Hostname: page-component-f554764f5-qhdkw Total loading time: 0 Render date: 2025-04-22T01:25:08.312Z Has data issue: false hasContentIssue false

MIXED BOUNDARY VALUE PROBLEM IN MICROFLUIDICS: THE AVER’YANOV–BLUNT MODEL REVISITED

Published online by Cambridge University Press:  07 April 2025

ANVAR KACIMOV*
Affiliation:
Sultan Qaboos University, Seeb, Sultanate of Oman; e-mail: [email protected]
YURII OBNOSOV
Affiliation:
Kazan Federal University and Institute of Forest Science, Russian Academy of Sciences, Moscow, Russia; e-mail: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

Macroscopically, a Darcian unsaturated moisture flow in the top soil is usually represented by an one-dimensional volume scale of evaporation from a static water table. On the microscale, simple pore-level models posit bundles of small-radius capillary tubes of a constant circular cross-section, fully occupied by mobile water moving in the Hagen–Poiseuille (HP) regime, while large-diameter pores are occupied by stagnant air. In our paper, cross-sections of cylindrical pores are polygonal. Steady, laminar, fully developed two-dimensional flows of Newtonian water in prismatic conduits, driven by a constant pressure gradient along a pore gradient, are more complex than the HP formula; this is based on the fact that the pores are only partially occupied by water and immobile air. The Poisson equation in a circular tetragon, with no-slip or mixed (no-shear-stress) boundary conditions on the two adjacent pore walls and two menisci, is solved by the methods of complex analysis. The velocity distribution is obtained via the Keldysh–Sedov type of singular integrals, and the flow rate is evaluated for several sets of meniscus radii by integrating the velocity over the corresponding tetragons.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1 Introduction

The motivation to apply microfluidic models to the hydrology of “solonchaks” is presented in this section. The fluxes of all pore fluids (liquid $\mathrm {H_2O}$ and soil gases) in the topsoil, especially in the vadose zone, are of utmost importance in soil hydrology, civil engineering and agro-environmental management of soil resources, in particular in hot dry lands [Reference Kovda23]. In arid deserts, the topsoil is usually very dry and the water table is deep (5+ m), that is, evaporation is negligible. However, at many urban locales of the Middle Eastern and North African countries, a new phenomenon of shallow (perched) water tables was recently discovered and mathematically modelled [Reference Avkhadiev and Kacimov7, Reference Kacimov, Al-Maktoumi, Al-Ismaily, Al-Mayahi, Al-Shukaili, Obnosov and Osman20]. The capillary fringe of these aquifers is a zone extensively emitting water by evapotranspiration to the desert atmosphere. This zone is only 10 to 50 cm thick. Mitigating the catastrophic consequences of waterlogging is imperative for the governments and public in the affected regions, especially in the urbanized catchment of Oman and the other Gulf countries. Dialectically, there are also ecohydrologically benign effects of shallow water tables (in particular, if groundwater is brackish or saline): for example, phreatophytes and halophytes thrive and deposit/sequester CO ${}_2$ (see [Reference Kacimov19]). The main question that soil physicists have to answer is: “How does one evaluate the ascending flux of moisture from a shallow, almost static, water table”?

We recall that direct measurements of evaporation require buried lysimeters (not used in Oman and other Gulf countries), which are expensive and cumbersome. In Australia, where shallow water tables and the ensuing waterlogging became a national disaster in the 1970s [Reference Thorburn33], the measured moisture fluxes (maintained by evapotranspiration and conceptualized as one-dimensional on the representative elementary volume (REV) scale) were 500+ mm/year, that is, hydrologically huge. Similarly, in the USSR (Central Asia) irrigation engineers confronted similar hyperintensive evaporation and salinization of agricultural lands [Reference Aver’yanov5]. The microfluidic models of this process were, however, not developed by Australians or Soviets.

Our paper analyses the microscopic process of two-dimensional pore water motion, which takes place in desert soils classified as “solonchaks” (“sabkhas” in Arabic) [Reference Chhabra11, Reference Kovda23]. The analytic solutions give integral representations (Keldysh–Sedov type [Reference Henrici18]) via singular integrals for an analytic function, found in a canonical domain (half-plane). Modelling of the pore water dynamics at this scale can be an overture to further field measurements of moisture motion in thin vadose zones and to developing the means of optimally controlling the macroscopic evapotranspiration.

The monitoring of perched aquifers in Oman [Reference Al-Battashi, Al-Maktoumi, Kacimov, Al-Mayahi, Al-Shukaili and Al-Ismaily2] evidenced that the depth of the water table in most of them is not only small but also almost constant. Consequently, although the atmospheric conditions vary with seasons, the annually averaged evaporation flux in our model is assumed to be steady state. We also assume that the mobile soil water phase is continuous, that is, it moves within soil pores under a constant pressure gradient, whereas the air phase is static. In soil physics, most models assume connected moisture pathway flow at the pore scale that implies the functional form of the two-phase Darcy equations. In problems with rapidly fluctuating water tables caused by, for example, infiltration, rather than evaporation, the water retention functions and water relative permeability are dynamic (in particular, hysteretic), that is, associated with rapidly moving breathing menisci and water–air contact lines. The rapidity of infiltration of irrigated water into the topsoil, as well as the propagation of the oil–water contacts in secondary recovery of crude oils in formations [Reference Blunt9], are very different from the “rapidity” of salinization, as pinpointed by Aver $'$ yanov [Reference Aver’yanov5]. A typical timescale of infiltration is minutes and hours, while for Aver $'$ yanov”s evaporative fluxes and the ensuing primary/secondary salinization and hydromorphism of “solonchaks” it is months and years.

Macroscopically, such flows are classified as “variably saturated” (see [Reference Allen3, Section 6.2] for more details), and obey the Richards–Richardson [Reference Radcliffe and Šimůnek28] nonlinear parabolic partial differential equation (PDE), which can be solved analytically for steady-state flows, as in two-dimensional evaporation in the vicinity of either a ponded land surface (application to flood irrigation [Reference Weir35]) or an impeding surface (buildings foundation in urban hydrology [Reference Buchwald and Viera10]). In case of transient seepage, the PDE is numerically solved. On the pore scale, water motion obeys the Navier–Stokes PDE, which, under the simplifications of our model, is reduced to the Poisson equation [Reference Avkhadiev and Kacimov6, Reference Shah and London31].

