-
Notifications
You must be signed in to change notification settings - Fork 0
/
thermalfluids.tex
983 lines (873 loc) · 60.5 KB
/
thermalfluids.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
\label{ch:thermalfluids}
This chapter focuses on the thermal-fluids modeling of prismatic HTGRs using Moltres.
This chapter comprises the following sections:
Section \ref{sec:thermal-prelim} presents a verification and validation of the model in simplified geometries,
Section \ref{sec:fuelcol} discusses the implementation of the model in a fuel column of a prismatic HTGR,
Section \ref{sec:ph1ex2} describes efforts in conducting Phase I Exercise 2 of the OECD/NEA MHTGR-350 Benchmark,
Section \ref{sec:thermal-coupling} introduces a simplified coupling scheme to solve Phase I Exercise 3 of the benchmark,
and Section \ref{sec:thermal-conc} concludes the chapter with a summary and the main points of the chapter.
\section{Preliminary studies}
\label{sec:thermal-prelim}
This section describes preliminary studies using Moltres and MOOSE heat conduction modules (Moltres/MOOSE) to simulate the heat transfer in prismatic HTGRs.
\subsection{Verification of the thermal-fluids model}
\label{sec:tf-ver}
This section analyzed a simplified cylindrical model whose analytical solution is known to verify the thermal-fluids model adopted in this work (see Section \ref{appendix:ver} for a description of the analytical solution).
Moltres/MOOSE obtained the numerical solution of the thermal-fluid equations from Section \ref{ch3:th}.
Figure \ref{fig:th-ver-mesh} displays the model geometry, which differentiates five subregions: fuel compact, helium gap, moderator, film, and coolant.
Table \ref{tab:th-ver-char} summarizes the geometry dimensions and the input parameters.
The model reference design was the GT-MHR.
The calculated moderator radius is the minimum distance between the fuel and coolant channels in the unit cell --- the fuel/coolant pitch minus the fuel compact and coolant channel radii.
The calculated coolant radius preserves the coolant channel volume.
The model assumed a sinusoidal power profile in the $z$-direction.
\begin{figure}[htbp!]
\centering
\includegraphics[width=0.50\linewidth]{figures-thermal/ver-mesh2}
\hfill
\caption{Model geometry axial layout.}
\label{fig:th-ver-mesh}
\end{figure}
\begin{table}[htbp!]
\centering
\caption{Problem characteristics.}
\label{tab:th-ver-char}
% \begin{tabular}{@{}l c S[table-format=2.2] c c}
\begin{tabular}{@{}l c c c c}
\toprule
\multicolumn{1}{c}{Parameter} & \multicolumn{1}{c}{Symbol} & \multicolumn{1}{c@{}}{Value} & \multicolumn{1}{c@{}}{Units} & \multicolumn{1}{c}{Reference} \\
\midrule
Fuel compact radius & R$_f$ & 0.6225 & cm & \cite{in_three-dimensional_2006} \\
Fuel channel radius & R$_g$ & 0.6350 & cm & \cite{in_three-dimensional_2006} \\
Coolant channel radius & - & 0.7950 & cm & \cite{in_three-dimensional_2006} \\
Fuel/coolant pitch & - & 1.8850 & cm & \cite{in_three-dimensional_2006} \\
Fuel column height & L & 793 & cm & \cite{in_three-dimensional_2006} \\
Coolant mass flow rate & $\dot{m}$ & 0.0176 & kg $\cdot s^{-1}$ & \cite{in_three-dimensional_2006} \\
Average power density & q$_{ave}$ & 35 & W $\cdot cm^{-3}$ & \cite{in_three-dimensional_2006} \\
Coolant inlet temperature & $T_{in}$ & 400 & $^{\circ}C$ & \cite{in_three-dimensional_2006} \\
Helium inlet pressure & P & 70 & bar & \cite{in_three-dimensional_2006} \\
Helium density & $\rho_c$ & 4.940 $\times 10^{-6}$ & kg $\cdot cm^{-3}$ & \cite{lemmon_thermophysical_2019} \\
Helium heat capacity & c$_{p,c}$ & 5188 & J $\cdot kg^{-1} \cdot K^{-1}$ & \cite{lemmon_thermophysical_2019} \\
Fuel compact thermal conductivity & k$_f$ & 0.07 & W $\cdot cm^{-1} \cdot K^{-1}$ & \cite{tak_numerical_2008} \\
Gap thermal conductivity & k$_g$ & 3 $\times 10^{-3}$ & W $\cdot cm^{-1} \cdot K^{-1}$ & \cite{tak_numerical_2008} \\
Moderator thermal conductivity & k$_m$ & 0.30 & W $\cdot cm^{-1} \cdot K^{-1}$ & \cite{tak_numerical_2008} \\
\midrule
\multicolumn{1}{c}{Calculated parameters} & & & & \\
\midrule
Calculated moderator radius & R$_m$ & 1.080 & cm & - \\
Coolant film radius & R$_{cf}$ & 1.090 & cm & - \\
Calculated coolant radius & R$_c$ & 1.349 & cm & - \\
Coolant average velocity & v$_c$ & 1794.33 & cm $\cdot s^{-1}$ & - \\
Film thermal conductivity & k$_{cf}$ & 1.722 $\times 10^{-3}$ & W $\cdot cm^{-1} \cdot K^{-1}$ & - \\
\bottomrule
\end{tabular}
\end{table}
Note that this is a simplified model only for verifying that the numerical solution agrees with the analytical solution.
Figure \ref{fig:th-ver-results} shows the axial and radial temperature profiles and demonstrates that both the analytical and numerical solutions exhibit good agreement.
The outlet coolant temperature is 770.2 $^{\circ}$C, whereas the maximum fuel temperature is 874.7 $^{\circ}$C.
\begin{figure}[htbp!]
\centering
\subfloat[Fuel centerline and bulk coolant axial temperatures.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/2D-preliminar-axial}
}
\subfloat[Radial temperature profile at z=L/2=396.5 cm.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/2D-preliminar-radial2}
}
\hfill
\caption{Comparison of the analytical and numerical temperature profiles.}
\label{fig:th-ver-results}
\end{figure}
\subsection{Unit cell problem}
\label{sec:unitcell}
This section solved the unit cell problem in the hot spot of the GT-MHR, which is the spot in the core with the largest power density.
This section intended to reproduce the results found by In et al. \cite{in_three-dimensional_2006} to validate the unit-cell model.
In et al.'s article solves a three-dimensional unit-cell model and gives one of the most thorough descriptions in the open literature.
Table \ref{tab:th-val-unit-char} presents the problem characteristics.
The article does not specify the solid's material properties, so the model used parameters from Tak et al. \cite{tak_numerical_2008}.
Figure \ref{fig:th-val-unit-model} displays the model geometry, and Figure \ref{fig:th-val-unit-model-b} the temperature-dependent material properties.
Additionally, In et al. used a chopped cosine as the power profile, while this study used the average value to simplify the analysis.
\begin{table}[htbp!]
