then defined as:
(3.35)
g -
ho =
(Tp- T1 )
The relation between q" and (q"o
1
),, can be found by accounting for the energy
needed to heat up the rising fluid layer above the decomposition temperature:
= qg
-
(j)M,,pc
where the average superficial rising fluid velocity
(jf)A
-
jf
TD)
(3.36)
is given by:
""
= p fAvg
HDecomp
(3.37)
Substituting equation (3.37) into (3.36), one can get:
(qqdown) pc _n.
o
[HDecomp
H*Decomp
(3.38)
where:
H*ecomp
HDecomp
+
Cf
(TI-TD)
(3.39)
The effective decomposition enthalpy Heccomp is obtained by assuming a parabolic
temperature profile, instead of a linear profile, exists in the rising fluid layer.
86
0
0
0
0
o0
0
0
molten pool
0
0
0
0
o
0
oU
o0
0
0
0
crust
0o
rising fluid
concrete
(a)
(b)
Figure 3.5 Analytical Picture of the Revised Periodic Contact Model
NA W"
Combining equations (3.34), (3.35) and (3.38) and defining a new parameter
r as follows:
(3.40)
r = HDecomp
Hecomp
gives:
-
-P
hP = ho
IT
(3.41)
r
- TD
3.3.1.2 Transient Heat Conduction
Based on transient heat conduction, the amount of heat transfer out of the
pool (through the pool/rising fluid interface) per unit area within a certain time
period
td
can be obtained as:
q
2p c
(3.42)
aPt(Tp-$I)
where a, is the thermal diffusivity of the pool material and
td
is the bubble
departure time. Therefore, the averaged downward heat flux qg as defined can be
written as:
q
to
q' = -
(3.43)
Combination of equations (3.35), (3.42) and (3.43) leads to:
ho = C1 . 3
(3.44)
where C1 is a constant and , = V/(pkc),.
3.3.1.3 Bubble Dynamics
As discussed in Ref. [L3], the bubble departure time can be obtained by the
integration of bubble growth rate and rising fluid generation rate, which gives:
td o(
^ d
L2If
88
(3.45)
The bubble departure radius Rd was derived based on the balance of surface
tension and buoyancy force as:
[r
1/2
Rd oc
=
(3.46)
A
g(p, - p5 )]
where A is the Laplace constant defined in equation (3.18).
The length L is the
dimension of a square area influenced by single bubble, and it was assumed to be
proportional to the Taylor wavelength as:
[~
L oc [ g(pp
11/2
)
(3.47)
-A
Substituting equations (3.46) and (3.47) into (3.45) and combining the result with
equation (3.44), leads to:
ho = C 2 3,
(
Oj A
(3.48)
"f)"Ag
where C 2 is a constant.
Based on equations (3.37) (3.38) and (3.40), the superficial rising fluid velocity
can be rewritten as:
(if )Avgpf H&comp = ho(Tp - T 1 )
(3.49)
or
(if )Avg =
(3.50)
ho(Tp - T)r
pf HDecomp
Substituting equation (3.50) into (3.48), after some manipulations, the heat transfer coefficient ho can be cast into the following form:
ho = Co(k,
A
p
cT
pf
89
-
HDecomp
r
I
(3.51)
Substituting this expression into equation (3.41), gives:
h)1
[CTP
A
Of Tp
- TD ..
-
TD)
Decomp
1
2
(3.52)
This equation can then be rearranged into a dimensionless form:
e
Nuf = CyO
(3.53)
TD)
Decomp_
where
Nuf = hpcA/kf;
yo
O [T-TI
Tp -TD
Of
1
3.3.1.4 Interface Temperature
As shown in equation (3.21), the average interface temperature Ti is obtained
by assuming direct contact between semi-infinite slabs of the pool material and
rising fluid. If one assumes that the pool material is instead brought into contact
with the solid concrete, the rising fluid property of should be replaced by the
concrete property /, in equation (3.21) to obtain the interface temperature, i.e.
[Tp -Tr
1
J
TP - TDI
ITP-TD
_
___
13p 13(3.54)
+ 0.s
However, due to the phase change of the eroded concrete in a periodic cycle, neither
equation (3.21) nor equation (3.54) gives proper representation of the average
interface termperature. Nevertheless, these two equations impose an upper and
lower bounds on the interface temperature. Therefore, one can employ a weighting
factor F for estimation of the interface temperature as:
T- TDK
TI
TP1.s
where 0.0 < F
F(
p
(1.0 - F)( 13)
+Of ) +
Op +
< 1.0.
90
(3.55)
In the original model, the F factor in effect was 1.0, which gives an upper
bound estimation of the interface temperature. For the new model, the F factor
is taken as:
F = 1.0 - r
(3.56)
where r is defined in equation (3.40). The weighting factor F is selected based on
the following argument. As the pool temperature drops, the interface temperature
T1 will approach TD, and little concrete will be decomposed. Under this condition,
the weighting factor F will approach zero. Therefore, the interface temperature
Tr will be determined as if the pool material is in contact with the solid concrete.
At higher pool temperature, more concrete will be decomposed. The increased
value of F leads to an increasing effect of the rising fluid on the determination of
the interface temperature. This leads to:
Tj = (1.0 - F)
3Tp +l,3TD
+
pTp + 3 fTD
3
+ F
p+ f(3.57)
An iteration is needed to determine Ti and r in equations (3.57) and (3.40),
respectively. The final form of the downward heat transfer coefficient of the revised
periodic contact model can be written as:
Nu = Co
2
cf (Tp -TD)]
(3.58)
+
(3.59)
with
7
(-
r
Equation (3.58) has been incorporated into the CORCON/MIT, and the only
unknown constant Co will be determined by curve fitting of the early BETA experimental data. The thermal conductivity ratio appearing in the original model
91
has been dropped and the constant n is kept at the theoretical value of 1.0 in the
revised model.
3.3.2 Transition Criteria of the Film Collapse Model
The first four sets of the BETA experiment were performed by February 1984.
The downward concrete penetration distances obtained from these tests have been
used as basis to determine the constants and multipliers appearing in the film
collapse and revised film collapse (combination of the revised periodic contact and
gas film models) models.
The experimental conditions as well as the measured initial (over the first
hundred seconds) downward penetration rates of the BETA tests are summarized
in Table 3.2. The measured downward penetration distances of those tests are
shown in Fig. 3.6. It is interesting to see that the initial downward penetration
rate of test VO.3 is quite different than those of tests VO.2 and V1.2 even though
all three tests have the same initial melt temperature. The V1.3 test has a lower
initial melt temperature, but double the initial penetration rate than those of
VO.2 and V1.2 tests.
Based on these observations, the assumption of the film
collapse model has been made. The transition criteria have to be determined by
distinguishing the differences among these tests.
In both the original and the revised periodic contact model, it can be seen
that the downward heat flux is barely affected by the density of the released gas.
However, examination of the existence of an initial stable gas film requires the
magnitude of the superficial gas velocity which is directly related to the released
gas density as shown in equation (3.32). In CORCON, the released gas density at
the bottom interface is calculated based on the ideal gas law:
Pg =
Pin
*
nRTint
(3.60)
where R is the gas constant and Pint and Tint are the bottom interface pressure
92
Table 3.2
Test Matrix of the Early BETA Experiments
Melt Composition Initial Melt
Oxidic Temperature
Test Metallic
(K)
(kg)
Planned
Power
(kW)
200
400
Average tErosion
Rate
Power
(kW)
(mm/s)
370
0.25
VO.2
Fe(300)
2473
VO.3
Fe(300) A1 2 0 3 (150)
2473
1700
1180
1.0
V1.2
Fe(200) A12 0 3 (150)
2473
Pulse
380
0.30
2173
1000
780
0.60
Fe(246) A12 0 3 (105)
V1.3
Cr( 30) SiO2 ( 45)
Ni( 24)
t
Average over the first one hundred seconds.
93
BETA EXPERIMENT
VO.2 TEST
VO.3 TEST
4 V1.2 TEST
-A V1.3 TEST
400
U
z
300
o200
-L
100
*
A
*A
0
-AL
0
0
A
A
100
0
300
200
400
500
TIME [s)
Figure 3.6 Downward Ablation Distances of Early BETA Tests
t
600
and temperature, respectively. Unfortunately, the original CORCON initializes
the interface pressure as atmospheric pressure without taking the pool weight
into account at the first time step. A modificatiort has been made in the COR-
CON/MIT by having:
Pt
= Patm + ppgH
where Pat, is the environment pressure and H is the pool height.
(3.61)
Given the
high density of the core melt, the pressure difference is significant in determining
whether an initial gas film exists.
A smaller pool mass will result in a lower
interface pressure and released gas density, thus leading to a higher superficial gas
velocity, which tends to stabilize the initial gas film.
Comparing the BETA VO.2, V1.2 and VO.3 experiments, all tests start with
the same temperature, but with different total masses. With the pressure modification in the CORCON/MIT, it can be predicted, with a unique multiplier MK,
that VO.2 and V1.2 would start with a stable gas film and VO.3 would start with
a periodic contact mode.
3.3.3 Post-Freezing Heat Transfer Model
The gas film is destabilized at a sufficiently low superficial gas velocity condition, which may occur earlier than the crust formation. The applicability of the
periodic contact model as discussed in the previous section is limited to the early
stages of the MCCI before any freezing occurs. Therefore, neither the gas film
nor the periodic contact model can be used to describe the downward heat transfer during the post-freezing stages of MCCL. In order to complete the downward
heat transfer model in the CORCON/MIT, a post-freezing model was developed.
Basically, this model assumes that thermal resistance between the pool boundary
and the concrete surface, i.e. resistance across the rising fluid, is continuous on
the basis of the superficial gas velocity that can be achieved after the formation
of a crust.
In Fig. 3.5(a), it is seen that the condition of a freezing crust initiation is:
T
where T
01
T30 1
(3.62)
.
is the solidus temperature of the molten pool materials. Imposing this
in equation (3.57), the following equation can
condition and replacing Tp by T
be obtained:
T
T
0
=
[1
- F) +
fF
+
,+Op
F
+ 0
o,+p
1- F
-
( OP
9
+
TD
(3.63)
)f
and a downward heat transfer coefficient h* based on the periodic contact model
at the condition of Tp = T; can then be calculated by equation (3.58).
Once the pool bulk temperature and its corresponding downward heat transfer
coefficient at the initiation freezing are obtained, the heat transfer coefficient across
the rising fluid can subsequently be defined as:
,
h*rf =
h*c(T -TD)
(Tsol - TD)
(3.64)
With this definition, the heat transfer coefficient, hrf, for the post-freezing stage
as shown in Fig. 3.5(b), is obtained by:
hrf = h*,-
(9
39
.. h*c(T
C=
-D
)0.5
(3.65)
where
-(X
PgHDecomp
and the superficial gas velocity
jg
for the post-freezing stage is defined by:
11
S.(3.67)
PgHDecomp
96
(3.66)
with the downward heat flux across the rising fluid/concrete interface given as:
q'oln
=
hrf(TI - TD).
(3.68)
The implementation of the post-freezing model in the CORCON/MIT is quite
similiar to the original gas film model.
Only the heat transfer across the gas
film (the combination of radiative and convective processes) has been replaced by
the heat transfer across the rising fluid described above. The pool internal heat
transfer coefficient hP,.I (see Fig. 3.5(b)) is determined by Greene's model, and
heat transfer across the solidified crust is described by heat conduction.
3.3.4 Summary
After several preliminary integral analyses of the BETA experiments using
CORCON/MIT, the constants and multipliers used in both the original and revised film collapse models were obtained by the best fitting to the experimental
data and are summarized in Table 3.3. Typical boiling curve of the film collapse
model, at both pre-freezing and post-freezing stages, with those specified transition limits can be seen in Fig. 3.7. The downward concrete ablation distances of the
BETA experiments predicted by the Film Collapse (FC), Revised Film Collapse
(RFC) and Gas Film (GF) heat transfer models compared to the experimental
data are shown in Figs. 3.8 through 3.11.
3.4 Downward Heat Flux Calculation
3.4.1 Cases Studied
Besides the downward heat transfer model, there are several parameters in
the MCCI which could affect the prediction on the magnitude of the downward
heat transfer. An analysis of the impact on the downward heat flux of some of the
parameters will be studied here. Some of the uncertain parameters are outlined
first.
97
Table 3.3
Empirical Constants of the Film Collapse Model
Original FC Model
Revised FC Model
CO
460.0
2.8
n
1.528
1.0
m
0.32
0.0
MK
0.65
0.40
MB
6.0
6.0
98
10
Kf K Concrete: T =1570 K
Melt Material: &eel Metal
9 r
108
LJ
106
0
105
0
800
400
Figure 3.7
1200
1600
TD[K]
Descriptive Downward Heat Flux of the Film Collapse Model
99
BETA EXPERIMENT VO.2
500
r
z
400
300
200
z
100
0 v
0
500
1000
1500
2000
2500
TIME [s]
Figure 3.8
Comparison between the Predicted and Measured Downward Erosion Distances of BETA Test VO.2
I~ N~
BETA EXPERIMENT VO.3
500-
400
z
300-
z
0
~0
o200
z
100
0
0
100
200
300
400
500
TIME [s]
Figure 3.9
Comparison between the Predicted and Measured Downward Erosion Distances of BETA Test VO.3
*~N~
BETA EXPERIMENT V1.2
500r,
400-
z
300
z
0
200
LQ
100
0
0
300
600
900
1200
1500
1800
2100
TIME [s]
Figuire 3.10
Comparison between the Predicted and Measured Downward Erosion Distances of BETA Test V1.2
Now&
BETA EXPERIMENT V1.3
500
r
z
400
300
0
CO)
0
0
200
100
Q
0
0
100
200
300
400
500
600
TIME [s]
Figure 3.11
Comparison between the Predicted and Measured Downward Erosion Distances of BETA Test V1.3
3.4.1.1 Molten Core Configuration in Concrete Cavity
The amount of heat that can be transferred downward during the MCCI
depends on the temperature and properties of the'melt directly in contact with
the horizontal concrete surface. The molten materials involved in the MCCI can be
classified into oxidic and metallic materials. The oxidic materials primarily consist
of uranium dioxide and zirconium dioxide. The constituents of the metallic phase
are structural materials, such as steel and zirconium. These oxidic and metallic
materials are immiscible, and the differences in the physical properties can be
large.
In the CORCON code, these two immiscible melts are assumed to be stratified
into layers in the reactor cavity. Physical orientation of the core debris conceived
in CORCON depends on the layers' densities. At the start of core debris attack on
concrete, the oxidic phase is heavier and is assumed to form a bottom layer of the
molten pool. The less dense metallic material forms a layer above this dense oxide.
As concrete ablation progresses, the decomposed concrete oxide, miscible with the
molten core oxide, will be incorporated into the oxidic layer thus reducing the
bulk density of this layer. On the other hand, the density of the metallic phase
may increase due to the inclusion of melted reinforcing steel and exclusion of
zirconium and chromium which are oxidized by the evolved gases. At some point
in time, the upper and lower pools of this debris configuration may flip due to
the density changes so that the oxidic layer will float atop the metallic phase. A
second hypothesis states that uranium metal may be present in the metallic phase
due to the reducing reaction of uranium dioxide with zirconium metal before core
meltdown. In this case, the density of the metallic phase may become heavier than
that of the oxidic phase at the start of the MCCI. This gives a debris configuration,
as depicted in the VANESA model, with the metallic layer at the bottom all the
time. However, the stratified layers configuration may in fact be destroyed by the
evolved gas if the superficial gas velocity is high enough. Gases sparging through
the melt will entrain and mix the oxidic and metallic phases into an approximately
l04
homogeneous mixture. This phenomenon could happen during the early stage of
the MCCI when the melt temperature is high.
Actual configuration of the core debris in the reactor cavity is not certain.
These possible assumptions could affect the prediction of the ex-vessel aerosols
release in two ways.
One is the downward heat transfer rate, another is the
vaporization potential of the melt. In this study, the downward heat fluxes from
different melts contacting the concrete will be analyzed.
3.4.1.2 Concrete Type of Reactor Cavity
All concretes used in construction have the same types of cement. What is
variable in the concrete is its aggregate composition and its free-water content.
When concrete is specified in the Final Safety Analysis Report for a plant, there
is no such data as the decomposition temperature, decomposition enthalpy, the
amount of water and the amount of carbon dioxide evolved upon heating the
concrete. However, these characteristics of the concrete are of crucial importance
to the analysis of MCCI.
There are three concretes, namely, Limestone/Common Sand, Limestone and
Basaltic, which have been widely adopted by severe accident analysts as representative of reactor concretes.
These concrete compositions are available as
user-selected default compositions in the CORCON code. In addition to these
concretes, a German silicate concrete, so called KfK concrete, was analyzed in
this study.
3.4.1.3 Solidus Temperature of Molten Core
In order to determine whether the concrete attack is by a molten or solid
material, well-known solidus temperatures of molten materials are needed. The
solidus temperature of any melt depends on its composition.
Available phase
diagrams are quite restricted to simple systems. However, the melt composition
105
that may develop in the MCCI is extremely complicated, therefore, precise prediction of the solidus temperature can be difficult or impossible.
In the CORCON/MOD2, simple procedures for estimating the liquidus and
solidus temperatures of various melts have been developed.
A sensitivity study
with various melt compositions will be analyzed to recognize the possible effect on
the downward heat transfer estimation.
3.4.2 Results and Discussions
The downward heat fluxes resulting from different conditions of the melt in the
melt/concrete interactions were calculated by using CORCON/MIT. The melts
chosen for this study are listed in Table 3.4. Melt properties shown in this table
are those calculated by the CORCON code.
Since the melt compositions will
change after the start of interactions, one-step calculations with different initial
melt temperatures were followed. Various concretes used as counterpart of the
interactions are listed in Table 3.5. The selection of melts and concretes has been
made to cover a range of materials that may result from various accident scenarios.
Physical properties, such as thermal conductivities of the melts and gas content
of the concretes, vary significantly among these selections.
3.4.2.1 Downward Heat Transfer Model
The calculated downward heat fluxes, based on the Gas Film (GF) model, the
Periodic Contact (PC) model and the Revised Periodic Contact (RPC) model, are
presented as follows.
In Fig. 3.12, the downward heat flux versus A T, (the difference between pool
bulk temperature and concrete decomposed temperature) are plotted for the case
of a core oxide overlying Limestone/Common Sand concrete. The solidus point of
the core oxide is shown in this figure. It is seen that there is about an order-ofmagnitude difference between the PC model and GF model predictions when the
nielt temperature is above its solidus point. The RPC model gives about 50%
106
Table 3.4
Compositions and Physical Properties of Various Melts
Steel
Steel+Zr
Core
Core+Concrete
Metal
Metal
Oxide
Oxide
Fe
76.0
55.5
-
Cr
18.0
13.5
-
Ni
8.0
6.0
-
25.0
-
Melt Composition(wt%):
Zr
-
-
80.0
76.6
Zr0 2
-
-
11.0
10.5
FeO
-
-
9.0
8.6
CaO
-
-
-
2.0
SiO 2
-
-
-
2.0
-
-
-
0.2
U0
2
A1 2 0
3
Melt Properties:
6767
6428
8206
7371
C, (J/kg K)
798
694
567
598
K (W/m K)
49.8
45.6
2.58
2.91
16400
14260
3465
3580
y (pPa- s)
3300
3000
6600
5200
a (N/m)
1.78
1.73
0.50
0.50
T901 (K)
1748
1748
2118
1941
Tiz, (K)
1758
1758
2744
2665
3
p (kg/m
j3 (J/m
2
)
so 'K)
107
Table 3.5
Compositions and Physical Properties of Concretes
Limestone/C.S.
Limestone
Basaltic
KfK
35.8
0.48
31.3
0.08
1.22
1.44
3.60
21.2
2.70
2.00
3.60
5.67
45.4
0.08
0.68
1.20
1.60
35.7
3.94
2.00
54.8
6.16
8.82
1.80
5.39
6.26
8.32
1.50
3.86
2.00
76.6
p (kg/m 3 )
C, (J/kg K)
K (W/m K)
2340
903
1.17
2340
979
1.17
0 sK)
0 (J/m 2so
HDecomp (MJ/kg)
TDecomP (K)
TSo
0 id. (K)
1572
2.500
1500
1420
1637
3.476
1750
1690
2340
913
1.59
1843
1.745
1450
1350
2300
910
1.59
1824
2.000
1573
1350
(K)
1670
'1875
1650
1650
p (kg/m 3 )
C, (J/kg K)
K (W/m K)
2666
1132
1.3
3245
1087
1.3
2456
1191
1.3
2346
1225
1.3
13.1
1596
0.203
65.1
7.08
1854
0.200
51.2
38.0
1319
0.214
103.6
32.5
1373
0.212
97.3
Compositions(wt%):
SiO2
MgO
CaO
Na 2 0
K20
Fe 2 0
A1 2 0
CO
3
3
2
H20 Evap
H20 Chem
-
9.22
-
5.32
2.92
4.22
1.77
Concrete:
TLiquidus
Slag:
Rising Fluid:
p (kg/m 3 )
C, (J/kg K)
K (W/m K)
0 (J/m 2 s 0 -'K)
108
Sand:
Limestone/Common
Melt Material: Core Oxide
TD= 1500 K
108
r_
0
400
800
-
AT,
Figure 3.12
1200
1600
TD [K]
Downward Heat Fluxes of the Various Models for Core Oxide Interacting with Limestone/Common Sand Concrete
109
higher downward heat flux than the PC model in this case. The differences among
various downward heat transfer models with different types of concretes can be
seen in Figs. 3.13, 3.14 and 3.15. It is noteworthy that the RPC model results in
three times higher downward heat flux than the PC model in the case of Limestone concrete. For silicate concrete (Basaltic and KfK), the downward heat flux
predicted by the RPC model is lower than the PC model by a very small value.
As predicted by all models, the downward heat fluxes drop dramatically near
the solidus temperature. The downward heat transfer is governed by conduction
through the crust, rather than a more effective convective controlled process. Additional thermal resistance across the solidified melt reduces the amount of heat that
can be transferred downward. Therefore, the calculated downward heat fluxes for
the post-freezing stage are rarely affected by different heat transfer models which
are used to predict the relatively small themal resistance at the pool/concrete interface. As shown in these figures, the downward heat fluxes calculated by various
models are almost equal at low debris temperatures.
3.4.2.2 Molten Material
The downward heat fluxes calculated for various melts interacting with Limestone/Common Sand concrete by the GF, PC and RPC models are shown in
Figs. 3.16, 3.17 and 3.18, respectively. At high melt temperature conditions, it is
interesting to note that the differences in the downward heat fluxes among these
melts are small even though the physical properties of the melts are widely different. Above the solidus points, the PC model predicts that the downward heat
flux of the metallic material is about 20% higher than the oxidic material. On
the contrary, the RPC model results in 30% higher downward heat flux with the
oxidic material. The differences predicted by the GF model are negligible. There
are reasonable explanations of these small differences for all models. In the GF
model, the downward heat transfer is predominately controlled by the gas film.
110
Lirnestone: TD= 1750 K
Melt Material: Core Oxide
108
E-1
106
LJ
0
400
AT
Figure 3.13
800
=Tp -
1200
1600
TD[K]
Downward Heat Fluxes of the Various Models for Core Oxide Interacting with Limestone Concrete
111
10
9
Basaltic: TD=1 4 5 0 K
Melt Material: Core Oxide
1-
1-
1
1
108
Lj
X
"U>
106
0
400
800
1200
1600
= Tp - TD [K]
Figure 3.14
Downward Heat Fluxes of the Various Models for Core Oxide Interacting with Basaltic Concrete
112
Kf K Concrete:
T= 1 5 7 0 K
Core Oxide
9 Melt Material:
r-n
C\1
106
Ae.
