Next Article in Journal
Deformation of the European Plate (58-0 Ma): Evidence from Calcite Twinning Strains
Next Article in Special Issue
Temperature Variation during Salt Migration in Frozen Hydrate-Bearing Sediments: Experimental Modeling
Previous Article in Journal
Dynamic Numerical Simulations of Dry-Stone Retaining Walls: Identification of the Seismic Behaviour Factor
Previous Article in Special Issue
Simulating Thermal Interaction of Gas Production Wells with Relict Gas Hydrate-Bearing Permafrost
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Changes in Unfrozen Water Contents in Warming Permafrost Soils

1
Center for Petroleum Science and Engineering, Skolkovo Institute of Science and Technology, 30, Build. 1, Bolshoi Boulevard, 121205 Moscow, Russia
2
Laboratory for Integrated Research of the Sea-Land Arctic System, Tomsk State University, 36, Lenin Avenue, 634050 Tomsk, Russia
*
Author to whom correspondence should be addressed.
Submission received: 8 May 2022 / Revised: 9 June 2022 / Accepted: 13 June 2022 / Published: 17 June 2022
(This article belongs to the Special Issue Permafrost and Gas Hydrate Response to Ground Temperature Rising)

Abstract

:
Climate warming in the Arctic, accompanied by changes in permafrost soil properties (mechanical, thermal, filtration, geophysical), is due to increasing unfrozen pore water content. The liquid component in frozen soils is an issue of key importance for permafrost engineering that has been extensively studied since the beginning of the 20th century. We suggest a synthesis and new classification of various experimental and calculation methods for the determination of unfrozen water content. Special focus is placed on the method of applying measurements to the water potential, which reveals the impact of permafrost warming on unfrozen water content. This method was applied to natural soil samples collected from shallow permafrost from northern West Siberia affected by climate change, and confirms the revealed trends. The obtained results confirm that unfrozen water content is sensitive not only temperature but also particle size distribution, salinity, and the organic matter content of permafrost soils.

1. Introduction

The rapid development of permafrost territories worldwide has caused problems associated with climate-induced permafrost warming, especially in recent decades. The warming and degrading of permafrost substantially weakens its bearing capacity and poses risks to the stability and safe operation of engineering constructions [1,2,3,4,5]. Although the temperature of the permafrost in the northern hemisphere reaches −12 to −10 °C, some parts of the pore moisture remain unfrozen [6]. The pore moisture of frozen ground consists of ice, unfrozen water, and a small amount of gas (including water vapor) that has only a minor influence on permafrost properties. The phase equilibrium in the pore moisture changes under climatic temperature effects: ice content increases upon cooling, and the amount of unfrozen water increases upon warming. The changing percentages of solid (ice) and liquid (water) components of pore moisture affect physical and chemical processes in the permafrost and its related thermal and mechanical properties [7]. Ongoing climate warming in the Arctic leads to disturbed ecosystems, and induces a slow warming of ground that has been frozen for hundreds of years or longer [8].
Permafrost retains an enormous resource of coldness, and is unlikely to thaw to a catastrophic extent in the coming decades. Nevertheless, the continuous warming of shallow permafrost may change the phase composition of pore moisture by increasing the amount of unfrozen water, causing the solid permafrost to adopt a more plastic behavior. The unfrozen water content may further increase following thermal impact from buildings and facility constructions, drilling operations, pipelines and others [9]. Therefore, it is critical to develop advanced methods and instruments for the prediction of possible permafrost responses to increasing unfrozen water content as a result of global warming, to ensure the stability of buildings and structures in the Arctic [10,11,12,13].

2. Materials for Estimating Unfrozen Water Content in Permafrost

Experiments in the early 20th century showed that some amount of pore water in the ground exposed to negative temperatures remained unfrozen [14,15]. The idea that some pore moisture in artificially frozen soil and in natural permafrost may remain liquid appeared in the handbook “General Geocryology” [16]. Later experiments by Tsitovich laid the foundation for the phase equilibrium principle of pore water in frozen ground [17,18]: frozen ground at any low negative temperature always contains some residual liquid water, the amount of which varies as a function of temperature and pressure.
Further instrumental advances in the 1950–1970s provided a wealth of experimental constraints on the sensitivity of pore moisture phase composition and grain size, mineralogy, and salinity of permafrost. Later, it was found that unfrozen water in the pores of frozen ground could exist in capillary, film, adsorbed, or other forms [19,20,21]. Its content was studied in terms of its effect on the thermal, mechanical and electrical properties of permafrost soils [22,23,24,25,26,27,28,29,30,31,32], as well as on various physical and chemical processes [18,33,34,35,36,37,38]. Research along these lines by many teams from Russia, China, Canada, Japan, Germany and the USA continues.
Experimental research in the 1960s, 1970s and later was complemented by modeling based on empirical equations that related the amount of unfrozen pore water in permafrost to other variables. Correspondingly, unfrozen water content was estimated from such parameters as plastic limit, temperature [39] and salinity [40], according to the Design Standards currently used in Russia (equivalent to the standards of the American Society for Testing and Materials or ASTM); pore pressure in freezing soils [41]; specific particle surface area [42,43,44]; liquid limit [45]; specific volume of water, partial specific heat capacity, apparent specific heat capacity, and total moisture content [46]; total moisture and temperature of soil and freezing point of pore water [47]; liquid limit moisture for saline soil [48]; freezing point and unfreezable water content [49]; moisture and grain size of soil [50], and potential (activity) of pore water [51].
The available methods for estimating the phase composition of pore moisture in frozen ground have been classified in different ways according to approaches to its determination [19,27,34,52]. We suggest an updated classification (Table 1), which combines all previous ones and divides the methods into experiments (direct measurements), modeling, and experiments + modeling (calculations based on measurements).
The methods of the first group include cryoscopy, hygroscopy, sublimation, and contact measurements of unfrozen water content. In the calorimetry, NMR, TDR, thermometry, desorption, and water potential methods (second group), water content is calculated from experimentally measured parameters. Modeling means the estimation of unfrozen water content using a correlation with other soil properties (e.g., grain size, specific surface area, salinity, plastic and liquid limits, freezing point, etc.).
The applicability of different methods depends on the lithology and temperature range of the soil, as well as on the required instrumental facilities and time, but most of these are labor- and time-consuming. Meanwhile, active construction on permafrost in Russia, China, Canada, and the US requires high-quality and high-performance engineering surveys and new or updated techniques for estimating the amount of unfrozen pore water.

