ArticlePDF Available

Interaction of basin-scale topography- and salinity-driven groundwater flow in synthetic and real hydrogeological systems

Authors:

Abstract

Salinization of groundwater has endangered e.g. drinking water supply, agricultural cultivation, groundwater-dependent ecosystems, geothermal energy supply, thermal and hydrocarbon well production to a rising degree. In order to investigate the problem of coupled topography- and salinity-driven groundwater flow on a basin-scale, a systematic simulation set has been carried out in a synthetic numerical model. Detailed sensitivity analysis was completed to reveal the effect of the salinity, permeability, permeability heterogeneity and anisotropy, mechanical dispersivity and water table head on the salt concentration field and the flow pattern. It was established that a saline dome with slow inner convection formed beneath the discharge zone in the base model due to the topography-driven regional fresh groundwater flow. An increase in the salinity or the anisotropy or decrease in the water table variation weakens the role of the forced convection driven by the topography, thus facilitating the formation of a saline, dense, sluggish layer in the deepest zone of the basin. In the studied parameter range, the variation in permeability and dispersivity affects the shape of the saltwater dome to less degree. However, the decrease in permeability and/or the increase in dispersivity advantage the homogenization of the salt concentration within the saline zone and strengthen the coupling between the saltwater and freshwater zone by growing the relative role of diffusion and transverse dispersion, respectively. The interaction of the topography-driven forced and salinity driven free convection was investigated along a real hydrological section in Hungary. Simulation elucidated the fresh, brackish and saline character of the water sampled the different hydrostratigraphic units by revealing the connection between the topography-driven upper siliciclastic aquifer and the lower confined karstic aquifer through faults in high-salinity clayey aquitard. The current study improves the understanding of the interaction between the topography-driven forced and the salinity-driven free convection, i.e. topohaline convection, especially in basin-scale groundwater flow systems.
Journal of Hydrology 609 (2022) 127695
Available online 9 March 2022
0022-1694/© 2022 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-
nc-nd/4.0/).
Research papers
Interaction of basin-scale topography- and salinity-driven groundwater
ow in synthetic and real hydrogeological systems
Attila Galsa
a
,
*
, ´
Ad´
am T´
oth
b
, M´
ark Szij´
art´
o
a
, Daniele Pedretti
c
, Judit M´
adl-Sz˝
onyi
b
a
Department of Geophysics and Space Science, Institute of Geography and Earth Sciences, ELTE E¨
otv¨
os Lor´
and University, Budapest, Hungary
b
J´
ozsef and Erzs´
ebet T´
oth Endowed Hydrogeology Chair, Department of Geology, Institute of Geography and Earth Sciences, ELTE E¨
otv¨
os Lor´
and University, Budapest,
Hungary
c
Dipartimento di Scienze della Terra A. Desio, Universit`
a degli Studi di Milano, Milan, Italy
ARTICLE INFO
This manuscript was handled by C. Corradini,
Editor-in-Chief
Keywords:
Mixed topohaline convection
Salinity-driven groundwater ow
Topography-driven groundwater ow
Numerical modelling
ABSTRACT
Salinization of groundwater has endangered e.g. drinking water supply, agricultural cultivation, groundwater-
dependent ecosystems, geothermal energy supply, thermal and hydrocarbon well production to a rising de-
gree. In order to investigate the problem of coupled topography- and salinity-driven groundwater ow on a
basin-scale, a systematic simulation set has been carried out in a synthetic numerical model. Detailed sensitivity
analysis was completed to reveal the effect of the salinity, permeability, permeability heterogeneity and
anisotropy, mechanical dispersivity and water table head on the salt concentration eld and the ow pattern. It
was established that a saline dome with slow inner convection formed beneath the discharge zone in the base
model due to the topography-driven regional fresh groundwater ow. An increase in the salinity or the
anisotropy or decrease in the water table variation weakens the role of the forced convection driven by the
topography, thus facilitating the formation of a saline, dense, sluggish layer in the deepest zone of the basin. In
the studied parameter range, the variation in permeability and dispersivity affects the shape of the saltwater
dome to less degree. However, the decrease in permeability and/or the increase in dispersivity advantage the
homogenization of the salt concentration within the saline zone and strengthen the coupling between the salt-
water and freshwater zone by growing the relative role of diffusion and transverse dispersion, respectively. The
interaction of the topography-driven forced and salinity driven free convection was investigated along a real
hydrological section in Hungary. Simulation elucidated the fresh, brackish and saline character of the water
sampled the different hydrostratigraphic units by revealing the connection between the topography-driven upper
siliciclastic aquifer and the lower conned karstic aquifer through faults in high-salinity clayey aquitard. The
current study improves the understanding of the interaction between the topography-driven forced and the
salinity-driven free convection, i.e. topohaline convection, especially in basin-scale groundwater ow systems.
1. Introduction
Hydrogeological problems that we have to solve in practical tasks
often refer to complex systems, but our solutions are necessarily
simplied. Unfortunately, it means that we neglect many effects and
processes and, on the contrary, put in the focus what we think is sig-
nicant. To exclude these uncertainties and better understand some
physical uid ow processes and their interactions, we started a sys-
tematic analysis of superimposed uid driving forces and their impacts
on basin-scale ow patterns. We have already examined the combina-
tion of topography and thermal buoyancy that drive groundwater ow
in a theoretical basin to better understand the inuencing factors and
the consequences of the interaction (Szij´
art´
o et al., 2019). The lesson
learnt from this systematic theoretical analysis, we applied to a real
hydrogeological system focusing on these driving forces (Szij´
art´
o et al.,
2021). In this paper, we continue the analysis with the interactions of
topography and salinity as driving forces of uids on basin-scale ow
systems and demonstrate it for an actual hydrogeological situation
without trying to extend the problem to any other driving force.
The importance and understanding of salt transport processes in
groundwater ow systems has developed parallel to the increasing
natural resource demands of humankind. In coastal environments, the
rising number and yield of wells producing freshwater entail the sub-
surface transgression of sea water and subsequent deterioration of water
* Corresponding author.
E-mail address: attila.galsa@ttk.elte.hu (A. Galsa).
Contents lists available at ScienceDirect
Journal of Hydrology
journal homepage: www.elsevier.com/locate/jhydrol
https://doi.org/10.1016/j.jhydrol.2022.127695
Received 12 May 2021; Received in revised form 16 February 2022; Accepted 4 March 2022
Journal of Hydrology 609 (2022) 127695
2
quality has become a daily problem of inhabitants living near the
coastlines (e.g. Barlow and Reichard, 2010; Werner et al., 2013; Hussain
et al., 2019). In an inland environment, salinization triggered by
groundwater overexploitation and pumping deeper and more saline
reserves gives difculties for agriculture, drinking water supply and the
groundwater-dependent ecosystem protection (e.g. Jalali, 2007; Rodrí-
guez-Rodríguez and Benavente, 2008; Herbert et al., 2015; McFarlane
et al., 2017; Payen et al., 2016; Schuler et al., 2018; Singh, 2020).
Interaction of salt and fresh groundwater can lead to carbonate disso-
lution resulting in permeability increase and extended reservoirs for
hydrocarbon and thermal water production (Dublyansky, 1995; Kamb-
esis and Coke, 2013; He et al., 2017; Zhang et al., 2020a; Smith et al.,
2021). On the other hand, it can trigger sinkhole development causing
serious environmental problems and damage (Oz et al., 2016; Karaki
et al., 2019). The long-term negative impact of salinization on aquifers
may be exacerbated by the effects of climate change (Havril et al., 2018),
e.g. in terms of sea level rising or change in intensitydurationfre-
quency of precipitation events (e.g. De Paola et al., 2014; Sun et al.,
2019). In addition, lithium recovery from brine attempts to satisfy the
increasing requirements of the battery industry (e.g. Kesler et al., 2012;
Marazuela et al., 2018; García-Gil et al., 2019).
Dissolved solids content increases the water density; hence it affects
the groundwater ow driven by buoyancy force. Nield (1968) analyti-
cally determined the conditions of the onset of haline convection in a
homogeneous, horizontal innite porous layer, and Van Dam et al.
(2009) uncovered the rst eld documentation of precipitation-induced
saltwater ngering as a manifestation of natural free convection in a
porous medium. However, the water table elevation varies from place to
place, thus the topography-driven groundwater ow is always present in
the upper part of the terrestrial basins (T´
oth, 2009). Therefore, it also
acts in these terrestrial environments with elevated dissolved solid
content. Consequently, these cases require mathematical models (typi-
cally, numerical models) to evaluate the coupled effects of topography-
and salinity-driven groundwater ow.
The concept of coupled topography- and salinity-driven groundwater
ow has been adopted to explain the high salt content of the submarine
groundwater discharge (SGD) on a local scale (Konikow et al., 2013),
and to elucidate the variability and the distance of SGD in a heteroge-
neous coastal volcanic aquifer near the Big Island of Hawaii, USA
(Kreyns et al., 2020). Duffy and Al-Hassan (1988) revealed the
groundwater circulation in a homogeneous medium as a balance be-
tween the rainwater recharge in surrounding mountains and the salt-
water counter ow beneath the playa in the Great Basin, Utah, USA. The
two-dimensional, homogeneous and isotropic numerical model was
recalculated using multicomponent reactive transport simulations to
infer the evolution of the evaporite precipitation and brine composition
(Hamann et al., 2015). Topography-driven forced and salinity-driven
free convection processes were modelled in a simple geological envi-
ronment to trace the brine inltration beneath a saline disposal basin in
Australia (Simmons and Narayan, 1997). Zhang et al. (2020b) presented
a sensitivity analysis to study the effect of the salinity, the hydraulic
conductivity and the dispersivity on the positions of stagnation points
and the hierarchy of the groundwater ow system. However, it seems
from the published papers that topography effects are examined on a
local scale or for simple basins, but they are commonly neglected.
Therefore, some of the models remained oversimplied in terms of
omittance of variations in topography and water table.
Although numerical models are acknowledged tools to quantify the
interaction of topography- and salinity-driven groundwater ow,
several questions are still unresolved regarding the sensitivity of these
models to the key parameters feeding the governing equations. Small
variations in critical parameters, such as relative density differences due
to salinity, variations of intrinsic permeability, anisotropy and me-
chanical dispersivity or change in water table amplitude may generate
great deviations in model outputs. As such, given the ubiquitously
difcult aquifer characterization that renders model parameterization
always uncertain, high sensitivity to model parameters can create high
uncertainty in the decisions to be made about the water management of
an aquifer. While disentangling the model response to the input pa-
rameters is therefore of the uppermost importance, this task is in turn
severely complicated by the computational burden required to resolve
the nonlinear fully coupled density-dependent groundwater ow
equations.
The main aim of this paper is to investigate, through detailed nu-
merical modelling, how the interaction of topography-driven forced and
salinity-driven free convection and the different model parameters in-
uence the salt concentration and groundwater ow pattern at basin-
scales and thus how the decisions made from the model responses
may vary depending on the model sensitivity. Other uid driving forces,
e.g. temperature difference, compaction, compression, were neglected
in order to focus on and highlight the interaction of topography- and
salinity-driven regional groundwater ow and systematically examine
the effects of this interaction rstly. Once the processes and phenomena
are comprehensively understood and synthesized, other uid driving
processes can also be coupled. Therefore, in the rst part of the paper, a
two-dimensional, synthetic simulation set is presented, where the effect
of salinity (dissolved solid content), permeability, depth-dependence
and anisotropy of permeability, dispersivity and water table amplitude
is qualied and quantied systematically by monitoring parameters, e.g.
the Darcy ux and the salt concentration for both the entire basin and
the fresh/saltwater zones. In the second part, the mixed topohaline
convection simulation is carried out to demonstrate its effect in a real
environment, along a hydrogeological section in Hungary based on
M´
adl-Sz˝
onyi et al. (2019). The fundamental goal is to reveal the evo-
lution of marine-origin pore water and to explain the formation of the
fresh-, brackish and saltwater saturated units on a basin scale.
2. Materials and methods
2.1. Model development
The solution of the combined topography- and salinity-driven
groundwater ow requires the coupled handling of the continuity
equation (1), the Darcys Law (2) and the mass transport equation (3)
governing the mass conservation, the momentum conservation and the
transport of the dissolved solid matter by advection, molecular diffusion
and mechanical dispersion, respectively (Delgado, 2012; Nield and
Bejan, 2017),
Φ
∂ρ
w
t+ ∇(
ρ
wq) = 0(1)
q= − k
μ
[∇p
ρ
wg](2)
Φ
c
t= − qc+ ∇ DdiffI+Ddisp c(3)
where q, p and c denote the unknown Darcy ux, the pressure and the
dissolved salt concentration, while Φ, k,
μ
, g, D
diff
, D
disp
, t and I are the
porosity, the diagonal permeability tensor, the dynamic viscosity of the
water, the gravitational acceleration, the molecular diffusion, the me-
chanical dispersion, the time and the identity matrix, respectively. The
parameters of the base model are summarized in Table 1. The salinity-
dependence of the viscosity was neglected in the course of the simula-
tion, the dispersion tensor was dened by the longitudinal and the
transverse dispersivity,
α
L
and
α
T
, respectively. The transverse dis-
persivity was afxed to the longitudinal one,
α
T
=
α
L
/10. The salt con-
tent was characterized by the non-dimensional concentration ranging
between c =0 and 1 to focus on the density variation of the water. A
simple linear relation was supposed between the water density and the
concentration (e.g. Duffy and Al-Hassan 1988, Nield and Bejan 2017),
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
3
ρ
w=
ρ
0(1+βc)(4)
where
ρ
0
=1000 kg/m
3
is the freshwater density and β denotes the
relative density difference between freshwater (c =0) and saltwater (c
=1).
A simple two-dimensional numerical model was built to reveal the
physical background of salinity transport. The basin-scale model is
characterized with a length of L =40 km and a mean depth of d =5 km
(Fig. 1) to test the outcomes of synthetic simulations for a real hydro-
geological system. The boundary conditions for the ow were no-ow
along the sidewalls and the bottom, while the regional groundwater
ow was driven by a cosinusoid water table (e.g. Domenico and Pal-
ciauskas, 1973; Wang et al., 2015; Zhang et al., 2018),
zwt(x) = Acos
π
x
L(5)
with an amplitude of A =50 m, where x denotes the horizontal coor-
dinate. The water table, z
wt
corresponds to the top of the model domain.
