Next Article in Journal
Daily Water Mapping and Spatiotemporal Dynamics Analysis over the Tibetan Plateau
Previous Article in Journal
GMesh: A Flexible Voronoi-Based Mesh Generator with Local Refinement for Watershed Hydrological Modeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Geothermal Heating on Deep-Water Temperature in Lake Baikal

Computational Geophysics Laboratory, Tomsk State University, Tomsk 634050, Russia
Hydrology 2025, 12(10), 256; https://doi.org/10.3390/hydrology12100256
Submission received: 24 July 2025 / Revised: 22 September 2025 / Accepted: 28 September 2025 / Published: 30 September 2025
(This article belongs to the Section Hydrological and Hydrodynamic Processes and Modelling)

Abstract

Geothermal heating that emanates from the interior of the Earth, including the Baikal Rift Zone, produces potential energy for water movement. The basic concept behind the mechanism of deep-water renewal in Lake Baikal is conditional instability, which is a consequence of the joint effects of temperature and pressure on water density. However, an exact trigger of this instability is unknown. In this study, based on a non-hydrostatic 2.5D numerical model taking into account the intraday variability of atmospheric conditions, it was shown that, due to geothermal heating, the water column near the lake bed becomes slightly warmer (0.1–0.2 °C) than ambient waters, which can lead to instability. Simulated temperature distributions showed that 3.4 °C waters gradually shifted along the bed slope to ~650 m on day 1, ~750 m on day 3, ~830 m on day 5, and >1200 m on day 10 in the presence of geothermal heat flux; however, in its absence these waters remained at the level of ~600 m. In view of these findings, a conceptual model of deep convection and a map with potential zones of high ventilation processes in Lake Baikal are proposed. According to the map developed, deep-water renewal is expected to be the most intense at the eastern shore of Lake Baikal because of abnormally high heat release.

1. Introduction

Lake Baikal, located nearly in the center of the Asian continent in the southern part of eastern Siberia, is not only the world’s oldest and deepest lake, but also home to an enormous number (>1500) of endemic plant and animal species and the largest by volume (23,013 km3) freshwater lake on our planet (~20% of global fresh surface water). Despite its great depth, Lake Baikal’s waters are enriched with oxygen (>9.5 mg L−1) throughout the water column [1,2], and its water age according to tritium/helium [3,4] and fluorochlorocarbons [5] distributions does not exceed 18 years. The short time period of deep-water renewal, contrary to a permanent stable stratification below ~250 m (the mesothermal maximum where water temperature is equal to the temperature of maximum density and thermal expansion coefficient changes sign [6]), suggests the existence of a special mechanism that controls water exchange processes between the surface and deep regions of the lake.
The mechanism of deep-water renewal, or ‘ventilation’, in Lake Baikal has been a mystery for many years. Despite remarkable efforts of scientists to solve it, a clear description of the physical mechanism of the processes of vertical water exchange does not yet exist, and many results of direct observations are difficult to explain. The basic concept behind the ventilation of Lake Baikal waters is conditional instability that depends on the joint effects of temperature and pressure on water density [5,7]. The temperature of maximum density, Tmd, is ~4 °C at the lake surface and decreases with depth by ~0.2 °C per 100 m [8,9]. Because a water column of Lake Baikal below the mesothermal maximum horizon has a relatively homogeneous temperature (3.2–3.6 °C) [2] and is permanently stratified, the stability of this column is expected to be very sensitive to small variations in water temperature.
However, it is unclear what triggers this instability in Lake Baikal. The most popular ideas are wind-forced mixing [5,10,11], highly mineralized inflow waters [12], and cabbeling near a thermal bar front [13] due to the density increase under the mixing of water masses of different temperatures [14]. Also, processes of ventilation in Lake Baikal are believed to be linked to near-coast downwelling and deep-water intrusions [9,15].
During spring and autumn transitions, the role of a unique phenomenon of nature, the thermal bar—a sinking of maximum density waters in a narrow zone [16,17,18]—is crucial in the ecological functioning of a temperate lake [19]. In Lake Baikal, thermal bars are observed in bays (Chivyrkujskij, Maloe More), shallow water zones (Selenga), and close to relatively steep littoral areas [20,21,22]. Vertical flows forming in the location of a thermal bar may promote the transport of oxygen, microorganisms, and suspended matter to deeper layers [6,23,24]. Tributaries are very important for the formation of thermal bars in Lake Baikal [23] because river waters entering the lake are involved in general cyclonic round-the-Baikal circulation [20,25,26].
Numerical modeling of the effects of geothermal heating during a thermal bar in temperate lakes has not been undertaken. Also, the novelty of the developed model consists in consideration of a cross-section of the example of the Lake Baikal deep-water bathymetry, taking into account the intraday variability of atmospheric conditions. The aim of this study was to estimate the contribution of geothermal heat flux to conditional instability that causes deep-water renewal in Lake Baikal. Herein, on the basis of numerical modeling and analysis of observational data, a conceptual model of deep convection in Lake Baikal and a map with potential zones of high ventilation are proposed.