\centering
\caption{Problem characteristics.}
\label{tab:th-val-unit-char}
% \begin{tabular}{@{}l c S[table-format=2.2] c c}
\begin{tabular}{@{}l c c c c}
\toprule
\multicolumn{1}{c}{Parameter} & \multicolumn{1}{c}{Symbol} & \multicolumn{1}{c@{}}{Value} & \multicolumn{1}{c@{}}{Units} & \multicolumn{1}{c}{Reference} \\
\midrule
Fuel compact radius & R$_f$ & 0.6225 & cm & \cite{in_three-dimensional_2006} \\
Fuel channel radius & R$_g$ & 0.6350 & cm & \cite{in_three-dimensional_2006} \\
Coolant channel radius & R$_c$ & 0.7950 & cm & \cite{in_three-dimensional_2006} \\
Fuel/coolant pitch & p & 1.8850 & cm & \cite{in_three-dimensional_2006} \\
Fuel column height & L & 793 & cm & \cite{in_three-dimensional_2006} \\
% input parameter characteristics
Coolant channel mass flow rate & $\dot{m}$ & 0.0176 & kg $\cdot s^{-1}$ & \cite{in_three-dimensional_2006} \\
Average power density & q$_{ave}$ & 35 & W $\cdot cm^{-3}$ & \cite{in_three-dimensional_2006} \\
Inlet coolant temperature & T$_{in}$ & 400 & $^{\circ}$C & \cite{in_three-dimensional_2006} \\
Helium inlet pressure & P & 70 & bar & \cite{in_three-dimensional_2006} \\
Helium density & $\rho$ & 4.94 $\times 10^{-6}$ & kg $\cdot cm^{-3}$ & \cite{lemmon_thermophysical_2019} \\
Helium heat capacity & c$_p$ & 5188 & J $\cdot kg^{-1} \cdot K^{-1}$ & \cite{lemmon_thermophysical_2019} \\
\midrule
\multicolumn{1}{c}{Calculated parameters} & & & & \\
\midrule
Coolant film radius & R$_{cf}$ & 0.8050 & cm & - \\
Coolant average velocity & v$_c$ & 1794.33 & cm $\cdot s^{-1}$ & - \\
Film thermal conductivity & k$_{cf}$ & 1.731 $\times 10^{-3}$ & W $\cdot cm^{-1} \cdot K^{-1}$ & - \\
\bottomrule
\end{tabular}
\end{table}
\begin{figure}[htbp!]
\centering
\includegraphics[width=0.40\linewidth]{figures-thermal/unitcell2}
\hfill
\caption{Axial layout of the model geometry.}
\label{fig:th-val-unit-model}
\end{figure}
\begin{figure}[htbp!]
\centering
\includegraphics[width=0.60\linewidth]{figures-thermal/val-unit-matprop}
\hfill
\caption{Temperature-dependent material properties \cite{tak_numerical_2008}.}
\label{fig:th-val-unit-model-b}
\end{figure}
Figure \ref{fig:th-val-unit-temps} shows the temperature profiles.
From the top to the bottom of the reactor, the axial temperatures increase, the moderator and coolant temperatures remain parallel, and the difference between the fuel and moderator temperatures decreases.
The model assumes a film thermal conductivity independent of the temperature; thus, the moderator-to-coolant temperature difference is constant.
The solids' thermal conductivity increases with temperature; as shown in Figure \ref{fig:th-val-unit-model-b}, the thermal resistance between the moderator and the fuel decreases, so their temperature difference decreases as well.
\begin{figure}[htbp!]
\centering
\subfloat[Maximum fuel, moderator, and bulk coolant axial temperatures.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/in-2006-5-axial}
}
\subfloat[Outlet plane (z=0 cm) temperature distribution.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/val-unit-outlet-plane}
}
\hfill
\caption{Unit cell temperature profile calculated by Moltres.}
\label{fig:th-val-unit-temps}
\end{figure}
Table \ref{tab:th-val-unit-results} summarizes the results.
Equation \ref{eq:rel-error} evaluates the relative difference to the reference results
\begin{align}
& \Delta_T = \left| \frac{T_M-T_{ref}}{T_{ref}} \right| \cdot 100 \label{eq:rel-error}
\intertext{where}
& \Delta_T = \mbox{relative difference } [\%] \notag \\
& T_M = \mbox{temperature determined by this work} [^{\circ}C] \notag \\
& T_{ref} = \mbox{reference temperature} [^{\circ}C]. \notag
\end{align}
Small discrepancies arise in the results: Moltres/MOOSE coolant temperature is smaller than In's by 4$^{\circ}$C.
The moderator temperature is larger by 9$^{\circ}$C, and the fuel temperature is larger by 22$^{\circ}$C.
The cause of the fuel temperature discrepancy is the power profile simplification.
As the previous section's analysis has shown, the fuel-to-coolant temperature difference is small at the outlet for a sinusoidal power profile.
The opposite extreme scenario is the uniform power profile, where the fuel-to-coolant temperature difference is larger.
In et al. used a chopped cosine power profile, which is between the two former cases.
Hence, this model yields a fuel-to-coolant temperature difference at the outlet larger than In et al.
Overall, Moltres results are close to In's results as the maximum difference between them is 22$^{\circ}$C, which corresponds to less than 2\% relative difference.
\begin{table}[htbp!]
\centering
\caption{Comparison between In et al. \cite{in_three-dimensional_2006} and Moltres/MOOSE results.}
\label{tab:th-val-unit-results}
\begin{tabular}{@{}l c c c}
\toprule
Parameter & In et al. [$^{\circ}$C] & Moltres/MOOSE [$^{\circ}$C] & $\Delta_T$ [\%]\\
\midrule
Maximum coolant temperature & 1144 & 1140 & 0.3 \\
Maximum moderator temperature & 1250 & 1259 & 0.7 \\
Maximum fuel temperature & 1295 & 1317 & 1.7 \\
\bottomrule
\end{tabular}
\end{table}
\section{Fuel column}
\label{sec:fuelcol}
This section discusses an HTGR fuel column analysis aimed at validating the fuel column model by reproducing some of Sato et al. \cite{sato_computational_2010} analyses.
This article gives a thorough description of the problem, making it possible to reproduce.
The first analysis studied a column with no bypass-gap between assemblies, while a second analysis studied a column with a 3mm bypass-gap between assemblies.
Figure \ref{fig:th-val-assem-model-a} displays the model geometry.
The model only considered a $1/12^{th}$ portion of the column due to symmetry.
The study used the GT-MHR as the reference reactor for the calculations.
The GT-MHR shares the geometry dimensions with the MHTGR, specified in Table \ref{tab:element-characteristics}.
\begin{figure}[htbp!]
\centering
\includegraphics[width=0.25\linewidth]{figures-thermal/val-assem-mesh}
\hfill
\caption{Model geometry axial layout.}
\label{fig:th-val-assem-model-a}
\end{figure}
Equation \ref{eq:matcoeff} evaluates the solid material properties \cite{johnson_cfd_2009}
\begin{align}
\phi(T) = A_1 + A_2 T + A_3 T^2 + A_4 T^3 + A_5 T^4. \label{eq:matcoeff}
\end{align}
Table \ref{tab:th-val-assem-mat} displays the fuel compact and moderator coefficients of equation \ref{eq:matcoeff}.
% Figure \ref{fig:th-val-assem-model-b} shows the temperature-dependent material properties.
\begin{table}[htbp!]
\centering
\caption{Thermal conductivity coefficients \cite{johnson_cfd_2009}.}
\label{tab:th-val-assem-mat}
\begin{tabular}{c| S[table-format=2.2] S[table-format=2.2e-1] S[table-format=3.3] |S[table-format=2.2e-1]}
% \begin{tabular}{c|ccc|c}
\toprule
& \multicolumn{3}{c|}{Moderator} & \multicolumn{1}{c}{Fuel compact} \\
Temperature range {[}K{]} & \multicolumn{1}{c}{255.6-816} & \multicolumn{1}{c}{816-1644.4} & \multicolumn{1}{c|}{1644.4-1922.2} & \multicolumn{1}{c}{255.6-2200} \\
\midrule
A1 & 28.6 & 1.24e+2 & 41.5 & 3.94 \\
A2 & & -3.32e-1 & & 3.59e-3 \\
A3 & & 4.09e-4 & & -1.98e-9 \\
A4 & & -2.11e-7 & & 3.19e-12 \\
A5 & & 4.02e-11 & & -9.77e-16 \\
\bottomrule
\end{tabular}
\end{table}
% \begin{figure}[htbp!]
% \centering
% \includegraphics[width=0.45\linewidth]{figures-thermal/val-assem-matprop}
% \hfill
% \caption{Temperature dependent material properties \cite{johnson_cfd_2009}.}
% \label{fig:th-val-assem-model-b}
% \end{figure}
Table \ref{tab:th-val-assem-char} lists the remaining input parameters.