3. Water Potential Method and Its Advantages

The water potential method is advantageous for estimating the amount of unfrozen pore water in permafrost [74,75,76,77,78]. It consists of measurements of pore water potential (or water activity) in samples with stepwise decreasing water content and subsequent thermodynamic calculations using measured data. The instruments used in this method provide automatic monitoring of measured water potential, which saves time and allows express measurements [78]. The method was applied to laboratory soil samples across a large range of negative temperatures reaching −15 °C and to natural samples collected during geological engineering surveys in northern West Siberia. The results agree well with data from other methods, including NMR and contact measurements, which are especially widespread in Russia [78].
The activity of pore water can be measured on a WP4 water potential meter (Pullman, WA, USA) [79] designed by the METER Group (formerly Decagon Devices). The details of this method are described in detail by Campbell et al. [79], and the thermodynamic background is provided by Istomin et al. [80]. The system (Figure 1) is based on sampling the pressure of water vapor over wet soil using the dew-point method.
The temperature dependence of unfrozen water content in frozen soil is obtained in several runs at different total moisture values measured by weighing the soil samples on analytical scales, to a precision of ±0.003 g, before and after water potential measurements are taken. First, the pore water potential ( ψ ) is measured in samples with known moisture content, and the results are converted to activity ( α ) which is further converted to the temperature of ice—pore water equilibrium or the equilibrium temperature ( t e q , °C)—which actually corresponds to the freezing point at the given total moisture.
The pore water activity in the range from 0.6–0.7 to 1.0 is related to the equilibrium temperature ( t e q , °C) as [75].
t e q = 103.25 ln α + 5.57 1 α 2 ,
Water activity ( α ) can be estimated by direct measurement at +25 °C or by thermodynamic calculations, as per [78].
ln α = ψ M R T = 0.0073 ψ ,
The measurements of pore water potential begin in a soil sample with initial natural moisture or with pore moisture versus the total moisture (for laboratory-made samples). While the sample is being dried, the potential is measured over six or seven runs at each step of progressively decreasing content of unfrozen water. The acquisition commonly takes 20 to 30 min (60 min, the longest, for dry mud), which is the time required for soil moisture and vapor-bearing air in the sample chamber to reach full equilibrium.

4. Main Controls of Unfrozen Water Content in Permafrost

The amount of unfrozen pore water depends on the temperature and composition of permafrost soil, including grain size, mineralogy, salinity, organic matter content, etc. Temperature-dependent variations of unfrozen water content can be plotted as in Figure 2. The geometry of the curve is approximately the same (Figure 2) at any lithology, mineralogy, salinity, peat content, and other properties of permafrost, and will vary only in the slope and specific values along the axes. At some subzero temperature ( t f 1 ~ 0 °C), the amount of unfrozen pore water may reach a point ( W 0 W 1 ) at which the phase composition of pore moisture no longer changes, because almost all the pore ice has already melted. In this case, permafrost will behave like unfrozen soil. However, it is at a negative temperature, and will lose the mechanical strength that was provided by the ice cement at lower temperatures ( t f 2 ,   t f 3 ,   t f 4 ). Furthermore, the amount of residual pore water in frozen soil is proportional to its salinity and to the content of organic matter and clay particles, which has been confirmed in several studies [19,78,81,82].
The sensitivity of unfrozen water content to organic matter (peat) content can be illustrated with silt–clay soil samples from northern West Siberia. The samples (Table 2) share similarity in grain size and mineralogy, consisting of 40% quartz and clay minerals (montmorillonite, kaolinite, hydromica, chlorite and mixed-layer groups). They are non-saline but contain different percentages of organic matter (peat). More details of these soils have been reported previously [78].
The natural soils of northern West Siberia with greater peat percentages ( I r , %) contain more liquid pore water, other things being equal (Figure 3). Specifically, unfrozen water content at −5 °C is ~3.5% in peat-free lean clay ( I r   = 0.4%), 5% at I r   = 2.7%, and up to 6% and 8% at the peat content of I r   = 3.1% and 6.2%, respectively.
The samples were collected from −5 to −6 °C permafrost. Its unfrozen pore water content increases only slightly if the permafrost becomes 1–2 °C warmer (Figure 3) but may reach notably greater percentages upon further warming toward 0 °C, especially at greater peat content. The permafrost with the highest percentages of peat (6.2%) will contain 13–15% of unfrozen pore water at −0.5 to −1 °C (Figure 3).
The content of unfrozen pore water is also sensitive to the salinity of permafrost (Figure 4), expressed as dry-weight salt percentage ( D s a l ). The salinity dependence was studied for saline lean clay and sand soils interacting with seawater, from marine terrace I in northern West Siberia [82].
The unfrozen water content increases proportionally with salinity in both sand and lean clay samples. Upon warming from −10 °C to −2 °C, the increase is moderate in non-saline soils but increases by several times at maximum salinity (0.8% in sand and 1% in lean clay), from 2.5% to 12.5% in sand and from 10% to 25% in lean clay.

