-
Notifications
You must be signed in to change notification settings - Fork 0
/
reference_old.bib
1637 lines (1537 loc) · 173 KB
/
reference_old.bib
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
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
@article{mcphee_dynamics_1987,
title = {Dynamics and {Thermodynamics} of the {Ice}/{Upper} {Ocean} {System} in the {Marginal} {Ice} {Zone} of the {Greenland} {Sea}},
volume = {92},
url = {http://www.agu.org/pubs/crossref/1987/JC092iC07p07017.shtml},
doi = {10.1029/JC092iC07p07017},
number = {C7},
journal = {Journal of Geophysical Research},
author = {McPhee, Miles G. and Maykut, Gary A. and Morison, James H.},
year = {1987},
pages = {7017--7031},
file = {McPhee et al. - 1987 - Dynamics and Thermodynamics of the Ice-Upper Ocean System in the Marginal Ice.pdf:/Users/cbegeman/Zotero/storage/PIJ4YHNS/McPhee et al. - 1987 - Dynamics and Thermodynamics of the Ice-Upper Ocean System in the Marginal Ice.pdf:application/pdf},
}
@article{mcphee_analytic_1981,
title = {An analytic similarity theory for the planetary boundary layer stabilized by surface buoyancy},
volume = {21},
issn = {0006-8314},
url = {http://www.springerlink.com/index/10.1007/BF00119277},
doi = {10.1007/BF00119277},
number = {3},
journal = {Boundary-Layer Meteorology},
author = {McPhee, Miles G.},
month = nov,
year = {1981},
keywords = {sea ice},
pages = {325--339},
file = {Attachment:/Users/cbegeman/Zotero/storage/8MJMWPJC/McPhee - 1981 - An analytic similarity theory for the planetary boundary layer stabilized by surface buoyancy.pdf:application/pdf},
}
@article{mcphee_revisiting_2008,
title = {Revisiting heat and salt exchange at the ice-ocean interface: {Ocean} flux and modeling considerations},
volume = {113},
issn = {0148-0227},
url = {http://www.agu.org/pubs/crossref/2008/2007JC004383.shtml},
doi = {10.1029/2007JC004383},
number = {C6},
journal = {Journal of Geophysical Research},
author = {McPhee, M. G. and Morison, J. H. and Nilsen, F.},
month = jun,
year = {2008},
keywords = {modeling, ice-ocean interactions, sea ice},
pages = {1--10},
file = {McPhee et al. - 2008 - Revisiting heat and salt exchange at the ice-ocean interface.pdf:/Users/cbegeman/Zotero/storage/EJBAELMT/McPhee et al. - 2008 - Revisiting heat and salt exchange at the ice-ocean interface.pdf:application/pdf},
}
@article{vreugdenhil_stratification_2019,
title = {Stratification effects in the turbulent boundary layer beneath a melting ice shelf: insights from resolved large-eddy simulations},
issn = {0022-3670},
shorttitle = {Stratification effects in the turbulent boundary layer beneath a melting ice shelf},
url = {https://journals.ametsoc.org/doi/abs/10.1175/JPO-D-18-0252.1},
doi = {10.1175/JPO-D-18-0252.1},
abstract = {Ocean turbulence contributes to the basal melting and dissolution of ice shelves by transporting heat and salt towards the ice. The meltwater causes a stable salinity stratification to form beneath the ice that suppresses turbulence. Here we use large-eddy simulations motivated by the ice-shelf/ocean boundary layer (ISOBL) to examine the inherently linked processes of turbulence and stratification, and their influence on the melt rate. Our rectangular domain is bounded from above by the ice base where a dynamic melt condition is imposed. By varying the speed of the flow and the ambient temperature, we identify a fully turbulent, well-mixed regime and an intermittently turbulent, strongly stratified regime. The transition between regimes can be characterised by comparing the Obukhov length, which provides a measure of the distance away from the ice base where stratification begins to dominate the flow, to the viscous length scale of the interfacial sublayer. Upper limits on simulated turbulent transfer coefficients are used to predict the transition from fully to intermittently turbulent flow. The predicted melt rate is sensitive to the choice of the heat and salt transfer coefficients and the drag coefficient. For example, when coefficients characteristic of fully-developed turbulence are applied to intermittent flow, the parameterized three-equation model overestimates the basal melt rate by almost a factor of ten. These insights may help to guide when existing parameterisations of ice melt are appropriate for use in regional or large-scale ocean models, and may also have implications for other ice-ocean interactions such as fast ice or drifting ice.},
urldate = {2019-05-09},
journal = {Journal of Physical Oceanography},
author = {Vreugdenhil, Catherine A. and Taylor, John R.},
month = may,
year = {2019},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/UNXSK9SP/Vreugdenhil and Taylor - 2019 - Stratification effects in the turbulent boundary l.pdf:application/pdf;Vreugdenhil Taylor - 2019 - Stratification effects in the turbulent boundary layer beneath a melting ice.pdf:/Users/cbegeman/Zotero/storage/6GQ65U3Z/Vreugdenhil Taylor - 2019 - Stratification effects in the turbulent boundary layer beneath a melting ice.pdf:application/pdf},
}
@article{mcconnochie_testing_2017,
title = {Testing a common ice-ocean parameterization with laboratory experiments},
volume = {122},
copyright = {© 2017. American Geophysical Union. All Rights Reserved.},
issn = {2169-9291},
url = {https://agupubs.onlinelibrary.wiley.com/doi/abs/10.1002/2017JC012918},
doi = {10.1002/2017JC012918},
abstract = {Numerical models of ice-ocean interactions typically rely upon a parameterization for the transport of heat and salt to the ice face that has not been satisfactorily validated by observational or experimental data. We compare laboratory experiments of ice-saltwater interactions to a common numerical parameterization and find a significant disagreement in the dependence of the melt rate on the fluid velocity. We suggest a resolution to this disagreement based on a theoretical analysis of the boundary layer next to a vertical heated plate, which results in a threshold fluid velocity of approximately 4 cm/s at driving temperatures between 0.5 and C, above which the form of the parameterization should be valid.},
language = {en},
number = {7},
urldate = {2019-10-18},
journal = {Journal of Geophysical Research: Oceans},
author = {McConnochie, C. D. and Kerr, R. C.},
year = {2017},
keywords = {ice-ocean interactions, parameterization, laboratory experiments},
pages = {5905--5915},
file = {McConnochie_Kerr - 2017 - Testing a common ice-ocean parameterization with laboratory experiments.pdf:/Users/cbegeman/Zotero/storage/Z4QL269L/McConnochie_Kerr - 2017 - Testing a common ice-ocean parameterization with laboratory experiments.pdf:application/pdf},
}
@article{ramudu_large_2018,
title = {Large {Eddy} {Simulation} of {Heat} {Entrainment} {Under} {Arctic} {Sea} {Ice}},
volume = {123},
issn = {2169-9291},
url = {https://agupubs.onlinelibrary.wiley.com/doi/abs/10.1002/2017JC013267},
doi = {10.1002/2017JC013267},
abstract = {Arctic sea ice has declined rapidly in recent decades. The faster than projected retreat suggests that free-running large-scale climate models may not be accurately representing some key processes. The small-scale turbulent entrainment of heat from the mixed layer could be one such process. To better understand this mechanism, we model the Arctic Ocean's Canada Basin, which is characterized by a perennial anomalously warm Pacific Summer Water (PSW) layer residing at the base of the mixed layer and a summertime Near-Surface Temperature Maximum (NSTM) within the mixed layer trapping heat from solar radiation. We use large eddy simulation (LES) to investigate heat entrainment for different ice-drift velocities and different initial temperature profiles. The value of LES is that the resolved turbulent fluxes are greater than the subgrid-scale fluxes for most of our parameter space. The results show that the presence of the NSTM enhances heat entrainment from the mixed layer. Additionally there is no PSW heat entrained under the parameter space considered. We propose a scaling law for the ocean-to-ice heat flux which depends on the initial temperature anomaly in the NSTM layer and the ice-drift velocity. A case study of “The Great Arctic Cyclone of 2012” gives a turbulent heat flux from the mixed layer that is approximately 70\% of the total ocean-to-ice heat flux estimated from the PIOMAS model often used for short-term predictions. Present results highlight the need for large-scale climate models to account for the NSTM layer.},
language = {en},
number = {1},
urldate = {2018-09-17},
journal = {Journal of Geophysical Research: Oceans},
author = {Ramudu, Eshwan and Gelderloos, Renske and Yang, Di and Meneveau, Charles and Gnanadesikan, Anand},
month = jan,
year = {2018},
keywords = {ice-ocean interaction, large eddy simulation, Near-Surface Temperature Maximum, Pacific Summer Water, Ocean / Arctic / mixed layer, Ocean / Heat entrainment},
pages = {287--304},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/DM5CJH8C/Ramudu et al. - 2018 - Large Eddy Simulation of Heat Entrainment Under Ar.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/EVLGEPUG/2017JC013267.html:text/html},
}
@article{wells_geophysical-scale_2008,
title = {A geophysical-scale model of vertical natural convection boundary layers},
volume = {609},
issn = {0022-1120},
url = {http://www.journals.cambridge.org/abstract_S0022112008002346},
doi = {10.1017/S0022112008002346},
journal = {Journal of Fluid Mechanics},
author = {Wells, Andrew J. and Worster, M. Grae},
month = jul,
year = {2008},
keywords = {modeling, general climate},
pages = {111--137},
file = {Wells_Worster - 2008 - A geophysical-scale model of vertical natural convection boundary layers.pdf:/Users/cbegeman/Zotero/storage/22E6L6ZS/Wells_Worster - 2008 - A geophysical-scale model of vertical natural convection boundary layers.pdf:application/pdf},
}
@article{maronga_overview_2020,
title = {Overview of the {PALM} model system 6.0},
volume = {13},
issn = {1991-959X},
url = {https://www.geosci-model-dev.net/13/1335/2020/},
doi = {https://doi.org/10.5194/gmd-13-1335-2020},
abstract = {{\textless}p{\textgreater}{\textless}strong{\textgreater}Abstract.{\textless}/strong{\textgreater} In this paper, we describe the PALM model system 6.0. PALM (formerly an abbreviation for Parallelized Large-eddy Simulation Model and now an independent name) is a Fortran-based code and has been applied for studying a variety of atmospheric and oceanic boundary layers for about 20 years. The model is optimized for use on massively parallel computer architectures. This is a follow-up paper to the PALM 4.0 model description in {\textless}span class="cit" id="xref\_text.1"{\textgreater}{\textless}a href="\#bib1.bibx73"{\textgreater}Maronga et al.{\textless}/a{\textgreater} ({\textless}a href="\#bib1.bibx73"{\textgreater}2015{\textless}/a{\textgreater}){\textless}/span{\textgreater}. During the last years, PALM has been significantly improved and now offers a variety of new components. In particular, much effort was made to enhance the model with components needed for applications in urban environments, like fully interactive land surface and radiation schemes, chemistry, and an indoor model. This paper serves as an overview paper of the PALM 6.0 model system and we describe its current model core. The individual components for urban applications, case studies, validation runs, and issues with suitable input data are presented and discussed in a series of companion papers in this special issue.{\textless}/p{\textgreater}},
language = {English},
number = {3},
urldate = {2020-03-27},
journal = {Geoscientific Model Development},
author = {Maronga, Björn and Banzhaf, Sabine and Burmeister, Cornelia and Esch, Thomas and Forkel, Renate and Fröhlich, Dominik and Fuka, Vladimir and Gehrke, Katrin Frieda and Geletič, Jan and Giersch, Sebastian and Gronemeier, Tobias and Groß, Günter and Heldens, Wieke and Hellsten, Antti and Hoffmann, Fabian and Inagaki, Atsushi and Kadasch, Eckhard and Kanani-Sühring, Farah and Ketelsen, Klaus and Khan, Basit Ali and Knigge, Christoph and Knoop, Helge and Krč, Pavel and Kurppa, Mona and Maamari, Halim and Matzarakis, Andreas and Mauder, Matthias and Pallasch, Matthias and Pavlik, Dirk and Pfafferott, Jens and Resler, Jaroslav and Rissmann, Sascha and Russo, Emmanuele and Salim, Mohamed and Schrempf, Michael and Schwenkel, Johannes and Seckmeyer, Gunther and Schubert, Sebastian and Sühring, Matthias and Tils, Robert von and Vollmer, Lukas and Ward, Simon and Witha, Björn and Wurps, Hauke and Zeidler, Julian and Raasch, Siegfried},
month = mar,
year = {2020},
note = {Publisher: Copernicus GmbH},
pages = {1335--1372},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/XQZUF3Z8/Maronga et al. - 2020 - Overview of the PALM model system 6.0.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/7ZHMKNFB/2020.html:text/html},
}
@article{davis_turbulence_2019,
title = {Turbulence {Observations} beneath {Larsen} {C} {Ice} {Shelf}, {Antarctica}},
volume = {0},
copyright = {This article is protected by copyright. All rights reserved.},
issn = {2169-9291},
url = {https://agupubs.onlinelibrary.wiley.com/doi/abs/10.1029/2019JC015164},
doi = {10.1029/2019JC015164},