Figure \ref{fig:th-val-assem-model-a} assigns a number as a label to each coolant channel and the bypass-gap.
The calculations adopted the mass flow distribution from Sato et al \cite{sato_computational_2010}, shown in Table \ref{tab:th-val-assem-massflow}.
\begin{table}[htbp!]
\centering
\caption{Constant problem characteristics.}
\label{tab:th-val-assem-char}
% \begin{tabular}{@{}l c S[table-format=2.2] c c}
\begin{tabular}{@{}l c c c c}
\toprule
\multicolumn{1}{c}{Parameter} & \multicolumn{1}{c}{Symbol} & \multicolumn{1}{c@{}}{Value} & \multicolumn{1}{c@{}}{Units} & \multicolumn{1}{c}{Reference} \\
\midrule
Inlet coolant temperature & T$_{in}$ & 490 & $^{\circ}$C & \cite{sato_computational_2010} \\
Helium inlet pressure & P & 70 & bar & \cite{sato_computational_2010} \\
Helium density & $\rho$ & 4.37 $\times 10^{-6}$ & kg $\cdot cm^{-3}$ & \cite{lemmon_thermophysical_2019} \\
Helium heat capacity & c$_p$ & 5188 & J $\cdot kg \cdot K$ & \cite{lemmon_thermophysical_2019} \\
Average power density & q$_{ave}$ & 27.88 & W $\cdot cm^{-3}$ & \cite{sato_computational_2010} \\
\midrule
\multicolumn{1}{c}{Calculated parameters} & & & & \\
\midrule
Coolant film radius & R$_{cf}$ & 0.804 & cm & - \\
Film thermal conductivity & k$_{cf}$ & 2.09 $\times 10^{-3}$ & W $\cdot cm^{-1} \cdot K^{-1}$ & - \\
\bottomrule
\end{tabular}
\end{table}
\begin{table}[htbp!]
\centering
\caption{Mass flow rate determined from Sato's calculation \cite{sato_computational_2010}. Values expressed in [g/s].}
\label{tab:th-val-assem-massflow}
\begin{tabular}{l|ccccccc}
\toprule
Channel & 1 & 2 & 3 & 4 & 5 & 6 & 7 \\
\midrule
No bypass-gap & 6.18 & 11.34 & 11.37 & 11.38 & 11.43 & 11.33 & 22.70 \\
3mm bypass-gap & 5.88 & 10.80 & 10.85 & 10.91 & 11.08 & 10.80 & 21.58 \\
\midrule
Channel & 8 & 9 & 10 & 11 & 12 & 13 & Gap \\
\midrule
No bypass-gap & 22.73 & 22.73 & 11.38 & 22.77 & 22.91 & 11.44 & - \\
3mm bypass-gap & 21.67 & 21.83 & 10.88 & 21.81 & 22.20 & 11.10 & 16.56 \\
\bottomrule
\end{tabular}
\end{table}
Table \ref{tab:th-val-assem-results} compares the results using Moltres/MOOSE to Sato's results \cite{sato_computational_2010}.
In the case with no bypass-gap, Moltres/MOOSE's maximum coolant temperature is 2 $^{\circ}$C lower, and the maximum fuel temperature is 4$^{\circ}$C higher.
In the case with a 3mm bypass-gap, Moltres/MOOSE's maximum coolant temperature is 2 $^{\circ}$C lower, and the maximum fuel temperature is 1$^{\circ}$C lower.
The results are within the 1\% difference to those in \cite{sato_computational_2010}.
Figure \ref{fig:th-val-assem-temps} shows the outlet temperature along lines A-B and A-C from Figure \ref{fig:th-val-assem-model-a}.
The bypass-flow reduces the peripheral temperature causing the center's temperature to rise.
The presence of the gap between assemblies produces a larger temperature gradient in the assembly.
\begin{table}[htbp!]
\centering
\caption{Comparison of the maximum temperatures calculated by this work and the reference values from Sato et al. \cite{sato_computational_2010}. Temperature values expressed in [$^{\circ}$C].}
\label{tab:th-val-assem-results}
\begin{tabular}{lcccccc}
\toprule
& \multicolumn{3}{c}{No gap} & \multicolumn{3}{c}{3mm gap} \\
\midrule
& $T_{ref}$ & Moltres/MOOSE & $\Delta_T$ [\%] & $T_{ref}$ & Moltres/MOOSE & $\Delta_T$ [\%] \\
\midrule
Coolant & 985 & 983 & 0.2 & 1007 & 1005 & 0.2 \\
Fuel & 1090 & 1094 & 0.4 & 1115 & 1114 & 0.1 \\
\bottomrule
\end{tabular}
\end{table}
\begin{figure}[htbp!]
\centering
% \hspace*{\fill}
\subfloat[Line A-B.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/val-assem-line-AB}
}
\subfloat[Line A-C.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/val-assem-line-AC}
}
\hfill
\caption{Outlet plane temperature along the line A-B and line A-C.}
\label{fig:th-val-assem-temps}
\end{figure}
\subsection{Flow distribution analysis}
\label{sec:flowdistrib}
As described in Section \ref{sec:litrev-thermalf}, several methods are available to calculate the coolant flow distribution in prismatic HTGRs.
This section compares a few such methods.
The metrics chosen for this comparison are the maximum coolant and fuel temperatures and the mass flow distribution.
Figure \ref{fig:th-val-assem-model-a} assigns a number as a label to each coolant channel and the bypass-gap.
Note that Channel 1 is a small coolant channel while Channels 2 to 13 are large coolant channels.
This study adopted the mass flow distribution from Sato et al. \cite{sato_computational_2010} as the reference and analyzed the following cases:
\begin{itemize}
\item Case 1: uses a flat velocity profile (every channel and gap has the same velocity).
\item Case 2: uses the incompressible flow model with temperature-independent helium viscosity.
\item Case 3: uses the incompressible flow model with a temperature-dependent helium viscosity.
\item Case 4: uses the low-Mach number model with a temperature-dependent helium viscosity.
\end{itemize}
Case 1 assumes a flat velocity profile in all coolant paths and calculated the mass flow using equation \ref{eq:case1mdot}
\begin{align}
& \dot{m}_i = \frac{A_i}{\sum_j A_j} \dot{m}_T \label{eq:case1mdot}
\intertext{where}
& \dot{m}_i = \mbox{channel $i$ mass flow rate } [kg \cdot s^{-1}] \notag \\
& A_i = \mbox{channel $i$ cross-sectional area } [cm^2] \notag \\
& \dot{m}_T = \mbox{total mass flow rate } [kg \cdot s^{-1}]. \notag
\end{align}
Case 2 and 3 used equation \ref{eq:case2mdot} \cite{melese_thermal_1984} to calculate the pressure drop
\begin{align}
& \Delta P = \frac{\dot{m}_i^2}{\rho_i A_i^2} f \frac{2 L}{D_h} \label{eq:case2mdot} \\
& B_i = \frac{1}{\rho_i A_i^2} f \frac{2 L}{D_h}
\intertext{where}
% & \dot{m}_i = \mbox{channel $i$ mass flow rate } [kg \cdot s^{-1}] \notag \\
% & A_i = \mbox{channel $i$ cross-sectional area } [cm^2] \notag \\
& \Delta P = \mbox{pressure drop } [Pa] \notag \\
& \rho_i = \mbox{channel inlet coolant density } [kg \cdot m^{-3}] \notag \\
& f = \mbox{friction factor } [-] \notag \\
& L = \mbox{channel length } [m] \notag \\
& D_h = \mbox{hydraulic diameter } [m]. \notag
\end{align}
Case 3 differs from Case 2 as the friction factor $f$ depends on the average channel temperature.