5
0
Q
104
0
400
800
-
Figure 3.15
1200
1600
TD [K]
Downward Heat Fluxes of the Various Models for Core Oxide Interacting with KfK Concrete
113
GAS FILM MODEL
Limestone/Common
Sand; TD=1
50 0
K
108
E-- 106
LJ
0
400
800
1600
- TD [K)
ATp
Figure 3.16
1200
Downward Heat Fluxes of the Gas Film Model for Various Melts
Interacting with Limestone/Common Sand Concrete
114
PERIODIC CONTACT MODEL
Limestone/ Common Sand: TD=1
10 9
E-
500
K
106
10 5
0
400
800
-
Figure 3.17
1200
1600
TD [K]
Downward Heat Fluxes of the Periodic Contact Model for Various
Melts Interacting with Limestone/Common Sand Concrete
115
REVISED PERIODIC CONTACT MODEL
Limestone/Common
Sand:
TD= 1500 K
108
E-L 106
105
0
400
800
-
ATp
Figure 3.18
1200
1600
TD [K]
Downward Heat Fluxes of the Revised Periodic Contact Model for
Various Melts Interacting with Limestone/CS Concrete
116
Physical properties of the melt play only a minor role. The downward heat flux
obtained from the PC model increases with the thermal conductivity of the melt
as well as the periodic contact frequency. The periodic contact frequency is directly reverse proportional to the Laplace constant A. Typical values of several
parameters used in the periodic contact model associated with the thermal properties of molten pool materials are shown in Table 3.6. Melt consisting of metallic
material has higher themal conductivity but lower frequency of periodic contact
compared with the oxidic melt. In the PC model, the dependence of the downward heat transfer coefficient on the pool properties has been weakened by the
approximate estimation of the interface temperature (notice the small variation
of the first column in Table 3.6), and then it has been strengthened by including
the thermal conductivity ratio. Combination of these effects results in a higher
downward heat flux of the metallic pool with the PC model. For the RPC model,
with the combination of those parameters in the first, second and fourth columns
of Table 3.6 (see equations (3.58) and (3.59)), the calculated downward heat flux
of an oxidic pool is higher than that of a metallic one.
The thermal conductivity of the melt can be, however, an important factor
in determining the downward heat flux at low melt temperature condition. Below
the solidus temperatures of the melts, the calculated downward heat fluxes of the
metallic material, shown in Figs. 3.16 through 3.18, are about 6 to 10 times higher
than the oxidic material.
3.4.2.3 Concrete Type
At a melt temperature above the solidus point the downward heat flux will be
determined by the temperature of the heat sink, i.e. the temperature of the decomposed concrete surface. The higher the decomposition temperature the lower is the
heat flux that can be transferred downward. With the lowest decomposition temperature of the Basaltic concrete, as shown in Table 3.5, the calculated downward
heat flux is the highest among the various concretes. Limestone concrete
117
Table 3.6
Typical Values of the Parameters Used
in the Periodic Contact Model
1/(1+ k)
()/(
+
)
k,/k.
A (mm)
Oxidic Pool:
Limestone/Common Sand
0.982
16.6
2.21
2.50
Limestone
0.985
21.7
2.21
2.49
Basaltic
0.971
11.6
1.62
2.50
KfK
0.973
12.3
1.62
2.50
Limestone/Common Sand
0.996
22.0
42.6
5.19
Limestone
0.997
29.1
42.6
5.18
Basaltic
0.994
16.0
31.3
5.20
KfK
0.994
18.6
31.3
5.19
Metallic Pool:
118
with the highest decomposition temperature gives the lowest heat flux.
Both
the GF and PC models predicted the same trend for these concretes, but the
differences among these concretes calculated by the PC model are larger than
the GF model, see Figs. 3.19 and 3.20. However, in the RPC model, the effect
of ATp is no longer a dominant factor when the melt temperature is relatively
high. The dependence of -yo shown in equation (3.59) on the thermal properties
of the concrete and its decomposed materials (rising fluid) can be seen in Table
3.6. Among various concretes, the Limestone concrete has the largest -yo value,
therefore, the RPC model gives Limestone concrete the highest heat flux at high
melt temperature condition, see Fig. 3.21.
3.4.2.4 Solidus Temperature of Molten Material
In figures 3.16 to 3.21, it can also be seen that the solidus temperature is an
important boundary in the prediction of the downward heat fluxes. The solidification model in the CORCON/MOD2 assumes that a crust forms on any surface
whose temperature falls below the solidus temperature of the molten material.
In CORCON/MOD2, the simple procedures used for estimating the liquidus and
solidus temperatures of various melts give two questionable results.
First, the
solidus temperature of a metallic material is determined by a simple fit to the
Fe - Cr - Ni ternary phase diagram without considering the effect of the presence of other metals, such as zirconium. This leads to an equal solidus temperature
for the steel and the steel+Zr melts in this study. The second interrogative result
can be seen in Fig. 3.22. The solidus temperature of the core oxide, specified in the
Table 3.4, is affected by the concrete assumed in the interactions. The differences
of the calculated solidus temperature can be as large as 200 K between the cases
of Limestone/Common Sand and Basaltic concretes. Intuitively, there is no reason
for the solidus temperature of the melt to depend on the concrete. This curious
result is caused by the iron oxide in the oxidic material since CORCON treats
all oxides other than fuel-oxides as decomposed concrete oxides. The solidus and
liquidus temperatures of the oxidic melt are then determined by those of fuel119
GAS FILM MODEL
1
9
Melt Material:
LI0
Steel Metal
I
II
I-
TONE /COMMON
LIMES TONE
-. -BASAL TIC
108
SAND
KfK
gji
04
i6
106
r
-7
-
-
-
105
I
1500
I
1800
I
I
II
I
I
2100
2400
I
I
II
2700
3000
TPool [K]
Figure 3.19
Downward Heat Fluxes of the Gas Film Model for Steel Melt Interacting with Various Concretes
120
PERIODIC CONTACT MODEL
10
9
Melt Material;
I I
Steel Metal
I
I
I
I
I
I
I
108
E-n
106
X0
1030
1500
1800
2100
2400
2700
3000
TPool [K]
Figure 3.20
Downward Heat Fluxes of the Periodic Contact Model for Steel
Melt Interacting with Various Concretes
121
REVISED PERIODIC CONTACT MODEL
10
9
Melt Material:
LI
I
Steel Metal
I
I
I
I
2100
2400
I
I
I
Z
108
Emi
106
0--
103L_
1500
1800
2700
3000
TPool [K]
Figure 3.21
Downward Heat Fluxes of the Revised Periodic Contact Model for
Steel Melt Interacting with Various Concretes
122
REVISED PERIODIC CONTACT MODEL
10 9
Melt Material:
Core Oxide
10 8
E--1
10 6
1500
1800
2100
2400
2700
3000
TPool [K]
Figure 3.22
1
Downward Heat Fluxes of the Revised Periodic Contact Model for
Core Oxide Interacting with Various Concretes
123
oxides and those of concrete weighted by the molar fractions of the non-fuel and
fuel-oxides in the melt.
However, the iron oxide, an element that very likely
participates in the oxidic melt before the start of MCCI, has nothing to do with the
concrete properties. The way that CORCON calculates the solidus temperature
may lead to a fallacious prediction of formation of a bottom crust at the initiation
of the MCCI. This could have a significant impact on the prediction of ex-vessel
aerosols behavior if the initial melt temperature is close to the calculated solidus
point.
3.5 Summary
1. A revised periodic contact model has been developed based on more complete
theoretical consideration than earlier derivations. Under certain conditions,
there are considerable differences between the downward heat fluxes predicted
by the revised and original periodic contact models.
2. When the debris temperature is above its solidus point, the downward heat
flux predicted by the PC model is about an order-of-magnitude higher than
the GF model.
3. With a bottom crust or an entirely solid layer, the downward heat fluxes
predicted by all models are about equal.
4. There is an order-of-magnitude reduction in the downward heat flux when the
melt temperature drops across its solidus point. A more precise estimation of
the solidus temperature of melt is warranted.
5. At high debris temperature, the downward heat flux is not affected significantly by the composition of the melt which may contact the horizontal
concrete surface. When the debris has a low temperature, allowing formation
of a bottom crust, a metallic pool with higher thermal conductivity has ten
times higher downward heat transfer than an oxidic one.
6. For a given condition of the molten pool materials, the differences in the
calculated downward heat fluxes for different concretes can be large.
124
CHAPTER 4
VALIDATION OF HEAT TRANSFER MODEL BY
COMPARISONS TO INTEGRAL EXPERIMENTS
4.1 Introduction
To describe the heat transfer processes involved in the melt/concrete interaction, computer codes (such as CORCON and WECHSL) have been developed
during recent years, based mainly on small scale separate effect simulant material
experiments or on transient tests with prototypic materials. To gain a wider data
base for validation of the existing models, integral experiments considered to be
reliable extrapolation of real reactor situations were later undertaken.
The BETA program at Kernforschungszentrum Karlsruhe (KfK), West Germany is a key experimental program for core melt/concrete interactions.
This
program carried out successful experiments between early 1984 and early 1986.
Data with regard to the concrete ablation, gas generation and aerosol releases
were obtained.
Phenomena important to the modeling of MCCI were observed
as well. At Sandia National Laboratory, an experimental program to investigate
melt/concrete interaction was initiated in July 1975. Various experiments using
real materials have been performed to identify characteristics of the melt/concrete
interaction.
Some conclusive statements have been reached based on the obser-
vations of the experiments, however, much of the data which can be used to
corroborate theoretical work remains unpublished.
These real material experiments will be reviewed in the following sections, and
the measured downward concrete ablation distances will be used to validate various
downward heat transfer models developed for the melt/concrete interaction.
125
4.2 Review of Real Material Experiments
4.2.1 BETA Experiments
4.2.1.1 Descriptions of BETA Facility
The BETA facility is a large scale high power inductive heating experiment
to allow sustained heating of a simulated core melt in a concrete crucible.
A
schematic diagram of the BETA facility is shown in Fig. 4.1.
To fulfill the requirements of code verification with various needs of model development, different interaction regimes were studied by a sequence of experiments
at various temperatures and power levels. The controlled quasi-steady conditions
of the power level were selected to result. in a heat flux to the concrete of the same
magnitude expected in reactor accidents.
The inner diameter of the cylindrical concrete crucible which contains hot
melt is 380 mm.
This dimension was selected to allow mutually undisturbed
downward and sideward melt propagations. This diameter also guarantees that
the gas release at the bottom and the stabilization of a top crust are not affected by
the vertical wall. The effects of the crucible diameter were examined in the V4 test
with a 600 mm crucible diameter. The detailed concrete crucible diagram is shown
in Fig. 4.2. The dots in the figure are the locations of embedded thermocouples.
The propagation of the melt front was inferred from the time of failure of the
thermocouple in the concrete.
The experiment concrete in the BETA tests was a silicate concrete of the type
commonly used in German reactors. This concrete is called here the KfK concrete.
Additionally, in the V3 test series, three crucibles of high carbonate content have
been used, farbricated from limestone material imported from USA. The materials
were Limestone concrete for V3.2 test, and Limestone/Common Sand concrete for
V3.1 and V3.3 tests. The compositions and the physical properties of the various
concretes can be found in Table 3.4.
126
MOLTEN POOL
PERIPHERY SYSTEMS
fThM~cUPIAs
lpflssm
m
1
2
3
CONCRETE CRUCIBLE
INDUCTION COIL
OFFGAS SYSTEM
Figure 4.1
4
5
THERMITE REACTION TANK
CONTAINER FOR MEASUREMENT
PROBES
Schematic Diagram of BETA Experiment (Ref.[A5])
127
IwoI
-
om
-O
00
ILII
Figure 4.2
Dimensions of Concrete Cavity of BETA Experiment (Ref.[A5])
128
High power inductive heating of the melt -
one of the main characteristics of
the BETA facility with up to 1900 kW net input power -
can be induced in the
metallic layer. This high power capability allows observation of various physical
and chemical processes during an extended period of time under quasi-steady
conditions.
Up to 350 kg metallic and 150 kg oxidic melt could be generated outside the
crucible by a thermite reaction and then poured into the concrete crucible. The
initial melt temperature was close to 2000
0C
or higher, as determined by the
composition of thermite without having direct measurement. The simulated melt
was initially composed of a steel melt with Fe, Cr, and Ni, and in most of the
tests, oxidic melt of Al 2 03 and Si0 2 . CaO was added to some of the oxidic melts
to lower the viscosity and solidification temperature of the oxide. The melt did
not contain simulants for the radioactive fission products.
Dip-in material sampling and temperature measurement instruments for melt
analysis were available above the crucible. The temperature of the melt was measured at predefined time intervals by W - Re thermocouples which were dipped
simultaneously into the metallic and oxidic phases. Also, the lances of dip-in temperature of the melt were measured continuously by a ratio pyrometer installed
on top of the crucible.
The hood and offgas pipe collected all gaseous products generated during
melt/concrete interaction for physical and chemical analysis. Two methods were
used independently for determination of the composition of the gas phase: On-line
measurements were realized by means of a quadrupole mass spectrometer (MS).
Off-line investigations of the gas phase by application of grab samples, filled during
the experiment and analyzed by gas chromatograph (GC) later on.
Aerosol concentration in the offgas line was analyzed by an on-line scattering
device (laser photometer) and by probe analysis.
129
Throughout the experiments, the behavior of the melt surface was observed
by a TV camera which provided valuable information on gas flow through the
melt, surface crust formation and aerosol production. After each experiment, the
crucible was sectioned and the post-test cavity shape observed.
4.2.1.2 Experimental Results and Discussions
A listing of BETA experiments is shown in Table 4.1. There were 19 tests
divided into the following categories:
" VO-series : Test of facility
" Vi-series : High power tests
" V2-series : Low power tests
" V3-series : U.S. concretes tests
e V4-series : Large cavity test
The results of these tests will be reviewed and discussed in what follows.
(i)
Concrete Erosion
For melts with high power input and small influence of crusts at the melt/
concrete interface, propagation of the melt was predominantly downward. Typical
downward and sideward melt front propagations are given in Fig. 4.3. It shows
a nearly constant and very high downward erosion rate of 1.0 mm/s in V1.8
test. Sideward erosion was much slower and ended after some 100 seconds. The
dominant downward erosion indicated a very effective heat transfer mechanism at
the bottom of the crucible, which was different from the sideward heat transfer
mechanism.
Experiments with low power input showed the role of the solidification process.
Sideward and downward propagations were found to be more balanced due to the
heat resistance of the metal crust at the bottom of the crucible. Fig. 4.4 shows
the erosion rates for V2.3 test, with the downward erosion rate of some 0.1 mm/s
being a factor of 10 less than that of V1.8 test.
130
Table 4.1
Test Matrix of the BETA Experiments
Test
Melt
Planned Power (kW)
VO.1
Iron
0
VO.2
Iron
400
VO.3
Iron+Oxide
1700
V1.1
Iron
Pulsed
Pour Failed
V1.2
Iron+Oxide
Pulsed
Lorentz-Forces
V1.3
Steel+Oxide
1000
V1.4
Steel
0
V1.5
Steel
450
V1.6
Steel+Oxide
1000
V1.7
Steel+Oxide
1700
V1.8
Steel+Oxide
1900
V1.9
Steel+Oxide
200-400
V2.1
Steel+Oxide
120-150
V2.2
Steel+Oxide
50-90
CaO Added
V2.3
Steel+Oxide
240
CaO Added
V3.1
Steel+Oxide
1700-2500
Remarks
Transient
II
No Dispersion (CaO)
CaO Added
Limestone/Common Sand
Heating from 0-66 seconds
V3.2
Steel+Oxide
40- > 1000
V3.3
Steel+Oxide
400-600
Limestone/Common Sand
V4.1
Steel+Oxide
300-1000
600 mm Diameter
Limestone
300 kg Oxide with Cao
550 kg Steel
131
BETA EXPERIMENT V1.8
500
I
II
IIIII
* AXIAL
o RADIAL
400
A
-
z
300
7,-
z
0
OWN
200
100
0
100
0
200
300
400
TIME [s]
Figure 4.3
I ON
Measured Concrete Erosion Distances of BETA Test V1.8
500
BETA EXPERIMENT V2.3
500 * AXI AL
LJ
40
RADIAL
L-j400 --
z
3000
8200
100
-
0
0
1000
2000
3000
4000
TIME [s]
Figure 4.4 Measured Concrete Erosion Distances of BETA Test V2.3
5000
In the test with Limestone/Common Sand concrete, the dominating downward erosion was also observed. However, in comparison with KfK concrete, the
sideward penetration was more pronounced.
(ii)
Melt Temperature
Fast initial cooling of the melt was always observed in the BETA tests. With-
out considering the possible uncertainty of the high temperature measurement,
the melt cooled down from its high initial temperature (as high as 2573 K) to its
solidus point (around 1870 K) typically in less than few hundred seconds, even
with the power input per unit volume to the melt being an order of magnitude
higher than decay power. This fast cooling is believed to be caused by the considerable downward heat transfer.
The long-term temperatures of the melts were close to the freezing temperature of the metal.
(iii)
Release Gas Composition
Gas analysis showed the major releases from melt/concrete interaction were
H2 , H 2 0, CO and CO2 with a small amount of CH 4 . With silicate concrete, H2
and H 2 0 were the dominating gases, while in the limestone concrete test, CO and
CO 2 were the main components.
Thermodynamic calculations based on the minimization of Gibbs free energy
were performed. It was found that at the beginning of the experiments, released
gases were in equilibrium (correspondence temperature 1200 K).
for both kinds of concrete.
This is valid
After the initial phase of the experiments the gas
composition deviated from equilibrium.
(iv)
Aerosol Release
The aerosol release in the BETA experiments with KfK concrete was charac-
terized by a short dense aerosol peak during, and some seconds after, pouring of
the melt, with the main components being iron and small amounts of chromium,
134
silicate, and other species from the decomposed concrete. Due to the small size
of the aerosols, the aerosol releases were judged to be formed by evaporation and
condensation of volatile species, while sparging was a negligible aerosol generation
process. A typical rate of 0.1 g aerosol per mole gas during the heating period, with
a reduction of an order-of-magnitude after shutting the power off, was measured in
the high power tests. In V2.3 test with low power input, the aerosol concentration
reduced from a short initial peak to 0.01 g per mole gas throughout the experiment.
For V3.2, the Limestone concrete test, the release of aerosols was considerably
more intense than in all other experiments. The aerosols were mainly CaO crystals
with a typical 1 micron size.
During the heating period, however, the aerosol
photometer gave no data because of the very high aerosol concentration.
The
minimum concentration in that period was estimated to be 1.2 g per mole gas. The
Limestone/Common Sand concrete showed considerably less aerosol generation
rates but of the same constitution as Limestone concrete. The aerosol generation
rate of Limeston/Common Sand concrete was comparable to that of KfK concrete
experiments, even for the high power heating phase.
The aerosols production in the limestone concrete test was attributed to the
process of lime burning. The calcium oxide forming at the decomposing concrete
surface can be easily powdered and transferred as solid particles into the gas
stream. For Limestone/Common Sand concrete, the presence of a relatively large
amount of SiO2 reduced this process by the formation of low melting SiO2 CaO mixtures. Therefore, the aerosol generation of the Limestone/Common Sand
concrete was not as intense as the Limestone concrete.
(v)
Boundary Crust Formation
The video observation of the melt surface during test V2.3 showed that, some
minutes after pouring, a thin unstable surface crust formed at the top of the oxidic
melt.
This crust, however, did not influence the gas release.
With additional
melting and admixture of silica, the crusts disappeared and the melt became less
135
viscous. Throughout the experiment, the heat transfer in the oxide was controlled
by the percolating gas. The steel melt, underlying the oxide, was at solidification
temperature and had probably formed a crust at. the concrete interface.
The
bottom crust gave a reasonable explanation of the relativly low downward erosion
rate in test V2.3.
(vi) Layer Mixing
Some of the experiments of the high power test series showed the occurence of
dispersion of the metal into the oxidic melt. The process of dispersion was driven
by the high gas flux evolving from the concrete, and resulted in fine metal droplets
distributed all over the oxidic melt. In BETA, dispersion was detected by the loss
of heating efficiency as the dispersed metal did not couple to the induction heater.
The low density difference between the metallic and oxidic species and the high
viscosity of the oxide facilitated dispersion.
In test V1.8, the addition of CaO to the oxidic melt avoided the occurence
of dispersion even for very high gas release by a reduction of the oxidic viscosity.
Under similar experimental conditions but with higher viscosity, test V1.7 ended
with the metallic melt completely dispersed. In all low power experiments, the
metallic and oxidic melts remained segregated, as the gas flow was low.
(vii) Melt Splashing
In the V3.1 test, due to the extremely vigorous gas release and agitation of
the melt, a considerable amount of the melt was splashed to the upper crucible
wall and formed a metallic layer. This metal layer cylinder increased the coupling
efficiency of the induction coil. The net power in the melt increased to 2500 kW,
the highest rate ever obtained in BETA, and caused failure of the power control
system by running out of range at 66 seconds.
Various attempts to reactivate
the induction system were not successful and freezing of the melt occurred very
rapidly.
136
(viii) Spallation
As limestone concrete interacts with a high temperature melt, the aggregates
are reduced by the lime burning process to CaO which has virtually no mechanical
strength. The concrete with a certain amount of CaO may break by the pressure
generated inside the concrete. In V3.2 test, some thermocouples failure by mechanical breakup were detected. In V3.3 test, spalled material from the upper
crucible was observed to drop into the melt.
The removal of the CaO by spallation may reduce the effective decomposition
temperature of the concrete, and result in a higher concrete erosion rate.
4.2.2 Sandia Experiments
The most intensive experimental study in the USA on the melt/concrete interaction was initiated about a decade ago at Sandia National Laboratory. Various
experiments [C1,M2,P1,P5,P6,P12,S3,S4] have been performed by using real materials or prototypic materials at different scales to investigate the phenomena of
the melt/concrete interaction. Unfortunately, the details and the results of most of
these experiments have not been systematically published. Only few quantitative
results are available at this time. Recently, results of two of the programs, SWISS
and TURC, were brought into attention since the results of these experiments were
adopted for validation of the models developed at Sandia. Since then, more information about these experiments has been released [B2,G4,G5,G12]. These two
experimental programs will be discussed and analyzed in the following sections.
4.2.2.1 Descriptions of Sandia Experiments
(i)
Sustained Melt-Concrete Interaction with Overlying Water (SWISS)
The objective of the SWISS program was to study the interaction between
molten debris, concrete and an overlying water pool. Principal observations included concrete erosion, crust formation, heat transfer, gas generation, aerosol
transport and overlying water effect.
37
A schematic view of the experimental apparatus is shown in Fig. 4.5.
A
crucible designed with magnesia oxide (MgO) sidewall and Limestone/Common
Sand concrete bottom was utilized in order to limit .melt/concrete heat transfer to
the axial direction only. Radial heat losses from the melt were determined from
the response of thermocouples embedded in the MgO walls. Water coolant was
injected on top of melt at planned times during the experiments.
About 45 kg
of molten stainless steel generated by induction heating in a melt generator was
poured into the crucible. Power input to the melt by an induction coil surrounding
outside the MgO wall was sustained at approximately 100 kW level. Release gas
was guided into an instrumented flow tube on top of the crucible to measure the
flow rate and aerosol mass generation rate.
In the SWISS program, a series of experiments was planned (see Table 4.2),
however, only the first two experiments (SWISS-1 and 2) were documented. The
major difference of these two experiments was the water quench time. The water
flow was activated late (~35 minutes after melt pouring) in the SWISS-1 test.
While in SWISS-2 test, the water was introduced in less than 2 minutes. Both
experiments were extended for a period of 40 minutes. These two experiments
provided valuable information about the effects of overlying water.
(ii)
Transient Urania Concrete Experiments (TURC)
The TURC facility was designed to identify melt/concrete interaction and
quantify the physical source term. This experiment could be of great interest to
reactor accident analysis since it contains several unique features:
" Well characterized experimental measurement of the released aerosols.
" Fuel-oxide (U0
2 /Zr02)
melt/concrete interaction.
" The melt was doped with fission product mockup (such as Te, La, Ce and
Ba).
A schematic diagram of the TURC facility is shown in Fig. 4.6. The crucible
was once again configured with MgO sidewall to ensure the melt attack on the
138
SWISS EXPERIMENTAL
APPARATUS
TOP HAT
INSTRUMENTED FLOW TUBE
AERSOLS
GAS SAMPLING
FLOW RATE
CHAMBER
MELT PENETRATOR
CYLINDER
MELT GENERATOR (MgO)
ZIRCONIA INSULATION
25.4
INDUCTION
COIL
STAINLESS STEEL 46 kg WATER INLET
CRUCIBLE (MgO)
- WATER EXIT
WATER POOL
Figure 4.5
SWISS Experimental Apparatus (Ref.[G 12])
139
r
.,:2
Table 4.2
Test Matrix of the SWISS Experiments
Test
Crucible
SWISS-1
LCSt
Purpose
Melt
Water
Quench Temperature
Late
-
2000 K
Base Line Data
Thick Crust Quench
SWISS-2
LCS
Early
-
2000 K
Melt/Water Interaction
Thin Crust Quench
SWISS-3
MgO
Early
-
Melt/Water Interaction
For Small jg:
Heat Transfer
Fission Product Transport
t
Limestone/Common Sand concrete
140
TURC ISS EXPERIMENT FACILITY
I
Figure 4.6 TURC-1SS Experiment Facility (Ref.[G4])
141
concrete was one-dimensional.
The radius of the MgO cylinder was 0.208 m
enclosing a Limestone/Common Sand concrete block at bottom.
