Show simple item record

Active Layer Groundwater Flow: The Interrelated Effects of Stratigraphy, Thaw, and Topography

dc.contributor.authorO’Connor, Michael T.
dc.contributor.authorCardenas, M. Bayani
dc.contributor.authorNeilson, Bethany T.
dc.contributor.authorNicholaides, Kindra D.
dc.contributor.authorKling, George W.
dc.date.accessioned2019-10-30T15:30:34Z
dc.date.availableWITHHELD_11_MONTHS
dc.date.available2019-10-30T15:30:34Z
dc.date.issued2019-08
dc.identifier.citationO’Connor, Michael T.; Cardenas, M. Bayani; Neilson, Bethany T.; Nicholaides, Kindra D.; Kling, George W. (2019). "Active Layer Groundwater Flow: The Interrelated Effects of Stratigraphy, Thaw, and Topography." Water Resources Research 55(8): 6555-6576.
dc.identifier.issn0043-1397
dc.identifier.issn1944-7973
dc.identifier.urihttps://hdl.handle.net/2027.42/151862
dc.description.abstractThe external drivers and internal controls of groundwater flow in the thawed “active layer” above permafrost are poorly constrained because they are dynamic and spatially variable. Understanding these controls is critical because groundwater can supply solutes such as dissolved organic matter to surface water bodies. We calculated steady‐state three‐dimensional suprapermafrost groundwater flow through the active layer using measurements of aquifer geometry, saturated thickness, and hydraulic properties collected from two major landscape types over time within a first‐order Arctic watershed. The depth position and thickness of the saturated zone is the dominant control of groundwater flow variability between sites and during different times of year. The effect of water table depth on groundwater flow dwarfs the effect of thaw depth. In landscapes with low land‐surface slopes (2–4%), a combination of higher water tables and thicker, permeable peat deposits cause relatively constant groundwater flows between the early and late thawed seasons. Landscapes with larger land‐surface slopes (4–10%) have both deeper water tables and thinner peat deposits; here the commonly observed permeability decrease with depth is more pronounced than in flatter areas, and groundwater flows decrease significantly between early and late summer as the water table drops. Groundwater flows are also affected by microtopographic features that retain groundwater that could otherwise be released as the active layer deepens. The dominant sources of groundwater, and thus dissolved organic matter, are likely wet, flatter regions with thick organic layers. This finding informs fluid flow and solute transport dynamics for the present and future Arctic.Plain Language SummaryGroundwater flow in permafrost watersheds is potentially a key component of global carbon budgets because permafrost soil stores vast amounts of carbon that could be mobilized due to a warming climate and the corresponding increase in soil thaw. In addition to carrying carbon, groundwater can supply important nutrients and solutes to surface waters. However, we do not yet understand the factors that control groundwater flow in soils above permafrost because saturation changes rapidly and continuously, and soil hydraulic properties are largely unknown. We created measurement‐informed calculations of groundwater flow from areas of permafrost with different characteristics and found that soil types, which vary based on the slope of the land surface, are the most important control. Near‐surface soils were identical in hillslopes and valleys, whereas deeper soils in hillslopes allowed for less groundwater flow than in valleys. In early summer, when only the near‐surface soils were thawed, groundwater flows in the hillslopes and valley were similar. In late summer, when the deeper soil was thawed, groundwater flow in the valley remained high, but flow in the hillslope was negligible. Our observations also showed that small mounds on the land surface caused groundwater to be trapped behind underground ice dams.Key PointsDetailed measurements of hydraulic head, hydraulic conductivity, and saturated thicknesses in active layers were made over time and spaceThree main soil layers consistently comprise the stratigraphy of the active layer across the studied Arctic watershedGroundwater flow depends most on the depth of the water table and the subsurface stratigraphy, which varies based on landscape type
dc.publisherAcademic Press
dc.publisherWiley Periodicals, Inc.
