1. Introduction
Large fractions of the Greenland and Antarctic ice sheets are covered in porous firn (Verjans and others, Reference Verjans2021; Noël and others, Reference Noël, Lenaerts, Lipscomb, Thayer-Calder and van den Broeke2022). Increasingly, this firn is experiencing periods of surface melting and even rainfall and infiltration of (melt) water (Van Angelen and others, Reference Van Angelen, Lenaerts, Van Den Broeke, Fettweis and Van Meijgaard2013; Bell and others, Reference Bell, Banwell, Trusel and Kingslake2018; Harper and others, Reference Harper, Saito and Humphrey2023). Refreezing of this melt within the firn has the capacity to absorb a significant fraction of this water and prevent mass loss to the ocean (Harper and others, Reference Harper, Humphrey, Pfeffer, Brown and Fettweis2012; Van Angelen and others, Reference Van Angelen, Lenaerts, Van Den Broeke, Fettweis and Van Meijgaard2013; de la Peña and others, Reference de la Pena2015). However, the increasing frequency of infiltration events had led to a rapid increase in large-scale ice layers within the firn that may buffer its water storage capacity, leading to increased lateral water flow and eventually mass loss from the ice sheet to the ocean (Pfeffer and others, Reference Pfeffer, Meier and Illangasekare1991; Van Angelen and others, Reference Van Angelen, Lenaerts, Van Den Broeke, Fettweis and Van Meijgaard2013; Noël and others, Reference Noël2017). Additionally, significant amounts of liquid water are stored in firn aquifers, which heat the surrounding firn and delay meltwater runoff (Forster and others, Reference Forster2014; Amory and others, Reference Amory2024). The critical role of firn in modulating melt runoff has led to increased interest in flow and transport processes in wet firn or firn hydrology (Amory and others, Reference Amory2024).
Of particular interest is the formation of ice layers which requires the localization of freezing in a narrow vertical interval within the firn. Localization can either occur by ponding of melt on pre-existing discontinuities within the firn (Marsh and Woo, Reference Marsh and Woo1984; Pfeffer and Humphrey, Reference Pfeffer and Humphrey1998; Wever and others, Reference Wever, Würzer, Fierz and Lehning2016; Humphrey and others, Reference Humphrey, Harper and Meierbachtol2021) or due to the rapid arrest of the wetting front after a melting event as the liquid water content declines and conduction becomes dominant to freeze melt in place (Shadab and others, Reference Shadab, Adhikari, Rutishauser, Grima and Hesse2024a). Impermeable ice layers likely form gradually and require multiple refreezing events to reduce porosity below the pore close-off. As the permeability of the ice layer decreases, melt percolation slows sufficiently for the infiltrating melt to pond and form a perched aquifer that spreads laterally. Within the perched aquifer, the melt will fully saturate the firn. To understand the evolution of ice layers and aquifers, it is therefore important to understand the interaction of melt with layers of lower permeability, the transition from unsaturated to saturated flow and the conditions that reduce ice layer permeability to zero. Currently, there is no physics-based model that can describe the formation of an ice layer with zero permeability (hereafter referred to as an impermeable ice layer). Here we aim to develop a first-order model that captures this process, based on a theoretical framework for the coupling of energy and mass transport in firn.
1.1. Modeling melt infiltration into firn
Modeling the infiltration of water from both rainfall and surface melting into firn is still a challenging problem and model predictions diverge due to different approaches to representing melt percolation in firn and poorly characterized constitutive relations (Stevens and others, Reference Stevens2020; Vandecrux and others, Reference Vandecrux2020; Amory and others, Reference Amory2024). Models for firn hydrology must include both mass and energy transport and incorporate both the nonlinearity inherent in unsaturated flow in porous media and the nonlinearity of the phase change during freezing and melting.
The most commonly used approach to unsaturated flow in firn, so-called bucket models, combine mass conservation with a discrete percolation model, based on the concept that the melt needs to reach a maximum holding capacity or irreducible water content before it advances to the next layer of the firn (Coléou and Lesaffre, Reference Coléou and Lesaffre1998; Bartelt and Lehning, Reference Bartelt and Lehning2002; Ligtenberg and others, Reference Ligtenberg, Helsen and van den Broeke2011; Vionnet and others, Reference Vionnet2012; Kuipers Munneke and others, Reference Kuipers Munneke, Ligtenberg, Suder and Van Den Broeke2015; Verjans and others, Reference Verjans, Leeson, Stevens, MacFerrin, Noël and van den Broeke2019). A second approach combines mass balance with the two-phase extension of Darcy’s law to obtain a kinematic wave model (Colbeck, Reference Colbeck1974a; Jordan, Reference Jordan1991; Singh, Reference Singh1997; Clark and others, Reference Clark, Nijssen and Luce2017; Shadab and others, Reference Shadab, Adhikari, Rutishauser, Grima and Hesse2024a). This approach can be extended to include the effects of capillary suction leading to Richards’ equation (Illangasekare and others, Reference Illangasekare, Walter, Meier and Pfeffer1990; Wever and others, Reference Wever, Fierz, Mitterer, Hirashima and Lehning2014; Meyer and Hewitt, Reference Meyer and Hewitt2017). More complex and multidimensional models have been developed to capture the process of preferential flow and ice pipe formation in snow and firn (Marsh and Woo, Reference Marsh and Woo1984; Schneebeli, Reference Schneebeli1995; Katsushima and others, Reference Katsushima, Yamaguchi, Kumakura and Sato2013). These models build upon Richards’ equation by extending it to include either an imbibition water entry pressure (Hirashima and others, Reference Hirashima, Yamaguchi and Katsushima2014; Leroux and Pomeroy, Reference Leroux and Pomeroy2017), a dynamic capillary pressure (Hassanizadeh and others, Reference Hassanizadeh, Celia and Dahle2002; Leroux and Pomeroy, Reference Leroux and Pomeroy2019) or apparent surface tension at the wetting front (Cueto-Felgueroso and Juanes, Reference Cueto-Felgueroso and Juanes2008; Moure and others, Reference Moure, Jones, Pawlak, Meyer and Fu2023).
Similarly, phase change and the associated latent heat is handled with different approaches in firn hydrology. Most models assume thermal equilibrium between melt and ice and the main difficulty in treating the phase change is that temperature is not an independent variable at the melting point (Anderson and Crerar, Reference Anderson and Crerar1993). Percolation or bucket models are inherently discrete and handle the phase change with a discrete algorithm comparing the heat content of the melt with the cold content of the firn layer (Bartelt and Lehning, Reference Bartelt and Lehning2002; Ligtenberg and others, Reference Ligtenberg, Helsen and van den Broeke2011; Kuipers Munneke and others, Reference Kuipers Munneke, Ligtenberg, Suder and Van Den Broeke2015). Darcy models treat phase change either discretely (Illangasekare and others, Reference Illangasekare, Walter, Meier and Pfeffer1990), avoid the degeneracy by distributing the phase change over a finite temperature interval (Clark and others, Reference Clark, Nijssen and Luce2017), employ the enthalpy method (Jordan, Reference Jordan1991; Meyer and Hewitt, Reference Meyer and Hewitt2017; Shadab and others, Reference Shadab, Adhikari, Rutishauser, Grima and Hesse2024a) or allow for thermal disequilibrium (Moure and others, Reference Moure, Jones, Pawlak, Meyer and Fu2023; Jones and others, Reference Jones, Moure and Fu2024).
The coupling between energy and mass transport and the inherent non-linearities of both unsaturated flow and phase change, together with the different modeling strategies summarized above lead to wide discrepancies between predictions of different firn models (Vandecrux and others, Reference Vandecrux2020). Improving these models, therefore, requires a comprehensive and systematic evaluation strategy (Clark and others, Reference Clark, Nijssen and Luce2017). Analytic solutions to synthetic test cases are an important element of model evaluation because they allow evaluation of numerical implementation, such as the discretization and coupling strategies. In this context, kinematic wave models are useful for firn hydrology because they contain both the essential nonlinearities of unsaturated flow and phase change as well as the coupling between mass and energy transport, yet they allow for analytical solutions for simple initial conditions. Past work on kinematic waves has been restricted to unsaturated conditions, but understanding the formation of impermeable ice layers that lead to ponding and the formation of saturated regions is crucial for estimating meltwater runoff versus storage in firn. Shadab and Hesse (Reference Shadab and Hesse2022) have shown that these saturated regions can be included in an extended kinematic theory that allows us to investigate these processes. The purpose of this paper is to derive a set of self-similar analytic solutions and determine the conditions that lead to the formation of impermeable ice layers and perched aquifers.
1.2. Kinematic wave theory for firn hydrology
Due to the large porosity and grain size of firn, the vertical flow of water may be primarily governed by gravity, while capillary suction/diffusion may play a minor role (Colbeck, Reference Colbeck1972; Reference Colbeck1974b). Neglecting capillary diffusion results in a nonlinear kinematic wave model that describes the evolution of the water/melt saturation in the firn (Colbeck, Reference Colbeck1972). Kinematic wave theory is a framework used to describe the propagation of waves in systems where the wave motion is influenced primarily by the kinematics, or the motion of particles, rather than by the dynamics or forces acting on them (Lighthill and Whitham, Reference Lighthill and Whitham1955). In case of melt infiltration, the kinematic wave model results in a hyperbolic partial differential equation (PDE) that allows analytic solutions in one dimension using the method of characteristics (MOC) (Lighthill and Whitham, Reference Lighthill and Whitham1955; LeVeque, Reference LeVeque1992). These analytic solutions describe nonlinear waves, e.g. wetting and drying fronts, and their interaction that capture the main features of field observations, for example from melt infiltration in the Seward glacier firn on the St Elias mountains in Canada (Sharp, Reference Sharp1951).
Initial work on kinematic models for melt migration focused on temperate snow where phase change is not important and unsaturated flow models could be adapted directly (Colbeck, 1971; Reference Colbeck1972; Colbeck and Davidson, Reference Colbeck and Davidson1973). Later, Colbeck (1976) added thermodynamics to the kinematic theory to study infiltration into cold firn that requires refreezing of the melt at the wetting front. He shows that the retardation of the wetting front is relatively minor in snow because the latent heat of fusion is large. This model has been applied to analyze the effects of an impermeable basal boundary (Colbeck, Reference Colbeck1974a), the retention of water in snow (Colbeck, 1976) and the effects of layering and heterogeneity (Colbeck, 1979; Reference Colbeck1991). The theory has also been used to estimate the permeability of snow using lysimeter data (Colbeck and Anderson, Reference Colbeck and Anderson1982). Singh and others Reference Singh, Bengtsson and Westerstrom(1997) and Clark and others Reference Clark, Nijssen and Luce(2017) use kinematic theory to study the interaction of drying and wetting fronts which was recently identified as key to the initiation of ice layer formation (Shadab and others, Reference Shadab, Adhikari, Rutishauser, Grima and Hesse2024a). Singh and others Reference Singh, Bengtsson and Westerstrom(1997) also investigate the effect of temporal variability in rainfall rate (for review see also Singh, Reference Singh1997).
1.3. Contribution of this manuscript
To understand the formation of impermeable ice layers, it is important to model variably saturated flows, i.e. flows that transition from unsaturated to saturated and vice versa. The conventional kinematic wave theory for infiltration fails in a fully saturated region because the flow of water becomes pressure driven rather than gravity driven and thus, the model equations are not valid anymore (see Shadab and Hesse, Reference Shadab and Hesse2022). Therefore, the kinematic wave theory needs an extension to capture fully saturated regions. Recently, Shadab and Hesse (Reference Shadab and Hesse2022) showed that saturated regions can be incorporated into an extended kinematic theory for simple problems, such as a step change in firn porosity. This allows the analysis of a rising perched water table, a pre-requisite for the formation of impermeable ice layers.
To address the dynamics of ice layer formation and to provide additional analytic solutions for the evaluation of the models in firn hydrology this manuscript makes the following contributions:
∘ Formulate coupled mass and energy transport as a system of nonlinear hyperbolic conservation equations and analyze their coupling.
∘ Use MOC to develop a set of self-similar analytic solutions for problems with an initial step change in volume fractions of ice or water (Riemann problems).
∘ Apply unified kinematic wave theory to firn hydrology to describe the formation of perched aquifers and derive conditions for impermeable ice layer formation.
We call it a unified theory, as it is not derived from prior kinematic wave theories for melt infiltration, but it still unifies most previous work and includes extensions to perched aquifers. The remainder of this paper is divided into four sections. Section 2 presents the model formulation. Section 3 considers the problem of melt transport across a discontinuity and documents 12 nature-inspired cases with their analytical solutions. Section 4 applies and validates this theory to study a multilayered firn leading to formation of a perched firn aquifer and discusses the general limitations of this theory. Finally, Section 5 concludes the paper.
2. Continuum model formulation
In this section, we first define the conserved quantities, then introduce the governing equations and constitutive models and finally provide the resulting dimensionless continuum model. The related assumptions will be introduced in this work when required.
2.1. Conserved quantities
We model firn as a three-phase system comprising liquid water (w), ice (i) and gas (g). These three phases are composed of two components, namely water (H2O) and air (∼Nitrogen gas, N2). The water component partitions into the liquid and ice, while the air component is confined to the gas phase. The first conserved variable is the water (
$\mathrm{H}_2\mathrm{O}$) composition (kg/m3), C, defined as the total mass of water component per unit representative elemental volume (REV) as