abstract = {Increased ocean-driven basal melting beneath Antarctic ice shelves causes grounded ice to flow into the ocean at an accelerated rate, with consequences for global sea level. The turbulent transfer of heat through the ice shelf-ocean boundary layer is critical in setting the basal melt rate, yet the processes controlling this transfer are poorly understood and inadequately represented in global climate models. This creates large uncertainties in predictions of future sea-level rise. Using a hot-water drilled access hole, two turbulence instrument clusters (TICs) were deployed 2.5 and 13.5 meters beneath Larsen C Ice Shelf in December 2011. Both instruments returned a year-long record of turbulent velocity fluctuations, providing a unique opportunity to explore the turbulent processes within the ice shelf-ocean boundary layer. Although the scaling between the turbulent kinetic energy (TKE) dissipation rate and mean flow speed varies with distance from the ice shelf base, at both TICs the TKE dissipation rate is balanced entirely by the rate of shear production. The freshwater released by basal melting plays no role in the TKE balance. When the upper TIC is within the log-layer, we derive an under-ice drag coefficient of 0.0022 and a roughness length of 0.44 mm, indicating that the ice base is smooth. Finally, we demonstrate that although the canonical three-equation melt rate parameterization can accurately predict the melt rate for this example of smooth ice underlain by a cold, tidally-forced boundary layer, the law of the wall assumption employed by the parameterization does not hold at low flow speeds.},
language = {en},
number = {ja},
urldate = {2019-07-17},
journal = {Journal of Geophysical Research: Oceans},
author = {Davis, Peter E. D. and Nicholls, Keith W.},
year = {2019},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/MT32TRMS/Davis and Nicholls - Turbulence Observations beneath Larsen C Ice Shelf.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/9HEFARRB/2019JC015164.html:text/html},
}
@article{arya_buoyancy_1975,
title = {Buoyancy effects in a horizontal flat-plate boundary layer},
volume = {68},
issn = {1469-7645, 0022-1120},
url = {https://www.cambridge.org/core/journals/journal-of-fluid-mechanics/article/abs/buoyancy-effects-in-a-horizontal-flatplate-boundary-layer/1212C3130504795ADF39FED9FDB0AC40},
doi = {10.1017/S0022112075000833},
abstract = {Observations made in a well-developed, thermally stratified, horizontal, flat- plate boundary layer are used to study the effects of buoyancy on the mean flow and turbulence structure. These are represented in a similarity framework obtained from the concept of local equilibrium in a fully developed turbulent flow. Mean velocity and temperature profiles in both the inner and outer layers are strongly dependent on the thermal stratification, the former suggesting an increase in the thickness of the viscous sublayer with increasing stability. The coefficients of skin friction and heat transfer, on the other hand, decrease with increasing stability.Normalized turbulent intensities, fluxes and their correlation coefficients also vary with buoyancy. In stable conditions, turbulence becomes rapidly suppressed with increasing stability as more and more energy has to be expended in over- coming buoyancy forces. The buoyancy effects are found to be more dominant in the stress budget than in the turbulent energy budget. The horizontal heat flux is much greater than the vertical heat flux and their ratio increases with stability. The ratio of the eddy diffusivities of heat and momentum, on the other hand, decreases with increasing stability. The spectra of velocity and temperature fluctuations indicate no buoyancy subrange, but the wavenumber corresponding to peak energy is found to increase with increasing stability.},
language = {en},
number = {2},
urldate = {2021-05-13},
journal = {Journal of Fluid Mechanics},
author = {Arya, S. P. S.},
month = mar,
year = {1975},
note = {Publisher: Cambridge University Press},
pages = {321--343},
file = {Snapshot:/Users/cbegeman/Zotero/storage/9FNV76RD/1212C3130504795ADF39FED9FDB0AC40.html:text/html},
}
@article{raasch_palm-large-eddy_2001,
title = {{PALM}-{A} large-eddy simulation model performing on massively parallel computers},
volume = {10},
doi = {10.1127/0941-2948/2001/0010-0363},
abstract = {An existing code of a large-eddy simulation (LES) model for the study of turbulent processes in the atmo-spheric and oceanic boundary layer has been completely recoded for use on massively parallel systems with distributed memory. Parallelization is achieved by two-dimensional domain decomposition and communica-tion is realized by the message passing interface (MPI). Periodic boundary conditions, which are used in both horizontal directions, helped to minimize the parallelization effort. The performance of the new PArallelized LES Model (PALM) is excellent on SGI/Cray-T3E systems and an almost linear speed-up is achieved up to very large numbers of processors. Parallelization strategy and model performance is discussed and validation experiments as well as future applications are presented. Zusammenfassung Für Untersuchungen turbulenter Prozesse in der atmosphärischen und ozeanischen Grenzschicht wurde ein bereits existierender LES-Code zur Nutzung auf massiv parallelen Systemen mit verteiltem Speicher vollständig neu implementiert. Die Parallelisierung geschieht mittels zweidimensionaler Gebietszerlegung, wobei die Kommunikation zwischen den Prozessoren durch das Message Passing Interface (MPI) reali-siert ist. Die im Modell in beiden horizontalen Richtungen verwendeten periodischen Randbedingunge n tra-gen dazu bei, den Parallelisierungsaufwand zu minimieren. Auf SGI/Cray-T3E Systemen skaliert das neue PArallelisierte LES Modell (PALM) hervorragend: selbst bis zu einer großen Anzahl verwendeter Prozes-soren wird ein nahezu linearer Speed-Up erreicht. In dieser Arbeit werden neben der Parallelisierungsstrategie sowohl die Ergebnisse von Skalierungstests und Validierungsrechnungen präsentiert und diskutiert, als auch zukünftige Anwendungen des Modells aufgezeigt.},
journal = {Meteorol. Z.},
author = {Raasch, Siegfried and Sch, Michael},
month = nov,
year = {2001},
pages = {363--372},
}
@article{maronga_parallelized_2015,
title = {The {Parallelized} {Large}-{Eddy} {Simulation} {Model} ({PALM}) version 4.0 for atmospheric and oceanic flows: model formulation, recent developments, and future perspectives},
volume = {8},
issn = {1991-9603},
shorttitle = {The {Parallelized} {Large}-{Eddy} {Simulation} {Model} ({PALM}) version 4.0 for atmospheric and oceanic flows},
url = {https://www.geosci-model-dev.net/8/2515/2015/},
doi = {10.5194/gmd-8-2515-2015},
abstract = {In this paper we present the current version of the Parallelized Large-Eddy Simulation Model (PALM) whose core has been developed at the Institute of Meteorology and Climatology at Leibniz Universität Hannover (Germany). PALM is a Fortran 95-based code with some Fortran 2003 extensions and has been applied for the simulation of a variety of atmospheric and oceanic boundary layers for more than 15 years. PALM is optimized for use on massively parallel computer architectures and was recently ported to general-purpose graphics processing units. In the present paper we give a detailed description of the current version of the model and its features, such as an embedded Lagrangian cloud model and the possibility to use Cartesian topography. Moreover, we discuss recent model developments and future perspectives for LES applications.},
language = {en},
number = {8},
urldate = {2018-08-23},
journal = {Geoscientific Model Development},
author = {Maronga, B. and Gryschka, M. and Heinze, R. and Hoffmann, F. and Kanani-Sühring, F. and Keck, M. and Ketelsen, K. and Letzel, M. O. and Sühring, M. and Raasch, S.},
month = aug,
year = {2015},
pages = {2515--2551},
}
@article{rozema_minimum-dissipation_2015,
title = {Minimum-dissipation models for large-eddy simulation},
volume = {27},
issn = {1070-6631},
url = {https://aip.scitation.org/doi/full/10.1063/1.4928700},
doi = {10.1063/1.4928700},
abstract = {Minimum-dissipation eddy-viscosity models are a class of sub-filter models for large-eddy simulation that give the minimum eddy dissipation required to dissipate the energy of sub-filter scales. A previously derived minimum-dissipation model is the QR model. This model is based on the invariants of the resolved rate-of-strain tensor and has many desirable properties. It appropriately switches off for laminar and transitional flows, has low computational complexity, and is consistent with the exact sub-filter tensor on isotropic grids. However, the QR model proposed in the literature gives insufficient eddy dissipation. It is demonstrated that this can be corrected by increasing the model constant. The corrected QR model gives good results in simulations of decaying grid turbulence on an isotropic grid. On anisotropic grids the QR model is not consistent with the exact sub-filter tensor and requires an approximation of the filter width. It is demonstrated that the results of the QR model on anisotropic grids are primarily determined by the used filter width approximation, and that no approximation gives satisfactory results in simulations of both a temporal mixing layer and turbulent channel flow. A new minimum-dissipation model for anisotropic grids is proposed. This anisotropic minimum-dissipation (AMD) model generalizes the desirable practical and theoretical properties of the QR model to anisotropic grids and does not require an approximation of the filter width. The AMD model is successfully applied in simulations of decaying grid turbulence on an isotropic grid and in simulations of a temporal mixing layer and turbulent channel flow on anisotropic grids.},
number = {8},
journal = {Physics of Fluids},
author = {Rozema, Wybe and Bae, Hyun J. and Moin, Parviz and Verstappen, Roel},
year = {2015},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/BMCDGG7I/Rozema et al. - 2015 - Minimum-dissipation models for large-eddy simulati.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/K5CD878N/1.html:text/html},
}
@article{abkar_minimum-dissipation_2016,
title = {Minimum-dissipation scalar transport model for large-eddy simulation of turbulent flows},
volume = {1},
url = {https://link.aps.org/doi/10.1103/PhysRevFluids.1.041701},
doi = {10.1103/PhysRevFluids.1.041701},
abstract = {Minimum-dissipation models are a simple alternative to the Smagorinsky-type approaches to parametrize the subfilter turbulent fluxes in large-eddy simulation. A recently derived model of this type for subfilter stress tensor is the anisotropic minimum-dissipation (AMD) model [Rozema et al., Phys. Fluids 27, 085107 (2015)], which has many desirable properties. It is more cost effective than the dynamic Smagorinsky model, it appropriately switches off in laminar and transitional flows, and it is consistent with the exact subfilter stress tensor on both isotropic and anisotropic grids. In this study, an extension of this approach to modeling the subfilter scalar flux is proposed. The performance of the AMD model is tested in the simulation of a high-Reynolds-number rough-wall boundary-layer flow with a constant and uniform surface scalar flux. The simulation results obtained from the AMD model show good agreement with well-established empirical correlations and theoretical predictions of the resolved flow statistics. In particular, the AMD model is capable of accurately predicting the expected surface-layer similarity profiles and power spectra for both velocity and scalar concentration.},
number = {4},
urldate = {2020-01-10},
journal = {Physical Review Fluids},
author = {Abkar, Mahdi and Bae, Hyun J. and Moin, Parviz},
month = aug,
year = {2016},
pages = {041701},
file = {APS Snapshot:/Users/cbegeman/Zotero/storage/3HPTTJ2D/PhysRevFluids.1.html:text/html;Full Text PDF:/Users/cbegeman/Zotero/storage/C9HHN7LU/Abkar et al. - 2016 - Minimum-dissipation scalar transport model for lar.pdf:application/pdf},
}
@article{abkar_large-eddy_2017,
title = {Large-{Eddy} {Simulation} of {Thermally} {Stratified} {Atmospheric} {Boundary}-{Layer} {Flow} {Using} a {Minimum} {Dissipation} {Model}},
volume = {165},
issn = {1573-1472},
url = {https://doi.org/10.1007/s10546-017-0288-4},
doi = {10.1007/s10546-017-0288-4},
abstract = {A generalized form of a recently developed minimum dissipation model for subfilter turbulent fluxes is proposed and implemented in the simulation of thermally stratified atmospheric boundary-layer flows. Compared with the original model, the generalized model includes the contribution of buoyant forces, in addition to shear, to the production or suppression of turbulence, with a number of desirable practical and theoretical properties. Specifically, the model has a low computational complexity, appropriately switches off in laminar and transitional flows, does not require any ad hoc shear and stability corrections, and is consistent with theoretical subfilter turbulent fluxes. The simulation results show remarkable agreement with well-established empirical correlations, theoretical predictions, and field observations in the atmosphere. In addition, the results show very little sensitivity to the grid resolution, demonstrating the robustness of the model in the simulation of the atmospheric boundary layer, even with relatively coarse resolutions.},
language = {en},
number = {3},
urldate = {2020-05-13},
journal = {Boundary-Layer Meteorology},
author = {Abkar, Mahdi and Moin, Parviz},
month = dec,
year = {2017},
pages = {405--419},
file = {Springer Full Text PDF:/Users/cbegeman/Zotero/storage/AZ3VSWQ4/Abkar and Moin - 2017 - Large-Eddy Simulation of Thermally Stratified Atmo.pdf:application/pdf},
}
@article{asay-davis_experimental_2016,
title = {Experimental design for three interrelated marine ice sheet and ocean model intercomparison projects: {MISMIP} v. 3 ({MISMIP} +), {ISOMIP} v. 2 ({ISOMIP} +) and {MISOMIP} v. 1 ({MISOMIP1})},
volume = {9},
issn = {1991-9603},
shorttitle = {Experimental design for three interrelated marine ice sheet and ocean model intercomparison projects},
url = {http://www.geosci-model-dev.net/9/2471/2016/},
doi = {10.5194/gmd-9-2471-2016},