5. Discussion

Unfrozen water is a ubiquitous component of interstitial moisture in permafrost. It may occur as lenses or layers of brines called cryopegs, or as thin film around ice crystals and mineral particles. Locally, saline unfrozen water may fill pores and voids in the soil skeleton, making miniature cryopegs. The amount of unfrozen water is proportional to the organic matter (peat) content and to salinity, because saline water has a lower freezing point than fresh water. The permafrost soil gains mechanical strength and stability as its liquid pore moisture decreases upon cooling to lower negative temperatures and, vice versa, loses strength and acquires plastic properties at greater percentages of residual unfrozen water. Pore ice can melt under external effects, i.e., mechanical loading [25,84], where films of unfrozen water on mineral particles and ice crystals become thicker. In this case, frozen sediments behave like ductile material, with their mechanical properties approaching those of unfrozen rocks. Frozen ground can resist brief (even strong) impact but undergoes strain and creep under prolonged, increasing loads. Creep shows up as increasing plasticity and large strain following relatively small but long-lasting loading.
Additionally, in some cases, it is assumed that the existing highly mineralized intrapermafrost waters (cryopegs) are desalting via newly formed fresh waters that have arisen with an increase in the temperature of the ice-rich permafrost. This will possibly occur in the case of a hydraulic connection between different horizons of thawing permafrost.
The permafrost of the western part of the Russian Arctic has been experiencing such an effect from global warming since the 1970s (Figure 5).
The long-term warming of air temperatures in the Arctic correlates with the gradual warming of shallow permafrost (Figure 6) and the related increase of unfrozen water content at depths attainable by heat waves. The process cannot be classified as thawing because the frozen sediments do not become fully unfrozen, and remain at negative temperatures, despite containing greater amounts of unfrozen pore water. However, the permafrost is actually degrading.
The temperature patterns of permafrost are mainly characterized by their mean annual values measured at the depth of zero annual amplitudes, which is commonly 10–15 m in northern West Siberia. The 2.8 °C warming from 1970 to 2018 led to a mean annual increase of 0.056 °C/yr for permafrost colder than −3 °C in this Arctic region, and 0.04 °C/yr in Russia’s European north [85]. Thus, the low-temperature Arctic permafrost heats up faster and can persist longer than that further in the south, due to its larger cold resources, while the “southern” permafrost, with its subzero mean annual temperature, will degrade faster. Therefore, a smaller amount of heat is required to warm up the permafrost until it starts thawing. See some temperature logging data from West Siberia in Figure 6.
In general, permafrost warming is very slow, but its consequences will be felt even in areas that are free from anthropogenic effects, provided that the climate trends of recent decades persist for a long time into the future. Mean annual temperature trends of shallow permafrost can be predicted by modeling. For instance, a model for the Marre–Sale area (western Yamal Peninsula, northern West Siberia) predicts that changes in mean annual ground temperature may reach a depth of 50 m in the coming 50 years [86]. Specifically, the permafrost may become 1.3 °C and 23 °C warmer at depths of 20 m and 7–10 m, respectively, while its shallowest 5 m may warm up as much as above 0 °C.
However, temperature variations are controlled by the composition of permafrost, as well as the effect of deeply penetrating heat pulses. Our results, as well as published data [21,25,27,34,38,78,82], show that the amount of unfrozen pore water varies as a function of salinity and peat content: it is much greater in saline and peat-rich sediments (Figure 3 and Figure 4) than in non-saline rocks poor in organic matter. This fact must be taken into account in the operation of existing engineering structures and in new projects.
Climate-induced heat waves are especially expected to influence shallow permafrost, the effect being controlled by grain size, total moisture/ice content, salinity, and peat percentage. These data can be used to predict the response of permafrost to warming at a certain depth. For instance, probable changes can be inferred from the depth profiles of permafrost properties (up to 100 m), as made for the area of the Bovanenkovo gas field on the Yamal Peninsula (Figure 7). The sampled 20 m of Yamal permafrost is composed mainly of clay lying over ground ice that encloses soil material and has a sand layer at the base. Below 20 m, there follows clay silt with massive lens-like cryostructures. The permafrost temperature below 5 m varies from −4.0 to −5.5 °C and reaches a minimum at 40–60 m. The total moisture of soil remains almost invariable (30–40%) below 40 m, but ranges from 30 to 130% at shallower depths due to the presence of ground ice lenses and layers (Figure 7).
Deposition and glaciation in the area have produced quite complex depth-dependent salinity and organic matter patterns. However, the history of permafrost is much less important for the predictions of water–ice phase changes than the actual depth profiles of its variables. The presence of saline and organic-rich layers in the upper 10–20 m of permafrost, where warming pulses will be able to penetrate within a few decades, may cause a considerable increase in the amount of unfrozen pore water. Meanwhile, as shown by experiments, this amount may increase even upon a minor warming of such permafrost, which may make it ductile. The situation may be worse in the presence of zones with high total moisture (ice content). In this case, special permafrost stabilization operations should be carried out during the design of engineering constructions (thermosiphons, thermal insulation covers, cryogels etc.).

6. Conclusions

The design of engineering structures on permafrost and economic planning in Arctic areas require accounting for the trend of global warming in recent decades and the related slight warming of negative permafrost temperatures. The warming of permafrost may lead to changes in its properties due to its increasing content of unfrozen pore water, which may be unevenly distributed in the heterogeneous permafrost. Meanwhile, despite the heterogeneity, unfrozen water may exist across a large range of negative temperatures.
The liquid component of pore moisture has been a subject of research worldwide for about 100 years, but many aspects of the problem remain poorly understood, in particular the control of the permafrost phase composition. The suggested overview and classification of methods for estimating unfrozen water content in permafrost may be of interest to permafrost scientists. Such estimation can be performed successfully with thermodynamic calculations based on water potential measurements, which is a simple, fast, and high-performance tool.
The amount of unfrozen pore water is sensitive to the salinity and peat content of permafrost, as illustrated by data on the natural samples of saline and organic-rich soil from northern West Siberia. The revealed pore water and mean annual temperature variations in shallow permafrost from the Yamal Peninsula were used to predict possible phase changes in pore moisture in response to ongoing climate warming. These results can be used when designing engineering constructions and facility buildings for new gas fields in the north of Western Siberia (South Tambey, Salman, Kharasavey, etc.) to extend their service period and reduce geotechnical and environmental risk.