No-ux boundary conditions for the concentration were prescribed
along the sidewalls, while the concentration was xed along the hori-
zontal boundaries. It was c =1 at the bottom and c =0 on the surface,
thus diffusive and dispersive ux was allowed across the horizontal
boundaries. The former represents salinity inux from a deep imper-
vious formation, and the latter symbolizes e.g. the freshwater inltration
and the salt precipitation from the groundwater. The initial condition for
the ow was obtained from Eqs. (1)(2) with reference water density
(Fig. 1.a), and the initial condition for the concentration was calculated
from Eq. (3) without the advection term (q =0), which is the solution of
the diffuse problem resulting in linear vertical concentration prole
(Fig. 1.b).
The coupled problem (Eq. (1)(3)) was solved by COMSOL Multi-
physics v4.2, a nite element numerical software package (Zimmer-
mann, 2006. The model domain was discretized by triangle elements
with a maximum size of 100 m and 8 boundary layer elements were used
to minimize the inaccuracy due to the weak resolution of the high
gradient zones. Thus, the mesh consisted of 57,159 nite elements for
the base model, which varied between 56,88080,023 as the water table
amplitude was changed from 0 m to 500 m. The pressure and the con-
centration were discretized within the elements by quadratic and linear
polynomials, respectively. Simulations were run until models reached
the stationary solution, which required typically 1 Myr, ranging from
0.5 to 50 Myr (e.g. for low permeability, high permeability anisotropy or
slight water table amplitude). Initial time steps were increased expo-
nentially to 1000 yr to stabilize the transition effects, after that the
maximum time step was xed as 100 yr. Simulations were run on an
Intel Workstation with 24 cores, requiring approx. 12 days CPU time
and 4 GB memory for one model. Before the synthetic simulation set, the
coupled model was expansively tested and conrmed in a horizontal and
inclined layer, and solutions were compared with the analytical and
numerical results of Weatherill et al. (2004). Tests resulted in perfect
accordance focusing on the onset (critical Rayleigh number) and the
form (transition from multi- to unicellular convection) of free haline
convection.
Table 1
Parameters of the base model.
Description Symbol Value Unit
Porosity Φ 0.1 1
Reference water density
ρ
0
1000 kg/m
3
Dynamic viscosity of the water
μ
10
-3
Pa s
Molecular diffusion D
diff
10
-9
m
2
/s
Relative density difference β 0.01 1
Horizontal permeability of the matrix k
x
10
-12
m
2
Factor of depth-dependent permeability γ 1 1
Permeability anisotropy coefcient
ε
1 1
Longitudinal dispersivity
α
L
0 m
Transverse dispersivity
α
T
0 m
Water level amplitude A 50 m
Fig. 1. Boundary and initial conditions for (a) the ow (blue) and (b) the salt concentration (green) in the base model (Table 1). (a) Streamlines of the Darcy ux
(white) are magnitude controlled, red arrows illustrate the ow direction. (b) Isocontour of c =0.5 (black line) represents the boundary between the saltwater and
the freshwater zone. The model length is 40 km, the average model depth is 5 km, the vertical exaggeration is 1.
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
4
2.2. Model parameterization
Firstly, we systematically investigated the effect of the following
relevant parameters (Table 2) resulting in 56 time-dependent models:.
1. the relative density difference due to salinity, β;
2. the permeability, k;
3. the factor of depth-dependence of permeability, γ;
4. the permeability anisotropy,
ε
;
5. the mechanical dispersivity,
α
L
and
α
T
; and
6. the water level amplitude, A.
The six selected parameters are usually uncertain given the limited
characterization and exhibit strong scale dependency. For instance, the
salinity of groundwaters may vary both in time and space. In Hungary, it
is typical that the salt concentration increases with depth owing to the
deep, marine sedimentary environment, while salinity is lower in
shallow strata complexes due to lacustrine sedimentation and the
intense meteoric water inltration (e.g. Vars´
anyi and ´
O.Kov´
acs, 2009;
Szocs et al., 2013). In general, the groundwater salinity can vary from
0 to even 40% including fresh, brackish, saline and brine waters
(Deming, 2002). In this paper, the effect of salinity on the water density
is parameterized by a factor of β, which presents the relative density
increase of the water. The aquifer permeability (or hydraulic conduc-
tivity) can vary over several orders of magnitude in relatively short
spatial scales to the effects of geological heterogeneity (Sanchez-Vila
et al., 2006; Pedretti et al., 2016; Liu et al., 2016). One of the most
common heterogeneities of permeability is its pressure-dependence
(Ehrenberg and Nadeau, 2005). Permeability decreasing exponentially
with depth was dened by
k(z) = k0explnγz
d(6)
to characterize the pressure-dependence of the permeability (Jiang et al.
2009; Jiang et al., 2011), where k
0
denotes the surface permeability and
z is the vertical coordinate pointing upwards. Anisotropy also controls
the time evolution of ow and transport patterns in aquifers and de-
pends on the nature and connectivity of the hydrogeological bodies
(Cirpka et al., 2015; Pedretti et al., 2014). In siliciclastic sedimentary
basins, where the sub-horizontal aquifer and aquitard layers alternate,
the effective hydraulic conductivity strongly depends on the ow di-
rection. The horizontal permeability, in such environments, can exceed
the vertical permeability by several orders of magnitude (e.g. Galsa,
1997; Hoque and Burgess, 2020). Mechanical dispersion is present in the
subsurface both in microscopic scale (grain size) and local scale (e.g.
quasi-impermeable zones, lens). In aquifers (e.g. fractured media, sili-
ciclastic and uvial sediments), the longitudinal dispersivity is scale-
dependent (Gao et al., 2012) and varies within a wide range (Gelhar
et al., 1992; Schulze-Makuch, 2005; Vanderborght and Vereecken,
2007). The variation of the water level inuences both the ow pattern
and the salinity of the groundwater as one of the main driving forces (e.
g. T´
oth, 2009).
The inuence of forming dense, saline zone/layer on the concen-
tration eld and the ow pattern was studied both qualitatively and
quantitatively. Monitoring parameters were dened to characterize the
groundwater ow system by the average Darcy ux (q
av
), concentration
(c
av
), hydraulic head (h
av
), as well as the relative area (A
s
, A
f
), the Darcy
ux (q
s
, q
f
) and the concentration (c
s
, c
f
) computed for both the saltwater
(subscript s) and the freshwater (subscript f) zone. The salt- and fresh-
water zones were separated by an arbitrary boundary of the solute
content of c =0.5. After the synthetic numerical simulations, the
topography- and salinity-driven groundwater ow model was applied
along a 2D hydrogeological section in Hungary taking into consideration
the outcome of the synthetic model runs.
3. Synthetic model results
3.1. Effect of relative density difference
As a rst step, the effect of density increase due to salinity on the
concentration eld and the ow pattern was studied. The results show
that the saltwater is swept out from the basin by the topography-driven
groundwater ow, when β =0, i.e. the water density is independent of
the salt content (Fig. 2.a). A shrank zone beneath the discharge area can
survive the regional freshwater circulation near the stagnation point
(see e.g. An et al., 2015). Even though the saltwater is denser by only
1%, an extended saline zone (c 0.5, contoured by a black line) forms
beneath the discharge area in the base model (Fig. 2.b, Table 1). The
concentration distribution within the saline zone is rather diffuse sug-
gesting that there is no intense inner ow. As the salt content (i.e. water
density due to salinity) increases, a dense layer evolves right above the
model bottom (Fig. 2.dg), which consequently reduces the diffuse
transport from the base below resulting in an average concentration
decrease in the basin.
Fig. 3 illustrates how the dense, saline zone modies the regional,
unidirectional ow pattern. When the concentration-dependence of the
water density is neglected (Fig. 3.a), the Darcy ux eld is equivalent to
the solution of the simple Darcy ow (see Fig. 1.a). When the water
density is not neglected (Fig. 3.bd), the ow eld strongly depends on
the presence of the saline zone. The most striking inuence of the saline
zone on the ow pattern is that the topography-driven regional ow is
suppressed, but the magnitude of the Darcy ux is not affected by the
saline dome. On the other hand, slow circulation forms within the
saltwater domain. This sluggish inner convection has opposite vorticity
compared to the regional groundwater ow. The pattern of the inner
convection depends strongly on the saline content, as measured by β.
The higher the salinity contrast with the freshwater, the more pro-
nounced the recirculation zone. Note that β =3% produces an inner
convection zone about two times more extended compared to β =1%.
The model is therefore highly sensitive to limited salinity changes.
The given problem of topography- and salinity-driven groundwater
ow tends towards a stationary solution, which is plotted against the
relative density difference, β (Fig. 4). As the water density increases (see
Eq. (4)) from β =0 to 0.2, the domain-averaged Darcy ux (q
av
) is
reduced by 43% (Fig. 4.a). The domain-averaged concentration (c
av
)
shows a non-monotonic character with a maximum at about β =23%
(Fig. 4.b), when a considerable saltwater zone forms beneath the
discharge area (Fig. 2.c). At a lower value of β, the salt content cannot
resist the topography-driven groundwater ow, while at a higher value
of β a thin, dense salty layer evolves in the deepest part of the basin. The
domain-averaged hydraulic head (h
av
) becomes more negative owing to
the increasing water density (Fig. 4.c).
The specic behaviour of the saltwater and freshwater zone sepa-
rated by the arbitrary isocontour of c =0.5 was also investigated. As β
increases, the relative area of the dense, saline zone (A
s
, the area normed
by the total area of the basin) shrinks, apart from the model of β =0,
when the saltwater is swept out from the basin (Fig. 2.a and 4.d). On the
contrary, the relative area of the complementary freshwater zone, A
f
(also normed by the total area of the basin) increases in the same way as
A
s
decreases. Fig. 4.e illustrates that the inner convective velocity of the
saltwater zone (q
s
) is lower up to about two orders of magnitude
Table 2
Studied parameters.
Description Symbol Value Unit
Relative density difference β 00.2 1
Matrix permeability k 10
-14
10
-12
m
2
Factor of depth-dependent permeability γ 1100 1
Permeability anisotropy coefcient
ε
11000 1
Longitudinal dispersivity
α
L
(=10
α
T
) 0100 m
Water level amplitude A 0500 m
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
5
compared to the velocity of the freshwater zone (q
f
), which remains
quite constant (q
f
210
-8
m/s). Specically, the minimum saltwater
velocity of q
s
=310
-10
m/s was found when β =2%, i.e. at the peak of
maximum domain-averaged concentration, c
av
. This point also corre-
sponds to the lowest average concentration of the dense zone (c
s
0.65)
(Fig. 4.f), when an extended saltwater dome with thick transition zone
forms beneath the discharge zone (Fig. 2.c).
Fig. 2. Stationary solution of salt concentration at different relative density differences, β =020%. Other parameters are dened in the base model (Table 1). Black
isocontour of c =0.5 represents the boundary of the saltwater and the freshwater zones.
Fig. 3. Darcy ux for different relative density differences due to salinity at t =100 kyr. Streamlines (white not magnitude controlled) and arrows (red
normalized) represent the ow direction.
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
6
3.2. Effect of permeability
3.2.1. Permeability in a homogeneous and isotropic medium
The role of rock permeability on the salinity was investigated using
the base model (Table 1), while the permeability (horizontal and ver-
tical) was decreased from k =10
12
m
2
to 10
14
m
2
. Fig. 5 illustrates that
the change in permeability has no drastic inuence on the salt concen-
tration eld. By decreasing the permeability, the ow linearly slows
down (Fig. 6.a), according to Eq. (2). When the Darcy ux is reduced in
the freshwater zone, the diffusion ux across the boundary of the salt-
and freshwater zones is retained. Lower diffusion ux from the saltwater
zone results in higher salt content (Fig. 6.b) and a more homogeneous
concentration eld (Fig. 5.ad). As the average salt content increases,
the hydraulic head decreases due to the enhanced water density (Fig. 6.
c).
Decreasing the permeability from k =10
-13
m
2
to 10
-14
m
2
, the
diffusion ux becomes commensurable to the order of advection, thus
the diffusion zone between the salt- and freshwater thickens (Fig. 5.dg).
A thickening diffusion zone appears above the model bottom, as well.
While the change in permeability by two orders of magnitude drives an
equal reduction in the average Darcy ux, the maximum of the diffusion
ux reduces by only a factor of 5, specically from 810
-11
to 1.410
-
11
m/s.
The permeability reduction inuences the Darcy ux of the saltwater
and freshwater zones in a different manner (Fig. 6.e). The ow slows
down in both zones, but the gradient of the Darcy ux curve is higher for
the freshwater zone than the saltwater zone (i.e. dq
f
/dk >dq
s
/dk),
consequently, the difference in the Darcy ux decreases to one order of
magnitude. The effect of two processes appears both in the concentra-
tion and the relative area curves (Fig. 6.d and f). The homogenization (in
the cases of k =10
-12
10
-13
m
2
) increases the salt content and the area of
the saline zone. Then the diffusion zone thickening (in the cases of k =
10
-13
10
-14
m
2
) hinders the rate of the area growth, and it leads to a
slight concentration decrease within the saltwater zone. Nevertheless,
the average salt content of the basin shows a slight increase (Fig. 6.b)
owing to the slowing down regional groundwater ow.
3.2.2. Permeability heterogeneity
Owing to the increasing pressure, the permeability shows a decrease
downwards both in carbonates and sandstones, which is frequently
Fig. 4. Stationary solutions of the monitoring parameters plotted against the relative density contrast, β. (a) The average Darcy ux, (b) the average salt concen-
tration and (c) the average hydraulic head. (d) The relative area, (e) the Darcy ux and (f) the concentration averaged for the saltwater (red) and the freshwater
(blue) zones.
Fig. 5. Stationary solution of salt concentration at different permeability of the matrix, k =10
12
10
-14
m
2
. Other parameters are dened in the base model (Table 1).
Black isocontour of c =0.5 represents the boundary of the saltwater and the freshwater zones.
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
7
approximated by an exponential trend. In Eq. (6) γ represents the factor
of the permeability decay, that is the ratio of the surface and bottom
permeability. Fig. S1 (Supplementary material) illustrates the effect of γ
on the concentration eld within the interval of γ =1100. Increasing γ
has a very similar inuence on salinity as decreasing permeability in the
previous part. When the permeability decreases with depth by a factor of
110 (Fig. S1.ad), a homogenization occurs in the saltwater zone due to
the lower Darcy ux in the freshwater zone (Fig. S2.e in Supplementary
material) and the reduced diffusion ux across the boundary of the salt-
and freshwater zones. The only difference between Fig. 5 and Fig. S1 is
that the homogenization evolves not in the whole saltwater dome, but it
has a depth-dependence (γ >1). This process results in a considerable
increase in average concentration (Fig. S2.b), the area and the concen-
tration of the saltwater zone (Fig. S2.d and f).