Several tests
involved transient interaction of about 150 kg metallic or oxidic melt with the
concrete. see Table 4.3.
In TURC, a great deal of work has concentrated on the aerosol/fission product source term measurement by using filter samplers, impactors, cyclone and
photometer.
The mass source rate, distribution and the composition of the re-
lease aerosol quantified by the analysis of aerosol sampling of TURC test can be
a valuable data base and well suited for VANESA model validation.
4.2.2.2 Results and Conclusions
Some of the more significant reported observations in the Sandia experiments
are summarized below:
(i) SWISS Tests
" Concrete erosion rate:
-
0.08 mm/s.
" The effect of overlying water on concrete erosion rate was negligible.
" Significant, reduction by a factor of 10 to 20 of the aerosol mass generation
rate after water was added.
* Neither steam explosion nor apparant fragmentation was observed.
" Reduction of aerosol release due to the formation of top crust was detected.
(ii) TURC Tests
" The concrete erosion rates of TURC-1SS and TURC-1T were 0.4 and 0.1
mm/s, respectively. While in the oxidic melt test, TURC-2, the formation
of stable crust reduced the concrete erosion rate to 0.008 mm/s.
" Combustible gases at thermal equilibrium were detected in both metallic and
oxidic melt tests.
" Aerosol generation rates measured in TURC-1SS test as well as the VANESA
predictions are shown in Fig. 4.7.
14-2
Table 4.3
Test Matrix of the TURC Experiments
Test
Melt Mass
(kg)
Melt Composition
Melt Temperature
(K)
TURC-1T
150
Fe - Al 2 03 (Thermite)
2700
TURC-1SS
147
S.S. 304
2350
TURC-2
144
LTO 2 , ZrO2
2820
TURC-3
200
U0
2
, ZrO2 , Zr
143
2670
10
0.1
-
TURC-188 MEASUREMENT$
0-01
PREDICTIONS BASED ON
OE-0' VANESA
EXPERIMENTAL TEMPERATURES.
GAS FLOWS. AND EROSION
2
1
3
TIME (Minutes)
Figure 4.7
Comparisons of the TURC-lSS Test Data and the Predictions of
the VANESA Code (Ref. [P13])
144
4.2.3 Summary
The experimental programs have succeeded in enhancing the understanding
of melt/concrete interaction and the resulting consequences, important in reactor
severe accident analysis. A qualitative understanding of the melt/concrete interaction has been obtained, and quantitative data have been obtained for many specific
phenomena. It is likely that the most significant phenomena that might be present
in a core meltdown accident have already been identified in these programs.
4.3 Experiment Analysis and Model Validation
As discussed in the previous chapter, the downward heat fluxes calculated by
the various heat transfer models are different by an order-of-magnitude. It is important that the accuracy of these analytical models be justified based upon the
real material experimental data. In the real material experiments, the physical
phenomena at the pool bottom interface cannot be directly observed. However,
the downward concrete erosion rate directly related to the magnitude of the downward heat flux can be measured to validate the precision of the analytical model.
Integral analyses based on CORCON/MIT using different downward heat transfer
models have been done to predict the downward erosion of the BETA and Sandia
experiments. The erosion distances measured at the end of BETA tests are used
to quantify the standard deviations of the various heat transfer models.
4.3.1 Experiments Analyzed and Input Parameters Used
Twelve BETA tests were analyzed with different downward heat transfer models. The experimental conditions and melt compositions of these tests used in the
CORCON/MIT calculation are listed in Tables 4.4 and 4.5, respectively.
The
experiment duration was determined by the maximum of the concrete erosion distance to be allowed. When the melt front reached a certain range, the power was
turned off and the experiment was terminated before meltthrough of the concrete
crucible. Typical power input histories are shown in Figs. 4.8 and 4.9.
145
Table 4.4
Test Conditions of the BETA Experiments
Melt
Mass
Initial Melt
Temperature
tExperiment
Duration
tAverage
Power
Total
Energy
(kg)
(K)
(s)
(kW)
(MJ)
VO.2
300
2473
2500
370
930
VO.3
450
2473
462
1180
540
V1.2
350
2473
2075
380
780
V1.3
450
2173
510
780
400
V1.5
300
2273
1300
330
430
V1.6
350
2273
742
710
530
V1.7
380
2373
375
1010
380
V1.8
490
2073
386
1600
620
V1.9
410
2173
2420
260
630
V2.1
450
2273
4620
140
640
V2.3
400
2173
4480
190
860
V3.3
350
2573
2930
350
1020
Test
t Determined
by the last data point of the axial erosion distance.
Average power input over the experiment duration.
146
Table 4.5
Melt Compositions of the BETA Experiments
A12 0 3
(kg)
-
-
150
-
-
200
-
-
150
-
-
V1.3
246
30
24
105
45
-
V1.5
246
30
24
-
-
-
V1.6
246
30
24
45
5
-
V1.7
246
30
24
72
8
-
V1.8
315
17.5
17.5
91
-
39
V1.9
315
17.5
17.5
42
-
18
V2.1
270
-
30
105
45
-
V2.3
270
15
15
70
10
20
V3.3
246
30
24
50
-
-
Fe
(kg)
VO.2
300
VO.3
300
V1.2
Cr
(kg)
147
SiO 2
(kg)
CaO
(kg)
Ni
(kg)
Test
BETA EXPERIMENT V1.3
1500
1000
z
oc
500
0 L
0
500
1000
1500
TIME [s]
Figure 4.8 Power Input History of BETA Test V1.3
I~m~
2000
BETA EXPERIMENT V2.3
800
700
600
500
z
400
300
200
100
0
0
1000
2000
3000
4000
TIME [s]
Figure 4.9 Power Input History of BETA Test V2.3
'I L f
5000
For Sandia experiments, the conditions of the analyzed tests are shown in
Table 4.6. The one-dimensional erosion mode was treated in the CORCON/MIT
by cutting off the sideward heat transfer. The sideward heat loss to the MgO wall
was compensated for by reduction of the actual power input to the code calculation.
The actual power input and the one for code calculation of the SWISS-1 test are
shown in Fig. 4.10.
To initiate the CORCON calculation, parameter values other than the specifications of the experiment are required to specify the boundary conditions and
emissivities of various components. The values commonly used for these parameters are listed in Table 4.7.
4.3.2 Results and Discussions
4.3.2.1 BETA Experiments
The experimental results and analytical predictions of the concrete erosion
of early BETA tests have been shown in the previous chapter, while those of the
other BETA tests are shown in Figs. 4.11 through 4.18. In general, the GF model
significantly underestimates the downward erosion, and the FC model overpredicts
some of the BETA tests, especially in the low power tests. As one can see the
RFC model gives the best agreement to the experimental data. The precisions of
these heat transfer models will be quantified in the following section.
The melt temperature responses of some of the BETA tests are shown in
Figs. 4.19 through 4.27. As indicated, most of the experiments showed a rapid
initial melt temperature drop for both metallic and oxidic layers. In the calculated
results, the FC model gives the highest rate of the initial melt temperature drop
among various models. However, the melt temperature is still greatly overpredicted by the FC model both at the initial transient period and at the steadystate. This may be caused by an error in the melt temperature measurement. As
indicated in these figures, most of the measured metallic temperatures drop below
the solidus temperature of the metallic layer, and some of them are even lower
150
Table 4.6
Test Conditions of the Sandia Experiments
SWISS-1 SWISS-2 TURC-1T TURC-ISS TURC-2
Melt (kg)
Fe
33.3
32.3
Ni
8.1
7.9
Ni
3.6
108.8
-
-
26.5
-
3.5
-
11.8
-
-
-
-
-
Zr0 2
-
-
-
-
A12 0
-
-
U0
2
120.0
100.8
43.2
Total
45.0
43.7
30.0
150.0
Initial Melt
Temperature (K)
2000
2000
2700
2350
2820
Concrete Crucible
Radius (m)
0.108
0.108
0.208
0.208
0.208
Water Quench
Late
Early
No
No
No
Power Input (kW)
-
0
0
0
Concrete
LCSt
LCS
LCS
LCS
LCS
Erosion Mode
1-D
1-D
1-D
1-D
1-D
40
40
30
30
15
2400
2400
735
120
600
3
Heat Loss to
MgO Wall (kW)
Duration (s)
t
90
100
-
Limestone/Common Sand concrete
151
-
-
147.0
144.0
SNADIA SWISS- I EXPERIMENT
150
I
-1--Epe
-
I
I
Dt
Data
-Experiment
......... Code Calculation
-/
100
/
- --- . J
I
Lz...
I.
50
I:
I:
Q1
I:
0
-50
I
I
0
I
I
20
10
I
I
30
TIME [min]
Figure 4.10
11"WWOMP
Power Input History of Sandia SWISS-1 Test
It A*t4
I
-
40
Table 4.7
Input Parameters Used in CORCON/MIT for Experiment Analysis
Initial Concrete Temperature
300 K
Initial Coolant Temperature
SWISS Tests
300 K
Surrounding Pressure
1.0 atm (Constant)
Surrounding Temperature
Beginning
300 K
End
500 ~ 800 K
Emissivity
Concrete
Melt
Surrounding
0.6
0.8
0.8
153
BETA EXPERIMENT V1.5
500
i
0
r,
400
-
I
I
I
I
EXPERIMENT AXIAL
EXPERIMENT RADIAL
FC MODEL AXIAL
FC MODEL RADIAL
RFC MODEL AXIAL
RFC MODEL RADIAL
GF MODEL AXIAL
GF MODEL RADIAL
300
S
200
--
100
---
O
0 &
0
---
I
-I
300
-_-
900
600
1200
1500
TIME [s]
Figure 4.11
Comparison between the Predicted and Measured Erosion Distances of BETA Test V1.5
ft
BETA EXPERIMENT V1.6
500
r
400
z
300
0
0
CI)
200
100
0
0
100
200
400
300
500
600
700
800
TIME [s]
Figure 4.12
Comparison between the Predicted and Measured Downward Erosion Distances of BETA Test V1.6
14 At
BETA EXPERIMENT V1.7
500
400
z
E--
300
Do
z
0
p-
z
200
100
0
0
100
200
300
400
TIME [s]
Figure 4.13
Comparison between the Predicted and Measured Downward Erosion Distances of BETA Test V1.7
Rp
MONlftItO*h#
BETA EXPERIMENT V1.8
500
r
-
S--
EXPERIMENT
FC MODEL
RFC MODEL
400
.......
GF MODEL
z
300-
-
z
o200
100
0
0
100
200
300
400
TIME [s]
Figure 4.14 Comparison between the Predicted and Measured Downward Erosion Distances of BETA Test V1.8
BETA EXPERIMENT V1.9
500
r
z
400
300
0
200
00
z
100
0
0
500
1000
1500
2000
2500
TIME [s]
Figure 4.15
Comparison between the Predicted and Measured Downward Erosion Distances of BETA Test V1.9
ROM
-vow
-PWOW -
BETA EXPERIMENT V2.1
500
400
z
U)1
300
z
0
Now
200
100
OL
0
1000
2000
3000
4000
5000
TIME [s]
Figure 4.16
Comparison between the Predicted and Measured Downward Erosion Distances of BETA Test V2.1
It.A
BETA EXPERIMENT V2.3
500
400
z
300
z0
WE"
200
0
z
100
0
k
0
1000
2000
3000
4000
5000
TIME [s]
Figure 4.17
Comparison between the Predicted and Measured Downward Erosion Distances of BETA Test V2.3
BETA EXPERIMENT V3.3
500
r
z
400
300
0
-
200
z
100
0k
0
500
1500
1000
2000
2500
3000
TIME [s]
Figure 4.18
Comparison between the Predicted and Measured Downward Erosion Distances of BETA Test V3.3
..lost
BETA EXPERIMENT VO.2
2800
II
'
I
j
I
I
'
'
I
v
I
EXPERIMENT
r-w
-
2600
--
-FC
-
- ---
MODEL
RC MODEL
GF MODEL
2400
2200
2000
I,
-.
E-
1800
-
1600
Solidu.
.
Temper ture
.-
4
Concrete Decomposition Temperature
1400
0
100
200
300
400
500
600
TIME [s]
Figure 4.19
Comparison between the Predicted and Measured Metallic Layer Temperatures of BETA Test VO.2
losIt
BETA EXPERIMENT V1.3
2600
*
I,
--
2400
-
-.-.-.
EXPERIMENT
FC MODEL
RFC MODEL
GF MODEL
2200
Emu
2000
CA3
Q
1800
-
Soliusu
1600
-
Concrete Decomposition Temperature
Temperature
1400
0
400
200
600
800
TIME [s]
Figure 4.20
Comparison between the Predicted and Measured Metallic Layer Temperatures of BETA Test V1.3
'I Iat
BETA EXPERIMENT V1.5
2600
r-,
2400
-
-eEXPERIMENT
FC MODEL
- RFC MODEL
--
GF MODEL
2200
E- 2000
\
S1800
-~.*
__
1600
Solidus
..
Temperature
Concrete Decomposition Temperature
1400
12001000
F
0
t
600
1200
1800
2400
3000
3600
TIME [s]
Figure 4.21
Comparison between the Predicted and Measured Metallic Layer Temperatures of BETA Test V1.5
BETA EXPERIMENT V1.6
2600
EXPERIMENT
-FC MODEL
-
2400
-
-
-.--.
RFC MODEL
GF MODEL
2200
m
2000
E-
1800
Solidus Temperature
1600
Concrete
o
--
C)'
Decompoition
Temperature
1400
0)
400
200
600
800
TIME [s]
Figure 4.22
Comparison between the Predicted and Measured Metallic Layer Temperatures of BETA Test V1.6
1.
10 a
BETA EXPERIMENT V1.7
2800
EXPERIMENT
-
2600
--
-FC
-
-
MODEL
RFC MODEL
-.---- GF MODEL
2400 --
-
2200 2000
1800S1600
Solidus Temperature
-Concrete
Decomposition
Temperature
100
200
1400
0
300
400
500
TIME [s]
Figure 4.23
Comparison between the Predicted and Measured Metallic Layer Temperatures of BETA Test V1.7
BETA EXPERIMENT V1.8
2600
EXPERIMENT
*
- - FC MODEL
-
2400
-
-.
RFC MODEL
GF MODEL
-
S 2200 --.--.
02000E---
1800
i
Solidus Temperature
-
1600
-
Concrete DecompoSition Temperature
14001
0
100
200
300
400
500
TIME [s]
Figure 4.24
Comparison between the Predicted and Measured Metallic Layer Temperatures of BETA Test V1.8
-
BETA EXPERIMENT V1.9
2600
EXPERIMENT
FC MODEL
2400
-
RFC MODEL
---.--- GF MODEL
-'Q
A
2200
-
2000
Solidus
1800
1600
Temperature
-
Concrete Decomposition Temperature
1400
0
500
1000
1500
2000
2500
TIME [s]
Figure 4.25 Comparison between the Predicted and Measured Metallic Layer Temperatures of BETA Test V1.9
losta
-
-
.WIKl"
BETA EXPERIMENT V2.3
2600
-F
~' 2400 --
I
EXPERIMENT
FC MODEL
- RFC MODEL
.....
GF MODEL
2200
Q>
2000
1800
Temperature
-Solidus
1600
Concrete Decomposition Temperature
1400
II
0
400
'
I
I
I
800
1200
1600
2000
TIME [s]
Figure 4.26
Comparison between the Predicted and Measured Metallic Layer Temperatures of BETA Test V2.3
dw
BETA EXPERIMENT V3.3
3000
,
1
,
-OEXPERIMENT
FC MODEL
2800
-
-
RFC MODEL
2600 ....... GF MODEL
2400
2200
o 2000
S1800
Solidus Temperature
1600
--
16Concrete
Decomposition Temperature
14001200
0
100
200
300
400
500
600
TIME [s]
Figure 4.27
Comparison between the Predicted and Measured Metallic Layer Temperatures of BETA Test V3.3
than the concrete decomposition temperature. It is believed that when the thermocouples dipped into the melt, a rapid formation of melt crust surrounding the
thermocouple resulted in a melt temperature rea:ding which is lower than the actual melt temperature. Another possible reason for the initial cooling of the melt
which was not reproduced in the calculation could be splashing of the melt on the
wall when it was poured into the cavity. This can effectively cool down the melt,
but is not considered by the computer model. Of course, it may be also caused by
underestimation of the sideward and upward heat losses in the code calculation.
The amounts of the released gases predicted by the FC model are shown from
Figs. 4.28 through 4.31 for several typical tests. It shows that CO and CO 2 are
the major components of the released gas with the KfK concrete, while H2 and
H 2 0 are dominant releases in the Limestone/Common Sand concrete test. This
prediction agrees qualitatively with the experimental observation.
(i) BETA V1.5 Test
r
In the analysis of test V1.5, both the FC and RFC models overpredict the
axial erosion at early times and then converge to experimental results later on.
The radial erosion is slightly underestimated by these models. In the GF model
application, the axial erosion is underpredicted especially at the end of the experiment.
The experimental results showed that the interaction started with a stable
gas film and then collapsed at about 400 seconds later. However, this initial stable
gas film cannot be predicted by the film collapse models based on the criterion
obtained from the data fitting of the first four BETA tests. This disagreement
may be caused by the uncertainty of the initial melt temperature of V1.5.
The GF model gives roughly equal downward and sideward erosion. The FC
and RFC models predict a dominant downward erosion which agrees with the
171
BETA EXPERIMENT V1.3
101
100
0
s
10-1
z
10-3 L
0
500
1000
1500
2000
TIME [s]
Figure 4.28
Gas Generation Rates of BETA Test V1.3 Calculated by the Film Collapse Model
BETA EXPERIMENT V1.9
101
0
w>
100
LJ
F"
zI
0
10-1
-1
P
10-2
0
500
1000
1500
2000
2500
TIME [s]
Figure 4.29
Gas Generation Rates of BETA Test V1.9 Calculated by the Film Collapse Model
BETA EXPERIMENT V2.3
101
0
0
100
L-j
z
10 -1
-1
z
10-2
0n
0
1000
3000
2000
4000
5000
6000
TIME [s)
Figure 4.30 Gas Generation Rates of BETA Test V2.3 Calculated by the Film Collapse Model
*I~pit
BETA EXPERIMENT V3.3
101
mCV2
U,
4)
0
S
100
L~J
z
1 -1
0
z
102
0
(I)
0
1000
2000
3000
4000
TIME [s]
Figure 4.31
Gas Generation Rates of BETA Test V3.3 Calculated by the Film Collapse Model
experimental observation. This is also true for the analyses of the other BETA
experiments.
(ii) BETA V1.6 Test
The power input of test V1.6 was initially sustained at 1000 kW. After 300
seconds, the power was reduced to below 500 kW until the end of the experiment.
Before the power reduction, the calculated erosion distances based on the RFC
model are in fairly good agreement with the experimental data, while the FC
model overpredicts and GF model underestimates the concrete erosion. After the
power reduction, the experimental data showed a significant reduction in the axial
erosion rate (from 0.6 mm/s to 0.12 mm/s) which can not be reproduced with any
of the heat transfer models. This could be caused by overestimation of the melt
temperatures. In Fig. 4.22, the calculated temperature histories show an enhanced
cooling at 300 second, but still not enough to follow the measured cooling rate of
the melt.
(iii) BETA V1.7 Test
With an initial melt temperature of 2373 K in test V1.7, the interaction
predicted by the film collapse models starts with a periodic contact mode. The FC
model significantly overpredicts while the RFC model moderately overestimates
the downward erosion distance. Nevertheless, the experimental data do not show
the existence of an initial stable gas film . The measured erosion rates, during both
the initial transient period and at steady-state, are higher than the predictions of
the GF model by a factor of two to three.
In test V1.7, a power reduction from more than 1500 kW to 200 kW occurred
at 250 seconds. However, the experimental data showed a smooth decrease in the
downward erosion rate after the power reduction. This is quite different from the
V1.6 observation.
176
(iv) BETA V1.8 Test
The measured axial erosion rate of test V1.8 was 1.0 mm/s, which is the highest erosion rate ever measured in real material experiments. All the heat transfer
models used in the analysis cannot predict such a high erosion rate. The best
prediction is calculated by the FC model (no initial gas film) which gives a downward erosion rate of 0.8 mm/s. The RFC and GF models further underpredict
the erosion distance with erosion rates of 0.6 and 0.2 mm/s, respectively.
Comparison between the tests results of V1.7 and V1.8 shows that V1.8 has
twice the erosion rate while its initial melt tempeature is 300 K lower than that of
V1.7. The reason for the extremely high erosion rate in test V1.8 may be its high
power input. However, in the code prediction, the axial erosion is affected mildly
by the power input.
The calculated melt temperatures shown in Fig. 4.24 remain constant or even
rise slightly for the first 300 seconds, while the experimental data showed a rapid
initial cooling rate despite the high power input.
Both the axial erosion and
melt temperature data of test V1.8 suggest that there is an extremely effective
downward heat transfer between the melt and concrete, and this heat transfer is
higher than the prediction of any of the models. The FC model overestimates the
downward erosion of most other experiments, but still falls short in the prediction
of the concrete erosion of test V1.8.
(v) BETA V1.9 Test
Test Vi.9 was a good experiment in terms of the duration and the power input
density. This test was sustained for 2400 seconds with initial melt temperature
of 2173 K.
The average power density induced in the melt was 0.65 kW per kg
melt mass which is comparable with the decay power in a real reactor accident
situation.
177
As predicted by the film collapsed models, the interaction starts with the
periodic contact mode. The agreement between the experimental data and the
results of the RFC model, both in the axial erosion and melt temperature, is
excellent, while the calculated results based on the FC model are less accurate.
The GF model, as usual, underpredicts the axial erosion significantly.
(vi) BETA V2.1 and V2.3 Tests
The erosion data of the low power test showed that a fast erosion of 0.4 mm/s
occurred during the first one hundred seconds. After the initial periods, the erosion
rate was reduced by an order-of-magnitude to 0.03 mm/s for about 2000 seconds,
and then started increasing to 0.08 mm/s and remained approximately the same
until the end of the experiment at 4500 seconds. One possible scenario is that the
steel melt formed a crust at these low power levels soon after contacting the cold
concrete. After a while, the crust would begin to remelt and the concrete ablation
rate would increase.
However, the measured temperatures showed no trend of
remelting.
The possibility of an initial stable gas film is ruled out due to the relatively
low initial melt temperature (about 300 K lower than required for stabilization of
a gas film). If an initial gas film was stabilized, it would have been collapsed at
earlier time, before 2000 seconds, due to the low power input.
Compared to the experimental data, the RFC model significantly overpredicts
the downward erosion at the initial periods, and then converges to experimental
results later on. This is due to the delay of the solidification. In the RFC model
application, the calculation starts with a periodic contact mode and predicts the
formation of a bottom crust at 300 seconds later without remelting throughout
the experiment. The axial erosion rates predicted by the RFC model before and
after the melt solidification are 0.4 and 0.08 mm/s, respectively. The FC model
predicts the same trend with a little higher erosion rate than the RFC model. The
178
GF model shows the best agreement with the low power test data. It follows the
slow erosion at the initial periods, but can not match the increase of the measured
erosion rate at later times.
(vii) BETA V3.3 Test
A nonuniform concrete erosion pattern was observed in test V3.3. As shown
in Fig. 4.18, certain erosion levels have more than one data point, and these
correspond to the failure times of thermocouples located at the same vertical but
different horizontal positions.
In this test, an initially stable gas film collapsed at 570 seconds, as also predicted by the RFC and FC models. The axial erosion is underestimated for the
initial periods before the gas film collapsed. After the film collapse, the calculated
erosion distances increase and follow the trend of experimental data. At the end
of the experiment, the RFC and FC models slightly underestimate the total axial erosion distance. Unlike the other analyses, the RFC model predicts a higher
downward erosion than the FC model in the Limestone/Common Sand concrete
test. It is interesting to note that the RPC model, as discussed in the previous
chapter, predicts a lower heat flux with the KfK concrete but a higher heat flux
with the Limestone/Common Sand concrete than the PC model. Based on the
experimental observation, it can be seen that the revised model is more adequate
in predicting the cases with different concretes.
4.3.2.2 Sandia Experiments
The downward concrete erosion distances of the Sandia experiments are shown
from Figs. 4.32 through 4.36. In general, the RFC and GF models are better than
the FC model in the predictions of the downward erosion distances of the Sandia
experiments.