Author Contributions

Conceptualization, methodology, supervision, E.C.; investigation, E.C., N.S. and B.B.; writing—original draft preparation, E.C. and N.S.; writing—review and editing, E.C., N.S. and B.B.; visualization, E.C. and B.B. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Tomsk State University Development Program (Priority 2030).

Data Availability Statement

Not Applicable.

Acknowledgments

The authors are grateful to the Skolkovo Institute of Science and Technology for research support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jorgenson, M.T.; Racine, C.H.; Walters, J.C.; Osterkamp, T.E. Permafrost degradation and ecological changes associated with a warming climate in Central Alaska. Clim. Chang. 2001, 48, 551–579. [Google Scholar] [CrossRef]
  2. Wang, D.; Tighe, S.L.; Yin, S. Preliminary analysis of permafrost degradation in Ingraham Trail, Northwest Territories. In Proceedings of the Canadian Society of Civil Engineering Annual Conference 2021; CSCE 2021; Lecture Notes in Civil Engineering, 240; Springer: Singapore, 2023. [Google Scholar] [CrossRef]
  3. Maadani, O.; Shafiee, M.; Egorov, I. Climate change challenges for flexible pavement in Canada: An overview. J. Cold Reg. Eng. 2021, 35, 03121002. [Google Scholar] [CrossRef]
  4. Melnikov, V.P.; Osipov, V.I.; Brouchkov, A.V.; Falaleeva, A.A.; Badina, S.V.; Zheleznyak, M.N.; Sadurtdinov, M.R.; Ostrakov, N.A.; Drozdov, D.S.; Osokin, A.B.; et al. Climate warming and permafrost thaw in the Russian Arctic: Potential economic impacts on public infrastructure by 2050. Nat. Hazards 2022, 112, 231–251. [Google Scholar] [CrossRef]
  5. Hjort, J.; Karjalainen, O.; Aalto, J.; Westermann, S.; Romanovsky, V.E.; Nelson, F.E.; Etzelmüller, B.; Luoto, M. Degrading permafrost puts Arctic infrastructure at risk by mid-century. Nat. Commun. 2018, 9, 5147. [Google Scholar] [CrossRef] [PubMed]
  6. Romanovsky, V.E.; Drozdov, D.S.; Oberman, N.G.; Malkova, G.V.; Kholodov, A.L.; Marchenko, S.S.; Moskalenko, N.G.; Sergeev, D.O.; Ukraintseva, N.G.; Abramov, A.A.; et al. Thermal state of permafrost in Russia. Permafr. Periglac. Processes 2010, 21, 136–155. [Google Scholar] [CrossRef]
  7. Harris, S.A.; Brouchkov, A.; Goudong, C. Geocryology. Characteristics and Use of Frozen Ground and Permafrost Landforms; CRC Press, Taylor & Francis Group: London, UK, 2018; 765p. [Google Scholar]
  8. Macdonald, R.W.; Harner, T.; Fyfe, J. Recent climate change in the Arctic and its impact on contaminant pathways and interpretation of temporal trend data. Sci. Total Environ. 2005, 342, 5–86. [Google Scholar] [CrossRef]
  9. Yu, W.; Zhang, T.; Lu, Y.; Han, F.; Zhou, Y.; Hu, D. Engineering risk analysis in cold regions: State of the art and perspectives. Cold Reg. Sci. Technol. 2020, 171, 102963. [Google Scholar] [CrossRef]
  10. Peng, H.; Ma, W.; Mu, Y.L.; Jin, L. Impact of permafrost degradation on embankment deformation of Qinghai-Tibet Highway in permafrost regions. J. Cent. South Univ. 2015, 22, 1079–1086. [Google Scholar] [CrossRef]
  11. Kong, X.; Doré, G. Thermal stabilization of embankments built on thaw-sensitive permafrost. J. Cold Reg. Eng. 2019, 35, 04021010. [Google Scholar] [CrossRef]
  12. Buslaev, G.; Tsvetkov, P.; Lavrik, A.; Kunshin, A.; Loseva, E.; Sidorov, D. Ensuring the sustainability of Arctic industrial facilities under conditions of global climate change. Resources 2021, 10, 128. [Google Scholar] [CrossRef]
  13. Khrustalev, L.N. Fundamentals of Geotechnics in Cryolithozone; MSU Press: Moscow, Russia, 2005; 544p. (In Russian) [Google Scholar]
  14. Boyoucous, G.J. Degree of temperature to which soils can be cooled without freezing. J. Agric. Res. 1920, 20, 267–269. [Google Scholar] [CrossRef]
  15. Andrianov, P.I. The freezing point of soils. Akad. Nauk SSSR 1936, 1, 17–54. (In Russian) [Google Scholar]
  16. Sumgin, M.I.; Kachurin, S.P.; Tolstikhin, N.I.; Tumel, V.F. General permafrost science. Akad. Nauk SSSR 1940, 340. (In Russian) [Google Scholar]
  17. Tsytovich, N.A. On the theory of water equilibrium in frozen soils. Izv. AN SSSR Ser. Geogr. Geofiz. 1945, 9, 493–502. [Google Scholar]
  18. Are, F.E. Thermal aspects of Tsytovich principle of water and ice equilibrium state in frozen ground. Kriosf. Zemli 2014, XVIII, 47–56. [Google Scholar]
  19. Yershov, E.D.; Akimov, Y.P.; Cheverev, V.G.; Kuchukov, E.Z. Phase Composition of Pore Moisture in Permafrost; Moscow University Press: Moscow, Russia, 1979; 189p. (In Russian) [Google Scholar]
  20. Williams, P.J.; Smith, M.W. The Frozen Earth. Fundamentals of Geocryology; Carleton University: Ottawa, ON, Canada, 1989; 306p. [Google Scholar]
  21. Cheverev, V.G. Properties of bound water in cryogenic soils. Kriosf. Zemli 2003, VII, 30–41. [Google Scholar]