dc.subject.otherpermafrost
dc.subject.othermicrotopography
dc.subject.otherpermeability
dc.subject.othergroundwater
dc.subject.otherArctic
dc.subject.otherbaseflow
dc.titleActive Layer Groundwater Flow: The Interrelated Effects of Stratigraphy, Thaw, and Topography
dc.typeArticle
dc.rights.robotsIndexNoFollow
dc.subject.hlbsecondlevelNatural Resources and Environment
dc.subject.hlbtoplevelScience
dc.description.peerreviewedPeer Reviewed
dc.description.bitstreamurlhttps://deepblue.lib.umich.edu/bitstream/2027.42/151862/1/wrcr24085_am.pdf
dc.description.bitstreamurlhttps://deepblue.lib.umich.edu/bitstream/2027.42/151862/2/wrcr24085.pdf
dc.identifier.doi10.1029/2018WR024636
dc.identifier.sourceWater Resources Research
dc.identifier.citedreferenceSmith, S. L., Wolfe, S. A., Riseborough, D. W., & Nixon, F. M. ( 2009 ). Active‐layer characteristics and summer climatic indices, Mackenzie Valley, Northwest Territories, Canada. Permafrost and Periglacial Processes, 20 ( 2 ), 201 – 220. https://doi.org/10.1002/ppp.651
dc.identifier.citedreferenceQuinton, W. L., Carey, S. K., & Goeller, N. T. ( 2004 ). Snowmelt runoff from northern alpine tundra hillslopes: Major processes and methods of simulation. Hydrology and Earth System Sciences, 8 ( 5 ), 877 – 890. https://doi.org/10.5194/hess‐8‐877‐2004
dc.identifier.citedreferenceQuinton, W. L., & Gray, D. M. ( 2003 ). Subsurface drainage from organic soils in permafrost terrain: The major factors to be represented in a runoff model. In Proceedings of the Eighth International Conference on Permafrost, Davos, Switzerland, 6pp.
dc.identifier.citedreferenceQuinton, W. L., Gray, D. M., & Marsh, P. ( 2000 ). Subsurface drainage from hummock‐covered hillslopes in the Arctic tundra. Journal of Hydrology, 237 ( 1–2 ), 113 – 125. https://doi.org/10.1016/S0022‐1694(00)00304‐8
dc.identifier.citedreferenceQuinton, W. L., Hayashi, M., & Carey, S. K. ( 2008 ). Peat hydraulic conductivity in cold regions and its relation to pore size and geometry. Hydrological Processes, 22 ( 15 ), 2829 – 2837. https://doi.org/10.1002/hyp.7027
dc.identifier.citedreferenceQuinton, W. L., Shirazi, T., Carey, S. K., & Pomeroy, J. W. ( 2005 ). Soil water storage and active‐layer development in a sub‐alpine tundra hillslope, southern Yukon Territory, Canada. Permafrost and Periglacial Processes, 16 ( 4 ), 369 – 382. https://doi.org/10.1002/ppp.543
dc.identifier.citedreferenceRoulet, N. T., Ash, R., Quinton, W., & Moore, T. ( 1993 ). Methane flux from drained northern peatlands: Effect of a persistent water table lowering on flux. Global Biogeochemical Cycles, 7 ( 4 ), 749 – 769. https://doi.org/10.1029/93GB01931
dc.identifier.citedreferenceRushlow, C. R., & Godsey, S. E. ( 2017 ). Rainfall–runoff responses on Arctic hillslopes underlain by continuous permafrost, North Slope, Alaska, USA. Hydrological Processes, 31 ( 23 ), 4092 – 4106. https://doi.org/10.1002/hyp.11294
dc.identifier.citedreferenceSchramm, I., Boike, J., Bolton, W. R., & Hinzman, L. D. ( 2007 ). Application of TopoFlow, a spatially distributed hydrological model, to the Imnavait Creek watershed, Alaska. Journal of Geophysical Research, 112, G04S46. https://doi.org/10.1029/2006JG000326
dc.identifier.citedreferenceSchuh, C., Frampton, A., & Christiansen, H. H. ( 2017 ). Soil moisture redistribution and its effect on inter‐annual active layer temperature and thickness variations in a dry loess terrace in Adventdalen, Svalbard. The Cryosphere, 11 ( 1 ), 635 – 651. https://doi.org/https://doi.org/10.5194/tc‐11‐635‐2017