Decreasing permeability with depth by 12 orders of magnitude (γ =
10100) brings forward another phenomenon (Fig. S1.dg). The Darcy
ux decrease with depth is so pronounced that the diffusion approxi-
mates the advection in the mass transport, thus a diffusion zone forms
between the fresh- and saltwater zones, as well as along the bottom
boundary. It enhances the salinity in the freshwater zone (Fig. S2.f).
Because of the depth-dependent permeability, the diffusion zone thins
upwards, and in the near-surface region, the intense ow effectively
erodes the top of the saltwater dome. Consequently, a slight reduction
can be observed in the average concentration (Fig. S2.b), the area and
the concentration of the saltwater dome (Fig. S2.d and f). Depth-
dependent permeability induces a continuous Darcy ux and hydrau-
lic head decrease in the whole basin (Fig. S2.a and c), however, the ow
slows down less in the freshwater zone, which occupies the shallower
zones of the basin (Fig. S2.e).
3.2.3. Permeability anisotropy
The permeability anisotropy coefcient, which is the ratio of the
horizontal to vertical permeability, was varied ranging between
ε
=
11000, while the value of the horizontal permeability was xed, k
x
=
10
-12
m
2
. Weakening the vertical permeability leads to slower ow
which extends the simulation time. For models, where 2 Myr was not
enough to reach the stationary solution, and the variation of the moni-
toring parameters was very slow, a steady-state solution was calculated
from the end of the time-depending solution.
It was found that the anisotropy strongly modies the shape of the
saltwater zone. Fig. 7 displays the stationary solution of the salt con-
centration at different permeability anisotropies. As the anisotropy in-
creases from
ε
=1 to 50 (Fig. 7.ad), the boundary between the salt- and
the freshwater zones changes from convex to concave, since the
groundwater ow becomes constrained to shallower zones. Growing the
anisotropy beyond the value of
ε
=100, the penetration of the regional
groundwater ow decreases, and a sluggish layer evolves in the deep
part of the basin. In the case of
ε
=1000, this sluggish zone becomes
thicker (approx. 1.52 km). The concentration in the salt layer shows a
diffuse pattern indicating that there is no relevant advective mass
transport.
Monitoring parameters point out the dichotomy of the effect of
anisotropy. As
ε
increases, the average Darcy ux decreases with an
increasing gradient (Fig. 8.a). The change in the morphology of the
saltwater dome (
ε
=150) is reected in a slight decrease of the average
concentration and the relative area of the saline zone (Fig. 8.b and d).
Within this interval, the Darcy ux in the saltwater zone is nearly con-
stant. At higher anisotropy values (
ε
50) the topography-driven fresh
groundwater does not reach the bottom of the basin, which facilitates
the formation of the deep, saline, sluggish, dense layer above the bot-
tom. This thickening layer enhances the average salt content (Fig. 8.b),
which results in a relevant head decrease (Fig. 8.c). The size of the saline
zone increases (Fig. 8.d), and the Darcy ux decreases steeply (Fig. 8.e).
As a result of the appearance of the diffuse transition across the salt- and
freshwater boundary, the salt content of the freshwater zone grows
(Fig. 8.f).
3.3. Effect of dispersivity
Longitudinal dispersivity was raised from the value of the base
model,
α
L
=0 to 100 m (Fig. 9), meanwhile the transverse dispersivity
varied together with
α
L
as
α
T
=
α
L
/10. Even though the longitudinal
dispersivity is only
α
L
=1 m, the concentration distribution within the
saline dome becomes more homogeneous (Fig. 9.b). By increasing
α
L
to
10 m, the salt content is getting higher owing to (1) the enhanced lon-
gitudinal dispersion of inner convection within the saltwater zone, and
(2) the higher transverse dispersion across the bottom below the saline
dome (Fig. 9.ae). As the dispersivity was increased further (
α
L
=
10100 m), the boundary separating the salt- and the freshwater zone
thickens due to the incremented transverse dispersion. In addition, a
Fig. 6. Stationary solutions of the monitoring parameters plotted against the matrix permeability, k. (a) The average Darcy ux, (b) the average salt concentration
and (c) the average hydraulic head. (d) The relative area, (e) the Darcy ux and (f) the concentration averaged for the saltwater (red) and the freshwater (blue) zones.
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
8
thin salty layer evolves along the bottom of the model beneath the
freshwater zone. Mechanical dispersion facilitates the solute content
mixing both within and between the zones, and consequently, it en-
hances salt transport. The surface Sherwood number (Weatherill et al.,
2004) increases from 146 to 1410 within the range of
α
L
=0100 m.
Fig. 10 quanties the above-mentioned processes. Within the inter-
val of
α
L
=010 m, the saltwater zone enlarges from A
s
=20% to 30%
(Fig. 10.d), its solute content increases by dispersive mixing (Fig. 10.f),
and the ow intensies due to the stronger linkage between the salt
dome and the topography-driven freshwater zone (Fig. 10.e). The
broadening salt dome reduces the average Darcy ux of the basin ow
by approx. 10% (Fig. 10.a), enhances the average salt content consid-
erably (Fig. 10.b), which results in lower hydraulic heads (Fig. 10.c).
When the longitudinal dispersivity is ranged from
α
L
=10 to 100 m,
there is no such a signicant variation in the monitoring parameters. As
the transition zone thickens due to the effective transverse dispersion
ux (between the salt- and freshwater zones, and along the bottom of the
basin), the Darcy ux difference and the solute content difference
between the two zones moderate (Fig. 10.e and f).
3.4. Effect of water table amplitude
The inuence of topography-driven groundwater ow can be
directly stimulated by the water table gradient. In the synthetic model,
the amplitude of the water table was varied between A =0 and 500 m to
observe the modication of the saltwater zone. In the models where the
amplitude was higher than that of the base model (A >50 m), the
regional groundwater ow was intense enough to press the saline dome
to the left side (Fig. 11.eg). The saline zone shrank toward the stag-
nation point and its density increased to be able to resist the regional
freshwater ow. For lower water table amplitudes (A <50 m) the
topography-driven ow weakens, thus the saline zone tends to pervade
the deeper part of the basin to form a dense, saline layer (Fig. 11.ac). It
is worth noting when the amplitude is approx. A =10 m, the
topography-driven groundwater ow is weak enough to press the saline
dome to the discharge zone, but strong enough to advect the brackish
Fig. 7. Stationary solution of salt concentration at different permeability anisotropy,
ε
=11000. Other parameters are dened in the base model (Table 1). Black
isocontour of c =0.5 represents the boundary of the saltwater and the freshwater zones.
Fig. 8. Stationary solutions of the monitoring parameters plotted against the permeability anisotropy,
ε
. (a) The average Darcy ux, (b) the average salt concen-
tration and (c) the average hydraulic head. (d) The relative area, (e) the Darcy ux and (f) the concentration averaged for the saltwater (red) and the freshwater
(blue) zones. The end of the time-dependent solution is denoted by diamond, the stationary solution computed from the end of the time-dependent solution is
denoted by a circle.
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
9
water from the model (Fig. 11.b). As a result, the size of the saline zone is
reduced. At A =1 m, the regional ow is weak, which results in a nearly
diffuse density-based layering (Fig. 11.a).
Stationary numerical solutions show a linear relationship between
the average Darcy ux and the water table head, following Darcys law
formulated in Eq. (2) (Fig. 12.a). The more intense the regional ow is,
the lower the average salt concentration in the basin is noticed (Fig. 12.
b), and lower salt content causes higher hydraulic heads (Fig. 12.c).
Hydraulic head higher than zero is the consequence of that the area
dominated by recharge is enhanced, while the area dominated by
discharge is reduced. At A =50 m the relative area of the zone having h
>0 is 50.3%, while it is 53.1% at A =500 m. As the water table
amplitude decreases, the average head becomes more negative, for the
model of A =1 m, only 15.1% of the model domain is characterized by a
hydraulic head higher than zero.
Decreasing water table amplitude eventuates in that the Darcy ux is
reduced both in the fresh- and the saltwater zone (Fig. 12.e). The ow
within the saline zone is slower by 12 orders of magnitude. The Darcy
ux in the model without water table variation (A =0) tends to zero,
therefore it is not displayed in Fig. 12.e. The size of the saline zone in-
creases by weakening the topography-driven ow (A), and it converges
to A
s
=A
f
=0.5 at A =0 (diffuse solution) in Fig. 12.d. However, a local
extremum appears in the curves near A =10 m, where a transition exists
in the morphology of the saltwater zone, and the saline dome transforms
into a deep, saline layer. The salt content of the freshwater zone is
practically zero if A >50 m, the regional ow sweeps out the saltwater
from the topography-dominated groundwater zone (Fig. 12.f). Since the
salt content of the saline zone increases by A, a sharp boundary evolves
between the two zones in the absence of mechanical dispersion (Fig. 11.
g). On the contrary, reducing A the concentration of the two zones
converge to the diffuse solution, c
s
=0.75 and c
f
=0.25.
Fig. 9. Stationary solution of salt concentration at different longitudinal dispersivity,
α
L
=0100 m. Transverse dispersivity is
α
T
=
α
L
/10, other parameters are
dened in the base model (Table 1). Black isocontour of c =0.5 represents the boundary of the saltwater and the freshwater zones.
Fig. 10. Stationary solutions of the monitoring parameters plotted against the longitudinal dispersivity,
α
L
. (a) The average Darcy ux, (b) the average salt con-
centration and (c) the average hydraulic head. (d) The relative area, (e) the Darcy ux and (f) the concentration averaged for the saltwater (red) and the freshwater
(blue) zones.
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
10
4. Demonstration in a realistic environment
A two-dimensional section was selected to demonstrate the interac-
tion of topography- and salinity-driven groundwater ows in a real
hydrological environment in Hungary. The area was extensively inves-
tigated by M´
adln´
e Sz˝
onyi et al. (2018), M´
adl-Sz˝
onyi et al. (2019) and
M´
adln´
e Sz˝
onyi (2020) studying the geological and structural settings,
hydrostratigraphy, examining the hydrocarbon and hydrogeological
systems, analysing pressure and salinity data from wells etc. in detail. In
addition, a two-dimensional numerical simulation was carried out tak-
ing into the consideration only the effect of topography and the advec-
tive heat transport (M´
adl-Sz˝
onyi et al., 2019). Therefore, in this part,
only the most relevant information is summarized, which is necessary to
elucidate the background of the numerical ow and transport simulation
and time-variation of salinity.
Fig. 13 illustrates the location of the section in the NW part of the
Great Hungarian Plain crossing the G¨
od¨
oll˝
o Hills from SW to NE direc-
tion in Hungary. The section length is 48.5 km, the depth of the model
varies between 3590 and 3625 m depending on the water table topog-
raphy. The geology of the area was grouped into ve hydrostratigraphic
units (HsU) complemented by the original pore uid content (Table 3):
the karstied Triassic carbonate (HsU1); the Eocene limestone (HsU2);
the clayey Oligocene aquitard (HsU3); the siliciclastic Miocene aquifer-
aquitard group (HsU4) and the undifferentiated Upper Miocene-
Pliocene-Quaternary group (HsU5) (M´
adln´
e Sz˝
onyi, 2020). In the
hydrostratigraphic division, HsU1 and HsU5 are considered as the main
aquifers, and HsU3 represents the main aquitard based on geological
and hydrogeological descriptions and evaluation of pumping test data
(Rman and T´
oth, 2011; Garamhegyi et al., 2020). The hydrostrati-
graphic section of the study area and the location of faults are originally
based on seismic interpretation of MOL Plc. (M´
adln´
e Sz˝
onyi et al.,
2013).
The most signicant difference between the present model and the
model of M´
adl-Sz˝
onyi et al. (2019) is that the evolution of the salinity is
taken into account here following Eq. (4). Since most hydrostratigraphic
units were deposited in a marine environment, we simplied the real
Fig. 11. Stationary solution of salt concentration at different water table amplitudes, A =0500 m. Other parameters are dened in the base model (Table 1). Black
isocontour of c =0.5 represents the boundary of the saltwater and the freshwater zones.
Fig. 12. Stationary solutions of the monitoring parameters plotted against the water table amplitude, A. (a) The average Darcy ux, (b) the average salt concen-
tration and (c) the average hydraulic head. (d) The relative area, (e) the Darcy ux and (f) the concentration averaged for the saltwater (red) and the freshwater
(blue) zones. (e) The values of q
s
and q
f
at A =0 tend to be zero, thus they are not marked in the logarithmic scale.
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
11
situation and the initial condition for salt concentration was c =1 in
each unit at a relative density difference of β =3% (TDS of approx.
40,000 mg/l). The boundary conditions for the ow are a linearized
surface water table with a maximum elevation of 125 m at x =20 km
representing the G¨
od¨
oll˝
o Hills (Fig. 14.a). Along the left and right
boundary of the model the head is xed at 100 m and 90 m, respectively,
showing the ow-through character of the model. The model is closed
from below by a no-ow boundary. Boundary conditions for the salt
Fig. 13. The location of the study area: (a) Hungary in Europe, (b) the topography of the Pannonian Basin and the broad vicinity of the study area (Horv´
ath et al.,
2006) and (c) the topography and the outcrops of Mesozoic carbonates with the simulated hydrogeological cross-section in EOV coordinate system (National
Hungarian Grid in meter unit).
Table 3
Description and model parameters of the hydrostratigraphic units, faults based on M´
adl-Sz˝
onyi et al. (2019).
Hydrostratigraphic
units
Age Mean
depth [m]
Sedimentary
environment
Pore uid Horizontal hydraulic
conductivity [m/s]
Porosity
[1]
Anisotropy
[1]
HsU5 Upper Miocene-
Pliocene-Quaternary
530 undifferentiated
siliciclastics
from brackish to
meteoric
10
-5
0.15 100
HsU4 Miocene 1480 undifferentiated siliciclastic
carbonate
marine 10
-6
0.08 100
HsU3 Oligocene 1990 deep water shale marine 10
-9
0.05 100
HsU2 Eocene-Oligocene 2430 siliciclastic carbonate with
karstication
marine &
meteoric
10
-6
0.2 100
HsU1 Paleozoic-Triassic-
Jurassic
3000 fractured, karstied marine
carbonate
marine 10
-5
0.2 10
Faults 10
-5
0.1 1
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
12
concentration are c =0 on the surface, knowing that the study area is a
recharge zone, while the bottom boundary is no-ux, representing the
impervious base rock. Along the side walls c =0 in HsU4 and HsU5,
where the freshwater outow dominates, and c =0.1 in HsU1HsU3
units, which have no direct connection to meteoric waters. Faults are
implicated in the model as domains with an average aperture of 10 m
and characterized with hydrogeological parameters listed in Table 3.