  22. Yershov, E.D. (Ed.) Fundamentals of Geocryology. Part 1. Physico-Chemical Geocryology; Moscow University Press: Moscow, Russia, 1995; 368p. (In Russian) [Google Scholar]
  23. Yershov, E.D. General Geocryology; Cambridge University Press: Cambridge, UK, 1998; 580p. [Google Scholar]
  24. Frolov, A.D. Electrical and Elastic Properties of Frozen Rocks and Ice; ONTI PNC RAN: Pushchino, Russia, 1998; 515p. (In Russian) [Google Scholar]
  25. Roman, L.T. Mechanic Properties of Frozen Ground; Nauka/Interperiodika: Moscow, Russia, 2002; 426p. (In Russian) [Google Scholar]
  26. Andersland, O.B.; Ladanyi, B. Frozen Ground Engineering; John Wiley: New York, NY, USA, 2003; 363p. [Google Scholar]
  27. Cheverev, V.G. Cryogenic Properties of Soils; Nauchniy Mir: Moscow, Russia, 2004; 234p. (In Russian) [Google Scholar]
  28. Sun, Q.J.; Li, D.W.; Wang, S.Y. Research on uniaxial compressive strength of frozen sand containing salt with different water contents. Coal Technol. 2015, 34, 86–88. [Google Scholar]
  29. Xu, X.T.; Wang, Y.B.; Bai, R.Q.; Fan, C.X.; Hua, S.G. Comparative studies on mechanical behavior of frozen natural saline silty sand and frozen desalted silty sand. Cold Reg. Sci. Technol. 2016, 132, 81–88. [Google Scholar] [CrossRef]
  30. Zhao, X.; Zhou, G.; Lu, G. Strain responses of frozen clay with thermal gradient under triaxial creep. Acta Geotech. 2017, 12, 183–193. [Google Scholar] [CrossRef]
  31. Tang, R.; Zhou, G.; Wang, J.; Zhao, G.; Lai, Z.; Jiu, F. A new method for estimating salt expansion in saturated saline soils during cooling based on electrical conductivity. Cold Reg. Sci. Technol. 2020, 170, 102943. [Google Scholar] [CrossRef]
  32. Liu, J.; Yang, P.; Yang, Z. Electrical properties of frozen saline clay and their relationship with unfrozen water content. Cold Reg. Sci. Technol. 2020, 178, 103127. [Google Scholar] [CrossRef]
  33. Grechishchev, S.E.; Chistotinov, L.V.; Shur, Y.L. Cryogenic Physical and Geological Processes and their Prediction; Nedra: Moscow, Russia, 1980; 384p. (In Russian) [Google Scholar]
  34. Komarov, I.A. Thermodynamics and Heat and Mass Transfer in Fine-Grained Porous Rocks; Nauchnyi Mir: Moscow, Russia, 2003; 608p. (In Russian) [Google Scholar]
  35. Starostin, E.G. Determination of the amount of unfrozen water, based on the crystallization kinetics. Kriosf. Zemli 2008, XII, 60–64. [Google Scholar]
  36. Liao, M.; Lai, Y.; Liu, E.; Wan, X. A fractional order creep constitutive model of warm frozen silt. Acta Geotech. 2017, 12, 377–389. [Google Scholar] [CrossRef]
  37. Kong, L.; Wang, Y.; Sun, W.; Qi, J. Influence of plasticity on unfrozen water content of frozen soils as determined by nuclear magnetic resonance. Cold Reg. Sci. Technol. 2020, 172, 102993. [Google Scholar] [CrossRef]
  38. Wang, L.; Liu, J.; Yu, X.; Wang, T.; Feng, R. A simplified model for the phase composition curve of saline soils considering the second phase transition. Water Resour. Res. 2021, 57, e2020WR028556. [Google Scholar] [CrossRef]
  39. Design Standards #II-B.6-66; Basements and Foundations of Buildings and Structures on Permafrost. Gosstroyizdat: Moscow, Russia, 1967; 33p. (In Russian)
  40. Working Document SNIP 2.02.04-88; Design Standards. Basements and Foundations on Permafrost. Gosstroy: Moscow, Russia, 2005; 52p. (In Russian)
  41. Koopmans, R.W.R.; Miller, R. Soil Freezing and Soil Water Characteristic Curves 1. Soil Sci. Soc. Am. J. 1966, 30, 680–685. [Google Scholar] [CrossRef]
  42. Dillon, H.B.; Andersland, O.B. Predicting unfrozen water contents in frozen soils. Can. Geotech. J. 1966, 3, 53–60. [Google Scholar] [CrossRef]
  43. Anderson, D.M.; Tice, A.R. Predicting unfrozen water contents in frozen soils from surface area measurements. Highw. Res. Rec. 1972, 393, 12–18. [Google Scholar]
  44. Anderson, D.M.; Tice, A.R.; McKim, H.L. The unfrozen water and the apparent specific heat capacity of frozen soils. In Proceedings of the Second International Conference on Permafrost, Jakutsk, Russia, 16–28 July 1973; North American Contribution National Academy of Science: Washington, DC, USA, 1973; pp. 289–294. [Google Scholar]
  45. Tice, A.R.; Anderson, D.M.; Banin, A. The Prediction of Unfrozen Water Contents in Frozen Soils from Liquid Limit Determinations; CRREL Report 79-8; Cold Regions Research and Engineering Laboratory: Hanover, NH, USA, 1976; 13p. [Google Scholar]
  46. Low, P.F.; Anderson, D.M.; Duwayne, M.; Hoekstra, P. Some thermodynamic relationships for soils at or below the freezing point. 1. Freezing point depression and heat capacity. Water Resour. Res. 1968, 4, 379–394. [Google Scholar] [CrossRef] [Green Version]
  47. Roman, L.T. Frozen Peat-Bearing Foundation Soils; Nauka: Novosibirsk, Russia, 1987; 222p. (In Russian) [Google Scholar]