2. Materials and Methods

2.1. Numerical Model

The non-hydrostatic 2.5D numerical model [27] for the reproduction of thermohydrodynamic processes takes into account the intraday variability of the fluxes of short- and longwave radiation and latent and sensible heat [28], as well as wind action at the water–air interface. The model, including the equations of momentum, continuity, energy, and salinity balance, is based on the Reynolds-averaged Navier–Stokes equations in the Boussinesq approximation with appropriate initial and boundary conditions, and considers the influence of river inflow [29] and the effect of the Coriolis force related to the Earth’s rotation [30]:
u t + u 2 x + u w z = 1 ρ 0 p x + x K x u x + z K z u z + 2 Ω z v 2 Ω y w ;
v t + u v x + w v z = x K x v x + z K z v z + 2 Ω x w 2 Ω z u ;
w t + u w x + w 2 z = 1 ρ 0 p z + x K x w x + z K z w z g ρ ρ 0 + 2 Ω y u 2 Ω x v ;
u x + w z = 0 ;
T t + u T x + w T z = x D x T x + z D z T z + 1 ρ 0 c p H s o l z ;
S t + u S x + w S z = x D x S x + z D z S z ,
where u and v are the horizontal velocity components along the Ox and Oy axes, respectively; w is the vertical velocity component; T is the temperature; S is the salinity; p is the pressure; Ωx, Ωy, and Ωz are the vector components of the Earth rotation’s angular velocity; g is the acceleration of gravity; cp is the specific heat capacity; and ρ0 is the water density at standard atmospheric pressure.
The initial conditions for the model equations were set as
u = 0 ;   v = 0 ;   w = 0 ;   T = T L ;   S = S L ,
where TL and SL are the lake’s water temperature and salinity, respectively.
Boundary conditions were considered, as follows:
(a)
At the surface of the lake
K z u z = τ s u r f u ρ 0 ;   K z v z = τ s u r f v ρ 0 ;   w = 0 ;   D z T z = H n e t ρ 0 c p ;   S z = 0 ,
where Hnet is the heat flux of latent and sensible heat and longwave radiation [28];
(b)
At the solid boundaries
u = 0 ;   v = 0 ;   w = 0 ;   T n = 0 ;   S n = 0 ,
where n is the direction of the outward normal to the domain;
(c)
At the river inflow boundary
u = u R ;   v = 0 ;   w = 0 ;   T = T R ;   S = S R ,
where uR is the river inflow velocity; TR and SR are the temperature and salinity in the river, respectively;
(d)
At the open boundary, where the following conditions for the radiation type [31] were set:
ϕ t + c ϕ ϕ x = 0   ϕ = u , v , T , S ;   w x = 0 ,
The phase velocity cϕ is calculated here from the space and time trends ϕ in the domain near the boundary.
To close the set of equations, Wilcox’s two-parameter k–ω turbulence model [32] consisting of equations for kinetic energy and turbulent fluctuation frequency and algebraic relations to find eddy diffusivity were used. The horizontal diffusion coefficients are set to constants Kx = Dx = 5.0 m2/s [33]. Conditions of no-slip are set on solid boundaries [33,34,35]. The Chen and Millero equation [36] connecting water density with temperature (T), salinity (S), and pressure (p), and valid within the range of 0 ≤ T ≤ 30 °C, 0 ≤ S ≤ 0.6 g/kg, 0 ≤ p ≤ 180 bar, was taken as the state equation.
The problem is solved using the finite volume method [37]. The scalar quantities (temperature, salinity, etc.) are calculated in the center of a grid cell, and the velocity vector components at the midpoints of the cell boundaries. The numerical algorithm for finding the flow and temperature fields is based on a Crank–Nicolson finite difference scheme. The convective terms in the equations are approximated with a second-order upstream scheme, QUICK [38]. The velocity and pressure fields calculated are correlated by a procedure for buoyant flows, SIMPLED (Semi-Implicit Method for Pressure Linked Equations with Density correction) [39]. The initial pressure field is specified by solving the state and hydrostatic equations with the boundary condition p = pa on the surface (pa is the atmospheric pressure) by a fourth-order Runge–Kutta method. The calculation domain with a length of 10,000 m and depth of 1200 m (Figure 1b) was covered by a uniform orthogonal grid with steps hx = 50 m and hz = 5 m. The time step is 30 s.
Validation of the model of the hydrodynamics of the thermal bar was conducted in [27,39]. Analyses of the effects of wind, surface heat fluxes, tributary water mineralization, and Coriolis force were presented in previous studies [27,28,29,30,40,41,42,43].