abstract = {Coupled ice sheet–ocean models capable of simulating moving grounding lines are just becoming available. Such models have a broad range of potential applications in studying the dynamics of marine ice sheets and tidewater glaciers, from process studies to future projections of ice mass loss and sea level rise. The Marine Ice Sheet–Ocean Model Intercomparison Project (MISOMIP) is a community effort aimed at designing and coordinating a series of model intercomparison projects (MIPs) for model evaluation in idealized setups, model verification based on observations, and future projections for key regions of the West Antarctic Ice Sheet (WAIS). Here we describe computational experiments constituting three interrelated MIPs for marine ice sheet models and regional ocean circulation models incorporating ice shelf cavities. These consist of ice sheet experiments under the Marine Ice Sheet MIP third phase (MISMIP+), ocean experiments under the Ice Shelf-Ocean MIP second phase (ISOMIP+) and coupled ice sheet–ocean experiments under the MISOMIP first phase (MISOMIP1). All three MIPs use a shared domain with idealized bedrock topography and forcing, allowing the coupled simulations (MISOMIP1) to be compared directly to the individual component simulations (MISMIP+ and ISOMIP+). The experiments, which have qualitative similarities to Pine Island Glacier Ice Shelf and the adjacent region of the Amundsen Sea, are designed to explore the effects of changes in ocean conditions, specifically the temperature at depth, on basal melting and ice dynamics. In future work, differences between model results will form the basis for the evaluation of the participating models.},
number = {7},
urldate = {2017-05-15},
journal = {Geosci. Model Dev.},
author = {Asay-Davis, X. S. and Cornford, S. L. and Durand, G. and Galton-Fenzi, B. K. and Gladstone, R. M. and Gudmundsson, G. H. and Hattermann, Tore and Holland, D. M. and Holland, D. and Holland, P. R. and Martin, D. F. and Mathiot, P. and Pattyn, F. and Seroussi, H.},
month = jul,
year = {2016},
note = {00003},
pages = {2471--2497},
file = {Geosci. Model Dev. PDF:/Users/cbegeman/Zotero/storage/RMNZ3WBG/Asay-Davis et al. - 2016 - Experimental design for three interrelated marine .pdf:application/pdf},
}
@article{jackett_algorithms_2006,
title = {Algorithms for {Density}, {Potential} {Temperature}, {Conservative} {Temperature}, and the {Freezing} {Temperature} of {Seawater}},
volume = {23},
issn = {0739-0572},
url = {https://journals.ametsoc.org/doi/full/10.1175/JTECH1946.1},
doi = {10.1175/JTECH1946.1},
abstract = {Algorithms are presented for density, potential temperature, conservative temperature, and the freezing temperature of seawater. The algorithms for potential temperature and density (in terms of potential temperature) are updates to routines recently published by McDougall et al., while the algorithms involving conservative temperature and the freezing temperatures of seawater are new. The McDougall et al. algorithms were based on the thermodynamic potential of Feistel and Hagen; the algorithms in this study are all based on the “new extended Gibbs thermodynamic potential of seawater” of Feistel. The algorithm for the computation of density in terms of salinity, pressure, and conservative temperature produces errors in density and in the corresponding thermal expansion coefficient of the same order as errors for the density equation using potential temperature, both being twice as accurate as the International Equation of State when compared with Feistel’s new equation of state. An inverse function relating potential temperature to conservative temperature is also provided. The difference between practical salinity and absolute salinity is discussed, and it is shown that the present practice of essentially ignoring the difference between these two different salinities is unlikely to cause significant errors in ocean models.},
number = {12},
urldate = {2019-04-24},
journal = {Journal of Atmospheric and Oceanic Technology},
author = {Jackett, David R. and McDougall, Trevor J. and Feistel, Rainer and Wright, Daniel G. and Griffies, Stephen M.},
month = dec,
year = {2006},
pages = {1709--1728},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/F7MIQ8RL/Jackett et al. - 2006 - Algorithms for Density, Potential Temperature, Con.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/QV2N5FCF/JTECH1946.html:text/html},
}
@article{holland_modeling_1999,
title = {Modeling {Thermodynamic} {Ice}–{Ocean} {Interactions} at the {Base} of an {Ice} {Shelf}},
volume = {29},
issn = {0022-3670},
url = {http://journals.ametsoc.org.oca.ucsc.edu/doi/abs/10.1175/1520-0485(1999)029%3C1787:MTIOIA%3E2.0.CO%3B2},
doi = {10.1175/1520-0485(1999)029<1787:MTIOIA>2.0.CO;2},
abstract = {Abstract Models of ocean circulation beneath ice shelves are driven primarily by the heat and freshwater fluxes that are associated with phase changes at the ice–ocean boundary. Their behavior is therefore closely linked to the mathematical description of the interaction between ice and ocean that is included in the code. An hierarchy of formulations that could be used to describe this interaction is presented. The main difference between them is the treatment of turbulent transfer within the oceanic boundary layer. The computed response to various levels of thermal driving and turbulent agitation in the mixed layer is discussed, as is the effect of various treatments of the conductive heat flux into the ice shelf. The performance of the different formulations that have been used in models of sub-ice-shelf circulation is assessed in comparison with observations of the turbulent heat flux beneath sea ice. Formulations that include an explicit parameterization of the oceanic boundary layer give results that lie within about 30\% of observation. Formulations that use constant bulk transfer coefficients entail a definite assumption about the level of turbulence in the water column and give melt/freeze rates that vary by a factor of 5, implying very different forcing on the respective ocean models.},
number = {8},
urldate = {2015-05-26},
journal = {Journal of Physical Oceanography},
author = {Holland, David M. and Jenkins, Adrian},
month = aug,
year = {1999},
note = {bibtex: holland\_modeling\_1999},
pages = {1787--1800},
file = {Snapshot:/Users/cbegeman/Zotero/storage/KCHFKVZJ/1520-0485(1999)0291787MTIOIA2.0.html:text/html},
}
@article{holland_effects_2006,
title = {The {Effects} of {Rotation} and {Ice} {Shelf} {Topography} on {Frazil}-{Laden} {Ice} {Shelf} {Water} {Plumes}},
volume = {36},
issn = {0022-3670},
url = {http://journals.ametsoc.org/doi/abs/10.1175/JPO2970.1},
doi = {10.1175/JPO2970.1},
abstract = {A model of the dynamics and thermodynamics of a plume of meltwater at the base of an ice shelf is presented. Such ice shelf water plumes may become supercooled and deposit marine ice if they rise (because of the pressure decrease in the in situ freezing temperature), so the model incorporates both melting and freezing at the ice shelf base and a multiple-size-class model of frazil ice dynamics and deposition. The plume is considered in two horizontal dimensions, so the influence of Coriolis forces is incorporated for the first time. It is found that rotation is extremely influential, with simulated plumes flowing in near-geostrophy because of the low friction at a smooth ice shelf base. As a result, an ice shelf water plume will only rise and become supercooled (and thus deposit marine ice) if it is constrained to flow upslope by topography. This result agrees with the observed distribution of marine ice under Filchner–Ronne Ice Shelf, Antarctica. In addition, it is found that the model only produces reasonable marine ice formation rates when an accurate ice shelf draft is used, implying that the characteristics of real ice shelf water plumes can only be captured using models with both rotation and a realistic topography.},
number = {12},
urldate = {2016-09-14},
journal = {Journal of Physical Oceanography},
author = {Holland, Paul R. and Feltham, Daniel L.},
month = dec,
year = {2006},
note = {00037},
pages = {2312--2327},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/Y8SMEWZY/Holland and Feltham - 2006 - The Effects of Rotation and Ice Shelf Topography o.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/TGS3BRH6/JPO2970.html:text/html},
}
@article{holland_response_2008,
title = {The {Response} of {Ice} {Shelf} {Basal} {Melting} to {Variations} in {Ocean} {Temperature}},
volume = {21},
issn = {0894-8755},
url = {http://journals.ametsoc.org/doi/abs/10.1175/2007JCLI1909.1},
doi = {10.1175/2007JCLI1909.1},
abstract = {A three-dimensional ocean general circulation model is used to study the response of idealized ice shelves to a series of ocean-warming scenarios. The model predicts that the total ice shelf basal melt increases quadratically as the ocean offshore of the ice front warms. This occurs because the melt rate is proportional to the product of ocean flow speed and temperature in the mixed layer directly beneath the ice shelf, both of which are found to increase linearly with ocean warming. The behavior of this complex primitive equation model can be described surprisingly well with recourse to an idealized reduced system of equations, and it is shown that this system supports a melt rate response to warming that is generally quadratic in nature. This study confirms and unifies several previous examinations of the relation between melt rate and ocean temperature but disagrees with other results, particularly the claim that a single melt rate sensitivity to warming is universally valid. The hypothesized warming does not necessarily require a heat input to the ocean, as warmer waters (or larger volumes of “warm” water) may reach ice shelves purely through a shift in ocean circulation. Since ice shelves link the Antarctic Ice Sheet to the climate of the Southern Ocean, this finding of an above-linear rise in ice shelf mass loss as the ocean steadily warms is of significant importance to understanding ice sheet evolution and sea level rise.},
number = {11},
urldate = {2016-08-22},
journal = {Journal of Climate},
author = {Holland, Paul R. and Jenkins, Adrian and Holland, David M.},
month = jun,
year = {2008},
note = {00133},
pages = {2558--2572},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/6JU7D4M6/Holland et al. - 2008 - The Response of Ice Shelf Basal Melting to Variati.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/A69YS376/2007JCLI1909.html:text/html},
}
@article{macayeal_numerical_1983,
title = {Numerical modeling of ice-shelf motion},
volume = {3},
url = {http://adsabs.harvard.edu/abs/1983AnGla...3..189M},
abstract = {Not Available},
urldate = {2015-06-01},
journal = {Annals of Glaciology},
author = {Macayeal, D. R. and Thomas, R. H.},
year = {1983},
note = {bibtex: macayeal\_numerical\_1983},
pages = {189--194},
}
@article{macayeal_numerical_1984,
title = {Numerical simulations of the {Ross} {Sea} tides},
volume = {89},
issn = {2156-2202},
url = {http://onlinelibrary.wiley.com/doi/10.1029/JC089iC01p00607/abstract},
doi = {10.1029/JC089iC01p00607},
abstract = {Tidal currents below the floating Ross Ice shelf are reconstructed by using a numerical tidal model. They are predominantly diurnal, achieve maximum strength in regions near where the ice shelf runs aground, and are significantly enhanced by topographic Rossby wave propagation along the ice front. A comparison with observations of the vertical motion of the ice shelf surface indicates that the model reproduces the diurnal tidal characteristics within 20\%. Similar agreement for the relatively weak semi-diurnal tides was not obtained, and this calls attention to possible errors of the open boundary forcing obtained from global-ocean tidal simulations and to possible errors in mapping zones of ice shelf grounding. Air-sea contact below the ice shelf is eliminated by the thick ice cover. The dominant sub-ice-shelf circulation may thus be tidally induced. A preliminary assessment of sub-ice-shelf conditions based on the numerical tidal simulations suggests that (1) strong barotropic circulation is driven along the ice front and (2) tidal fronts may form in the sub-ice-shelf cavity where the water column is thin and where the buoyancy input is weak.},
language = {en},
number = {C1},
urldate = {2016-08-30},
journal = {Journal of Geophysical Research: Oceans},
author = {MacAyeal, Douglas Reed},
month = jan,
year = {1984},
note = {00160},
pages = {607--615},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/IVVD7IED/MacAyeal - 1984 - Numerical simulations of the Ross Sea tides.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/3LP4MXBD/abstract.html:text/html},
}
@article{nicholls_measurements_2006,
title = {Measurements beneath an {Antarctic} ice shelf using an autonomous underwater vehicle},
volume = {33},
issn = {1944-8007},
url = {http://onlinelibrary.wiley.com/doi/10.1029/2006GL025998/abstract},
doi = {10.1029/2006GL025998},
abstract = {The cavities beneath Antarctic ice shelves are among the least studied regions of the World Ocean, yet they are sites of globally important water mass transformations. Here we report results from a mission beneath Fimbul Ice Shelf of an autonomous underwater vehicle. The data reveal a spatially complex oceanographic environment, an ice base with widely varying roughness, and a cavity periodically exposed to water with a temperature significantly above the surface freezing point. The results of this, the briefest of glimpses of conditions in this extraordinary environment, are already reforming our view of the topographic and oceanographic conditions beneath ice shelves, holding out great promises for future missions from similar platforms.},
language = {en},
number = {8},
urldate = {2016-09-14},
journal = {Geophysical Research Letters},