where ρα is the density (kg/m3) and ϕα refers to volume fraction of the phase
$\alpha \in \{w,i,g \}$. The formulation assumes that the water vapor component is negligible in the gas phase and that ice and water phases are pure. Assuming the same density for ice and water,
$\rho_i\approx \rho_w = \rho$, and the volume fraction constraint,
$\phi_i+\phi_w+\phi_g=1$, Eqn (1) further simplifies to

The second conserved variable is related to the thermodynamics of the system. Since phase change is involved when ice melts, temperature does not accurately represent all states of the system considered (Jordan, Reference Jordan1991; Alexiades and Solomon, Reference Alexiades and Solomon1993; Aschwanden and others, Reference Aschwanden, Bueler, Khroulev and Blatter2012; Carnahan and others, Reference Carnahan, Wolfenbarger, Jordan and Hesse2021). Therefore, the second conserved variable is chosen to be the enthalpy of the system (J/m3), H, which is defined as

where hα is the specific enthalpy (J/kg) of the phase α, which is a piecewise linear function of temperature (K), T. For simplicity, we fix the reference enthalpy at the solidus to be H = 0 where the system is at the melting temperature,
$T=T_m$. The specific enthalpy hα of each phase
$\alpha \in \{w,i,g\}$ can then be defined as



Here
$c_{p,\alpha}$ is the specific heat capacity at constant pressure (J/kg·K) for phase α, Tm is the melting temperature (K) and L is the latent heat of fusion of water (J/kg). The density and specific heat capacity of gas are much lower than those of liquid water or ice, i.e.
$\rho_g \ll \rho$ and
$c_{p,g} \lt c_{p,i}$ or
$c_{p,w}$ (see Table 1). Hence, we make the simplification that the gas phase contribution to the total enthalpy of the system is negligible. After substituting Eqns (4) and (5) into Eqn (3), H can be ultimately formulated as

Table 1. A summary of simplified thermodynamic properties as well as flow properties of water in porous ice used in present work

Here the maximum enthalpy limit for the three-phase region is the product of composition, C, and the latent heat of fusion, L because it is the enthalpy of the fully molten system at the melting point (see Fig. 1a). The boundaries of the three-phase region are not included in the region
$0 \lt H \lt CL$ because it strictly refers to the three-phase region. From the above formulation, we classify three regions, where region 1 (
$H \leqslant 0$) is comprised of ice and gas, region 2 (
$0 \lt H \lt CL$) contains all three phases and region 3 (
$H \geqslant CL$) corresponds to a no-matrix state consisting of only water and gas phases.

Figure 1. The dependence of temperature and volume fractions on dimensional and dimensionless enthalpy and composition, (
$C,H$) and (
$\mathcal{C},\mathcal{H}$), respectively. Dimensional
$C,H$: (a) temperature and volume fractions of (b) water, (c) ice and (d) gas phases. Dimensionless
$\mathcal{C},\mathcal{H}$: (e) scaled temperature and volume fractions of (f) water, (g) ice and (h) gas phases. The contours are restricted to
$T\in[-100^\circ\textrm{C},100^\circ\textrm{C}]$ to avoid phase change at boiling as well as keep the contour levels consistent. Solid black lines are the level sets, whereas the dashed lines show the boundaries of the regions, i.e. H = 0 or
$\mathcal{H}=0$ and H = CL or
$\mathcal{H}=\mathcal{C}$. The three regions are labeled in panels a and e.
The temperature and volume fractions of water, ice and gas phases can be evaluated from composition, C, and enthalpy, H, as shown in Fig. 1a–d, respectively. As shown in Eqn (2), the volume fraction of gas, ϕg, only depends on the composition, C, as shown in Fig. 1d. Next we formulate the governing equations for this model corresponding to the two conserved variables.
2.2. Transport model
In a reference frame moving with ice, the conservation equations for water composition and system enthalpy are, respectively, given as


where q is the volumetric flux of water phase (m3/m
$^2\cdot$s) relative to ice phase. The effective thermal conductivity of the mixture
$\overline{\kappa}$, typically a weighted average of the thermal conductivities of the phases, κα (W/m·K).
2.3. Constitutive relations
The volumetric flux of water relative to ice, q, can be written using extended Darcy’s law,

where
$\mathrm{k}$ is the absolute permeability (m2) which is a function of porosity φ
$(\varphi=\phi_w+\phi_g =1-\phi_i)$, the ratio of void volume to the bulk volume, p is water pressure (Pa), µ is the viscosity of water (Pa·s) and g is the acceleration due to gravity vector (m2/s). The relative permeability for multiphase flow, krw, display complex hysteresis (Blunt, Reference Blunt2017), but here we only consider the simplest case with power law dependence. The relative permeability of the water phase, krw, is a function of the water saturation, s, which is the ratio of water phase volume to void volume,
$s= \phi_w/1-\phi_i$. We assume that the water phase becomes immobile below a certain residual water saturation,
$s_{w r}$. Similarly, the gas phase becomes immobile below the residual gas saturation, sgr. As a result, the two-phase fluid flow of both gas and water phases is restricted to regions where
$s_{wr} \lt s \lt 1-s_{gr}$. We will refer to regions with
$s=1-s_{gr}$ as saturated in the remainder of this paper. The residual water saturation during drainage has been determined to be
$s_{w r}\sim0.07$ from lysimeter (Colbeck, 1976) and calorimeter (Coléou and Lesaffre, Reference Coléou and Lesaffre1998) techniques. However, as the ice is water-wet with a near-zero contact angle at the ice–water–air interface (Knight, Reference Knight1971), the residual water saturation during saturation rise (imbibition) is zero due to hysteresis in the relative permeability–capillary pressure curve (Carlson, Reference Carlson1981; Blunt, Reference Blunt2017). The phenomenon of hysteresis has largely been neglected in the firn hydrology literature but it will affect the speeds of the meltwater fronts.
Next we assume the problem is gravity dominated in unsaturated regions (Colbeck, Reference Colbeck1972), such that the spatial variations in the difference between the water and air pressure (capillary pressure) are negligible at the problem length scales. See Smith (Reference Smith1983), Shadab and Hesse (Reference Shadab and Hesse2022, Reference Shadab and Hesse2024) for a more detailed discussion on neglecting the capillary pressure term in the context of soils using scaling analysis. As a result, the pressure of the water phase in the unsaturated regions becomes a constant, equal to the reference gas pressure, i.e. p = 0 (Colbeck, Reference Colbeck1972; Shadab and Hesse, Reference Shadab and Hesse2022). Plugging it in Eqn (10) eliminates the diffusive, pressure term. The volumetric flux of water, q, finally takes the gravity-driven form

The absolute permeability of ice (m2),
$\mathrm{k}$, and the relative permeability of water, krw, are assumed to be power laws (Kozeny, Reference Kozeny1927; Carman, 1937; Brooks and Corey, Reference Brooks and Corey1964; Bear, Reference Bear2013; Meyer and Hewitt, Reference Meyer and Hewitt2017) defined as


where
$\mathrm{k}_0$ is a model constant, which can be considered as an absolute permeability (m2) when there is no ice matrix, and
$k^0_{rw}$ is end point relative permeability of water phase. Here we have also assumed that the residual saturations of both water and gas phases are zero, i.e.
$s_{wr}=s_{gr}=0$. It will provide accurate speeds for the wetting fronts moving into dry firn, due to hysteresis in the relative permeability. Plugging Eqns (12) and (13) in Eqn (11) finally gives

where the acceleration due to gravity vector is
$\textbf{g}=g\hat{\textbf{g}}$ with
$\hat{\textbf{g}}$ being the unit vector in the direction of gravity. The symbol
$K_h = \frac{\mathrm{k}_0 k^0_{rw}}{\mu}\rho g$ is a known constant which can be considered as the maximum gravity-dependent volumetric flux of water,
$|\textbf{q}|$, at unity porosity. Note that the dynamics at unity porosity is not Darcy-type as the constitutive relationships (10–12) are only valid for a porous medium. Therefore, we will restrict our analysis to the porous media where φ < 1 and the flow is laminar, thus, Darcy’s law is applicable (Tek, Reference Tek1957).
2.4. Scaling
We nondimensionalize the model to make it scale-independent and find dominant terms governing the physics of the problem. The model is scaled using dimensionless variables for composition,
$\mathcal{C}$, enthalpy,
$\mathcal{H}$, temperature,
$\mathcal{T}$, depth, ζ and time, τ, which are definedFootnote 1 as

Here the spatial coordinates (e.g. the depth coordinate z) are nondimensionalized by length scale of heterogeneity or the REV scale of the problem, δ. Time variable is scaled by the shortest time of water seepage across the characteristic length through a medium with unity porosity, i.e.
$\delta /K_h$. The definitions of conserved quantities C and H, given in Eqns (2) and (7), respectively, thus transform into the dimensionless forms


where Ste is the Stefan number defined as ratio of sensible heat of water at melting temperature to the latent heat of fusion of
$\mathrm{H}_2\mathrm{O}$, i.e. Ste
$=c_{p,w} T_m / L$ and
$\mathfrak{c}_{p,r}=c_{p,i}/c_{p,w}$ is the ratio of specific heat of ice to that of water. From the formulations of dimensionless enthalpy (17) and dimensional-specific enthalpy of water phase (5), the dimensionless temperature,
$\mathcal{T}$, and dimensionless specific enthalpy of water phase,
$\mathfrak{h}_w=h_w/L$, can be derived as