Additional modications were made compared to the model of M´
adl-
Sz˝
onyi et al. (2019). Mechanical dispersion was introduced due to the
coupled topography- and salinity-driven problem. Based on the former
synthetic model results, the value of the longitudinal dispersivity was
dened as
α
L
=100 m and the transverse dispersivity was
α
T
=10 m. In
order to handle the effective molecular diffusion in porous medium D
eff
,
the Bruggeman model was used (Wu et al., 2019), that is D
eff
=Φ
3/2
D
diff
.
Using additional thermal well hydraulic data, the horizontal hydraulic
conductivity of HsU1 and HsU2 was reduced by one order of magnitude.
Also, the values of anisotropic coefcients were increased from
ε
=10 to
100 in units HsU2HsU5, where the layered siliciclastic sediments
effectively retard the vertical groundwater ow (e.g. Galsa, 1997; Rman
and T´
oth, 2011; Hoque and Burgess, 2020).
The evolution of the salt concentration is presented for four different
time steps up to the maximum simulation time of 1 Myr (Fig. 14.be).
Only 1 kyr is enough for the meteoric water to displace the saltwater in
the upper part of HsU5. After 10 kyr, saltwater is swept out from HsU5
and the upper part of HsU4. Dense, salty plumes descend from the clayey
HsU3 aquitard into the karstied HsU1 at x =310 km, where the
aquitard is the thinnest (Fig. 14.c). The faults at x6 and 34 km, where
the salt content decrease appears rstly, facilitate the hydraulic
connection between the aquifers above and below HsU3. After 100 kyr
HsU4 and HsU5 are fully saturated by freshwater, while HsU1 is satu-
rated by brackish water owing through this domain from SW to NE
(Fig. 14.d). Small-scaled salty downwellings are noticeable in HsU2,
directly below the Oligocene aquitard (HsU3). After 1 Myr a consider-
able part of the original salt content has already been preserved in the
lower part of HsU3, however, the salinity of the layer has not been
uniform (Fig. 14.e). The concentration is higher in the zones, which are
thicker and are far from the hydraulically conductive faults. If the saline
water reaches a fault or the boundary of HsU3 by diffusion, it will be
effectively advected and displaced by fresh/brackish groundwater.
Fig. 14.f emphasizes that HsU4 and HsU5 are dominated by topography-
driven meteoric groundwater ow, while HsU1 and HsU2 are prevailed
by brackish through-ow. However, the upper and lower units are not
perfectly separated, because the two groundwater types communicate
through faults and the thin parts of the HsU3 aquitard. Finally, we note
that a hydraulic head maximum (120125 m) can be noticed in the
surroundings of the topographic divide, G¨
od¨
oll˝
o Hills (x =20 km)
(Fig. 14g). On the other hand, distinctive head minima appear in the
lower part of HsU3 correlating with the salt content. Within these zones,
even a 100 m head decrease is present compared to the freshwater zone.
5. Interpretation and discussion
5.1. Synthetic scenarios
Based on the synthetic simulations, it was established that a salt-
water zone evolves beneath the discharge zone as a result of a dynamic
Fig. 14. (a) The horizontal hydraulic conductivity (K
x
) of the hydrostratigraphic units and the location of faults (magenta) with the boundary conditions for the ow
(blue) and the salt concentration (green). Concentration distribution along the section at (b) t =1 kyr, (c) t =10 kyr, (d) t =100 kyr and (e) t =1 Myr. (f) The Darcy
ux magnitude and (g) the hydraulic head eld at t =1 Myr. (f) Normalized black arrows illustrate the ow direction. Vertical exaggeration is 1.5.
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
13
equilibrium between the negative buoyancy force of saltwater and the
topography-driven fresh groundwater ow. The size and shape of the
saline zone depend on the studied parameters. In general, if the effect of
salinity becomes dominant, the saltwater zone forms a dense, salty layer
in the deeper part of the basin. If the salinity-driven buoyancy prevails
against the regional ow, the salt concentration tends to the diffuse
distribution without advective solute content transport. This is the case,
when the relative density contrast (β 10%, Fig. 2.dg) and the
permeability anisotropy (
ε
100, Fig. 7.eg) are enhanced or the water
table amplitude is reduced (A 5 m, Fig. 11.ab). However, if the
topography-driven groundwater ow becomes dominant, the saltwater
zone shrinks, and it is pressed toward the stagnation point beneath the
discharge zone. The diffusion from below is not sufcient to maintain an
extended saline zone against the regional ow. This solution is obtained
from the simulations by decreasing the salinity (β <1%, Fig. 2.a) or by
increasing the water table amplitude (A >200 m, Fig. 11.g).
Monitoring parameters quantify the effects of studied parameters on
Darcy ux, the salt concentration and the hydraulic head etc. computed
for the entire basin and/or the salt- and freshwater zone. Besides, they
highlight the modications in the extent and morphology of the salt-
water zone (saline zone to layer at
ε
=100, Fig. 8.bd), or the transition
from advection- to diffusion-dominated mass transport system (at A =
10 m, Fig. 12.b and d). Darcy ux values averaged for the different zones
indicate that sluggish circulation occurs within the saltwater zone. The
ow in the saline zone is slower by approx. 12 orders of magnitude, but
q
s
and q
f
seem not to be independent of each other. The diffusion
mechanism is supposed to couple these two zones. Near the zone
boundaries, the diffusive ux is high due to the large concentration
gradient, and the salt content in the saline zone decreases resulting in
buoyancy force and upward motion being parallel to the regional
groundwater ow (Fig. 3). We note that this weak linkage can be
overtaken by small-scale heterogeneities within the saltwater zone,
which can lead to more complex inner convection (Fig. S3 in Supple-
mentary material).
In the studied parameter interval, the variation of permeability and
dispersivity do not cause relevant modications in the shape of the sa-
line water zone. It is worth mentioning that qualitatively, the perme-
ability decrease and the dispersivity increase affect the structure of the
saline zone in a similar way. First, the salt content increases by ho-
mogenization (Fig. 5.ad and Fig. 9.ad), then it decreases by thickening
the transition between the two zones (Fig. 5.eg and. Fig. 9.eh). The
latter process can be attributed to the increase of the relative effect of
diffusion (reducing the permeability retains the role of advection) and to
the amplication of transverse dispersion. Both phenomena strengthen
the coupling between the two zones, as it is supported by the fact that
the difference between q
f
and q
s
decreases from 2 to 1 orders of
magnitude (Fig. 6.e and Fig. 10.e). The effect of the permeability
decreasing exponentially with depth is very similar to the case of
permeability decrease, but its inuence weakens toward the surface.
Dimensionless numbers facilitate the simultaneous interpretation of
the effects of different model parameters on the behavior of the
groundwater ow system. Sherwood number (Sh) describes the ratio of
the surface concentration ux and the concentration ux due to purely
diffusion. It means that the larger the value of Sh, the stronger the role of
advection/convection in the solute transport. P´
eclet number (Pe) ex-
presses the convective transport (i.e. q
av
, the average Darcy ux in
simulations) compared to diffusion and dispersive solute transport,
Pe =qavd
Ddiff +
α
Lqav
.(7)
Both the diffusive and dispersive uxes are used in the denominator
since both hinder the convection (e.g. Simmons and Narayan, 1997).
Fig. 15.a shows the Sherwood number plotted against the P´
eclet
number for each synthetic simulation. Increasing salinity (β) retains the
ow slightly and reduces the surface solute ux by almost one order of
magnitude. Decreasing permeability in the homogeneous (k), the depth-
dependent (γ) and the anisotropic (
ε
) models slow down the ow and
lessens the surface concentration ux. When the effective permeability
decreases with depth (γ,
ε
), where the salt content is higher, a reduction
in Sh is more pronounced. Although the dispersivity (
α
L
) decreases Pe
(by enhancing the denominator of (7)), Sh increases owing to the
stronger coupling between the salt- and freshwater zones by higher
transverse dispersion. Higher water table amplitude (A) intensies the
advective transport and raises the surface concentration ux. While
lower water table amplitude decreases both Pe and Sh, and drives the
system toward the thin saline layer and then toward the diffusion state.
In the most numerical solutions, a dense, sluggish, saline dome forms
beneath the discharge zone, but decreasing the water table amplitude
(A) or increasing permeability anisotropy (
ε
) and salinity (β), the system
transforms into a thin layer or a diffusion regime (solid symbols).
Szij´
art´
o et al. (2019) dened a modied P´
eclet number to separate
the effect of thermal buoyancy from the total convective Darcy ux in
synthetic groundwater ow models driven by mixed thermal convec-
tion. In this paper, we reinterpreted the former denition, because (1)
advection competes with molecular diffusion and mechanical dispersion
(instead of thermal diffusion), and (2) the haline convection retards the
ow in the studied models. Thus, the modied P´
eclet number is
Pe*=(qadv qav)d
Ddiff +
α
Lqav
(8)
using the difference between the advective (q
adv
) and the total convec-
tive (q
av
) Darcy ux. Advective Darcy ux was calculated for each
simulation at β =0, that is the negative buoyancy force of salinity was
Fig. 15. (a) Sherwood number (Sh) plotted against the P´
eclet number (Pe) calculated for each synthetic simulation. The saline layer/diffusion regime is separated
from the saline dome regime by shading. (b) Non-dimensional numbers obtained from the P´
eclet number and the modied P´
eclet number (Pe*) calculated for each
synthetic simulation. The studied model parameters are denoted by different symbols in (b) and their extremums are labelled. The intersection of dashed lines shows
the base model (Table 1).
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
14
neglected.
The relation between the topography-driven and the buoyancy-
driven groundwater ow is elucidated in Fig. 15.b, where Pe +Pe* is
proportional to the advective Darcy ux and Pe/(Pe +Pe*) shows the
ratio of the total convective and advective Darcy ux (see Eq. (7) and
(8)). When the salinity does not inuence the ow (β =0), the advective
(q
adv
) and convective ux (q
av
) are identical, Pe/(Pe +Pe*) =1. The
increase in salinity retains the regional groundwater ow, e.g. the
convective Darcy ux at a value of β =0.2 is reduced by 43%. The
decrease of permeability in a homogeneous medium (k) and the increase
in mechanical dispersion (
α
L
) similarly affect the convective ux (see
Figs. 5 and 9). Decrease in k and increase in
α
L
considerably reduce the
advective transport compared to diffusion and dispersion (Pe +Pe*),
respectively, but these effects moderate the convective ow relative to
the base model by only some percent. Effective permeability decreasing
with depth (γ,
ε
) slows down the ow and lessens the role of salinity-
induced buoyancy force, as Pe/(Pe +Pe*) tends to 1. Increasing the
water table amplitude (A) strengthens the advective ux and weakens
the inuence of salinity on the ow. On the other hand, decreasing water
table amplitude accentuates the negative buoyancy force, as, for
instance, the convective ux at A =1 m is reduced to 23%.
Zhang et al. (2020b) noted that the increase of salinity retards the
nested groundwater ow intensity analogously to our ndings ,
although their hierarchical synthetic models have different boundary
conditions. Qualitatively, a similar brine zone with counter-rotating
inner convection forms beneath playas, as an interaction between the
freshwater inltration in the surrounding mountains and the redis-
solution of evaporites in playas both on regional (Duffy and Al-Hassan,
1988) and on a local scale (Hamann et al., 2015). However, in these
models, the saltwater recharge across the surface destabilizes the system
and stimulates the groundwater ow, while in our models, the salt
transport occurs from below, which stabilizes the system and hinders the
groundwater ow. At the same time, the dispersive ux through the
boundary of the salt- and freshwater zones appears in both simulations.
The transverse dispersion between the two zones enhances the coupling
between the ow systems by decoupling the ow from the salt transport,
as was concluded by Wen et al. (2018), who investigated the role of
dispersion on CO
2
storage in deep saline aquifers.
In an examined realistic hydrogeological situation (β =2.5%, k
x
=
10
-12
m
2
,
ε
=100,
α
L
=100 m and A =50 m) a thin saline layer (approx.
17,50035,000 mg/l) formed due to the high anisotropy coefcient and
it occupied the lower 7.6% of the basin (lower 380 m). The average TDS
content of the saline layer and the topography-driven upper zone was
26,100 mg/l and 1,240 mg/l, respectively. The brackish character of the
upper zone is explained by the large value of transverse dispersivity,
α
T
=10 m, which enhanced the salinity ux into the freshwater zone. The
maximum TDS content near the surface approximates only 630 mg/l in
the leftmost part of the discharge zone, however, this value reaches
4000 mg/l at a depth of 200 m, which can seriously deteriorate the
quality of the water in case of drinking and irrigation utilization. The
Darcy ux in the upper zone (q
f
6.810
-9
m/s) is higher than that in the
saline, sluggish layer by a factor of approx. 110. The average water
density in the saline water layer is 1018 kg/m
3
resulting in a 50 m hy-
draulic head decrease in the entire layer.
5.2. Realistic scenario
Numerical modelling of the studied hydrogeological section illus-
trates that the high salt content trapped in the clayey Oligocene aquitard
(HsU3) can survive over 1 Myr. Based on samples of hydrocarbon and
thermal wells, high TDS content in the Oligocene aquitard
(30,00050,000 mg/l) and brackish pore water observed in Triassic
carbonates (1,00010,000 mg/l) (M´
adl-Sz˝
onyi et al., 2019) agree with
the numerical model results (Fig. 14.e). The elevated TDS content ex-
plains the hydraulic head decit measured in HsU3, and it is in harmony
with the nding, that the thicker HsU3, the lower the hydraulic head.
However, higher salt content cannot elucidate the underpressure
(M´
adl-Sz˝
onyi et al., 2015), which might be the result of decompaction
and insufcient groundwater replenishment due to the uplifting of
G¨
od¨
oll˝
o Hills, which started 34 Myr ago (Ruszkiczay-Rüdiger et al.,
2007).
During the last 34 Myr, G¨
od¨
oll˝
o Hills became a topographic divide,
which facilitated the signicance of a topography-driven groundwater
ow system. Since the entire section is a recharge area, the upper
Neogene and Quaternary units (HsU4 and HsU5) are characterized by
low TDS. Based on data evaluation, the fractured and conned Triassic
carbonate is likely to be in connection with unconned carbonate on the
western side of Danube River (Fig. 13.c), so it is part of the Buda
Thermal Karst (BTK) system (M´
adl-Sz˝
onyi and T´
oth, 2015; Szij´
art´
o
et al., 2021). However, the connection and isolation are tectonically
inuenced, thus it can differ from place to place (M´
adln´
e Sz˝
onyi, 2020).