author = {Nicholls, K. W. and Abrahamsen, E. P. and Buck, J. J. H. and Dodd, P. A. and Goldblatt, C. and Griffiths, G. and Heywood, K. J. and Hughes, N. E. and Kaletzky, A. and Lane-Serff, G. F. and McPhail, S. D. and Millard, N. W. and Oliver, K. I. C. and Perrett, J. and Price, M. R. and Pudsey, C. J. and Saw, K. and Stansfield, K. and Stott, M. J. and Wadhams, P. and Webb, A. T. and Wilkinson, J. P.},
month = apr,
year = {2006},
note = {00073},
keywords = {Antarctica, Instruments and techniques, Oceanography / Arctic and Antarctic, Ocean / Continental shelf and slope processes},
pages = {L08612},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/NIZW2HD4/Nicholls et al. - 2006 - Measurements beneath an Antarctic ice shelf using .pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/8DEQVW9K/abstract.html:text/html},
}
@article{zonta_stably_2018,
title = {Stably {Stratified} {Wall}-{Bounded} {Turbulence}},
volume = {70},
issn = {0003-6900},
url = {https://doi.org/10.1115/1.4040838},
doi = {10.1115/1.4040838},
abstract = {Stably stratified wall-bounded turbulence is commonly encountered in many industrial and environmental processes. The interaction between turbulence and stratification induces remarkable modifications on the entire flow field, which in turn influence the overall transfer rates of mass, momentum, and heat. Although a vast proportion of the parameter range of wall-bounded stably stratified turbulence is still unexplored (in particular when stratification is strong), numerical simulations and experiments have recently developed a fairly robust picture of the flow structure, also providing essential ground for addressing more complex problems of paramount technological, environmental and geophysical importance. In this paper, we review models used to describe the influence of stratification on turbulence, as well as numerical and experimental methods and flow configurations for studying the resulting dynamics. Conclusions with a view on current open issues will be also provided.},
number = {040801},
urldate = {2021-05-04},
journal = {Applied Mechanics Reviews},
author = {Zonta, Francesco and Soldati, Alfredo},
month = aug,
year = {2018},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/HH9KRKI2/Zonta and Soldati - 2018 - Stably Stratified Wall-Bounded Turbulence.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/CXQKAF8Z/366076.html:text/html},
}
@article{flores_analysis_2011,
title = {Analysis of {Turbulence} {Collapse} in the {Stably} {Stratified} {Surface} {Layer} {Using} {Direct} {Numerical} {Simulation}},
volume = {139},
issn = {1573-1472},
url = {https://doi.org/10.1007/s10546-011-9588-2},
doi = {10.1007/s10546-011-9588-2},
abstract = {The nocturnal atmospheric boundary layer (ABL) poses several challenges to standard turbulence and dispersion models, since the stable stratification imposed by the radiative cooling of the ground modifies the flow turbulence in ways that are not yet completely understood. In the present work we perform direct numerical simulation of a turbulent open channel flow with a constant (cooling) heat flux imposed at the ground. This configuration provides a very simplified model for the surface layer at night. As a result of the ground cooling, the Reynolds stresses and the turbulent fluctuations near the ground re-adjust on times of the order of L/uτ, where L is the Obukhov length scale and uτis the friction velocity. For relatively weak cooling turbulence survives, but when \$\$\{Re\_L=Lu\_{\textbackslash}tau/{\textbackslash}nu {\textbackslash}lesssim 100\}\$\$turbulence collapses, a situation that is also observed in the ABL. This criterion, which can be locally measured in the field, is justified in terms of the scale separation between the largest and smallest structures of the dynamic sublayer.},
language = {en},
number = {2},
urldate = {2021-05-13},
journal = {Boundary-Layer Meteorology},
author = {Flores, O. and Riley, J. J.},
month = may,
year = {2011},
pages = {241--259},
file = {Springer Full Text PDF:/Users/cbegeman/Zotero/storage/D4KD66JK/Flores and Riley - 2011 - Analysis of Turbulence Collapse in the Stably Stra.pdf:application/pdf},
}
@article{nieuwstadt_direct_2005,
title = {Direct {Numerical} {Simulation} of {Stable} {Channel} {Flow} at {Large} {Stability}},
volume = {116},
issn = {1573-1472},
url = {https://doi.org/10.1007/s10546-004-2818-0},
doi = {10.1007/s10546-004-2818-0},
abstract = {We consider a model for the stable atmospheric boundary at large stability, i.e. near the limit where turbulence is no longer able to survive. The model is a plane horizontally homogeneous channel flow, which is driven by a constant pressure gradient and which has a no-slip wall at the bottom and a free-slip wall at the top. At the lower wall a constant negative temperature flux is imposed. First, we consider a direct numerical simulation of the same channel flow. The simulation is computed with the neutral channel flow as initial condition and computed as a function of time for various values of the stability parameter h/L, where h is the channel height and L is related to the Obukhov length. We find that a turbulent solution is only possible for h/L {\textless} 1.25 and for larger values turbulence decays. Next, we consider a theoretical model for this channel flow based on a simple gradient transfer closure. The resulting equations allow an exact solution for the case of a stationary flow. The velocity profile for this solution is almost linear as a function of height in most of the channel. In the limit of infinite Reynolds number, the temperature profile has a logarithmic singularity at the upper wall of the channel. For the cases where a turbulent flow is maintained in the numerical simulation, we find that the velocity and temperature profiles are in good agreement with the results of the theoretical model when the effects of the surface layer on the exchange coefficients are taken into account.},
language = {en},
number = {2},
urldate = {2021-05-13},
journal = {Boundary-Layer Meteorology},
author = {Nieuwstadt, F. T. M.},
month = aug,
year = {2005},
pages = {277--299},
file = {Springer Full Text PDF:/Users/cbegeman/Zotero/storage/VCAAHZQZ/Nieuwstadt - 2005 - Direct Numerical Simulation of Stable Channel Flow.pdf:application/pdf},
}
@article{wiel_cessation_2012,
title = {The {Cessation} of {Continuous} {Turbulence} as {Precursor} of the {Very} {Stable} {Nocturnal} {Boundary} {Layer}},
volume = {69},
issn = {0022-4928, 1520-0469},
url = {https://journals.ametsoc.org/view/journals/atsc/69/11/jas-d-12-064.1.xml},
doi = {10.1175/JAS-D-12-064.1},
abstract = {{\textless}section class="abstract"{\textgreater}{\textless}h2 class="abstractTitle text-title my-1" id="d34386520e72"{\textgreater}Abstract{\textless}/h2{\textgreater}{\textless}p{\textgreater}The mechanism behind the collapse of turbulence in the evening as a precursor to the onset of the very stable boundary layer is investigated. To this end a cooled, pressure-driven flow is investigated by means of a local similarity model. Simulations reveal a temporary collapse of turbulence whenever the surface heat extraction, expressed in its nondimensional form {\textless}em{\textgreater}h{\textless}/em{\textgreater}/{\textless}em{\textgreater}L{\textless}/em{\textgreater}, exceeds a critical value. As any temporary reduction of turbulent friction is followed by flow acceleration, the long-term state is unconditionally turbulent. In contrast, the temporary cessation of turbulence, which may actually last for several hours in the nocturnal boundary layer, can be understood from the fact that the time scale for boundary layer diffusion is much smaller than the time scale for flow acceleration. This limits the available momentum that can be used for downward heat transport. In case the surface heat extraction exceeds the so-called maximum sustainable heat flux (MSHF), the near-surface inversion rapidly increases. Finally, turbulent activity is largely suppressed by the intense density stratification that supports the emergence of a different, calmer boundary layer regime.{\textless}/p{\textgreater}{\textless}/section{\textgreater}},
language = {EN},
number = {11},
urldate = {2021-05-13},
journal = {Journal of the Atmospheric Sciences},
author = {Wiel, B. J. H. Van de and Moene, A. F. and Jonker, H. J. J.},
month = nov,
year = {2012},
note = {Publisher: American Meteorological Society
Section: Journal of the Atmospheric Sciences},
pages = {3097--3115},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/33UUCCVH/Wiel et al. - 2012 - The Cessation of Continuous Turbulence as Precurso.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/6CBGMVRG/jas-d-12-064.1.html:text/html},
}
@article{komori_turbulence_1983,
title = {Turbulence structure in stably stratified open-channel flow},
volume = {130},
issn = {1469-7645, 0022-1120},
url = {https://www.cambridge.org/core/journals/journal-of-fluid-mechanics/article/abs/turbulence-structure-in-stably-stratified-openchannel-flow/74C716F46BB78899BC053ED6A0046F61},
doi = {10.1017/S0022112083000944},
abstract = {The effects of stable stratification on turbulence structure have been experimentally investigated in stratified open-channel flow and a theoretical spectral-equation model has been applied to the stably stratified flow. The measurements were made in the outer layer of open-channel flow with strongly stable density gradient, where the wall effect was small. Velocity and temperature fluctuations were simultaneously measured by a laser-Doppler velocimeter and a cold-film probe. Measurements include turbulent intensities, correlation coefficients of turbulent fluxes and coherence–phase relationships. These turbulent quantities were correlated with the local gradient Richardson number and compared with the values calculated using a spectral-equation model and with other laboratory measurements. In stable conditions, turbulent motions approach wavelike motions, and negative heat and momentum transfer against the mean temperature and velocity gradient occurs in strongly stable stratification.},
language = {en},
urldate = {2021-05-13},
journal = {Journal of Fluid Mechanics},
author = {Komori, Satoru and Ueda, Hiromasa and Ogino, Fumimaru and Mizushina, Tokuro},
month = may,
year = {1983},
note = {Publisher: Cambridge University Press},
pages = {13--26},
file = {Snapshot:/Users/cbegeman/Zotero/storage/DTL528IS/74C716F46BB78899BC053ED6A0046F61.html:text/html},
}
@article{donda_collapse_2015,
title = {Collapse of turbulence in stably stratified channel flow: a transient phenomenon},
volume = {141},
copyright = {© 2014 Royal Meteorological Society},
issn = {1477-870X},
shorttitle = {Collapse of turbulence in stably stratified channel flow},
url = {https://rmets.onlinelibrary.wiley.com/doi/abs/10.1002/qj.2511},
doi = {https://doi.org/10.1002/qj.2511},
abstract = {The collapse of turbulence in a pressure-driven, cooled channel flow is studied by using 3D direct numerical simulations (DNS) in combination with theoretical analysis using a local similarity model. Previous studies with DNS reported a definite collapse of turbulence in cases when the normalized surface cooling h/L (with h the channel depth and L the Obukhov length) exceeded a value of 0.5. A recent study by the present authors succeeded in explaining this collapse using the so-called maximum sustainable heat flux (MSHF) theory. This states that collapse may occur when the ambient momentum of the flow is too weak to transport enough heat downward to compensate for the surface cooling. The MSHF theory predicts that, in pressure-driven flows, acceleration of the fluid after collapse will eventually cause a regeneration of turbulence, in contrast with the aforementioned DNS results. It also predicts that the flow should be able to survive ‘supercritical’ cooling rates, in cases when sufficient momentum is applied to the initial state. Here, both predictions are confirmed using DNS simulations. It is also shown that in DNS a recovery of turbulence will occur naturally, provided that perturbations of finite amplitude are imposed on the laminarized state and provided that sufficient time for flow acceleration is allowed. As such, we conclude that the collapse of turbulence in this configuration is a temporary, transient phenomenon for which a universal cooling rate does not exist. Finally, in the present work a one-to-one comparison between a parametrized, local similarity model and the turbulence-resolving model (DNS) is made. Although local similarity originates from observations that represent much larger Reynolds numbers than those covered by our DNS simulations, both methods appear to predict very similar mean velocity (and temperature) profiles. This suggests that in-depth analysis with DNS can be an attractive complementary tool with which to study atmospheric physics, in addition to tools that are able to represent high Reynolds number flows like large-eddy simulations.},
language = {en},
number = {691},
urldate = {2021-05-13},
journal = {Quarterly Journal of the Royal Meteorological Society},
author = {Donda, J. M. M. and Hooijdonk, I. G. S. van and Moene, A. F. and Jonker, H. J. J. and Heijst, G. J. F. van and Clercx, H. J. H. and Wiel, B. J. H. van de},
year = {2015},
note = {\_eprint: https://rmets.onlinelibrary.wiley.com/doi/pdf/10.1002/qj.2511},
keywords = {direct numerical simulation, collapse of turbulence, nocturnal boundary layer, revival of turbulence},
pages = {2137--2147},
file = {Snapshot:/Users/cbegeman/Zotero/storage/SKYWLPCB/qj.html:text/html},
}
@article{stoll_large-eddy_2008,
title = {Large-{Eddy} {Simulation} of the {Stable} {Atmospheric} {Boundary} {Layer} using {Dynamic} {Models} with {Different} {Averaging} {Schemes}},
volume = {126},
issn = {1573-1472},
url = {https://doi.org/10.1007/s10546-007-9207-4},
doi = {10.1007/s10546-007-9207-4},