  48. Sheikin, I.V. Calculated relationships between water-physical variables of permafrost. In Formation of Frozen Ground and Prediction of Cryogenic Processes; Nauka: Moscow, Russia, 1986; pp. 151–156. (In Russian) [Google Scholar]
  49. Kozlowski, T. A semi-empirical model for phase composition of water in clay-water systems. Cold Reg. Sci. Technol. 2007, 49, 226–236. [Google Scholar] [CrossRef]
  50. Qin, Y.; Li, G.; Qu, G. A formula for the unfrozen water content and temperature of frozen soils. In Proceedings of the 14th Conference on Cold Regions Engineering, Duluth, MN, USA, 31 August–2 September 2009; pp. 155–161. [Google Scholar] [CrossRef]
  51. Istomin, V.; Chuvilin, E.; Makhonina, N.; Bukhanov, B. Temperature dependence of unfrozen water content in sediments on the water potential measurements. Kriosf. Zemli 2009, XIII, 35–43. [Google Scholar]
  52. Yershov, E.D. (Ed.) Methods of Geocryological Research; Moscow University Press: Moscow, Russia, 2004; 512p. (In Russian) [Google Scholar]
  53. Nersesova, Z.A. Phase composition of Pore Moisture in Freezing and Thawing Soils. In Laboratory Studies of Frozen Soils; Idatel’stvo AN SSSR: Moscow, Russia, 1953; Volume 1, pp. 37–51. (In Russian) [Google Scholar]
  54. Nersesova, Z.A.; Tsytovich, H.A. Unfrozen water in frozen ground. In Proceedings of the International Conference on Permafrost, Lafayette, IN, USA, 11–15 November 1963; Section 4. Phase Equilibrium and Transitions. Izd. AN SSSR: Moscow, Russia, 1963; pp. 62–70. (In Russian). [Google Scholar]
  55. Williams, P.J. Specific Heat and Unfrozen Water Contents of Frozen Soils; Technical Memorandum 76; Natural Research Council of Canada, Associate Committee on Soil and Snow Mechanics: Ottawa, ON, Canada, 1963; pp. 109–126. [Google Scholar]
  56. Williams, P.J. Experimental determination of apparent specific heat of frozen soils. Geotechnique 1964, 14, 133–142. [Google Scholar] [CrossRef] [Green Version]
  57. Sargsyan, R.M.; Nersesova, Z.A.; Vyalov, S.S.; Zazarnaya, A.G. (Eds.) Guidelines for the Estimation of Physical, Thermal, and Mechanic Properties of Frozen Soils; Gosstoyizdat: Moscow, Russia, 1973; 194p. (In Russian) [Google Scholar]
  58. Kvlividze, V.I.; Kiselev, V.F.; Ushakova, L.A. NMR of the mobile water phase on the ice surface. Bull. Mosc. State Univ. 1974, 6, 736–738. [Google Scholar]
  59. Ananyan, A.A. Phase composition of water in frozen fine-grained rocks determined by NMR relaxometry. In Frozen Soils and Snow Cover; Nauka: Moscow, Russia, 1977; pp. 82–91. (In Russian) [Google Scholar]
  60. Tice, A.R.; Barrous, C.M.; Anderson, D.M. Phase composition measurements on soils at very high water contents by the pulsed Nuclear Magnetic Resonance technique. Transp. Res. Rec. 1978, 675, 11–14. [Google Scholar] [CrossRef]
  61. Tice, A.R.; Anderson, D.M.; Sterrett, K.F. Unfrozen water contents of submarine permafrost determined by nuclear magnetic resonance. Eng. Geol. 1981, 18, 135–146. [Google Scholar] [CrossRef]
  62. Oliphant, J.L.; Tice, A.R. Comparison of unfrozen water contents measured by DSC and NMR. In Proceedings of the 3rd International Symposium on Ground Freezing, Hanover, NH, USA, 22–24 June 1982; pp. 115–121. [Google Scholar]
  63. Tice, A.R.; Oliphant, J.L. The effect of magnetic particles on the unfrozen water contents of frozen soils determined by Nuclear Magnetic Resonance. Soil Sci. 1984, 138, 63–73. [Google Scholar] [CrossRef]
  64. Votyakov, I.N. Mechanic Properties of Permafrost in Yakutia; Nauka: Novosibirsk, Russia, 1975; 176p. (In Russian) [Google Scholar]
  65. Frolov, A.D. Electrical and Elastic Properties of Permafrost; Nedra: Moscow, Russia, 1976; 254p. (In Russian) [Google Scholar]
  66. Paterson, D.E.; Smith, M.W. The use of time domain reflectometry for the measurement of unfrozen water content in frozen soils. Cold Reg. Sci. Technol. 1980, 3, 205–210. [Google Scholar] [CrossRef]
  67. Paterson, D.E.; Smith, M.W. The measurement of unfrozen water content by time domain reflectometry: Results from laboratory tests. Can. Geotech. J. 1981, 18, 131–144. [Google Scholar] [CrossRef]
  68. Gurov, V.V. Methodology and some experimental data on dielectric properties of frozen soils. Merzlotnye Issled. 1983, 21, 170–178. [Google Scholar]
  69. Stein, J.; Kane, D.L. Monitoring the unfrozen water content of soil and snow using time domain reflectometry. Water Resour. Res. 1983, 19, 1573–1584. [Google Scholar] [CrossRef]
  70. Spaans, E.J.A.; Baker, J.M. The soil freezing characteristics: Its measurement and similarity to the soil moisture characteristic. Soil Sci. Soc. Am. J. 1996, 60, 13–19. [Google Scholar] [CrossRef]
  71. Starostin, E.G.; Petrov, E.E. Crystallization kinetic studies of unfrozen water contents in soils. Nauka i Obraz. 2013, 1, 19–23. [Google Scholar]