Models for relative permeabilities of the water $k_{rw}$ and air $k_{ra}$ phases in macroscopically (REV-scale) two-phase steady-state Darcian flows through porous media are applied in soil physics, and in reservoir, geotechnical and chemical engineering. In these models, semi-empirical dependencies of $k_{rw}$ and $k_{ra}$ on the degrees of saturation with respect to water, $S_w$ , or on the capillary pressure (matric potential/suction) head, p, are used [Reference Blunt9, Reference Dullien13, Reference Fredlund and Rahardjo16, Reference Philip24]). In the Richards–Richardson PDE for the mobile phase only, the functions $k_{rw}(S_w)$ or $k_{rw}(p)$ were developed in [Reference Aver’yanov5, Reference Fredlund and Rahardjo16, Reference Philip25] among others. On a microscale, Aver $'$ yanov [Reference Aver’yanov4] considered a partially filled circular pore and solved the Poisson equation (see, for example, [Reference Polubarinova-Kochina26] for details of Aver $'$ yanov”s derivations) in an annulus with a no-slip condition for water velocity along the pore wall (a circle with a large radius) and a no-shear condition along a meniscus (a concentric circle with a smaller radius). Aver $'$ yanov found that $k_{rw}=k_w/k_0=S_w^{m_A}$ , where $0\leq S_W\leq 1$ , and $k_w$ and $k_0$ are the unsaturated and saturated (intrinsic) hydraulic conductivities of the soil, respectively. Aver $'$ yanov”s analytically derived exponent $m_A \approx 3.5$ (see further empirical analysis in [Reference Daneshian, Habibagahi and Nikooee12, Reference Salugin29]) is consistent with the empirical Campbell analysis, which was obtained for pores of various radii but fully occupied by capillary water. The Van Genuchten relative permeability function [Reference Radcliffe and Šimůnek28] for pore water is

$$ \begin{align*}k_{rw} =S_{w}^{1/2} \{1-(1-S_{w} ^{1/m_{VG} } )^{m_{VG} } \}^{2}, \end{align*} $$

where $m_{VG}$ is a soil-specific empirical constant. Aver $'$ yanov”s exact and explicit solutions were generalized to the case of two Poisson equations in [Reference El-Khatib14], where the conjugation condition was imposed on a circular interface (meniscus), which separates water and oil phases concurrently moving in a steady-state regime.

In our view, there is an issue with the REV versus pore-level upscaling–downscaling: bundles of conduits are most often postulated to be circles (in their cross-sections perpendicular to the macroscopic pressure gradient). The real pore channels are obviously not circular, even if cylindrical. Such “circularization” of water conduits stems, obviously, from the availability of the Hagen–Poiseuille (HP) formula [Reference Allen3] for a one-phase flow through a pipe. In unsaturated moisture flow models [Reference Radcliffe and Šimůnek28], microscopic tubes of sufficiently small radii are believed to convey the pore water in the HP regime, while the air is static in large-diameter empty pores. In transient quasi-HP regimes of rapid imbibition/drainage, for example, in the Green–Ampt infiltration model [Reference Philip24], the HP formula is usually adjusted by assuming a propagating meniscus.

Microfluidic models (see Figures 2.16, 3.4, 3.6, 4.1, 4.2 and 5.5 in [Reference Blunt9]) consider polygons (triangles and rectangles) rather than circles as cross-sections of the pore channels making parallel bundles. The vertices of these polygons (angular pores) allow accumulation of the water phase in the crevices of pores (the case tackled by Kacimov et al. [Reference Kacimov, Maklakov, Kayumov and Al-Futaisi22]) or in water “bridges”, sandwiched between immobile bulk air phases by two zero-shear stress menisci (see Figures 2.29, 5.6 and 5.16 in [Reference Blunt9] and our Figure 1).

Figure 1 (a) Cross-section of an arbitrary triangular pore. Three thin mobile water “bridges” are sandwiched between immobile air. (b) 3D diagram with zoomed menisci $CNB$ and $DMA$ near one pore corner. (c) Boundary conditions in the flow domain, a circular tetragon $G_{z}$ .

In this paper, we use the theory of holomorphic functions [Reference Gakhov17, Reference Henrici18], and we obtain a new integral solution to a mixed boundary value problem (BVP) for the Poisson equation in a circular quadrangle $G_z$ , as shown in Figure 1, which has zero pore velocity on the straight sides of the quadrangle and zero-shear boundary condition on two menisci. In other words, our paper is a confluence of the Aver $'$ yanov combination of the boundary conditions (tackled analytically) with Blunt”s more realistic geometry of pore cross-sections.

We recall that an area integration of the Prandtl function, which also obeys the Poisson equation, gives a torsional rigidity of an elastic bar [Reference Timoshenko and Goodier34]. In microfluidics, a mathematically equivalent integration of the magnitude of the pore water velocity over $G_z$ gives the volumetric discharge, provided the air phase is stagnant. In Section 2, a mixed (no-slip versus no-shear conditions) BVP is solved in a curvilinear quadrangle for the Poisson equation, which the velocity of the water phase obeys. The moisture content within the triangular pores is relatively low and the air phase is static. In Section 3, the obtained solution is interpreted as an approximation of water flow in the case of high moisture content and stagnant air confined near the corners of the pore, as well as for the case of the dominating gas phase flowing through the bulk volume of the pores, while the immobile water phase is trapped near the corners. In Section 4, the limitations and perspectives of our analytical model are discussed.

2 Analytical solution to a mixed BVP, Poisson”s PDE

In a cross-section perpendicular to the macroscopically vertical direction Of (ascending evaporation), we assume that the pore makes a triangle $OO_{1}O_{2}$ having arbitrary sides $c_{1}$ , $c_{2}$ , $c_{3}$ , (Figure 1(a)). As a geometrical input of our model, we consider the vicinity of one corner (O) of this triangle where water is contained in a “bridge” bounded by two menisci, $AMD$ and $CNB$ , and two pore walls making an angle $\alpha $ , $0 <\alpha <\pi $ (Figure 1(b)). The menisci are circular arcs of constant radii, $r_{1}$ and $r_{2}$ , respectively. We assume that the contact angle $\theta =\pi /2$ .