2.2. Study Area and Problem Parameters

The Boldakov River–Maloye More Strait cross-section, located 70 km north of the Selenga River delta, in the Central Baikal, was taken for this study (Figure 1a). The bottom topography (Figure 1b) was taken from Likhoshway et al. [23].
The initial temperature field in simulations approximately corresponded to the average data of the thermal regime of the Central Basin of Lake Baikal in October [2]. In the model, the initial water temperature in the Boldakov River was 1.5 °C and decreased by 0.02 °C every day. The river flowed into the lake at a velocity of 5 mm/s. Water mineralization in the lake was 96 mg/kg [2]; in the river it was set to 128.2 mg/kg, which is the mean value for all tributaries of Lake Baikal [44]. Heat flux components across the lake surface were calculated based on data on air temperature, relative humidity, air pressure, cloud cover, and wind speed and direction taken from the weather conditions archives of the meteorological station in Goryachinsk village from 1 November 2015, to 30 November 2015 [45]. Note that the initial conditions in simulations are the same, except for geothermal heat flux. Despite the short period of existence of a thermal bar phenomenon, earlier model results obtained for Lake Baikal are consistent with observational data [27,39].

3. Results and Discussion

3.1. Geothermal Heating and Deep Convection

Geothermal heat flux, Hgeo, emanating from the interior of the Earth produces potential energy. In Lake Kivu, for example, convective mixing at the lake bottom induced by extensive geothermal springs can become the dominant mixing process [46]. In Frolikha Bay of Lake Baikal, the temperature in the lowest 20 m of water increases by 0.1–0.15 °C [47]. Regional scale studies of ocean circulation show that geothermal heating can modify local water properties [48,49,50] and contribute to an overall warming of bottom waters by ~0.4 °C [51].
Geothermal heat flux in the Baikal Rift Zone, including the Baikal depression, is sufficiently non-uniform [52,53,54]. Hgeo is fairly low and poorly variable in the North basin of Lake Baikal and has positive anomalies located near the eastern shore in the South and Central basins [54]. The average value Hgeo is 71 ± 21 mW/m2 [53,54]. An abnormally high heat release (100–200 mW/m2) occurs only on separate sites—in fault zones against the lake shore; extreme values (200–3000 mW/m2) were registered in local centers of discharge of fissure hydrotherms at the lake bottom [52,54]. The value of Hgeo = 138 mW/m2 used in this study corresponds to observational data [55] at a region under simulation—a 10 km transect directed from the Boldakov River to the Moloye More Strait (Figure 1).
Simulated temperature distributions show that in the presence of geothermal heat flux the 3.4 °C waters along the bed slope gradually shift to ~650 m on day 1, ~750 m on day 3, ~830 m on day 5, and > 1200 m on day 10 (Figure 2b), and in the absence of geothermal heat flux they remain at the level of ~600 m (Figure 2a).
Due to geothermal heating, waters along the bed slope (Figure 2b) become slightly warmer (less compressible and less dense) than ambient waters. As a result, the water column near the slope becomes unstable, and dense (ambient) waters from the overlying layers move down the bed slope into the deeper part of the lake. Thus, deep convection occurs.
The results of simulations demonstrate that the velocity of the near-slope flows below 600 m is less than the velocity of the thermal bar circulation, which is ~0.3 cm/s. However, the impact of geothermal heat flux is clearly manifested on the temperature distributions in the near-slope zone (Figure 2b,d). It follows that geothermal heat flux can act in combination with and supplement other system parameters such as water mineralization, wind friction, or solar radiation, which can enhance the effect of thermobaricity. The circulation paths during thermal bar events on the Boldakov River–Maloye More Strait cross-section in November 2015 are presented in [43].
A similar deep convection mechanism below ~600 m is, however, not observed in the absence of geothermal heat flux (Figure 2a,c). In this situation, we see only a formation of vertical temperature homogeneity in the upper 600 m layer between thermoactive and thermoinert regions of the lake (Figure 2c). Note that geothermal heat flux, in contrast to other external sources of force (wind, solar radiation, river inflow, etc.), has the important feature of acting throughout the entire water column. As simulations showed, the energy of geothermal heat flux was enough to heat up the near-slope deep waters by 0.1–0.2 °C (Figure 2d), which can lead to the effect of thermobaricity (caused by pressure dependence of the thermal expansion of water) [14] near the bed slope of the lake. As a consequence, the upper layer waters shift downslope to the deeper region. The effectiveness of deep ventilation can be affected by slope gradient, because in a current flowing down a slope the direct entrainment increases with slope [56]. Simulations with different slopes during the thermal bar were performed in [57].