Case 4 used equation \ref{eq:case4mdot} \cite{melese_thermal_1984} to calculate the pressure drop
\begin{align}
& \Delta P = \frac{\dot{m}_i^2}{2 \rho_i A_i^2} \left[ \frac{4 f L (T_i+T_o)}{2 D T_i} + \frac{T_o-T_i}{T_i} \right] \label{eq:case4mdot} \\
& B_i = \frac{1}{2 \rho_i A_i^2} \left[ \frac{4 f L (T_i+T_o)}{2 D T_i} + \frac{T_o-T_i}{T_i} \right]
\intertext{where}
& T_i = \mbox{channel inlet coolant temperature } [^{\circ}C] \notag \\
& T_o = \mbox{channel outlet coolant temperature } [^{\circ}C]. \notag \\
\end{align}
Equations \ref{eq:case2mdot} and \ref{eq:case4mdot} used an iterative solver to obtain $\dot{m}_i$ \cite{melese_thermal_1984}(for further detail refer to Section \ref{appendix:equations-fluid})
\begin{align}
& B_i^{(n)} = f \left(\dot{m}_i^{(n)} \right) \\
& \Delta P^{(n)} = \frac{\dot{m}_T}{\sum_i \frac{1}{\sqrt{ B_i^{(n)} }}} \\
& \dot{m}_i^{(n+1)} = \sqrt{\frac{\Delta P^{(n)}}{ B_i^{(n)} }}
\intertext{with a convergence criterion of}
& \left| \Delta P^{(n+1)} - \Delta P^{(n)} \right| < 10^{-4}.
\end{align}
The solution of Cases 3 and 4 depends on the temperature, requiring a second level of iterations that updates the material properties.
The convergence criteria were 1$^{\circ}$C for the maximum coolant and fuel temperatures.
These iterative solvers were implemented in a Python script external to Moltres.
Table \ref{tab:th-assem-flow-massflow} shows the results achieved in this thesis for the mass flow rates.
Equation \ref{eq:rel-error2} evaluates the relative difference to the reference results
\begin{align}
& \Delta_m = \left| \frac{\dot{m}_i-\dot{m}_{r, i}}{\dot{m}_{r, i}} \right| \cdot 100 \label{eq:rel-error2}
\intertext{where}
& \Delta_m = \mbox{relative difference } [\%] \notag \\
& \dot{m}_i = \mbox{channel $i$ mass flow rate determined by this work } [g \cdot s^{-1}] \notag \\
& \dot{m}_{r, i} = \mbox{reference mass flow rate } [g \cdot s^{-1}]. \notag
\end{align}
% Case 4 and case 5 converged after two and three iterations, respectively.
Case 1 yields the largest differences to reference results, yielding a 13.3\% difference to the reference small coolant channel mass flow.
Additionally, Case 1 yields the smallest large coolant channel mass flow, the differences being within 6.2\%.
Case 2 and 3 results are within the 3.3\% difference to the reference results.
Case 2 and 3 barely differ, proving that not considering the viscosity's temperature dependency yields a more straightforward method with acceptable accuracy.
Case 4 results are within 1\% of the reference results except for the gap mass flow that yields a difference of 4.2\%.
For the most part, Case 4 arrives at the values closest to the reference solution.
Table \ref{tab:th-assem-flow-results} summarizes the maximum temperatures and the relative difference to the reference values using equation \ref{eq:rel-error}.
Case 1 yields the largest difference for the maximum coolant temperature, which is less than 10$^{\circ}$C.
Case 4 yields the results closest to the reference values, showing a difference of 0.2\%.
For the maximum fuel temperature, Case 1 and 4 yield the best results.
Again, Case 2 and 3 results barely differ.
The low-Mach number model (Case 4) yielded the closest results to the reference solution.
However, such a method required a two-level iterative solver.
From a computational perspective, the Case 1 model is the most straightforward as it does not require an iterative solver.
Additionally, the Case 1 model yields the simplest Moltres input file.
For these reasons, the rest of the thesis adopts the flat velocity approximation for the fluid flow distribution.
\begin{table}[htbp!]
\centering
\caption{Comparison of the calculated flow rates and the reference values \cite{sato_computational_2010}. Mass flow rates values expressed in [$g \cdot s^{-1}$].}
\label{tab:th-assem-flow-massflow}
\begin{tabular}{c S S S S S S S S S}
\toprule
& & \multicolumn{2}{c}{Case 1} & \multicolumn{2}{c}{Case 2} & \multicolumn{2}{c}{Case 3} & \multicolumn{2}{c}{Case 4} \\
\midrule
Channel & \multicolumn{1}{c}{$\dot{m}_{r, i}$} & \multicolumn{1}{c}{$\dot{m}_{i}$} & \multicolumn{1}{c}{$\Delta_m$ [\%]} & \multicolumn{1}{c}{$\dot{m}_{i}$} & \multicolumn{1}{c}{$\Delta_m$ [\%]} & \multicolumn{1}{c}{$\dot{m}_{i}$} & \multicolumn{1}{c}{$\Delta_m$ [\%]} & \multicolumn{1}{c}{$\dot{m}_{i}$} & \multicolumn{1}{c}{$\Delta_m$ [\%]} \\
\midrule
1 & 5.88 & 6.66 & 13.3 & 5.98 & 1.7 & 5.97 & 1.5 & 5.83 & 0.9 \\
2 & 10.80 & 10.41 & 3.6 & 10.91 & 1.0 & 10.90 & 0.9 & 10.75 & 0.5 \\
3 & 10.85 & 10.41 & 4.1 & 10.91 & 0.6 & 10.90 & 0.5 & 10.81 & 0.4 \\
4 & 10.91 & 10.41 & 4.6 & 10.91 & 0.0 & 10.91 & 0.0 & 10.90 & 0.1 \\
5 & 11.08 & 10.41 & 6.0 & 10.91 & 1.5 & 10.92 & 1.4 & 11.09 & 0.1 \\
6 & 10.80 & 10.41 & 3.6 & 10.91 & 1.0 & 10.90 & 0.9 & 10.73 & 0.6 \\
7 & 21.58 & 20.82 & 3.5 & 21.82 & 1.1 & 21.80 & 1.0 & 21.58 & 0.0 \\
8 & 21.67 & 20.82 & 3.9 & 21.82 & 0.7 & 21.81 & 0.6 & 21.71 & 0.2 \\
9 & 21.83 & 20.82 & 4.6 & 21.82 & 0.0 & 21.83 & 0.0 & 21.92 & 0.4 \\
10 & 10.88 & 10.41 & 4.3 & 10.91 & 0.3 & 10.91 & 0.3 & 10.84 & 0.4 \\
11 & 21.81 & 20.82 & 4.5 & 21.82 & 0.0 & 21.82 & 0.0 & 21.87 & 0.3 \\
12 & 22.20 & 20.82 & 6.2 & 21.82 & 1.7 & 21.85 & 1.6 & 22.26 & 0.3 \\
13 & 11.10 & 10.41 & 6.2 & 10.91 & 1.7 & 10.92 & 1.6 & 11.08 & 0.2 \\
Half-gap & 8.28 & 8.20 & 1.0 & 8.55 & 3.3 & 8.55 & 3.3 & 8.63 & 4.2 \\
\bottomrule
\end{tabular}
\end{table}
\begin{table}[htbp!]
\centering
\caption{Comparison of the maximum temperatures between Moltres-derived and the reference results. Temperature values expressed in [$^{\circ}$C].}
\label{tab:th-assem-flow-results}
\begin{tabular}{l S[table-format=3.0] S[table-format=3.0] S[table-format=1.1] S[table-format=3.0] S[table-format=1.1] S[table-format=3.0] S[table-format=1.1] S[table-format=3.0] S[table-format=1.1]}
\toprule
& & \multicolumn{2}{c}{Case 1} & \multicolumn{2}{c}{Case 2} & \multicolumn{2}{c}{Case 3} & \multicolumn{2}{c}{Case 4} \\
\midrule
& \multicolumn{1}{c}{$T_{ref}$} & \multicolumn{1}{c}{T} & \multicolumn{1}{c}{$\Delta_T$ [\%]} & \multicolumn{1}{c}{T} & \multicolumn{1}{c}{$\Delta_T$ [\%]} & \multicolumn{1}{c}{T} & \multicolumn{1}{c}{$\Delta_T$ [\%]} & \multicolumn{1}{c}{T} & \multicolumn{1}{c}{$\Delta_T$ [\%]} \\
\midrule
Coolant & 1005 & 994 & 1.1 & 999 & 0.6 & 1000 & 0.5 & 1007 & 0.2 \\
Fuel & 1114 & 1116 & 0.2 & 1109 & 0.4 & 1110 & 0.4 & 1116 & 0.2 \\
\bottomrule
\end{tabular}
\end{table}
\subsection{Mesh convergence analysis}
\label{sec:meshconverge}
% I might not include this in the final document
The remainder of this chapter intends to solve the full-core problem.