The gradient of soil water pressure $\nabla p_{w} $ is assumed to be constant, as in [Reference Kacimov, Maklakov, Kayumov and Al-Futaisi22]. Such gradients emerge due to, for example, an increase in the disjoining and capillary pressure caused by the above-mentioned gradual “evaporative thinning” (that is, the decrease of $r_{2}$ and increase of $r_{1}$ in the Of direction) of the water bridges. A higher evaporation rate from menisci of larger diameters (just above the capillary fringe in desert solonchaks), as compared with the vicinity of the super hot desert surface, is well documented.

We introduce a complex physical coordinate $z=x+{i}y=r{e}^{{i}\varphi } $ . The water microfluidic velocity obeys the Poisson equation [Reference Avkhadiev and Kacimov6]: that is

(2.1) $$ \begin{align} \Delta u(r,\varphi )=\frac{\partial ^{2} u}{\partial r^{2} } +\frac{1}{r} \frac{\partial u}{\partial r} +\frac{1}{r^{2} } \frac{\partial ^{2} u}{\partial \varphi ^{2} } =-e,\quad e=\nabla p_{w} /\mu _{w}, \end{align} $$

where $\mu _{w}$ is the viscosity of the pore water. The right-hand side, e, in equation (2.1) is a given constant as a model hydrodynamic input (see [Reference Kacimov, Maklakov, Kayumov and Al-Futaisi22, Reference Philip25] for more details of similar flows in a circular digon and triangle).

The domain $G_{z}$ in Figure 1(c) is a curvilinear tetragon made of two straight segments $DC$ and $BA$ and two arcs $AMD$ and $CND$ . At the boundary of $G_{z}$ , the following conditions hold:

(2.2) $$ \begin{align} \begin{array}{ll} u(r{e}^{\pm {i}\alpha /2} )=0 & r_{1} < r< r_{2} , \\ \dfrac{\partial u(r_{k} {e}^{{i}\varphi } )}{\partial r} =0 & k=1,2,\quad -\alpha /2<\varphi <\alpha /2, \end{array} \end{align} $$

where r is a radial coordinate from the corner O and $\varphi $ is an angular coordinate counted from the axis of symmetry of the bridge, that is, $OMN$ . The pair $r_{1}$ and $r_{2}$ provide another geometrical input of our model. The no-slip (the first line in equation (2.1)) and no-shear-stress (second line) conditions physically mean that the rigid skeleton of our porous medium is static (always true) and $\mu _{a}$ is almost zero, as was assumed in [Reference Kacimov, Maklakov, Kayumov and Al-Futaisi22]. The latter assumption is justified by the two-order difference in viscosities of air and pore water. The main output of the model given by (2.1)–(2.2) is the velocity distribution $u(r, \varphi )$ and ensuing areal integral of this quantity over $G_{z}$ , which gives the volumetric flow rate, and hence the water phase permeability. Extension of the model given by (2.1)–(2.2) to more complex menisci can be found in [Reference Finn15] among other applications to microgravity and soil physics.

The solution of a BVP with the $u = 0$ condition along the whole wall of our $G_z$ was obtained by Sparrow et al. [Reference Sparrow, Chen and Jonsson32], and it was reported by Shah and London [Reference Shah and London31]. Unfortunately, no details of the solution are presented in [Reference Shah and London31, Reference Sparrow, Chen and Jonsson32]. Moreover, equation (394) in [Reference Shah and London31] does not match equation (7) in [Reference Sparrow, Chen and Jonsson32]. Sparrow et al. used the method of separation of variables for solving their BVP for a harmonic function obeying Dirichlet”s boundary conditions, that is, they did not use the theory of holomorphic functions.

Aver $'$ yanov”s [Reference Aver’yanov4] solution to PDE (2.1) can be readily reformulated for $G_{z}$ in the following manner: the boundary conditions are no slip along $DA$ in Figure 1(c) and no shear along $DC$ , $CB$ and $CA$ .

Figure 2 (a) Rectangle $G_{z_1}$ . (b) Reference half-plane $G_{\zeta }$ .

Similar to the work in the papers [Reference Kacimov, Yakimov and Šimůnek21, Reference Kacimov, Maklakov, Kayumov and Al-Futaisi22, Reference Sparrow, Chen and Jonsson32], we reduce the BVP (2.1)–(2.2) for the Poisson equation to a BVP for the Laplace equation. For this purpose, we use the transformation

(2.3) $$ \begin{align} \begin{array}{ll} {z_{1} =\log (z/r_{1} )=x_{1} +{i}y_{1} =\log (r/r_{1} )+{i}\varphi , \quad x_{1} =\log (r/r_{1} ),\quad y_{1} =\varphi ,} \\[.2cm] {\Phi (z)=u(r,\varphi )+er^{2} /4, \quad \Phi _{1} (z_{1} )=\Phi (r_{1} \exp \, z_{1} ),} \end{array} \end{align} $$

which makes the function u in equation (2.1) harmonic in a rectangle $G_{z_{1} } $ (Figure 2(a)): that is,

(2.4) $$ \begin{align} \Delta \Phi _{1} (x_{1} ,y_{1} )=0 \quad \textrm{for} \quad z_{1} =x_{1} +{i}y_{1} \in G_{z_{1}}. \end{align} $$

The boundary conditions (2.2), (2.3) in $G_{z_{1} } $ become

(2.5) $$ \begin{align} AB, DC: &\, \Phi _{1} (x_{1} ,\pm \alpha )=0.25\mathrm{\; }e\mathrm{\; }r_{1}^{2} {e}^{2x_{1} } \quad 0<x_{1} <l=\log (r_{2} /r_{1} ), \nonumber\\ AD: & \,\dfrac{\partial \Phi _{1} (0,y_{1} )}{\partial x_{1} } =\frac{e\mathrm{\; }r_{1}^{2} }{2} \quad -\alpha /2<y_{1} <\alpha /2, \\ BC:& \, \dfrac{\partial \Phi _{1} (l,y_{1} )}{\partial x_{1} } =\frac{e\mathrm{\; }r_{2}^{2} }{2} \quad -\alpha /2<y_{1} <\alpha /2.\nonumber \end{align} $$

Clearly, the function $\Phi _{1} (x_{1} ,-y_{1} )$ is a solution of the mixed BVP (2.4) and (2.5) together with $\Phi _{1} (x_{1} ,y_{1} )$ . Hence,

$$ \begin{align*}\Phi _{1} (x_{1} ,-y_{1} )=\Phi _{1} (x_{1} ,y_{1} ),\end{align*} $$

due to the uniqueness theorem [Reference Henrici18].