3.2. Impact of the Thermal Bar

The thermal bar originated by the end of day 2 (Figure 3a), but the amplification of northwest wind up to 8 m/s (Figure 4) blowing opposite to the direction of the thermal bar propagation on day 3 led to the destruction of the thermal bar front; after the wind weakened it was formed again at the beginning of day 4. When colder river water (<Tmd) mixes with warmer lake water (>Tmd) denser water (~Tmd) forms, and due to the effect of cabbeling (caused by the temperature dependence of the thermal expansion of water [14]) it sinks at a front of the thermal bar to the horizon of the mesothermal maximum (Figure 3b). In the absence of geothermal heat flux, the density plume may drop below this horizon, but it is unable to penetrate to greater depths; the denser water reached a depth of ~600 m on day 10 (Figure 3b). The average velocity of water sinking in this case was ~0.3 cm/s, which is consistent with the value estimated by Shimaraev et al. [13].
If there is geothermal heat flux, denser surface water near the thermal bar can sink along the bed slope directly to the bottom, reaching the bed of the lake and then flowing down along the bed slope (Figure 3c). Along-slope downwelling, generated by geothermal heating, attracts the denser water that forms at the thermal bar. For this reason, the horizontal propagation of the thermal bar is more sensitive to atmospheric impacts. The position of the temperature of maximum density on the lake surface indicates that the front of the thermal bar on day 10 was located at 2.3 km in the absence of geothermal heat flux and 1.7 km from the shore in the presence of geothermal heat flux (Figure 3a), and this difference at the end of the day decreased by 2.4 km and 2.0 km, respectively (Figure 3b,c). As the distance from the river mouth to the thermal bar front increases, the influence of atmospheric parameters strengthens (after day 8) (Figure 3a).
Note that during the existence of the thermal bar, cold surface waters located relative to the thermal bar front in the shore side in autumn, or in the open lake side in spring, can directly move down to the abyss. The combination of cabbeling and geothermal heating can force the cold-water plume to sink to the lake bottom. Presumably, a similar picture can be observed in the case of a classic thermal bar (without river inflow), but the intensity of deep convection can be low because of the lack of additional mechanical energy from river inflow.

3.3. Mechanism of Water Renewal

This study suggests that the mechanism of water renewal in Lake Baikal during the thermal bar (Figure 5) can consist of three phases. The first phase (up to ~200 m) is the transfer of surface waters to the depth of the mesothermal maximum due to active wind forcing. Simulations with various wind speeds demonstrated that if winds with a speed < 10.8 m/s had an impact on water temperature in the upper 100 m layer of Lake Baikal, strong wind (10.8–13.8 m/s) had an impact only in the upper 200 m layer [40]. The second phase (up to ~600 m) can occur during the existence of a thermal bar, in spring and autumn transition periods, due to the effect of cabbeling. The third phase (up to maximum depths) is associated with the effect of non-linearities in the equation of state for fresh water; namely, a difference in the compressibility of water leads to conditional instability—the effect of thermobaricity—triggered by the high mineralization of tributary waters [43] and/or geothermal heating. Note that the phases listed can occur either individually or in combinations.
Because water is denser at the horizon of the mesothermal maximum, a parcel of water generates a compensatory circulation above this horizon during vertical mixing. Also, the formation of compensatory circulations and intrusions above the boundary between relatively flat and steep areas of lake bed can be affected by the joint effect of wind and lake morphology [41]. Along-slope downwelling, reaching the lake bottom, can induce bottom intrusions under the action of Coriolis forces [58].

3.4. Potential Zones with High Ventilation

On the basis of these findings and analysis of values of heat fluxes measured at the bottom of Lake Baikal [54], a Lake Baikal map was developed indicating zones where deep-water renewal can be especially pronounced (Figure 6). In this map, the marked points correspond to zones with high heat flux (>120 mW/m2). For reasons of specificity of heterogenic distribution of heat flow, the eastern shore of Lake Baikal produces more favorable conditions for generating conditional instability. Nevertheless, in the North basin (near Capes Malyi Solontsovyi, Zavorotnyi, and Tonkii) active water renewal processes are viable at the western shore. However, the effect of thermobaricity in the North basin appears to be manifested to a lesser degree due to lower depth (the average depth is 576 m [59]) in comparison with other basins of Lake Baikal.
Note that the existing in situ registrations of deep-water renewal in Lake Baikal [6,13,15] coincide with one of the points marked on the map. At the same time, numerical modeling with Hgeo = 50 mW/m2 (geothermal heat flux through the bottom of Lake Baikal exceeds 50 mW/m2 almost everywhere [52]) also demonstrates deep convection, but its efficiency has been significantly reduced.
The results obtained are consistent with a hypothesis on the relationship of deep-water renewal and coastal downwelling [9,60]. In ocean studies, geothermal heating can lead to large-scale overturning circulation [61,62] and can increase the temperature by ~0.5 °C [48,63].

4. Conclusions

This study suggests that the mechanism of water renewal in Lake Baikal is carried out through
(i)
Wind forcing (up to ~200 m);
(ii)
Cabbeling at the thermal bar (up to ~600 m);
(iii)
Tributary mineralization and/or geothermal heating (up to maximum depths).
Based on simulations conducted here, we can conclude that geothermal heat flux can trigger conditional instability near the bed slope of Lake Baikal. Due to abnormally high heat release, deep-water renewal in Lake Baikal can be especially pronounced at the eastern shore. These results could be a starting point for more in situ observations on the study of the nature and extent of influence of geothermal heating on Baikal water renewal. Also, they have implications for selecting an effective strategy to prevent deep-water pollution in Lake Baikal, the largest repository of freshwater on our planet, in light of the noticeable worsening of its littoral ecological status in recent decades.