Subsequently, the volume fractions of the phases, ϕα, and the porosity of the medium, φ, can be rewritten as functions of
$\mathcal{C}$ and
$\mathcal{H}$ as


The scaled temperature and volume fractions of water, ice and gas phases can be evaluated from dimensionless composition,
$\mathcal{C}$, and dimensionless enthalpy,
$\mathcal{H}$, as shown in Fig. 1e–h, respectively. As shown in Eqn (2), the volume fraction of gas, ϕg, only depends on the dimensionless composition as illustrated in Fig. 1h.
The composition and enthalpy transport equations (8 and 9) thus take the dimensionless form


Here κw is the thermal conductivity of the water phase. The ratio of heat convected to heat diffused is defined as the Peclet number for enthalpy equation,
$Pe_{\mathcal{H}}=K_h \delta/\alpha_T$, where
$\alpha_T={\kappa_w}/{\rho c_{p,w}}$ (m2/s) is the thermal diffusivity of water. Moreover, the divergence and gradient operators are now scaled with inverse of characteristic depth,
$1/\delta$.
While both conduction and advection can be important heat transport processes in firn (Shadab and others, Reference Shadab, Adhikari, Rutishauser, Grima and Hesse2024a), conduction does not affect the steady propagation of wetting and drying fronts analyzed here. This can be shown using thermodynamic and fluid flow parameters from Table 1. The value of Peclet number,
$Pe_{\mathcal{H}}$, comes out to be
$4\cdot 10^{3}$ for δ = 1 m. The ratio
$\overline{\kappa}/\kappa_w \leqslant 1$ as its maximum value of unity is achieved when the REV only contains liquid water. The Stefan number Ste is a constant of value 3.43. Therefore, the value of
$\frac{\overline{\kappa}}{\kappa_w}\frac{\text{Ste}}{Pe_{\mathcal{H}}}$ indicates about three orders of magnitude higher heat advection compared to heat conduction for gravity-driven infiltration in firn. Therefore, we can neglect the second-order heat conduction term in Eqn (22), which is also the necessary condition in the three-phase region (
$0 \lt \mathcal{H} \lt \mathcal{C}$ or
$\mathcal{T}=0$) as
$\nabla \mathcal{T} = 0$. Assuming local thermodynamic equilibrium, in the limit
$Pe_{\mathcal{H}}\to \infty$, the system of dimensionless governing equations (21 and 22) then reduces to quasi-linear system of coupled hyperbolic equations,

where
$\textbf{u} = [\mathcal{C},\mathcal{H} ]^T$ is the vector of dimensionless conserved variables and
${\boldsymbol{f}}(\textbf{u}) = [ {\boldsymbol{f}}_{\mathcal{C}},{\boldsymbol{f}}_{\mathcal{H}}]^T$ is the vector of their corresponding nonlinear flux vectors. Here the flux vector functions for the dimensionless composition and enthalpy are given as

The above analysis shows the distinct system behaviors in the different regions based on fluxes (see Fig. 2). In this work, region 3 with no solid matrix is not considered since the constitutive relation for volumetric flux (Darcy’s law) is not valid anymore. From Eqn (24), it can be observed that the fluxes of enthalpy and composition in the system of governing equation (23) are identical in regions 1 and 2 defined by the symbol
${\boldsymbol{f}}$ for brevity. This is the result of scaling owing to the fact that the composition changes only when water infiltrates or convects while carrying the enthalpy in the form of latent heat (and specific heat) along with it.

Figure 2. The dimensionless flux of composition or enthalpy in
$\mathcal{C}\mathcal{H}$ phase space for m = 3 and n = 2. In region 1 consisting of water and gas region (
$\mathcal{H} \leqslant 0$) as well as region 3 comprising of three-phase region (
$0 \lt \mathcal{H} \lt \mathcal{C}$), the flux of dimensionless enthalpy and composition are identical, i.e.
$f_{\mathcal{C}}=f_{\mathcal{H}}$. Region 3 with water and gas (
$\mathcal{H}\geqslant \mathcal{C}$) is not considered in the present work.
In the next section, we will consider a simple problem of melt transport across a discontinuity to utilize the MOC for solving the system of hyperbolic PDEs (Lighthill and Whitham, Reference Lighthill and Whitham1955; LeVeque, Reference LeVeque1992) given in Eqn (23).
3. Melt transport across a discontinuity in firn
This section considers the reaction front arising from melt flow across a discontinuity in dimensionless composition and enthalpy at a depth, say ζ = 0, as shown in Fig. 3. The dynamics of such problems can be understood using hyperbolic analysis of the coupled system of PDEs in one dimension (Lighthill and Whitham, Reference Lighthill and Whitham1955; LeVeque, Reference LeVeque1992; Venkatraman and others, Reference Venkatraman, Hesse, Lake and Johns2014; Jordan and Hesse, Reference Jordan and Hesse2015; Ghaderi Zefreh and others, Reference Ghaderi Zefreh, Doster and Hesse2019).

Figure 3. One-dimensional Riemann problem: (a) Schematic representation of u across a discontinuity within or at the boundary of a porous firn. Initial conditions for the Riemann problem for conserved variables (b)
$\mathcal{C}$ and (c)
$\mathcal{H}$ plotted against dimensionless depth coordinate, ζ.
3.1. General structure of reaction fronts
Consider the following one-dimensional initial value problem with two constant states, known as a Riemann problem. See LeVeque Reference LeVeque1992 for a pedagogical introduction to Riemann problems and their analysis. Let the spatial dimension be the direction of gravity,
$\hat{\textbf{g}}$, which aligns with the (dimensionless) depth coordinate, z (ζ). In that case, the dimensionless flux vector (24) reduces to

The system of dimensionless composition and enthalpy conservation equations (23) can be written in one-dimensional depth coordinates, ζ, as

with initial conditions

where the flux vector in ζ direction is
$\textbf{f}(\textbf{u})=[f,f]^T$ and the subscripts τ and ζ refer to the partial derivatives with respect to the dimensionless time and depth, respectively. The subscripts l and r refer to the state of the system on the left and right sides of a discontinuity. During melt infiltration into firn the left (right) state corresponds to the top (bottom) layer around a discontinuity. An example of an initial discontinuity with left state, ul, and right state, ur, is shown in Fig. 3b and c. The flow is toward the direction of gravity, assumed to be in
$+ \zeta$-direction. The solution to the Riemann problem for well-behaved systems of two coupled nonlinear PDEs is characterized by the formation of an intermediate state, ui, bounded by two waves
$\mathscr{W}_1$ and
$\mathscr{W}_2$ (LeVeque, Reference LeVeque1992). This solution structure, observed in Fig. 4, can be represented as


Figure 4. Solution of the Riemann problem introduced in Figure 3. (a) Evolution of dimensionless composition,
$\mathcal{C}$, in space, ζ, and time, τ. (b) The same self-similar solution plotted as a function of similarity variable
$\eta=\zeta/\tau$.
In the context of reactive meltwater transport, the waves
$\mathscr{W}_1$ and
$\mathscr{W}_2$ are the reaction fronts and the intermediate state, ui, corresponds to a state between the fronts. The system (26) can be recast into a quasilinear form using chain rule as

where
$\nabla_{\textbf{u}} \textbf{f}(\textbf{u})$ is the gradient of flux,
$\textbf{f}(\textbf{u})$, with respect to the conserved variables, u, which takes the matrix form

where the subscripts
$,\mathcal{C}$ and
$,\mathcal{H}$ refer to the partial derivatives with respect to dimensionless composition and enthalpy, respectively. The derivatives of the flux gradient above can be evaluated explicitly, and which are given in Appendix A. The system of advection equations (29) results in waves (fronts) propagating with their characteristic velocities. These fronts have self-similar stretching patterns because of their own characteristic velocities given by the flux gradient.
3.2. Self-similarity of reaction fronts
The recognition of the constant stretching morphology of the reaction fronts from an initial step change allows the introduction of the similarity variable

Physically, η describes the dimensionless propagation velocity of the reaction front. The solution generally collapses into a single profile when plotted as a function of η (see Fig. 4b). Therefore, the system of PDEs (29) can be transformed into a system of ordinary differential equations by considering the nonlinear eigenvalue problem

where the flux gradient is
$\textbf{A}=\nabla_{\textbf{u}} \textbf{f}(\textbf{u})$ and the eigenvector is
$\textbf{r}_p=\mathrm{d} \textbf{u}/\mathrm{d} \eta$ corresponding to the eigenvalue λp. Here the eigenvalues λ 1 and λ 2 are the characteristic propagation speeds of the waves
$\mathscr{W}_1$ and
$\mathscr{W}_2$, respectively. The associated eigenvectors,
$\textbf{r}_p=\mathrm{d} \textbf{u}/\mathrm{d} \eta$, give the pathways through the
$\mathcal{C}-\mathcal{H}$ plane, also referred to as the hodograph plane, that satisfy the conservation equations (see Fig. 5b, e.g.).
Constant solutions of the conservation laws (29) satisfy (32) trivially because
$\mathrm{d} \textbf{u}/\mathrm{d} \eta=0$. Solutions of the conservation laws (29) that vary continuously must instead satisfy the eigenvalue problem (32). Finally, discontinuities in the solution of (29) must satisfy the Rankine–Hugoniot (R–H) jump condition (LeVeque, Reference LeVeque1992), which is derived from the discrete conservation of mass and enthalpy around the discontinuity, and given by

where
$\Lambda_{\mathscr{S}}$ is the shock speed,
$[\,\cdot\,]$ refers to the jump condition across the shock and subscripts + and − refer to the state on the left and right sides of a shock wave. Note that the left (l) and right (r) states might not necessarily be the left (−) and right (+) sides of a shock front, due to the presence of an intermediate state.
3.3. Construction of the solution in the
$\mathcal{C}-\mathcal{H} $ hodograph plane
The self-similar solutions are constructed by identifying directions in the
$\mathcal{C}-\mathcal{H}$ hodograph plane that satisfy conservation laws and the equation of state. One such direction allows a continuous variation in u, which can be found by integrating the eigenvectors of the flux gradient. Another set of directions is determined by the nonlinear algebraic system of equations arising from the R–H jump condition (33), described by shock fronts. First, we consider a system where both left state,
$\textbf{u}_l=[\mathcal{C}_l,\mathcal{H}_l]^T$, and right state,
$\textbf{u}_r=[\mathcal{C}_r,\mathcal{H}_r]^T$, reside in the same region. Then we investigate more complicated cases where left and right states can lie in different regions. Lastly we discuss the cases where a fully saturated region forms, leading to the failure of the current hyperbolic PDE analysis. This theory is then further extended to analyze the formation and evolution of fully saturated regions which are governed by a different, elliptic PDE (Shadab and Hesse, Reference Shadab and Hesse2022). Although there can be at most one moving wave for simple cases where the medium remains unsaturated (
$C \lt \rho$,
$\mathcal{C} \lt 1$ or
$\phi_g \gt 0$), there can be two moving waves when a fully saturated region appears, i.e.
$\mathcal{C}=1$. Below we enumerate 12 distinct self-similar solutions to the Riemann problem that describe a variety of firn hydrological processes and summarize them in Table 2.
Table 2. Summary of all analytic solutions presented in this paper along with related works that either studied or observed the corresponding scenario.