Next, we introduce the complex potential function

(2.6) $$ \begin{align} \Omega _{1} (z_{1} )=\Phi _{1} (x_{1} ,y_{1} )+{i}\Psi _{1} (x_{1} ,y_{1} ), \end{align} $$

which is holomorphic in $G_{z_{1} } $ . Here the function $\Psi _{1} (x_{1} ,y_{1} )$ , being complex conjugate to $\Phi _{1} (x_{1} ,y_{1} )$ , satisfies the symmetry condition

$$ \begin{align*} \Psi _{1} (x_{1} ,-y_{1} )=-\Psi _{1} (x_{1} ,y_{1} ). \end{align*} $$

Due to equations (2.5) and the Riemann–Schwartz symmetry principle [Reference Gakhov17], the function (2.6) admits analytic continuations through the vertical sides of the boundary of the rectangle $G_{z_{1} } $ . Hence, the Cauchy–Riemann conditions (CRC) hold up to these sides. Using the CRC $\partial \Phi _{1} /\partial x_{1} =\partial \Psi _{1} /\partial y_{1} $ and the last two conditions of (2.5) and integrating, we get

(2.7) $$ \begin{align} \Psi _{1} (0,y_{1} )=0.5er_{1}^{2} y_{1} ,\quad \Psi _{1} (l,y_{1} )=0.5er_{2}^{2} y_{1} \quad -\alpha <y_{1} <\alpha. \end{align} $$

It is noteworthy that an arbitrary constant arising after integrating the CRC condition is equal to zero due to the above-mentioned symmetry property of the function $\Psi _{1} (x_{1} ,y_{1} )$ .

We map conformally $G_{z_{1} } $ onto the upper half-plane $\mathrm {Im}\zeta>0$ , $G_{\zeta }={\mathbb C}^+ $ (Figure 2(b)), with the correspondence of the points

$$ \begin{align*}C\to -1/\lambda , \quad D\to -1, \quad A\to 1, \quad B\to 1/\lambda , \end{align*} $$

where $0<\lambda <1$ . The Schwarz–Christoffel formula [Reference Polubarinova-Kochina26] gives

(2.8) $$ \begin{align} z_{1} (\zeta )=-\frac{{i}\alpha }{2K(\lambda )} \int _{0}^{\zeta }\frac{{d}t}{\sqrt{(1-t^{2} )(1-\lambda ^{2} t^{2} )} } =-\frac{{i}\alpha }{2K(\lambda )} \mathrm{F}(\arcsin \zeta ,\lambda ) , \end{align} $$

where $\mathrm {F}(\arcsin \zeta ,\lambda )$ and $K(\lambda )$ are, respectively, incomplete and complete elliptic integrals of the first kind (see [Reference Abramowitz and Stegun1, formulae 17.2.7, 17.3.1]). The modulus $\lambda $ is determined from the condition

(2.9) $$ \begin{align} K'(\lambda )/K(\lambda )=2l/\alpha, \end{align} $$

where $K'(\lambda )=K(\lambda ')$ , $\lambda '=\sqrt {1-\lambda ^{2} } $ . We use the FindRoot routine of Mathematica [Reference Wolfram36] and find $\lambda $ from equation (2.9).

Combining conditions (2.5) and (2.7), we obtain the following mixed BVP for the function $\Omega (\zeta )=\Omega _{1} (z(\zeta ))$ in $G_{\zeta } $

(2.10) $$ \begin{align} \begin{array}{ll} \mathrm{Re}\Omega (\xi )=0.25er_{1}^{2} {e}^{2x_{1} (\xi )} & 1<|\xi |<1/\lambda , \\[.2cm] \mathrm{Im}\Omega (\xi )=0.5er_{1}^{2} y_{1} (\xi ) & -1<\xi <1, \\[.2cm] \mathrm{Im}\Omega (\xi )=0.5er_{2}^{2} y_{1} (\xi ) & |\xi |>1/\lambda. \end{array} \end{align} $$

Here $x_{1} (\xi )$ and $y_{1} (\xi )$ are real and imaginary parts of the function (2.8), respectively: that is,

(2.11) $$ \begin{align} \begin{aligned} x_{1} (\xi )&=\dfrac{\alpha }{2K(\lambda )} \int_{1}^{\xi }\frac{{d}t}{\sqrt{(t^{2} -1)(1-\lambda ^{2} t^{2} )} } =\frac{\alpha }{2K(\lambda )} F\bigg(\!\arcsin \dfrac{\sqrt{1-\xi ^{-2} } }{\lambda '} ,\lambda '\bigg) \quad 1<\xi <1/\lambda , \\ y_{1} (\xi )&=-\dfrac{\alpha }{2K(\lambda )} \int_{0}^{\xi }\dfrac{{d}t}{\sqrt{(1-t^{2} )(1-\lambda ^{2} t^{2} )} } =\dfrac{\alpha }{2K(\lambda )} F(\arcsin \xi ,\lambda ) \quad-1<\xi <1, \\ y_{1} (\xi )&=\dfrac{\alpha }{2K(\lambda )}\! \int_{1/(\lambda \xi )}^{1}\!\dfrac{{d}t}{\sqrt{(1-t^{2} )(1-\lambda ^{2} t^{2} )} } -\dfrac{\alpha }{2} \\ &=\frac{\alpha }{2K(\lambda )} F\bigg(\!\arcsin \sqrt{\dfrac{\xi ^{2} -1/\lambda ^{2} }{\xi ^{2} -1} } ,\lambda \bigg)-\dfrac{\alpha }{2} \quad \xi>1/\lambda . \end{aligned} \end{align} $$

It is clear that $x_{1} (\xi )=x_{1} (-\xi )$ for $-1/\lambda <\xi <-1$ and $y_{1} (\xi )=-y_{1} (-\xi )$ for $\xi <-1/\lambda $ .