abstract = {Large-eddy simulation (LES) of a stable atmospheric boundary layer is performed using recently developed dynamic subgrid-scale (SGS) models. These models not only calculate the Smagorinsky coefficient and SGS Prandtl number dynamically based on the smallest resolved motions in the flow, they also allow for scale dependence of those coefficients. This dynamic calculation requires statistical averaging for numerical stability. Here, we evaluate three commonly used averaging schemes in stable atmospheric boundary-layer simulations: averaging over horizontal planes, over adjacent grid points, and following fluid particle trajectories. Particular attention is focused on assessing the effect of the different averaging methods on resolved flow statistics and SGS model coefficients. Our results indicate that averaging schemes that allow the coefficients to fluctuate locally give results that are in better agreement with boundary-layer similarity theory and previous LES studies. Even among models that are local, the averaging method is found to affect model coefficient probability density function distributions and turbulent spectra of the resolved velocity and temperature fields. Overall, averaging along fluid pathlines is found to produce the best combination of self consistent model coefficients, first- and second-order flow statistics and insensitivity to grid resolution.},
language = {en},
number = {1},
urldate = {2021-05-13},
journal = {Boundary-Layer Meteorology},
author = {Stoll, Rob and Porté-Agel, Fernando},
month = jan,
year = {2008},
pages = {1--28},
file = {Springer Full Text PDF:/Users/cbegeman/Zotero/storage/J3TT5I22/Stoll and Porté-Agel - 2008 - Large-Eddy Simulation of the Stable Atmospheric Bo.pdf:application/pdf},
}
@article{nicholls_oceanographic_2001,
title = {Oceanographic conditions south of {Berkner} {Island}, beneath {Filchner}-{Ronne} {Ice} {Shelf}, {Antarctica}},
volume = {106},
issn = {2156-2202},
url = {http://onlinelibrary.wiley.com/doi/10.1029/2000JC000350/abstract},
doi = {10.1029/2000JC000350},
abstract = {We have made oceanographic measurements at two sites beneath the southern Filchner-Ronne Ice Shelf. Hot-water drilled access holes were made during January 1999, allowing conductivity-temperature-depth (CTD) profiling and the deployment of instrument moorings. The CTD profiles show that the entire water column is below the surface freezing point. We estimate the (summer) flux of water between the two sites to be 2×106 m3 s−1. The summer potential temperature-salinity properties of the water column suggest that this flow is part of a recirculation in the deepest part of the subice shelf cavity and the Filchner Depression. The recirculation is driven by a combination of the melting of deep basal ice and the freezing that results from the depressurization of the cold buoyant water as it ascends the ice shelf base. The source of the water was high-salinity shelf water (HSSW) produced in the Ronne Depression. This is the water that provides the external heat necessary for the strong melting at the deep grounding lines in the vicinity of Foundation Ice Stream. Instruments moored at the drill sites show that during the winter HSSW formed on the Berkner Shelf flows beneath the ice shelf and largely displaces the recirculating water from the two sites. This provides an externally driven through flow that is warmer (nearer the surface freezing point) and slower than the internal recirculation and which is low enough in density to escape the Filchner Depression.},
language = {en},
number = {C6},
urldate = {2017-02-16},
journal = {Journal of Geophysical Research: Oceans},
author = {Nicholls, K. W. and Østerhus, S. and Makinson, K. and Johnson, M. R.},
month = jun,
year = {2001},
note = {00052},
keywords = {Oceanography, Oceanography / Analytical modeling and laboratory experiments, Oceanography / Descriptive and regional oceanography},
pages = {11481--11492},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/N6PUC7GN/Nicholls et al. - 2001 - Oceanographic conditions south of Berkner Island, .pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/QM7TQJJJ/abstract.html:text/html},
}
@article{hattermann_two_2012,
title = {Two years of oceanic observations below the {Fimbul} {Ice} {Shelf}, {Antarctica}},
volume = {39},
issn = {1944-8007},
url = {http://onlinelibrary.wiley.com.oca.ucsc.edu/doi/10.1029/2012GL051012/abstract},
doi = {10.1029/2012GL051012},
abstract = {The mechanisms by which heat is delivered to Antarctic ice shelves are a major source of uncertainty when assessing the response of the Antarctic ice sheet to climate change. Direct observations of the ice shelf-ocean interaction are extremely scarce and in many regions melt rates from ice shelf-ocean models are not constrained by measurements. Our two years of data (2010 and 2011) from three oceanic moorings below the Fimbul Ice Shelf in the Eastern Weddell Sea show cold cavity waters, with average temperatures of less than 0.1°C above the surface freezing point. This suggests low basal melt rates, consistent with remote sensing-based, steady-state mass balance estimates for this sector of the Antarctic coast. Oceanic heat for basal melting is found to be supplied by two sources of warm water entering below the ice: (i) eddy-like bursts of Modified Warm Deep Water that access the cavity at depth for eight months of the record; and (ii) fresh surface water that flushes parts of the ice base with temperatures above freezing during late summer and fall. This interplay of processes implies that basal melting at the Fimbul Ice Shelf cannot simply be parameterized by coastal deep ocean temperatures, but instead appears directly linked to both solar forcing at the surface as well as to the dynamics of the coastal current system.},
language = {en},
number = {12},
urldate = {2016-01-31},
journal = {Geophysical Research Letters},
author = {Hattermann, Tore and Nøst, Ole Anders and Lilly, Jonathan M. and Smedsrud, Lars H.},
month = jun,
year = {2012},
note = {bibtex: hattermann\_two\_2012},
keywords = {ice shelf basal melting, ice-ocean interaction, polar oceanography, Cryosphere / Ice shelves, Oceanography / Arctic and Antarctic, Oceanography / water masses, Ocean / Southern / water masses, Ocean / Continental shelf and slope processes, Ocean / Coastal processes},
pages = {L12605},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/JP2SBKUK/Hattermann et al. - 2012 - Two years of oceanic observations below the Fimbul.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/8WUDNLUU/abstract.html:text/html},
}
@article{malyarenko_synthesis_2020,
title = {A synthesis of thermodynamic ablation at ice-ocean interfaces from theory, observations and models},
issn = {1463-5003},
url = {http://www.sciencedirect.com/science/article/pii/S1463500320301943},
doi = {10.1016/j.ocemod.2020.101692},
abstract = {Thermodynamic ablation of ice in contact with the ocean is an essential element of ice sheet and ocean interactions but is challenging to model and quantify. Building on earlier observations of sea ice ablation, a variety of recent theoretical, experimental and observational studies have considered ice ablation in contrasting geometries, from vertical to near-horizontal ice faces, and reveal different scaling behaviour for predicted ablation rates in different dynamical regimes. However, uncertainties remain about when the contrasting results should be applied, as existing model parameterisations do not capture all relevant regimes of ice-ocean ablation. To progress towards improved models of ice-ocean interaction, we synthesise current understanding into a classification of ablation types. We examine the effect of the classification on the parameterisation of turbulent fluxes from the ocean towards the ice, and identify the dominant processes next to ice interfaces of different orientation. Four ablation types are defined: melting and dissolving based on ocean temperatures, and shear-controlled and buoyancy-controlled regimes based on the dynamics of the near-ice molecular sublayer. We describe existing observational and modelling studies of sea ice, ice shelves, and glacier termini, as well as laboratory studies, to show how they fit into this classification. Two sets of observations from the Ross and Ronne Ice Shelf cavities suggest that both the buoyancy-controlled and shear-controlled regimes may be relevant under different oceanographic conditions. Overall, buoyancy-controlled dynamics are more likely when the molecular sublayer has lower Reynolds number, and shear for higher Reynolds number, although the observations suggest some variability about this trend.},
language = {en},
urldate = {2020-09-08},
journal = {Ocean Modelling},
author = {Malyarenko, Alena and Wells, Andrew J. and Langhorne, Patricia J. and Robinson, Natalie J. and Williams, Michael J. M. and Nicholls, Keith W.},
month = aug,
year = {2020},
keywords = {Cryosphere, Ablation, Heat transfer, Ice melting, Ice shelf cavity observations, Ice-ocean modelling},
pages = {101692},
file = {ScienceDirect Snapshot:/Users/cbegeman/Zotero/storage/3QUCKGRE/S1463500320301943.html:text/html},
}
@article{makinson_influence_2011,
title = {Influence of tides on melting and freezing beneath {Filchner}-{Ronne} {Ice} {Shelf}, {Antarctica}},
volume = {38},
issn = {1944-8007},
doi = {10.1029/2010GL046462},
abstract = {An isopycnic coordinate ocean circulation model is applied to the ocean cavity beneath Filchner-Ronne Ice Shelf, investigating the role of tides on sub-ice shelf circulation and ice shelf basal mass balance. Including tidal forcing causes a significant intensification in the sub-ice shelf circulation, with an increase in melting (3-fold) and refreezing (6-fold); the net melt rate and seawater flux through the cavity approximately doubles. With tidal forcing, the spatial pattern and magnitude of basal melting and freezing generally match observations. The 0.22 m a−1 net melt rate is close to satellite-derived estimates and at the lower end of oceanographic values. The Ice Shelf Water outflow mixes with shelf waters, forming a cold ({\textless}−1.9°C), dense overflow (0.83 Sv) that spills down the continental slope. These results demonstrate that tidal forcing is fundamental to both ice shelf-ocean interactions and deep-water formation in the southern Weddell Sea.},
language = {en},
number = {6},
journal = {Geophysical Research Letters},
author = {Makinson, Keith and Holland, Paul R. and Jenkins, Adrian and Nicholls, Keith W. and Holland, David M.},
year = {2011},
keywords = {Antarctica, mass balance, ice shelf, Cryosphere / Ice shelves, Cryosphere / Mass balance, Oceanography / Arctic and Antarctic, Oceanography / Surface waves and tides, Ocean / Continental shelf and slope processes, Ocean / Tides},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/J7HRFP9T/Makinson et al. - 2011 - Influence of tides on melting and freezing beneath.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/V5VQI54T/abstract.html:text/html},
}
@article{makinson_modeling_2002,
title = {Modeling {Tidal} {Current} {Profiles} and {Vertical} {Mixing} beneath {Filchner}–{Ronne} {Ice} {Shelf}, {Antarctica}},
volume = {32},
issn = {0022-3670},
doi = {10.1175/1520-0485(2002)032<0202:MTCPAV>2.0.CO;2},
abstract = {One of the warmest water masses beneath Filchner–Ronne Ice Shelf (FRIS) is dense, high salinity shelf water (HSSW) that flows into the sub-ice-shelf cavity from the ice front and occupies the lower portion of the water column. A one-dimensional turbulence closure ocean model has been applied to this high latitude sub-ice-shelf environment to demonstrate that tidal currents mix HSSW vertically through the water column and cause melting at the bottom of the ice shelf. Significantly FRIS lies near the critical latitude for the semidiurnal tide, where the Coriolis frequency equals the tidal frequency, resulting in a strongly depth-dependent tidal current and thick boundary layers. Using the model, the effect of the critical latitude, stratification, and the polarization of the tidal current ellipse on boundary layer structure and subsequent vertical mixing are examined. The model shows that stratification significantly affects how the shape of the tidal current ellipse varies with depth and that both the depth to which the pycnocline initially develops and the longer term melt rates are highly dependent on tidal current ellipse polarization. The sensitivity to both the stratification and the polarization are due, in large part, to the proximity of the critical latitude. Positive polarizations (anticlockwise rotating current vectors) quickly develop deeper pycnoclines and maintain higher melt rates than negative polarizations (clockwise rotating current vectors). For many areas beneath FRIS the polarization ranges from −0.3 to +0.3; here the modeled pycnocline development is sensitive to polarization, though the effect on the time-averaged melt rate is suppressed for positive polarizations. However, in key areas where the polarization exceeds ±0.3 and the ellipses are more open and circular, the effects of polarization are significant. Levels of tidal mixing and associated melting vary by more than an order of magnitude over the whole tidal ellipse polarization range, showing that very different mixing and melting regimes are present beneath FRIS.},
number = {1},
journal = {Journal of Physical Oceanography},
author = {Makinson, Keith},
year = {2002},
pages = {202--215},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/KZCCP4JW/Makinson - 2002 - Modeling Tidal Current Profiles and Vertical Mixin.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/CRUPA3ZC/1520-0485(2002)0320202MTCPAV2.0.html:text/html},
}
@article{mueller_tidal_2018,
title = {Tidal influences on a future evolution of the {Filchner}-{Ronne} {Ice} {Shelf} cavity in the {Weddell} {Sea}, {Antarctica}},
volume = {12},