3.3.1. Region 2 only (three-phase region)
Region 2 (
$0 \lt \mathcal{H} \lt \mathcal{C}$) consists of all three phases, which is relevant for temperate glaciers where
$\mathcal{T}=0$ or
$T=T_m$. In the three-phase region, the eigenvalues of the flux gradient (30) are

where the subscripts
${\mathcal{C}}$ and
${\mathcal{H}}$ refer to the partial derivatives with respect to the corresponding conserved variable.

Figure 5. Contour plots of (a) second eigenvalue, λ 2, and (b) propagation velocity of second front with respect to the melt given by
$\phi_w\lambda_2$ in the
$\mathcal{C}-\mathcal{H}$ hodograph plane for m = 3 and n = 2. The slow path r1 and fast path r2 are shown with dashed and solid lines, respectively, in panel b.
The first reaction front is a stationary contact discontinuity as
$\lambda_1=0$. The eigenvalue λ 2 gives a dimensionless propagation speed of the second reaction front
$\mathscr{W}_2$ as a function of
$\mathcal{C}$ and
$\mathcal{H}$, as plotted in Fig. 5a for m = 3 and n = 2. Therefore, all solutions governed by hyperbolic PDEs (26) will have at most a single moving reaction front. Due to variable porosity, the dimensionless system is scaled with respect to the largest saturated hydraulic conductivity, Kh, which corresponds to the volumetric flux of water (Darcy’s flux) rather than the melt velocity. Hence, the propagation speed of the second reaction front relative to the melt is
$\phi_w \lambda_2$, as shown in Fig. 5b. As expected, the propagation velocity
$\phi_w \lambda_2$ decreases toward the lower boundary of the three-phase region (
$\mathcal{H}=0$). These eigenvalues yield two corresponding, linearly independent eigenvectors in the
$\mathcal{C}-\mathcal{H}$ hodograph plane given by

These eigenvectors can be used to find the integral curves using the system of ODEs

which can be further integrated to obtain the solution pathways
$\textbf{u}(\textbf{u}_0,\eta) $ as

These paths in the
$\mathcal{C}-\mathcal{H}$ hodograph plane comprise the set of states that can be connected to the initial state u0 by a reaction front with a continuous variation in u. In the three-phase region, the family of integral curves corresponding to first eigenvector r1 is referred to as slow path as
$\lambda_1 \lt \lambda_2$ and is given by

where
$\mathfrak{C}$ is the constant of integration, which can be found for the initial point u0. The slow path lines are the same as constant flux lines, as shown in Fig. 5b. Next, the family of integral curves corresponding to the second eigenvector is known as fast path and is given by

The speed of second characteristic λ 2 is nonnegative and increases monotonically in the direction of integral curves corresponding to the second eigenvector r2 (fast paths) in three-phase region for n > 1 (see Lemma A.1 in Appendix A and Fig. 5b). Additionally, the slow path corresponds to constant flux contours as
$\lambda_1=0$ and the fast path corresponds to constant porosity, φ, contours (Lemma A.2 in Appendix A).

Figure 6. The simple solutions of a Riemann problem leading to a contact discontinuity
$\mathscr{C}$ (green), rarefaction
$\mathscr{R}$ (blue) and shock waves
$\mathscr{S}$ (red). (a) Construction of solution in the hodograph plane and their corresponding self-similar analytical solutions for (b) dimensionless composition and (c) dimensionless enthalpy with dimensionless velocity η. The evolution of the volume fractions of the three phases in the system for different configurations at dimensionless times
$\tau=0,0.5,1.0$: (d–f) Case I—contact discontinuity, (g–i) Case II—drying front/rarefaction wave, and (j–l) Case III—wetting front/shock wave.
Solutions in the three-phase region
We will now discuss the different analytical solutions within the three-phase region, tailored to glaciological scenarios. All scenarios are summarized in Table 2. The discussion below assumes that u0 is the left state, ul, and describes the set of permissible right states ur. We begin with cases that lead to solutions with a single front before considering cases leading to two fronts and the formation of an intermediate state.
(a) Stationary linear reaction front (Case I)
This case resembles a steady meltwater flux into a temperate firn with a step reduction in porosity at shallow depth (Fig. 6d–f). The integral curves associated with λ 1 and r1, known as the first characteristic field (
$\lambda_1,\textbf{r}_1$), are constant flux lines in the
$\mathcal{C}-\mathcal{H}$ hodograph plane. Any right state ur, along the integral curve, connected to u0 by a stationary discontinuity is a weak solution of Equation (29). Because
$\lambda_1=0$, the first wave is a stationary contact discontinuity
$\mathscr{C}_1$. The fluxes of
$\mathcal{C}$ and
$\mathcal{H}$ on both sides are the same so that the melt transport does not change
$\mathcal{C}$ and
$\mathcal{H}$ and the front does not evolve. So, a contact discontinuity is the solution for the left and right states lying on the slow path (38) (constant flux lines) satisfying

The complete solution in this case takes the form

Fig. 6a–c (green color) illustrate an example of a system that results in a stationary contact discontinuity. This system corresponds to a steady meltwater flux of f = 0.112 inside temperate porous firn with a jump in porosity from 70% at ζ < 0 to 55.3% in ζ > 0 (coarse-to-fine transition in firn) leading to liquid water contents of 0.4 and 0.45, respectively, in these regions. These values correspond to
$\textbf{u}_l=[\mathcal{C}_l,\mathcal{H}_l]^T=[0.7,0.4]^T$ and right state
$\textbf{u}_r=[0.897,0.45]^T$. The resulting system illustrated by the volume fractions of the three phases highlights a porosity jump in the temperate firn that has a constant steady meltwater flux on both sides, as shown in Fig. 6d–f at different times.
(b) Moving nonlinear reaction front
The integral curves associated with the second characteristic field (
$\lambda_2,\textbf{r}_2$) are the constant porosity φ contours. Thus, a nonlinear characteristic wave is the solution for the left and right states lying on the fast path (39) satisfying

(i) Rarefaction wave (Case II)
This case resembles a sudden drop of meltwater influx into a temperate firn with constant porosity (Fig. 6g–i). If the characteristic speed λ 2 varies smoothly from left to right state, any right state u along the fast path is connected to left state u0 by a continuously varying saturation front (Fig. 6a, blue lines). The propagation velocity, λ 2, along these continuous reaction fronts increases monotonically such that the reaction front spreads with time. These self-smoothening drying fronts are referred to as rarefaction waves, denoted by the symbol
$\mathscr{R}$. The term drying front refers to drying due to meltwater drainage instead of refreezing (Clark and others, Reference Clark, Nijssen and Luce2017). Rarefaction waves are a weak solution of Eqn (32) if the resultant profile of u is single-valued. This condition is satisfied if u lies on the branch of the integral curve r2 emanating from u0 in the direction of increasing λ 2 (see Figs. 5 and 6a, blue lines). The analytical solution concerning the self-similar variable η for a rarefaction wave on an integral curve can be evaluated from Eqn (37), which comes out to be


The final solution in this case takes the form

where the speed of the second characteristic
$\lambda_2(\cdot)$ is evaluated from Eqn (34). An example of this case is when meltwater flux instantly drops, leading to a smoothing drainage front. Fig. 6a–c (blue line) show a moving rarefaction developed inside a 70% porous firn due to an instantaneous reduction in meltwater flux, captured by lower (40%) and higher (55%) liquid water content layers on top and bottom, respectively. These numbers translate to left (top)
$\textbf{u}_l=[0.7, 0.4]^T$ and right (bottom) states being
$\textbf{u}_r=[0.85,0.55]^T$. The resulting evolution of volume fractions is shown in Fig. 6g–i. The leading edge of the rarefaction front moves faster than the trailing edge connected by a gradual, linear smoothening. In this case, it is a linear profile (straight line) as the power law exponents in Eqn (45) are m = 3 and n = 2. This drying/rarefaction front has been studied by Clark and others Reference Clark, Nijssen and Luce(2017) and was observed in models studying the Dye-2 site in Greenland on 12 August 2016 after the meltwater flux ceased (Samimi and others, Reference Samimi, Marshall and MacFerrin2020; Vandecrux and others, Reference Vandecrux2020; Colliander and others, Reference Colliander2022; Shadab and others, Reference Shadab, Adhikari, Rutishauser, Grima and Hesse2024a).
(ii) Shock wave (Case III)
This case resembles a sudden increase of meltwater influx into a temperate firn with constant porosity (Fig. 6j–l). If the right state u lies on the opposite branch of the integral curve (shown by red line in Fig. 6a, e.g.), a continuous reaction front would result in unphysical solutions as the characteristics will cross each other. Therefore, in this case, u is connected to left state u0 by a discontinuous reaction front that propagates with a velocity,
$\Lambda_{\mathscr{S}}(\textbf{u}_0,\textbf{u})$, which can be calculated from the R–H jump condition (33) using initial conditions (
$\textbf{u}_l,\textbf{u}_r$). Such fronts are referred to as wetting/shock fronts, denoted by the symbol
$\mathscr{S}$. The set of permissible right states u that can be connected to the left state u0 by shocks lie on the segment of the Hugoniot-locus that satisfies the entropy condition. In the three-phase region for the system of equations considered, the Hugoniot-locus is the same as the integral curve, which is found to be the fast path r2 (39) from the Hugoniot jump condition (33), since the flux of enthalpy and composition are the same. The dimensionless velocity of the shock from Eqn (33) is then

where the subscript − refers to the left (top) state and + is the right (bottom) state for this particular configuration where
$f(\textbf{u}_l) \gt f(\textbf{u}_r)$. The final solution thus takes the form