dc.identifier.citedreferenceStieglitz, M., Shaman, J., McNamara, J., Engel, V., Shanley, J., & Kling, G. W. ( 2003 ). An approach to understanding hydrologic connectivity on the hillslope and the implications for nutrient transport. Global Biogeochemical Cycles, 17 ( 4 ), 1105. https://doi.org/10.1029/2003GB002041
dc.identifier.citedreferenceSurridge, B. W. J., Baird, A. J., & Heathwaite, A. L. ( 2005 ). Evaluating the quality of hydraulic conductivity estimates from piezometer slug tests in peat. Hydrological Processes, 19 ( 6 ), 1227 – 1244. https://doi.org/10.1002/hyp.5653
dc.identifier.citedreferenceTrexler, J. C., & Travis, J. ( 1993 ). Nontraditional regression analyses. Ecology, 74 ( 6 ), 1629 – 1637. https://doi.org/10.2307/1939921
dc.identifier.citedreferenceTromp‐van Meerveld, H. J., & McDonnell, J. J. ( 2006 ). Threshold relations in subsurface stormflow: 2. The fill and spill hypothesis. Water Resources Research, 42, W02411. Retrieved from. https://doi.org/10.1029/2004WR003800/full
dc.identifier.citedreferencevan Asselen, S., Stouthamer, E., & van Asch, T. W. J. ( 2009 ). Effects of peat compaction on delta evolution: A review on processes, responses, measuring and modeling. Earth‐Science Reviews, 92 ( 1–2 ), 35 – 51. https://doi.org/10.1016/j.earscirev.2008.11.001
dc.identifier.citedreferenceVoytek, E., Rushlow, C., Godsey, S., & Singha, K. ( 2016 ). Identifying hydrologic flowpaths on Arctic hillslopes using electrical resistivity and self potential. Geophysics, 81 ( 1 ), WA225 – WA232. https://doi.org/10.1190/geo2015‐0172.1
dc.identifier.citedreferenceWahrhaftig, C. ( 1965 ). Physiographic divisions of Alaska. Washington, DC: US Government Printing Office.
dc.identifier.citedreferenceWalker, D. A., & Everett, K. R. ( 1991 ). Loess ecosystems of Northern Alaska: Regional gradient and toposequence at Prudhoe Bay. Ecological Monographs, 61 ( 4 ), 437 – 464. https://doi.org/10.2307/2937050
dc.identifier.citedreferenceWalker, D. A., Jia, G. J., Epstein, H. E., Raynolds, M. K., Chapin, F. S. III, Copass, C., Hinzman, L. D., Knudson, J. A., Maier, H. A., Michaelson, G. J., Nelson, F., Ping, C. L., Romanovsky, V. E., & Shiklomanov, N. ( 2003 ). Vegetation‐soil‐thaw‐depth relationships along a low‐Arctic bioclimate gradient, Alaska: Synthesis of information from the ATLAS studies. Permafrost and Periglacial Processes, 14 ( 2 ), 103 – 123. https://doi.org/10.1002/ppp.452
dc.identifier.citedreferenceWalker, D. A., & Walker, M. D. ( 1996 ). Terrain and vegetation of the Imnavait Creek watershed. In Landscape function and disturbance in Arctic tundra (pp. 73 – 108 ). New York: Springer. Retrieved from https://doi.org/10.1007/978‐3‐662‐01145‐4_4
dc.identifier.citedreferenceWalvoord, M. A., & Kurylyk, B. L. ( 2016 ). Hydrologic impacts of thawing permafrost—A review. Vadose Zone Journal, 15 ( 6 ). https://doi.org/10.2136/vzj2016.01.0010
dc.identifier.citedreferenceWalvoord, M. A., & Striegl, R. G. ( 2007 ). Increased groundwater to stream discharge from permafrost thawing in the Yukon River basin: Potential impacts on lateral export of carbon and nitrogen. Geophysical Research Letters, 34, L12402. https://doi.org/10.1029/2007GL030216
dc.identifier.citedreferenceWang, H. F., & Anderson, M. P. ( 1995 ). Introduction to groundwater modeling: Finite difference and finite element methods. London: Academic Press.