For sustained heating experiments (SWISS), the RFC model is able to produce
a fairly good agreement in the axial erosion distance. While in the transient tet
179
SANDIA SWISS- I EXPERIMENT
500
400
z
C,)
300
z0
200
O-A
z
100
0
0
L
0
500
1000
1500
2000
2500
TIME [s]
Figure 4.32 Comparison between the Predicted and Measured Downward Erosion Distances of SWISS-1 Test
SANDIA SWISS-2
EXPERIMENT
500
400
z
300
z
0
U)
I A
200
100
0
0
500
1000
1500
2000
2500
TIME [s]
Figure 4.33
Comparison between the Predicted and Measured Downward Erosion Distances of SWISS-2 Test
SANDIA TURC-IT EXPERIMENT
120
r,1
7 100
14
80
z
U)
0
60
40
z
20
0
0
200
400
600
800
1000
TIME [s]
Figure 4.34 Comparison between the Predicted and Measured Downward Erosion Distances of TURC-1T Test
lost
SANDIA TURC-1SS EXPERIMENT
50
r
40
z
30
C)
20
z
0
10
9z
OL
0
40
60
80
100
120
TIME [s]
Figure 4.35
Comparison between the Predicted and Measured Downward Erosion Distances of TURC-1SS Test
SANDIA TURC-2
EXPERIMENT
10
8
z
6
z0
Ul)
00
0
4
2
0
0
100
300
200
400
500
600
TIME [s]
Figure 4.36
Comparison between the Predicted and Measured Downward Erosion Distances of TURC-2 Test
lost
(TURC) analyses, the RFC model cannot predict the axial erosion very well in
most cases.
(i) Sandia SWISS Tests
The temperature histories of test SWISS-1 are shown in Fig. 4.37. The experimental data shown in this figure from different publications [G11,G12] are
inconsistent. It is important to notice that the initial melt temperature indicated
by Ref. [G11] and [G12] was 1925 K and 2000 K, respectively. This initial temperature difference is considered to be small and within the uncertainty range.
However, it has significant effect in determining the stabilization of an initial gas
film. With the initial melt temperature of 1925 K, both the FC and RFC models
will predict a periodic contact mode for SWISS tests. In the code calculations,
the initial melt temperatures of the SWISS tests (both SWISS-1 and SWISS-2)
were specified as 2000 K.
At this initial temperature, the RFC model predicts
an initial stable gas film while the FC model predicts a periodic contact mode.
The reasons for these different predictions of the initial gas film between the FC
and RFC models are: (1) the downward heat flux at the initial conditions of the
SWISS test predicted by the PC model is lower than the RPC model; (2) the
multiplier of the Kutateladze's limit used in the FC model is higher than that in
the RFC model.
In the RFC model calculations, the gas film stabilized initially is sustained
throughout the whole period of experiment without collapsing.
Therefore, the
calculated axial erosion distances of the RFC model are exactly the same as those
of the GF model, and are in fairly good agreement to the experimental data.
However, the melt temperatures are significantly overpredicted by both the RFC
and GF models (see Fig. 4.37). In the FC model calculations, the axial erosion
distances of SWISS tests are overpredicted by a factor of two compared to the
experimental results, even though the calculated melt temperatures are close to
the experimental data.
185
SNADIA SWISS-1
EXPERIMENT
2400
2300
2200
'Un
2100
2000
r
1900
1800
1700
1600
1500
0
5
10
15
20
25
30
35
40
TIME [min]
Figure 4.37 Comparison between the Predicted and Measured Melt Temperature Histories of SWISS-1 Test
186
(ii)
Sandia TURC Tests
Since both the FC and RFC models predict a stable gas film throughout the
whole period of tests TURC-1T and TURC-1SS, the calculated erosion distances
by various models are exactly the same. The predicted erosion distances of test
TURC-1T agree closely with the experimental data. While in the TURC-1SS
analysis, the predicted erosion distances are significantly lower than the experimental data by a factor of three. The highly irregular erosion data of TURC-1SS
indicate that the interaction may have started with a gas film and switched to
the periodic contact mode 20 seconds later. The failure of the predictions may be
caused by the uncertainties of the initial melt temperature and the heat losses to
the MgO sidewall.
TURC-2 test is the only real material experiment with oxidic melt.
The
solidus temperature of the oxidic melt calculated by CORCON is 2825 K, and the
reported initial melt temperature was 2820 K. Therefore, in the code calculations
the melt/concrete interaction is governed by a solidified pool. As discussed in the
pervious chapter, the downward heat fluxes of the post-freezing stage predicted by
the various models are about the same, therefore, the calculated erosion distances
of the TURC-2 test with different models are almost equal.
These predicted
erosion distances are higher than the experimental data by a factor of two.
It
seems that the downward heat transfer of a solidified pool is overestimated by
the post-freezing models. However, before more data corresponding to the postfreezing regime become available, a firm conclusion is hard to be made.
4.3.3 Model Validation
The average downward erosion rates of the BETA tests, both the experimental
and calculational results, are summarized in Table 4.8. The total erosion distances
predicted by the various heat transfer models versus the measured distances are
shown in Fig. 4.38.
187
Table 4.8
Axial Concrete Erosion Ratest of the BETA Experiments
t
Test
Experiment
Data
GF
Model
PC
Model
FC
Model
RPC
Model
RFC
Model
VO.2
0.160
0.106
0.233
0.185
0.223
0.196
VO.3
0.866
0.186
0.800
0.800
0.624
0.624
V1.2
0.241
0.0911
0.234
0.181
0.219
0.180
V1.3
0.500
0.151
0.551
0.551
0.450
0.450
V1.5
0.231
0.114
0.280
0.280
0.266
0.266
V1.6
0.337
0.159
0.580
0.580
0.486
0.486
V1.7
0.533
0.197
0.907
0.907
0.706
0.706
V1.8
1.036
0.184
0.815
0.815
0.622
0.622
V1.9
0.145
0.0786
0.165
0.165
0.153
0.153
V2.1
0.0649
0.0502
0.0902
0.0902
0.0868
0.0868
V2.3
0.0893
0.0678
0.108
0.108
0.105
0.105
V3.3
0.136
0.0626
0.229
0.109
0.215
0.118
Average erosion rate in units of mm/s over the experiment duration.
188
DOWNWARD
EROSION DISTANCE
700
600
500
-'400
z0
ip
Z
300
200
100
0
0
100
200
300
400
500
600
700
EXPERIMENT [mm]
Figure 4.38
Downward Erosion Distances of BETA Tests (Prediction versus
Experiment)
189
In order to examine the trend of erosion with respect to the power input and
the initial melt temperature, a decomposable concrete volume, VDC, is defined
and plotted against the axial erosion distance for .each BETA test in Fig. 4.39.
The decomposable concrete volume is defined by:
VDC =
QTotal
PconcHDecomp
(4.1)
where QTotal is total available energy and given by:
QTotal = Qinput
where
Qinput
+ [(mcp)metal + (mcp)ozide] ATi
(4.2)
is the total energy input shown in Table 4.4, and ATi is the temper-
ature difference between the initial melt temperature and a reference temperature.
Since the melt temperatures approached the metallic solidus point at the end of
the BETA tests, the reference temperature in the calculation of the sensible heat
is selected to be the metallic solidus temperature (1760 K). The heat capacities
of the metallic and oxidic materials are 790 and 1780 J/kg K, respectively. Based
on this definition, the decomposable concrete volume appears to be a proper combined index of the power input and the initial melt temperature. In Fig. 4.39, it
can be seen that the axial erosion increases with the increase of the power input
and the initial melt temperature. Least square fits of the predictions of various
models are also shown in the figure. Both the FC and RFC models show good
agreement, within the data spread. Individual erosion distances of the predictions
relative to the experimental data are shown in Fig. 4.40.
By using the definitions specified in Table 4.9, standard deviations in the
predicted downward erosion are calculated for various heat transfer models.
As
indicated in the table, the GF model shows a significant discrepancy with the
experimental data. It can also be seen that the revised periodic contact model is
better than the original one. Among various downward heat transfer models, the
190
0.6
I
I
I
I
I
I
I
I
I
I
I
EXPERIMENTAL DATA
GF MODEL
FC MODEL
RFC MODEL
0.5
0. 4
z
Cn)
z0
0.3
r
0.2
0.1
0.0L
0.10
I
I
I
I
I
I
I
0.15
I
I
I
0.20
I
I
I
I
0.25
DECOMPOSABLE CONCRETE VOLUME [m3]
Figure 4.39
Least Square Fits of the Various Heat Transfer Models on the Predictions of the Downward Erosion of BETA Tests
191
-I
2.0
z
1.8
E- 1.6
1 .4
C-)
cn 1.2
1.0
Q
0
z
0.8
0.6
cn 0.4
0 .2
0.0I0.10
0.15
0.20
0.25
DECOMPOSABLE CONCRETE VOLUME [M 3 ]
Figure 4.40 Relative Downward Erosion Distances of BETA Tests
192
Table 4.9
Statistics for the Various Heat Transfer Models
in the Calculations of the BETA Results
GF Model
PC Model
FC Model
RPC Model
RFC Model
0.564
0.409
0.349
0.319
0.266
-0.533
0.275
0.158
0.133
0.046
Standard Deviation
(6ca)i - (6ep)i]2
\n
( 6 ep)i
Mean Relative Error
1
n
[( 6 cal)i - (bezp)i
n
(bep)i
193
RFC model is the best in the prediction of the downward concrete erosion of the
real material experiment.
4.4 Conclusions
1. The gas film model, developed earlier and commonly used in severe accident
analysis, significantly underestimates the downward heat transfer in many
experiments.
2. The film collapse model is more adequate than either of the gas film or the
periodic contact model alone.
3. The revised film collapse model is better than the original one in predictions
of both BETA and Sandia experiments.
4. The revised film collapse model is capable of producing erosion results of the
BETA experiments with a mean error of 5% and a standard deviation of
27%. Less accuracy is found in the predictions of transient, one-dimensional
experiments performed at Sandia National Laboratory.
5. Based on the revised and original film collapse models, the increase in axial
erosion with power input and initial melt temperature is predicted, within
the data spread.
6. The results of certain low power tests (BETA V2.1 and V2.3 in particular)
do not fit the pattern set by the other experiments. A reexamination of data
uncertainty is warranted.
7. The initial rapid cooling of the melt observed in BETA tests cannot be successfully predicted.
Further investigations of the initial heat losses in the
sideward and upward directions are necessary. Models for the splashing and
layer mixing phenomena may be needed to explain the initial melt temperature behavior.
194
CHAPTER 5
SENSITIVITY STUDY OF THE EX-VESSEL SOURCE TERM
5.1 Objective
The interaction of molten core with the concrete cavity of a reactor containment is an important phase in a severe reactor accident. Major concerns in the
safety assessment include: the possibility of a basemat meltthrough; containment
over-pressurization due to the generation of noncondensible gases; and most importantly the release of the fission products from the corium pool if containment
fails. During the molten core/concrete interaction, radioactive fission products as
well as nonradioactive materials can be released as aerosols from the molten pool
into the containment atmosphere by either evaporation or mechanical processes.
This ex-vessel source of aerosols has a direct impact on the estimation of the radiological consequence of a severe reactor accident. The magnitude, physcial and
chemical characteristics, and timing of the ex-vessel aerosol release, determined
largely by the heat transfer between the molten core and the concrete, are important factors affecting the potential amount of radioactive fission product that
could be released from the containment building if it fails.
The primary purpose of this chapter is to identify the impact of the downward
heat transfer model on the concrete erosion, gas generation and ex-vessel aerosol
release in analysis of a real reactor case. A parametric study on the significant
variables, such as (1)initial melt temperature; (2)concrete properties; (3)amount of
unoxidized zirconium; (4)amount of melt; (5)decay heat; and (6)layering potential
of melt constituents, is performed to identify the important source of uncertainties
in calculation of the ex-vessel aerosol release.
195
5.2 Introduction
5.2.1 Characteristics of Ex-Vessel Source Term
Exposure of concrete to the high temperature melt will result in the decomposition of concrete and release of steam and noncondensible gases. As the steam
and noncondensible gases bubble through the molten debris, they react chemically
with the melt constituents, and also provide a tremendous free surface within the
pool to enhance the vaporization process of the high vapor pressure constituents
of the melt.
The vaporized materials, both radioactive and nonradioactive, af-
ter emerging from the molten debris into the containment atmosphere, can be
condensed as aerosol particles due to the relatively low temperature of the containment atmosphere.
This is the evaporation process of forming the ex-vessel
aerosols. Once gas bubbles reach the molten pool surface, some amount of surface
melt can be injected into the containment atmosphere as aerosol-sized droplets
due to bubble burst. This is a mechanical process of producing ex-vessel aerosols.
Aerosol production during MCCI is not as intense but lasts far longer than
aerosol production during in-vessel phases of an accident. Aerosols produced during MCCI are predicted to consist primarily of nonradioactive material, from control rods, structures, and concrete. These nonradioactive aerosols are important
in the prediction of radionuclide aerosol behavior in the containment. The aerosols
evolved from the corium pool will mix with the radioactive aerosols released earlier
during the in-vessel phase of an accident, and then enhance the agglomeration and
sedimentation in the containment atmosphere. The amount of radioactive aerosols
suspended at the time of containment failure is the source that could be released
from the containment to the public. In order to properly assess the radiological
consequence of a severe reactor accident, it is necessary to understand the phenomena governing the physical and chemical processes involved which could lead
to the release of ex-vessel aerosols.
196
5.2.2 Background
Early research on severe accidents has pointed out the need to integrate the
analysis of complex severe accident phenomena'to obtain realistic estimates of
source terms. While recent studies have permitted advances in the analysis of
source terms, the improved understanding has also allowed a more clear identification of the major uncertainties.
In the QUEST study [L1), uncertainties were grouped into two categories:
code input uncertainties and code phenomena uncertainties. Code input uncertainties were investigated by allowing certain important user-input parameters to
vary over a reasonable range. Phenomena uncertainties were examined by varying
values for the important phenomena described in the code. Since there were numerous parameters involved in the BMI-2104 code set, it was impossible to have
all parameters analyzed in detail. A scheme was employed to determine which
parameters warranted inclusion in the study. This scheme addresses the codes in
the BMI-2104 suite in reverse order starting with NAUA. A sensitivity study of
NAUA was performed first by varying one input parameter by output from another code or directly by the user -
whether provided
at a time about a base
case. If a parameter had a strong influence on the interested output, for example
the suspended aerosol mass as calculated by NAUA, it was identified as an important parameter. Once an input parameter of a downstream code provided by
the output of a upstream code was identified to be important, sensitivity study
of the upstream code was performed in the same fashion to determine the effect
of each input parameter on that specific output. In this manner, the entire suite
of codes was examined to determine a list of the most important parameters for
further quantitative evaluation.
In the BMI-2104 code suite, CORCON/MOD1 and VANESA were used to
predict the behavior of ex-vessel aerosol during the MCCI. The estimates obtained
from the VANESA model are sensitive to the features of the plant and accident
197
in question. The VANESA calculations are quite dependent on initial conditions
specified as input to the model. These initial conditions are typically obtained
from models such as MARCH and CORSOR. The calculations are also somewhat
sensitive to the boundary conditions, such as the melt temperatures, gas generation rates which depend on the modeling of corium/concrete interactions and the
nature of concrete assumed present in the plant.
More specifically, the aerosol
releases calculated by the VANESA model increase exponentially with increasing
temperature of the melt, and vary approximately linearly with gas generation rate
and concentration of volatile species in the melt. Therefore, anything that could
cause an increase in the melt temperature, the gas generation rate or the concentration of volatile species in the melt would increase the ex-vessel aerosol release.
For example, the increase of the heat transfer into the concrete will increase the
concrete ablation and gas generation rate, thus it will increase the aerosol release.
On the other hand, the melt temperature and the concentration of volatile species
will be depressed due to the increased heat transfer, therefore, the aerosol generation will be reduced. The total effect of such parameters needs to be examined
by integral analyses.
The input parameters examined by the QUEST study were rank-ordered in
terms of decreasing effect on the ex-vessel source term as follows [L1]:
o Amount of core involved in the MCCI
o Amount of zirconium in the melt
e Amount of steel in the melt
o Initial debris temperature
e Reactor pressure vessel failure time
* Concrete type
. Water content of the concrete
* Free energies of gas species
. Cavity floor area
As mentioned in QUEST, the results of the sensitivity study and the above rank
196
ordering apply only to calculations involving single-parameter variation around
the TMLB' accident at the Surry plant and ought not be generalized.
Since the updated version CORCON/MOD2 was not available at the time
of the QUEST study, CORCON/MOD1 was used. Unfortunately, it should be
noted that while the initial melt temperature specified in the QUEST for most of
the cases is far below the solidus temperature of the melt, the CORCON/MOD1
is applicable only to the high temperature phase of the MCCI when the melt is
hot enough to be entirely liquid to erode the concrete at a relatively rapid rate.
There are reasons, as discussed in the QUEST report, to believe that the CORCON/MOD1 code does not, in fact, calculate reliable melt temperatures. In some
cases, such as the high zirconium content and low steel mass cases, the melt experiences a dramatic temperature excursion during the MCCI. Melt temperatures can
reach completely unjustifiable levels -
in excess of 3200 K during this excursion.
Therefore, the effects of those parameters on the ex-vessel source term could not
be reliably ascertained in QUEST.
Furthermore, the completeness of the list of important parameters in QUEST
needs to be examined. Because only limited cases have been analyzed, it is likely
that some other parameters which could be important were not identified.
For
instance, the importance of the amount of FeO or the solidus temperature of the
oxidic melt related to the user-input uncertainties could be underestimated due to
improper post-freezing modeling in CORCON/MOD1. Also, the effects of various
parameters on the source term based on different initial melt temperatures still
remain questionable.
With respect to the code phenomena uncertainties, the effect of the downward
heat transfer model has never been studied. As identified in the previous chapter,
the gas film model adopted in the original CORCON code is not adequate in
describing the downward heat transfer of the melt/concrete interaction. Both the
periodic contact and film collapse models are better than the gas film model in
199
calculation of the experimentally observed concrete erosion and melt temperatures.
The downward heat flux effect on the calculation of the ex-vessel source term based
on the CORCON/MIT is described in this study.
5.3 Formulation of This Study
5.3.1 Analysis Methods and Computer Codes
The methods used to identify the uncertainties in the calculation of ex-vessel
aerosol resulting from various components are similar to the QUEST study, in
which calculations were performed by varying one input parameter at a time about
a base case.
Instead of the CORCON/MOD1 used in the QUEST study, the revised version CORCON/MIT was used in this study to predict the behavior of the corium/
concrete interaction. The VANESA code was used to calculate aerosol generation
and the releases of radionuclides based on the time dependent melt temperatures,
gas flow rates, concrete erosion mass and aerosol release area from the CORCON/MIT output.
To avoid possible typo-error during data transfer, VANESA has been coupled
with CORCON/MIT in a simple way. In the simplified linkage, necessary data calculated by the CORCON/MIT are stored temporarily during execution and then
transferred to VANESA at the end of the CORCON/MIT calculation. No attempt
has been made to couple the calculations of these two codes step by step. The releases of radionuclides and non-radioactive materials calculated by the VANESA
model do not feedback to the CORCON/MIT during each time step to update
the amount of melt and the decay heat level in the concrete cavity. However, the
amounts of the released materials during the MCCI are relatively small compared
to the total inventory in the concrete cavity. The possible error due to the simplified procedure could be negligible. A more significant error from the simplified
coupling belongs to the chemical reaction heat. Since the chemical reaction packages in the CORCON and VANESA are different, inconsistent calculations of thie
200
chemical reaction heat generation and zirconium depletion time would occur. In
CORCON, the chemical reacton is calculated by thermodynamic analysis based
on minimization of the Gibbs function of the system. In other words, the released
gases and melt constituents are assumed to be in thermodynamic equilibrium and
the chemical reaction in the system achieves a maximum rate. There can be, however, barriers that prevent or retard achieving the maximum reaction rate defined
by the thermodynamic analysis. In VANESA, the extent of the chemical reaction
is treated by kinetics theory which would reduce the reaction rate and decrease
the zirconium depletion rate if the system deviates from thermodynamic equilibrium. However, there is no hard evidence at this moment to indicate a significant
deviation from equilibrium during the corium/concrete interaction.
Because of
the elevated temperature, the impact of the chemical reaction difference on the
ex-vessel source term could be small.
The first correction set for CORCON/MOD2 (version 2.00) was issued from
the Sandia National Laboratory in July 1986. This correction set (version 2.01)
eliminates many of the coding and modeling errors that have been identified since
the release of the code. From calculations for the CORCON Sample Problem, the
results of the versions 2.00 and 2.01 were not widely different. However, it is not
inconceivable that a particular set of input might trigger differences between the
two versions. In this study, major parts of analysis were done based on the version
2.00 before the revised version became available. In addition, the differences between the results of the PC and RPC models were analyzed in several cases based
on the version 2.01.
5.3.2 Input Parameters
5.3.2.1 Specifications of the Base Case
Parameters used as the base case in this study are shown in Table 5.1. The
initial fission products inventory is the same as in the base case of QUEST, and
is given in Table 5.2. The base case conditions that have been chosen in this
201
Table 5.1
Parameters Used in the Base Case Study
2600 K
Initial Melt Temperature (T)
Fraction of Unoxidized Zr in Melt
50.0%
Coolant (H20) Above Corium
60 Mg
Radius of Concrete Cavity
3.0 m
Timing of MCCI after Shutdown
180 minutes
0
Limestone/Common Sand
Type of Concrete
Concrete Decomposition
Temperature (TD)
1500 K
Reactor Nominal Power
3000 MWt
58.8%
Fraction of Core Meltdown
Emissivity of
Concrete
Melt
Surrounding
0.6
0.8
0.8
202
Table 5.2
Melt Inventory of the Base Case (MI) at the Start of MCCI
Species
Mass (kg)
Uranium Oxide (U02)
60000
Zirconium Oxide (ZrO2 )
Ferrous Oxide (FeO)
Iron (Fe)
Chromium (Cr)
Nickel (Ni)
Zirconium (Zr)
Cesium (Cs)
Iodine (I)
Tellurium (Te)
Barium (Ba)
Tin (Sn)
Ruthenium (Ru)
Molybdenum (Mo)
Strontium (Sr)
Rubdium (Rb)
Yttrium (Y)
Technetium (Tc)
Rhodium (Rh)
Palladium (Pd)
Lanthanum (La)
Cerium (Ce)
Praseodymium (Pr)
Neodymium (Nd)
Samarium (Sm)
Plutonium (Pu)
Antimony (Sb)
Niobium (Nb)
Silver (Ag)
8400
3600
18700
2200
1100
6000
0.7
0.1
16.4
49.1
152.0
103.0
140.0
43.7
0.1
22.9
36.7
20.7
52.0
62.3
131.0
50.7
171.0
34.0
469.0
0.31
4.0
1460.0
203
study were similar to the conditions specified in the CSNI benchmark problem
No. 1 [N7]. It has a relatively small amount of metallic melt and high initial melt
temperature compared to the base case of the TMLB' sequence in the QUEST
study.
5.3.2.2 Cases Analyzed
In this study, analyses will primarily focus on the differences that might occur
in the calculation of the ex-vessel aerosol release between the PC model and the
GF model. Both the PC model and GF model were applied to all cases to study
the maximum effects on the aerosol release due to the heat transfer models. The
FC model, a combination of these two, was applied only in some cases to illustrate
its relative tendency with respect to the PC and GF models. Limited cases have
been analyzed based on the RPC model to show the differences compared to the
PC model.
The parameters studied and their application range used in the test are given
in Table 5.3. Detailed melt compositions used in the cases can be found in Table
5.4. Each case has been named by a short-hand representation in the following
sections. For instance, PCM1 Base Case is the one with the specifications of the
base case and melt compositions of M1, and calculated by the periodic contact
model. Another example, GFM4 T = 2320 K represents the case of M4 melt
with initial temperature of 2320 K and calculated by the gas film model, while
the other parameters are the same as the base case.
5.4 Results and Discussions
Effects of uncertain input parameters on the ex-vessel aerosol release analyzed
by the CORCON/MIT-VANESA codes will be discussed and compared to the
QUEST results. Not only the calculated aerosol generation but also those factors
affecting the calculation of the aerosol generation will be discussed. The results of
the various cases presented in the followings include: concrete erosion, gas
204
Table 5.3
Phenomena and Parameters Range Used in
the CORCON/MIT-VANESA Sensitivity Study
Phenomena and Parameters
Case
Downward Heat Transfer Model
Periodic Contact (PC & RPC)
Film Collapse
(FC & RFC)
Gas Film
(GF)
Initial Melt Temperature (Ti)
2600
2400
2320
2000
1807
Melt Composition
K
K
K
K
K
Base Case
(Mi)
Enriched Core Oxide (M2)
Enriched Metal Steel (M3)
Enriched FeO
(M4)
Fraction of Unoxidized Zr in Melt
50 %
20%
0%
Concrete Type
Limestone/Common Sand
Limestone
Basaltic
Concrete Decomposition
Temperature (TD)
1500 K
1670 K
1420 K
Initial Layer Configuration
Metal - aver - Oxide
Oxide - over - Metal
Decay Heat
CORCON
ANS Standard
205
Table 5.4
Compositions of Various Melts Used in
the CORCON/MIT-VANSEA Sensitivity Study
M1 Case
M2 Case
M3 Case
M4 Case
(kg)
(kg)
(kg)
(kg)
60000
90000
60000
60000
ZrO2
8400
12600
8400
8400
FeO
3600
3600
3600
6000
Fe
18700
18700
37400
18700
Cr
2200
2200
4400
2200
Ni
1100
1100
2200
1100
Zr
6000
9000
6000
6000
Oxidic Layer:
U0
2
Metallic Layer:
* Other materials are the same as in Table 5.2.