When meltwater flux instantly increases, it leads to a sharp wetting front propagating in the direction of gravity. Fig. 6j–k show a moving shock front developed inside a 70% porous firn due to an instantaneous increase in meltwater flux, captured by higher (40%) and lower (20%) liquid water content ϕw in top and bottom layers, respectively. These numbers translate to left (top) state being
$\textbf{u}_l=[0.7, 0.4]^T$ and right (bottom) state being
$\textbf{u}_r=[0.5,0.2]^T$ sketched analytically in Fig. 6a–c. The wetting front has been discussed (Colbeck, Reference Colbeck1972; Humphrey and others, Reference Humphrey, Harper and Pfeffer2012; Clark and others, Reference Clark, Nijssen and Luce2017; Meyer and Hewitt, Reference Meyer and Hewitt2017, e.g.) and observed in models (Vandecrux and others, Reference Vandecrux2020; Samimi and others, Reference Samimi, Marshall, Vandecrux and MacFerrin2021; Colliander and others, Reference Colliander2022, e.g.) and field observations including at the Dye-2 site in Greenland on 9 August 2016 when meltwater percolates in temperate firn (Heilig and others, Reference Heilig, Eisen, MacFerrin, Tedesco and Fettweis2018; Samimi and others, Reference Samimi, Marshall and MacFerrin2020).
(c) Two fronts with an intermediate state
The solution profile contains a single reaction front if ul and ur share the same integral curve or Hugoniot-locus. In all other cases in the three-phase region without complete saturation, a different state than left or right state forms which is referred to as an intermediate state, ui, in the hodograph plane. At this intermediate state, the solution switches from the first characteristic field
$(\lambda_1,\textbf{r}_1)$ to the second
$(\lambda_2,\textbf{r}_2)$. In other words, at the intermediate state ui, the solution changes from the stationary front along constant flux lines (slow path) to the advancing reaction front along the path of constant porosity contours (fast path) that are parallel to the upper boundary of the three-phase region (
$\mathcal{H}=\mathcal{C}$). The two possible intermediate states are given by the intersections of the integral curves emanating from ul and ur (see Fig. 7a, e.g.). Only one intersection yields a physically realistic single-value solution. The correct intersection is selected by requiring that the propagation speed increases monotonically from ul and ur. A single-valued solution is ensured if and only if ul and ui lie on the slow path first and then ui is connected to ur along the fast path. The reactive melt transport system considered here only allows two solutions for this case:

because the first characteristic is linearly degenerate and the reaction front along the slow path is always a contact discontinuity
$\mathscr{C}_1$. The reactive melt transport across an initial discontinuity is characterized by the formation of a reacted zone corresponding to ui that is bounded between a stationary front
$\mathscr{C}_1$, and an advancing front that is either a rarefaction wave
$\mathscr{R}_2$ or a shock wave
$\mathscr{S}_2$. Below we will discuss these two cases.

Figure 7. Formation of an intermediate state ui or
$\textbf{u}^*_i$ for Case IV—
$\mathscr{C}_1\mathscr{R}_2$ or Case V—
$\mathscr{C}_1\mathscr{S}_2$, respectively. An asterisk is used to differentiate the two intermediate states corresponding to the two cases. (a) Construction of solution in the hodograph plane and their corresponding self-similar analytical solutions for (b) dimensionless composition and (c) dimensionless enthalpy with dimensionless velocity η. The evolution of the volume fractions of the three phases in the system at dimensionless times
$\tau=0,0.5,1.0$ for the two configurations: (d–f) Case IV—
$\mathscr{C}_1\mathscr{R}_2$ and (g–i) Case V—
$\mathscr{C}_1\mathscr{S}_2$.
(i) 1-Contact discontinuity and 2-Rarefaction (Case IV)
This case resembles a sudden drop in meltwater flux inside temperate firn where porosity also reduces with depth (Fig. 7d–f). In this case,
$f(\textbf{u}_l) \lt f(\textbf{u}_r)$ and the resulting first characteristic wave is a contact discontinuity,
$\mathscr{C}_1$, which satisfies Eqn (38) for left state ul and the intermediate state ui. The second characteristic wave is a rarefaction which is governed by Eqns (39), (43) and (44) for intermediate state ui and right state ur. Combining all these equations results in a nonlinear algebraic equation to evaluate
$\mathcal{H}_i$ that is given by

which can be re-written in terms of porosities, φ, as

Next, the composition at intermediate state,
$\mathcal{C}_i$, can be computed from Eqns (39) and (34) which corresponds to the fast path and the speed of the second characteristic, λ 2, respectively. The final solution in this case takes the form

Physically, the first contact discontinuity represents the constant flux of water entering the bottom, low porosity layer (Fig. 7e and f). The rarefaction shows the drainage of the wetter firn due to gravity. Fig. 7a shows the construction of the solution
$\mathscr{C}_1 \mathscr{R}_2$ in the blue line for the left (top) state which is more porous
$\varphi_l=80\%$ and has less water content (LWC)
$\phi_{w,l}=0.1$. The right (bottom) state is less porous
$\varphi_r=58\%$ but has more liquid water content
$\phi_{w,r}=0.528$. These values correspond to left and right states being
$\textbf{u}_l=(0.3,0.1)^T$ and
$\textbf{u}_r=(0.948,0.528)^T$ respectively shown in Fig. 7a–c. The intermediate state comes out to be
$\textbf{u}_i=(0.538,0.117)^T$ which corresponds to LWC
$\phi_{w,i}=11.7\%$ and the same porosity as the right state, i.e.
$\varphi_i=58.0\%$. Fig. 7b and c show the corresponding self-similar analytical solutions for composition and enthalpy, respectively, for this case with blue lines which only depend on the dimensionless velocity. The rarefaction moves down with a characteristic velocity that can be computed analytically. Fig. 7d–f show the resulting evolution of volume fraction of each phase at different times showing self-similar expansion of the rarefaction wave.
(ii) 1-Contact discontinuity and 2-Shock (Case V)
This case is similar to Case IV but with a sudden rise in meltwater flux (Fig. 7g–i), and as such
$f(\textbf{u}_l) \gt f(\textbf{u}_r)$. As a result, the first characteristic wave is a contact discontinuity,
$\mathscr{C}_1$, that satisfies Eqn (38) for the left state ul and intermediate state ui. Since the second characteristic lies on the Hugoniot locus, the result is a shockwave,
$\mathscr{S}_2$, which satisfies the Hugoniot-jump condition (33) for the intermediate state and right state. Combining Eqns (38) and (39) gives a simple relation for dimensionless enthalpy at intermediate state,
$\mathcal{H}_i$, to be

Then Eqn (39) for the intermediate state, ui, and right state, ur, provides the value of dimensionless composition at intermediate state,
$\mathcal{C}_i$. The final solution in this case takes the form

Physically, the first contact discontinuity represents the increased, constant flux of water entering the bottom, low porosity layer. The second shock shows the wetting front advancing the water content to dryer firn because of gravity (Fig. 7h and i). Fig. 7a shows the construction of the solution
$\mathscr{C}_1 \mathscr{S}_2$ with red line for the left state which is less porous, i.e.
$\varphi_l=58.0\%$, but has more water content (
$\phi_{w,l}=52.8\%$), and the right state is more porous and less wet corresponding to
$\varphi_r = 80\%$ and LWC
$\phi_{w,r}=10\%$. These values correspond to
$\textbf{u}_l=(0.948,0.528)^T$ and
$\textbf{u}_r=(0.3,0.1)^T$ in Fig. 7a–c. The intermediate state comes out to be
$\textbf{u}^*_i=(0.648,0.448)^T$ which corresponds to the wetter intermediate region behind the wetting front with LWC
$\phi_{w,i}=44.8\%$ but the same porosity as the right state, i.e.
$\varphi_i=80\%$ (see Fig. 7g–i). Fig. 7b and c show the corresponding self-similar analytical solutions for composition and enthalpy, respectively, for this case with red lines which only depend on the dimensionless velocity.
(d) Two fronts and a jump with two intermediate states (formation of a saturated region, Case VI)
This case resembles a temperate firn with a step decrease in porosity at depth receiving a sudden increase of meltwater influx (Fig. 8d–f), which results in ponding at the porosity contrast and a rising water table. The initial conditions of this case are similar to Case V (
$\mathscr{C}_1\mathscr{S}_2$). However, in this case, the slow path emanating from the left state does not intersect with the fast path from the right state in the three-phase region where
$\mathcal{C} \lt 1$, as shown in Fig. 8a. In other words, the bottom layer is unable to accommodate the flux of meltwater from the top layer, as discussed for soils in Shadab and Hesse (Reference Shadab and Hesse2022). If the intermediate state leads to complete saturation, i.e.
$\mathcal{C}_i=1$, then the proposed hyperbolic PDE solution framework breaks down. This happens because inside the fully saturated region, the dynamics of the water phase change from gravity-driven to pressure-driven as the governing model changes from hyperbolic (local) to elliptic (global) PDEs (see Shadab and Hesse, Reference Shadab and Hesse2022 for a detailed analysis). In this case, the dynamics becomes complicated as the complete solution cannot be directly interpreted from the hodograph plane because the flux in the saturated region may not simply be the hydraulic conductivity (24) anymore.
Nevertheless, we can still construct a full analytical solution to this problem using the extended kinematic wave approximation proposed by Shadab and Hesse (Reference Shadab and Hesse2022). In this case, the solution consists of three waves including a backfilling shock moving upwards (Fig. 8d–f), denoted by symbol
$\mathscr{S}_1^*$, a stationary jump at the initial location of the jump, denoted by
$\mathscr{J}_2$, and a downward moving shock (wetting front),
$\mathscr{S}_3$, into the less porous and temperate layer. Note that the jump
$\mathscr{J}_2$ lies in the saturated region and therefore does not represent any hyperbolic wave. Therefore, it is represented by a broken arrow in the full solution given by@@@

The solution is explained in detail as follows and illustrated in Fig. 8a–c. First, the backfilling shock on the fast path (constant porosity line) connects to the first intermediate state, i 1, which lies on the line
$\mathcal{C}=1$. Therefore, the first intermediate state variables are

This state is observed right next to the left state and can be considered as the rising perched water table in the region ζ < 0. The speed of this backfilling, upper shock in dimensionless form is again given by the R–H jump condition as

where ζU is the location of the upper shock. The shock moves upwards due to choking as the numerator is positive, because flux
$f(\textbf{u}_l)$ is larger than the time-dependent dimensionless flux in the saturated region,
$q_s(\tau)$, which is also scaled by Kh. Note that the flux in the saturated region
$q_s(\tau)$ is not the saturated hydraulic conductivity but instead, it is governed by the dynamics of the saturated region.