Brackish pore water in HsU1 and HsU2 can be the consequence of the
meteoric water inltration in the unconned part of BTK, the hydraulic
connection between the unconned and conned carbonates and the
weak saltwater leakage from HsU3. Based on the numerical simulation,
the linkage between the topography-driven fresh groundwater (HsU4
and HsU5) and the compositionally mixed brackish karst water (HsU1
and HsU2) is plausible through faults and the thinnest parts of the
Oligocene aquitard (HsU3).
These simulations, like each numerical or conceptual model, include
simplications and have limitations, which can modify the outcomes.
(1) One of them is that source/sink terms (precipitation and dissolution
of salts (e.g. V´
asquez et al., 2013; Hamann et al., 2015), evaporite for-
mation etc.) are not built into the simulations, however, the exit of the
salt content is allowed through the upper boundary of the synthetic
model by diffusion. However, it is supposed that these local-scale pro-
cesses cannot considerably change the received basin-scale ow pattern.
(2) Although equivalent porous media (EPM) approximation can be
accepted in the karstied carbonates on basin-scale (e.g. Long et al.,
1982; Scanlon et al., 2003; M´
adl-Sz˝
onyi and T´
oth, 2015), the effect of
relevant hydraulically conductive faults was taken into account (Berre
et al., 2019). (3) For the sake of simplicity, the effect of thermal buoy-
ancy was neglected in the simulations, which might be another impor-
tant factor in deep, karstied, basin-scale carbonate systems (e.g. Havril
et al., 2016; Szij´
art´
o et al., 2019, 2021). However, a recent study showed
that in the presence of topography-driven groundwater ow advective
heat transport exists, but free thermal convection might occur only at
higher Rayleigh numbers (Szij´
art´
o et al. 2019). For the purposes of this
paper, omittance of buoyancy due to temperature differences can be an
acceptable assumption. The fully coupled solutions (topography plus
salinity plus temperature) indicate a plausible perspective of further
simulations. Nevertheless, a preliminary simulation was completed
along the hydrogeological section crossing G¨
od¨
oll˝
o Hills in order to ash
the process of topothermohaline convection, where the ow pattern and
the salt concentration is controlled by both the water table topography
and the buoyancy force due to the temperature and the salinity variation
(Supplementary material). As a working hypothesis for further simula-
tions, it can be established that owing to the elevated geothermal
gradient characterizing the area, thermal convection formed in the deep,
partly conned karstied aquifers (HSU1 and HSU2), which intensied
the ow, strengthened the hydrodynamic connection between the upper
(HSU4 and HSU5) and lower aquifers, and facilitated the saltwater
drainage from the clayey Oligocene aquitard (HSU3) (Fig. S4 in Sup-
plementary material). (4) In both the synthetic simulations and the case
study, the upper boundary condition for the ow was a simple cosinu-
soid or a linear hydraulic head. However, local variations in the
topography and the water table can disturb the ow pattern and the
salinity eld, as well. A hierarchically nested groundwater system could
induce a more complex distribution of the salt concentration and the
temperature (Zhang et al., 2020b; An et al., 2015). In general, higher
salinity and surface heat ux are expected in the regional discharge area
rather than in local discharge/recharge areas. The lowest salinity and
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
15
heat ux are probable in regional recharge zones. Nonetheless, the
formation of time-dependent thermal buoyancy, which strongly de-
pends on the physical and geometrical model parameters (Szij´
art´
o et al.,
2019), is able to modulate or even overwrite the above-mentioned
simplied consideration resulting in a complex nested system where
the ow, the salinity and the heat vary in time and space.
6. Conclusions
Interaction of topography- and salinity-driven groundwater ow was
systematically investigated in synthetic two-dimensional, basin-scale
numerical models and then in a realistic scenario based on the
siliciclastic-carbonate basin in Hungary.
It was established in the synthetic analyses, that the effect of
increasing water density due to salinity and decreasing water table
amplitude on the saltwater distribution is similar. The rst enhances the
buoyancy-driven free convection, while the second reduces the
topography-driven forced convection. Thus, the numerical solutions
trend from a topography-dominated regional freshwater system towards
a diffusion-dominated layered system. The increase of permeability
anisotropy shows similar progress towards the formation of a dense,
saline layer in the deeper part of the basin. In the studied parameter
domain, both the decrease of permeability and the increase of dis-
persivity extinguish the relative role of advection by enhancing the
diffusion or dispersion, respectively. The two phenomena strengthen the
coupling between the fresh- and saltwater zone.
Based on the synthetic simulation set, the coupled topography- and
salinity-driven model was applied along a 2D hydrogeological section in
Hungary. After 100 kyr, the original marine pore water is outplaced
from both the upper siliciclastic aquifers by topography-driven meteoric
freshwater and the lower karstied carbonates by the brackish
groundwater through-ow. However, the marine-origin pore water in
the middle aquitard can survive over 1 Myr resulting in a signicant
head minimum. The connection between the upper and lower pervious
units are revealed by the presence of conductive faults, while the
connection between the clayey cover and conned carbonates is repre-
sented by dense, saline nger-like downwellings.
The current study improves the understanding of the interaction
between the topography-driven forced and the salinity-driven free
convection, especially in basin-scale groundwater ow systems. On the
one hand, the sensitivity analysis of the synthetic models highlights the
role of the most important model parameters inuencing the coupled
ow system. On the other hand, the phenomenon of mixed convection
was investigated and proved along a simplied 2D hydrogeological
section in Hungary to demonstrate the mixing of fresh, brackish and
saline water. Nevertheless, the combined topography- and salinity-
driven groundwater ow plays an important role in the Atacama salt
at aquifer, Chile (Tejeda et al., 2003; V´
asquez et al., 2013), in the
surroundings of playas in Pilot Valley, USA (Duffy and Al-Hassan, 1988;
Fan et al., 1997), below saline disposal basins in the MurrayDarling
Basin, Australia (Simmons and Narayan, 1997), in the salt accumulation
in the beach of Bon Secour National Wildlife Refuge, USA (Geng and
Boufadel, 2015) etc. Therefore, the results of the study can be used with
the necessary adaptations in further terrestrial hydrogeological envi-
ronments as well.
CRediT authorship contribution statement
Attila Galsa: Methodology, Validation, Investigation, Writing
original draft, Writing review & editing. ´
Ad´
am T´
oth: Data curation,
Supervision, Writing review & editing, Funding acquisition. M´
ark
Szij´
art´
o: Visualization, Supervision, Writing review & editing, Fund-
ing acquisition. Daniele Pedretti: Supervision, Funding acquisition.
Judit M´
adl-Sz˝
onyi: Data curation, Conceptualization, Writing review
& editing, Supervision, Funding acquisition.
Declaration of Competing Interest
The authors declare that they have no known competing nancial
interests or personal relationships that could have appeared to inuence
the work reported in this paper.
Acknowledgements
Authors are grateful to the anonymous reviewers for the instructive
comments and inspiring suggestions. This research is part of a project
that has received funding from the European Unions Horizon 2020
research and innovation program under grant agreement No 810980.
The project was supported by the ÚNKP-19-3 and ÚNKP-19-4 New Na-
tional Excellence Program of the Ministry for Innovation and Technol-
ogy, by the Hungarian Scientic Research Fund (K 129279), by the
J´
anos Bolyai Research Scholarship of the Hungarian Academy of Sci-
ences and by the ÚNKP-21-4 New National Excellence Program of the
Ministry for Innovation and Technology from the source of the National
Research, Development and Innovation Fund. The paper was prepared
with the professional support of the Doctoral Student Scholarship Pro-
gram of the Co-operative Doctoral Program of the Ministry of Innovation
and Technology nanced from the National Research, Development and
Innovation Fund.
Appendix A. Supplementary data
Supplementary data to this article can be found online at https://doi.
org/10.1016/j.jhydrol.2022.127695.
References
An, R., Jiang, X.-W., Wang, J.-Z., Wan, L., Wang, X.-S., Li, H., 2015. A theoretical analysis
of basin-scale groundwater temperature distribution. Hydrogeol. J. 23 (2), 397404.
https://doi.org/10.1007/s10040-014-1197-y.
Barlow, P.M., Reichard, E.G., 2010. Saltwater intrusion in coastal regions of North
America. Hydrogeol. J. 18, 247260. https://doi.org/10.1007/s10040-009-0514-3.
Berre, I., Doster, F., Keilegavlen, E., 2019. Flow in fractured porous media: A review of
conceptual models and discretization approaches. Transp. Porous Media 130,
215236. https://doi.org/10.1007/s11242-018-1171-6.
Cirpka, O.A., Chiogna, G., Rolle, M., Bellin, A., 2015. Transverse mixing in three-
dimensional nonstationary anisotropic heterogeneous porous media. Water Resour.
Res. 51, 241260. https://doi.org/10.1002/2014WR015331.
De Paola, F., Giugni, M., Topa, M.E., Bucchignani, E., 2014. Intensity-Duration-
Frequency (IDF) rainfall curves, for data series and climate projection in African
cities. SpringerPlus 3, 133. https://doi.org/10.1186/2193-1801-3-133.
Delgado, J.M.P.Q., 2012. Heat and Mass Transfer in Porous Media. Springer, Berlin,
Heidelberg, p. 266.
Deming, D., 2002. Introduction to Hydrogeology. McGraw Hill, New York, pp. 480,
ISBN-10: 0072326220.
Domenico, P.A., Palciauskas, V.V., 1973. Theoretical analysis of forced convective heat
transfer in regional ground-water ow. GSA Bull. 84 (12), 38033814. https://doi.
org/10.1130/0016-7606.
Dublyansky, Y.V., 1995. Speleogenetic history of the Hungarian hydrothermal karst.
Environ. Geol. 25, 2435. https://doi.org/10.1007/BF01061827.
Duffy, C.J., Al-Hassan, S., 1988. Groundwater circulation in a closed desert basin:
topographic scaling and climatic forcing. Water Resour. Res. 24 (10), 16751688.
https://doi.org/10.1029/WR024i010p01675.
Ehrenberg, S.N., Nadeau, P.H., 2005. Sandstone vs. carbonate petroleum reservoirs: A
global perspective on porosity-depth and porosity-permeability relationships. AAPG
Bull. 89 (4), 435445. https://doi.org/10.1306/11230404071.
Fan, Y., Duffy, C.J., Oliver Jr, D.S., 1997. Density-driven groundwater ow in closed
desert basins: eld investigations and numerical experiments. J. Hydrology 196,
139184. https://doi.org/10.1016/S0022-1694(96)03292-1.
Galsa, A., 1997. Modelling of groundwater ow along a section in the Great Hungarian
Plain using hydraulic heads measured in wells (in Hungarian with English abstract).
Magyar Geozika 38 (4), 245256. http://epa.oszk.hu/03400/03436/00151/pdf/
EPA03436_magyar_geozika_1997_04_245-256.pdf.
Gao, G., Zhan, H., Feng, S., Fu, B., Huang, G., 2012. A mobileimmobile model with an
asymptotic scale-dependent dispersion function. J. Hydrology 424425, 172183.
https://doi.org/10.1016/j.jhydrol.2011.12.041.
Garamhegyi, T., Sz´
ekely, F., Carrillo-Rivera, J.J., M´
adl-Sz˝
onyi, J., 2020. Revision of
archive recovery tests using analytical and numerical methods on thermal water
wells in sandstone and fractured carbonate aquifers in the vicinity of Budapest.
Hungary. Environmental Earth Science. 79, 129. https://doi.org/10.1007/s12665-
020-8835-6.
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
16
García-Gil, A., V´
azquez-Su˜
n´
e, E., Ayora, C., Tore, C., Henríquez, ´
A., Y´
a˜
nez, J., 2019.
Impacts of the transient skin effect during brine extraction operations in a crystalline
halite aquifer, J. Hydrology, 577, paper: 123912, 10.1016/j.jhydrol.2019.123912.
Gelhar, L.W., Welty, C., Rehfeldt, K.R., 1992. A critical review of data on eld-scale
dispersion in aquifers. Water Resour. Res. 28 (7), 19551974. https://doi.org/
10.1029/92WR00607.
Geng, X., Boufadel, M.C., 2015. Impacts of evaporation on subsurface ow and salt
accumulation in a tidally inuenced beach. Water Resour. Res. 51, 55475565.
https://doi.org/10.1002/2015WR016886.
Hamann, E., Post, V., Kohfahl, C., Prommer, H., Simmons, C.T., 2015. Numerical
investigation of coupled density-driven ow and hydrogeochemical processes below
playas. Water Resour. Res. 51, 93389352. https://doi.org/10.1002/
2015WR017833.
Havril, T., Molson, J.W., M´
adl-Sz˝
onyi, J., 2016. Evolution of uid ow and heat
distribution over geological time scales at the margin of unconned and conned
carbonate sequences - A numerical investigation based on the Buda Thermal Karst
analogue. Mar. Pet. Geol. 78, 738749. https://doi.org/10.1016/j.
marpetgeo.2016.10.001.
Havril, T., T´
oth, ´
A., Molson, J.W., Galsa, A., M´
adl-Sz˝
onyi, J., 2018. Impacts of predicted
climate change on groundwater ow systems: Can wetlands disappear due to
recharge reduction? J. Hydrol. 563, 11691180. https://doi.org/10.1016/j.
jhydrol.2017.09.020.
He, Z., Ding, Q., Wo, Y., Zhang, J., Fan, M., Yue, X., 2017. Experiment of carbonate
dissolution: implication for high quality carbonate reservoir formation in deep and
ultradeep basins. Geouids, 2017, paper: 8439259, 10.1155/2017/8439259.
Herbert, E.R., Boon, P., Burgin, A.J., Neubauer, S.C., Franklin, R.B., Ard´
on, M.,
Hopfensperger, K.N., Lamers, L.P.M., Gell, P., 2015. A global perspective on wetland
salinization: ecological consequences of a growing threat to freshwater wetlands.
Ecosphere 6 (10), paper: 206. https://doi.org/10.1890/ES14-00534.1.
Hoque, M.A., Burgess, W.G., 2020. Representing heterogeneity of uvio-deltaic aquifers
in models of groundwater ow and solute transport: A multi-model investigation in
the Bengal Basin. J. Hydrology, 590, paper: 125507, 10.1016/j.
jhydrol.2020.125507.