Funding

This research was funded by the Russian Science Foundation, grant number 23-71-10020.

Data Availability Statement

Datasets are available on request from the author.

Acknowledgments

The author is very grateful to Jean Kollantai, MSW, for style review.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Beeton, A.M. Large freshwater lakes: Present state, trends, and future. Environ. Conserv. 2002, 29, 21–38. [Google Scholar] [CrossRef]
  2. Shimaraev, M.N.; Verbolov, V.I.; Granin, N.G.; Sherstyankin, P.P. Physical Limnology of Lake Baikal: A Review; BYCER: Irkutsk, Russia; Okayama, Japan, 1994. [Google Scholar]
  3. Craig, H. Helium, tritium, methane, and temperature in deep rift-valley lakes: Tanganyika, Kivu, and Baikal. EOS Trans. AGU 1989, 70, 1137. [Google Scholar]
  4. Hohmann, R.; Hofer, M.; Kipfer, R.; Peeters, F.; Imboden, D.M.; Baur, H.; Shimaraev, M.N. Distribution of helium and tritium in Lake Baikal. J. Geophys. Res. 1998, 103, 12823–12838. [Google Scholar] [CrossRef]
  5. Weiss, R.; Carmack, E.; Koropalov, V. Deep-water renewal and biological production in Lake Baikal. Nature 1991, 349, 665–669. [Google Scholar] [CrossRef]
  6. Shimaraev, M.N.; Grachev, M.A.; Imboden, D.M.; Okuda, S.; Granin, N.G.; Kipfer, R.; Levin, L.A.; Endo, S. International hydrophysical experiment in Lake Baikal: Spring deep water renewal. Dokl. Acad. Nauk 1995, 343, 824–827. [Google Scholar]
  7. Carmack, E.C.; Farmer, D.M. Cooling processes in deep, temperate lakes: A review with examples from two lakes in British Columbia. J. Mar. Res. 1982, 40, 85–111. [Google Scholar]
  8. Eklund, H. Stability of lakes near the temperature of maximum density. Science 1965, 149, 632–633. [Google Scholar] [CrossRef]
  9. Schmid, M.; Budnev, N.M.; Granin, N.G.; Sturm, M.; Schurter, M.; Wüest, A. Lake Baikal deepwater renewal mystery solved. Geophys. Res. Lett. 2008, 35, L09605. [Google Scholar] [CrossRef]
  10. Botte, V.; Kay, A. A model of the wind-driven circulation in Lake Baikal. Dynam. Atmos. Oceans 2002, 35, 131–152. [Google Scholar] [CrossRef]
  11. Killworth, P.D.; Carmack, E.C.; Weiss, R.F.; Matear, R. Modeling deep-water renewal in Lake Baikal. Limnol. Oceanogr. 1996, 41, 1521–1538. [Google Scholar] [CrossRef]
  12. Hohmann, R.; Kipfer, R.; Peeters, F.; Piepke, G.; Shimaraev, M.N.; Imboden, D.M. Processes of deep-water renewal in Lake Baikal. Limnol. Oceanogr. 1997, 42, 841–855. [Google Scholar] [CrossRef]
  13. Shimaraev, M.N.; Granin, N.G.; Zhdanov, A.A. Deep ventilation of Lake Baikal waters due to spring thermal bars. Limnol. Oceanogr. 1993, 38, 1068–1072. [Google Scholar] [CrossRef]
  14. McDougall, T.J. The relative roles of diapycnal and isopycnal mixing on subsurface water mass conversion. J. Phys. Oceanogr. 1984, 14, 1577–1589. [Google Scholar] [CrossRef]
  15. Wüest, A.; Ravens, T.M.; Granin, N.G.; Kosis, O.; Schurter, M.; Sturm, M. Cold intrusions in Lake Baikal: Direct observational evidence for deep-water renewal. Limnol. Oceanogr. 2005, 50, 184–196. [Google Scholar] [CrossRef]
  16. Blokhina, N.S.; Ordanovich, A.E.; Savel’eva, O.S. Model of formation and development of spring thermal bar. Water Resour. 2001, 28, 201–204. [Google Scholar] [CrossRef]
  17. Chubarenko, I.P.; Demchenko, N.Y. Laboratory modeling of the structure of a thermal bar and related circulation in a basin with a sloping bottom. Oceanology 2008, 48, 327–339. [Google Scholar] [CrossRef]
  18. Farrow, D.E. A numerical model of the hydrodynamics of the thermal bar. J. Fluid Mech. 1995, 303, 279–295. [Google Scholar] [CrossRef]
  19. Holland, P.R.; Kay, A. A review of the physics and ecological implications of the thermal bar circulation. Limnologica 2003, 33, 153–162. [Google Scholar] [CrossRef]
  20. Verbolov, V.I.; Sokolnikov, V.M.; Shimaraev, M.N. Hydrometeorological Regime and Heat Balance of Lake Baikal; Nauka: Moscow, Russia, 1965. [Google Scholar]
  21. Sherstyankin, P.P. Spatial distribution of transparency in Maloe More and its connection with water dynamics. In Baikal Productivity and Anthropogenic Changes of Its Nature; ISU: Irkutsk, Russia, 1974; pp. 54–62. [Google Scholar]