This section aims to identify some possible problems introduced by the full-core problem's large mesh size requirement.
This section studies a mesh convergence analysis of the full-fuel column problem.
Figure \ref{fig:th-full-assem-mesh} displays the model geometry, and Table \ref{tab:th-full-assem-results} presents the results.
The convergence criteria were 1$^{\circ}$C for the maximum coolant and fuel temperatures.
The coolant temperature converged for the fifth mesh, but the fuel temperature did not converge.
Further refinement of the mesh was not possible as the simulation memory requirements were too high.
\begin{figure}[htbp!]
\centering
\includegraphics[width=0.50\linewidth]{figures-thermal/full-assem-mesh2}
\hfill
\caption{Axial view of the full-fuel column model geometry.}
\label{fig:th-full-assem-mesh}
\end{figure}
\begin{table}[htbp!]
\centering
\caption{Mesh characteristics and maximum temperatures of the full-fuel column model.}
\label{tab:th-full-assem-results}
\begin{tabular}{lccccc}
\toprule
& Mesh 1 & Mesh 2 & Mesh 3 & Mesh 4 & Mesh 5 \\
\midrule
Number of elements [$\times 10^{6}$] & 1.025 & 1.306 & 1.833 & 2.596 & 3.006 \\
Number of DoFs [$\times 10^{6}$] & 0.524 & 0.666 & 0.932 & 1.317 & 1.525 \\
Maximum coolant temperature & 1060.41 & 1062.23 & 1064.00 & 1065.13 & 1065.32 \\
Maximum fuel temperature & 1204.49 & 1217.32 & 1225.57 & 1233.44 & 1234.93 \\
\bottomrule
\end{tabular}
\end{table}
This analysis reveals a potential problem.
The high level of detail in the geometry requires a large number of elements in the mesh.
In the full-scale problem, the dimensions and the number of elements in the mesh both increase.
This method will potentially be unable to solve the three-dimensional full-scale problem due to a high memory requirement.
For this reason, the next section intends to solve the full-scale problem using a two-dimensional cylindrical model.
% Maybe talk about that the number of nodes in the coolant channels could be reduced as we are using the fluid 1D equations.
% However, the coolant channel mesh requires enough points in the circle to properly model the heat transfer between the solid and the coolant.
% The ideal mesh would have many points in the boundary of the cylinder and one point in the cylinder center.
% But that is not numerically good, as the jacobian of those triangles is too as mall and could cause issues on the simulations.
% \section{Full-core 3D}
% This section will extend the methodology to a full-core problem and it will intend to solve Exercise 2 of Phase I of the OECD/NEA MHTGR-350 Benchmark.
\section{OECD/NEA MHTGR-350 Benchmark: Phase I Exercise 2}
\label{sec:ph1ex2}
This section describes Phase I Exercise 2 of the OECD/NEA MHTGR-350 Benchmark \cite{oecd_nea_benchmark_2017} and the exercise's results calculated by Moltres.
Phase I Exercise 2 defines a thermal-fluid stand-alone calculation.
The exercise aims to ensure that the thermal-fluid model differences between participants are negligible and will not affect the coupled exercises.
The benchmark specifies the power density of each fuel region and defines the type of mass flow distribution and material properties for four sub-cases:
\begin{itemize}
\item Exercise 2a: No bypass flow and fixed thermo-physical properties. The model does not account for the bypass flow, and the thermo-physical properties are constant.
\item Exercise 2b: Bypass flow type I and fixed thermo-physical properties. This exercise prescribes the bypass flow distribution, and the thermo-physical properties are constant.
\item Exercise 2c: Bypass flow type I and variable thermo-physical properties. This exercise prescribes the bypass flow distribution, and the thermo-physical properties depend on different simulation parameters.
\item Exercise 2d: Bypass flow type II and variable thermo-physical properties. This exercise solves the bypass flow distribution through the explicit modeling of the bypass gaps. The thermo-physical properties depend on different simulation parameters.
\end{itemize}
The exercise requires reporting the average and maximum temperature values of the reflector, \gls{RPV}, fuel, moderator, and coolant.
It also requires reporting the RPV heat flux and the mass flow rate distribution in the coolant channels and the bypass gaps.
These data are helpful to trace the source of possible differences in the primary parameters between participants.
Since OECD/NEA did not publish the results for this exercise, this section compares Moltres/MOOSE results against INL's benchmark results \cite{strydom_inl_2013}.
This section presents only a subset of available data that illustrate the main characteristics of the exercise.
Figure \ref{fig:ex2a-1st-model} displays the model geometry.
The model uses several simplifications to conduct the exercise.
In the axial direction, it does not consider the upper plenum and the outlet plenum, and the axial boundaries are the top reflector's upper face and the bottom reflector's lower face.
The top reflector's upper face and the inlet coolant flow are at 259 $^{\circ}$C, while the bottom reflector's lower face is adiabatic.
The outlet coolant uses an outflow boundary condition.
In the radial direction, the model does not consider the core barrel and the helium gap.
The model includes the RPV, followed by the outside air.
The outside air region outer boundary is at 30 $^{\circ}$C.
% model H figure
\begin{figure}[htbp!]
\centering
\includegraphics[width=0.50\linewidth]{figures-thermal/ex2a-meshD2}
\hfill
\caption{Phase I Exercise 2 model geometry for simulation in Moltres.}
\label{fig:ex2a-1st-model}
\end{figure}
The core model geometry uses INL's model as the reference design.
Nine rings define the solid structures in the core, and the other nine rings the coolant flow in the core.
The radii of the rings preserve the assemblies' volume.
The coolant thickness is constant for all the rings and preserves the coolant volume in the core.
% missing description of the power density distribution
Table \ref{tab:th-ex2a} presents the material properties.
All the graphite regions assume the grade H-451 graphite material properties.
\begin{table}[htbp!]
\centering
\caption{Problem characteristics.}
\label{tab:th-ex2a}
% \begin{tabular}{@{}l S[table-format=3.5] c c}
\begin{tabular}{@{}l c c c}
\toprule
\multicolumn{1}{c}{Parameter} & \multicolumn{1}{c@{}}{Value} & \multicolumn{1}{c@{}}{Units} & \multicolumn{1}{c}{Reference} \\
\midrule
Fuel compact thermal conductivity & 20 & W $\cdot m^{-1} \cdot K^{-1}$ & \cite{oecd_nea_benchmark_2017} \\
Fuel block thermal conductivity & 37 & W $\cdot m^{-1} \cdot K^{-1}$ & \cite{oecd_nea_benchmark_2017} \\
Graphite thermal conductivity & 66 & W $\cdot m^{-1} \cdot K^{-1}$ & \cite{oecd_nea_benchmark_2017} \\
\gls{RPV} thermal conductivity & 40 & W $\cdot m^{-1} \cdot K^{-1}$ & \cite{oecd_nea_benchmark_2017} \\
Coolant thermal conductivity & 0.41 & W $\cdot m^{-1} \cdot K^{-1}$ & \cite{oecd_nea_benchmark_2017} \\
Air thermal conductivity & 0.068 & W $\cdot m^{-1} \cdot K^{-1}$ & \cite{oecd_nea_benchmark_2017} \\
Helium density & 5.703 & $\times 10^{-6}$ kg $\cdot cm^{-3}$ & \cite{lemmon_thermophysical_2019} \\
Helium heat capacity & 5188 & J $\cdot kg^{-1} \cdot K^{-1}$ & \cite{lemmon_thermophysical_2019} \\
\bottomrule
\end{tabular}
\end{table}
Table \ref{tab:th-ex2a-1st-results} displays the results.