Horv´
ath, F., Bada, G., Windhoffer, G., Csontos, L., Dombr´
adi, E., D¨
ov´
enyi, P., Fodor, L.,
Grenerczy, G.y., Síkhegyi, F., Sza´
an, P., Sz´
ekely, B., Tim´
ar, G., T´
oth, L., T´
oth, T.,
2006. Atlas of the present-day geodynamics of the Pannonian basin: Euroconform
maps with explanatory text (in Hungarian with English abstract). Magyar Geozika
47 (4), 133137. http://epa.niif.hu/03400/03436/00186/pdf/EPA03436_magyar_
geozika_2006_04_133-137.pdf.
Hussain, M.S., Abd-Elhamid, H.F., Javadi, A.A., Sherif, M.M., 2019. Management of
Seawater Intrusion in Coastal Aquifers: A Review. Water, 11 (12), paper: 2467,
10.3390/w11122467.
Jalali, M., 2007. Salinization of groundwater in arid and semi-arid zones: an example
from Tajarak, western Iran. Environ. Geol. 52, 11331149. https://doi.org/10.1007/
s00254-006-0551-3.
Jiang, X-W., Wan, L., Wang, X-S., Ge, S., Liu, J., 2009. Effect of exponential decay in
hydraulic conductivity with depth on regional groundwater ow. Geophys. Res.
Lett., 36 (24), paper: L24402, 10.1029/2009GL041251.
Jiang, X-W., Wang, X-S., Wan, L., Ge, S., 2011. An analytical study on stagnation points
in nested ow systems in basins with depth-decaying hydraulic conductivity. Water
Resour. Res., 47 (1), paper: W01512, 10.1029/2010WR009346.
Kambesis, P.N., Coke, J.G., 2013. Overview of the Controls on Eogenetic Cave and Karst
Development in Quintana Roo, Mexico. In: Coastal Karst Landforms (ed. Lace, M.J.,
Mylroie, J.E.). Springer, Dordrecht, 347373, ISBN: 978-94-007-5016-6, 10.1007/
978-94-007-5016-6_16.
Karaki, N.A., Fiaschi, S., Paenen, K., Al-Awabdeh, M., Closson, D., 2019. Exposure of
tourism development to salt karst hazards along the Jordanian Dead Sea shore.
Hydrol. Earth Syst. Sci. 23, 21112127. https://doi.org/10.5194/hess-23-2111-
2019.
Kesler, S.E., Gruber, P.W., Medina, P.A:, Keoleian, G.A., Everson, M.P., Wallington, T.J.,
2012. Global lithium resources: Relative importance of pegmatite, brine and other
deposits. Ore Geol. Rev., 48, 5569, 10.1016/j.oregeorev.2012.05.006.
Konikow, L.F., Akhavan, M., Langevin, C.D., Michael, H.A., Sawyer, A.H., 2013.
Seawater circulation in sediments driven by interactions between seabed topography
and uid density. Water Resour. Res. 49, 13861399. https://doi.org/10.1002/
wrcr.20121.
Kreyns, P., Geng, X., Michael, H.A., 2020. The inuence of connected heterogeneity on
groundwater ow and salinity distributions in coastal volcanic aquifers. J.
Hydrology, 586, paper: 124863, 10.1016/j.jhydrol.2020.124863.
Liu, R., Li, B., Jiang, Y., Huang, N., 2016. Review: Mathematical expressions for
estimating equivalent permeability of rock fracture networks. Hydrogeol. J. 24,
16231649. https://doi.org/10.1007/s10040-016-1441-8.
Long, J.C.S., Remer, J.S., Wilson, C.R., Witherspoon, P.A., 1982. Porous media
equivalents for networks of discontinuous fractures. Water Resour. Res. 18 (3),
645658. https://doi.org/10.1029/WR018i003p00645.
M´
adln´
e Sz˝
onyi, J., 2020. Pattern of groundwater ow at the boundary of unconned and
conned carbonate systems on the example of Buda Thermal Karst and its
surroundings (in Hungarian). DSc Thesis, pp. 150, https://scholar.google.com/
scholar?q=Pattern%20of%20Groundwater%20Flow%20at%20the%20Boundary%
20of%20Unconned%20and%20Conned%20Carbonate%20Systems%20on%
20the%20Example%20of%20Buda%20Thermal%20Karst%20and%20its%
20Surroundings.
M´
adln´
e Sz˝
onyi, J., Czauner, B., Er˝
oss, A., Simon, S., 2013. Karbon´
atos ´
es Csatlakoz´
o
Üled´
ekes Medenceterületek Fluidumdinamikai ¨
Osszefügg´
eseinek Vizsg´
alata a
Sz´
enhidrog´
en Kutat´
as Hat´
ekonys´
ag´
anak Javít´
asa ´
Erdek´
eben a Paleog´
en-medenc´
eben
Z´
ar´
ojelent´
es. Investigation of uid-dynamic connection of carbonate and adjoining
basin in hydrocarbon exploration purposes Final Report. Hungarian Oil Company,
Budapest.
M´
adln´
e Sz˝
onyi, J., Er˝
oss, A., Havril, T., Poros, Zs., Gy˝
ori, O., T´
oth, ´
A., Csoma, A., Ronchi,
P., Mindszenty, A., 2018. Fluids, ow systems and their minerological imprints in
the Buda Thermal Karst (in Hungarian with English abstract). F¨
oldtani K¨
ozl¨
ony, 148
(2), 7596, 10.23928/foldt.kozl.2018.148.1.75.
M´
adl-Sz˝
onyi, J., Czauner, B., Iv´
an, V., T´
oth, ´
A., Simon, S.z., Er˝
oss, A., Bodor, P.,
Havril, T., Boncz, L., S˝
oreg, V., 2019. Conned carbonates? regional-scale
hydraulic interaction or isolation? Mar. Pet. Geol. 107, 591612. https://doi.org/
10.1016/j.marpetgeo.2017.06.006.
M´
adl-Sz˝
onyi, J., Pulay, E., T´
oth, ´
A., Bodor, P., 2015. Regional underpressure: a factor of
uncertainty in the geothermal exploration of deep carbonates, G¨
od¨
oll˝
o Region,
Hungary. Environ. Earth Sci. 74 (12), 75237538. https://doi.org/10.1007/s12665-
015-4608-z.
M´
adl-Sz˝
onyi, J., T´
oth, ´
A., 2015. Basin-scale conceptual groundwater ow model for an
unconned and conned thick carbonate region. Hydrogeol. J. 23, 13591380.
https://doi.org/10.1007/s10040-015-1274-x.
Marazuela, M.A., V´
azquez-Su˜
n´
e, E., Custodio, E., Palma, T., García-Gil, A., Ayora, C.,
2018. 3D mapping, hydrodynamics and modelling of the freshwater-brine mixing
zone in salt ats similar to the Salar de Atacama (Chile). J. Hydrology 561, 223235.
https://doi.org/10.1016/j.jhydrol.2018.04.010.
McFarlane, D.J., George, R.J., Barrett-Lennard, E.G., Gilfedder, M., 2017. Salinity in
dryland agricultural systems: Challenges and opportunities. In: Innovations in
Dryland Agriculture (ed. Farooq, M., Siddique, K.H.M.). Springer International
Publishing, 521547, ISBN: 978-3-319-47928-6, 10.1007/978-3-319-47928-6_19.
Nield, D.A., 1968. Onset of thermohaline convection in a porous medium. Water Resour.
Res. 4 (3), 553560. https://doi.org/10.1029/WR004i003p00553.
Nield, D.A., Bejan, A., 2017. Convection in Porous Media, 5th ed. Springer Int, p. 988.
Oz, I., Eyal, S., Yoseph, Y., Ittai, G., Elad, L., Haim, G., 2016. Salt dissolution and sinkhole
formation: Results of laboratory experiments. J. Geophys. Res. Earth Surf. 121,
17461762. https://doi.org/10.1002/2016JF003902.
Payen, S., Basset-Mens, C., ˜
nez, M., Follain, S., Grünberger, O., Marlet, S., Perret, S.,
Roux, P., 2016. Salinisation impacts in life cycle assessment: a review of challenges
and options towards their consistent integration. Int. J. Life Cycle Assess. 21,
577594. https://doi.org/10.1007/s11367-016-1040-x.
Pedretti, D., Russian, A., Sanchez-Vila, X., Dentz, M., 2016. Scale dependence of the
hydraulic properties of a fractured aquifer estimated using transfer functions. Water
Resour. Res. 52, 50085024. https://doi.org/10.1002/2016WR018660.
Pedretti, D., Fern`
andez-Garcia, D., Sanchez-Vila, X., Bolster, D., Benson, D.A., 2014.
Apparent directional mass-transfer capacity coefcients in three-dimensional
anisotropic heterogeneous aquifers under radial convergent transport. Water Resour.
Res. 50, 12051224. https://doi.org/10.1002/2013WR014578.
Rman, N., T´
oth, Gy., 2011. T-JAM Vízf¨
oldtani koncepcion´
alis modell (in Hungarian), pp.
25.
Rodríguez-Rodríguez, M., Benavente, J., 2008. Denition of wetland typology for hydro-
morphological elements within the WFD. A case study from southern Spain. Water
Resour. Manage. 22, 797821. https://doi.org/10.1007/s11269-007-9193-9.
Ruszkiczay-Rüdiger, Z., Fodor, L.I., Horv´
ath, E., 2007. Neotectonics and quaternary
landscape evolution of the G¨
od¨
oll˝
o Hills, Central Pannonian Basin. Hungary. Global .
Planet. Change 58 (1), 181196. https://doi.org/10.1016/j.gloplacha.2007.02.010.
Sanchez-Vila, X., Guadagnini, A., Carrera, J., 2006. Representative hydraulic
conductivities in saturated groundwater ow. Rev. Geophys., 44, paper: RG3002,
10.1029/2005RG000169.
Scanlon, B.R., Mace, R.E., Barrett, M.E., Smith, B., 2003. Can we simulate regional
groundwater ow in a karst system using equivalent porous media models? Case
study, Barton Springs Edwards aquifer, USA. J. Hydrology 276, 137158. https://
doi.org/10.1016/S0022-1694(03)00064-7.
Schuler, M.S., Ca˜
nedo-Argüelles, M., Hintz, W.D., Dyack, B., Birk, S., Relyea, R.A., 2018.
Regulations are needed to protect freshwater ecosystems from salinization.
Philosophical Transactions of the Royal Society B, Biological Sciences, 374, paper:
20180019, 10.1098/rstb.2018.0019.
Schulze-Makuch, D., 2005. Longitudinal dispersivity data and implications for scaling
behavior. Groundwater 43 (3), 443456. https://doi.org/10.1111/j.1745-
6584.2005.0051.x.
Simmons, C.T., Narayan, K.A., 1997. Mixed convection processes below a saline disposal
basin. J. Hydrology 194, 263285. https://doi.org/10.1016/S0022-1694(96)03204-
0.
Singh, A., 2020. Salinization and drainage problems of agricultural land. Irrig. Drain. 69
(4), 844853. https://doi.org/10.1002/ird.2477.
Smith, M.E., Wynn, J.G., Scharping, R.J., Moore, E.W., Garey, J.R., Onac, B.P., 2021.
Source of saline groundwater on tidally inuenced blue holes on San Salvador
Island, Bahamas. Hydrogeol. J. 29, 429441. https://doi.org/10.1007/s10040-020-
02266-z.
Sun, Y., Wendi, D., Kim, D.E., Liong, S.-Y., 2019. Deriving intensitydurationfrequency
(IDF) curves using downscaled in situ rainfall assimilated with remote sensing data.
Geosci. Lett. 6, 17. https://doi.org/10.1186/s40562-019-0147-x.
Szij´
art´
o, M., Galsa, A., T´
oth, ´
A., M´
adl-Sz˝
onyi, J., 2019. Numerical investigation of the
combined effect of forced and free thermal convection in synthetic groundwater
basins. J. Hydrol. 572, 364379. https://doi.org/10.1016/j.jhydrol.2019.03.003.
Szij´
art´
o, M., Galsa, A., T´
oth, ´
A., M´
adl-Sz˝
onyi, J., 2021. Numerical analysis of the
potential for mixed thermal convection in the Buda Thermal Karst, Hungary. J.
Hydrology: Regional Studies, 34, paper: 100783, 10.1016/j.ejrh.2021.100783.
Szocs, T., Rman, N., Süveges, M., Palcsu, L., T´
oth, G.y., Lapanje, A., 2013. The
application of isotope and chemical analyses in managing transboundary
groundwater resources. Appl. Geochem. 32, 95107. https://doi.org/10.1016/j.
apgeochem.2012.10.006.
A. Galsa et al.
Journal of Hydrology 609 (2022) 127695
17
Tejeda, I., Cienfuegos, R., Mu˜
noz, J.F., Dur´
an, M., 2003. Numerical modeling of saline
intrusion in Salar de Atacama. J. Hydrol. Eng. 8 (1), 2534. https://doi.org/
10.1061/(ASCE)1084-0699(2003)8:1(25).
T´
oth, J., 2009. Gravitational Systems of Groundwater Flow: Theory, Evaluation,
Utilization. Cambridge University Press, United Kingdom, p. 297.
Van Dam, R.L., Simmons, C.T., Hyndman, D.W., Wood, W.W., 2009. Natural free
convection in porous media: First eld documentation in groundwater. Geophys.
Res. Lett., 36, paper: LL11403, 10.1029/2008GL036906.
Vanderborght, J., Vereecken, H., 2007. Review of dispersivities for transport modeling in
soils. Vadose Zone J. 6 (1), 2952. https://doi.org/10.2136/vzj2006.0096.
Vars´
anyi, I., ´
O.Kov´
acs, L., 2009. Origin, chemical and isotopic evolution of formation
water in geopressured zones in the Pannonian Basin, Hungary. Chem. Geol. 264
(14), 187196. https://doi.org/10.1016/j.chemgeo.2009.03.006.
V´
asquez, C., Ortiz, C., Su´
arez, F., Mu˜
noz, J.F., 2013. Modeling ow and reactive
transport to explain mineral zoning in the Atacama salt at aquifer, Chile. J. Hydrol.
490, 114125. https://doi.org/10.1016/j.jhydrol.2013.03.028.
Wang, J.-Z., Jiang, X.-W., Wan, L., W¨
orman, A., Wang, H., Wang, X.-S., Li, H., 2015. An
analytical study on artesian ow conditions in unconned-aquifer drainage basins.