Thus, we have to find $\Omega (\zeta )$ , which is holomorphic in $G_{\zeta } $ . This function must satisfy boundary conditions (2.10). It must be bounded at all transition points $\pm 1$ , $\pm 1/\lambda $ and at infinity in Figure 2(b). The index of the BVP (2.10) in the indicated class of functions is $-2$ , and, generally speaking, one solvability condition must be satisfied [Reference Gakhov17]. As will be shown below, this solvability condition takes place automatically due to the properties of functions (2.11).

Solution of problem (2.10) bounded at transition points, but generally, not at infinity, [Reference Henrici18] is

$$ \begin{align*} \Omega (\zeta )=\frac{er_{1}^{2} \omega \, (\zeta )}{2\pi } \bigg(\!\int_{-1}^{1}\frac{y_{1} (\xi )}{\omega \, (\xi )} \frac{{d}\xi }{\xi -\zeta } +\int_{L_\lambda }\frac{{e}^{2x_{1} (\xi )} }{2|\omega \, (\xi )|} \frac{{d}\xi }{\xi -\zeta } +{e}^{2l} \int_{1/\lambda }^{-1/\lambda }\frac{y_{1} (\xi )}{\omega \, (\xi )} \frac{{d}\xi }{\xi -\zeta } \bigg),\end{align*} $$

where $L_\lambda =(-1,-1/\lambda )\cup (1,1/\lambda )$ , and the last integral is taken from $1/\lambda $ to $-1/\lambda $ through infinity. The branch of the function

(2.12) $$ \begin{align}\omega (\zeta )=\sqrt{(1-\zeta ^{2} )(1-\lambda ^{2} \zeta ^{2} )} , \end{align} $$

is fixed in the upper half-plane by the condition of its positivity in the interval (-1,1). This branch is real and negative at $\zeta =\xi \in (-\infty ,-1/\lambda )\cup (1/\lambda ,\infty )$ . The branch is pure imaginary with negative and positive imaginary parts at the intervals $(1,1/\lambda )$ and $(-1/\lambda ,-1)$ , respectively. It may seem that the $\Omega (\zeta )$ obtained in equation (2.12) is unbounded at infinity. If, however, we take into account the properties of functions (2.11) and (2.12), this solution can be rewritten as

(2.13) $$ \begin{align} \Omega (\zeta )=\frac{er_{1}^{2} \omega (\zeta )}{\pi } \bigg(\!\int_{0}^{1}\frac{\xi y_{1} (\xi )}{\omega (\xi )} \frac{{d}\xi }{\xi ^{2} -\zeta ^{2} } +\frac{1}{2} \int_{1}^{1/\lambda }\frac{\xi {e}^{2x_{1} (\xi )} }{|\omega (\xi )|} \frac{{d}\xi }{\xi ^{2} -\zeta ^{2} } +{e}^{2l} \int_{1/\lambda }^{\infty }\frac{\xi y_{1} (\xi )}{\omega (\xi )} \frac{{d}\xi }{\xi ^{2} -\zeta ^{2} } \bigg). \end{align} $$

From equation (2.13), it is clear that our solution is bounded at $\zeta \to \infty $ , as it should be. Note that the functions (2.8), (2.12) and (2.13) satisfy the same symmetry condition

(2.14) $$ \begin{align}z_{1} (\zeta )\equiv \overline{z_{1} (-\bar{\zeta })}, \quad \omega (\zeta )\equiv \overline{\omega (-\bar{\zeta })}, \quad \Omega (\zeta )\equiv \overline{\Omega (-\bar{\zeta })}. \end{align} $$

The total flow rate through $G_{z}$ , due to equation (2.3), is

$$ \begin{align*} Q=\int_{G_{z} }u(x,y)\,{d}s =\int _{G_{z} }\Phi (x,y)\,{d}s -\frac{e\alpha }{16} (r_{2}^{4} -r_{1}^{4} ). \end{align*} $$

The last integral can be transformed as

$$ \begin{align*} Q_{0} =\int _{G_{z} }\Phi (x,y)\,{d}s =\frac{{i}}{2} \int _{G_{z} }\Phi (x,y)\,{d}z\wedge {\kern 1pt} {d}\bar{z}=\frac{{i}r_{1}^{2} }{2} \int _{G_{z_{1} } }\Phi _{1} (x_{1} ,y_{1} )e^{2x_{1} } \,{d}z_{1} \wedge {\kern 1pt} {d}\bar{z}_{1} , \end{align*} $$

where $\wedge $ is the symbol for the exterior or wedge product. Using the transformation (2.8), we obtain

(2.15) $$ \begin{align} Q_{0} =\frac{{i}r_{1}^{2} \alpha ^{2} }{16K^{2} } \int_{{\mathbb C}^{+} }\frac{\Omega (\zeta )+\overline{\Omega (\zeta )}}{\omega (\zeta )\overline{\omega (\zeta )}} e^{2\mathrm{Re} z_{1} (\zeta )} \,{d}\zeta \wedge {d}\bar{\zeta }=\frac{r_{1}^{2} \alpha ^{2} }{4K^{2} } \int_{-\infty }^{\infty }\int_{0}^{\infty }\frac{\mathrm{Re}\Omega (\xi +{i}\eta )}{|\omega (\xi +{i}\eta )|^{2} } e^{2x_{1} (\xi +{i}\eta )} \,{d}\eta\, {d}\xi. \end{align} $$

Using the Stokes formula, $Q_{0}$ is transformed into a contour integral along the real axis

$$ \begin{align*}Q_{0} =\frac{{i}r_{1}^{2} \alpha ^{2} }{16K^{2} } \bigg(\!\int_{-\infty }^{\infty }\frac{e^{\overline{z_{1} (\xi )}} }{\overline{\omega (\xi )}} \int_{0}^{\xi }\frac{\Omega (t)e^{z_{1} (t)} }{\omega (t)} {d}t\,{d}\xi -\int_{-\infty }^{\infty }\frac{e^{z_{1} (\xi )} }{\omega (\xi )} \int_{0}^{\xi }\frac{\overline{\Omega (t)}e^{\overline{z_{1} (t)}} }{\overline{\omega (t)}} {d}t\,{d}\xi \bigg).\end{align*} $$