206
generation, melt temperatures, aerosol generation rate, fission products release
fraction and accumulated release of the ex-vessel aerosol.
It should be mentioned that, in the following presentations, the decontamination factor due to the scrubbing effect of the overlying coolant has been applied
to the total aerosol (radioactive and non-radioactive) release. But this factor has
not been applied to correct individual fission product release.
5.4.1 Impact of the Downward Heat Transfer Models
Estimated melt behavior using different heat transfer models for the base case
is described in Table 5.5.
It is seen that significant differences in the predicted
melt behavior can result from the various models. The downward heat transfer
based on the PC model is higher than the GF model. Therefore, the axial erosion
rate as well as the melt cooling down rate are larger based on the PC model (see
Figs. 5.1 and 5.2).
Results of the FC model before the film collapse, which is
predicted at 860 seconds after the start of MCCI, are the same as those of the GF
model. After the film collapse, the results of the FC model fall between those of
the PC and GF models.
It should be noted that the temperature trends of the oxidic and metallic
layers are essentially identical. An initial sharp decrease of the melt temperature
followed by a slow-varying quasi-steady temperature is calculated with all heat
transfer models in the base case study. Both phases reach below 1800 K almost
at the same time as shown in Fig. 5.2. The temperature difference between the
oxidic and metallic layers at the quasi-steady state calculated by the GF model is
about 50 K, and the temperature differences calculated by the PC and FC models
are even smaller. While in the QUEST study, the temperature behavior of these
two layers calculated by the CORCON-MOD1 models were quite different. The
calculated oxidic temperature was sustained above 2100 K for several hours, while
the metallic temperature dropped rapidly to its liquidus point (1790 K).
207
Table 5.5
Phenomena and Timing of Events of the Base Case
GF
Model
PC
Model
FC
Model
Yes
No
Yes
Zirconium depletion (min)
11
2
11
Layer flip (min)
12
2
12
Initial stable gas film
Film collapse (min)
14
Formation of bottom oxidic
crust before layer flip
No
No
No
Time of bottom metallic
crust formation (min)
40
15
25
Time for oxidic temperature
drops below 1720 K (min)
63
27
36
Coolant left after 3 hours
interaction (kg)
1900
8400
3900
Oxidic solidus temperature
after 3 hours interaction (K)
1657
1615
1642
208
800
700
600
500
z
400
z0
300
200
100
0
0
30
60
90
120
150
180
TIME [min]
Figure 5.1
Concrete Ablation Distances Predicted by Different Heat Transfer
Models
209
3000
2800
2600
I,
2400
2200
2000
1800
1600
1400
0
10
20
30
40
50
60
TIME [min]
Figure 5.2
Melt Temperature Histories Predicted by Different Heat Transfer
Models
210
The calculated gas generation rates are given in Fig. 5.3. It is shown that
initially the gas generation rate calculated by the PC model is five times higher
than the GF model. However, at the quasi-steady .state, the gas generation rate
of the PC model is lower than the GF model because lower melt temperature is
calculated by the former model. A sudden increase of the gas generation due to
the coking effect after the depletion of zirconium can be seen in the GF and FC
models. However, this sudden surge of the gas generation is compensated for by
the sharp decrease of melt temperature calculated by the PC model, therefore,
there is no peak in the gas generation rate in the PC model predictions. Increased
gas generation rate due to film collapse calculated by the FC model also can be
seen in Fig. 5.3.
In Fig. 5.4, the aerosol generation rate calculated with the PC model is 1 kg/s
initially, and it drops to below 10 g/s in 15 minutes of interaction due to significant
decrease in the melt temperature. In the GF and FC applications, an initial aerosol
generation rate of 200 g/s, 5 times less than that of the PC model, is calculated.
However, in less than 10 minutes, the aerosol generation rates calculated by both
the GF and FC models are higher than that of PC model. The coking effect of
the increase in gas generation gives an increase in the aerosol generation rate at
11 minutes after the start of MCCL. With both GF and FC models, the aerosol
generation rates drop below 10 g/s at 35 and 26 minutes, respectively.
Using all heat transfer models, a dramatic decrease, about an order-ofmagnitude, in the areosol release rate is calculated at different times (27 minutes
for PC model, 36 minutes for FC model and 63 minutes for GF model). The reason
for this sharp decrease is that there exists a critical oxidic temperature (1720 K)
in the VANESA model, above which the vaporization of calcium oxide is viable.
For the base case analysis, the aerosol composition calculated by VANESA is dominated by the calcium oxide. As soon as the oxidic temperature drops below 1720
K, the vaporization process of calcium oxide and hence the areosol generation rate
211
C
L-s
zC
r
106
z
0
U)
0
105
0
30
60
90
120
150
180
TIME [min]
Figure 5.3
Gas Generation Rates Predicted by Different Heat Transfer Models
212
II
Lj
102
z
101
0
C12
100
10-1
10-2L
0
30
90
60
120
150
180
TIME [min]
Figure 5.4
Aerosol Generation Rates Predicted by Different Heat Transfer
Models
213
is reduced significantly. The timing of this significant drop in the aerosol generation
rates calculated by various models coincides with the time required for the oxidic
temperature to reach below 1720 K as listed in Table 5.5. After this sharp decrease,
the aerosol generation rates from the various models fall between 0.2 and 0.4
g/s.
In the QUEST study, because of the relatively high oxidic temperature
calculated by the CORCON-MOD1, the calculated aerosol generation rates in all
cases remain above 1 g/s.
In Fig. 5.5, it is shown that the total aerosol mass production is mildly affected by the different heat transfer models except in the first ten minutes. The
accumulated aerosol release calculated by the PC model shows a rapid initial rise
and reaches 180 kg in 15 minutes. After that, it almost stays constant for three
hours of MCCI. As for the GF model calculation, the accumulated aerosol achieves
200 kg release at 40 minutes without further significant increase. Results of the
FC model are about the same as the GF model.
However, the radionuclide releases resulting from different heat transfer models have significant differences.
As indicated in Fig. 5.6, the initial lanthanum
release rate calculated by the PC model is five times that of the GF model. The
accumulated release fractions of lanthanum are shown in Fig. 5.7. It is seen that
the PC model gives three times higher total lanthanum release at 3 hours than the
GF model. Differences of the lanthanum release between the calculations of the
FC and GF models are negligible. The releases of other radionuclides are shown
in Figs. 5.8, 5.9 and 5.10. All these figures show similar trends for the impact of
the different heat transfer models.
The rate of the radionuclide release from the vaporization process depends
not only on the melt temperature but also the available free surface. At the initial
temperature of the base case, the gas generation rate, and hence the available
free surface calculated by the PC model is larger than the GF model. Therefore,
higher radionuclides release rates are calculated by the PC model at the start of
214
300
250
C,)
200
C,)
150
100
50
0h
0
20
60
40
80
100
120
TIME [min]
Figure 5.5
Accumulated Aerosol Releases Predicted by Different Heat Transfer Models
215
101
100
r,
10-1
10-2
10 -3
10~5_
z
E--
0-6
10-7
10-8
0
20
60
40
80
100
120
TIME [min]
Figure 5.6
Lanthanum Release Rates Predicted by Different Heat Transfer
Models
216
101
I
-E-
---- 3
RELEASE FRACTION
(La)
LANTHANUM
I
I
PCM1 BASE CASE
GFM 1 BASE CASE
FCMI BASE CASE
100
0
30
90
60
120
150
180
TIME [min]
Figure 5.7
Lanthanum Release Fractions Predicted by Different Heat Transfer
Models
217
TELLURIUM (Te)
10
2
1
1
1
RELEASE
FRACTION
1
Ct)
p
z
so
101
0
30
90
60
120
150
180
TIME [min]
Figure 5.8
Tellurium Release Fractions Predicted by Different Heat Transfer
Models
218
ANTIMONY
(Sb) RELEASE FRACTION
100
z
0
10 -1
0
30
60
90
120
150
180
TIME [min]
Figure 5.9
Antimony Release Fractions Predicted by Different Heat Transfer
Models
219
STRONTIUM (Sr) RELEASE FRACTION
102 1
1
1
1
z
101
0
30
60
90
120
150
180
TIME [min]
Figure 5.10
Strontium Release F~actions Predicted by Different Heat Transfer
Models
220
MCCI. As predicted, the initial gas film assumed by the FC model collapses at
a relatively low melt temperature condition, in other words, it collapses at the
moment when the radionuclide release rate is reduced significantly.
Therefore,
the differences in the calculation of radionuclide release between the FC and GF
models are negligible. Because the melt temperatures calculated by the PC model
drop quickly at the start of MCCI, the radionuclides release rates are promptly
reduced, and no major releases are calculated after the first time step (3 minutes
internval) of the VANESA calculation. While in the GF model application, the
time of the accumulating of radionuclides can extend as long as 30 minutes.
5.4.2 Effect of Concrete Decomposition Temperature
The phase change of the concrete mixture is characterized by its solidus and
liquidus temperatures.
In CORCON, the concrete is always assumed to be de-
composed and melted at a user specified temperature.
The best choice of this
decomposition temperature was suggested as the solidus temperature plus one
third of the difference between the solidus and liquidus temperatures of the concrete
[C3];
however, it may be affected by the gas pressure created below the
decomposed concrete surface. In addition to the suggested value, the solidus and
liquidus points of the Limestone/Common Sand concrete were chosen in this test
to illustrate the effect of the decomposition temperature variation.
The most direct effect of the decomposition temperature is the heat transfer
between the melt and the concrete. The lower the decomposition temperature the
higher is the downward heat flux and, therefore, the higher is the initial gas generation rate. Integral results of the axial ablation distances, the melt temperatures
and the gas generation rates are shown in Fig. 5.11, 5.12 and 5.13, respectively. As
noted, the quasi-steady oxidic temperature is also affected by the decomposition
temperature. The time for the oxidic melt to reach below the critical temperature
(1720 K), and hence the time of the significant decrease of the aerosol generation
rate is delayed by an increased decomposition temperature (see Fig. 5.14). For the
221
800
700
600
z
5 0 0
CC,
*400
z
rI
300
S200
100
0
0
30
60
90
120
150
180
TIME [min]
Figure 5.11
Downward Ablation Distances for Different Concrete Decomposition Temperatures
222
3000
2800
2600
a:' 2400
2200
E-
2000
1800
1600
1400
[
0
10
20
30
40
50
60
TIME [min]
Figure 5.12
Melt Temperature Histories for Different Concrete Decomposition
Temperatures
223
108
0
r
A.
z
E
106
0
30
60
90
120
150
180
TIME [min]
Figure 5.13
Gas Generation Rates for Different Concrete Decomposition Temperatures
224
105
104
0
101
CI)
100
10~1
10-2
0
30
60
90
120
150
180
TIME [min]
Figure 5.14
Aerosol Generation Rates for Different Concrete
Decomposition
Temperatures
225
high decomposition temperature case, the areosol generation rates calculated by
the GF model remain above 2 g/s for three hours of MCCI. The initial aerosol
generation rate increasing with the decrease of the .decomposition temperature is
calculated by the PC model, while in the GF model, the differences of the initial
areosol release rate are small. In Fig. 5.15, it is shown that the total aerosol mass
production is not affected by the variation of the decomposition temperature. A
more significant effect can be found in the lanthanum release. Based on the PC
model, the lanthanum release of the base case is a factor of two higher than the
high decomposition temperature case, while the low decomposition temperature
case has the same results as the base case (see Fig. 5.16). In the GF model, less
significant effect is calculated for the decomposition temperature variation.
In general, the effect of the decomposition temperature on the melt behavior
calculated by the PC model is more significant than that of the GF model, since
the rate of heat transfer in the PC model is more sensitive to the variation of
(Tp - TD) than the GF model (see Fig. 3.7).
5.4.3 Effect of Concrete Type
This test was done to illustrate the effect of using different concretes in the
calculation of the ex-vessel aerosol release.
Three generic concretes specified as
default choices in the CORCON were used in this study. Physical properties of
these concretes can be found in Table 3.4.
Basaltic concrete has the lowest decomposition temperature and enthalpy
which result in the highest axial ablation distance and melt cooling rate, as shown
in Figs. 5.17 and 5.18.
On the other hand, the Limestone concrete, with the
highest decompositon temperature and enthalpy, has the lowest axial ablation
distance and melt cooling rate. It is interesting to note that the quasi-steady
oxidic temperatures for the Limestone concrete calculated by both heat transfer
models remain above 1850 K while those of the other concretes drop below the
critical temperature (1720 K).
226
300
1
-&---
PCM1
PCM1
PCM1
GFM1
GFM1
GFM1
-E-
---A--
250
---
I
I
BASE CASE
TD=1 4 2 0 K
T=16 7 0 K
BASE CASE
TD=1 4 2 0 K
TD=1 6 7 0 K
200
0r
0
150
E100
50
0
0
20
40
60
80
100
120
TIME [min]
Figure 5.15
Accumulated Aerosol Releases for Different Concrete Decomposition Temperatures
227
LANTHANUM
e
I
---- E3-4---101 -U
I
RELEASE FRACTION
(La)
I
II
I
I
I
'
I
BASE CASE
TD=1 4 2 0 K
TI=1670 K
BASE CASE
TD=1420 K
67 O K
GFM1 TD=
PCM1
PCM1
PCM1
GFMI
GFM1
00
z
00
0
30
60
90
120
150
180
TIME [min]
Figure 5.16
Lanthanum Release Fractions for Different Concrete Decomposition Temperatures
228
900
800
700
600
z
500
z
400
300
200
100
0
0
30
60
90
120
150
180
TIME [min]
Figure 5.17
Downward Ablation Distances for Different Types of Concrete
229
3000
I
O
~s-8---
2800
-A-
---
PCMI
PCM1
PCM1
GFM1
GFM1
GFM1
I
I
I
I
BASE CASE
LIMESTONE
BASALTIC
BASE CASE
LIMESTONE
BASALTIC
2600
2400
m 2200
r
2000
0
1800
1600
1400
0
10
20
30
40
50
TIME [min]
Figure 5.18 Melt Temperature Histories for Different Types of Concrete
230
60
Because of the high gas content of the Limestone concrete, the gas generation
rate for this concrete is higher even though less concrete is being ablated (see
Fig. 5.19).
An important effect of the gas content of concrete is the depletion
time of zirconium in the metallic phase. As calculated by the GF model, 6 tons of
zirconium can be completely oxidized in 10 minutes for the Limestone/Common
Sand and Limestone cases. While in the Basaltic concrete case, due to its relatively
low gas content, it takes 2 hours to deplete the same amount of zirconium. In the
GF model, the surges of the gas generation rates at the moment of zirconium
depletion for various concretes can be seen in Fig. 5.19.
The aerosol gerneation rates and the total releases are shown in Figs. 5.20
and 5.21, respectively. For the Limestone concrete, not only because of its higher
gas content but also because of the higher quasi-steady oxidic temperature, the
aerosol generation rates remain at relatively high values compared to the other
concretes. Also, in general, the largest fission product release is calculated in the
Limestone concrete case while the smallest release is calculated with the Basaltic
concrete. In Fig. 5.22, the total tellurium releases for the three cases calculated
by the GF model span a range of a factor of three. A smaller difference among
these releases is resulted by the PC model.
5.4.4 Effect of Zirconium Metal
To investigate the sensitivity of the calculations of the ex-vessel source term
to the amount of metallic zirconium in the core debris initially, three cases with
different amounts of zirconium metal were analyzed. Based on the M1 compositions listed in Table 6, these cases were specified by having 6 tons (50%), 2.4 tons
(20%) and zero ton (0%) of unoxidized zirconium metal.
The case with 6 tons
zirconium metal was used as the base case.
The most direct impact of the amount of zirconium metal is the temperature
of the melt. The melt temperature can be affected by the heat of the chemical
reaction of the zirconium metal with the gas evolved from the decomposed con231
108
C.)
z0
r
ar
106
z
C-,
Cl)
0
10 5
0
30
60
90
120
150
TIME [min]
Figure 5.19
Gas Generation Rates for Different Types of Concrete
232
180
z 102
Er
101
Cf'2
100
10-1
10-2 L
0
30
60
90
120
150
180
TIME [min]
Figure 5.20
Aerosol Generation Rates for Different Types of Concrete
233
400
1
1
--e--
1
1
1-
PCM1 BASE CASE
-3-5-
350 ~--
-U-
1
-
PCM1 LIMESTONE
PCM1 BASALTIC
GFM1 BASE CASE
GFM1 LIMESTONE
GFM1 BASALTIC
300 -
250C12
0
200
150
D 100
50
0
0
20
40
60
80
100
120
TIME [min]
Figure 5.21
Accumulated Aerosol Releases for Different Types of Concrete
234
RELEASE FRACTION
TELLURIUM (Te)
102
Cn2
z
0
101
0
30
60
90
120
150
180
TIME [min]
Figure 5.22
Tellurium Release Fractions for Different Types of Concrete
235
crete. The more zirconium metal in the melt, the more chemical reaction heat is
generated, thus, the more is the concrete ablation (see Fig. 5.23). The temperature
histories of the oxidic melt with different amounts of zirconium are shown in
Fig. 5.24. Up to about 30 minutes, the smaller the amount of zirconium, the lower
is the temperature of the melt. After about 30 minutes, the oxidic temperatures
reverse; the larger the initial amount of zirconium the lower the oxidic temperature.
This inversion is caused by the enhanced concrete dilution of the melt. However,
the temperature differences among these cases at late times are small. At early
times, the temperature differences calculated by the GF model are larger than the
PC model.
In Fig. 5.25, it is shown that the gas generation rate calculated by the PC
model is mildly affected by the amount of zirconium. In the GF model, it can be
seen that the surge of the gas generation rate is affected by different amounts of
zirconium. The larger the amount of zirconium, the later is the surge of the gas
generation and the larger is the peak of the gas generation rate. After the gas
generation surge, the effect of the amount of zirconium is weakened.
The aerosol generation rates and the integral releases are shown in Figs. 5.26
and 5.27, respectively. It is seen that the larger the amount of zirconium, the more
areosol release is calculated. At 3 hours, the accumulated aerosol release of the
base case is higher than the 0% Zr case by a factor of 2 in the PC model and 2.4
in the GF model.
As calculated by both the PC and GF models, the larger the amount of
zirconium metal the greater is the fission product release. In Fig. 5.28, the base
case release of lanthanum is orders-of-magnitude greater than the 0% Zr case
despite similar melt temperatures and gas gerneration rates. The depressed release
of the lanthanum in the 0% Zr case is caused by the high oxygen potential of gases
evolved from the concrete. In the base case, reactions of H2 0 and CO2 from the
concrete proceed far enough to depress the oxygen potential of the gas to the point
236
800
700
r,
600
z
500
let
z
0
400
.
300
200
100
0
0
30
60
90
120
150
180
TIME [min]
Figure 5.23
Downward Ablation Distances for Different Amounts of Zirconium
237
3000
2800
2600
2400
m 2200
E-
2000
1800
1600
1400
0
10
20
30
40
50
60
TIME [min]
Figure 5.24
Melt Temperature Histories for Different Amounts of Zirconium
238
108
E-
z
106
E-
zU)
r0r
0
30
90
60
120
150
180
TIME [min]
Figure 5.25
Gas Generation Rates for Different Amounts of Zirconium
239
10 4
102
z
r
0
101
100
10-1
10-
2
[
0
30
90
60
120
150
180
TIME [min]
Figure 5.26
Aerosol Generation Rates for Different Amounts of Zirconium
240
300
250
II
Cn2
200
CI)
r0r 150
100
EQ
50
0
0
20
40
60
80
100
120
TIME [min]
Figure 5.27
Accumulated Aerosol Releases for Different Amounts of Zirconium
241
LANTHANUM
PCM1I
PCM1
PCM1
--4--GFM1
GFM 1
-AGFM1
----&
---
(La)
RELEASE FRACTION
BASE CASE
20% Zr
0% Zr
BASE CASE
20% Zr
0% Zr
S101
0
010
z
z
E--L
1 0-1
10-2
0
30
60
90
120
150
180
TIME [min]
Figure 5.28 Lanthanum Release Fractions for Different Amounts of Zirconium
242
where LaO(g) is stable in the gas phase. The lanthanum release of the 20% Zr
case falls between the other two cases. The orders-of-magnitude difference is also
found in the barium and strontium releases for-different amounts of zirconium
metal.
It is important to note that the significant effect of the zirconium metal on
the fission products release is based on the debris configuration assumed in the
VANESA model, in which the oxidic layer stays on top of the metallic layer. The
gas generation data calculated by the CORCON are the amounts of the gases
-
CO, C0 2 , H 2 and H 2 0 -
emerging from the core melt. These data are
accepted by the VANESA and then converted to the amounts of CO 2 and H20
liberated directly from the decomposed concrete. No matter at what location
(either sidewall or basemat of the concrete cavity) the gases are released, VANESA
assumes that all the released gases penetrate the metallic phase and react with
the zirconium metal before entering the oxidic layer. In the meantime, the major
forms of these fission products assumed in VANESA are BaO, SrO2 and La 2 0 3 .
These fission product oxides in the oxidic layer can only react with the gas which
is filtered by the bottom metallic layer. Therefore, the calculated fission product
releases are affected significantly by the amount of zirconium in the metallic phase.
However, this significant effect may disappear if the oxidic layer stays at the bottom
initially or float atop the metallic layer after the zirconium depletion as depicted
in the CORCON model.
The tellurium release, as shown in Fig. 5.29, is mildly affected by the amount
of zirconium metal. In the VANESA model, tellurium is assumed to be present as
Te(l) in the metal phase. Since tellurium is, by itself, quite volatile, its release in
the form of Te(g) is not affected by the oxygen potential differences in these cases.
5.4.5 Effect of Initial Debris Temperature
The initial temperature of core debris in the reactor cavity will be determined
by the temperature of the debris when it falls from the reactor vessel, i.e. it
243
I
TELLURIUM (Te)
RELEASE FRACTION
102
E-a
z
101
0
30
90
60
150
180
TIME [min]
Figure 5.29
Tellurium Release Fractions for Different Amounts of Zirconium
244
depends on the melt progression during the in-vessel phases of an accident. The
possible range of the initial temperature of a molten core, as discussed in QUEST,
is from 1807 K to 2600 K. In this study, five temperatures, 1807 K, 2000 K,
2320 K, 2400 K and 2600 K, were chosen to study the effect of the initial debris
temperature. Among these initial temperatures, 2320 K is the solidus temperature
of the oxidic melt used in the base case, and 1807 K is the liquidus temperature
of the metallic phase. Initial melt temperatures higher than 2320 K are therefore
categorized as high initial temperature while the others are referred to as low
initial temperature.
The axial ablation distances calculated by the PC and GF models are shown
in Figs. 5.30 and 5.31, respectively. It is interesting to note that the axial ablation
distances calculated by the GF model for all cases are almost equal at three hours.
In the PC model, the axial ablation distances of the high initial temperature cases
are higher than the low initial temperature cases. For the low initial temperature
cases, the axial ablation distances appear as an S-curve. The initial ablation is
limited by the formation of an oxidic crust at the bottom. Later, the ablation is
accelerated because of the layer flip. When the layer flip occurs, the metallic layer
with a temperature higher than its solidus point is brought down to the bottom.
The conduction controlled downward heat transfer is then replaced by the more
effective convective process, and the axial ablation rate is enhanced.
The melt temperature response can be found in Figs. 5.32 and 5.33. For the
Ti=1807 K and 2000 K cases, the debris temperature increases initially because
the heat generated in the melt is more than the heat being dissipated.
situation is reversed when the debris temperature reaches 2320 K.