issn = {1994-0424},
doi = {10.5194/tc-12-453-2018},
abstract = {Recent modeling studies of ocean circulation in the southern Weddell Sea, Antarctica, project an increase over this century of ocean heat into the cavity beneath Filchner–Ronne Ice Shelf (FRIS). This increase in ocean heat would lead to more basal melting and a modification of the FRIS ice draft. The corresponding change in cavity shape will affect advective pathways and the spatial distribution of tidal currents, which play important roles in basal melting under FRIS. These feedbacks between heat flux, basal melting, and tides will affect the evolution of FRIS under the influence of a changing climate. We explore these feedbacks with a three-dimensional ocean model of the southern Weddell Sea that is forced by thermodynamic exchange beneath the ice shelf and tides along the open boundaries. Our results show regionally dependent feedbacks that, in some areas, substantially modify the melt rates near the grounding lines of buttressed ice streams that flow into FRIS. These feedbacks are introduced by variations in meltwater production as well as the circulation of this meltwater within the FRIS cavity; they are influenced locally by sensitivity of tidal currents to water column thickness (wct) and non-locally by changes in circulation pathways that transport an integrated history of mixing and meltwater entrainment along flow paths. Our results highlight the importance of including explicit tidal forcing in models of future mass loss from FRIS and from the adjacent grounded ice sheet as individual ice-stream grounding zones experience different responses to warming of the ocean inflow.},
number = {2},
journal = {The Cryosphere},
author = {Mueller, R. D. and Hattermann, Tore and Howard, S. L. and Padman, L.},
year = {2018},
pages = {453--476},
}
@article{gwyther_cold_2020,
title = {Cold {Ocean} {Cavity} and {Weak} {Basal} {Melting} of the {Sørsdal} {Ice} {Shelf} {Revealed} by {Surveys} {Using} {Autonomous} {Platforms}},
volume = {125},
copyright = {©2020. The Authors.},
issn = {2169-9291},
url = {https://agupubs.onlinelibrary.wiley.com/doi/abs/10.1029/2019JC015882},
doi = {https://doi.org/10.1029/2019JC015882},
abstract = {Basal melting of ice shelves is inherently difficult to quantify through direct observations, yet it is a critical factor controlling Antarctic mass balance and global sea-level rise. While much research attention is paid to larger ice shelves and those experiencing the most rapid change, many smaller, unstudied ice shelves offer valuable insights. Here, we investigate the oceanographic conditions and melting beneath the Sørsdal ice shelf, East Antarctica. We present results from the 2018/2019 Sørsdal deployment of the University of Tasmania's autonomous underwater vehicle nupiri muka. Oceanography adjacent to and beneath the ice shelf front shows a cold and relatively saline environment dominated by Winter Water and Dense Shelf Water, while bathymetry measurements show a deep (∼1,200 m) trough running into the ice shelf cavity. Two multiyear deployments of Autonomous Phase-sensitive Radar Echo Sounders on the surface of the ice shelf show weak melt rates (average of 1.6 and 2.3 m yr−1) with low temporal variability. These observations are supported by numerical ocean model and satellite estimates of melting. We speculate that the presence of a ∼825 m thick (350 m to at least 1,175 m) homogeneous layer of cold, dense water blocks access from warmer waters that intrude into Prydz Bay from offshore, resulting in weak melt rates. However, the newly identified trough means that the ice shelf is vulnerable to any decrease in polynya activity that allows warm water to enter the cavity. This could lead to increased basal melting and mass loss through this sector of Antarctica.},
language = {en},
number = {6},
urldate = {2021-05-12},
journal = {Journal of Geophysical Research: Oceans},
author = {Gwyther, David E. and Spain, Erica A. and King, Peter and Guihen, Damien and Williams, Guy D. and Evans, Eleri and Cook, Sue and Richter, Ole and Galton‐Fenzi, Benjamin K. and Coleman, Richard},
year = {2020},
note = {\_eprint: https://agupubs.onlinelibrary.wiley.com/doi/pdf/10.1029/2019JC015882},
keywords = {Antarctica, ice shelves, oceanography, basal melting, APRES, AUV},
pages = {e2019JC015882},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/8TH79XFU/Gwyther et al. - 2020 - Cold Ocean Cavity and Weak Basal Melting of the Sø.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/THGVASEJ/2019JC015882.html:text/html},
}
@article{ohya_wind-tunnel_2001,
title = {Wind-{Tunnel} {Study} {Of} {Atmospheric} {Stable} {Boundary} {Layers} {Over} {A} {Rough} {Surface}},
volume = {98},
issn = {1573-1472},
url = {https://doi.org/10.1023/A:1018767829067},
doi = {10.1023/A:1018767829067},
abstract = {Wind-tunnel simulations of theatmospheric stable boundary layer (SBL) developedover a rough surface were conducted by using athermally stratified wind tunnel at the Research Institutefor Applied Mechanics (RIAM), Kyushu University. Thepresent experiment is a continuation of the workcarried out in a wind tunnel at Colorado StateUniversity (CSU), where the SBL flows were developed over asmooth surface. Stably stratified flows were createdby heating the wind-tunnel airflow to a temperature ofabout 40–50°and by cooling the test-section floor toa temperature of about 10°. To simulate therough surface, a chain roughness was placed over thetest-section floor. We have investigated the buoyancyeffect on the turbulent boundary layer developed overthis rough surface for a wide range of stability,particularly focusing on the turbulence structure andtransport process in the very stable boundary layer.The present experimental results broadly confirm theresults obtained in the CSU experiment with the smoothsurface, and emphasizes the following features: thevertical profiles of turbulence statistics exhibitdifferent behaviour in two distinct stability regimes with weak and strong stability,corresponding to the difference in the verticalprofiles of the local Richardson number. The tworegimes are separated by the critical Richardsonnumber. The magnitudes in turbulence intensities andturbulent fluxes for the weak stability regime aremuch greater than those of the CSU experiments becauseof the greater surface roughness. For the very stableboundary layer, the turbulent fluxes of momentum andheat tend to vanish and wave-like motions due to theKelvin–Helmholtz instability and the rolling up andbreaking of those waves can be observed. Furthermore,the appearance of internal gravity waves is suggestedfrom cross-spectrum analyses.},
language = {en},
number = {1},
urldate = {2021-05-13},
journal = {Boundary-Layer Meteorology},
author = {Ohya, Yuji},
month = jan,
year = {2001},
pages = {57--82},
file = {Springer Full Text PDF:/Users/cbegeman/Zotero/storage/WYJUBBD5/Ohya - 2001 - Wind-Tunnel Study Of Atmospheric Stable Boundary L.pdf:application/pdf},
}
@article{middleton_numerical_2021,
title = {Numerical {Simulations} of {Melt}-{Driven} {Double}-{Diffusive} {Fluxes} in a {Turbulent} {Boundary} {Layer} beneath an {Ice} {Shelf}},
volume = {51},
issn = {0022-3670, 1520-0485},
url = {https://journals.ametsoc.org/view/journals/phoc/51/2/jpo-d-20-0114.1.xml},
doi = {10.1175/JPO-D-20-0114.1},
abstract = {{\textless}section class="abstract"{\textgreater}{\textless}h2 class="abstractTitle text-title my-1" id="d63656025e126"{\textgreater}Abstract{\textless}/h2{\textgreater}{\textless}p{\textgreater}The transport of heat and salt through turbulent ice shelf–ocean boundary layers is a large source of uncertainty within ocean models of ice shelf cavities. This study uses small-scale, high-resolution, 3D numerical simulations to model an idealized boundary layer beneath a melting ice shelf to investigate the influence of ambient turbulence on double-diffusive convection (i.e., convection driven by the difference in diffusivities between salinity and temperature). Isotropic turbulence is forced throughout the simulations and the temperature and salinity are initialized with homogeneous values similar to observations. The initial temperature and the strength of forced turbulence are varied as controlling parameters within an oceanographically relevant parameter space. Two contrasting regimes are identified. In one regime double-diffusive convection dominates, and in the other convection is inhibited by the forced turbulence. The convective regime occurs for high temperatures and low turbulence levels, where it is long lived and affects the flow, melt rate, and melt pattern. A criterion for identifying convection in terms of the temperature and salinity profiles, and the turbulent dissipation rate, is proposed. This criterion may be applied to observations and theoretical models to quantify the effect of double-diffusive convection on ice shelf melt rates.{\textless}/p{\textgreater}{\textless}/section{\textgreater}},
language = {EN},
number = {2},
journal = {Journal of Physical Oceanography},
author = {Middleton, Leo and Vreugdenhil, Catherine A. and Holland, Paul R. and Taylor, John R.},
month = jan,
year = {2021},
note = {Publisher: American Meteorological Society
Section: Journal of Physical Oceanography},
pages = {403--418},
file = {Snapshot:/Users/cbegeman/Zotero/storage/3GYP2QKA/jpo-d-20-0114.1.html:text/html},
}
@article{rohr_growth_1988,
title = {Growth and decay of turbulence in a stably stratified shear flow},
volume = {195},
issn = {1469-7645, 0022-1120},
url = {https://www.cambridge.org/core/journals/journal-of-fluid-mechanics/article/growth-and-decay-of-turbulence-in-a-stably-stratified-shear-flow/1BD7CF9BC8351B89BEFEE7A22BBBF00E},
doi = {10.1017/S0022112088002332},
abstract = {The behaviour of an evolving, stably stratified turbulent shear flow was investigated in a ten-layer, closed-loop, salt-stratified water channel. Simultaneous single-point measurements of the mean and fluctuating density and longitudinal and vertical velocities were made over a wide range of downstream positions. For strong stability, i.e. a mean gradient Richardson number Ri greater than a critical value of Ricr ≈ 0.25, there is no observed growth of turbulence and the buoyancy effects are similar to those in the unsheared experiments of Stillinger, Helland \& Van Atta (1983) and Itsweire, Helland \& Van Atta (1986). For values of Richardson number less than Ricr the turbulence grows at a rate depending on Ri and for large evolution times the ratio between the Ozmidov and turbulent lengthscale approaches a constant value which is also a function of Richardson number.Normalized velocity and density power spectra for the present experiments conform to normalized spectra from previous moderate- to high-Reynolds-number studies. With increasing \${\textbackslash}tau = (x/{\textbackslash}overline\{U\}) ({\textbackslash}partial {\textbackslash}overline\{U\}/{\textbackslash}partial z)\$ or decreasing stability, the stratified shear spectra exhibit greater portions of the universal non-stratified spectrum curve. The shapes of the shear-stress and buoyancy-flux cospectra confirm that they act as sources and sinks for the velocity and density fluctuations.},
urldate = {2017-10-24},
journal = {Journal of Fluid Mechanics},
author = {Rohr, J. J. and Itsweire, E. C. and Helland, K. N. and Atta, C. W. Van},
month = oct,
year = {1988},
pages = {77--111},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/RPZXAAL9/Rohr et al. - 1988 - Growth and decay of turbulence in a stably stratif.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/XKIPBQY7/1BD7CF9BC8351B89BEFEE7A22BBBF00E.html:text/html},
}
@misc{holt_numerical_1992,
title = {A numerical study of the evolution and structure of homogeneous stably stratified sheared turbulence},
url = {/core/journals/journal-of-fluid-mechanics/article/numerical-study-of-the-evolution-and-structure-of-homogeneous-stably-stratified-sheared-turbulence/340AEE0415A0EA2F4915C871505E44E1},
abstract = {{\textless}div class="title"{\textgreater}A numerical study of the evolution and structure of homogeneous stably stratified sheared turbulence{\textless}/div{\textgreater} - Volume 237 - Steven E. Holt, Jeffrey R. Koseff, Joel H. Ferziger},
urldate = {2017-10-24},
journal = {Journal of Fluid Mechanics},
author = {Holt, Steven E. and Koseff, Jeffrey R. and Ferziger, Joel H.},
month = apr,
year = {1992},
doi = {10.1017/S0022112092003513},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/XVNMTYL5/Holt et al. - 1992 - A numerical study of the evolution and structure o.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/EQTRRDPR/340AEE0415A0EA2F4915C871505E44E1.html:text/html},
}
@article{armenio_investigation_2002,
title = {An investigation of stably stratified turbulent channel flow using large-eddy simulation},
volume = {459},
issn = {0022-1120, 1469-7645},
url = {https://www.cambridge.org/core/product/identifier/S0022112002007851/type/journal_article},
doi = {10.1017/S0022112002007851},
abstract = {Boundary-forced stratified turbulence is studied in the prototypical case of turbulent
channel flow subject to stable stratification. The large-eddy simulation approach is
used with a mixed subgrid model that involves a dynamic eddy viscosity component
and a scale-similarity component. After an initial transient, the flow reaches a new
balanced state corresponding to active wall-bounded turbulence with reduced vertical
transport which, for the cases in our study with moderate-to-large levels of stratification,
coexists with internal wave activity in the core of the channel. A systematic
reduction of turbulence levels, density fluctuations and associated vertical transport
with increasing stratification is observed. Countergradient buoyancy flux is observed
in the outer region for sufficiently high stratification.