dc.identifier.citedreferenceWoo, M. ( 2012 ). Permafrost hydrology. Berlin Heidelberg: Springer‐Verlag. Retrieved from. https://www.springer.com/us/book/9783642234613, https://doi.org/10.1007/978‐3‐642‐23462‐0
dc.identifier.citedreferenceWoo, M., & Steer, P. ( 1983 ). Slope hydrology as influenced by thawing of the active layer, Resolute, N.W.T. Canadian Journal of Earth Sciences, 20 ( 6 ), 978 – 986. https://doi.org/10.1139/e83‐087
dc.identifier.citedreferenceWright, N., Hayashi, M., & Quinton, W. L. ( 2009 ). Spatial and temporal variations in active layer thawing and their implication on runoff generation in peat‐covered permafrost terrain. Water Resources Research, 45, W05414. https://doi.org/10.1029/2008WR006880
dc.identifier.citedreferenceZhang, Y. ( 2014 ). Thermal‐hydro‐mechanical model for freezing and thawing of soils. Retrieved from http://deepblue.lib.umich.edu/handle/2027.42/108828
dc.identifier.citedreferenceAtchley, A. L., Painter, S. L., Harp, D. R., Coon, E. T., Wilson, C. J., Liljedahl, A. K., & Romanovsky, V. E. ( 2015 ). Using field observations to inform thermal hydrology models of permafrost dynamics with ATS (v0.83). Geoscientific Model Development, 8 ( 9 ), 2701 – 2722. https://doi.org/10.5194/gmd‐8‐2701‐2015
dc.identifier.citedreferenceBeckwith, C. W., Baird, A. J., & Heathwaite, A. L. ( 2003 ). Anisotropy and depth‐related heterogeneity of hydraulic conductivity in a bog peat. I: Laboratory measurements. Hydrological Processes, 17 ( 1 ), 89 – 101. https://doi.org/10.1002/hyp.1116
dc.identifier.citedreferenceBlaen, P. J. ( 2013 ). Water source dynamics of high Arctic river basins. Hydrological Processes, 28 ( 10 ), 3521 – 3538. https://doi.org/10.1002/hyp.9891
dc.identifier.citedreferenceBockheim, J. G. ( 2007 ). Importance of cryoturbation in redistributing organic carbon in permafrost‐affected soils. Soil Science Society of America Journal, 71 ( 4 ), 1335. https://doi.org/10.2136/sssaj2006.0414N
dc.identifier.citedreferenceBouwer, H., & Rice, R. C. ( 1976 ). A slug test for determining hydraulic conductivity of unconfined aquifers with completely or partially penetrating wells. Water Resources Research, 12 ( 3 ), 423 – 428. https://doi.org/10.1029/WR012i003p00423
dc.identifier.citedreferenceCardenas, M. B. ( 2010 ). Lessons from and assessment of Boussinesq aquifer modeling of a large fluvial island in a dam‐regulated river. Advances in Water Resources, 33 ( 11 ), 1359 – 1366. https://doi.org/10.1016/j.advwatres.2010.03.015
dc.identifier.citedreferenceCarman, P. C. ( 1956 ). Flow of gases through porous media. New York: Academic Press.
dc.identifier.citedreferenceDetterman, R. L., Bowsher, A. L., & Dutro, J. T. ( 1958 ). Glaciation on the Arctic Slope of the Brooks Range, Northern Alaska. Arctic, 11 ( 1 ), 43 – 61. https://doi.org/10.14430/arctic3732
dc.identifier.citedreferenceEbel, B. A., Koch, J. C., & Walvoord, M. A. ( 2019 ). Soil physical, hydraulic, and thermal properties in Interior Alaska, USA: Implications for hydrologic response to thawing permafrost conditions. Water Resources Research, 55, 4427 – 4447. https://doi.org/10.1029/2018WR023673
dc.identifier.citedreferenceEvans, S., & Ge, S. ( 2017 ). Contrasting hydrogeologic responses to warming in permafrost and seasonally frozen ground hillslopes. Geophysical Research Letters, 44, 1803 – 1813. https://doi.org/10.1002/2016GL072009
dc.identifier.citedreferenceFrampton, A., Painter, S., Lyon, S. W., & Destouni, G. ( 2011 ). Non‐isothermal, three‐phase simulations of near‐surface flows in a model permafrost system under seasonal variability and climate change. Journal of Hydrology, 403 ( 3–4 ), 352 – 359. https://doi.org/10.1016/j.jhydrol.2011.04.010
dc.identifier.citedreferenceFreeze, R. A., & Cherry, J. A. ( 1979 ). Groundwater. Englewood Cliffs, N.J: Prentice‐Hall.