The model considerably under-predicts the inner reflector rings temperature, while it over-predicts the coolant rings temperature.
With the purpose of better understanding this behavior, Figure \ref{fig:ex2a-1st-model-across} shows the temperature across the reactor at the bottom of the active core (z=200 cm).
This figure exposes that the temperatures in the active core are well above the other region temperatures.
Although such behavior is expected, the temperature profile reveals some lack of heat transfer between the different rings.
In other words, the heat transfer from the fuel rings to the rest of the core structures is smaller than the heat transfer to the coolant rings in the active core region.
Consequently, the reflector temperatures are too low, and the coolant temperature in the fuel rings too high.
These results indicate that Moltres' model fails to capture some heat transfer mechanisms that INL's model captures.
The most obvious difference between the models is the inclusion of the radiative heat transfer.
Future Moltres development efforts will focus on confirming this supposition and adding the capability to model the radiative heat transfer.
\begin{table}[htbp!]
\centering
\caption{Comparison of this work and INL \cite{strydom_inl_2013} first bottom core level average temperatures.}
\label{tab:th-ex2a-1st-results}
\begin{tabular}{lcccccc}
\toprule
& \multicolumn{3}{c}{Reflector} & \multicolumn{3}{c}{Coolant} \\
& Ring 1 & Ring 2 & Ring 3 & Ring 4 & Ring 5 & Ring 6 \\
\midrule
INL [$^{\circ}C$] & 790 & 794 & 802 & 797 & 636 & 673 \\
Moltres/MOOSE [$^{\circ}C$] & 268 & 313 & 769 & 1424 & 1597 & 1157 \\
$\Delta_T$ [\%] & 66 & 61 & 4 & 79 & 151 & 72 \\
\bottomrule
\end{tabular}
\end{table}
\begin{figure}[htbp!]
\centering
\includegraphics[width=0.50\linewidth]{figures-thermal/ex2a-across}
\hfill
\caption{Radial temperature at the bottom of the active core (z=200 cm) determined by Moltres.}
\label{fig:ex2a-1st-model-across}
\end{figure}
% left here on review
To circumvent this barrier, the following analysis considered a second model based on Stainsby's approach \cite{stainsby_investigation_2008}.
% which defines only one coolant ring for each fuel ring.
The calculation used a global model to obtain the coolant temperature in the core and then a sub-channel model to get the fuel and moderator average temperatures.
The sub-channel model used half of the unit cell from Section \ref{sec:unitcell}.
Figure \ref{fig:ex2a-2nd-model} displays the global model geometry.
Three fuel rings represent the three rings of fuel assemblies in the MHTGR-350 (see Figure \ref{fig:layout}).
In the middle of each fuel ring, a coolant ring defines the coolant flow in that fuel ring.
The rings' radii preserve the assemblies' volume, and the coolant ring volumes preserve the coolant volume in each of the fuel rings.
The model uses the material properties from Table \ref{tab:th-ex2a}.
Figure \ref{fig:ex2a-temps} presents the coolant and average solid temperatures in the different fuel rings.
The coolant temperature increases from the top to the bottom of the active core, and it is constant in the bottom and top reflector regions.
The highest coolant, fuel, and moderator temperatures are in Fuel Ring 1.
Table \ref{tab:th-ex2a-results} compares Moltres/MOOSE results to INL's \cite{strydom_inl_2013}.
For the coolant temperature, Moltres results are 9.9\% lower in Fuel Ring 1, and 1.6\% and 3.4\% higher in Fuel Rings 2 and 3.
Although the discrepancies are small, the global model fails to correctly distribute the heat produced in the fuel rings into the coolant rings.
A possible cause of this discrepancy is the flat velocity approximation.
Section \ref{sec:flowdistrib} showed that a flat velocity distribution of the coolant is reasonable in the fuel column model; however, one of the assumptions of that model was a uniform power density.
In this exercise, the power density is not uniform, which could explain the discrepancies.
Further studies will confirm their cause.
Additionally, Moltres predicts a smaller coolant-to-moderator temperature difference for Fuel Rings 1 and 2.
The most considerable discrepancy is in Fuel Ring 1.
INL's model coolant-to-moderator temperature difference is 46$^{\circ}$C while Moltres/MOOSE predicts a difference of 30$^{\circ}$C.
The flat velocity approximation might be again the source of these discrepancies.
Finally, the moderator-to-fuel temperature differences from INL and Moltres/MOOSE are close.
In Fuel Ring 1, the differences are 20$^{\circ}$C and 22$^{\circ}$C for INL and Moltres/MOOSE results.
In fuel ring 2, the differences are 16$^{\circ}$C and 17$^{\circ}$C for INL and Moltres/MOOSE results.
% model G figure
\begin{figure}[htbp!]
\centering
\includegraphics[width=0.50\linewidth]{figures-thermal/ex2a-meshC2}
\hfill
\caption{Model geometry.}
\label{fig:ex2a-2nd-model}
\end{figure}
% model G results
\begin{figure}[htbp!]
\centering
\subfloat[Coolant temperature in the different fuel rings.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/ex2a-fullcore-cool}
}
\subfloat[Average moderator and fuel temperatures in the different fuel rings.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/ex2a-solid}
}
\hfill
\caption{Axial temperatures determined by Moltres.}
\label{fig:ex2a-temps}
\end{figure}
\begin{table}[htbp!]
\centering
\caption{Comparison of this work and INL \cite{strydom_inl_2013} first bottom core level average temperatures. Temperature values expressed in [$^{\circ}$C].}
\label{tab:th-ex2a-results}
\begin{tabular}{l|ccc|ccc|ccc}
\toprule
& \multicolumn{3}{c|}{Fuel Ring 1} & \multicolumn{3}{c|}{Fuel Ring 2} & \multicolumn{3}{c}{Fuel Ring 3} \\ %\cline{2-10}
& Coolant & Moderator & Fuel & Coolant & Moderator & Fuel & Coolant & Moderator & Fuel \\
\midrule
INL & 797 & 843 & 863 & 636 & 672 & 688 & 673 & Not shown & 722 \\
Moltres & 718 & 748 & 770 & 646 & 669 & 686 & 696 & 721 & 739 \\
$\Delta_T$ [\%] & 9.9 & 11.3 & 10.8 & 1.6 & 0.4 & 0.3 & 3.4 & - & 2.4 \\
\bottomrule
\end{tabular}
\end{table}
\section{Coupled simulations}
\label{sec:thermal-coupling}
This section sets the roadmap for conducting neutronics and thermal-fluids coupled simulations of prismatic HTGRs in Moltres.
\subsection{OECD/NEA MHTGR-350 Benchmark: Phase I Exercise 3}
\label{sec:ph1ex3}
% description of the exercise
This section presents Phase I Exercise 3 of the OECD/NEA MHTGR-350 Benchmark and describes the exercise conducted with Moltres/MOOSE.
Exercise 3 combines all the data from the first two exercises of the benchmark, in which the participants need to determine a coupled neutronic and thermal-fluids solution.
The exercise requires the reporting of the same parameters reported in Exercises 1 and 2 combined.
The benchmark specifies the group constants necessary to conduct the exercise.
The group constants depend on four state parameters: moderator ($T_m$) and fuel temperature ($T_f$) and xenon-135 ($C_{Xe}$) and hydrogen concentration ($C_H$).
In addition to these data, the benchmark provides fluence maps to determine the thermal conductivity of graphite.