Water Resour. Res. 51 (10), 86588667. https://doi.org/10.1002/2015WR017104.
Weatherill, D., Simmons, C.T., Voss, C.I., Robinson, N.I., 2004. Testing density-
dependent groundwater models: two-dimensional steady state unstable convection
in innite, nite and inclined porous layers. Adv. Water Resour. 27, 547562.
https://doi.org/10.1016/j.advwatres.2004.01.003.
Wen, B., Chang, K. W., Hesse, M.A., 2018. Rayleigh-Darcy convection with
hydrodynamic dispersion. Phys. Rev. Fluids, 3, paper: 123801, 10.1103/
PhysRevFluids.3.123801.
Werner, A.D., Bakker, M., Post, V.E.A., Vandenbohede, A., Lu, C., Ataie-Ashtiani, B.,
Simmons, C.T., Barry, D.A., 2013. Seawater intrusion processes, investigation and
management: Recent advances and future challenges. Adv. Water Resour. 51, 326.
https://doi.org/10.1016/j.advwatres.2012.03.004.
Wu, H., Fang, W.-Z., Kang, Q., Tao, W.-Q., Qiao, R., 2019. Predicting effective diffusivity
of porous media from images by deep learning. Sci. Rep. 9, 20387. https://doi.org/
10.1038/s41598-019-56309-x.
Zhang, J., Zhou, J., Fu, L., Li, H., Lou, D., 2020a. Karstication of Ordovician carbonate
reservoirs in Huanghua depression and its control factors. Carbonates Evaporites 35,
paper: 42. https://doi.org/10.1007/s13146-020-00572-x.
Zhang, X., Jiao, J.J., Li, H., Luo, X., Kuang, X., 2020b. Effects of downward intrusion of
saline water on nested groundwater ow systems. Water Resour. Res., 56, paper:
e2020WR028377, 10.1029/2020WR028377.
Zhang, Z.-Y., Jiang, X.-W., Wang, X.-S., Wan, L., Wang, J.-Z., 2018. A numerical study on
the occurrence of owing wells in the discharge area of basins due to the upward
hydraulic gradient induced wellbore ow. Hydrol. Process. 32 (11), 16821694.
https://doi.org/10.1002/hyp.11623.
Zimmermann, W.B.J., 2006. Multiphysics Modeling With Finite Element Methods. World
Scientic Publishing Company, Singapore, p. 422.
A. Galsa et al.
... The surface of the finite element mesh is approximately consisted with the groundwater table generally following the topography [29] with the head boundary condition. In this model, the water depression cone resulting from climate change and epeirogenetic geology were not taken into consideration. ...
... The hydraulic water head h can be extended with the Girinski potential for groundwater flow in stratified aquifers [28], which allows for the application of Equations (1) and (2) to the rocks with layered heterogeneity. The surface of the finite element mesh is approximately consisted with the groundwater table generally following the topography [29] with the head boundary condition. In this model, the water depression cone resulting from climate change and epeirogenetic geol-ogy were not taken into consideration. ...
Article
Full-text available
Mine water injection into deep formations is one of the effective approaches for reducing the drainage from coal mines in the arid and semi-arid region of the Ordos basin, China. Many coal mines are attempting to execute the related projects. Under the influence of groundwater protection, the understanding of regional groundwater flow is becoming highly important to the mine water monitoring, whereas quite few academic research teams focus on the deep groundwater flow pattern by mine water injection. This paper reveals the spatial distribution of Liujiagou Formation that is in positive correlation with the terrain, and its local thickness is influenced by the dominant W-E and NE-SW directions of geological structures. Only a part of sandstone rocks consists of aquifers, the rest 61.9% of relatively dry rock provide the enhanced storage space and partial mudstone aquicludes decrease the possibility of the vertical leakage for mine water. The dynamic storage capacity is evaluated at 2.36 Mm3 per 1 km2 and over 25.10 billion m3 in this study area. Two hydrogeologic cross-sections of basin-scale identify the W-E and N-S regional groundwater flow directions, with the lower Yellow River catchment becoming the discharged region. The hierarchically and steadily nested flow systems containing coal mining claims are influenced by coal mining activity. The groundwater depression cone in a shallow coal measure aquifer is caused by mine water drainage whereas the groundwater mound in Liujiagou Formation is generated by mine water injection activity. The numerical simulation revealed that the groundwater head rebound is slightly decreased and will not recover to its initial baseline within 500 years due to its low porosity and permeability. This study elucidates the deep groundwater flow patterns induced by mine water injection and provides a practical methodology for the management and pollution monitoring of mine water injection activity.
... A robust geologically based redefinition of hydrostratigraphy, allows a better understanding of the regional groundwater flow in the Pannonian basin and will support further hydrogeological studies e.g. to reveal groundwater flow systems for drinking water and geothermal exploration (Horváth et al., 2015;Nádor et al., 2019;Szijártó et al., 2019Szijártó et al., , 2021. It helps to explain the origin and migration pathways of saline waters (Galsa et al., 2022), to interpret biogenic gas accumulations (Czauner and Mádl-Szőnyi, 2013;Molnár et al., 2015) and allow the elaboration of more reliable numerical models with a better approximation of the subsurface fluid flow. ...
Article
Full-text available
Heterogeneity and anisotropy of lithological units influences hydraulic conductivity on several scales. The purpose of this study is to develop a novel method supporting hydrostratigraphic classification by considering horizontal and vertical variations of sand content in relation with the paleo-depositional environment. The workflow was applied for the uppermost ca. 1800 m thick fluvio-deltaic deposits (Great Plain Aquifer) of late Neogene to Quaternary age basin-fill succession in the eastern part of the Pannonian basin in Hungary. Five combined 3D seismic volumes, seven master horizons, and 30 well logs were analyzed. First, RMS amplitude maps were extracted to interpret the seismic geomorphological features and depositional architectures. Their associated lithology was inferred from wireline logs (GR and SP) by calculating the shale volume and the net-to-gross sand ratio for 30 m thick intervals. These were used to calibrate the seismic facies as a proxy for the horizontal distribution of sand versus shale. This method allowed the identification of sand bodies, i.e.: deltaic lobes, complex channel belts, simple fluvial channels behaving as aquifers, and the dominantly muddy delta plain to flood plain suits as aquitard. The vertical pattern of sand distribution was also evaluated. Three major stratiform and some corridor-like minor hydrostratigraphic units were defined instead of the former regional (basin) scale aquifer unit. 1) Laterally extended interval of stacked deltaic lobes of high sand ratios and high rate of connectedness, at the bottom of the studied interval. 2) General presence of extended muddy floodplains with anastomosing river systems, characterized by 100–200 m wide channels, low sand ratio, and limited connectedness: 3) widespread meandering and/or braided river systems with high sand ratios and high connectedness in the Quaternary. Within unit 2) spatially and temporally variable appearance of major 500–3600 m wide meandering channel belts produce locally high sand-ratio corridors in the NE-SW direction. A workflow adapted from the oil-and-gas industry was successfully applied to distinguish units of varying sand content and hydraulic conductivity. This approach can be used on basin to local scale to build a spatially complex facies-based model of hydrostratigraphy. Thus, a robust heterogeneous geological model serves as a base for investigating fluid flow on the required scale.
... Buoyant flow is a phenomenon that occurs when fluid motion is induced by differences in densities of displacing and displaced fluids [23,24]. Research on buoyancy-driven flow in porous media has numerous applications, including in contaminants transport, saltwater intrusion, tracer dispersion, and accumulation of hydrocarbons in traps [25][26][27][28]. Buoyant flow also plays a crucial role in gas storage processes, especially in CO 2 storage in geological formations [29,30]. ...
Article
Understanding buoyancy-driven flow is essential for ensuring the safety, reliability, and efficiency of underground hydrogen storage in saline aquifers. In this study, we develop a three-dimensional (3D) numerical model of an aquifer for hydrogen storage, with an isothermal system at 323.15 K, to investigate the behavior of hydrogen buoyant flow. We perform a series of sensitivity simulations to examine the influence of storage formation heterogeneity, including Dykstra-Parsons coefficient, autocorrelation length, and permeability anisot-ropy, as well as fluid-rock interaction parameters, such as capillary pressure and relative permeability, on storage efficiency. We evaluate the efficiency of geologic storage containment by quantifying a defined metric "escape ratio," which is the mass fraction of hydrogen in a presumed escape region over the total mass of hydrogen. Our results indicate that hydrogen escape ratios and migration characteristics are primarily influenced by formation heterogeneities. The escape ratio decreases as Dykstra-Parsons coefficient increases and permeability anisotropy ratio decreases. The specific hydrogen escape path is controlled by spatial structures (i.e., autocorrelation length). Inclusion of capillary heterogeneity results in a decrease in the escape ratio, and the escape region exhibits a smoother appearance in the absence of capillary heterogeneity trapping as compared to when it is modeled. As for fluid-rock interactions, the hydrogen escape ratio is highly susceptible to the exponent of Brooks-Corey relative permeability function. Nevertheless, the magnitude of the escape ratio variation is predominantly determined by formation heterogeneity.
... A robust geologically based redefinition of hydrostratigraphy, allows a better understanding of the regional groundwater flow in the Pannonian basin and will support further hydrogeological studies e.g. to reveal groundwater flow systems for drinking water and geothermal exploration (Horváth et al., 2015;Nádor et al., 2019;Szijártó et al., 2019Szijártó et al., , 2021. It helps to explain the origin and migration pathways of saline waters (Galsa et al., 2022), to interpret biogenic gas accumulations (Czauner and Mádl-Szőnyi, 2013;Molnár et al., 2015) and allow the elaboration of more reliable numerical models with a better approximation of the subsurface fluid flow. ...
Conference Paper
Hydrostratigraphic evaluation is an asset for the overall understanding of hydrogeological context, the buildup of reliable numerical models and making proper decisions related to regional groundwater issues. The fluvio-deltaic environment is one of the most complex to evaluate, due to the intense heterogeneity of sediments, induced by the rapid channels’ migration. This study aims to introduce a novel approach for hydrostratigraphic decomposition in fluvio-deltaic sequence in eastern Pannonian Basin and a comparison between geostatistical and geophysical based methods. 3D seismic data was interpreted; shale volume logs were generated from GR and SP wireline logs, net-to-gross sand thickness of 30m thick intervals was used to highlight the vertical distribution and to calibrate the seismic attribute maps. Root Mean Square maps were reproduced from key horizons and identified the spatio-temporal variation of sand content. Later, a cluster analysis, as well as geophysical evaluation of porosity, permeability and hydraulic conductivity by neutron, density and sonic logs were conducted. Both results identified the following 4 major hydrostratigraphical classification. 1) an aquifer dominated by delta front sand lobes with 67% sand; 2) an aquitard formed by delta plain sediments with a lower sand content 43% sand; 3) an aquitard formed by a mud-prone fluvial environment with 35% sand content; 4) sand-dominated fluvial succession forming a good quality aquifer with 60% sand. These approaches are applicable to any study region and successfully delineated the regional hydrostratigraphy, however, geophysical analysis showed detailed results by providing precise hydro-physical parameters compared to statistical analysis, which are depth-dependent (cluster estimations).
... By creating an integration model, the possibility of sustainable management of surface and groundwater is considered [Rajanayaka et al., 2021]. The interaction of forced convection due to relief and free convection due to salinity was studied in a real hydrological section in Hungary [Galsa et al., 2022]. The modeling of groundwater depletion was considered concerning the territory of Great Britain [Marchant and Bloomfield, 2018]. ...
Article
Full-text available
The West Kazakhstan region of the Republic of Kazakhstan occupies an area equal to 151,339 km2. In the land structure, 69.7% of the area is occupied by agricultural land. The region has great prospects for the development of the livestock industry. However, uneven territorial availability of water resources is a limiting factor in increasing the amount of livestock in the region. The purpose of the study is to monitor underground water sources in the West Kazakhstan region of the Republic of Kazakhstan to assess the zonality of their placement. The boundaries of natural and climatic zones on the territory of the region were laid over the publicly available cartographic materials on the hydrological data of the distribution of groundwater. The water source monitoring was carried out by examining their actual condition in specific geographical locations, including using remote sensing methods, with a further determination of quantitative and qualitative parameters. The paper considers the state and problems of water supply at the pastures in the natural and climatic zones of the West Kazakhstan region. The region is characterized by the use of groundwater in the water supply of pasture lands. Underground springs have a certain zonality in their location, manifest themselves at different depths corresponding to different geological horizons, and differ in a wide variation of water mineralization. In the dry steppe zone, it is recommended to use the aquiferous mid-upper quaternary alluvial, aquiferous upper Pliocene Akchagyl, and aquiferous upper cretaceous Maastricht horizons. The water sources used have depths of up to 120 meters, and the mineralization varies from 0.2 to 9.1 g/dm³. In the semi-desert zone, the upper-quaternary aquiferous marine Khvalynsky and the lower-middle-quaternary aquiferous marine Baku-Khazar horizons are recommended. The water sources used have depths of up to 90 meters, and the mineralization varies from 0.2 to 11.8 g/dm³. The semi-desert zone is characterized by the use of springs with depths up to 80 meters. The mineralization of water in the permeable modern Aeolian horizon is more often low (0.11–0.9 g/dm³) and rarely brackish (1.1–9.36 g/dm³)
Article
The problem of limestone karst water inrush from coal seam floors is becoming more and more prominent, and gradually becoming the focus of water control work in coal mines. Taking the floor of 10th coal seam in the third mining area of Zhuxianzhuang Coal Mine as the research object, seven indicators were identified as the main controlling factors in three aspects, such as aquifer, water barrier, geological structure. Subjective and objective weights were determined by the analytic hierarchy process and the entropy weighting method, and vulnerability zoning evaluation was carried out based on the geographic information system using the vulnerability index method. The study area was classified into five zones based on the graded thresholds calculated by the statistical analysis. The study shows that the area of Taiyuan Formation limestone karst water outburst gradually transitions from the safe zone to the vulnerable zone from the north to the east. This study has a certain guiding effect on the water inrush of Taiyuan Formation limestone karst from the floor of the 10th coal seam in the study area.