Mixing of the density field in stratified channel flow results from turbulent events
generated near the boundaries that couple with the outer, more stable flow. The
vertical density structure is thus of interest for analogous boundary-forced mixing
situations in geophysical flows. It is found that, with increasing stratification, the
mean density profile becomes sharper in the central region between the two turbulent
layers at the upper and lower walls, similar to observations in field measurements as
well as laboratory experiments with analogous density-mixing situations.
Channel flow is strongly inhomogeneous with alternative choices for the Richardson
number. In spite of these complications, the gradient Richardson number,
Ri
g
, appears
to be the important local determinant of buoyancy effects. All simulated cases show
that correlation coefficients associated with vertical transport collapse from their nominal unstratified
values over a narrow range, 0.15 {\textless}
Ri
g
{\textless} 0.25. The vertical turbulent
Froude number,
Fr
w
, has an
O
(1) value across most of the channel. It
is remarkable that stratified channel flow, with such a large variation of overall
density difference (factor of 26) between cases, shows a relatively universal behaviour
of correlation coefficients and vertical Froude number when plotted as a function of
Ri
g
. The visualizations show wavy motion in the core region where the gradient
Richardson number,
Ri
g
, is large and low-speed streaks in the near-wall region,
typical of unstratified channel flow, where
Ri
g
is small. It appears from the
visualizations that, with increasing stratification, the region with wavy motion progressively encroaches
into the zone with active turbulence; the location of
Ri
g
≃ 0.2 roughly
corresponds to the boundary between the two zones.},
language = {en},
urldate = {2021-05-07},
journal = {Journal of Fluid Mechanics},
author = {Armenio, Vincenzo and Sarkar, Sutanu},
month = may,
year = {2002},
pages = {1--42},
}
@article{peltier_mixing_2003,
title = {Mixing {Efficiency} in {Stratified} {Shear} {Flows}},
volume = {35},
url = {http://dx.doi.org/10.1146/annurev.fluid.35.101101.161144},
doi = {10.1146/annurev.fluid.35.101101.161144},
abstract = {The issue of the physical mechanism(s) that control the efficiency with which the density field in stably stratified fluid is mixed by turbulent processes has remained enigmatic. Similarly enigmatic has been an explanation of the numerical value of ∼0.2, which is observed to characterize this efficiency experimentally. We review recent work on the turbulence transition in stratified parallel flows that demonstrates that this value is not only numerically predictable but also that it is expected to be a nonmonotonic function of the Richardson number that characterizes preturbulent stratification strength. This value of the mixing efficiency appears to be characteristic of the late-time behavior of the turbulent flow that develops after an initially laminar shear flow has undergone the transition to turbulence through an intermediate instability of Kelvin-Helmholtz type.},
number = {1},
urldate = {2016-08-19},
journal = {Annual Review of Fluid Mechanics},
author = {Peltier, W. R. and Caulfield, C. P.},
year = {2003},
keywords = {stratified flows, transition},
pages = {135--167},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/TANEA9SZ/Peltier and Caulfield - 2003 - Mixing Efficiency in Stratified Shear Flows.pdf:application/pdf},
}
@article{rignot_rapid_2002,
title = {Rapid {Bottom} {Melting} {Widespread} near {Antarctic} {Ice} {Sheet} {Grounding} {Lines}},
volume = {296},
url = {http://www.sciencemag.org/content/296/5575/2020.abstract},
doi = {10.1126/science.1070942},
abstract = {As continental ice from Antarctica reaches the grounding line and begins to float, its underside melts into the ocean. Results obtained with satellite radar interferometry reveal that bottom melt rates experienced by large outlet glaciers near their grounding lines are far higher than generally assumed. The melting rate is positively correlated with thermal forcing, increasing by 1 meter per year for each 0.1°C rise in ocean temperature. Where deep water has direct access to grounding lines, glaciers and ice shelves are vulnerable to ongoing increases in ocean temperature.},
number = {5575},
urldate = {2011-12-01},
journal = {Science},
author = {Rignot, Eric and Jacobs, Stanley S.},
month = jun,
year = {2002},
note = {bibtex: rignot\_rapid\_2002},
pages = {2020 --2023},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/3XM8R7SB/Rignot and Jacobs - 2002 - Rapid Bottom Melting Widespread near Antarctic Ice.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/MDAUFUN4/2020.html:text/html},
}
@article{mcphee_turbulent_1983,
title = {Turbulent heat and momentum transfer in the oceanic boundary layer under melting pack ice},
volume = {88},
issn = {2156-2202},
url = {http://onlinelibrary.wiley.com.oca.ucsc.edu/doi/10.1029/JC088iC05p02827/abstract},
doi = {10.1029/JC088iC05p02827},
abstract = {A theory for momentum flux in the planetary boundary layer (PBL) stabilized by continuous surface buoyancy is extended to include turbulent flux of an arbitrary scalar contaminant and then used to estimate how wind-driven sea ice melts as it encounters temperatures in the oceanic boundary layer that are above the melting point. Given wind stress and temperature difference Δθ across the oceanic PBL, the theory predicts melt rate and ice drift velocity. Results indicate that the effect of buoyancy on PBL turbulence is significant even with small values of Δθ({\textless}0.5 K). Curves of melt rate and ice speed as functions of u*, the interfacial friction velocity, show that melting is strongly dependent on stress at the interface and that the effective drag on the ice undersurface is significantly reduced at oceanic temperatures commonly observed in the marginal ice zone. The latter suggests that divergence will occur at the ice margin when off-ice winds advect the pack over water above the melting temperature. The structure of mean currents beneath the ice is also investigated; results indicate that advection will play an important, if not dominant, role in determining water column properties near an ice edge front.},
language = {en},
number = {C5},
urldate = {2016-01-05},
journal = {Journal of Geophysical Research: Oceans},
author = {McPhee, Miles G.},
month = mar,
year = {1983},
note = {bibtex: mcphee\_turbulent\_1983},
pages = {2827--2835},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/NQXGI8CT/McPhee - 1983 - Turbulent heat and momentum transfer in the oceani.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/9KN2SB7C/abstract.html:text/html},
}
@book{mcphee_air-ice-ocean_2008,
address = {Dordrecht},
title = {Air-ice-ocean interaction: turbulent ocean boundary layer exchange processes},
isbn = {978-0-387-78334-5 0-387-78334-2 978-0-387-78334-5 978-0-387-78335-2},
shorttitle = {Air-ice-ocean interaction},
language = {eng},
publisher = {Springer},
author = {McPhee, Miles G.},
year = {2008},
note = {bibtex: mcphee\_air-ice-ocean\_2008},
keywords = {Oceanography, Geography, Ocean-atmosphere interaction, Eis, Environmental Physics, Grenzschicht, Mechanics, Fluids, Thermodynamics, Physical geography, Polargebiete, Geophysics, Atmosphere, Fluid Dynamics / Turbulent boundary layer},
}
@article{mcphee_turbulent_1992,
title = {Turbulent heat flux in the upper ocean under sea ice},
volume = {97},
issn = {2156-2202},
url = {http://onlinelibrary.wiley.com.oca.ucsc.edu/doi/10.1029/92JC00239/abstract},
doi = {10.1029/92JC00239},
abstract = {Turbulence data from three Arctic drift station experiments demonstrate features of turbulent heat transfer in the oceanic boundary layer. Time series analysis of several w′T′ records shows that heat and momentum flux occur at nearly the same scales, typically by turbulent eddies of the order of 10–20 m in horizontal extent and a few meters in vertical extent. Probability distribution functions of w′T′ have large skewness and kurtosis, where the latter confirms that most of the flux occurs in intermittent “events” with positive and negative excursions an order of magnitude larger than the mean value. An estimate of the eddy heat diffusivity in the outer (Ekman) part of the boundary layer, based on measured heat flux and temperature gradient during a diurnal tidal cycle over the Yermak Plateau slope north of Fram Strait, agrees reasonably well with the eddy viscosity, with values as high as 0.15 m2 s−1. An analysis of measurements made near the ice-ocean interface at the three stations shows that heat flux increases with both temperature elevation above freezing and with friction velocity at the interface. It also reveals a surprising uniformity in parameters describing the heat and mass transfer: e.g., the thickness of the “transition sublayer” (from a modified version of the Yaglom-Kader theory) is about 10 cm at all three sites, despite nearly a fivefold difference in the under-ice roughness z0, which ranges from approximately 2 to 9 cm. A much simplified model for heat and mass transfer at the ice-ocean interface, suggested by the relative uniformity of the heat transfer coefficients at the three sites, is outlined.},
language = {en},
number = {C4},
urldate = {2016-01-31},
journal = {Journal of Geophysical Research: Oceans},
author = {McPhee, Miles G.},
month = apr,
year = {1992},
note = {bibtex: mcphee\_turbulent\_1992},
keywords = {Oceanography / Arctic and Antarctic, Cryosphere / Ice mechanics and air/sea/ice exchange processes, Oceanography / Upper ocean and mixed layer processes, Fluid Dynamics / Turbulence, diffusion, and mixing processes},
pages = {5365--5379},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/U6UB9CK9/McPhee - 1992 - Turbulent heat flux in the upper ocean under sea i.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/D8984G3K/abstract.html:text/html},
}
@article{mcphee_revisiting_2008-1,
title = {Revisiting heat and salt exchange at the ice-ocean interface: {Ocean} flux and modeling considerations},
volume = {113},
issn = {2156-2202},
shorttitle = {Revisiting heat and salt exchange at the ice-ocean interface},
url = {http://onlinelibrary.wiley.com.oca.ucsc.edu/doi/10.1029/2007JC004383/abstract},
doi = {10.1029/2007JC004383},
abstract = {Properly describing heat and salt flux at the ice/ocean interface is essential for understanding and modeling the energy and mass balance of drifting sea ice. Basal growth or ablation depends on the ratio, R, of the interface heat exchange coefficient to that of salt, such that as R increases so does the rate-limiting impact of salt diffusion. Observations of relatively slow melt rates in above freezing seawater (plus migration of summer “false bottoms”) suggest by analogy with laboratory studies that double diffusion of heat and salt from the ocean is important during the melting process, with numeric values for R estimated to range from 35 to 70. If the same double-diffusive principles apply for ice growth as for melting (i.e., if the process is symmetric), supercooling (possibly relieved by frazil crystal production) would occur under rapidly growing ice, yet neither extensive supercooling nor frazil accumulation is found in Arctic pack ice with limited atmospheric contact. Physical properties and turbulent fluxes of heat and salt were measured in the relatively controlled setting of a tidally driven Svalbard fjord, under growing fast ice in late winter. The data failed to show supercooling, frazil production, or enhanced heat flux, suggesting that the double diffusive process is asymmetric. Modeling results compared with measured turbulent fluxes imply that R = 1 when ice freezes; i.e., that double-diffusive tendencies are relieved at or above the immediate interface. An algorithm for calculating ice/ocean heat and salt flux accommodating the different processes is presented, along with recommended ranges for the interface exchange coefficients.},