Two sub-cases compose Exercise 3:
\begin{itemize}
\item Exercise 3a: Same thermal-fluids problem definition from Exercise 2c. This exercise prescribes the bypass flow distribution and the thermo-physical properties depend on the temperature and fluence.
\item Exercise 3b: Same thermal-fluids problem definition from Exercise 2d. This exercise solves the bypass flow distribution through the explicit modeling of the bypass gaps. The thermo-physical properties depend on the temperature and fluence.
\end{itemize}
% initial coupling: description of what I've done
Section \ref{sec:ph1ex2} solved the thermal-fluids stand-alone problem using a global and a sub-channel model.
Those simulations required running two separate input files, where the output of the global model served as an input to the sub-channel model.
Exercise 3 requires modeling the temperature feedback.
Using the Section \ref{sec:ph1ex2} approach for the temperature feedback would require the fully-coupled simulation of both input files.
However, MOOSE framework did not have the capability to couple these specific input files when carrying out this exercise.
To solve Exercise 3, the analysis used a simplified global model, requiring the simulation of only one input file.
The global model homogenizes the fuel and moderator materials.
The explicit modeling of the fuel in the global model requires the conversion of the fuel dimensions in the three-dimensional geometry to the simplified cylindrical model.
This conversion requires the conservation of two parameters: the fuel volume and the fuel thickness.
The fuel volume preservation ensures that the reactor power does not change in the approximation.
On the other hand, the maximum fuel temperature is sensitive to the fuel thickness, requiring its conservation.
A cylindrical model cannot preserve both parameters simultaneously, and the global model homogenizes the fuel and moderator materials.
Hence, the solution of this model did not differentiate between moderator and fuel.
% describe the model
Figure \ref{fig:coupled-mesh} presents the model geometry.
The model includes 28 fuel and 29 coolant rings.
The radii of the rings were calculated by preserving the assemblies' and the coolant volume.
The coolant ring pitch is the coolant channel pitch in a fuel assembly.
Exercise 3a requires using material properties from Exercise 2c.
This simplified model used the material properties from Exercise 2a, as shown in Table \ref{tab:th-ex2a}.
Regarding the group-constants, the benchmark prescribes them for 232 regions in the reactor.
Table \ref{tab:coupled-rz} shows what benchmark sub-domains integrate each region in the model.
The model did not include the control rod region (sub-domain 232).
The simulations used a three energy-group structure.
Table \ref{tab:coupled-eg} indicates what benchmark group numbers integrate each model group.
Conducting this exercise required developing a Python tool to translate the benchmark group constants to Moltres format.
The tool was responsible for homogenizing and collapsing the group constants as well.
For further detail on the group constants handling, refer to Sections \ref{appendix:group-const-homo} to \ref{appendix:group-const-bench}.
The problem assumed the same boundary conditions from Section \ref{sec:ph1ex2}.
Because Moltres can decouple the neutronics from the thermal-fluid effects, and for the sake of comparison, the exercise was conducted with and without thermal feedback.
The calculations without thermal feedback assumed the group constants at 550 $^{\circ}$C.
Figure \ref{fig:coupled-results-flux} and \ref{fig:coupled-results-th} display the results on two selected lines across the core, axial line on $r=85 cm$ and radial line on $z=556.5 cm$.
The thermal feedback affects the flux, which consequently affects the thermal-fluids.
The axial flux peak moves towards the reactor top, where the temperatures are lower than the bottom.
Thus, the heat production shifts towards the reactor top, and the temperatures near the reactor outlet decrease.
The radial temperature profile has a similar shape to the thermal flux, as this group is the main contributor to fission and thus the power profile.
The radial flux does not change considerably; hence the temperature profile does not change much either.
The thermal feedback moves the radial temperature profile up because the heat production shifts towards the top.
\begin{figure}[htbp!]
\centering
\includegraphics[width=0.50\linewidth]{figures-thermal/ex3-mesh}
\hfill
\caption{Axial view of the model geometry.}
\label{fig:coupled-mesh}
\end{figure}
\begin{table}[htbp!]
\centering
\caption{Homogenization scheme.}
\label{tab:coupled-rz}
\begin{tabular}{l c}
\toprule
Model homogenized region & Benchmark sub-domains \\
\midrule
Fuel & 1-220 \\
Bottom reflector & 221-224 \\
Top reflector & 228-231 \\
Inner reflector & 225 \\
Outer reflector & 226-227 \\
\bottomrule
\end{tabular}
\end{table}
\begin{table}[htbp!]
\centering
\caption{Energy group condensation scheme.}
\label{tab:coupled-eg}
\begin{tabular}{c c}
\toprule
Model group number & Benchmark group number \\
\midrule
1 & 1-4 \\
2 & 5-15 \\
3 & 16-26 \\
\bottomrule
\end{tabular}
\end{table}
\begin{figure}[htbp!]
\centering
\subfloat[Axial flux over the line at r=85 cm.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/coupledD-3G-axial}
}
\subfloat[Radial temperature at z=L/2=556.5 cm.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/coupledD-3G-radial}
}
\hfill
\caption{Flux profile comparison. F: thermal feedback, NF: no thermal feedback.}
\label{fig:coupled-results-flux}
\end{figure}
\begin{figure}[htbp!]
\centering
\subfloat[Axial temperature at r=85 cm.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/coupledD-3G-axialT}
}
\subfloat[Radial temperature at z=L/2=556.5 cm.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/coupledD-3G-radialT}
}
\hfill
\caption{Temperature profile comparison. F: thermal feedback, NF: no thermal feedback.}
\label{fig:coupled-results-th}
\end{figure}
\subsection{Discussion}
\label{sec:discuss}
This section analyzes Moltres' areas of improvement for conducting prismatic HTGR coupled simulations.
% Describe moltres heterogenous capabilities as an MSR solver
Moltres' initial development targeted MSRs, allowing them to rely on heterogeneous diffusion calculations.
In a heterogeneous solver, each mesh node holds the information of each variable.
In previous work \cite{lindsay_introduction_2018}\cite{pater_multiphysics_2019}, the authors defined two materials in Moltres input files: the moderator and the fuel.
For such a configuration, a moderator mesh node holds the neutronics and temperature information only of the moderator.
The same is true for the fuel.
For these calculations, Moltres defines the neutron flux and the temperature on each node, and each node uses its temperature to compute the thermal feedback for the neutron flux.
% Moltres should use average temperatures
As discussed in Chapter \ref{ch:neutronics}, in a prismatic HTGR simulation, the neutronics calculation requires an assembly-level homogenization of the group constants.
Because of the homogenization, a mesh node contains the information of the combined group constants' from the different materials in the assembly --- only the fuel and the moderator as the coolant does not contribute considerably.
As the fuel group constants depended on the fuel temperature, and the moderator group constants depended on the moderator temperature, the thermal feedback depends on both temperatures.
Hence, a mesh node should hold both temperatures.
Section \ref{sec:ph1ex3} used a heterogeneous thermal-fluid model to solve the temperature in the reactor.
The problem with using a heterogeneous thermal-fluid model in a coupled simulation is that each node only holds the temperature's value only in that particular material.
For example, a coolant node holds the coolant temperature information and computes the thermal feedback with such information instead of the moderator and fuel temperatures.
Computing the thermal feedback with the coolant temperature is essentially wrong, as the group constants do not depend on the coolant temperature.
For this reason, Moltres should use the average assembly-level fuel and moderator temperatures instead of the point-wise temperature to compute the thermal feedback.
% Moltres should differentiate Tmod and Tfuel
The thermal-fluids model in Section \ref{sec:ph1ex3} homogenized the fuel and the moderator into one material.
Such homogenization assumes that both the moderator and fuel are in thermal equilibrium, and therefore have the same temperature.
Consequently, the model does not differentiate the fuel temperature from the moderator temperature.
However, a coupling model cannot correctly calculate the thermal feedback if it does not differentiate between the moderator and the fuel temperatures \cite{damian_vhtr_2008}.