Finally, using the symmetry properties (2.14), we obtain

(2.16) $$ \begin{align} Q=-\frac{r_{1}^{2} \alpha ^{2} }{4K^{2} } \mathrm{Im}\int_{0}^{\infty }\frac{e^{\overline{z_{1} (\xi )}} }{\overline{\omega (\xi )}} \int_{0}^{\xi }\frac{\Omega (t)e^{z_{1} (t)} }{\omega (t)} {d}t\,{d}\xi -\frac{e\alpha }{16} (r_{2}^{4} -r_{1}^{4} ). \end{align} $$

We used the NIntegrate routine of Wolfram”s Mathematica [Reference Wolfram36] to compute the integrals in equations (2.15) and (2.16). It turned out that computation Q using equation (2.16) required up to three days (for example, on a PC Intel(R) Core(TM) i7-4770 CPU 3.40GHz, 3401 IHz, cores: 4, logical processors: 8, which we used) for each value of $r_{1} $ . If one uses equation (2.15) (as we did) the computations take up to an hour.

We introduce the dimensional quantities $Q^{*}=Q/( e r_{2}^{4})$ and $r^{*}=r_{1}/r_{2}$ . Table 1 presents the values of $Q^{*}$ for $\alpha =\pi /3$ and five values of $r^{*}$ .

Table 1 Dimensionless flow rate $Q^{*}$ computed using equations (2.15) and (2.16) for tetragons $G_{z}$ having $\alpha =\pi /3$ and five values of $r^{*}$ .

Remark 1 In addition, for simplification of the expression $\Omega (\zeta )$ in equation (2.15), we used the substitution $\xi \to -1/(\lambda \xi )$ and transformed the last term in equation (2.13) to give

$$ \begin{align*}{e}^{2l} \int_{1/\lambda }^{\infty }\frac{\xi y_{1} (\xi )}{\omega (\xi )} \frac{{d}\xi }{\xi ^{2} -\zeta ^{2} } ={e}^{2l} \int_{-\infty }^{-1/\lambda }\frac{\xi y_{1} (\xi )}{\omega (\xi )} \frac{{d}\xi }{\xi ^{2} -\zeta ^{2} } =-\lambda {e}^{2l} \int_{0}^{1}\frac{\xi y_{1} (\xi )}{\omega (\xi )} \frac{{d}\xi }{1-\lambda ^{2} \xi ^{2} \zeta ^{2} } .\end{align*} $$

To obtain this result, we used the identities

$$ \begin{align*}\omega (-1/(\lambda \zeta ))\equiv -\omega (\zeta )/(\lambda \zeta ^{2} ),\quad z_{1} (-1/(\lambda \zeta ))\equiv l-z_{1} (\zeta ). \end{align*} $$

3 Other applications of water flow at high $S_{w}$ in the pore core with static air in the corners

A curvilinear tetragon, as considered in Section 2 for the case of “thin bridges” and small $S_{w}$ , can be used as a generic mobile phase conduit in the opposite limit of large $S_{w}$ (close to 1). Indeed, let viscous water occupy almost the whole pore (see an equilateral triangle in Figure 3(a)) and move in the Of direction (not shown in Figure 3). Tiny pockets of immobile air are “entrapped” near the three corners in Figure 3(a) (see [Reference Blunt9]) which, in applications to reservoir engineering, can be considered (at the REV-scale) as an early (primary) stage of crude oil recovery with a minor proportion of gas formation in the rock. The water flow domain in Figure 3(a) is a circular sextagon, with point E being its centre. Owing to symmetry, we can consider the curvilinear pentagon $ADE_{C}EE_{B}A$ , that is, the third part of $G_{z}$ in Figure 3(a). Along the broken line $E_{C}EE_{B}$ the shear stress is zero (owing to symmetry). We can approximate this curvilinear pentagon $DE_{C}EE_{B}A$ by a curvilinear tetragon $DE_{C}NE_{B}A$ , in which the wedge $E_{C}EE_{B}$ is replaced by a circular arc $E_{C}NE_{B}$ , that is, we arrive at $G_{z}$ in our Figure 1. Consequently, in $ADE_{C}NE_{B}A$ of Figure 3, all analysis in Section 2 is valid.

Figure 3 (a) Flow of water at high $S_{w}$ in a cylindrical pore channel, the cross-section of which is an equilateral triangle with immobile air pockets entrapped near the corners. (b) Flow of air at low $S_{w}$ in a cylindrical pore channel, the cross-section of which is an equilateral triangle with immobile water pockets entrapped near the corners.

In Figure 3(b), another limiting case of very low $S_{w}$ is depicted. Now the pore gas is moving in the Of direction. The right-hand side in equation (2.1) becomes $e=\nabla p_{a} /\mu _{a} $ , where $\mu _{a}$ is the viscosity of the air and $\nabla p_{a} $ is the gradient of the gas pressure along the pore. Water in Figure 3(b) is entrapped near the triangle corners, that is, it is immobile. Similarly to the case in Figure 3(a), we make a circular tetragon with mixed boundary conditions on its sides such that the analysis in Section 2 is valid. Applications of this pore-scale model are to the emission of in-pore CO ${}_{2}$ from the root zone, which has much higher concentrations of CO ${}_{2}$ than the superjacent atmospheric air. Evaluation of the fluxes of gases through the vadose zone are of utmost importance (see [Reference Ben-Noah and Friedman8, Reference Scanlon, Nicot, Massmann and Warrick30]), especially in the context of the impact of allegedly deleterious greenhouse gases (generated by rhizomes in the soil) on the human race and terrestrial ecosystems.

4 Concluding remarks on limitations and perspectives of our model

Our model and machinery of solving mixed BVPs for holomorphic functions allows us to use a combination of slip and no-slip boundary conditions, which is applicable when one phase is an immobile gas while the pore is polygonal rather than the Aver”yanov variably wet circle. In other words, we extend the cases of digon and trigon with mixed boundary conditions, considered in [Reference Kacimov, Maklakov, Kayumov and Al-Futaisi22, Reference Philip25], to the case of a circular tetragon, on the opposite sides of which these conditions are imposed. The motion of moisture in a “bridge”, which is modelled by such a tetragon, is described by Poisson”s equation. With the help of the Keldysh–Sedov-type formula [Reference Henrici18], we obtained an integral representation for the in-pore velocity.