Figure 8. Formation of a fully saturated region in temperate firn (Case VI): (a) Construction of solution in the hodograph plane and their corresponding self-similar analytical solutions for (b) dimensionless composition and (c) dimensionless enthalpy with dimensionless velocity η. The result shown with dark blue line consists of a backfilling shock,
$\mathscr{S}^*_1$, a jump,
$\mathscr{J}_2$, and another wetting shock,
$\mathscr{S}_3$, along with two intermediate states
$\textbf{u}_{i_1}=(\mathcal{C}_{i_1},\mathcal{H}_{i_1})^T$ and
$\textbf{u}_{i_2}=(\mathcal{C}_{i_2},\mathcal{H}_{i_2})^T$. The left and right states are
$\textbf{u}_l=(0.9,0.4)^T$ and
$\textbf{u}_r=(0.8,0.1)^T$, respectively. The first and second intermediate states,
$\textbf{u}_{i_1}=(1,0.5)^T$ and
$\textbf{u}_{i_2}=(1,0.3)^T$, respectively. The evolution of the volume fractions of the three phases in the resulting system at dimensionless times (d) τ = 0, (e) 0.5, (f) 1.0.
Next, the first intermediate state i 1 is connected to the second intermediate state i 2 through a stationary jump
$\mathscr{J}_2$ at the location of initial jump at ζ = 0. Both intermediate states lie in the fully saturated region and therefore the jump
$\mathscr{J}_2$ does not represent a hyperbolic wave. The flux inside the isothermal saturated region
$q_s(\tau)$ is found to be uniform in this case (Shadab and Hesse, Reference Shadab and Hesse2022). Therefore, the flux between the two intermediate states is also same, equal to
$q_s(\tau)$. The state variables
$\mathcal{C}$ and
$\mathcal{H}$ for the second intermediate state, i 2, are provided by the right state. The second intermediate state also lies at
$\mathcal{C}=1$ on the fast path (39) (Hugoniot locus) emanating from right state, i.e.
$\mathcal{C}_r-\mathcal{H}_r=\mathcal{C}_{i_2}-\mathcal{H}_{i_2}$. Therefore, the second intermediate state variables are simply,

Similarly, the velocity of the downward-moving lower shock (wetting front) is

where ζL is the dimensionless location of the lower shock. Similar to a two-layer soil discussed in Shadab and Hesse (Reference Shadab and Hesse2022), the dimensionless flux in the saturated region is a depth-based harmonic mean of the dimensionless saturated hydraulic conductivities at the two intermediate states given by

where
$K_{i_1}$ and
$K_{i_2}$ are the dimensionless saturated hydraulic conductivities at the first and second intermediate states given by

using Eqns (54) and (56). Note that the porosities at the first and second intermediate states are the porosities of the left and right state, respectively. Solving the system of coupled ordinary differential Eqns. (57) and (55) along with the definition of flux in the saturated region (58) gives the location of the shocks and the flux in the saturated region. Similar to the case of two-layered soils in Shadab and Hesse (Reference Shadab and Hesse2022), to find the analytic value of
$q_s(\tau)$, the ratio of shock speeds can be considered a constant as an ansatz given by

where
$\mathfrak{R}$ is the constant ratio of shock speeds which is negative and τp is the dimensionless time of ponding when the upward moving shock reaches the surface. In the special case when this jump condition happens to exist at the surface,
$\mathfrak{R}=0$,
$\tau_p=0$ and
$q_s=K_{i_2}$ which is saturated hydraulic conductivity of the second intermediate state. Otherwise, by substituting equations from the shock speed ratio definition (60), shock speeds (55) and (57), and flux in the saturated region (58),
$\mathfrak{R}$ comes out to be the solution of a quadratic equation


Figure 9. Solutions when the left state lies in region 1 (ice and gas): either only a contact discontinuity appears (Case VII) or an intermediate state, ui, along with a rarefaction wave
$\mathscr{R}_2$ also forms in a
$\mathscr{C}_1\mathscr{R}_2$ fashion (Case VIII). (a) Construction of solution in the hodograph plane and their corresponding self-similar analytical solutions for (b) dimensionless composition and (c) dimensionless enthalpy with dimensionless velocity η. Blue and red lines, respectively, show the solutions when the right states are in Regions 1 and 2, respectively. The evolution of the volume fractions of the three phases in the resulting system at dimensionless times
$\tau=0,0.5,1.0$ for the two configurations: (d–f) Case VII—
$\mathscr{C}$ and (g–i) Case VIII—
$\mathscr{C}_1\mathscr{R}_2$.
Subsequently, substituting Eqns (59) and (60) in (58) gives the time-invariant dimensionless flux in the saturated region qs during
$0 \lt \tau \leqslant \tau_p$ as

The time and space invariant flux qs leads to the result using Eqns (55) and (57) that the shock speeds are constant for
$0 \lt \tau \leqslant \tau_p$. Therefore, the shock locations before ponding vary linearly with time. Lastly, the full solution for this case can be summarized as

Here the shock speeds
$\Lambda_{\mathscr{S}^*_1}$ and
$\Lambda_{\mathscr{S}_3}$ depend on both left and right states due to formation of the saturated region as the flux is governed by an elliptic PDE. In terms of firn processes, the lower (low porosity) layer is unable to accommodate the flux of water from the upper layer and therefore leads to a rising perched water table as well as a wetting front (Fig. 8e and f). In this example, the top layer has 50% porosity and 40% liquid water content (
$\phi_{w,l}$) and the bottom layer is 30% porous and has only 10% LWC. These values correspond to left and right states being
$\textbf{u}_l=(0.9,0.4)^T$ and
$\textbf{u}_r=(0.8,0.1)^T$, respectively (Fig. 8a–c). Consequently, the first and second intermediate states,
$\textbf{u}_{i_1}=(1,0.5)^T$ and
$\textbf{u}_{i_2}=(1,0.3)^T$, respectively, lie in the expanding saturated region (Fig. 8d–f). The speeds of both fronts
$\mathscr{S}_1^*$ and
$\mathscr{S}_3$ are constant before ponding occurs.
3.3.2. Region 1 only (ice and gas region), Case VII
While the previous cases consisted of temperate firn, this case resembles a cold firn with a step reduction in porosity at depth (e.g. porous firn on top of less porous firn or glacier ice), and no meltwater influx, resulting in a dry, static system (Fig. 9d–f). In region 1 (
$\mathcal{H}\leqslant 0$), since both fluxes are zero the system is not strictly hyperbolic and leads to a single wave. As the characteristic speed, λp, is constant, this wave is linearly degenerate and since
$\lambda_1=\lambda_2=0$, the characteristic is stationary. The resulting wave is a stationary contact discontinuity
$\mathscr{C}$. There will not be any transport between the two states since the fluxes of both composition and enthalpy are zero on either side. In other words, ul will be connected to ur by a stationary contact discontinuity
$\mathscr{C}$ as

The resulting solution thus takes the form

In this example, a temperate, 40% porous firn lies on top of a cold (
$T=-19.8^\circ$C), 20% porous firn corresponding to
$\textbf{u}_l=(0.6,0.0)^T$ and
$\textbf{u}_r=(0.8,-0.1)^T$, as shown in Fig. 9a–c (blue lines) and d–f. However, in reality, the firn may compact due to overburden (Cuffey and Paterson, Reference Cuffey and Paterson2010) which is not considered in the present model. Note that the temperatures (not shown) in the two layers are the same as the initial condition with a sharp transition at ζ = 0 due to the absence of heat conduction.
3.4. Solutions transitioning between regions
3.4.1. From region 1 (ice and gas) to region 2 (three-phase region), Case VIII
This case resembles a dry, porous cold firn layer on top of wet temperate firn (Fig. 9g–f), without additional meltwater influx. In this case, the left state lies in region 1 corresponding to the cold firn and the right state resides in the three-phase region (region 2). The result is a contact discontinuity
$\mathscr{C}_1$ onto the lower boundary of the three-phase region (
$\mathcal{H}=0$) where the intermediate state lies (see Fig. 9a, e.g.), i.e.

Moreover, the intermediate state,
$\mathcal{C}_i$, lies on the fast path (constant porosity line) in region 2 satisfying Eqn (39) which gives

Simultaneously, the intermediate state is connected to the right state on the fast path, resulting in a rarefaction wave
$\mathscr{R}_2$. It is important to note that the second wave,
$\mathscr{W}_2$, is supposed to be faster than the first and that is why the slow path is avoided in the three-phase region (region 2). The final solution to this case is the same as given in Eqn (51).
As an example, the left state corresponds to a 30% porous cold layer at T
$=-22.63^\circ$C lying on top of the layer corresponding to wet, temperate firn similar to liquid storage with 80% porosity and 60% LWC (
$\phi_{w,r}$), as shown in Fig. 9g. This configuration corresponds to
$\textbf{u}_l=(0.7,-0.1)^T$ and
$\textbf{u}_r=(0.8,0.6)^T$ with intermediate state
$\textbf{u}_i=(0.2,0.0)^T$ (Fig. 9a, red line). As a result, the porosity jump remains stationary but the liquid storage drains downwards due to gravity forming a self-similar rarefaction wave as shown in Fig. 9b–c (red lines) and g–i. In a nutshell, this case describes the evolution of a more saturated firn layer below a previously formed less permeable, cold frozen fringe.
3.4.2. From region 2 (three-phase region) to region 1 (ice and gas)
This corresponds to temperate and wet firn overlying initially cold firn (see Figs 10–12). In this case, the left state corresponding to the top layer lies in the three-phase region (region 2) and the right state corresponding to the bottom layer lies in region 1. Note that the temperature remains subzero only in the cold region (region 1,
$\mathcal{H} \lt 0$) which lies only in the right state with a sharp transition from the left or intermediate state as heat conduction is not considered in this model. There can be four scenarios corresponding to this initial condition:

Figure 10. Solutions when the left state lies in region 2 (three-phase) and the right state lies in region 1 without formation of saturated regions: either only a refreezing front
$\mathscr{S}$ (Case IX) appears or a contact discontinuity,
$\mathscr{C}_1$, an intermediate state, ui, along with a refreezing front
$\mathscr{S}_2$ forms in a
$\mathscr{C}_1\mathscr{S}_2$ (Case X) fashion. (a) Construction of solution in the hodograph plane and their corresponding self-similar analytical solutions for (b) dimensionless composition and (c) dimensionless enthalpy with dimensionless velocity η. Red and blue lines, respectively, show the solutions when the right states are in region 2 and region 1, respectively. The evolution of the volume fractions of the three phases in the resulting system at dimensionless times
$\tau=0,0.5,1.0$ for the two configurations: (d–f) Case VII—
$\mathscr{C}$ and (g–i) Case VIII—
$\mathscr{C}_1\mathscr{R}_2$.
(i) Shock (Case IX)
This case corresponds to a sudden increase in meltwater flux into temperate, wet firn overlying cold firn, resulting in meltwater percolation into the deeper layers and formation of a frozen fringe (Fig. 10d–f). When the right state in region 1 lies on the fast path in region 2 (three-phase region) extended to region 1 (ice and gas region) referred to as extended fast path, it results in only a single moving shock
$\mathscr{S}$ (see Fig. 10a, blue line). Note that the extended fast paths are not the constant porosity contours in region 1 (Fig. 1c and g). Mathematically, the left and right states are connected by the relation

The shock speed
$\Lambda_{\mathscr{S}}$ can then be evaluated using the Rankine-Hugoniot condition (46). The final solution of this case is provided in Eqn (47). Physically, the water seeps from the top layer to the bottom layer due to gravity and a part of it precipitates due to heat loss to the surrounding cold firn. Since the flux in the right state is zero, the evaluation depends on the difference of either dimensionless composition or enthalpy. Fig. 10a–c (blue line) show an example of the solution where temperate firn with 45% porosity (φl) and 25% LWC (
$\phi_{w,l}$) initially lies on a cold layer of porosity 50% at T
$=-15.84^\circ$C (Fig. 10d). These initial conditions correspond to a left state,
$\textbf{u}_l=(0.8,0.25)^T$ and a right state
$\textbf{u}_r=(0.5,-0.05)^T$ (Fig. 10a–c, blue line). A refreezing shock
$\mathscr{S}$ that results moves downwards while warming up the surrounding snow by partially refreezing (Fig. 10d–f). This reduces the porosity behind the wetting front from
$50\%$ to
$45\%$, equal to the same value as the left state thereby extending the previously formed frozen fringe in the top layer into the bottom layer.
(ii) 1-Contact discontinuity and 2-Shock (Case X)
This is similar to Case IX (wet firn overlying cold firn), but with a decreased porosity in the underlying cold firn leading to the formation of a frozen fringe (Fig. 10g–i). Here, the left state cannot be directly connected to the right state along the extended fast path as discussed in Case IX. So in this case, the left state ul in region 2 is more porous and connected to an intermediate state, ui, along the slow path where the dimensionless composition of the intermediate state,
$\mathcal{C}_i$, is less than unity, to keep the medium unsaturated (see Fig. 10a, red line). Similar to the only shock
$\mathscr{S}$ case (Case IX), the intermediate state lies on the extended fast path from region 2 to region 1. Therefore, the solution is a combination of a stationary contact discontinuity
$\mathscr{C}_1$ and a moving shock
$\mathscr{S}_2$. The first characteristic wave, contact discontinuity
$\mathscr{C}_1$, satisfies Eqn (38) for left state ul and intermediate state ui. Since the second characteristic lies on the Hugoniot locus, a shockwave
$\mathscr{S}_2$ results, that satisfies the Hugoniot-jump condition (33) for the intermediate state ui and right state ur. Combining all these equations gives a simple relation for dimensionless enthalpy and composition at the intermediate state as