This thesis denominates this issue as the 'homogenization dilemma.'
To correctly compute neutronics, the diffusion solver must use homogenized parameters that depend on both moderator and fuel temperatures.
Simultaneously, to accurately calculate the thermal feedback, a mesh node should differentiate the moderator temperature from the fuel temperature, which requires a heterogeneous calculation in the thermal-fluids model.
A thermal-fluid heterogeneous calculation is still valid, but the thermal feedback should use fuel and moderator average temperatures instead.
\subsection{Coupled exercise with average temperature thermal feedback}
\label{sec:coupled-average}
The previous discussion led to the development of this section.
This analysis used the solid average temperatures for calculating the thermal feedback.
This section compares the results to Section \ref{sec:ph1ex3}'s results, which used the point-wise temperatures for calculating the thermal feedback.
Figure \ref{fig:coupled-ave-results-th2} exhibits the point-wise temperature and the average temperatures across the reactor.
The model used the average temperatures to compute the thermal feedback in the core.
The simulations did not compute the average temperature in the RPV and the outside air.
Figure \ref{fig:coupled-ave-results-flux} displays the neutron flux profiles, and Figure \ref{fig:coupled-ave-results-th}, the temperature profiles.
The results are almost identical.
The only noticeable difference is in the axial temperature profile, which differs by less than 3$^{\circ}$C in the outlet.
The radial temperature profile is not affected.
This analysis expected more variations in the results.
The point-wise approach arrived at similar results because the difference between the solid and coolant temperature is small (less than 100$^{\circ}$C).
However, in transient analyses, the temperature difference between the fuel and the coolant may increase, causing more variability in the results.
% Although using the point-wise approach is easier for the input file definition,
Further studies should analyze this behavior.
\begin{figure}[htbp!]
\centering
\subfloat[Temperature profile.]{
\includegraphics[width=0.40\textwidth]{figures-thermal/coupledH-temp}
}
\subfloat[Average temperature in each assembly.]{
\includegraphics[width=0.40\textwidth]{figures-thermal/coupledH-auxtemp}
}
\hfill
\caption{Temperature distribution.}
\label{fig:coupled-ave-results-th2}
\end{figure}
\begin{figure}[htbp!]
\centering
\subfloat[Axial flux over the line at r=85 cm.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/coupledD-H-axial}
}
\subfloat[Radial temperature at z=L/2=556.5 cm.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/coupledD-H-radial}
}
\hfill
\caption{Flux profile comparison. AVE: uses average temperature, P-W: uses point-wise temperature.}
\label{fig:coupled-ave-results-flux}
\end{figure}
\begin{figure}[htbp!]
\centering
\subfloat[Axial temperature at r=85 cm.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/coupledD-H-axialT}
}
\subfloat[Radial temperature at z=L/2=556.5 cm.]{
\includegraphics[width=0.45\textwidth]{figures-thermal/coupledD-H-radialT}
}
\hfill
\caption{Temperature profile comparison.}
\label{fig:coupled-ave-results-th}
\end{figure}
\section{Conclusions}
\label{sec:thermal-conc}
This chapter addressed several objectives of this thesis by presenting the results of several stand-alone thermal-fluids simulations of prismatic HTGRs using Moltres.
The preliminary studies focused on the verification of the thermal-fluids model and the validation of the unit-cell model.
The verification of the thermal-fluids model studied a simplified cylindrical model comparing the Moltres/MOOSE numerical solution to the problem analytical solution - both solutions were in agreement.
To validate the unit cell model, Section \ref{sec:unitcell} compared Moltres/MOOSE results to the results from a published article.
Moltres/MOOSE results presented some discrepancies, being within 2\% of the reference results.
Section \ref{sec:fuelcol} focus of the analysis was a $1/12^{th}$ section of an HTGR fuel column.
To validate the model, Section \ref{sec:fuelcol} analysis reproduced a published article's results.
The analyses comprised two studies: one with no bypass-gap, and one with a 3mm-bypass-gap.
To simplify the analysis, the model adopted the mass flow rates from Sato's article.
For both case studies, the coolant and the fuel's maximum temperatures were within a 1\% difference to the reference temperatures.
Section \ref{sec:fuelcol} also presented the temperature profile in two of the edges of the geometry, exhibiting the effects of the bypass flow on the temperature distribution.
The bypass-gap makes the center temperature rise while reducing the peripheral temperature, overall, increasing the temperature gradient inside the column.
The next analysis studied different calculation methods for the mass flow distribution in the fuel column.
Section \ref{sec:flowdistrib} adopted known mass flow distribution values as a reference and calculated the mass flow distribution using four different methods.
From Cases 1 to 4, the methods' complexity increased as some required iterative solvers.
Overall, Case 1 proved to be the simplest method, and its application did not considerably deteriorate the results' accuracy.
For that reason, the following studies adopted the Case 1 mass flow distribution method.
Section \ref{sec:meshconverge} studied the feasibility of extending this methodology to larger meshes and conduct full-scale simulations.
Section \ref{sec:meshconverge} studied the mesh convergence of the full-fuel column problem.
As the model uses the one-dimensional coolant equations, the coolant temperature converges relatively fast compared to the fuel temperature.
The fuel temperature did not reach convergence in this analysis.
As the mesh discretization increased, it imposed a high memory requirement on the simulations.
This analysis concluded that modeling thermal-fluids with such a detailed level is computationally too expensive, and it suggests searching for other approaches.
The following sections adopted a two-dimensional cylindrical model for the full-core analyses.
Section \ref{sec:ph1ex2} described Phase I Exercise 2 of the OECD/NEA MHTGR-350 Benchmark.
This exercise encompasses four sub-cases with different definitions of bypass-flow and material properties.
The simplest sub-case is Exercise 2a, as it does not model the bypass flow and defines the material properties.
Section \ref{sec:ph1ex2} used Moltres/MOOSE to conduct Exercise 2a.
As OECD/NEA did not publish this exercise’s results, Section \ref{sec:ph1ex2} compared Moltres/MOOSE results to INL benchmark results \cite{strydom_inl_2013}.
Moltres' thermal-fluids model based its definition on INL's model.
The large discrepancies between Moltres/MOOSE and INL results suggest that our model does not capture some heat transfer mechanism that the INL model does.
One of the known differences between the models is the inclusion of the radiative heat transfer mechanism.
As Moltres does not model that mechanism, it could be the cause of the discrepancies.
A second study in Section \ref{sec:ph1ex2} used a two-level approach, based on a global and a unit cell model \cite{stainsby_investigation_2008}.
The global model calculated the coolant temperature, while the unit cell model focused on the moderator and fuel temperatures.
With this approach, Moltres/MOOSE results were closer to INL results.
The results showed some discrepancies, which indicates that some assumption/model simplification is not accurate enough.
A flat velocity approximation might be the cause of such discrepancies.
Further studies should analyze their origin.
Section \ref{sec:ph1ex3} developed a global model to solve a simplified version of Phase 1 Exercise 3 of the OECD/NEA benchmark.
Although the model was simple, it allowed visualizing some of the essential aspects of a prismatic HTGR multi-physics simulation in Moltres.
This exercise led to Section \ref{sec:discuss}, which described some of the limitations found in the model.
Additionally, Section \ref{sec:coupled-average} addressed some of the limitations found in the model and set the basis for conducting prismatic HTGR coupled exercises in Moltres.
This chapter met the objective of calculating accurate temperature distributions in prismatic HTGRs; however, it achieved it partially.
This work proved Moltres' capability to conduct steady-state thermal-fluid calculations in local models such as the unit cell and the fuel column problems.
On the other hand, the full-core results showed several inconsistencies that will require further analysis.
This chapter arrives at a similar conclusion for the objective of conducting coupled simulation.
Nevertheless, an inaccurate thermal-fluids calculation yields inaccurate coupled simulations.
For this reason, Section \ref{sec:ph1ex3} carried out a simplified exercise with the objective of establishing the roadmap for future simulations.