We have doubts about the solution obtained by Sparrow et al. [Reference Sparrow, Chen and Jonsson32] to the BVP with $u= 0$ on the whole boundary of the tetragon in Figure 1(c). The lack of comparisons in [Reference Sparrow, Chen and Jonsson32] (as well as in [Reference Shah and London31]) with the earlier explicit solution to the Poisson equation in a circular triangle, obtained by Saint–Venant [Reference Timoshenko and Goodier34] is confusing. Indeed, the solution obtained in [Reference Sparrow, Chen and Jonsson32] should degenerate into Saint–Venant”s solution when the small-radius meniscus in Figure 1 vanishes, which was not demonstrated by Sparrow et al. [Reference Sparrow, Chen and Jonsson32].

The estimates of the flow rate Q through polygonal cross-sections with no-shear menisci can be attempted with the help of isoperimetric inequalities [Reference Pólya and Szegö27].

Recent video recording techniques of microfluidics processes can be used as experimental tools for measuring the in-pore parameters $\alpha $ , $r_1$ , $r_2$ in Figure 1.

Solute transport on the microscale was not studied in our paper. In future, the advective dispersion equation on the macroscale (see [Reference Aver’yanov5]) can be downscaled to the pore level. The root zone salinization, caused by evapotranspiration from shallow water tables, is a daunting problem in Australia and other arid countries, where applied mathematicians may, like Aver $'$ yanov, transcend from microfluidics to the scale of engineered drainage of large salinized crop fields.

The limitations of our flow model in Section 2 are as follows.

  1. (1) Viscosity of water is constant (in reality, it varies along the Of-axis, because the soil temperature of deserts has a strong diurnal gradient in the vertical direction).

  2. (2) The flow domain $G_{z }$ does not vary in the Of direction (in reality, water evaporates into the pore air and $G_{z }$ dwindles from the capillary fringe to the topsoil).

  3. (3) The water pressure gradient is constant (in reality, it is not because the decrease of $r_{1}$ and $r_{2}$ ) in the Of direction causes variation of this gradient.

  4. (4) The pores are cylindrical and flow is two-dimensional (in reality, the pores are not cylindrical and flow is microscopically three-dimensional).

  5. (5) The contact angle is $90^{\circ }$ (in reality, most soils are either hydrophilic or hydrophobic).

  6. (6) In Section 2, air is immobile (in reality, both water and air move, and not always co-currently).

  7. (7) Water “bridges” are assumed to be stable to potential perturbations (in reality, the menisci in Figure 1 may break into Avery $'$ anov-type films that simply “coat” the wall of the polygonal pore, which gives the case of flow considered by Kacimov et al. [Reference Kacimov, Maklakov, Kayumov and Al-Futaisi22]).

  8. (8) Gravity is ignored, that is, capillarity dominates the fluid behaviour, along with the hydrodynamic pressure gradient along Of (in thick vadose zones, the gravity force is explicitly counted on the right-hand side of the Poisson equation).

  9. (9) Menisci are circular arcs (in reality, they are not [Reference Finn15]).

Acknowledgements

The comments/critique of two anonymous referees are appreciated. A. Kacimov is supported by projects DR∖RG∖17; IG/AGR/SWAE/22/02, IG/VC/WRC/21/01, Oman. Y. Obnosov and A. Kacimov are supported by the Russian Scientific Foundation, interdisciplinary project no. 23-64-10002.