  22. Shimaraev, M.N. Elements of Heat Regime of Lake Baikal; Nauka: Novosibirsk, Russia, 1977. [Google Scholar]
  23. Likhoshway, Y.V.; Kuzmina, A.Y.; Potyemkina, T.G.; Potyemkin, V.L.; Shimaraev, M.N. The distribution of diatoms near a thermal bar in Lake Baikal. J. Great Lakes Res. 1996, 22, 5–14. [Google Scholar] [CrossRef]
  24. Parfenova, V.V.; Shimaraev, M.N.; Kostornova, T.Y.; Domysheva, V.M.; Levin, L.A.; Dryukker, V.V.; Zhdanov, A.A.; Gnatovskii, R.Y.; Tsekhanovskii, V.V.; Logacheva, N.F. On the vertical distribution of microorganisms in Lake Baikal during spring deep-water renewal. Mikrobiologiya 2000, 69, 433–440. [Google Scholar] [CrossRef]
  25. Krotova, V.A.; Mankovskij, B.I. Report on the Study of Currents in Lake Baikal in 1962; LIN Library Funds: Irkutsk, Russia, 1962. [Google Scholar]
  26. Sokolnikov, V.M. Currents and water exchange in Lake Baikal. In Elements of Hydrometeorological Regime of Lake Baikal; Nauka: Moscow–Leningrad, Russia, 1964; pp. 5–21. [Google Scholar]
  27. Tsydenov, B.O. A numerical study of the thermal bar in shallow water during the autumn cooling. J. Great Lakes Res. 2019, 45, 715–725. [Google Scholar] [CrossRef]
  28. Tsydenov, B.O. Effects of heat fluxes on the phytoplankton distribution in a freshwater lake. Atmos. Ocean. Opt. 2021, 34, 603–610. [Google Scholar] [CrossRef]
  29. Tsydenov, B.O. Numerical modeling of the effect of inflow water mineralization in the dynamics of the autumnal thermal bar in Kamloops Lake. Mosc. Univ. Phys. Bull. 2018, 73, 435–440. [Google Scholar] [CrossRef]
  30. Tsydenov, B.O. The effect of the Coriolis force and wind on the dynamics of the autumn thermal bar. Moscow Univ. Phys. Bull. 2019, 74, 70–76. [Google Scholar] [CrossRef]
  31. Orlanski, I. A simple boundary condition for unbounded hyperbolic flows. J. Comput. Phys. 1976, 21, 251–269. [Google Scholar] [CrossRef]
  32. Wilcox, D. Reassessment of the scale-determining equation for advanced turbulence models. AIAA J. 1988, 26, 1299–1310. [Google Scholar] [CrossRef]
  33. Malm, J. Spring circulation associated with the thermal bar in large temperate lakes. Nord. Hydrol. 1995, 26, 331–358. [Google Scholar] [CrossRef]
  34. Tsvetova, E.A. Mathematical modelling of Lake Baikal hydrodynamics. Hydrobiologia 1999, 407, 37–43. [Google Scholar] [CrossRef]
  35. Holland, P.R.; Kay, A.; Botte, V. A numerical study of the dynamics of the riverine thermal bar in a deep lake. J. Environ. Fluid Mech. 2001, 1, 311–332. [Google Scholar] [CrossRef]
  36. Chen, C.T.; Millero, F.G. Precise thermodynamic properties for natural waters covering only limnologies range. Limnol. Oceanogr. 1986, 31, 657–662. [Google Scholar] [CrossRef]
  37. Patankar, S. Numerical Heat Transfer and Fluid Flow; Hemisphere Publishing Corporation: Washington, DC, USA, 1980. [Google Scholar]
  38. Leonard, B. A stable and accurate convective modelling procedure based on quadratic upstream interpolation. Comput. Meth. Appl. Mech. Engineer. 1979, 19, 59–98. [Google Scholar] [CrossRef]
  39. Tsydenov, B.O.; Kay, A.; Starchenko, A.V. Numerical modeling of the spring thermal bar and pollutant transport in a large lake. Ocean Model. 2016, 104, 73–83. [Google Scholar] [CrossRef]
  40. Tsydenov, B.O. The impact of easterly wind force on Lake Baikal water temperature in autumn. In Proceedings of the All-Russian Scientific Conference “Water and Ecological Problems of Siberia and Central Asia”, Barnaul, Russia, 29 August–3 September 2022; Volume 1, pp. 257–262. [Google Scholar]
  41. Tsydenov, B.O. The fall thermal bar dynamics under a differential wind load. Mosc. Univ. Phys. 2022, 77, 61–67. [Google Scholar] [CrossRef]
  42. Tsydenov, B.O.; Trunov, N.S.; Churuksaeva, V.V.; Degi, D.V. Wind effects on deep convection in Lake Baikal during the autumnal thermal bar. Mosc. Univ. Phys. Bull. 2024, 79, 283–290. [Google Scholar] [CrossRef]