This
In the PC
model, the temperature drops at 50 minutes and 40 minutes after the initiation of
the 1807 K and 2000 K cases, respectively. This temperature drop is caused
245
800
700
600
S500
z
~400
E0
300
200
100
0
0
30
60
90
120
150
180
TIME [min]
Figure 5.30
Downward Ablation Distances Predicted by the Periodic Contact
Model with Different Initial Debris Temperatures
246
800
700
600
z
5 0 0
C,)
E400
z
300
200
100
0
0
30
60
90
120
150
180
TIME [min]
Figure 5.31
Downward Ablation Distances Predicted by the Gas Film Model
with Different Initial Debris Temperatures
24 7
3000
2800
2600
2400
E"
2200
r
2000
1800
1600
1400
0
10
20
30
40
50
60
TIME [min]
Figure 5.32 Melt Temperature Histories Predicted by the Periodic Contact
Model with Different Initial Debris Temperatures
248
3000
2800
2600
2400
2200
0
2000
1800
1600
1400
0
10
20
30
40
50
60
TIME [min]
Figure 5.33
Melt Temperature Histories Predicted by the Gas Film Model with
Different Initial Debris Temperatures
249
by the layer flip. For the 2320 K case, the layer flip occurs at an earlier time, 30
minutes after the initiation. In the GF model, the oxidic temperature responses
are similar to those of the PC model for the low initial temperature cases.
The gas generation rates are shown in Figs. 5.34 and 5.35 for the PC and
GF models, respectively. Before 90 minutes, the differences of the gas generation
rates between the low initial temperature and the high initial temperature cases
are orders-of-magnitude. For the low initial temperature cases, again similar results are calculated by the PC and GF models, except sharper peaks of the gas
generation rates are calculated by the PC model. These peaks are caused by the
combination effects of the layer flip and coking phenomena.
The aerosol generation rates and accumulated releases are given in Figs. 5.36
through 5.39. It is seen that the initial aerosol release rate is significantly affected
by the initial debris temperature. After the first hour, the accumulated aerosol
releases for various initial temperature cases differ only by small amounts. For the
PC model, it is interesting to note that the accumulated aerosol releases of the
low initial temperature cases can become higher than the high initial temperature
cases after few hours interaction.
The initial debris temperature has significant effect on the calculation of the
fission product release. Some typical release rates resulted from different initial
melt temperatures are shown in Fig. 5.40. As a result, the total lanthanum releases
of the high initial temperature cases are higher than the low initial temperature
cases by orders-of-magnitude, see Figs. 5.41 and 5.42. The tellurium releases for
various cases are shown in Figs. 5.43 and 5.44. After the first hour, all cases with
low initial temperatures result in the same fission products release. It is even more
interesting that in the cases of low initial temperature, the fission product release
is not affected by the different heat transfer models.
250
108
z
106
0n
0
30
60
90
120
150
180
TIME [min]
Figure 5.34
Gas Generation Rates Predicted by the Periodic Contact Model
with Different Initial Debris Temperatures
251
I
108
L-1
z0
rI
z
10 5
0
30
60
90
120
150
180
TIME [min]
Figure 5.35
Gas Generation Rates Predicted by the Gas Film Model with Different Initial Debris Temperatures
252
E-
z0
z
10 1
ra
0
10 0
10 -1
10-
2
L
0
30
60
90
120
150
180
TIME [min]
Figure 5.36
Aerosol Generation Rates Predicted by the Periodic Contact Model
with Different Initial Debris Temperatures
253
E-
z 102
0
z
101
100
10-1
1 0-2
I
I
0
30
i
60
t
90
120
150
180
TIME [min]
Figure 5.37
Aerosol Generation Rates Predicted by the Gas Film Model with
Different Initial Debris Temperatures
254
300
250
r1
200
C
C
150
1)
100
50
0
0
20
40
60
80
100
120
TIME [min]
Figure 5.38 Accumulated Aerosol Releases Predicted by the Periodic Contact
Model with Different Initial Debris Temperatures
255
300
250
II
C'2
200
U)
150
EU)
100
50
0 k
0
20
40
60
80
100
120
TIME [min]
Figure 5.39
Accumulated Aerosol Releases Predicted by the Gas Film Model
with Different Initial Debris Temperatures
256
101
100
10-1
10-2
Em 10-3
10 -4
0
105
M
10-6
Cf)
10~7
10 -8
0
20
40
60
80
100
120
TIME [min]
Figure 5.40
Fission Products Release Rates Predicted by the Periodic Contact
Model with Different Initial Debris Temperatures
257
LANTHANUM
RELEASE FRACTION
(La)
102
101
0
100
10 -1
z*
10 -2
10 -3
0
30
60
90
120
150
180
TIME [min]
Figure 5.41
Lanthanum Release Fractions Predicted by the Periodic Contact
Model with Different Initial Debris Temperatures
258
LANTHANUM
(La) RELEASE FRACTION
102
101
U)
100
10 -1
Ez
10 -2
"a
10 -3
10 -4
10 -5
0
30
90
60
120
150
180
TIME [min]
Figure 5.42
Lanthanum Release Fractions Predicted by the Gas Film Model
with Different Initial Debris Temperatures
259
i
TELLURIUM (Te)
RELEASE FRACTION
102
101
z
100
me
10-1
10-2
0
30
60
90
120
150
180
TIME [min]
Figure 5.43 Tellurium Release Fractions Predicted by the Periodic Contact
Model with Different Initial Debris Temperatures
260
TELLURIUM (Te) RELEASE FRACTION
102
Q
101
z
100
0
10-1
10-
2
L
0
30
60
90
120
150
180
TIME [min]
Figure 5.44
Tellurium Release Fractions Predicted by the Gas Film Model with
Different Initial Debris Temperatures
261
5.4.6 Effect of Amount of Melts
Besides the base case (Ml), an enriched core oxide case (M2), and an enriched
metal steel case (M3), listed in Table 5.4, have been used to study the effect of
the melt amount. The amount of zirconium metal in the enriched core oxide was
increased correspondingly.
The effect of the amount of melt on the concrete erosion, on the melt temperature and on the gas generation is shown in Figs. 5.45, 5.46 and 5.47, respectively.
The oxidic temperature of the enriched core oxide case is the highest because of its
increased zirconium. However, the differences of the oxidic temperatures among
these cases are small.
In Fig. 5.48, the aerosol generation rate is increased by increasing the oxidic
material while it is only mildly affected by increasing the metallic material. The
total aerosol release of the enriched core oxide case is higher than the base case
by 60% and 40% for the GF and PC models, respectively (see Fig. 5.49). The
lanthanum release as shown in Fig. 5.50 is not affected significantly by the amount
of melts.
5.4.7 Effect of Ferrous Oxide
Here, different amounts of FeO have been used to study its impact on the
aerosol release. The enriched FeO case (M4), listed in Table 5.4, has been analyzed
and compared to the M1 cases. It should be noted that the amounts of FeO
are only small fractions of the oxidic layer in both cases. The most direct effect
resulting from the variation of the amount of FeO is the solidus temperature of the
oxidic melt. Increased FeO in the oxidic phase depresses the solidus temperature.
Considering the possible effect caused by the formation of the initial bottom crust,
three different initial debris temperatures (2000 K, 2320 K, and 2600 K) have been
applied to the enriched FeO case.
262
900
800
700
600
z
500
z
400
300
200
100
0
0
30
60
90
120
150
180
TIME [min]
Figure 5.45
Downward Ablation Distances for Different Amounts of Melt
263
3000
2800
2600
II
LJ
2400
2200
2000
0
1800
1600
1400
0
10
30
20
40
50
60
TIME [min]
Figure 5.46
Melt Temperature Histories for Different Amounts of Melt
264
10io1
GFM3 BASE CASE
L1j
4107.
z
10 6
z
105
10 4
0
I
30
I |
60
|
|
90
120
150
TIME [min]
Figure 5.47 Gas Generation Rates for Different Amounts of Melt
265
180
GFM3 BASE
r---
CASE
03
z102
m 10 0
-
10~1
I
10-2I
0
30
60
90
120
150
TIME [min]
Figure 5.48
Aerosol Generation Rates for Different Amounts of Melt
266
180
450
400
m
LJ
ci)
U)
350
300
0
ci) 250
0
200
150
Q
Q
100
50
0
0
20
40
60
80
100
120
TIME [min]
Figure 5.49
Accumulated Aerosol Releases for Different Amounts of Melt
267
I
LANTHANUM
(La) RELEASE FRACTION
10 1
Cl)
z
zi
W"
100
L
0
30
60
90
120
150
180
TIME [min]
Figure 5.50 Lanthanum Release Fractions for Different Amounts of Melt
268
In Table 5.6, the timing of the important events calculated by CORCON is
listed for both the M1 and M4 cases. It is seen that the zirconium depletion and
layer flip are delayed by a significant period of time when the initial bottom oxidic
crust is formed. When the initial temperature is 2600 K, there is no initial bottom
crust in either the M1 or M4 case; at 2000 K, the initial bottom crust is formed
in both cases; while at 2320 K, the initial bottom crust is formed only in the M1
case. This is because the oxidic solidus temperature of the M4 melt is lowered
by the increased FeO. Therefore, the most significant effect of the amount of
FeO on the behaviors of the corium/concrete interaction is found at the initial
temperature of 2320 K (see Figs. 5.51 and 5.52).
Similar trend is found in the calculation of the lanthanum release. As calculated by the PC model, the lanthanum release of the M4 case is about ten times
higher than the M1 case when the initial temperature is 2320 K, see Fig. 5.53.
While at the initial temperature of 2600 K, the M1 and M4 cases have exactly
the same lanthanum release. A similar trend is found for the calculations using
the GF model. The results of the lanthanum releases calculated by the GF model
are shown in Fig. 5.54.
It should be noted that the effect of the amount of FeO depends on the
formation of the initial bottom crust, which in turn, depends on the prediction
of the solidus temperature of the oxidic material. Therefore, if the initial debris
temperature is close to its solidus point, it is important to know precisely the
solidus point to determine the aerosol release.
5.4.8 Effect of Decay Heat
The entire corium/concrete interaction process is driven by decay heat generated in the pool. Because of the loss of some of the more volatile fission products
before the pool is formed, CORCON calculates the decay heat based on specified retention fraction of each element to account for partial loss of those volatile
species. Apparently, this will result in a lower decay heat generation compared
269
Table 5.6
Timing of Events of the Cases with Different Initial
Melt Temperatures and Different Amounts of FeO
Zr Depleted
Layer Flip
Initial Bottom
(min)
(min)
Crust
PCM1 T = 2600 K
2.0
2.0
No
PCM1 T = 2400 K
2.3
2.8
No
PCM1 T = 2320 K
36.0
39.0
Yes
PCM1 Ti = 2000 K
44.0
46.0
Yes
PCM1 Ti = 1807 K
51.0
52.5
Yes
PCM4 T =2600 K
2.1
1.8
No
PCM4 Ti = 2320 K
2.7
3.1
No
PCM4 T = 2000 K
25.3
26.5
Yes
GFM1 T = 2600 K
10.7
12.3
No
GFM1 T = 2400 K
11.0
14.0
No
GFM1 T = 2320 K
34.0
40.0
Yes
GFM1 T = 2000 K
46.0
51.0
Yes
GFM1 T = 1807 K
55.5
60.5
Yes
GFM4 T = 2600 K
10.3
11.3
No
GFM4 T = 2320 K
10.5
12.0
No
GFM4 T = 2000 K
33.0
37.5
Yes
Case
270
r
800
700
600
500
z
400
r
0
300
200
100
0
0
30
60
90
120
150
180
TIME [min]
Figure 5.51
Downward Ablation Distances Predicted by the Periodic Contact
Model with Different Amounts of FeO
271
300
---
250
Cl,
Cl,
'~U--
PCM4 T8=2320 K
PCM4 T =2000 K
200
Cl,
50
r
00
-~100
S50
0
20
40
60
80
100
120
TIME [min]
Figure 5.52 Accumulated Aerosol Releases Predicted by the Periodic Contact
Model with Different Amounts of FeO
272
LANTHANUM
(La) RELEASE FRACTION
101
100
10 -1
(*2
E-
10 -2
10 -3
10 -5
0
30
60
90
120
150
180
TIME [min]
Figure 5.53
Lanthanum Release Fractions Predicted by the Periodic Contact
Model with Different Amounts of FeO
273
I
LANTHANUM
(La) RELEASE FRACTION
102
101
C12
100
10-1
z
A
10-2
0
0
30
60
90
120
150
180
TIME [min]
Figure 5.54
Lanthanum Release Fractions Predicted by the Gas Film Model
with Different Amounts of FeO
274
to the ANS standard decay curve. In the base case analysis, it is found that the
decay powers calculated by CORCON are 11.4 and 9.86 MW at respectively 3 and
6 hours after scram. These values are about 50%.lower than those of the ANS
standard. To study the effect of decay heat, two cases with initial melt temperature
of 2600 K (HDHHT) and 1807 K (HDHLT) were analyzed by increasing the decay
heat to the ANS standard level.
In Fig. 5.55, the oxidic temperature of the increased decay heat case is seen to
be higher than the base case initially. After a period of time, due to the concrete
dilution of the melt, the situation is reversed.
About two hours later, a slow
increase in the oxidic temperature is calculated with increased decay power. In
general, there is no significant effect on the melt temperature due to the 50% decay
heat increase.
The accumulated aerosol releases of the HDHHT and HDHLT cases are shown
in Figs. 5.56 and 5.57, respectively. It is seen that the higher the decay heat, the
more aerosols are released; however, the difference is small. For the low initial
temperature, significant release of the ex-vessel aerosol in the case with increased
decay heat occurs earlier (about 20 minutes) than the base case.
The lanthanum release, as shown in Fig. 5.58, is not affected by the decay
power variation in the high initial temperature case.
While in the low initial
temperature case, the initial heatup rate of the oxide material is increased with
increasing decay heat. The difference in the lanthanum release can be as large as
an order-of-magnitude due to the difference in the oxide temperature response for
the first 30 minutes (see Fig. 5.59). After one hour, the lanthanum release of the
increased decay heat case is about 50% higher than the base case.
In general, the effect of the magnitude of the decay heat on the aerosol release
is less in the PC model than in the GF model. Because in the PC model the pool
temperature is primarily driven by the heat loss in the downward direction, the
275
3000
I
I
-E
PCM1
PCM1
-A8- GFM1
GFM1
-k-
-
2800
I i
1
I i1
i
1
BASE CASE
HDHHT
BASE CASE
HDHHT
2600
2400
w 2200
2000
1800
1600
1400
I
0
I
10
I
I
20
|I
30
|
40
50
60
TIME [min]
Figure 5.55
Melt Temperature Histories of the High Initial Debris Temperature
Cases with Different Amounts of Decay Heat
276
300
250
200
.W
150
E14 100
50
0
0
20
40
60
80
100
120
TIME [min]
Figure 5.56 Accumulated Aerosol Releases of the High Initial Debris Temperature Cases with Different Amounts of Decay Heat
277
I
300
I
-E-
PCM1 T-=1807 K
-A-
GFM1 T-=1807 K
GFM1 HDHLT
I
-4-- PCM1 HbHLT
-A--
250
C,)
200
C,)
0
150
r
A.
-
100
50
0
20
60
40
80
100
120
TIME [min]
Figure 5.57
Accumulated Aerosol Releases of the Low Initial Debris Temperature Cases with Different Amounts of Decay Heat
278
LANTHANUM
(La)
RELEASE FRACTION
I I I I
PCM I BASE CASE
-k- PCM 1 HDHHT
-AT
GFM1 BASE CASE
GFM1 HDHHT
101
0
r
z
On
A
A
10 0
0
30
60
90
120
150
180
TIME [min]
Figure 5.58
Lanthanum Release Fractions of the High Initial Debris Temperature Cases with Different Amounts of Decay Heat
279
LANTHANUM
(La) RELEASE FRACTION
100
E PCM1 T.=1807
---
PCM1 HDHLT
-A-
GFM1 T=1807
-i--GFM1 HDH LT
10-1-
10-2
r
Ar
z
10-5 L
0
30
60
90
120
150
180
TIME [min]
Figure 5.59
Lanthanum Release Fractions of the Low Initial Debris Temperature Cases with Different Amounts of Decay Heat
280
decay heat only contributes a relatively small fraction to the energy balance of the
corium pool.
5.4.9 Effect of Layer Configuration
As mentioned before, CORCON initially places the oxidic phase below the
metallic phase due to the density difference.
This configuration persists for a
period of time before the density of the oxide is reduced by ablated concrete
to a value less than that of the metallic material.
WECHSL, where the reference density of U0
2
This is not the case with
has been sufficiently reduced from
the handbook value to yield an initial oxide-over-metal configuration. However,
there are no direct measurements of the densities of the liquids involved. The
relative densities of metal and oxide, and hence the layer ordering should be viewed
as uncertain before conclusive results can be established. The most direct effect
based on different layer configuration will be the potential of forming an initial
bottom crust. The solidus temperature of the metallic phase is hundreds of degrees
Kelvin lower than that of the oxidic phase. At certain melt temperature (between
the solidus points of the metallic and oxidic materials), the bottom concrete attack
could be relatively violent if the metallic material stays at the bottom. On the
other hand, if the oxidic material settles below the metallic phase, the concrete
attack could be quite restricted due to the formation of a bottom crust.
In this test, the metallic phase is forced to be a bottom layer even though
the oxide, based on the CORCON calculation, is more dense than the metal initially. Two cases with initial melt temperatures of 2600 K (MSBHT) and 2320 K
(MSBLT) were analyzed by using the PC, GF and FC models. These cases will
be compared with the base case (initial metal-over-oxide) to illustrate the possible
effect resulting from different layer configurations.
The effects on the axial ablation distance and the melt temperature are illustrated in Figs. 5.60 and 5.61, respectively. As discussed in the previous chapter,
281
800
1
700
~*
-7~-*-
-e- PCM1 BASE CASE
PCM1 MSBHT
PCM1 Ti=2320 K
PCM1 MSBLT
'600
500
C14
400
z
r
300
200
100
0
0
30
60
90
120
150
180
TIME [min]
Figure 5.60
Downward Ablation Distances for Different Layer Configurations
282
3000
2800
2600
r-1
2400
2200
2000
1800
1600
1400 L
0
10
30
40
50
60
TIME [min]
Figure 5.61
Melt Temperature Histories for Different Layer Configurations
283
the downward heat flux is not sensitive to the properties of the pool materials in the
pre-freezing stages. It is seen that the axial ablation distance is mildly affected
by the layer configuration when the initial melt -temperature is high enough to
prevent any initial freezing in both the oxidic and metallic phases. In the case of
T = 2320 K, the layer configuration has significant effects on the calculations of
the concrete erosion and the melt temperature.
The aerosol generation rate and total release are shown in Figs. 5.62 and
5.63, respectively. Compared to the base case (in Figs. 5.36 through 5.39), it is
interesting to note that the aerosol release is not affected by the layer ordering if the
initial melt temperature is above the solidus points of both phases. While at the
low initial melt temperature (2320 K), there is an order-of-magnitude difference
on the aerosol generation rate. For the fission product release, a significant effect
is also found only in the low initial melt temperature case. The lanthanum release
at three hours for the PCM1 MSBLT case as shown in Fig. 5.64 is about eight
times higher than that of the PCM1 Ti = 2320 K case in Fig. 5.41.
It is interesting to note the differences in the fission product releases based
on different heat transfer models. The fission product release calculated by the
FC model will be about the same as the GF model under one of the following
two conditions. First, the initial melt temperature is high enough to stabilize an
initial gas film. Second, the initial melt temperature is low enough to form an
initial bottom crust. In the second condition, even the PC model gives about
the same fission product release as the GF model. In the MSBLT case, neither
an initial gas film nor an initial bottom crust is formed, the lanthanum release
calculated by the FC model, exactly the same as that of the PC model, is about
five times higher than the calculation of the GF model. In the MSBHT case,
an initial gas film but no initial bottom crust is formed, the lanthanum release
calculated by the FC model is about the same as that of the GF model, and it is
three times less than the calculation of the PC model.
284
ha
z
0
10 2
r
z
0
101
100
10 -1
10 -2
0
30
60
90
120
150
180
TIME [min]
Figure 5.62
Aerosol Generation Rates for Different Layer Configurations
285
300
-*-
PCM1 MSBHT
-A---
GFM1 MSBHT
-e-
PCM1 MSBLT
GFM1 MSBLT
FCMI MSBLT
----
250
-9-
FCM1 MSBHT
Cl)A
Ci)
<200
Cl)
150
w
r
100
50
0
0
20
40
60
80
100
120
TIME [min]
Figure 5.63
Accumulated Aerosol Releases for Different Layer Configurations
286
LANTHANUM
I
-------E)
-A-E-
I
PCMI
GFM1
FCM1
PCM1
GFM1
FCM1
(La) RELEASE FRACTION
I
T
MSBHT
MSBHT
MSBHT
MSBLT
MSBLT
MSBLT
101
100
z
ME
10-1
0
30
60
90
120
150
180
TIME [min]
Figure 5.64
Lanthanum Release Fractions for Different Layer Configurations
287
5.4.10 Effect of CORCON/MOD2 Version 2.01
The only changes included in the version 2.01 concern the treatment of partially solidified layers. As a result significant differences between the version 2.00
and 2.01 results may only occur either early if the core debris is partially solidified, or later when the molten core material is beginning to solidify. Three cases,
PCM1 Base Case, PCM1 0% Zr and PCM1 T, = 1807 K, have been analyzed
with version 2.01 to ascertain the impact of the code correction on the calculated
results.
Comparsions between the version 2.00 and 2.01 results are shown in Figs. 5.65
through 5.70. As expected, significant differences of these output parameters associated with the code correction are found only in the PCM1 Tj = 1807 K case.
However, some other parameters such as the radial ablation distance and released
gas composition are affected significantly in all the three cases.
As shown in
Figs. 5.71 and 5.72, the radial ablation distance and the accumulated CO 2 release
are reduced by a factor of two due to the code correction. These differences, however, do not have significant effects on the integral results of the ex-vessel source
terms.
5.4.11 Revised Periodic Contact Model
In the previous chapter, it was shown that the most significant difference in
the calculation of the downward heat flux between the PC and RPC models is
found in the case of the Limestone concrete, while the differences are relatively
small in the Limestone/Common Sand and Basaltic concrete cases. In order to see
the impact of the revised model on the ex-vessel source term calculation, several
cases (the Base Case, Limestone and Ti = 1807 K)
RPC model.
have been analyzed by the
In these cases, both the calculations of the RPC and PC models
were performed using the version 2.01.
The results of these calculations are shown in Figs. 5.73 through 5.78.
As
shown in these figures, although there are some effects associated with the modifi-
288
800
1
-e----A-
700
-*-
-E-
--
1
1
PCM1
PCM1
PCM1
PCM1
PCM1
PCM1
BASE CASE V2.00
BASE CASE V2.01
0% Zr V2.00
0% Zr V2.01
Ti=1807 K V2.00
Tj=1807 K V2.01
6 00
500
z
E-~400-
300
200
100
0
0
30
60
90
120
150
180
TIME [min]
Figure 5.65
Effect of the Corrected Version of CORCON/MOD2 on the Prediction of the Downward Ablation Distance
289
3000
2800
2600
2400
LJ
2200
E-
2000
1800
1600
1400
0
10
20
30
40
50
60
TIME [min]
Figure 5.66
Effect of the Corrected Version of CORCON/MOD2 on the Prediction of the Melt Temperature Histories
290
108
z
0
106
z
105
0
30
60
90
120
150
180
TIME [min]
Figure 5.67
Effect of the Corrected Version of CORCON/MOD2 on the Prediction of the Gas Generation Rate
291
10 5
E-
z
102
r
101
E-
100
10-1
10-2
0
30
60
90
120
150
180
TIME [min]
Figure 5.68
Effect of the Corrected Version of CORCON/MOD2 on the Prediction of the Aerosol Generation Rate
292
300
250
C,)
U)
200
0n
150
r
100
50
0
0
20
60
40
80
100
120
TIME [min]
Figure 5.69
Effect of the Corrected Version of CORCON/MOD2 on the Prediction of the Accumulated Aerosol Release
293
LANTHANUM
(La) RELEASE FRACTION
102
101
100
10 -1
A.
z
10-2
0
30
60
90
120
150
180
TIME [min]
Figure 5.70 Effect of the Corrected Version of CORCON/MOD2 on the Prediction of the Lanthanum Release Fraction
294
300
250
200
z
z0
150
100
50
0
0
30
60
90
120
150
180
TIME [min]
Figure 5.71
Effect of the Corrected Version of CORCON/MOD2 on the Prediction of the Radial Ablation Distance
295
I
20
15
C
10
r
Cf2
0e
5
0
0
30
60
90
120
150
180
TIME [min]
Figure 5.72
Effect of the Corrected Version of CORCON/MOD2 on the Prediction of the Released Gas
296
800
700
r
600
LJ
S 500
z
E- 400
0)
300
0
200
100
0
0
30
60
90
120
150
180
TIME [min]
Figure 5.73
Effect of the Revised Periodic Contact Model on the Prediction of
the Downward Ablation Distance
297
3000
2800
2600
2400
E-
II
2200
r
ar
Q
2000
1800
1600
1400
0
10
20
30
40
50
60
TIME [min]
Figure 5.74 Effect of the Revised Periodic Contact Model on the Prediction of
the Melt Temperature Histories
298
E-
CE
z0
106
z
10 4
0
30
60
90
120
150
180
TIME [min]
Figure 5.75
Effect of the Revised Periodic Contact Model on the Prediction of
the Gas Generation Rate
299
z
r
z
101
100
10-1
10-
2
0
30
60
90
120
150
180
TIME [min]
Figure 5.76
Effect of the Revised Periodic Contact Model on the Prediction of
the Aerosol Generation Rate
300
300
250
200
E-1
150
0
100
50
0 6
0
20
40
60
80
100
120
TIME [min]
Figure 5.77
Effect of the Revised Periodic Contact Model on the Prediction of
the Accumulated Aerosol Release
301
LANTHANUM
---
(La ) RELEASE
BSI
I
PCM I BASE CASE
RPCM I BASE CASE
PCM1 LUMESTONE
RPCM 1 LIME STONE
I
I
FRACTION
I
I
I
101
C,)
A
z
z
"=I
"a
100L
0
30
60
90fi
90
120
150
180
TIME [min]
Figure 5.78
Effect of the Revised Periodic Contact Model on the Prediction of
the Lanthanum Release Fraction
302
cation of the periodic contact model, the differences in the calculated aerosol and
fission product releases between the RPC and PC models are in general small.