References

Abramowitz, M. and Stegun, I. A., Handbook of mathematical functions with formulas, graphs, and mathematical tables (National Bureau of Standards, Washington, DC, 1964).Google Scholar
Al-Battashi, M., Al-Maktoumi, A., Kacimov, A., Al-Mayahi, A., Al-Shukaili, A. and Al-Ismaily, S., “Shallow water table in arid urban zone: preliminary study at Sultan Qaboos University campus, Oman”, J. Agric. Marine Sci. 27(1) (2022) 6276; doi:10.53541/jams.vol27iss1pp62-76.Google Scholar
Allen, M. B., The mathematics of fluid flow through porous media (Wiley, Hoboken, NJ, 2021).CrossRefGoogle Scholar
Aver’yanov, S. F., “Dependence of soil water permeability on air content”, Doklady AN SSSR 69 (1949) 142144 (in Russian).Google Scholar
Aver’yanov, S. F., Combating salinization of irrigated lands (Kolos, Moscow, 1978) (in Russian).Google Scholar
Avkhadiev, F. G. and Kacimov, A. R., “Analytical solutions and estimates for microlevel flows”, J. Porous Media 8 (2005) 125148; doi:10.1615/JPorMedia.v8.i2.30.CrossRefGoogle Scholar
Avkhadiev, F. G. and Kacimov, A. R., “The Saint–Venant type isoperimetric inequalities for assessing saturated water storage in Lacunary shallow perched aquifers”, Z. Angew. Math. Mech. – J. Appl. Math. Mech. 103 (2023) Article ID: e202100069; doi:10.1002/zamm.202100069.Google Scholar
Ben-Noah, I. and Friedman, S. P., “Review and evaluation of root respiration and of natural and agricultural processes of soil aeration”, Vadose Zone J. 17 (2018) 147; doi:10.2136/vzj2017.06.0119.Google Scholar
Blunt, M. J., Multiphase flow in permeable media: a pore-scale perspective (Cambridge University Press, Cambridge, 2017); doi:10.1017/9781316145098.Google Scholar
Buchwald, V. T. and Viera, F., “Linearised evaporation from a soil of finite depth above a water table”, ANZIAM J. 39 (1998) 557576; doi:10.1017/S0334270000007803.Google Scholar
Chhabra, R., Salt-affected soils and marginal waters: global perspectives and sustainable management (Springer, Cham, 2022).Google Scholar
Daneshian, B., Habibagahi, G. and Nikooee, E., “Determination of unsaturated hydraulic conductivity of sandy soils: a new pore network approach”, Acta Geotechnica 16 (2021) 449466; doi:10.1007/s11440-020-01088-3.CrossRefGoogle Scholar
Dullien, F. A. L., Porous media: fluid transport and pore structure, 2nd edn (Academic Press, New York, 1992).Google Scholar
El-Khatib, N. A., “The relative permeabilities of an idealized model of two-phase flow in porous media”, Appl. Math. Model. 4 (1980) 463466; doi:10.1016/0307-904X(80)90179-1.Google Scholar
Finn, R., Equilibrium capillary surfaces (Springer, New York, 1986); doi:10.1007/978-1-4613-8584-4.CrossRefGoogle Scholar
Fredlund, D. G. and Rahardjo, H., Soil mechanics for unsaturated soils (Wiley, New York, 1993); doi:10.1002/9780470172759.CrossRefGoogle Scholar
Gakhov, F. D., Boundary value problems (Addison Wesley, New York, 1966).Google Scholar
Henrici, P., Applied and computational complex analysis, volume 3. Discrete Fourier analysis, Cauchy integrals, construction of conformal maps, univalent functions (Wiley, New York, 1993).Google Scholar
Kacimov, A., “Analytic solutions for quasi-3D seepage in a shallow unconfined aquifer as a plane composed of a transpiration-inducing park and its hydraulically commingled exterior”, J. Hydrology 626 (2023) Article ID: 130302; doi:10.1016/j.jhydrol.2023.130302.Google Scholar
Kacimov, A., Al-Maktoumi, A., Al-Ismaily, S., Al-Mayahi, A., Al-Shukaili, A., Obnosov, Yu. and Osman, A., “Water table rise in arid urban area due to evaporation impedance and its mitigation by intelligently designed capillary chimney siphons”, Environ. Earth Sci. 80 (2021) 117; doi:10.1007/s12665-021-09857-3.Google Scholar
Kacimov, A., Yakimov, N. D. and Šimůnek, J., “Phreatic seepage flow through an earth dam with an impeding strip”, Comput. Geosci. 24 (2020) 1735; doi:10.1007/s10596-019-09879-8.Google Scholar
Kacimov, A. R., Maklakov, D. V., Kayumov, I. R. and Al-Futaisi, A., “Free surface flow in a microfluidic corner and in an unconfined aquifer with accretion: the Signorini and Saint–Venant analytical techniques revisited”, Transp. Porous Media 116 (2017) 115142; doi:10.1007/s11242-016-0767-y.CrossRefGoogle Scholar
Kovda, V. A., Fundamental theory on soils. General theory of Pedogenic process (Nauka, Moscow, 1973) (in Russian).Google Scholar
Philip, J. R., “Flow in porous media”, Annu. Rev. Fluid Mech. 2 (1970) 177204; doi:10.1146/annurev.fl.02.010170.001141.Google Scholar
Philip, J. R., “Flows satisfying mixed no-slip and no-shear conditions”, J. Appl. Math. Phys. (ZAMP) 23 (1972) 353372; doi:10.1007/BF01595477.CrossRefGoogle Scholar
Polubarinova-Kochina, P. Y., Theory of ground water movement (Princeton University Press, Princeton, NJ, 1962), (Second edition of the book (in Russian) published in 1977, Nauka, Moscow).Google Scholar
Pólya, G. and Szegö, G., Isoperimetric inequalities in mathematical physics (Princeton University Press, Princeton, NJ, 1951).Google Scholar
Radcliffe, D. E. and Šimůnek, J., Soil physics with HYDRUS: modeling and applications (CRC Press, Boca Raton, FL, 2018); doi:10.1201/9781315275666.CrossRefGoogle Scholar
Salugin, A. N., “Modeling the evaporation process”, Arid Ecosystems 13 (2023) 133139; doi:10.1134/S2079096123020129.Google Scholar
Scanlon, B. R., Nicot, J. P. and Massmann, J. W., “Soil gas movement in unsaturated systems”, in: Soil physics companion (ed. Warrick, A. W.) (CRC Press, Boca Raton, FL, 2022) 297341.Google Scholar
Shah, R. K. and London, A. L., Laminar flow forced convection in Ducts: a source book for compact heat exchanger analytical data (Academic Press, New York, 1978).Google Scholar
Sparrow, E. M., Chen, T. S. and Jonsson, V. K., “Laminar flow and pressure drop in internally finned annular ducts”, Inter. J. Heat Mass Transfer 7 (1964) 583585; doi:10.1016/0017-9310(64)90056-0.Google Scholar
Thorburn, P. J., “Land management impacts on evaporation from shallow, saline water tables”, in: Subsurface hydrological responses to land cover and land use changes (ed. M. Taniguchi) (Springer, New York, 1997) 2134; doi:10.1007/978-1-4615-6141-5.Google Scholar
Timoshenko, S. P. and Goodier, J. C., Theory of elasticity (McGraw-Hill, New York, 1970).Google Scholar
Weir, G. J., “Linearised evaporation about a shallow half-plane pond”, ANZIAM J. 34 (1993) 355367; doi:10.1017/S0334270000008948.Google Scholar
Wolfram, S., Mathematica. A system for doing mathematics by computer (Addison-Wesley, Redwood City, CA, 1991).Google Scholar
Figure 0

Figure 1 (a) Cross-section of an arbitrary triangular pore. Three thin mobile water “bridges” are sandwiched between immobile air. (b) 3D diagram with zoomed menisci $CNB$ and $DMA$ near one pore corner. (c) Boundary conditions in the flow domain, a circular tetragon $G_{z}$.

Figure 1

Figure 2 (a) Rectangle $G_{z_1}$. (b) Reference half-plane $G_{\zeta }$.

Figure 2

Table 1 Dimensionless flow rate $Q^{*}$ computed using equations (2.15) and (2.16) for tetragons $G_{z}$ having $\alpha =\pi /3$ and five values of $r^{*}$.

Figure 3

Figure 3 (a) Flow of water at high $S_{w}$ in a cylindrical pore channel, the cross-section of which is an equilateral triangle with immobile air pockets entrapped near the corners. (b) Flow of air at low $S_{w}$ in a cylindrical pore channel, the cross-section of which is an equilateral triangle with immobile water pockets entrapped near the corners.