Article
In this paper, a numerical model is presented to simulate the density-driven flow in heterogeneous porous media with micro- and macro-fractures. The micro-fractures (like fissures) with relatively small fracture lengths compared to the reservoir are modeled using an equivalent continuum model. The macro-fractures (like faults) with large fracture lengths are modeled using the extended finite element method (XFEM). Governing equations are based on the mass conservation equation for both the matrix and fracture domains together with the proper constitutive laws for the fluid velocity, dispersive velocity of solute, and equation of state for density of fluid phase. Several parameter studies are performed to investigate the effects of fracture characteristics on the solute spread in the medium. It is demonstrated that in the absence of macro-fractures, the influence of micro-fractures’ aperture on the solute spread is fully substantial and predictable. However, in the presence of macro-fractures, there is a region where the macro-fracture performs like a vacuum diminishing the influence of nearby micro-fractures. Finally, a real hydrological problem of the Gödöllő Hills reservoir located in Hungary with five different layers and eight faults, which is full of saline water is numerically modeled to investigate the effect of pressure gradient, anisotropy, and density difference on the solute transport. It is illustrated that in the top layers of the Gödöllő Hills reservoir, the gravitational force and vertical permeability are mostly accountable for the solute flow across the faults, whereas in the bottom layers, the horizontal pressure gradient is primarily responsible for the solute transport.
Article
The distribution of groundwater temperature and flow can be used to describe the hydrogeological process in a geothermal system, which is of great significance to the exploitation of geothermal resources. The traditional gravity-driven groundwater system theory takes less consideration of thermal convection, resulting in a deviation in the investigation of geothermal systems. The rise of water table and the acceleration of the flow of geothermal water in the discharge section suggest the existence of geothermal buoyancy. The changes in temperature, salinity and viscosity comprise the physical basis for the formation of geothermal buoyancy. Geothermal buoyancy due to free heat convection in a fault-controlled discharge section is discussed in the Xinzhou geothermal field, South China. The geothermal buoyancy is the additional pressure head created by the increase in temperature, increase in salinity and decrease in viscosity. At the convection point, geothermal buoyancy appears and forms a maximum of + 417.6 m. Geothermal buoyancy gradually decreases in the discharge section due to temperature domination. The salinity effect and viscosity effect on geothermal buoyancy are minor and negligible, respectively. Comparing the vertical velocity of geothermal water in the discharge section (32.47 × 10−3 m/d) and the average vertical circulation velocity (3.15 × 10−3 m/d), the geothermal buoyancy has an obvious acceleration effect on groundwater flow in the discharge section. The geothermal buoyancy at typical points provides the framework and control points for geothermal water flow in the discharge section of a geothermal system.
Article
This review paper briefly summarizes the research results of the majority (∼70%) women team of the Hydrogeology Research Group of Eötvös Loránd University, Hungary, led by Judit Mádl-Szőnyi. The group had originally focused on basin-scale groundwater flow systems and the related processes and phenomena but extended its research activity to other geofluids in answer to global challenges such as the water crisis, climate change, and energy transition. However, the core concept of these studies remained the basin-scale system approach of groundwater flow, as these flow systems interact with the rock framework and all other geofluids resulting in a systematic distribution of the related environmental and geological processes and phenomena. The presented methodological developments and mostly general results have been and can be utilized in the future in any sedimentary basins. These cover the following fields of hydrogeology and geofluid research: carbonate and karst hydrogeology, asymmetric basin and flow pattern, geothermal and petroleum hydrogeology, radioactivity of groundwater, groundwater and surface water interaction, groundwater-dependent ecosystems, effects of climate change on groundwater flow systems, managed aquifer recharge.
Article
Full-text available
Study region Buda Thermal Karst system, Hungary. Study focus The pilot area has high geothermal potential characterized by prominent thermal anomalies, such as thermal springs and spas which tap the Triassic carbonate aquifers. Therefore, numerical simulations were carried out to examine the temperature field and flow pattern considering three successive heat transport mechanisms: thermal conduction, forced and mixed thermal convection in order to highlight the role of different driving forces of groundwater flow in the Buda Thermal Karst. New hydrological insights for the region Compared to thermal conduction, topography-driven heat advection increases the surface heat flux. The superimposed effect of free thermal convection facilitates the formation of time-dependent mixed thermal convection from the deep carbonate layers. The Nusselt number varied between Nu = 1.56 and 5.25, while the recharge rate (R) ranged from R = 178 mm/yr to 250 mm/yr. Radiogenic heat production and hydraulically conductive faults have only a minor influence on the basin-scale temperature field and flow pattern. Boundary conditions prescribed on the temperature and pressure can considerably affect the numerical results. In each scenario, independently of the model parameters, time-dependent mixed thermal convection evolved both in the deep and the confined parts of the karstified carbonates of the Buda Thermal Karst system.
Article
Full-text available
Stable hydrogen and oxygen isotope signatures of waters within Church and Inkwell blue holes are measured on San Salvador Island (Bahamas) to identify the origin of their fresh and saline waters. Stable isotope data, paired with a suite of physicochemical water parameters measured throughout the blue holes, as a function of both time and depth, provide a detailed understanding of the tidally influenced groundwater interactions on the island. Blue holes are prominent karst features in carbonate environments which serve as windows into subterranean hydrologic processes. Carbonate island hydrology is often complicated by complex interactions between the marine and meteoric water systems, as tidal pumping and water mixing result in diagenetic alteration of the bedrock, that in turn influence dissolution rates and preferential flow paths. Although the blue holes on the island are physically influenced by tidal forcing, the stable isotope data indicate that both their fresh and saline waters are of a meteoric origin rather than seawater, where the meteoric water is likely becoming saline through enrichment by aerosol-derived sea salts. Additionally, the physical profiles of each blue hole indicate differences in mixing processes driven by wind and tidal forcing, where stronger mixing can result in a disruption of the freshwater lens. The implications of this study are important for assessing mixing corrosion processes and dissolution effects, but more research and longer data sets are needed to show whether these results are applicable to other coastal carbonate environments.
Article
Full-text available
Nested groundwater flow systems (NGFS) are commonplace in various hydrogeological environments, including endorheic basins and coastal aquifers. The subsystems of the NGFS can be spatially separated by streamlines around the internal stagnation points. At the discharge zones of the topographic depressions, saline water often emerges due to high evaporation in endorheic drainage basins, or the combined effects of evaporation and intermittent seawater submersion in coastal areas. To date, there are limited studies that have considered the impact of local‐scale downward migration of saline plumes in the topographic depressions on the NGFS. In this study, the classic NGFS are revisited by considering saline water in their discharge zones. To quantify the effects of salinity in the discharge zones on the NGFS, scenarios of various salinities in the discharge zones are simulated. The displacements of the internal stagnation points are used to quantify the evolution of the NGFS in response to salinity changes in the discharge zones. The results show that, as the salinity in the discharge zones increases, the hydraulic gradient near the discharge zone can be significantly reduced, the internal stagnation points shift upward, and the local groundwater flow systems retreat upward so that their original spaces are replaced by intermediate or regional flow systems. The discharge zone is expanded and the overall groundwater flow velocity magnitude of the entire system decreases with salinity. This study may shed light on the management of saline wetlands, e.g. the control of groundwater salinization, evolution of saline groundwater basins, and seawater intrusion.
Article
Full-text available
Groundwater in alluvial terrain, supporting much of the global requirement for irrigation and domestic water, is at risk of sustained water level decline and/or contamination over large areas. In southern Bangladesh excessive arsenic (As) in shallow groundwater has led to deeper groundwater becoming the preferred alternative source of potable water. The vulnerability of deeper tube-wells to As breakthrough from shallow levels can be assessed using groundwater modelling, but representation of alluvial aquifer heterogeneities in large-scale groundwater models presents a challenge; what level of complexity is required? To assess the optimum level of complexity necessary in models of groundwater flow and solute transport in the Bengal Aquifer System (BAS), we explore a range of representations of the lithological heterogeneity using a multi-modelling approach. We use an array of geological information including drillers' logs (n = 589) and hydrocarbon exploration data (n = 11) across an area of 5000 km 2 as a basis for alternative representations of upscaled aquifer heterogeneity, characterising hydrogeological structure in a series of five groundwater models at increasing levels of complexity. We rank the models by comparing model outcomes of travel time with available data on groundwater age based on 14 C. The results demonstrate the importance of spatial heterogeneity and suggest the significance of incorporating vertical heterogeneity in model representations of the Bengal Aquifer System and similar spatially extensive fluvio-deltaic aquifers.
Article
Full-text available
With extensive ancient karsts and abundant oil and gas resources, the Ordovician carbonate reservoirs in the Huanghua depression have broad prospects of exploration and development. The types, characteristics, distribution and controlling factors of karsts were investigated by means of the analysis of cores, thin slices, logging, seismic, and geochemical data in this study. The study reveals that the Ordovician carbonate reservoirs in the Huanghua depression falls into three main categories, pore type, pore-vug-fracture type and fracture type, of which, pore-vug-fracture reservoirs account for 62%. The epigenetic uniform karst mainly developed in the Late Ordovician to Early Carboniferous, while the differential karst developed in the Mesozoic. The uniform karstification mainly include chemical dissolution of the original carbonate rock, while the differential karstification not only include chemical dissolution but also hydraulic erosion and biochemical dissolution by organic acid. The burial karstification mainly involved the dissolution of the Ordovician reservoir by large amount of acidic fluids generated by the overlying Upper Paleozoic coal measure source rock while generating hydrocarbon. The dissolution caused by upwelling magmatic-tectonic hydrothermal fluids along faults also contributed to the formation of burial karst reservoirs. The sedimentation-diagenesis and paleogeomorphology-tectonic movement have certain control effects on the karstification of the Ordovician reservoirs in the Huanghua depression, determining the reservoir quality.
Article
Full-text available
This study offers a reinterpretation of archive aquifer tests, predominantly on the basis of recovery data, from an original datasheet of thermal water wells located in carbonate and sandstone aquifer units in the vicinity of Budapest, Hungary. The study compares the hydraulic conductivity (K) and specific storage (Ss) values derived in the first instance from an aquifer test evaluation. This included an initial application of the classical analytical Cooper and Jacob method. Subsequently, the visual two-zone (VTZ) numerical method was applied, then third, a more complex model, namely, WT software. It was found that the simple analytical solution is not able to represent the field conditions accurately, while in the course of the application of the VTZ model, it proved possible to alter the various hydraulic parameters within reasonable limits to fit the field data. In the case of the VTZ model, the researcher is required to calculate the accuracy of the fitted model separately, while with the WT model, this is automatic, the software seeks out the best fit. In addition to VTZ parameters, the WT model can efficiently incorporate data on up to 500 model layers, water level, and pressure. The optimization of the parameters may be achieved by automatic calibration, improving the accuracy of the numerical results. Recovery tests for 12 wells were numerically simulated to obtain values for vertical and horizontal hydraulic conductivity and specific storage for Triassic and Eocene fractured carbonate and the Upper-Miocene-Pliocene granular sandstone aquifer units. When an analytical solution is applied, only average values could be obtained. The conclusion reached was that the results of the analytical solution can be improved by the use of numerical methods. These methods are able to incorporate basic information on well design, aquifer material and the hydrogeological environment in the course of the evaluation. The revision of the archive recovery data using numerical methods may assist in the quest for better data for numerical flow and transport simulations without the need to perform new tests. In addition, the methods employed here can explain cases in which the original analytical interpretations proved unable to yield reliable data and predictions.
Article
Full-text available
We report the application of machine learning methods for predicting the effective diffusivity (De) of two-dimensional porous media from images of their structures. Pore structures are built using reconstruction methods and represented as images, and their effective diffusivity is computed by lattice Boltzmann (LBM) simulations. The datasets thus generated are used to train convolutional neural network (CNN) models and evaluate their performance. The trained model predicts the effective diffusivity of porous structures with computational cost orders of magnitude lower than LBM simulations. The optimized model performs well on porous media with realistic topology, large variation of porosity (0.28–0.98), and effective diffusivity spanning more than one order of magnitude (0.1 ≲ De < 1), e.g., >95% of predicted De have truncated relative error of <10% when the true De is larger than 0.2. The CNN model provides better prediction than the empirical Bruggeman equation, especially for porous structure with small diffusivity. The relative error of CNN predictions, however, is rather high for structures with De < 0.1. To address this issue, the porosity of porous structures is encoded directly into the neural network but the performance is enhanced marginally. Further improvement, i.e., 70% of the CNN predictions for structures with true De < 0.1 have relative error <30%, is achieved by removing trapped regions and dead-end pathways using a simple algorithm. These results suggest that deep learning augmented by field knowledge can be a powerful technique for predicting the transport properties of porous media. Directions for future research of machine learning in porous media are discussed based on detailed analysis of the performance of CNN models in the present work.
Article
The extension of irrigated agriculture is essential to fulfilling the mounting demand for food and fibre from a growing population which is anticipated to reach about 10 billion individuals in 2050. Lacking the necessary drainage, this extension can bring about salinization issues in irrigated regions. Continuous irrigation over many years with no adequate drainage services has led to huge agricultural regions becoming unproductive. Salinization problems linked to poor drainage are widespread in agricultural zones over the world. The aforementioned drainage‐related salinization issues in agricultural land have been solved by using various methods and techniques. This paper presents an assessment of some non‐conventional methods and strategies utilized for dealing with salinization and drainage issues of irrigated land. An indication of the drainage and salinization issues of irrigated areas and the significance of the study are given. Reasoning and foundation of the issues are presented. Numerical and analytical solutions of the problems are described. The drainage water management technique for managing nutrient losses from a drainage system is discussed. The analysis revealed that conjunctive water use with an increased groundwater proportion can to some extent manage the salinization and root‐zone submergence problems of irrigated areas. © 2020 John Wiley & Sons, Ltd.
Article
Geological heterogeneity is a key factor affecting rates and patterns of groundwater flow and the evolution of salinity distributions in coastal aquifers. The hydrogeologic systems of volcanic aquifers are characterized by lava flows that can form connected geologic structures in the subsurface. Surface-based geostatistical techniques were adopted to generate geologically-realistic, statistically equivalent model realizations of a hydrogeologic system based on that of the Big Island of Hawaii (conduit models). The density-dependent groundwater flow and solute transport code SEAWAT was used to perform 3D simulations to investigate subsurface flow and salt transport through these random realizations. Statistically equivalent geological systems generated with sequential indicator simulation (SIS) and equivalent homogeneous systems were also simulated for comparison. Simulation results show that conduit realizations tend to have more complex salinity distributions with larger mixing zones for which the center of mass is farther offshore compared to homogeneous and SIS realizations. The zone of groundwater discharge is also farther offshore than for homogeneous and SIS models, with higher point fluxes of both fresh and saline water and greater spatial variability. This highlights the importance of the geometry of geologic features in coastal groundwater flow and solute transport processes in highly heterogeneous aquifers and effects of preferential flow, with implications for both coastal groundwater resource management and solute fluxes to the ocean.