In the Limestone concrete case, the RPC model predicts about 20% lower total
aerosol release and 25% higher total lanthanum release than the PC model.
5.4.12 Summary
Integral results of the PCM1 Base Case and GFM1 Base Case are shown in
Table 5.7. While those of the other cases are summarized in Tables 5.8 through
5.13. The values shown in these tables are relative to those of the M1 base case
stated in Table 5.7.
The lanthanum release shows the largest variation while the total aerosol
release has the smallest variation. The effect of the initial debris temperature on
the fission products release is the most significant, followed by the amount of Zr
present in the metallic phase.
The differences in the fission product releases between the PC model and GF
model are small when the melt is at a low temperature so that an initial bottom
crust can be formed. At high initial debris temperature, the PC model results in
higher fission product releases than the GF model. When an initial gas film or
a bottom crust can be formed, the results of the FC model are close to the GF
model. Otherwise, the FC model will follow exactly the PC model, and give higher
fission product releases than the GF model. The difference in the calculation of
the ex-vessel source term between the RPC and PC models is relatively small.
The layering order, i.e. whether the metallic or oxidic layer is in contact with
concrete, is unimportant if the initial melt temperature is high. However, at low
initial temperature, if the metallic layer contacts the concrete, a higher rate of
heat loss to concrete is possible, which leads to somewhat of an increased aerosol
generation when using the PC model. The effects of layering on the predicted
behavior by the GF model is very small.
303
Table 5.7
Integral Results of the Base Case at 3 Hours
after the Start of MCCI
GFM1
PCM1
Base Case
Base Case
CO
13.7
16.3
CO 2
2.25
3.19
H2
7.28
8.37
H20
1.36
2.19
Total
24.6
30.0
Axial
0.362
0.566
Radial
0.276
0.173
Sb
0.31
0.64
Te
37.7
63.9
Ba
19.8
25.0
Sr
35.4
50.3
La
2.08
6.36
Total Aerosol Release (kg)
207.8
184.4
Gas Release (104 moles)
Concrete Erosion Distance (m)
F.P. Release (% of Inventory)
304
Table 5.8
Radial and Axial Concrete Erosion Distances Relative to
PCM1 Base Case at 3 Hours after the Start of MCCI
Case
Radial Erosion
Axial Erosion
PCM1 Base Case
1.0
1.0
PCM1 Limestone
0.70
0.79
PCM1 Basaltic
1.00
1.28
PCM1 TD = 1420 K
1.02
1.14
PCM1 TD = 1670 K
0.91
0.70
PCM1 20% Zr
0.89
0.87
PCM1 0% Zr
0.82
0.77
PCM1 Tj = 2400 K
0.92
0.88
PCM1 Ti = 2320 K
1.00
0.66
PCM1 T = 2000 K
0.94
0.67
PCM1 T = 1807 K
0.87
0.70
PCM2 Base Case
1.58
1.30
PCM3 Base Case
1.20
1.12
PCM4 Base Case
1.06
1.02
PCM4 T = 2320 K
0.96
0.87
PCM4 T = 2000 K
0.90
0.70
PCM1 HDHHT
1.45
1.07
PCM1 HDHLT
1.12
0.78
PCM1 MSBHT
0.98
0.98
PCM1 MSBLT
0.82
0.83
FCM1 MSBHT
1.20
0.89
FCM1 MSBLT
0.82
0.83
GFM1 Base Case
1.60
0.64
FCM1 Base Case
1.49
0.72
305
Table 5.9
Radial and Axial Concrete Erosion Distances Relative to
GFM1 Base Case at 3 Hours after the Start of MCCI
Case
Radial Erosion
Axial Erosion
GFM1 Base Case
1.0
1.0
GFM1 Limestone
0.74
1.01
GFM1 Basaltic
0.95
1.42
GFM1 TD = 1420 K
1.13
1.12
GFM1 TD = 1670 K
0.94
0.80
GFM1 20% Zr
0.87
0.93
GFM1 0% Zr
0.78
0.88
GFM1 T = 2400 K
0.97
1.00
GFM1 T = 2320 K
0.71
0.97
GFM1 T = 2000 K
0.77
0.95
GFM1 T = 1807 K
0.86
0.94
GFM2 Base Case
1.56
1.24
GFM3 Base Case
1.23
0.98
GFM4 Base Case
1.08
0.98
GFM4 T = 2320 K
1.07
0.98
GFM4 T = 2000 K
0.89
0.92
GFM1 HDHHT
1.38
1.06
GFM1 HDHLT
1.07
1.00
GFM1 MSBHT
1.66
1.00
GFM1 MSBLT
0.76
1.00
306
Table 5.10
Accumulated Releases of Decomposition Gases Relative to
PCM1 Base Case at 3 Hours after the Start of MCCI
Case
CO
CO 2
H2
H20
Total
PCM1 Base Case
1.0
1.0
1.0
1.0
1.0
PCM1 Limestone
1.37
1.16
0.98
1.07
1.22
0.092
0.068
1.74
0.95
0.61
PCM1 TD = 1420 K
1.12
1.09
1.11
1.16
1.12
PCM1 TD = 1670 K
0.77
0.53
0.80
0.43
0.73
PCM1 20% Zr
0.85
0.93
0.80
1.11
0.87
PCM1 0% Zr
0.74
0.88
0.65
1.18
0.76
PCM1 Ti = 2400 K
0.91
0.74
0.93
0.71
0.88
PCM1 T = 2320 K
0.77
0.31
0.81
0.25
0.70
PCM1 T = 2000 K
0.77
0.32
0.81
0.26
0.70
PCM1 T = 1807 K
0.77
0.34
0.82
0.27
0.70
PCM2 Base Case
1.34
1.96
1.34
1.83
1.45
PCM3 Base Case
1.17
1.16
1.19
1.06
1.17
PCM4 Base Case
1.01
1.11
1.01
1.10
1.03
PCM4 T = 2320 K
0.88
0.81
0.90
0.74
0.87
PCM4 T = 2000 K
0.77
0.39
0.82
0.32
0.71
PCM1 HDHHT
1.07
1.55
1.05
1.53
1.15
PCM1 HDHLT
0.84
0.67
0.87
0.59
0.81
PCM1 MSBHT
0.98
1.02
0.98
0.99
0.98
PCM1 MSBLT
0.83
0.69
0.86
0.60
0.81
FCM1 MSBHT
0.91
1.03
0.93
0.95
0.94
FCM1 MSBLT
0.83
0.69
0.86
0.60
0.81
GFM1 Base Case
0.84
0.71
0.87
0.62
0.82
FCM1 Base Case
0.90
0.73
0.92
0.69
0.87
PCM1 Basaltic
307
Table 5.11
Accumulated Releases of Decomposition Gases Relative to
GFM1 Base Case at 3 Hours after the Start of MCCI
Case
CO
C02
H2
H2 0
Total
GFM1 Base Case
1.0
1.0
1.0
1.0
1.0
GFM1 Limestone
1.74
0.73
1.19
1.18
1.45
GFM1 Basaltic
0.092
0.075
1.69
1.18
0.62
GFM1 TD = 1420 K
1.09
1.27
1.08
1.34
1.12
GFM1 TD = 1670 K
0.82
0.68
0.83
0.62
0.80
GFM1 20% Zr
0.86
1.07
0.80
1.32
0.89
GFM1 0% Zr
0.77
1.09
0.67
1.60
0.81
GFM1 T = 2400 K
0.99
0.88
1.00
0.88
0.98
GFM1 T = 2320 K
0.94
0.54
0.96
0.52
0.89
GFM1 T = 2000 K
0.93
0.52
0.95
0.49
0.87
GFM1 T, = 1807 K
0.92
0.50
0.94
0.46
0.86
GFM2 Base Case
1.36
2.51
1.35
2.40
1.52
GFM3 Base Case
1.07
1.26
1.08
1.15
1.09
GFM4 Base Case
0.99
1.11
0.99
1.08
1.01
GFM4 Ti = 2320 K
0.80
1.15
0.78
1.19
0.85
GFM4 T = 2000 K
0.90
0.60
0.91
0.54
0.85
GFM1 HDHHT
1.06
1.79
1.04
1.78
1.16
GFM1 HDHLT
0.98
1.04
0.98
1.02
0.99
GFM1 MSBHT
0.89
1.47
0.90
1.32
0.97
GFM1 MSBLT
0.85
0.96
0.83
0.90
0.85
308
r
Table 5.12
Fission Products and Total Aerosol Releases Relative to
PCM1 Base Case at 3 Hours after the Start of MCCI
Sb
Te
Ba
Sr
La
Total
Aerosol
PCM1 Base Case
PCM1 Limestone
PCM1 Basaltic
1.0
1.18
0.55
1.0
1.15
0.61
1.0
1.08
0.71
1.0
1.0
1.0
1.04
0.74
0.86
0.62
1.40
0.74
PCM1 TD = 1420 K
1.14
1.09
0.99
0.99
1.00
0.91
PCM1 TD = 1670 K
PCM1 20% Zr
PCM1 0% Zr
PCM1 T = 2400 K
PCM1 T = 2320 K
0.70
0.79
1.09
1.11
0.99
0.95
0.92
0.30
0.047
0.87
0.31
0.0034
0.59
0.33
0.026
1.01
0.65
0.50
0.44
0.61
0.73
0.59
0.11
0.71
0.13
0.25
0.31
0.22
0.0049
1.04
PCM1 T = 2000 K
PCM1 T, = 1807 K
PCM2 Base Case
PCM3 Base Case
PCM4 Base Case
0.13
0.15
0.26
0.28
0.33
0.39
0.23
0.25
0.0034
0.0042
0.91
0.95
0.87
1.00
1.07
1.00
0.84
1.38
0.46
1.04
0.73
1.03
0.78
0.94
0.80
0.95
0.87
0.98
0.90
0.87-
PCM4 T = 2320 K
PCM4 T, = 2000 K
PCM1 HDHHT
PCM1 HDHLT
PCM1 MSBHT
0.35
0.53
0.62
0.45
0.042
0.66
0.12
0.24
0.28
0.19
0.0015
0.82
1.03
0.17
1.02
0.32
1.00
0.46
1.00
0.28
1.01
0.0058
1.04
0.97
0.99
0.99
1.00
1.00
1.01
0.98
PCM1 MSBLT
0.29
0.45
0.60
0.037
FCM1 MSBHT
0.51
0.64
0.82
0.44
0.72
0.31
0.74
1.02
FCM1 MSBLT
GFM1 Base Case
0.29
0.45
0.60
0.44
0.037
0.74
0.48
0.59
0.79
0.70
FCM1 Base Case
0.49
0.61
0.79
0.70
0.33
0.32
1.13
1.11
Case
309
Table 5.13
Fission Products and Total Aerosol Releases Relative to
GFM1 Base Case at 3 Hours after the Start of MCCI
Total
Case
Sb
Te
Ba
Sr
La
Aerosol
GFM1 Base Case
1.0
1.0
1.0
1.0
1.0
1.0
GFM1 Limestone
1.78
1.62
1.25
1.26
1.34
1.45
GFM1 Basaltic
0.57
0.59
0.65
0.69
0.71
0.79
GFM1 TD = 1420 K
1.04
1.04
1.01
1.02
1.08
1.02
GFM1 TD = 1670 K
0.90
0.91
0.96
0.94
0.83
1.07
GFM1 20% Zr
1.17
1.04
0.38
0.45
0.72
0.60
GFM1 0% Zr
1.31
1.06
0.029
0.0023
0.031
0.42
GFM1 T = 2400 K
0.59
0.75
0.76
0.62
0.11
0.92
GFM1 T = 2320 K
0.30
0.47
0.45
0.35
0.017
0.94
GFM1 T = 2000 K
0.31
0.49
0.50
0.36
0.013
0.92
GFM1 T, = 1807 K
0.33
0.51
0.55
0.38
0.015
0.91
GFM2 Base Case
1.16
1.25
1.17
1.13
1.16
1.62
GFM3 Base Case
0.41
0.68
0.81
0.84
0.99
1.03
GFM4 Base Case
1.04
1.02
1.05
0.96
1.01
1.01
GFM4 T = 2320 K
0.43
0.58
0.59
0.46
0.053
0.83
GFM4 T = 2000 K
0.29
0.45
0.43
0.31
0.0077
0.85
GFM1 HDHHT
1.04
1.05
1.02
1.02
1.02
1.07
GFM1 HDHLT
0.41
0.60
0.62
0.44
0.024
0.93
GFM1 MSBHT
0.84
0.87
0.90
0.91
0.88
1.02
GFM1 MSBLT
0.28
0.43
0.42
0.34
0.025
0.89
310
r
a.
5.5 Conclusions
Based on the results of the cases studied, the following conclusions are reached.
These conclusions reflect single-parameter variation around the specified base case.
1. In the base case, the initial aerosol generation rate calculated by the PC model
is five times higher than the GF model. In the low initial melt temperature
cases, the initial aerosol generation rates calculated by these two models are
close.
2. With the higher downward heat flux in the PC model, the melt temperature,
the gas generation rate, and the aerosol release rate decrease faster than the
GF model.
3. The fission product release calculated by the PC model is higher than the GF
model. The difference was as large as three times in the base case. While
in the oxide-over-metal configuration case, the release calculated by the PC
model was five times higher than the GF model.
4. When an initial stable gas film or an initial bottom crust calculated by the
FC model, the fission product release calculated by the FC model is about the
same as that of the GF model. Otherwise, the FC model results in a higher
(as much as five times) fission product release than the GF model.
5. The accumulated aerosol release (predominately non-radioactive materials)
calculated by the PC model is higher than the GF model initially. After the
transient stage, the GF model predicts a slightly higher aerosol release.
6. The difference in the calculation of the ex-vessel source term between the
RPC and PC models is relatively small.
7. Based on the VANESA calculation, the amount of zirconium metal has significant effect on the fission products release.
8. The concrete decomposition temperature is important in determining the
aerosol release rate during the quasi-steady state, because higher concrete
decomposition temperature gives higher oxidic temperature which results in
a higher aerosol release rate.
311
9. The aerosol release calculated with the Limestone concrete is higher than
the Limestone/Common Sand concrete, which is higher than the Basaltic
concrete.
10. Increasing the decay heat, the aerosol release is increased by a small amount
in the case of high initial melt temperature. While for low initial melt temperatures, there is a striking effect due to the decay heat variation on the
aerosol release in the initial period of time.
11. Significant effect of the layer ordering on the fission product release is found
only when the initial melt temperature is between the solidus points of the
metallic and oxidic materials.
12. The aerosol release is affected significantly by the formation of an initial bottom crust, therefore, the solidus temperature of the melt is important in
determining the aerosol release.
13. In general, a relatively low aerosol generation rate is calculated by using the
CORCON/MIT compared to the CORCON/MOD1. At quasi-steady state,
except for the Limestone concrete case, the aerosol generation rates calculated
by the CORCON/MIT drop below 1 g/s, while the CORCON/MOD1 used
in QUEST gives about an order-of-magnitude higher aerosol generation rate.
14. At a high initial melt temperature, the transient behavior of the fission products release calculated by using the CORCON/MIT span a much shorter
time than the CORCON/MOD1. The lanthanum release calculated by the
PC model can be saturated in 3 minutes.
15. Among these studied cases, the calculations from the VANESA model of the
ex-vessel source term can be rank-ordered in terms of decreasing sensitivity
to the parameters as follows: (1)Initial melt temperature; (2)amount of zirconium in melt; (3)downward heat transfer model; (4)amount of steel in melt;
(5)amount of core oxide involved; (6)amount of ferrous oxide in melt; (7)decay
heat and (8)layer ordering.
312
CHAPTER6
SUMMARY AND CONCLUSIONS
6.1 Summary of This Work
The radioactive aerosol formation and release in reactor severe accidents receive considerable attention in nuclear plant safety assessments as this is the most
significant potential hazard to the public. Progress in the understanding of the
aerosol phenomena has been made in recent years. Based on the new developments,
the calculated amounts of volatile radioactive material that could be released to
the environment is substantially smaller than was reported in the Reactor Safety
Study. This finding resulted from better understanding of containment integrity,
natural retention potential of reactor systems and chemistry of cesium iodine (CsI).
However, one mechanism that might, for some sequences, increase the radionuclide releases above those calculated in the Reactor Safety Study is the release
of nonvolatile radionuclides in the core-concrete interaction. The magnitude of
the contribution from the nonvolatile radionuclides which could be available for
long-term ex-vessel release is still open to question, primarily because of the modeling uncertainty of the MCCI. While the fundamental concepts of current models
have been generally accepted, heat transfer modeling is not fully developed and
validation is incomplete. In addition, calculations of the ex-vessel source term are
also known to depend strongly on details of core melt progression for which many
uncertainties still exist. All these have motivated this study focusing on modeling
of the phenomena involved in thermal hydraulics of the core/concrete interactions.
6.1.1 Experimental Observations
Simulant experiments were designed with air injection and cooling capabilities
of both a single-layer and a multi-layer liquid pool to investigate several important
physical processes of the core/concrete interaction, such as freezing, liquid/liquid
313
interfacial heat transfer, layer mixing, and droplet entrainment. Water and cyclohexane, which are immiscible, were used in the experiments to simulate the oxidic
and metallic materials of the core/concrete interaction.
The freezing phenomena experiments conducted in this study are of scoping
nature. It was observed that a bottom crust could be formed across the bubble
agitated horizontal liquid/solid interface, with gas velocities up to 126 mm/s.
This observation validates the freezing model used in the current MCCI integral
analysis codes. However, the liquid/liquid interface crust also assumed in the
analysis codes did not form in the simulant experiments. The stability of a top
crust is also called into question by the observations of this experiment. Therefore,
if the oxidic materials form a top layer, the freezing of the oxidic layer involved
in the MCCI could be in a slurry form rather than a crusting boundary. The
heat transfer from the oxidic layer to either the metallic layer or the containment
atmosphere at post-freezing stages will increase if there is no boundary crust. In
addition, the supercooling phenomenon observed in the experiments is not usually
accounted for in predicting the timing of freezing.
Several correlations have been developed based on the surface renewal concepts to calculate the heat transfer rate between the bubble agitated immiscible
oxidic and metallic layers. Significant differences of the interfacial heat transfer
among these models' predictions were found. In the simulant experiment, the
interfacial heat transfer between the water and cyclohexane layers was measured
under various superficial gas velocities. Comparisons of the data with the existing
models were made. The modified Szekely model used in the WECHSL code, and
the model developed by Lee and Kazimi agree well with the experimental data.
The Greene model incorporated in the CORCON/MOD2 seems to overpredict
the experimental results, and the modified Konsetov model used in the CORCON/MOD1 underestimates the experimental data by an order-of-magnitude.
314
r
In the layer mixing test, it was found that two immiscible liquids with density
ratio of 0.78 were entirely homogenized under a modest superficial gas velocity of
50 mm/s. The transition patterns of the mixing phenomena with different gas
velocities were observed as well. In the core/concrete interaction, the superficial
gas velocity could reach as high as 1 m/s when the melt temperature is high,
and the density ratio of the metallic and oxidic materials is about 0.8 initially
when the oxidic layer consists of heavy core oxide only. From the observation of
the simulant experiment, it is highly likely that the metallic and oxidic materials
could be mixed into a single layer during some periods of the MCCI.
Liquid droplets entrained by the flowing gas were quantified. The median and
maximum sizes of the water droplets entrained by a gas flow of j, = 8.0 mm/s
were found to be 2.0 and 20.0 pIm, respectively. The amount of entrainment was
found in good agreement with the Kataoka and Ishii model.
6.1.2 Development and Validation of Heat Transfer Models
The behavior of a pool of molten core materials in a concrete cavity is governed by an energy balance. The decay heat and chemical reaction heat generated
in the pool may be lost through its top surface to the containment atmosphere
or containment structure or to the surrounding concrete. The partition of energy
between concrete and the top surface is determined by the various thermal resistances from the pool of molten core materials to the surroundings. Among various
phenomenological heat transfer models, the one having the most direct impact on
the core-concrete interactions process is that describing the heat transfer across
the melt/concrete interface. The extent of concrete ablation, the melt temperature
response, the amount of decomposition gas release and therefore the amounts of
chemical heat and aerosol releases all directly depend on the amount of heat that
can be transferred across the melt/concrete interface.
In the first generation of MCCI integral analysis codes, a gas film model was
used for the downward heat transfer. This model assumed that the downward heat
315
transfer of the corium pool is governed by a stable gas film across the horizontal
corium/concrete interface. The stable gas film was inferred from water/dry ice
simulant tests. However, in air injection simulant .experiments, no gas film was
observed at the water/porous plate interface with superficial gas velocity up to
130 mm/s.
A preliminary periodic contact model was proposed to govern the
heat transfer process when a gas film cannot be sustained at the interface. This
model considered the heat transfer mechanism as a transient heat conduction
process of a periodic direct contact between the hot pool and the relatively cold
concrete surface. In this study, a revised periodic contact model was developed
based on more complete theoretical consideration to overcome a deficiency found
in its original derivation.
Both the gas film and periodic contact models are conceptualized based on assumed physical phenomena which cannot be directly observed in the real material
experiments. The actual heat transfer mode may depend on the melt temperature
and is not well understood. A film collapse model was proposed assuming that
the downward heat transfer may follow a combination of the gas film and periodic
contact models. If the melt initial temperature is high enough, a stable film may
exist, and the heat flux presumably follows the gas film model until a minimum
stable film limit is reached. When the film collapses, the heat transfer mode undergoes a transition to the periodic contact model. If the melt initial temperature
is too low to generate a stable film, the heat transfer will be governed by the
periodic contact model all the time until the solidus temperature is reached. The
transition criteria of a stable gas film used in the film collapse model are based on
hydrodynamic considerations. The film establishment criterion was related to the
Kutateladze's flooding limit, and the film collapse limit was related to Berenson's
minimum gas flux to stabilize the film. For the case of core/concrete interaction,
those limits differ by two orders-of-magnitude.
In the development of the film
collapse model, multiplication factors were applied to both limits in order to get
the best fit of the real material experimental data.
316
As the pool cools down, formation of a bottom crust provides an additional
thermal resistance to the downward heat flow path.
This conductive thermal
resistance will limit the amount of heat loss to concrete and reduce the amount
of gas gerneration.
The gas film is destabilized at a sufficiently low superficial
gas velocity condition, which may occur earlier than the crust formation.
The
applicability of the periodic contact model is limited to the early stages of the
MCCI before any freezing occurs. Therefore, 'neither the gas film nor the periodic
contact model can be used to describe the downward heat transfer during the
post-freezing stages of MCCI. In this study, a post-freezing model was developed
based on an assumption that thermal resistance between the pool boundary and
the concrete surface, i.e. resistance across the rising fluid, is continuous on the
basis of the superficial gas velocity that can be achieved after the formation of a
solidified bottom crust.