  72. Yershov, E.D.; Gurov, V.V.; Kuchukov, E.Z.; Komarov, I.A. Sublimation as a method for studying the properties of ice and permafrost. Merzlotnye Issled. 1976, 15, 243–247. [Google Scholar]
  73. Cheverev, V.G.; Vidyapin IYu Motenko, R.G.; Kondakov, M.V. Determination of unfrozen water content in soils by sorption-desorption isotherms. Kriosf. Zemli 2005, IX, 29–33. [Google Scholar]
  74. Istomin, V.A.; Chuvilin, E.M.; Makhonina, N.A.; Bukhanov, B.A. A method for calculating the unfrozen water curve from moisture potential. In Proceedings of the International Conference Cryogenic Resources of the Polar and Mountainous Regions: State of the Art and Prospects of Permafrost Engineering, Tyumen, Russia, 21–24 April 2008; pp. 398–401. (In Russian). [Google Scholar]
  75. Istomin, V.A.; Chuvilin, E.M.; Bukhanov, B.A. Fast estimation of unfrozen water content in frozen soils. Earth’s Cryosphere 2017, XXI, 134–139. [Google Scholar]
  76. Istomin, V.; Chuvilin, E.; Bukhanov, B.; Uchida, T. Pore water content in equilibrium with ice or gas hydrate in sediments. Cold Reg. Sci. Technol. 2017, 137, 60–67. [Google Scholar] [CrossRef]
  77. Chuvilin, E.M.; Sokolova, N.S.; Bukhanov, B.A.; Shevchik, F.A.; Istomin, V.A.; Mukhametdinova, A.Z.; Alekseev, A.G.; Grechishcheva, E.S. The water potential method for determination of unfrozen water contents in frozen soils of different compositions. Earth’s Cryosphere 2020, XXIV, 16–28. [Google Scholar]
  78. Chuvilin, E.M.; Bukhanov, B.A.; Mukhametdinova, A.Z.; Grechishcheva, E.S.; Sokolova, N.S.; Alekseev, A.G.; Istomin, V.A. Freezing point and unfrozen water contents of permafrost soils: Estimation by the water potential method. Cold Reg. Sci. Technol. 2022, 196, 103488. [Google Scholar] [CrossRef]
  79. Campbell, G.S.; Smith, D.M.; Teare, B.L. Application of a Dew Point Method to obtain the soil water characteristic. In Experimental Unsaturated Soil Mechanics; Springer Proceedings in Physics; Springer: Berlin/Heidelberg, Germany, 2007; Volume 112, pp. 71–77. [Google Scholar]
  80. Istomin, V.A.; Chuvilin, E.M.; Bukhanov, B.A. A Method for Determination of Unfrozen Water Content in Soils. Russian Federation. Patent No. 2654832, 22 May 2018. (In Russian). [Google Scholar]
  81. Chuvilin, E.M.; Sokolova, N.S.; Bukhanov, B.A.; Istomin, V.A.; Mingareeva, G.R. Determination of the freezing point of soils based on measurements of pore water potential. Earth’s Cryosphere 2020, XXIV, 9–16. [Google Scholar] [CrossRef]
  82. Aleksyutina, D.M.; Motenko, R.G. Effect of soil salinity and organic matter content on thermal properties and unfrozen water content of frozen soils in the western coast of Baydara Bay. Bull. Mosc. Univ. 2016, 4, 59–63. [Google Scholar] [CrossRef]
  83. ASTM D 2487–06; Standard Practice for Classification of Soils for Engineering Purposes (Unified Soil Classification System). ASTM: West Conshohocken, PA, USA, 2010; 12p.
  84. Roman, L.T.; Tsarapov, M.N.; Kotov, P.I.; Volokhov, S.S.; Motenko, R.G.; Cherkasov, A.M.; Shteyn, A.I.; Kostousov, A.I. A Guide for Estimation of Mechanic Properties of Freezing, Frozen and Thawing Fine-Grained Soils; Knizhnyi Dom Universitet: Moscow, Russia, 2018; 188p. (In Russian) [Google Scholar]
  85. Vasiliev, A.A.; Gravis, A.G.; Gubarkov, A.A.; Drozdov, D.S.; Korostelev, Y.V.; Malkova, G.V.; Oblogov, G.E.; Ponomareva, O.E.; Sadurtdinov, M.R.; Streletskaya, I.D.; et al. Permafrost degradation: Results of the long-term geocryological monitoring in the Western sector of Russian Arctic. Earth’s Cryosphere 2020, XXIV, 14–26. [Google Scholar] [CrossRef]
  86. Osipov, V.; Aksyutin, O.; Sergeev, D.; Tipenko, G.; Ishkov, A. Using the data of geocryological monitoring and geocryological forecast for risk assessment and adaptation to climate change. Energies 2022, 15, 879. [Google Scholar] [CrossRef]
  87. Chuvilin, E.M.; Perlova, E.V.; Baranov, Y.B.; Kondakov, V.V.; Osokin, A.B.; Yakushev, V.S. Structure and Properties of Permafrost in the Southern Bovanenkovo Gas-Condensate Field; GEOS: Moscow, Russia, 2007; 137p. (In Russian) [Google Scholar]
Figure 1. WP4-T (a) and WP4C (b) water potential meters.
Figure 1. WP4-T (a) and WP4C (b) water potential meters.
Geosciences 12 00253 g001
Figure 2. Temperature-dependent content of unfrozen pore water in permafrost. t f 1 ,   t f 2 ,   t f 3 and t f 4   are soil freezing points with moisture contents W 1 , W 2 , W 3 ,     W 4 , respectively. W 0   is the soil natural moisture content.
Figure 2. Temperature-dependent content of unfrozen pore water in permafrost. t f 1 ,   t f 2 ,   t f 3 and t f 4   are soil freezing points with moisture contents W 1 , W 2 , W 3 ,     W 4 , respectively. W 0   is the soil natural moisture content.