language = {en},
number = {C6},
urldate = {2016-01-31},
journal = {Journal of Geophysical Research: Oceans},
author = {McPhee, Miles G. and Morison, J. H. and Nilsen, F.},
month = jun,
year = {2008},
note = {bibtex: mcphee\_revisiting\_2008},
keywords = {Sea ice, ice growth and ablation, Cryosphere / Ice mechanics and air/sea/ice exchange processes, Double diffusion, Fluid Dynamics / Turbulence, Fluid Dynamics / Turbulence, diffusion, and mixing processes},
pages = {C06014},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/INC5W7PQ/McPhee et al. - 2008 - Revisiting heat and salt exchange at the ice-ocean.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/C24PZ4NW/abstract.html:text/html},
}
@article{mondal_ablation_2019,
title = {Ablation of sloping ice faces into polar seawater},
volume = {863},
issn = {0022-1120, 1469-7645},
url = {https://www.cambridge.org/core/journals/journal-of-fluid-mechanics/article/ablation-of-sloping-ice-faces-into-polar-seawater/E8256D328B03930905A20AF3CD9E5ED1},
doi = {10.1017/jfm.2018.970},
abstract = {The effects of the slope of an ice–seawater interface on the mechanisms and rate of ablation of the ice by natural convection are examined using turbulence-resolving simulations. Solutions are obtained for ice slopes
{\textbackslash}unicode[STIX]\{x1D703\}=2{\textasciicircum}\{{\textbackslash}circ \}\{-\}90{\textasciicircum}\{{\textbackslash}circ \}{\textbackslash}unicode[STIX]\{x1D703\}=2{\textasciicircum}\{{\textbackslash}circ \}\{-\}90{\textasciicircum}\{{\textbackslash}circ \}
, at a fixed ambient salinity and temperature, chosen to represent common Antarctic ocean conditions. For laminar boundary layers the ablation rate decreases with height, whereas in the turbulent regime the ablation rate is found to be height independent. The simulated laminar ablation rates scale with
({\textbackslash}sin {\textbackslash}unicode[STIX]\{x1D703\}){\textasciicircum}\{1/4\}({\textbackslash}sin {\textbackslash}unicode[STIX]\{x1D703\}){\textasciicircum}\{1/4\}
, whereas in the turbulent regime it follows a
({\textbackslash}sin {\textbackslash}unicode[STIX]\{x1D703\}){\textasciicircum}\{2/3\}({\textbackslash}sin {\textbackslash}unicode[STIX]\{x1D703\}){\textasciicircum}\{2/3\}
scaling, both consistent with the theoretical predictions developed here. The reduction in the ablation rate with shallower slopes arises as a result of the development of stable density stratification beneath the ice face, which reduces turbulent buoyancy fluxes to the ice. The turbulent kinetic energy budget of the flow shows that, for very steep slopes, both buoyancy and shear production are drivers of turbulence, whereas for shallower slopes shear production becomes the dominant mechanism for sustaining turbulence in the convective boundary layer.},
language = {en},
urldate = {2019-09-05},
journal = {Journal of Fluid Mechanics},
author = {Mondal, Mainak and Gayen, Bishakhdatta and Griffiths, Ross W. and Kerr, Ross C.},
month = mar,
year = {2019},
keywords = {ice sheets, buoyant boundary layers, turbulent convection},
pages = {545--571},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/MMW2W6UX/Mondal et al. - 2019 - Ablation of sloping ice faces into polar seawater.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/7CVVBFKH/E8256D328B03930905A20AF3CD9E5ED1.html:text/html},
}
@article{ramudu_turbulent_2016,
title = {Turbulent heat exchange between water and ice at an evolving ice–water interface},
volume = {798},
issn = {0022-1120, 1469-7645},
url = {https://www.cambridge.org/core/journals/journal-of-fluid-mechanics/article/turbulent-heat-exchange-between-water-and-ice-at-an-evolving-icewater-interface/E3C271817E23FD1859078E5E86FEBDFA},
doi = {10.1017/jfm.2016.321},
abstract = {We conduct laboratory experiments on the time evolution of an ice layer cooled from below and subjected to a turbulent shear flow of warm water from above. Our study is motivated by observations of warm water intrusion into the ocean cavity under Antarctic ice shelves, accelerating the melting of their basal surfaces. The strength of the applied turbulent shear flow in our experiments is represented in terms of its Reynolds number
\$Re\$
, which is varied over the range
\$2.0{\textbackslash}times 10{\textasciicircum}\{3\}{\textbackslash}leqslant Re{\textbackslash}leqslant 1.0{\textbackslash}times 10{\textasciicircum}\{4\}\$
. Depending on the water temperature, partial transient melting of the ice occurs at the lower end of this range of
\$Re\$
and complete transient melting of the ice occurs at the higher end. Following these episodes of transient melting, the ice reforms at a rate that is independent of
\$Re\$
. We fit our experimental measurements of ice thickness and temperature to a one-dimensional model for the evolution of the ice thickness in which the turbulent heat transfer is parameterized in terms of the friction velocity of the shear flow. Applying our model to field measurements at a site under the Antarctic Pine Island Glacier ice shelf yields a predicted melt rate that exceeds present-day observations.},
urldate = {2017-08-21},
journal = {Journal of Fluid Mechanics},
author = {Ramudu, Eshwan and Hirsh, Benjamin Henry and Olson, Peter and Gnanadesikan, Anand},
month = jul,
year = {2016},
keywords = {ocean processes, solidification/melting, Fluid Dynamics / Turbulent boundary layer},
pages = {572--597},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/WFBEFKB6/Ramudu et al. - 2016 - Turbulent heat exchange between water and ice at a.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/5UCURY4J/E3C271817E23FD1859078E5E86FEBDFA.html:text/html},
}
@article{jenkins_simple_2016,
title = {A {Simple} {Model} of the {Ice} {Shelf}-{Ocean} {Boundary} {Layer} and {Current}},
volume = {46},
issn = {0022-3670},
doi = {10.1175/JPO-D-15-0194.1},
abstract = {Ocean-forced basal melting has been implicated in the widespread thinning of Antarctic ice shelves, but an understanding of what determines melt rates is hampered by limited knowledge of the buoyancy- and frictionally controlled flows along the ice shelf base that regulate heat transfer from ocean to ice. In an attempt to address this deficiency, a simple model of a buoyant boundary flow, considering only the spatial dimension perpendicular to the boundary, is presented. Results indicate that two possible flow regimes exist: a weakly stratified, geostrophic cross-slope current with upslope flow within a buoyant Ekman layer or a strongly stratified, upslope current with a weak cross-slope flow. The latter regime, which is analogous to the steady solution for a katabatic wind, is most appropriate when the ice–ocean interface is steep. For the gentle slopes typical of Antarctic ice shelves, the buoyant Ekman regime, which has similarities with the case of an unstratified density current on a slope, provides some useful insight. When combined with a background flow, a range of possible near-ice current profiles emerge as a result of arrest or enhancement of the upslope Ekman transport. A simple expression for the upslope transport can be formed that is analogous to that for the wind-forced surface Ekman layer, with curvature of the ice shelf base replacing the wind stress curl in driving exchange between the Ekman layer and the geostrophic current below.},
number = {6},
journal = {Journal of Physical Oceanography},
author = {Jenkins, Adrian},
year = {2016},
pages = {1785--1803},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/DKPLC98F/Jenkins - 2016 - A Simple Model of the Ice Shelf–Ocean Boundary Lay.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/XANWJHXJ/JPO-D-15-0194.html:text/html},
}
@article{garcia-villalba_turbulence_2011,
title = {Turbulence modification by stable stratification in channel flow},
volume = {23},
issn = {1070-6631},
url = {https://aip.scitation.org/doi/abs/10.1063/1.3560359},
doi = {10.1063/1.3560359},
number = {4},
urldate = {2021-05-04},
journal = {Physics of Fluids},
author = {García-Villalba, Manuel and del Álamo, Juan C.},
month = apr,
year = {2011},
note = {Publisher: American Institute of Physics},
pages = {045104},
file = {Snapshot:/Users/cbegeman/Zotero/storage/48WWLGZL/1.html:text/html},
}
@article{monin_basic_1954,
title = {Basic laws of turbulent mixing in the surface layer of the atmosphere},
volume = {151},
url = {http://www.mcnaughty.com/keith/papers/Monin_and_Obukhov_1954.pdf},
number = {163},
urldate = {2017-05-26},
journal = {Contrib. Geophys. Inst. Acad. Sci. USSR},
author = {Monin, A. S. and Obukhov, A. M. F.},
year = {1954},
pages = {e187},
}
@article{jenkins_role_2001,
title = {The {Role} of {Meltwater} {Advection} in the {Formulation} of {Conservative} {Boundary} {Conditions} at an {Ice}–{Ocean} {Interface}},
volume = {31},
issn = {0022-3670},
url = {https://journals.ametsoc.org/doi/abs/10.1175/1520-0485(2001)031%3C0285:TROMAI%3E2.0.CO;2},
doi = {10.1175/1520-0485(2001)031<0285:TROMAI>2.0.CO;2},
abstract = {Upper boundary conditions for numerical models of the ocean are conventionally formulated under the premise that the boundary is a material surface. In the presence of an ice cover, such an assumption can lead to nonconservative equations for temperature, salinity, and other tracers. The problem arises because conditions at the ice–ocean interface differ from those in the water beneath. Advection of water with interfacial properties into the interior of the ocean therefore constitutes a tracer flux, neglect of which induces a drift in concentration that is most rapid for those tracers having the lowest diffusivities. If tracers are to be correctly conserved, either the kinematic boundary condition must explicitly allow advection across the interface, or the flux boundary condition must parameterize the effects of both vertical advection and diffusion in the boundary layer. In practice, the latter alternative is often implemented, although this is rarely done for all tracers.},
number = {1},
urldate = {2018-08-29},
journal = {Journal of Physical Oceanography},
author = {Jenkins, Adrian and Hellmer, Hartmut H. and Holland, David M.},
month = jan,
year = {2001},
pages = {285--296},
file = {Full Text PDF:/Users/cbegeman/Zotero/storage/LPXMZJJN/Jenkins et al. - 2001 - The Role of Meltwater Advection in the Formulation.pdf:application/pdf;Snapshot:/Users/cbegeman/Zotero/storage/HMPTGF4Y/1520-0485(2001)0310285TROMAI2.0.html:text/html},
}