An integral analysis computer code CORCON/MIT was developed by incorporating the proposed downward heat transfer models into CORCON/MOD2 to
analyze the integral behavior of the core/concrete interaction. Sensitivity study
was performed to scope out the important parameters, such as the melt temperature and compositions, and the concrete types, in the calculations of the downward
heat transfer. It was found that the downward heat flux of the core/concrete interaction predicted by the periodic contact model is about an order-of-magnitude
higher than the gas film model if the corium temperature is higher than its solidus
point. As predicted by all models, the downward heat fluxes drop dramatically
when the melt temperature drops across its solidus point. The calculated downward heat fluxes for the post-freezing stage is rarely affected by different heat
transfer models which are used to predict the relatively small thermal resistance
at the melt/concrete interface compared to the dominant thermal resistance across
the solidified melt. It was also found that the downward heat flux is not affected
significantly by the compositions of the melt which may contact the horizontal
concrete surface when the debris temperature is high. The possible implication
317
of this finding is that the downward heat transfer between the corium and concrete may not be affected significantly by the mixing of the metallic and oxidic
materials. At low debris temperature, allowing formation of a bottom crust, a
metallic pool with higher thermal conductivity has a higher downward heat transfer than an oxidic one. The differences in the calculated downward heat fluxes
among different types of concrete can be large.
The various models were validated by comparison to the German BETA experimental results. The BETA facility is a large scale (380 mm diameter concrete
crucible contains up to 350 kg metallic and 150 kg oxidic melt) high power inductive heating (up to 1900 kW) experiment. The major finding of the BETA tests
was that the dominant downward erosion indicates a very effective heat transfer
mechanism at the bottom of the concrete crucible, which is different from the
sideward heat transfer mechanism. The proposed downward heat transfer model
is capable of producing the downward erosion results of the BETA experiments
with a mean error of 5% (overestimation) and a standard deviation of 27%. Less
accuracy is found by the other existing models in the calculation of the erosion
data of the BETA tests. The gas film model, developed earlier and commonly
used in severe accident analysis, significantly underestimates the downward heat
transfer (with mean error of -53%, and standard deviation of 56%).
The large scale, real material experiments (SWISS and TURC) conducted at
the Sandia National Laboratory were also analyzed. It is found that the proposed
model is able to produce a fairly good agreement in the axial erosion for sustained
heating SWISS tests. While in the transient TURC tests, the proposed model
cannot predict the axial erosion data very well in most cases.
6.1.3 Impact of Heat Transfer Models
In view of the large differences among the downward heat fluxes predicted
by the various models, a sensitivity study based on CORCON/MIT-VANESA
calculation was performed to investigate the effect of the downward heat transfer
318
model on the concrete erosion, gas generation, melt temperature, and ex-vessel
aerosol release in a real reactor case.
It was found that the rate and history of the ex-vessel aerosol release is significantly affected by the downward heat transfer model used in the calculation. In
the base case analyzed in this study, the initial aerosol generation rate predicted
by the periodic contact model is five times higher than the gas film model because
the initial downward heat flux and gas generation rate calculated by the periodic
contact model are higher.
However, the melt temperature, the gas generation
rate, and the aerosol release rate decrease faster for the periodic contact model,
therefore, the accumulated aerosol (including radionuclides and non-radioactive
materials) release after three hours interaction calculated by the periodic contact
model is not significantly different from the gas film model. Nevertheless, the accumulated fission products release (mostly released at elevated melt temperature)
calculated by the periodic contact model was three times higher than the gas film
model if the initial melt temperature is high.
At low initial melt temperature,
when a bottom crust is initially formed, the various heat transfer models do not
lead to significant differences in the fission product release. The results of the film
collapse model depend on the initial condition of the core/concrete interaction. If
neither a bottom crust nor a gas film exists initially, the fission products release
predicted by the film collapse model can be as much as five times higher than the
gas film model.
6.1.4 Sensitivity Study on Ex-Vessel Aerosol Release
A parametric study on significant variables, such as (1)initial melt temperature; (2)concrete properties; (3)amount of unoxidized zirconium; (4)amount of
melt; (5)decay heat; and (6)layering potential of melt constituents, was performed
to identify the important source of the uncertainties in calculation of the ex-vessel
aerosol release.
319
It is found that the initial melt temperature is extremely important to the calculation of the ex-vessel aerosol release. In the cases studied, the total lanthanum
release of the high initial melt temperature case can be an order-of-magnitude
higher than the low initial melt temperature case.
The amount of unoxidized zirconium in the melt has little effect on the melt
temperature, however, it has significant effect on the calculation of the fission
product release. The more unoxidized zirconium in the melt, the lower the oxygen
potential of the reaction gases will be, which leads to a more stable vaporized
fission product in the gas phase, and therefore a higher fission product release
during core/concrete interaction.
The concrete properties, such as concrete decomposition temperature and gas
content, have certain effects in determining the ex-vessel release. The higher the
concrete decomposition temperature is, the higher the oxidic temperature will be
during the quasi-steady state, which will result in a higher aerosol release. Since
the ex-vessel aerosol release increase with the increasing of the gas generation
rate, the concrete with higher gas content will give higher aerosol release. In the
calculations of different concretes, it is found that the aerosol release predicted
with the Limestone concrete is higher than the Limestone/Common Sand concrete,
which is higher than the Basaltic concrete.
The layer ordering effect on the aerosol release is found to be important only
if the initial melt temperature is between the solidus points of the metallic (~1700
K) and oxidic (~2300 K) materials. The solidus temperature of the melt and the
amount of ferrous oxide in the melt, which can change the potential and timing of
forming a bottom crust, also affect the calculation of the aerosol release.
320
REFERENCES
[Al] American Nuclear Society, "Report of the Special Committee on Source
Terms," September 1984.
[A2] American Physics Society, "Report to the American Physical Society of the
Study Group on Radionuclide Release from Severe Accidents at Nuclear Power
Plants," Reviews of Morden Physics, 57, No.3, Part II, July 1985.
[A3] H. Alsmeyer et al., "A Model Describing the Interaction of A Core Melt with
Concrete," NUREG/TR-0039, September 1978.
[A4] S.Z. Ahmed and V.K. Dhir, "On the Simulation of Penetration of Solidified
Core Material into Concrete," Proceedings of the ANS/ENS International
Meeting on Light Water Reactor Severe Accident Evaluation, Cambridge,
MA, August 28 - September 1, 1983.
[A5] H. Alsmeyer, M. Riemann, and J.P. Hosemann, "Preliminary Results of the
KfK MCCI Experimental BETA Facility," Eleventh NRC Water Reactor
Safety Information Meeting, Gaithersburg, MD, October 1984.
[A6] H. Alsmeyer et al., "BETA Experimental Results on Melt/Concrete Interaction: Limestone Concrete Behavior," OECD/CSNI Specialist Meeting on
Core Debris/Concrete Interactions, Palo Alto, CA, September 3 - 5, 1986.
[A7] H. Alsmeyer et al., "BETA Experimental Results on Melt/Concrete Interaction: Silicate Concrete Behavior," OECD/CSNI Specialist Meeting on Core
Debris/Concrete Interactions, Palo Alto, CA, September 3 - 5, 1986.
[B1] L. Baker et al., "Interaction of LMFBR Core Debris with Concrete,"
Proceedings of the International Meeting on Fast Reactor Safety and
Related Physics, Chicago, Ill, October 5 - 8, 1976.
[B2] R.E. Blose et al., "SWISS: Sustained Metallic Melt-Concrete Interactions
with an Overlying Water Pool - Experiments and Analysis," Sandia National
Laboratory, SAND85-1546, to be published.
[B3] A.S. Benjamin, "Core-Concrete Pool Dynamics and Interfacial Heat Transfer," Proceedings of the ANS/ASME NRC Topical Meeting on Nuclear
Reactor Thermal Hydraulics, NUREG/CP-0014, Vol.2, Saratoga Spring,
NY, October 5 - 8, 1980.
[B4] F.G. Blottner, "Hydrodynamics and Heat Transfer Characteristic of Liquid
Pool with Bubble Agitation," Sandia National Laboratory, SAND79-1132,
NUREG/CR-0944, November 1979.
[B5] A. Butland et al., "The Vaporization of Chemical Species and the Production
of Aerosols During a Core Debris/Concrete Interactions," OECD/CSNI
Specialist Meeting on Core Debris/Concrete Interactions, Palo Alto, CA,
September 3 - 5, 1986.
321
[B6] D.R. Bradley and A.J. Suo-Anttila, "Improvements in CORCON Heat Transfer Modeling," OECD/CSNI Specialist Meeting on Core Debris/Concrete
Interactions, Palo Alto, CA, September 3 - 5, 1986.
[B7] H. Bunz, M. Kayro, and W. Schock, "NAUA-Mod4: A Code for Calculating Aerosol Behavior in LVR Core Melt Accidents," Kernforschungszentrum
Karlsruhe, FRG, KfK-3554, August 1983.
[B8] K.D. Bergeron et al., "User's Manual for CONTAIN 1.0, A Computer Code
for Severe Nuclear Reactor Accident Containment Analysis," Sandia National
Laboratory, NUREG/CR-4058, SAND84-1204, May 1985.
[B9] P.J. Berenson, "Film Boiling Heat Transfer from A Horizontal Surface,"
Journal of Heat Transfer, 38, pp. 3 5 1 - 358, 1961.
[Cl] T.Y. Chu, "Radiant Heat Evaluation of Concrete - A Study of the Erosion
of Concrete Due to Surface Heating," Sandia National Laboratory, SAND770922, January 1978.
[C2] E.R. Copus and T.M. Kerley, "Interaction of Hot Solid Core Debris with
Concrete," Meeting to Review NRC Core Concrete Experimental and
Analytical Programs, Silver Spring, MD, May 14 - 16, 1985.
[C3] R.K. Cole, Jr., D.P. Kelly, and M.A. Ellis, "CORCON/MOD2: A Computer
Program for Analysis of Molten-Core Concrete Interactions," Sandia National
Laboratory, NUREG/CR-3920, SAND84-1246, August 1984.
[C4] R.K. Cole, Jr., "A Crust Formation and Refreezing Model for Molten Fuel/
Concrete Interaction Codes," Proceedings of the ANS /ENS International
Meeting on Light Water Reactor Severe Accident Evaluation, Cambridge,
MA, August 28 - September 1, 1983.
[C5] P. Clough, S. Ramsdale, and P. Smith, "Aerosol Decontamination Factor in
Pools Overlying Molten Core-Concrete: Code Modeling," OECD/CSNI
Specialist Meeting on Core Debris/Concrete Interactions, Palo Alto, CA,
September 3 - 5, 1986.
[D1] V.K. Dhir, J.N. Castle, and I. Catton, "Role of Taylor Instability Sublimation
of Horizontal Slab of Dry Ice," Journal of Heat Transfer, 99, pp.411 - 418,
1977.
[D2] R.S. Denning et al., "Radionuclide Release Calculations for Selected Accident Scenarios," Battelle Columbus Laboratory, NUREG/CR-4624, BMI2139, July 1986.
[Fl] D.K. Felde, H.S. Kim, and S.I. Abdel-Khalik, "Convective Heat Transfer Correlation for Molten Core Debris Pools Growing in Concrete," Nucl. Eng. Des.,
58, pp.65 - 74, 1980.
[F2] Fauske and Associates, Inc., "MAAP: Modular Accident Analysis Program
User's Manual," Technical Report 16.2-3, IDCOR, August 1983.
322
[F3] Fauske and Associates, Inc., "MAAP Uncertainty Analysis," FAI/85-19
(IDCOR Subtask 23.4), April 1985.
[F4]
E.L. Fuller and Z.T. Mendoza, "Structured Sensitivity Analyses Using the
MAAP 2.0 Computer Program," Proceedings of the ANS/ENS International
Topical Meeting on Thermal Reactor Safety, San Diego, CA, February 1986.
[G1] J.A. Gieseke et al., "Radionuclide Release Under Specific LWR Accident Conditions," Battelle Columbus Laboratory, BMI-2104, July 1984.
[G2] G.A. Greene et al., "Corium-Concrete Interaction in the Mark I Containment
Drywell and Local Liner Failure," International Symposium on Source
Term Evaluation for Accident Conditions, IAEA-SM-281/36, Columbus,
OH, October 1985.
[G3] F. Gonzalez and M. Corradini, "Experimental Study of Pool Entrainment and
Mixing Between Two Immiscible Liquids with Gas Injection," OECD /CSNI
Specialist Meeting on Core Debris/Concrete Interactions, Palo Alto, CA,
September 3 - 5, 1986.
[G4] J.E. Gronager et al., "TURC1: Large Scale Metallic Melt-Concrete Interaction Experiments and Analysis," Sandia National Laboratory, SAND85-0707,
NUREG/CR-4420, 1985.
[G5] J.E. Gronager, A.J. Suo-Anttila, and J.E. Brockmann, "TURC2 and TURC3:
Large Scale U0 2 /ZrO 2 /Zr Melt-Concrete Interaction Experiments and Analysis," Sandia National Laboratory, SAND86-0318, NUREG/CR-4521, 1986.
[G6] R. Grief, "Heat Transfer with Gas Injection at the Surface," Intl. J. Heat
Mass Transfer, 8, pp.1253, 1965.
[G7] G.A. Greene et al., "Heat Transfer between Immiscible Liquid Enhanced by
Gas Bubbling," International Meeting on Thermal Nuclear Reactor Safety,
Chicago, Ill, August 29 - September 2, 1982.
[G8] T. Ginsberg and G.A. Greene, "BNL Program in Support of LWR Degraded
Core Accident Analysis," Proceedings of the U.S. NRC Tenth Water Reactor
Safety Research Information Meeting, NUREG/CP-0041, Gaithersburg,
MD, October 12 - 15, 1982.
[G9] T. Ginsberg, "Aerosol Generation from Sparging of Molten Pools of Corium
by Gases Released from Core-Concrete Interactions," Proceedings of the
ANS/ENS International Meeting on Light Water Reactor Severe Accident
Evaluation, Cambridge, MA, August 28 - September 1, 1983.
[G10] J.A. Gieseke et al., "Source Term Code Pakage: A User's Guide (Mod 1),"
Battelle Columbus Laboratory, NUREG/CR-4587, BMI-2138, July 1986.
[G11] G.A. Greene and Y. Sanborn, "Independent Analysis of Selected Core-Concrete Interaction and Fission Product Release Experiments with CORCON
-MOD2 and VANESA," Proceedings of the ANS/ENS International Topical
Meeting on Thermal Reactor Safety, San Diego, CA, Febuary 1986.
323
[G12] J. Gronager and A. Suo-Anttilla, "Sustained Water Interacting with Stainless
Steel (SWISS) Test Series Review," Meeting to Review NRC Core Concrete
Experimental and Analytical Programs, Silver Spring, MD, May 1985.
[HI] R.E. Henry, "A Model for Core--Concrete Interaction," Proceedings of the
ANS/ENS International Meeting on Light Water Reactor Severe Accident
Evaluation, Cambridge, MA, August 28 - September 1, 1983.
[I1] Technology for Energy Corp., " IDCOR Technical Summary Report: Nuiclear Power Plant Response to Severe Accidents," Atomic Industrial Forum,
November 1984.
[12] H.J. Ivey, "Relationships between Bubble Frequency, Departure Diameter and
Rise Velocity in Nucleate Boiling," Intl. J. Heat Mass Transfer, .0, pp.1023
- 1040, 1967.
[Ki] T.S. Kress et al., "Review of the Status of Validation of the Computer Codes
Used in the Severe Accident Source Term Reassessment Study (BMI-2104),"
Oak Ridge National Laboratory, ORNL/TM-8842, April 1985.
[K2] M. Khatib-Rahbar et al., "Independent Verification of Radionuclide Release
Calculations for Selected Accident Scenarios," Brookhaven National Laboratory, NUREG/CR-4629, BNL-NUREG-51998, July 1986.
[K3] S.S. Kutateladze and I. Malenkov, "Boiling and Bubbling Heat Transfer
Under the Conditions of Free and Forced Convection," Sixth International
Heat Transfer Conference, Toronto, Canada, 1978.
[K4] K.K. Konsetov, "Heat Transfer during Bubbling of Gas through Liquid,"
Intl. J. Heat Mass Transfer, 9, pp. 1 1 0 3 - 1108, 1966.
[K5] L.S. Kao, M. Lee, and M.S. Kazimi, "Assessment of Heat Transfer Models for
Corium-Concrete Interaction," Proceedings of the Third International
Topical Meeting on Reactor Thermal Hydraulics, Newport, RI, October 15
- 18, 1985.
[K6] L.S. Kao and M.S. Kazimi, "Application of CORCON/MIT to Analysis of
the BETA Experiments," Presentation at BETA Workshop, Karlsruhe, FRG,
June 12, 1986.
[K7] M.R. Kuhlman, D.J. Lemicke, and R.O. Meyer, "CORSOR User's Manual,"
Battelle Columbus Laboratory, NUREG/CR-4173, BMI-2122, March 1985.
[K8] L.S. Kao and M.S. Kazimi, "Containment Pressure Response to A Meltdown
Condition of the Light Water Reactor," Proceedings: The Sixth Information
Exchange Meeting on Debris Coolability, EPRI NP-4455, UCLA, Los Angeles, CA, November 7 - 9 1984.
[K9] I. Kataoka and M. Ishii, "Mechanistic Modeling and Correlations for Pool Entrainment Phenomenon," Argonne National Laboratory, NUREG/CR-3304,
ANL-83-37, April 1983.
324
[K10] L.S. Kao and M.S. Kazimi, "A Heat Transfer Model for Core/Concrete Interaction," AIChE/ANS 24th National Heat Transfer Conference, Pittsburgh,
PA, August 1987.
[LI] R.J. Lipinski et al., "Uncertainty in Radionuclide Release Under Specific LWR
Accident Conditions," Sandia National Laboratory, SAND84-0410, Vol.2,
February 1985.
[L2] M. Lee, "Modeling Core-Concrete Interaction," Ph.D Thesis, Department of
Nuclear Engineering, M.I.T., June 1985.
[L3] M. Lee, M.S. Kazimi, and G. Brown, "A Heat Transfer Model for the Corium/
Concrete Interface," Proceedings of the ANS/ENS International Meeting on
Light Water Reactor Severe Accident Evaluation, Cambridge, MA, August
28 - September 1, 1983.
[L4] M. Lee and M.S. Kazimi, "Modeling of Corium-Concrete Interactions,"
MITNE-267, Department of Nuclear Engineering, M.I.T., June 1985.
[Ml] J.F. Muir, "Response of Concrete Exposed to a High Heat Flux on One Surface," Sandia National Laboratory, SAND77-1467, November 1977.
[M2] J.F. Muir, D.A. Powers, and D.A. Dahlgren, "Studies on Molten Fuel-Concrete Interaction," Proceedings of the International Meeting on Fast Reactor
Safety and Related Physics, Chicago, Ill, October 5 - 8, 1976.
[M3] W.B. Murfin, "A Preliminary Model for Core/Concrete Interactions," Sandia
National Laboratory, SAND77-0370, August 1977.
[M4] J.F. Muir, et al., "CORCON/MOD1: An Improved Model for Molten Core/
Concrete Interaction," Sandia National Laboratory, NUREG/CR-2142,
SAND80-2415, July 1981.
[M5] J.F. Muir and A.S. Benjamin, "Modelling of Molten Core/Concrete Interactions," Proceedings of the ANS/ENS Topical Meeting on Thermal Reactor
Safety, CONF-800403, .I, pp.205, 1980.
[NI] USNRC, "Reactor Safety Study - An Assessment of Accident Risks in U.S.
Commercial Nuclear Power Plants," WASH-1400 (NUREG-75/014), October
1975.
[N2] USNRC, Advance Notice of Rulemaking, 45FR65474, October 2, 1980.
[N3] USNRC, Proposed Commission Policy Statement on Severe Accidents and
Releated Views on Nuclear Reactor Regulation, SECY-82-1B, 48FR16014,
April 13, 1983.
[N4] USNRC, NRC Policy Statement Regarding Future Reactor Design and Existing Plants, NUREG-1070, Draft, April 18, 1984.
[N5] USNRC, "Containment Performance Working Group Report," NUREG-1037,
Draft, May 1985.
325
[N6] USNRC, "Technical Bases for Estimating Fission Product Behavior During
LWR Accidents," NUREG-0772, June 1981.
[N7] "CSNI Benchmark Problems Related to Ex-Vessel Severe Accident ThermalHydraulics," Private Communication, NRC, December 1984.
[P1] D.A. Powers et al., "Exploratory Study of Molten Core Material/Concrete
Interactions," Sandia National Laboratory, SAND77-2042, February 1978.
[P2] M. Plys, "An Experimental Investigation of the Core-Concrete Interaction,"
Sc.D. Thesis, Department of Nuclear Engineering, M.I.T., February 1984.
[P3] D. Perinic et al., " Concrete Crucible Tests with Thermite Melts," Nuclear
Research Center, Karlsruhe, KfK-2572, NUREG/TR-669, 1979.
[P4] M. Peehs, A. Skokan, and M. Reimann, "The Behavior of Concrete in Contact
with Molten Corium in the Case of a Hypothetical Core Melt Accident,"
Nucl. Tech., 46, pp.192 - 198, December 1979.
[P5] D.A Powers and F.E. Arellano, "Large-Scale Transient Tests of the Interaction of Molten Steel with Concrete," Sandia National Laboratory, SAND81
-1753, NUREG/CR-2282, January 1982.
[P6] D.A. Powers, "Sustained Molten Steel/Concrete Interaction Tests, A Preliminary Report on the Feasibility of Experimental Technique," Sandia National
Laboratory, SAND77-1423, NUREG/CR-0166, June 1978.
[P7] D.A. Powers, J.E. Brockmann, and A.W. Shiver, "VANESA: A Mechanistic Model of Radionuclide Release and Aerosol Generation During Core Debris Interactions with Concrete," Sandia National Laboratory, NUREG/CR4308, SAND85-1370, June 1985.
[P8] M.G. Plys and R.E. Henry, "Ex-Vessel Fission Product Release Modeling,"
Trans. Am. Nucl. Soc., 50, pp.319, November 1985.
[P9] C. Park and M. Khatib-Rahbar, "Quantification and Uncertainty Analysis
of Source Terms for Severe Accidents in Light Water Reactors (QUASAR),
Part I - Methodology and Program Plan," Brookhaven National Laboratory,
NUREG /CR-4688, BNL-NUREG-52008, Vol.1, June 1986.
[P10] M.G. Plys, P.G. Ellison, and R.E. Henry, "Influence of Containment ThermalHydraulics on Source Term Compositions," Proceedings of the Third
International Topical Meeting on Reactor Thermal Hydraulics, Newport, RI,
October 15 - 18, 1985.
[P11] M.G. Plys, M.A. Kenton, and R.E. Henry, "Ex-Vessel Fission Product Release
Modeling within Integrated Accident Analysis," OECD/CSNI Specialist
Meeting on Core Debris/Concrete Interactions, Sept. 1986, Palo Alto, CA.
[P12] D.A. Powers and F.E. Arellano, "Direct Observation of Melt Behavior During
High Temperature Melt/Concrete Interactions," Sandia National Laboratory,
NUREG/CR-2283, SAND81-1754, January 1982.
326
[P13] D.A. Powers, "The Phenomena of the Ex-Vessel Source Term," International Symposium on Source Term Evaluation for Accident Conditions
Columbus, OH, October 1985.
[RI] M. Reimann and W.B. Murfin, "The WECHSL Code: A Computer Program
for the Interaction of Core Melt with Concrete," Karlsruhe, FRG, KfK-2890,
November 1981.
[S1] T.P. Speis et al., "Containment Loads Working Group Report," U.S. NRC,
NUREG-1079, Draft, July 1985.
[S2] M. Silberberg et al., "Reassessment of the Technical Bases for Estimating
Source Terms," U.S. NRC, NUREG-0956, July 1986.
[S3] "Light Water Reactor Safety Research Program Quarterly Report April June 1978," Sandia National Laboratory, SAND78-1901, NUREG/CR-0422,
January 1979.
[S4] "Light Water Reactor Safety Research Program Quarterly Report Jan. March 1980," Sandia National Laboratory, SAND80-1304, NUREG/CR1509, July 1980.
[S5] B.W. Spencer et al., "EPRI/ANL Investigations of MCCI Phenomena and
Aerosol Release," OECD/CSNI Specialist Meeting on Core Debris/Concrete
Interactions, Palo Alto, CA, September 3 - 5, 1986.
[S6] J. Szekely, "Mathematical Model for Heat and Mass Transfer at the BubbleStirred Interface of Two Immiscible Liquids," Intl. J. Heat Mass Transfer,
6, pp. 3 0 7 - 311, 1963.
[WI] H. Werle, "Enhancement of Heat Transfer between Two Horizontal Liquid
Layers by Gas Injection at the Bottom," Nucl. Tech., 59, pp.160-164, 1982.
[W2] R.O. Wooton and H.I. Avci, "MARCH: (Meltdown Accident Response
CHaracteristic) Code Description and User's Manual," Battelle Columbus
Laboratory, NUREG/CR-1711, BMI-2064, October 1980.
327
Download