The dimensionless velocity of the shock wave
$\mathscr{S}_2$ (refreezing front) is

The full solution is given in Eqn (53) with shock speed (69). Since
$\mathcal{H}_r \lt 0$,
$\mathcal{H}_i \gt 0$ and
$f(\textbf{u}_r)=0$, the speed of the refreezing front
$\mathscr{S}_2$ is slower than the temperate firn case (Case V) due to refreezing. Fig. 10a–c (red line) show an example of such a solution for light rainfall on a multilayered firn with coarse to fine transition. The left state corresponds to the wetter, temperate firn on top with a porosity of
$70\%$ and a liquid water content
$\phi_{w,l}=0.1$. The right state corresponds to a cold and dry firn layer with 35% porosity and a temperature of T
$=-19.49^\circ$C. These values for left and right states correspond to states being
$\textbf{u}_l=(0.4,0.1)^T$ and
$\textbf{u}_r=(0.65,-0.08)^T$, respectively. Here, a stationary contact discontinuity
$\mathscr{C}_1$ stays at the surface leading to a growing intermediate state
$\textbf{u}_i=(0.893,0.163)$ formed due to partial refreezing of the rainwater which warms the surrounding firn (Fig. 10g–i). The intermediate state with
$27\%$ porosity and
$16.3\%$ LWC expands with time as the refreezing front
$\mathscr{S}_2$ infiltrates further into the right state. Note that the porosity in the intermediate state is smaller than the right state which decreases further due to meltwater refreezing leading to the formation of a fresh frozen fringe. This phenomenon has been observed in the model by Meyer and Hewitt Reference Meyer and Hewitt(2017) who used the dimensional form of Eqn (69) with distinct densities of ice and water phases to explain the field data from Humphrey and others (Reference Humphrey, Harper and Pfeffer2012).

Figure 11. Formation of a fully saturated region when right state lies in region 1: (a) Construction of solution in the hodograph plane and their corresponding self-similar analytical solutions for (b) dimensionless composition and (c) dimensionless enthalpy with dimensionless velocity η. The result shown with dark blue line consists of a backfilling shock,
$\mathscr{S}^*_1$, a jump,
$\mathscr{J}_2$, and another refreezing shock,
$\mathscr{S}_3$, along with two intermediate states
$\textbf{u}_{i_1}=(\mathcal{C}_{i_1},\mathcal{H}_{i_1})^T$ and
$\textbf{u}_{i_2}=(\mathcal{C}_{i_2},\mathcal{H}_{i_2})^T$. The evolution of the volume fractions of the three phases in the resulting system at dimensionless times (d) τ = 0, (e) 0.1, (f) 0.2.
(iii) 1-Backfilling shock, 2-Jump and 3-Shock (formation of a saturated region, Case XI)
This is similar to Case X, where wet and temperate firn layer lies on top of cold and less porous firn, and the system receives a sudden increase in meltwater flux (Fig. 11d–f). However, the temperature of the underlying cold firn is reduced, and the liquid water content in the above temperate layer increases which leads to the formation of a rising perched water table. The slow path originating from the left state and the extended fast path emanating from the right state do not intersect where
$\mathcal{C} \lt 1$, making it different from Case X. If the intermediate state lies on the saturated region, then the hyperbolic nature of the solution breaks down, similar to Case VI for temperate firn (see Table 2). So the analytic solution (63) is the same as provided for the temperate firn (Case VI) consisting of a backfilling shock
$\mathscr{S}^*_1$ moving upwards with speed (55), a stationary jump
$\mathscr{J}_2$ at the location of initial jump ζ = 0 and a ‘refreezing’ front
$\mathscr{S}_3$ moving downwards with speed provided in Eqn (57). The flux in the saturated region qs is again provided by Eqn (58). Note that the flux at the right state is
$f(\textbf{u}_r)=0$ since it lies in region 1. Invoking the ansatz for a constant shock speed ratio, similar to Case VI, provides an analytic relation for a constant flux in the saturated region, qs, same as Eqn (62) with both shocks moving in opposite directions at constant speeds. However, since the flux at the right state is zero, the relation for the ratio of shock speeds,
$\mathfrak{R}$, earlier provided by Eqn (61), now simplifies to

Fig. 11 shows an example of this case where the left state corresponds to wet and temperate layer with 80% porosity (φl) and 65% LWC (
$\phi_{w,l}$) lying on top of very cold and less porous (T
$=-30^\circ$C,
$\varphi_r=50\%$) firn. This corresponds to the left and right states being
$\textbf{u}_l=(0.85,0.65)^T$ and
$\textbf{u}_r=(0.5,-0.095)^T$. The first and second intermediate states are
$\textbf{u}_{i_1}=(1,0.8)^T$ and
$\textbf{u}_{i_2}=(1,0.405)^T$, respectively. The upper shock, also called a rising perched water table, is almost four times faster than the lower, refreezing front (see Fig. 11b–f). Below the initial jump at ζ = 0 the porosity is further reduced by refreezing, leading to the formation of a frozen fringe (Fig. 11e–f). The second intermediate state below jump
$\mathscr{J}_2$ with a newly formed frozen fringe is unable to accommodate the whole volumetric flux of water, leading to the formation of a rising perched water table (see Fig. 11e–f). Once the rising perched water table reaches the surface, it will lead to ponding and can eventually form runoff.
(iv) 1-Backfilling shock, 2-Jump and 3-Contact discontinuity (formation of an impermeable ice layer, Case XII)
This final case captures the formation of an impermeable ice layer through advection of meltwater and associated latent heat. An ice layer forms through the construction in Case XI, if the extended fast path (
$\mathcal{H}=\mathcal{C}+\mathfrak{C}$) emanating from the right state reaches
$\mathcal{C}\geqslant 1$ at the solidus (
$\mathcal{H}=0$) or the porosity of the second intermediate state reaches 0. This case is entirely governed by the right state and an ice layer will form if and only if the right state satisfies


Figure 12. Impermeable ice layer formation: (a) Construction of solution in the hodograph plane and their corresponding self-similar analytical solutions for (b) dimensionless composition and (c) dimensionless enthalpy with dimensionless velocity η. The result shown with dark blue line consists of a backfilling shock,
$\mathscr{S}^*_1$, a jump,
$\mathscr{J}_2$ and a contact discontinuity,
$\mathscr{C}_3$ along with two intermediate states
$\textbf{u}_{i_1}=(\mathcal{C}_{i_1},\mathcal{H}_{i_1})^T$ and
$\textbf{u}_{i_2}=(\mathcal{C}_{i_2},\mathcal{H}_{i_2})^T$. The second intermediate state
$\textbf{u}_{i_2}$ corresponds to the impermeable ice layer. The grey region corresponds to
$1-\mathcal{C}+\mathcal{H}\leqslant 0$ where the right state resides to cause impermeable ice layer formation. The evolution of the volume fractions of the three phases in the resulting system at dimensionless times (d) τ = 0, (e) 0.02, (f) 0.04.
The exact location of the right state does not matter if it lies in the region of ice layer formation (grey region in Fig. 12a). In other words, an impermeable ice layer via heat advection will form if and only if the cold content of the firn exceeds the latent enthalpy of incoming meltwater. The mathematical condition (71) in dimensional form has been given in Humphrey and others (Reference Humphrey, Harper and Meierbachtol2021) for distinct densities of ice and water. In this limit, the solution for Case XI breaks down due to the formation of an impermeable ice layer. Fig. 12 shows the region of ice layer formation which requires either very low firn porosity or temperatures in the right state. The final solution for this case is a backfilling shock,
$\mathscr{S}^*_1$ to the intermediate saturated region i 1. The first saturated region is connected to the second intermediate state i 2 through the jump
$\mathscr{J}_2$. The second intermediate state i 2 is the infinitesimally thin ice layer (see Fig. 12e–f) that blocks further meltwater percolation. The solution in this case can be written as

where the state solution is quantified as

where
$\mathrm{d} \zeta$ is the infinitesimally small thickness of the ice layer.
In this case, only a single shock moves upwards similar to the filling of a bucket. The flux in the saturated region is simply zero from the mass balance at the ice layer, i.e.
$q_s=f(\textbf{u}_{i_1})=f(\textbf{u}_{i_2})=0$. The speed of the rising perched water table (first shock
$\mathscr{S}^*_1$) is then

which depends only on the left-state variables.
It can be deduced that either a cold firn (lower
$\mathcal{H}$) or a nearly non-porous firn with
$\mathcal{C}$ close to unity will induce the formation of an impermeable ice layer via refreezing caused by heat advection. For example, Fig. 12 shows such a solution when a wet and temperate layer of 70% porosity and 65% LWC lies on top of a cold layer with 5% porosity at T
$=-30^\circ$C. These conditions correspond to
$\textbf{u}_l = (0.95,0.65)^T$ and
$\textbf{u}_r = (0.95,-0.180)^T$ (Fig. 12a–c). Here the intermediate states lie at
$\textbf{u}_{i_1}=(1,0.7)^T$ and
$\textbf{u}_{i_2}=(1,0)^T$, respectively. An impermeable ice layer forms as a result (see Fig. 12e–f) that blocks the flow of meltwater downwards. Since there is no meltwater percolating in the lower layer, the backfilling occurs very rapidly.
4. Comparison against numerical solutions and limitations of the theory
All of the analytic solutions described in Sections 3.3 and 3.4 show excellent comparison against the corresponding numerical solutions of Equations (8) and (9) in the absence of heat conduction (not shown for brevity). In this section, we demonstrate the application of this theory to a more realistic multilayered firn leading to the formation of a perched firn aquifer. The analytic solutions are compared against the corresponding numerical solution without heat conduction. Finally, we summarize the potential limitations of the theory for consideration and future development.