  43. Tsydenov, B.; Churuksaeva, V.; Trunov, N.; Bart, A.; Degi, D. The impact of tributary mineralization on deep-water renewal in Lake Baikal during the thermal bar. Water 2025, 17, 1315. [Google Scholar] [CrossRef]
  44. Votintsev, K.K. Hydrochemistry. In Problems of Lake Baikal; Galazii, G.I., Votintsev, K.K., Eds.; Siberian Branch of the USSR Academy of Sciences: Novosibirsk, Russia, 1978; pp. 124–146. [Google Scholar]
  45. Reliable Prognosis. Available online: https://rp5.ru (accessed on 21 August 2023).
  46. Imboden, D.M.; Wüest, A. Mixing mechanisms in lakes. In Physics and Chemistry of Lakes; Lerman, A., Imboden, D., Gat, J., Eds.; Springer: Berlin/Heidelberg, Germany, 1995; pp. 83–138. [Google Scholar]
  47. Golubev, V.A. Geothermics of Lake Baikal; Nauka: Novosibirsk, Russia, 1982. [Google Scholar]
  48. Adcroft, A.; Scott, J.R.; Marotzke, J. Impact of geothermal heating on the global ocean circulation. Geophys. Res. Lett. 2001, 28, 1735–1738. [Google Scholar] [CrossRef]
  49. Joyce, T.M.; Speer, K.G. Modeling the large-scale influence of geothermal source on abyssal flow. J. Geophys. Res. 1987, 92, 2843–2850. [Google Scholar] [CrossRef]
  50. Thompson, L.; Johnson, G. Abyssal currents generated by diffusion and geothermal heating over rises. Deep Sea Res. 1996, 43, 193–211. [Google Scholar] [CrossRef]
  51. Hofmann, M.; Morales Maqueda, M.A. Geothermal heat flux and its influence on the oceanic abyssal circulation and radiocarbon distribution. Geophys. Res. Lett. 2009, 36, L03603. [Google Scholar] [CrossRef]
  52. Dorofeeva, R.P.; Lysak, S.V. Heat flow of Central Asia. In Proceedings of the World Geothermal Congress, Bali, Indonesia, 25–29 April 2010; Available online: https://www.geothermal-energy.org/pdf/IGAstandard/WGC/2010/1308.pdf (accessed on 29 August 2023).
  53. Duchkov, A.D.; Lysak, S.V.; Golubev, V.A.; Dorofeeva, R.P.; Sokolova, L.S. Heat flow and geothermal field of the Baikal region. Geol. Geofiz. 1999, 40, 287–303. [Google Scholar]
  54. Golubev, V.A.; Klerkx, I.; Kipfer, R. Heat flow, hydrothermal vents and static stability of discharging thermal water in Lake Baikal (South-Eastern Siberia). BCREDP Elf Aquitaine 1993, 17, 53–65. [Google Scholar]
  55. Golubev, V.A. Geothermal field and distribution for earthquake source depths in the Baikal Rift zone. Dokl. Earth Sci. 2010, 433, 1083–1087. [Google Scholar] [CrossRef]
  56. Simpson, J.E. Gravity currents in the laboratory, atmosphere, and ocean. Annu. Rev. Fluid Mech. 1982, 14, 213–234. [Google Scholar] [CrossRef]
  57. Tsydenov, B.O. Effect of underwater relief on the dynamics of the autumnal thermal bar. Mosc. Univ. Phys. Bull. 2023, 78, 689–696. [Google Scholar] [CrossRef]
  58. Tsydenov, B.O.; Starchenko, A.V. Numerical model of river-lake interaction in the case of a spring thermal bar in Kamloops Lake. Vestn. Tomsk. Gos. Univ. Mat. Mekh. 2013, 5, 102–115. [Google Scholar]
  59. Kolokoltseva, E.M. Morphometric characteristics of Baikal. In Mesozoic and Cenozoic Lakes of Siberia; Nauka: Moscow, Russia, 1968; pp. 183–188. [Google Scholar]
  60. Piccolroaz, S.; Toffolon, M. Deep water renewal in Lake Baikal: A model for long-term analyses. J. Geophys. Res. Oceans 2013, 118, 6717–6733. [Google Scholar] [CrossRef]
  61. Huang, R.X. Mixing and energetics of the oceanic thermohaline circulation. J. Phys. Oceanogr. 1999, 29, 727–746. [Google Scholar] [CrossRef]
  62. Wunsch, C.; Ferrari, R. Vertical mixing, energy and the general circulation of the oceans. Annu. Rev. Fluid Mech. 2004, 36, 281–314. [Google Scholar] [CrossRef]
  63. Scott, J.R.; Marotzke, J.; Adcroft, A.A. Geothermal heating and its influence on the meridional overturning circulation. J. Geophys. Res. 2001, 106, 31141–31154. [Google Scholar] [CrossRef]
Figure 1. Study area. (a) Cross-section of Lake Baikal. (b) Calculation domain (its location on the top panel is indicated by the Ox axis).
Figure 1. Study area. (a) Cross-section of Lake Baikal. (b) Calculation domain (its location on the top panel is indicated by the Ox axis).
Hydrology 12 00256 g001