dc.identifier.citedreferenceFrey, K. E., & McClelland, J. W. ( 2009 ). Impacts of permafrost degradation on Arctic river biogeochemistry. Hydrological Processes, 23 ( 1 ), 169 – 182. https://doi.org/10.1002/hyp.7196
dc.identifier.citedreferenceGe, S., McKenzie, J., Voss, C., & Wu, Q. ( 2011 ). Exchange of groundwater and surface‐water mediated by permafrost response to seasonal and long term air temperature variation. Geophysical Research Letters, 38, L14402. https://doi.org/10.1029/2011GL047911
dc.identifier.citedreferenceHamilton, T. D. ( 1982 ). A late Pleistocene glacial chronology for the southern Brooks Range: Stratigraphic record and regional significance. GSA Bulletin, 93 ( 8 ), 700 – 716. https://doi.org/10.1130/0016‐7606(1982)93<700:ALPGCF>2.0.CO;2
dc.identifier.citedreferenceHinkel, K. M., & Nelson, F. E. ( 2003 ). Spatial and temporal patterns of active layer thickness at Circumpolar Active Layer Monitoring (CALM) sites in northern Alaska, 1995–2000. Journal of Geophysical Research, 108 ( D2 ), 8168. https://doi.org/10.1029/2001JD000927
dc.identifier.citedreferenceHinzman, L. D., Kane, D. L., Gieck, R. E., & Everett, K. R. ( 1991 ). Hydrologic and thermal properties of the active layer in the Alaskan Arctic. Cold Regions Science and Technology, 19 ( 2 ), 95 – 110. https://doi.org/10.1016/0165‐232X(91)90001‐W
dc.identifier.citedreferenceHobbie, J., & Kling, G. ( 2014 ). Alaska’s changing Arctic: Ecological consequences for tundra, streams, and lakes (1 edition). Oxford; New York: Oxford University Press.
dc.identifier.citedreferenceHolden, J., & Burt, T. P. ( 2003 ). Hydrological studies on blanket peat: The significance of the acrotelm‐catotelm model. Journal of Ecology, 91 ( 1 ), 86 – 102. https://doi.org/10.1046/j.1365‐2745.2003.00748.x
dc.identifier.citedreferenceKoch, J., Kikuchi, C., Wickland, K., & Schuster, P. ( 2014 ). Runoff sources and flow paths in a partially burned, upland boreal catchment underlain by permafrost. Water Resources Research, 50, 8141 – 8158. https://doi.org/10.1002/2014WR015586
dc.identifier.citedreferenceKurylyk, B., Hayashi, M., Quinton, W., McKenzie, J., & Voss, C. ( 2016 ). Influence of vertical and lateral heat transfer on permafrost thaw, peatland landscape transition, and groundwater flow. Water Resources Research, 52, 1286 – 1305. https://doi.org/10.1002/2015WR018057
dc.identifier.citedreferenceLawrence, D. M., Slater, A. G., & Swenson, S. C. ( 2011 ). Simulation of present‐day and future permafrost and seasonally frozen ground conditions in CCSM4. Journal of Climate, 25 ( 7 ), 2207 – 2225. https://doi.org/10.1175/JCLI‐D‐11‐00334.1
dc.identifier.citedreferenceLiljedahl, A., & Hinzman, L. ( 2012 ). Ice‐wedge polygon type controls low‐gradient watershed‐scale hydrology. In Tenth International Conference on Permafrost Vol. 1: International Contributions, (Vol. 1, pp. 231 – 236 ). Salekhard, Russia: The Northern Publisher.
dc.identifier.citedreferenceLiu, Y., Tong, J., & Li, X. ( 2005 ). Analysing the silt particles with the Malvern Mastersizer 2000. Water Conservancy Science and Technology and Economy, 11 ( 6 ), 329 – 331.