Figure 13. Infiltration into a multilayered firn with porosity and temperature decay with depth: (a) Schematic diagram showing all of the layers (b) Construction of solution in the hodograph plane. The contours showing evolution of the firn (c) porosity φ, (d) liquid water content LWC or volume fraction of water ϕw and temperature T evaluated by the numerical simulator. Here all dashed lines show analytic solutions computed from the proposed theory. The thin, grey dashed lines show theoretically calculated dimensionless times of saturation τs and ponding τp. The theoretical evolution of the initial wetting front
$\mathscr{S}$ (red dashed line) is computed from Case III, whereas the dynamics of saturated region after wetting front
$\mathscr{S}$ reaches ζ = 1 shown by blue and green dashed lines is computed by Case XI. Here δ and
$t_c = \delta/K_h$ are characteristic times with former being calculated from their definition 15. For example, the characteristic depth is δ = 5 m and for
$K_h=5\cdot10^{-4}$ m/s (see Table 1), the characteristic time comes out to be
$t_c=2.53$ hours. The dimensionless wetting front speeds can be redimensionalized by multiplying with saturated (and no-matrix) hydraulic conductivity Kh.
4.1. Meltwater infiltration into a multilayered firn—formation of a perched water table
This final test shows the infiltration into multilayered firn after a melt event combining two cases proposed in Section 3 that are summarized in Table 2 (see Fig. 13a). This problem summates the commonly studied wetting front propagation in a temperate region (Case III) with the wetting front in a cold region and the formation of a perched water table (Case XI) that has not been studied in the firn literature. This problem demonstrates a delay in meltwater ponding at the surface due to a decay in both firn porosity and temperature with depth. The firn is 70% porous, dry and at 0∘C, above a dimensionless depth of ζ = 1 (Fig. 13a). At time τ = 0, the meltwater is generated at the surface (ζ = 0), which keeps the liquid water content (denoted by LWC or ϕw) to 40% making the firn very wet at the surface, ζ = 0. The analytic solution of this problem can be constructed in the
$\mathcal{C}-\mathcal{H} $ hodograph plane in two parts (Fig. 13b).
First, a wetting front
$\mathscr{S}$ initially propagates downwards with a constant dimensionless speed
$\Lambda_{\mathscr{S}}=0.28$, as given in Case III (Fig. 13c–e, red dashed line). The dimensionless time when the initial front reaches the depth of transition
$\zeta_t = 1$ to form a saturated region is
$\tau_s = \frac{\zeta_t}{\Lambda_{\mathscr{S}}}=1/0.28=3.57$, as shown in Case XI. Afterwards the saturated region forms due to large meltwater flux compared to the hydraulic conductivity of the second intermediate region formed by refreezing below the jump ζ > 1. This makes the flow pressure-driven instead of gravity-driven and enforces
$q_s(\tau) \leqslant K_{i_2}$. This saturated region expands in both directions as a perched water table (upper shock shown by green dashed line in Fig. 13) rising to the surface and the wetting front
${\mathscr{S}_3}$ that percolates in the cold region (lower shock shown with blue dashed line) while refreezing a part of meltwater to warm the surrounding firn. Thus, it shows a reduction in porosity from 30% to 21.2% for ζ > 1 behind the wetting front
$\mathscr{S}_3$ at
$\tau \gt \tau_s$ (Fig. 13c). The perched water table
${\mathscr{S}^*_1}$ rises upwards with constant dimensionless velocity
$\Lambda_{\mathscr{S}^*_1}=-0.268$ until ponding occurs. Meanwhile, the wetting front keeps percolating in the cold firn with reduced velocity
$\Lambda_{\mathscr{S}_3}=0.105$. Lastly, the ponding starts at a time when the rising perched water table reaches the surface so the dimensionless ponding time can be calculated theoretically as
$\tau_p = \zeta_t/\Lambda_{\mathscr{S}} + (-\zeta_t)/\Lambda_{\mathscr{S}^*_1} = 1/0.28 + (-1)/(-0.268)=7.30$. All of these dimensionless shock speeds and times are computed analytically and the resulting locations are graphed with dashed lines in Fig. 13c–e.
These theoretical results show an excellent comparison with the numerical solutions shown in contour plots. The numerical solutions are obtained by solving the governing model (26) performed in the absence of capillary effects in between water and gas phases as well as heat conduction along with the treatment in the saturated region given by Shadab and Hesse (Reference Shadab and Hesse2024) and implementation of the enthalpy method in Shadab and others Reference Shadab, Adhikari, Rutishauser, Grima and Hesse(2024a). Further, the densities of the water and the ice phases are assumed to be the same. The computational domain
$\zeta \in [0,2]$ is divided uniformly into 400 cells. The boundary condition at the top surface (ζ = 0) is prescribed as the ‘Top condition’ in Fig. 13a whereas the bottom boundary condition is not required. This problem constitutes a very specific benchmark test for firn hydrology simulators that are able to simulate variably saturated flows. It shows how vertical heterogeneity in firn may lead to the formation of perched aquifers that can cause ponding at a later stage. It also illustrates that an impermeable layer is not required to cause meltwater perching and subsequently, ponding.
4.2. Limitations of the present theory and further work
The unified kinematic wave theory makes necessary hydrologic and thermodynamic simplifications to make the system hyperbolic and amenable to solution using the method of characteristics. Some of these simplifications can be relaxed, but doing so will make the solutions more involved. On the hydrologic side, the theory neglects capillary forces in unsaturated firn, which may introduce significant errors at small spatial scales and grain sizes. Thus, it assumes that meltwater advection occurs faster as a wavefront than through diffusion via capillary suction. Capillary forces account for <10% of the total force, including gravity, when the meltwater flux is 10−8 m/s, although the percentage rapidly increases for smaller fluxes (Colbeck, Reference Colbeck1974b). The theory assumes that the density of water and ice are the same, which underestimates the porosity reduction due to refreezing when a wetting front percolates into a cold, dry firn. The kinematic model could easily be extended to account for different densities. The present model does not assume compaction of ice due to factors such as overburden, temperature, time and initial snow properties (e.g. crystal shape, packing geometry), which is present in both wet and dry firn (Bader, Reference Bader1954; Mellor, Reference Mellor1977; Herron and Langway, Reference Herron and Langway1980; Amory and others, Reference Amory2024). Lastly, the present theory assumes a one-dimensional flow in the direction of gravity with piecewise homogeneous firn and does not consider preferential flow which might enhance the speed and depth of meltwater percolation at higher melt fluxes (Wever and others, Reference Wever, Würzer, Fierz and Lehning2016; Vandecrux and others, Reference Vandecrux2020; Jones and others, Reference Jones, Moure and Fu2024).
On the thermodynamic side, the present theory assumes local thermodynamic equilibrium, which may not be a suitable assumption for rapid infiltration. Further, the theory neglects heat conduction, which may be dominant at smaller spatial scales and meltwater fluxes. Heat conduction leads to the formation of a conductive boundary layer ahead of the wetting front (Shadab and others, Reference Shadab, Adhikari, Rutishauser, Grima and Hesse2024a). The width of the thermal boundary layer is controlled by the Peclet number for enthalpy transport, as defined in Section 2.4. A high Peclet number typically leads to a thinner thermal boundary layer, indicating the dominance of heat advection over heat conduction.
However, the kinematic theory could be extended to incorporate several additional physical processes. It could be extended to account for hysteresis in relative permeabilities, which allows melt to enter dry snow without having to overcome a saturation threshold but to leave behind residual melt during drying. The method of characteristics can also be applied to study the kinematics of non-equilibrium processes, though the solutions are no longer self-similar. This could be used to analyze the percolation of melt that is not in thermal equilibrium with the ice or the coarsening of ice grains due to interaction with melt. The theory could be extended to add tracer transport that could be used to study changes in isotopic composition of the ice or dissolved gases in the melt and their effect on the climatic record in ice cores. Finally, the theory could be extended to incorporate salt and used to study the effect of colligative melting point depression in firn processes or sea-water intrusion into firn. As such, the kinematic theory is a useful tool to probe the physics of advective processes in firn and provide benchmark analytic solutions to a range of coupled non-linear processes.
5. Conclusions
This work introduces a unified kinematic wave theory for meltwater infiltration into firn that helps construct analytic solutions in the hodograph plane. The theory neglects heat conduction and capillary forces while assuming a constant density for water and ice phases. We provide a suite of 12 cases of melt infiltration into firn, inspired by nature, and construct their analytic solutions while connecting most of the cases given in the literature. The previous works were predominantly limited to unsaturated wetting fronts in cold and temperate firn. We consider cases that have not been studied previously such as the formation of a perched water table (Case VI and Case XI) and the formation of an impermeable ice layer (Case XII). The simple cases discussed here can be combined to study more realistic problems, as was demonstrated for the case of infiltration into multilayered firn. The combined solutions can help construct time-varying solutions with more complexity such as variable surface conditions. There are several consequences of this work. First, one can interpret the physics of the meltwater infiltration into firn without running expensive numerical simulations. This can be used to better constrain the process of firn densification, and the partitioning of meltwater runoff versus storage, which is crucial when deriving surface mass balance. Second, these analytic solutions can help in developing better, cost-effective physics-based firn hydrological models which can then be integrated with ice-sheet and Earth system models. Further, these problems can serve as a benchmark for the next generation of wet firn hydrological models which currently show significant deviation due to a lack of benchmark problems. This comprehensive framework can significantly enhance our understanding of wet firn hydrology, a component that has been poorly understood, ultimately aiding in constraining its contribution to surface mass balance loss from glaciers and, consequently, sea-level rise.
Acknowledgements
The authors would like to acknowledge Dr Surendra Adhikari and Dr Andreas Colliander for the discussions on meltwater infiltration at the Dye-2 site in Southwest Greenland. The code and data used to generate all figures of this paper are available in a permanent repository (https://doi.org/10.5281/zenodo.13936153) (Shadab and others, Reference Shadab, Rutishauser, Grima and Hesse2024b) as well as Github for reproducibility (https://github.com/mashadab/unified-kinematic-wave-theory). More information is provided in the Readme of the repository.
Funding
Support for this work was provided through NASA under Emerging World Grant number NASA 18−EW
$18\_ 2-0027$ and the University of Texas Institute for Geophysics under the Blue Sky Student Fellowship.
Appendix A. Flux gradient and lemma related to eigen-decomposition
The partial derivatives of the flux function given in Eqn (24) give:


Lemma. Prove that λ 2 is nonnegative and increases monotonically in the direction of integral curves corresponding to the second eigenvector r2 (fast paths) in the three-phase region,
$0 \lt \mathcal{H} \lt \mathcal{C}$.
Proof. Since
$\mathcal{H} \gt 0$,
$-\mathcal{C} \lt \mathcal{H}-\mathcal{C}$, and
$0 \leqslant \mathcal{C} \lt 1$,

Along the fast path,
$\mathcal{C}=\mathcal{H}+\mathfrak{C}$,



Now, taking the derivative with respect to
$\mathcal{H}$, we get

As
$\mathcal{H} \gt 0$ and
$\mathfrak{C}=\mathcal{C}-\mathcal{H} \lt 1$ in three-phase region,

Therefore, λ 2 is nonnegative and increases monotonically in the direction of r2 when n > 1.□
Lemma. Prove that fast paths are parallel to constant porosity φ contours in three-phase region,
$0 \lt \mathcal{H} \leqslant \mathcal{C}$.
Proof. In the three-phase region, the porosity
$\varphi=\phi_w+\phi_g=\mathcal{H}+(1-\mathcal{C})$. For a constant porosity
$\mathrm{d} \varphi=0$, therefore




where
$\mathfrak{C}$ again is an integration constant. Therefore, the constant porosity φ contours are the integral curves corresponding to second eigenvector r2 (fast paths) in three-phase region.□