Figure 2. Influence of geothermal heating on deep convection. Temperature distributions (°C) simulated without (a) and with (b) geothermal heat flux after days 1, 3, 5, and 10. Details at day 15 in simulations without (c) and with (d) geothermal heat flux.
Figure 2. Influence of geothermal heating on deep convection. Temperature distributions (°C) simulated without (a) and with (b) geothermal heat flux after days 1, 3, 5, and 10. Details at day 15 in simulations without (c) and with (d) geothermal heat flux.
Hydrology 12 00256 g002
Figure 3. Evolution of the thermal bar and its role in deep-water temperature distribution. (a) Dynamics of development of thermal bar fronts (displacement of maximum density temperature positions at the lake surface) in simulations without (blue) and with (red) geothermal heat flux. Isotherms (°C) and maximum density temperature contours (bold) without (b) and with (c) geothermal heat flux after day 10.
Figure 3. Evolution of the thermal bar and its role in deep-water temperature distribution. (a) Dynamics of development of thermal bar fronts (displacement of maximum density temperature positions at the lake surface) in simulations without (blue) and with (red) geothermal heat flux. Isotherms (°C) and maximum density temperature contours (bold) without (b) and with (c) geothermal heat flux after day 10.
Hydrology 12 00256 g003
Figure 4. Wind conditions from 1 November 2015, to 30 November 2015. (a) Wind speed. (b) Wind direction.
Figure 4. Wind conditions from 1 November 2015, to 30 November 2015. (a) Wind speed. (b) Wind direction.
Hydrology 12 00256 g004
Figure 5. A conceptual model of deep-water renewal in Lake Baikal. Surface waters can penetrate to ~200 m depth due to wind forcing, to ~600 m depth due to the effect of cabbeling at the thermal bar, and to maximum depths from thermobaricity triggered by the high mineralization of tributary waters and/or geothermal heating. Water density, lake morphology, and external factors affect the formation of compensatory circulations and intrusions.
Figure 5. A conceptual model of deep-water renewal in Lake Baikal. Surface waters can penetrate to ~200 m depth due to wind forcing, to ~600 m depth due to the effect of cabbeling at the thermal bar, and to maximum depths from thermobaricity triggered by the high mineralization of tributary waters and/or geothermal heating. Water density, lake morphology, and external factors affect the formation of compensatory circulations and intrusions.
Hydrology 12 00256 g005
Figure 6. A map of Lake Baikal with potential zones of high ventilation. Red points indicate zones where Hgeo > 120 mW/m2. Hypothetically, deep-water renewal is the most intense at the eastern shore of Lake Baikal due to abnormally high heat release. At the western shore of Lake Baikal, deep ventilation can occur in the North basin near Capes Malyi Solontsovyi, Zavorotnyi, and Tonkii.
Figure 6. A map of Lake Baikal with potential zones of high ventilation. Red points indicate zones where Hgeo > 120 mW/m2. Hypothetically, deep-water renewal is the most intense at the eastern shore of Lake Baikal due to abnormally high heat release. At the western shore of Lake Baikal, deep ventilation can occur in the North basin near Capes Malyi Solontsovyi, Zavorotnyi, and Tonkii.
Hydrology 12 00256 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tsydenov, B.O. Effect of Geothermal Heating on Deep-Water Temperature in Lake Baikal. Hydrology 2025, 12, 256. https://doi.org/10.3390/hydrology12100256

AMA Style

Tsydenov BO. Effect of Geothermal Heating on Deep-Water Temperature in Lake Baikal. Hydrology. 2025; 12(10):256. https://doi.org/10.3390/hydrology12100256

Chicago/Turabian Style

Tsydenov, Bair O. 2025. "Effect of Geothermal Heating on Deep-Water Temperature in Lake Baikal" Hydrology 12, no. 10: 256. https://doi.org/10.3390/hydrology12100256

APA Style

Tsydenov, B. O. (2025). Effect of Geothermal Heating on Deep-Water Temperature in Lake Baikal. Hydrology, 12(10), 256. https://doi.org/10.3390/hydrology12100256

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