dc.identifier.citedreferenceMcNamara, J. P., Kane, D. L., & Hinzman, L. D. ( 1997 ). Hydrograph separations in an Arctic watershed using mixing model and graphical techniques. Water Resources Research, 33 ( 7 ), 1707 – 1719. https://doi.org/10.1029/97WR01033
dc.identifier.citedreferenceMcNamara, J. P., Kane, D. L., & Hinzman, L. D. ( 1999 ). An analysis of an Arctic channel network using a digital elevation model. Geomorphology, 29 ( 3‐4 ), 339 – 353. https://doi.org/10.1016/S0169‐555X(99)00017‐3
dc.identifier.citedreferenceMerck, M. F., Neilson, B. T., Cory, R. M., & Kling, G. W. ( 2012 ). Variability of in‐stream and riparian storage in a beaded Arctic stream. Hydrological Processes, 26 ( 19 ), 2938 – 2950. https://doi.org/10.1002/hyp.8323
dc.identifier.citedreferenceMichaelson, G. J., Ping, C. L., & Kimble, J. M. ( 1996 ). Carbon storage and distribution in tundra soils of Arctic Alaska, U.S.A. Arctic and Alpine Research, 28 ( 4 ), 414 – 424. https://doi.org/10.2307/1551852
dc.identifier.citedreferenceMorris, P. J., Waddington, J. M., Benscoter, B. W., & Turetsky, M. R. ( 2011 ). Conceptual frameworks in peatland ecohydrology: Looking beyond the two‐layered (acrotelm–catotelm) model. Ecohydrology, 4 ( 1 ), 1 – 11. https://doi.org/10.1002/eco.191
dc.identifier.citedreferenceNeilson, B. T., Cardenas, M. B., O’Connor, M. T., Rasmussen, M. T., King, T. V., & Kling, G. W. ( 2018 ). Groundwater flow and exchange across the land surface explain carbon export patterns in continuous permafrost watersheds. Geophysical Research Letters, 45, 7596 – 7605. https://doi.org/10.1029/2018GL078140
dc.identifier.citedreferenceNelson, F. E., Shiklomanov, N. I., & Mueller, G. R. ( 1999 ). Variability of active‐layer thickness at multiple spatial scales, North‐Central Alaska, U.S.A. Arctic, Antarctic, and Alpine Research, 31 ( 2 ), 179 – 186. https://doi.org/10.2307/1552606
dc.identifier.citedreferenceOmernik, J. M., & Griffith, G. E. ( 2014 ). Ecoregions of the Conterminous United States: Evolution of a hierarchical spatial framework. Environmental Management, 54 ( 6 ), 1249 – 1266. https://doi.org/10.1007/s00267‐014‐0364‐1
dc.identifier.citedreferenceOsterkamp, T. E., & Payne, M. W. ( 1981 ). Estimates of permafrost thickness from well logs in northern Alaska. Cold Regions Science and Technology, 5 ( 1 ), 13 – 27. https://doi.org/10.1016/0165‐232X(81)90037‐9
dc.identifier.citedreferencePing, C.‐L., Michaelson, G. J., Jorgenson, M. T., Kimble, J. M., Epstein, H., Romanovsky, V. E., & Walker, D. A. ( 2008 ). High stocks of soil organic carbon in the North American Arctic region. Nature Geoscience, 1 ( 9 ), 615 – 619. https://doi.org/10.1038/ngeo284
dc.identifier.citedreferencePomeroy, J. W., Gray, D. M., Brown, T., Hedstrom, N. R., Quinton, W. L., Granger, R. J., & Carey, S. K. ( 2007 ). The cold regions hydrological model: A platform for basing process representation and model structure on physical evidence. Hydrological Processes, 21 ( 19 ), 2650 – 2667. https://doi.org/10.1002/hyp.6787
dc.owningcollnameInterdisciplinary and Peer-Reviewed


Files in this item

Show simple item record

Remediation of Harmful Language

The University of Michigan Library aims to describe library materials in a way that respects the people and communities who create, use, and are represented in our collections. Report harmful or offensive language in catalog records, finding aids, or elsewhere in our collections anonymously through our metadata feedback form. More information at Remediation of Harmful Language.

Accessibility

If you are unable to use this file in its current format, please select the Contact Us link and we can modify it to make it more accessible to you.