Geosciences 12 00253 g002
Figure 3. Temperature-dependent unfrozen pore water content in clay samples with peat contents ( I r ) of 0.4% (1); 2.7% (2); 3.1% (3); 6.2% (4).
Figure 3. Temperature-dependent unfrozen pore water content in clay samples with peat contents ( I r ) of 0.4% (1); 2.7% (2); 3.1% (3); 6.2% (4).
Geosciences 12 00253 g003
Figure 4. Salinity-dependent unfrozen pore water content in sand and lean clay soils at temperatures of: −2 °C; −4 °C; −10 °C, after Aleksyutina and Motenko [82].
Figure 4. Salinity-dependent unfrozen pore water content in sand and lean clay soils at temperatures of: −2 °C; −4 °C; −10 °C, after Aleksyutina and Motenko [82].
Geosciences 12 00253 g004
Figure 5. Geocryological station location in the western segment of the Russian Arctic (a) and time-dependent changes in air temperature at nearby meteorological stations (b): 1 = Marre–Sale; 2 = Cape Konstantinovsky; 3 = Vorkuta; 4 = Novy Urengoi; 5 = Nadym; 6 = Tarko–Sale, after Vasiliev et al. [82].
Figure 5. Geocryological station location in the western segment of the Russian Arctic (a) and time-dependent changes in air temperature at nearby meteorological stations (b): 1 = Marre–Sale; 2 = Cape Konstantinovsky; 3 = Vorkuta; 4 = Novy Urengoi; 5 = Nadym; 6 = Tarko–Sale, after Vasiliev et al. [82].
Geosciences 12 00253 g005
Figure 6. Variations of mean annual permafrost temperatures at the depth of zero annual amplitudes, from temperature logs at Marre–Sale (a) and North Urengoi (b), zones of typical and southern tundra, respectively [85]. Numbers on legends indicate borehole numbers. Single temperature measurements (grey and green points) are made in unnumbered wells.
Figure 6. Variations of mean annual permafrost temperatures at the depth of zero annual amplitudes, from temperature logs at Marre–Sale (a) and North Urengoi (b), zones of typical and southern tundra, respectively [85]. Numbers on legends indicate borehole numbers. Single temperature measurements (grey and green points) are made in unnumbered wells.
Geosciences 12 00253 g006
Figure 7. Depth profiles (to 100 m) of permafrost type (a), temperature (b), total moisture (c), salinity (d), and organic carbon content (e) estimated from log data for the area of the Bovanenkovo gas field in the Yamal Peninsula [87].
Figure 7. Depth profiles (to 100 m) of permafrost type (a), temperature (b), total moisture (c), salinity (d), and organic carbon content (e) estimated from log data for the area of the Bovanenkovo gas field in the Yamal Peninsula [87].
Geosciences 12 00253 g007
Table 1. Summary of methods for studying the phase composition of pore moisture in frozen soil.
Table 1. Summary of methods for studying the phase composition of pore moisture in frozen soil.
ExperimentCombined
(Experiments + Modeling)
Modeling Using Empirical
Relationships with Other
Soil Variables
Cryoscopy [15]Calorimetry [53,54,55,56,57]Plasticity, temperature, and salinity [39,40,53]
Nuclear Magnetic Resonance
(NMR) [58,59,60,61,62,63]
Total moisture and grain size [50]
Hygroscopic
absorption [64]
Time domain reflectometry
(TDR) [65,66,67,68,69,70]
Pore pressure [41]
Thermometry [35,71]Specific surface area [42,43,44]
Sublimation [72]Desorption [27,73]Total moisture and heat capacity [46]
Water potential [74,75,76,77,78]Total moisture, ground temperature, and freezing point [47,48]
Contact [19] Freezing point and unfreezable water content [49]
The references are to papers where the respective methods were first described.
Table 2. Salinity and organic content in soil samples from northern West Siberia.
Table 2. Salinity and organic content in soil samples from northern West Siberia.
SampleSoil Type *Salinity, %Organic Matter
Content **, %
1Lean clay (CL)0.050.4
2Sandy silty clay (CL-ML)0.042.7
3Lean clay (CL)0.063.1
4Silt (ML)0.066.2
* Description according to ASTM D2487-06 [83]. ** OM content calculated as organic matter-to-dry sediment weight ratio from loss on ignition at 525 °C for 4 h.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chuvilin, E.; Sokolova, N.; Bukhanov, B. Changes in Unfrozen Water Contents in Warming Permafrost Soils. Geosciences 2022, 12, 253. https://0-doi-org.brum.beds.ac.uk/10.3390/geosciences12060253

AMA Style

Chuvilin E, Sokolova N, Bukhanov B. Changes in Unfrozen Water Contents in Warming Permafrost Soils. Geosciences. 2022; 12(6):253. https://0-doi-org.brum.beds.ac.uk/10.3390/geosciences12060253

Chicago/Turabian Style

Chuvilin, Evgeny, Natalia Sokolova, and Boris Bukhanov. 2022. "Changes in Unfrozen Water Contents in Warming Permafrost Soils" Geosciences 12, no. 6: 253. https://0-doi-org.brum.beds.ac.uk/10.3390/geosciences12060253

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop