1 Introduction

Bingham fluids, or yield stress fluids, are encountered in a wide range of applications: toothpastes, cements, mortars, foams, muds, mayonnaise, etc. The fundamental character of these fluids is that they are able to deform indefinitely only if they are submitted to a stress above some critical value. Actually, toothpaste visually exhibits the fundamental character of such a fluid model: it flows when the applied stress exceeds a certain value, otherwise it does not flow, behaving as a solid body (actually below the critical stress toothpaste deforms in finite way). Despite the apparent simplicity in the constitutive modeling (especially within the implicit framework theory developed by Rajagopal and co-workers [22,23,24,25]), the flow characteristics of these materials are difficult to be predicted, since they involve unknown boundaries separating the liquid and the solid regions. Readers are referred to the books by Huilgol [14] and by Ionescu, Sofonea [15] where several issues concerning the Bingham model (constitutive equations, mathematical techniques, numerical methods and so on) are deeply analyzed.

A well known example of materials which are often modeled as Bingham Fluids are waxy crude oils (i.e. oils with an high paraffin content). These fluids are known to cause handling and pipelining difficulties. The flow properties depend strongly on the yield stress which, in turn, depends on the shear history [31]. This leads to a definable minimum operating point below which flow in a waxy crude oil pipeline would cease.

Familiar examples of non-Newtonian fluids described by the Bingham model include also mud (see [18] and the references therein cited), lubricated pipelining [13], and materials used in ceramic casting [16].

2 Constitutive Model

The simplest shear-stress experiment that characterizes the Bingham fluid is represented in Fig. 1. On the top surface of a layer of material a uniform shear force F is applied. If A is the area of the surface, the applied shear is FA. If the applied shear load (i.e. force per unit surface) does not exceed a certain threshold, τ o, the material does not move (the bottom of the layer is fixed on the “floor” ). When the applied shear load exceeds the τ o, the material flows as a linear viscous fluid.

Fig. 1
figure 1

A schematic representation of the 1D shear stress experiment. F is the shear force and A which is uniformly applied on a surface A so that the shear stress is FA. The threshold is τ o

Actually such a peculiar rheological behavior is well highlighted by the experimental tests performed in cylindrical viscometers. We just mention the recent review by Coussot [5] and the numerous experimental papers therein cited. Hence, considering a simple one-dimensional shear flow, if τ denote the modulus of the shear stress and \(\dot {\gamma }\) the modulus of the strain rate, the constitutive Bingham model writes as

$$\displaystyle \begin{aligned} \tau =\tau _{o}+\mu \dot{\gamma},\ \ \ \ \mathrm{if}\ \ \ \ \tau >\tau _{o},\ \ \ \ (\mathrm{flow),} {} \end{aligned} $$
(1)

and

$$\displaystyle \begin{aligned} \dot{\gamma}=0,\ \ \ \ \mathrm{if}\ \ \ \ \tau \leq \tau _{o},\ \ \ \ (\mathrm{no\ \ flow).} {} \end{aligned} $$
(2)

Indeed (2) means, from the physical point of view, rigid behavior [11]. The threshold τ o is usually defined shear yield stress, or simply yield stress, and μ is referred to as viscosity. In Fig. 2 we have reported the shear stress-shear rate curve for a Bingham fluid.

Fig. 2
figure 2

The shear stress—shear rate curve for the Bingham model

The model (1), (2) was introduced for the first time by E.C. Bingham [1, 2]. We note that in the Bingham model is much more natural to express the modulus of the shear rate \(\dot {\gamma }\) in terms of the modulus of the shear stress τ. Indeed (1), (2) can be rewritten (see also Fig. 3 )

$$\displaystyle \begin{aligned} \dot{\gamma}={\displaystyle {\left( \tau -\tau _{o}\right) _{+} \over \mu }}, {} \end{aligned} $$
(3)

where \(\left ( \ \ \ \ \right ) _{+}\) denotes the positive part, namely

$$\displaystyle \begin{aligned} \left( \tau -\tau _{o}\right) _{+}=\left\{ \begin{array}{l} \tau -\tau _{o},\ \ \mathrm{if}\ \ \ \tau \geq \tau _{o}, \\ \\ 0,\ \ \mathrm{if}\ \ \ \tau <\tau _{o}. \end{array} \right. \end{aligned}$$

In the example of Fig. 1 it is easy to identify the shear stress and formulate the yield criterion. But how can we formulate the yield criterion in the general 3D case? What is in this case the shear stress? Let us consider a point P and a small facet of area ds surrounding. The normal to the facet is n. If we denote by F the force acting on ds we have

$$\displaystyle \begin{aligned} \boldsymbol{F}=\mathbf{T}\left(P\right) \boldsymbol{n}, \end{aligned} $$

where \(\mathbf {T}\left ( P\right ) \) is the Cauchy stress. The force vector F can be splitted into its component along n and in a tangential vector τ (said tangential or shear force)

$$\displaystyle \begin{aligned} \boldsymbol{F}=\boldsymbol{\tau }+\sigma \boldsymbol{n}, \end{aligned} $$

where \(\sigma =\boldsymbol {F\cdot n}=\mathbf {T}\left ( P\right ) \boldsymbol { n\cdot n}\) (see Fig. 4). Obviously \(\left \vert \boldsymbol {F} \right \vert ^2 = \sigma ^2 +\tau ^2 \), where \(\tau =\left \vert \boldsymbol {\tau }\right \vert \).

Fig. 3
figure 3

Shear stress—shear rate according to (3)

Fig. 4
figure 4

Decomposition of the force F

We then consider a reference frame in which \(\mathbf {T}\left ( P\right ) \) is diagonal, namely

$$\displaystyle \begin{aligned} \mathbf{T}\left(P\right) =\left( \begin{array}{lll} T_{1} & 0 & 0 \\ 0 & T_{2} & 0 \\ 0 & 0 & T_{3} \end{array} \right) , \end{aligned} $$

and assume T 1 > T 2 > T 3. Hence

$$\displaystyle \begin{aligned} \begin{array}{rcl} \left\vert \boldsymbol{F}\right\vert ^{2} &\displaystyle =&\displaystyle \sigma ^{2}+\tau ^{2}=\left\vert \left( \begin{array}{lll} T_{1} &\displaystyle 0 &\displaystyle 0 \\ 0 &\displaystyle T_{2} &\displaystyle 0 \\ 0 &\displaystyle 0 &\displaystyle T_{3} \end{array} \right) \left( \begin{array}{l} n_{1} \\ n_{2} \\ n_{3} \end{array} \right) \right\vert ^{2} \\[0.05in] &\displaystyle =&\displaystyle T_{1}^{2}n_{1}^{2}+T_{2}^{2}n_{2}^{2}+T_{3}^{2}n_{3}^{2}, \end{array} \end{aligned} $$

and

$$\displaystyle \begin{aligned} \begin{array}{rcl} \sigma &\displaystyle =&\displaystyle \boldsymbol{F\cdot n}=\mathbf{T}\left(P\right) \boldsymbol{n\cdot n} \\[0.05in] &\displaystyle =&\displaystyle \left( \begin{array}{lll} T_{1} &\displaystyle 0 &\displaystyle 0 \\ 0 &\displaystyle T_{2} &\displaystyle 0 \\ 0 &\displaystyle 0 &\displaystyle T_{3} \end{array} \right) \left( \begin{array}{l} n_{1} \\ n_{2} \\ n_{3} \end{array} \right) \cdot \left( \begin{array}{l} n_{1} \\ n_{2} \\ n_{3} \end{array} \right) =T_{1}n_{1}^{2}+T_{2}n_{2}^{2}+T_{3}n_{3}^{2}~. \end{array} \end{aligned} $$

We thus get the system

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} \left\vert \boldsymbol{n}\right\vert ^{2}=1, \\ \\ \left\vert \boldsymbol{F}\right\vert ^{2}=\sigma ^{2}+\tau ^{2}, \\ \\ \boldsymbol{F\cdot n}=\sigma , \end{array} \right. \ \ \Longrightarrow \ \ \left\{ \begin{array}{l} n_{1}^{2}+n_{2}^{2}+n_{3}^{2}=1, \\ \\ T_{1}^{2}n_{1}^{2}+T_{2}^{2}n_{2}^{2}+T_{3}^{2}n_{3}^{2}=\sigma ^{2}+\tau ^{2}, \\ \\ T_{1}n_{1}^{2}+T_{2}n_{2}^{2}+T_{3}n_{3}^{2}=\sigma , \end{array} \right. \end{aligned}$$

which, once solved with respect to \(n_{1}^{2}\), \(n_{2}^{2}\), \(n_{3}^{2}\), gives

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} n_{1}^{2}={\displaystyle {{\displaystyle {\left( \sigma -{\displaystyle {T_{2}+T_{3} \over 2}}\right) ^{2}+\tau ^{2}-\left( {\displaystyle {T_{2}-T_{3} \over 2}}\right) ^{2} \over \left( T_{2}-T_{1}\right) \left( T_{3}-T_{1}\right) }}}}\geq 0, \\ \\ n_{2}^{2}=-{\displaystyle {{\displaystyle {\left( \sigma -{\displaystyle {T_{1}+T_{3} \over 2}}\right) ^{2}+\tau ^{2}-\left( {\displaystyle {T_{3}-T_{1} \over 2}}\right) ^{2} \over \left( T_{2}-T_{3}\right) \left( T_{1}-T_{2}\right) }}}}\geq 0, \\ \\ n_{3}^{2}={\displaystyle {{\displaystyle {\left( \sigma -{\displaystyle {T_{2}+T_{1} \over 2}}\right) ^{2}+\tau ^{2}-\left( {\displaystyle {T_{2}-T_{1} \over 2}}\right) ^{2} \over \left( T_{1}-T_{3}\right) \left( T_{2}-T_{3}\right) }}}}\geq 0. \end{array} \right. \end{aligned}$$

Thus, in the \(\left ( \sigma ,\tau \right ) \) plane a domain is defined

$$\displaystyle \begin{aligned} \begin{array}{rcl} \left( \sigma -{\displaystyle {T_{2}+T_{3} \over 2}}\right) ^{2}+\tau ^{2} &\displaystyle \geq &\displaystyle \left( {\displaystyle {T_{2}-T_{3} \over 2}}\right) ^{2}, \\ \left( \sigma -{\displaystyle {T_{1}+T_{3} \over 2}}\right) ^{2}+\tau ^{2} &\displaystyle \leq &\displaystyle \left( {\displaystyle {T_{3}-T_{1} \over 2}}\right) ^{2}, \\ \left( \sigma -{\displaystyle {T_{2}+T_{1} \over 2}}\right) ^{2}+\tau ^{2} &\displaystyle \geq &\displaystyle \left( {\displaystyle {T_{2}-T_{1} \over 2}}\right) ^{2}, \end{array} \end{aligned} $$

which corresponds to the area bordered by three circles, the so-called three Mohr circles (see Fig. 5).

Fig. 5
figure 5

Three points on the horizontal axis correspond to the three main stresses. The figure shows that the maximum shear is equal to the radius of the largest circle

Thus, since the maximum shear stress is τ MAX, a possible yield criterion is

$$\displaystyle \begin{aligned} \left\{ \begin{array}{lll} \tau _{MAX}\leq {\displaystyle {{\displaystyle {\tau _{o} \over 2}}}}, & & \text{no}\ \text{flow,} \\ & & \\ \tau _{MAX}>{\displaystyle {{\displaystyle {\tau _{o} \over 2}}}}, & & \text{flow,} \end{array} \right. {} \end{aligned} $$
(4)

where

$$\displaystyle \begin{aligned} \tau _{MAX}={\displaystyle {1 \over 2}}\max \left\{ \left\vert T_{1}-T_{2}\right\vert ,~\left\vert T_{2}-T_{3}\right\vert ,\ \left\vert T_{3}-T_{1}\right\vert \right\} . {} \end{aligned} $$
(5)

Such a criterion is known as Tresca criterion [3]. The representation in the principal stress space of the surface (5), known as Tresca Surface, or

$$\displaystyle \begin{aligned} \left[ \left( T_{1}-T_{2}\right) ^{2}-\tau _{o}^{2}\right] \left[ \left( T_{1}-T_{3}\right) ^{2}-\tau _{o}^{2}\right] \left[ \left( T_{3}-T_{2}\right) ^{2}-\tau _{o}^{2}\right] =0, \end{aligned}$$

is a cylindrical unbounded surface with hexagonal section and axis (1,1,1). So, according to the Tresca criterion, at every point of the material we have to compute the eigenvalues of the stress tensor. If all the eigenvalues are within the Tresca surface then the material is in a rigid state. if at least one of the eigenvalues is outside the surface then the material is in a fluid state.

The Tresca criterion, unfortunately, is not very practical to use. For this reason it is preferred to use the Von Mises criterion [30], based on the second invariant of the stress tensor. The maximum shear stress criterion (4) is replaced by

$$\displaystyle \begin{aligned} \left\{ \begin{array}{lll} \tau _{VM}\leq {\displaystyle{{\displaystyle {\tau _{o} \over 2}}}}, & & \text{no}\ \text{flow,} \\ & & \\ \tau _{VM}>{\displaystyle{{\displaystyle {\tau _{o} \over 2}}}}, & & \text{flow,} \end{array} \right. \end{aligned}$$

where

$$\displaystyle \begin{aligned} \tau _{VM}=\sqrt{{\displaystyle {1 \over 2}}\left[ \left( T_{1}-T_{2}\right) ^{2}+\left( T_{2}-T_{3}\right) ^{2}+\left( T_{3}-T_{1}\right) ^{2}\right] }. \end{aligned} $$

Thus, splitting the Cauchy stress as T = −P I + S, where P = 1∕3trT and S is the extra-stress, we immediately realize that σ V M coincides with the second invariant of the extra-stress, namely

$$\displaystyle \begin{aligned} \sigma_{VM}=II_{\mathbf{S}}=\sqrt{{\displaystyle {1 \over 2}}\text{tr }{\mathbf{S}}^{2}}. \end{aligned} $$

Therefore the generalization of the Bingham model (1), (2) to the 3D case is the following

$$\displaystyle \begin{aligned} \mathbf{S}=\left( 2\mu +{\displaystyle {\tau _{o} \over II_{\mathbf{D}}}}\right) \mathbf{D} ,\ \ \ \ \mathrm{if}\ \ \ \ II_{\mathbf{S}}>\tau _{o},\ \ \ \ (\mathrm{flow}), {} \end{aligned} $$
(6)

and

$$\displaystyle \begin{aligned} \mathbf{D}=0,\ \ \ \ \mathrm{if}\ \ \ \ \left\vert \tau \right\vert \leq \tau _{o},\ \ \ \ (\mathrm{no\ \ flow),} {} \end{aligned} $$
(7)

where

$$\displaystyle \begin{aligned}\mathbf{D}={\displaystyle{{\displaystyle {1 \over 2}}}}\left( \nabla \boldsymbol{v} +\nabla \boldsymbol{v}^{{}^{\mathrm{T}}}\right) \ \ \ \ \ \ \ II_{\mathbf{D}}= \sqrt{{\displaystyle{{\displaystyle {1 \over 2}}}}\,\mathrm{tr~}\mathbf{D}},\end{aligned} $$

and where v is the fluid velocity field. We remark that the Bingham constitutive equation can be written in the implicit form (see [22,23,24, 26])

$$\displaystyle \begin{aligned} \mathbf{D}=\left( {\displaystyle {II_{\mathbf{D}} \over 2\mu II_{\mathbf{D}}+{\tau _{o}}}} \right) \mathbf{S}. {} \end{aligned} $$
(8)

The above constitutive equation allows to express S as a function of D only when II S > τ o, while D = 0, entails only II S ≤ τ o, the stress being constitutively undetermined.

3 Flow in a Channel

We consider the flow of an incompressible Bingham fluid in a symmetric channel of length L and width 2H. We consider a laminar flow so that the velocity field is

$$\displaystyle \begin{aligned} \boldsymbol{v}=v(y,t)\boldsymbol{e}_{x}, \end{aligned}$$

where, as shown in Fig. 6, x, y are the longitudinal and transversal coordinates respectively.

Fig. 6
figure 6

Sketch of the channel

Due to symmetry we consider just the upper part of the channel. The the inlet pressure ΔP is prescribed and we rescale the outlet pressure to 0. Hence the pressure gradient driving the flow is given by

$$\displaystyle \begin{aligned} f_{o}={\displaystyle {\varDelta P \over L}}~. \end{aligned}$$

In principle \(f_{o}=f_{o}\left ( t\right )\) since ΔP may depend on time. In such a simple setting

$$\displaystyle \begin{aligned} \mathbf{D}=\left( \begin{array}{lll} 0 & v_{y} & 0 \\ v_{y} & 0 & 0 \\ 0 & 0 & 0 \end{array} \right) ,\ \ \ \ \ \mathbf{T}=\left( \begin{array}{ccc} -P & S_{12} & 0 \\ S_{12} & -P & 0 \\ 0 & 0 & -P \end{array} \right) , \end{aligned}$$

so that \(II_{\mathbf {S}}=\left \vert S_{12}\right \vert \) and \(II_{\mathbf {D} }=\left \vert v_{y}\right \vert \). Hence, (6), (7) give

$$\displaystyle \begin{aligned} \begin{array}{rcl} II_{\mathbf{S}} &\displaystyle >&\displaystyle \tau _{o},\ \ \ \ \Longrightarrow \ \ \ \ S_{12}=-\tau _{o}+\mu v_{y}, \\[0.04in] II_{\mathbf{S}} &\displaystyle \leq &\displaystyle \tau _{o},\ \ \ \ \Longrightarrow \ \ \ \ v_{y}=0. \end{array} \end{aligned} $$

Next, we assume that the viscous region, namely II S > τ o, and the rigid region, II S ≤ τ o, are separated by a sharp interface y = s(t) a priori unknown, i.e. a free boundary. In the fluid region, i.e. \(s\left ( t\right ) <y<H\), the motion equation reduces to

$$\displaystyle \begin{aligned} \rho v_{t}=\mu v_{yy}+f_{o}, \end{aligned}$$

coupled with the no-slip condition on y = H,

$$\displaystyle \begin{aligned} v\left( H,t\right) =0, \end{aligned}$$

and with the threshold condition on the free boundary \(s\left ( t\right ) \)

$$\displaystyle \begin{aligned} \left. S_{12}\right\vert _{y=s }=\tau _{o},\ \ \ \ \Longrightarrow \ \ \ \ \left. v_{y}~\right\vert _{y=s }=0. \end{aligned}$$

In particular, we assume no-slip also at the interface \(s\left ( t\right ) \), so that the velocity of the rigid core is

$$\displaystyle \begin{aligned} \kappa \left( t\right) =v\left( s\left( t\right) ,t\right) . \end{aligned}$$

Figure 7 represents a sketch of the problem we have to solve.

Fig. 7
figure 7

The position s(t) of the “fluid – rigid” is unknown. Indeed we deal with a free boundary problem

We need an evolution for the interface \(s\left ( t\right ) \). According to the approach developed in [12], which is based on the pioneering works by Safronchik [29] and Rubinstein [28], the unyielded region is treated as an evolving non material volume, whose motion is determined by using the integral (or global) momentum balance. The dynamics of the unyielded domain

$$\displaystyle \begin{aligned} \varOmega \left( t\right) =\left\{ 0<x<L,\ \ -s\left( t\right) <y<s\left( t\right) \right\} , \end{aligned}$$

is thus given by (see, e.g., [3])

$$\displaystyle \begin{aligned} {\displaystyle {d \over dt}}\int_{\varOmega \left( t\right) }\rho \boldsymbol{v} dV=\int_{\partial \varOmega \left( t\right) }\mathbf{T}\boldsymbol{n~} dS-\int_{\partial \varOmega \left( t\right) }\rho \boldsymbol{v}\left[ \left( \boldsymbol{v}-\boldsymbol{w}\right) \cdot \boldsymbol{n}\right] \boldsymbol{ ~}dS, {} \end{aligned} $$
(9)

where w is the velocity of the boundaryFootnote 1 ∂Ω and n its outward normal.

Taking Fig. 8 into account relation (9 ) reduces toFootnote 2

$$\displaystyle \begin{aligned} v_{t}\left( s\left( t\right) ,t\right) ={\displaystyle {1 \over \rho }}\left(f_{o}-{\displaystyle { \tau _{o} \over s\left( t\right) }}\right) . \end{aligned}$$
Fig. 8
figure 8

The evolving non-material domain Ω

We thus end up with this free boundary problem

$$\displaystyle \begin{aligned} \left\{ \begin{array}{lll} \rho v_{t}=\mu v_{yy}-f_{o}, & & s\left( t\right) <y<H,\ t>0 \\ \\ v\left( H,t\right) =0, & & t>0, \\ \\ v_{y}\left( s\left( t\right) ,t\right) =0, & & t>0, \\ \\ \rho v_{t}\left( s\left( t\right) ,t\right) =f_{o}-{\displaystyle {{\displaystyle { \tau _{o} \over s\left( t\right) }}}}, & & t>0, \\ \\ v\left(y,0\right) =v_{o}\left(y\right) & & 0<y<s_{o}, \\ \\ s\left( 0\right) =s_{o}. & & \end{array} \right. {} \end{aligned} $$
(10)

The domain of the problem is represented in Fig. 9.

Fig. 9
figure 9

Sketch of the domain of problem (10)

The problem (10) is a free boundary problem but not of Stefan type, because in the evolution equation for the free boundary \(\dot s\) does not appear. However if we consider as new dependent variable \( z\left (y.t\right ) =v_{t}\left (y,t\right ) \), problem (10 ) rewrites as

$$\displaystyle \begin{aligned} \left\{ \begin{array}{lll} \rho z_{t}=\mu z_{yy}, & & s\left( t\right) <y<H,\ t>0 \\ \\ z\left( H,t\right) =0, & & t>0, \\ \\ \rho z\left( s\left( t\right) ,t\right) =f_{o}-{\displaystyle {{\displaystyle {\tau _{o} \over s\left( t\right) }}}}, & & t>0, \\ \\ z_{y}\left( s\left( t\right) ,t\right) ={\displaystyle {{\displaystyle {\tau _{o} \over s\left( t\right) }}}}\dot s, & & t>0, \\ \\ v\left(y,0\right) =v_{o}\left(y\right) & & 0<y<s_{o}, \\ \\ s\left( 0\right) =s_{o}. & & \end{array} \right. {} \end{aligned} $$
(11)

In [4] global well posedness of problem (11) has been proved.

4 Bingham Model with Deformable Core

The Bingham model predicts that the material behaves as a rigid body if the shear stress is less than the threshold τ o. It is however evident that the schematization of the rigid body is not plausible from a physical point of view (think, for example, to the mayonnaise). Oldroyd [20] and Yoshimura et al. [32] have proposed to treat the “solid phase” as deformable. In [7] Fusi et al. have studied an extension of the 1D problem (10) to the case of an elastic core. The model has been developed within the context of the theory of natural configurations [24, 25]. In [10] Fusi et al. have extended the problem studied in [7] to a 2D channel flow where the channel amplitude is not uniform.

Actually, other models have been proposed for Bingham-like fluids whose unyielded core is deformable. In [8, 9], for instance, the material in the non-fluid region has been modeled as a visco-elastic fluid.

4.1 Channel Flow of a Bingham-Like Fluid with Linear Elastic Core

Here we consider a channel flow of a Bingham-like fluid driven by a known pressure gradient, assuming that the continuum behaves as a linear viscous fluid if the stress is above the yield stress and as a linear elastic solid when the stress is below such a threshold. In particular we assume that the domain may be split in two sub-domains, see Fig. 10. The inner core, in which the material behaves as a linear elastic material, and the outer part (i.e. the one close to the channel walls) where a linear viscous behavior occurs. The two regions are separated by the unknown sharp interfaces y = ±σ(x, t). Moreover the channel width varies along x, so that the channel walls are given by the function y = H(x). The channel is finite and we denote its length by L. As in Sect. 3, we limit our analysis to the upper part of the channel because of symmetry.

Fig. 10
figure 10

A schematic representation of the channel flow with deformable core

The motion equations are obtained by imposing the mass and momentum balance. The evolution equation of the interface y = σ(x, t), as well as the boundary conditions, are derived imposing Rankine-Hugoniot conditions and Von Mises yield criterion. The general mathematical problem is therefore a two-dimensional free boundary problem. We develop the model assuming that the characteristic height of the upper part H is far less than L, i.e. the aspect ratio

$$\displaystyle \begin{aligned}\varepsilon ={\displaystyle {H \over L}}\ll 1\end{aligned}$$

is very small.

4.2 Kinematics and Constitutive Equation

Consider a two dimensional setting and assume that the motion is given by

$$\displaystyle \begin{aligned} \boldsymbol{x}=\boldsymbol{\chi ~}(\boldsymbol{X},t), {} \end{aligned} $$
(12)

where χ is a differentiable and invertible mapping from R 2 →R 2. The vectors x, X are the Eulerian and Lagrangian coordinates, respectively. The deformation tensor isFootnote 3

$$\displaystyle \begin{aligned} \mathbf{F}=\mathrm{grad} \boldsymbol{\chi ~}(\boldsymbol{X},t) \end{aligned}$$

and mechanical incompressibility entails \(\det \mathbf {F}=1\). The Eulerian velocity and acceleration are defined as

$$\displaystyle \begin{aligned} \begin{array}{rcl} \boldsymbol{v}\left( \boldsymbol{x},t\right) &\displaystyle =&\displaystyle \left. {\displaystyle {{\displaystyle { \partial \boldsymbol{\chi } \over \partial t}}}}\right\vert _{\boldsymbol{X}= \boldsymbol{\chi }^{{}^{-1}}(\boldsymbol{x},t)} \\[0.04in] \boldsymbol{a}\left( \boldsymbol{x},t\right) &\displaystyle =&\displaystyle \left. {\displaystyle {{\displaystyle { \partial ^{2}\boldsymbol{\chi } \over \partial t^{{}^{2}}}}}}\right\vert _{ \boldsymbol{X}=\boldsymbol{\chi }^{{}^{-1}}(\boldsymbol{x},t)} \end{array} \end{aligned} $$

respectively. The strain rate tensor is

$$\displaystyle \begin{aligned} \mathbf{D}=\left( \begin{array}{cc} {\displaystyle {{\displaystyle {\partial v_{1} \over \partial x}}}} & {\displaystyle {{\displaystyle { 1 \over 2}}}}\left( {\displaystyle {{\displaystyle {\partial v_{1} \over \partial y}}}}+{ \displaystyle {{\displaystyle {\partial v_{2} \over \partial x}}}}\right) \\ & \\ {\displaystyle {{\displaystyle {1 \over 2}}}}\left( {\displaystyle {{\displaystyle {\partial v_{1} \over \partial y}}}}+{\displaystyle {{\displaystyle {\partial v_{2} \over \partial x}}}} \right) & {\displaystyle {{\displaystyle {\partial v_{2} \over \partial y}}}} \end{array} \right), \end{aligned}$$

and mechanical incompressibility gives

$$\displaystyle \begin{aligned} \mathrm{tr}\:\mathbf{D}={\displaystyle {{\displaystyle {\partial v_{1} \over \partial x}}}}+{ \displaystyle {{\displaystyle {\partial v_{2} \over \partial y}}}}=0. {} \end{aligned} $$
(13)

Splitting, as usual, the Cauchy stress as T = −P I + S, with P = 1∕3trT, we extend we extend the constitutive relation (6), (7) to the case in which the region II S < τ o behaves as a linear elastic materialFootnote 4 (see [7, 10]).

$$\displaystyle \begin{aligned} \left[ \mathbf{S}-\left( 2\mu +{\displaystyle {{\displaystyle {\tau _{o} \over II_{\mathbf{ D}}}}}}\right) \mathbf{D}\right] \varTheta \left( II_{\mathbf{S}}-\tau _{o}\right) +\left( \mathbf{S}-2\eta \mathbf{E}\right) \varTheta \left( \tau _{o}-II_{\mathbf{S}}\right) =0, {} \end{aligned} $$
(14)

where

$$\displaystyle \begin{aligned}\mathbf{E}={\displaystyle {{\displaystyle {1 \over 2}}}}\left[ \nabla \boldsymbol{f} +\left( \nabla \boldsymbol{f}\right) ^{\mathrm{T}}\right] \end{aligned}$$

is the linearized strain tensor, f = χ −X is the displacement , η is the elastic modulus and Θ is the Heaviside function

$$\displaystyle \begin{aligned} \varTheta (z)=\left\{ \begin{array}{ll} 0, & \mathrm{if\;\;}z<0, \\ & \\ 1, & \mathrm{if\;\;}z\geqslant 0. \end{array} \right. \end{aligned}$$

From (14) it is clear that, whenever II S ≤ τ o, the continuum behaves as a linear elastic solid whereas the viscous behavior occurs when II S ≥ τ o.

4.3 Flow in a Channel

We rescale the longitudinal variable as

$$\displaystyle \begin{aligned}\tilde{x}={\displaystyle {x \over L}} \ \ \ \ \ \ \ \ \tilde{h}(x)={\displaystyle {H(x) \over H}} \ \ \ \ \ \ \ \ \ \ \tilde{\sigma}={\displaystyle {\sigma \over H}} \ \ \ \ \ \ \ \ \ \ H=\max_{x\in \lbrack 0,L]}H(x)\end{aligned}$$

and we introduce the Reynolds and Bingham number

$$\displaystyle \begin{aligned} \mathsf{Re}={\displaystyle{{\displaystyle {\rho UH \over \mu }}}},\;\;\;\;\mathsf{Bn}={ \displaystyle{{\displaystyle {\tau _{o}H \over \mu U}}}}, {} \end{aligned} $$
(15)

where U is the characteristic velocity and ρ is the material density. The we set

$$\displaystyle \begin{aligned}\tilde{y}={\displaystyle {1 \over \varepsilon }}{\displaystyle {y \over L}} \ \ \ \ \ \ \tilde{t}={\displaystyle {t \over t_{c}}} \ \ \ \ \ \ t_{c}={\displaystyle {L \over U}}.\end{aligned}$$

Concerning velocity and pressure, we introduce

$$\displaystyle \begin{aligned} \tilde{v}_{1}={\displaystyle{{\displaystyle {v_{1} \over U}}}},\;\;\;\tilde{v}_{2}={\displaystyle {{\displaystyle {v_{2} \over \varepsilon U}}}},\;\;\;\tilde{P}={\displaystyle{{\displaystyle {P \over P_{c}}}}} \ \ \ P_{c}={\displaystyle{{\displaystyle {\mu UL \over H^{2}}}}} \end{aligned}$$

where P c comes from the classical Poiseuille formula. Rescaling S as

$$\displaystyle \begin{aligned} \mathbf{S}={\displaystyle {{\displaystyle {\mu U \over H}}}}\left( \begin{array}{cc} \tilde{S}_{11} & \tilde{S}_{12} \\ & \\ \tilde{S}_{12} & \tilde{S}_{22} \end{array} \right) , \end{aligned}$$

the motion equations becomeFootnote 5

$$\displaystyle \begin{aligned} {\displaystyle {\mathsf{Re} \over \varepsilon }}\left( {\displaystyle {{\displaystyle {\partial v_{1} \over \partial t}}}}+{\displaystyle {{\displaystyle {\partial v_{1} \over \partial x}}}}v_{1}+{ \displaystyle {{\displaystyle {\partial v_{1} \over \partial y}}}}v_{2}\right) =-{\displaystyle {1 \over \varepsilon ^{2}}}{\displaystyle {{\displaystyle {\partial P \over \partial x}}}}+{\displaystyle {1 \over \varepsilon }}{\displaystyle {{\displaystyle {\partial \over \partial x}}}}\left( S_{11}\right) +{\displaystyle {1 \over \varepsilon ^{2}}}{\displaystyle {{\displaystyle {\partial \over \partial y}}}}\left( S_{12}\right) , {} \end{aligned} $$
(16)
$$\displaystyle \begin{aligned} {\displaystyle {\mathsf{Re} \over \varepsilon }}\left( {\displaystyle {{\displaystyle {\partial v_{2} \over \partial t}}}}+{\displaystyle {{\displaystyle {\partial v_{2} \over \partial x}}}}v_{1}+{ \displaystyle {{\displaystyle {\partial v_{2} \over \partial y}}}}v_{2}\right) =-{\displaystyle {1 \over \varepsilon ^{4}}}{\displaystyle {{\displaystyle {\partial P \over \partial y}}}}+{\displaystyle {1 \over \varepsilon ^{2}}}\left[ {\displaystyle {{\displaystyle {\partial \over \partial x}}}}\left( S_{12}\right) +{\displaystyle {1 \over \varepsilon }}{\displaystyle {{\displaystyle {\partial \over \partial y}}}}\left( S_{22}\right) \right] , {} \end{aligned} $$
(17)

and mass conservation is

$$\displaystyle \begin{aligned} {\displaystyle {{\displaystyle {\partial v_{1} \over \partial x}}}}+{\displaystyle {{\displaystyle { \partial v_{2} \over \partial y}}}}=0. {} \end{aligned} $$
(18)

The displacement is rescaled as

$$\displaystyle \begin{aligned}\tilde{f}_{1}={\displaystyle {f_{1} \over L}} \ \ \ \ \ \ \ \ \tilde{f}_{2}={\displaystyle {1 \over \varepsilon }}{\displaystyle {f_{2} \over L}},\end{aligned}$$

so thatFootnote 6

$$\displaystyle \begin{aligned} \mathbf{E}=\left( \begin{array}{cc} {\displaystyle {{\displaystyle {\partial f_{1} \over \partial x}}}} & {\displaystyle {{\displaystyle { 1 \over 2}}}}\left( {\displaystyle {{\displaystyle {1 \over \varepsilon }}}}{\displaystyle { {\displaystyle {\partial f_{1} \over \partial y}}}}+\varepsilon {\displaystyle {{\displaystyle { \partial f_{2} \over \partial x}}}}\right) \\ & \\ {\displaystyle {{\displaystyle {1 \over 2}}}}\left( {\displaystyle {{\displaystyle {1 \over \varepsilon }}} }{\displaystyle {{\displaystyle {\partial f_{1} \over \partial y}}}}+\varepsilon { \displaystyle {{\displaystyle {\partial f_{2} \over \partial x}}}}\right) & {\displaystyle { {\displaystyle {\partial f_{2} \over \partial y}}}} \end{array} \right) . \end{aligned}$$

The sharp interface y = σ(x, t) separates the elastic domain from the viscous domain, as shown in Fig. 2. We remark that the interface σ is unknown and it is not material. The dimensionless normal velocity of the interface is

$$\displaystyle \begin{aligned} w={\displaystyle {\varepsilon \over \sqrt{1+\varepsilon ^{2}\left( {\displaystyle {{\displaystyle {\partial \sigma \over \partial x}}}}\right) ^{2}} }}{\displaystyle {{\displaystyle {\partial \sigma \over \partial t}}}}. \end{aligned}$$

The dimensionless strain rate tensor is

$$\displaystyle \begin{aligned} \mathbf{D}= \left( \begin{array}{cc} \varepsilon {\displaystyle {{\displaystyle {\partial v_{1} \over \partial x}}}} & { \displaystyle {{\displaystyle {1 \over 2}}}}\left( {\displaystyle {{\displaystyle {\partial v_{1} \over \partial y}}}}+\varepsilon ^{2}{\displaystyle {{\displaystyle {\partial v_{2} \over \partial x}}}}\right) \\ & \\ {\displaystyle {{\displaystyle {1 \over 2}}}}\left( {\displaystyle {{\displaystyle {\partial v_{1} \over \partial y}}}}+\varepsilon ^{2}{\displaystyle {{\displaystyle {\partial v_{2} \over \partial x}}}}\right) & \varepsilon {\displaystyle {{\displaystyle {\partial v_{2} \over \partial y}}}} \end{array} \right) , \end{aligned}$$

while the dimensionless invariant II D is

$$\displaystyle \begin{aligned} II_{\mathbf{D}}={\displaystyle {1 \over 2}}\sqrt{\left( {\displaystyle {\partial v_{1} \over \partial y}} +\varepsilon ^{2}{\displaystyle {\partial v_{2} \over \partial x}}\right) ^{2}+2\varepsilon ^{2}\left[ \left( {\displaystyle {{\displaystyle {\partial v_{1} \over \partial x}}}} \right) ^{2}+\left( {\displaystyle {{\displaystyle {\partial v_{2} \over \partial y}}}} \right) ^{2}\right] }. {} \end{aligned} $$
(19)

The yield criterion becomes

$$\displaystyle \begin{aligned} II_{\mathbf{S}}\lesseqgtr \mathsf{Bn}. {} \end{aligned} $$
(20)

4.3.1 Boundary Conditions

Let denotes the “jump” across the surface y = σ. Assuming no-slip on σ we get

(21)

The continuity of the stress gives

(22)

and

(23)

Remark 1

If we neglect \(\mathcal {O}\left ( \varepsilon \right ) \) terms, then , provided that S ij are bounded.

On the channel wall y = h the no-slip condition yields \(\boldsymbol {v}\left ( x,h ,t\right ) \equiv 0\), while the boundary conditions for the pressure are

$$\displaystyle \begin{aligned} P\left( 0,y,t\right) =P_{in}\left( t\right) ,\ \ \ \mathrm{and}\ \ \ P\left( 1,y,t\right) =P_{in}\left( t\right) -\varDelta P\left( t\right) ,\ \ \mathrm{ with}\ \ \varDelta P\geq 0. \end{aligned}$$

Finally, we impose the symmetry conditions:

$$\displaystyle \begin{aligned} {\displaystyle {{\displaystyle {\partial f_{1}\left( x,0,t\right) \over \partial y}}}} =0,\;\;\;\;\;f_{2}\left( x,0,t\right) =0,\;\;\;\;\;\;{\displaystyle {{\displaystyle { \partial f_{2}\left( x,0,t\right) \over \partial y}}}}=0,\;\;\;\;\;\;{ \displaystyle {{\displaystyle {\partial f_{2}\left( x,0,t\right) \over \partial t}}}}=0. {} \end{aligned} $$
(24)

4.3.2 The Elastic Domain and the Viscous Domain

Let us first consider the elastic domain y ∈ [0, σ]. According to (14)

$$\displaystyle \begin{aligned} \mathbf{S}={\displaystyle {\mathsf{Re} \over \lambda ^{2}}}\left( \begin{array}{cc} 2{\displaystyle {{\displaystyle {\partial f_{1} \over \partial x}}}} & {\displaystyle { {\displaystyle {1 \over \varepsilon }}}}{\displaystyle {{\displaystyle {\partial f_{1} \over \partial y}}}} +\varepsilon {\displaystyle {{\displaystyle {\partial f_{2} \over \partial x}}}} \\ & \\ {\displaystyle {{\displaystyle {1 \over \varepsilon }}}}{\displaystyle {{\displaystyle {\partial f_{1} \over \partial y}}}}+\varepsilon {\displaystyle {{\displaystyle {\partial f_{2} \over \partial x}}}} & 2{\displaystyle {{\displaystyle {\partial f_{2} \over \partial y}}}} \end{array} \right) \mathbf{,} \end{aligned}$$

with

$$\displaystyle \begin{aligned} \lambda ^{2}=\left( {\displaystyle {U \over c}}\right) ^{2}\ \ \ \ \ \ \ \ c^{~2}={\displaystyle {\eta \over \rho }}. {} \end{aligned} $$
(25)

The momentum balance yields

$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle &\displaystyle \mathsf{Re}\left( {\displaystyle {{\displaystyle {\partial ^{2}f_{1} \over \partial t^{2}}}} }+{\displaystyle {{\displaystyle {\partial ^{2}f_{1} \over \partial x\partial t}}}}{ \displaystyle {{\displaystyle {\partial f_{1} \over \partial t}}}}+{\displaystyle {{\displaystyle { \partial ^{2}f_{1} \over \partial y\partial t}}}}{\displaystyle {{\displaystyle {\partial f_{2} \over \partial t}}}}\right) \\ &\displaystyle &\displaystyle \quad =-{\displaystyle {1 \over \varepsilon }}{\displaystyle {{\displaystyle { \partial P \over \partial x}}}} +{\displaystyle {\mathsf{Re} \over \lambda ^{2}}}\left( 2{\displaystyle {{\displaystyle {\partial ^{2}f_{1} \over \partial x^{2}}}}}+{\displaystyle {1 \over \varepsilon ^{2}}}{\displaystyle { {\displaystyle {\partial ^{2}f_{1} \over \partial y^{2}}}}}+{\displaystyle {1 \over \varepsilon }}{ \displaystyle {{\displaystyle {\partial ^{2}f_{2} \over \partial y\partial x}}}}\right) , {} \end{array} \end{aligned} $$
(26)
$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle &\displaystyle \mathsf{Re}\left( {\displaystyle {{\displaystyle {\partial ^{2}f_{2} \over \partial t^{2}}}} }+{\displaystyle {{\displaystyle {\partial ^{2}f_{2} \over \partial x\partial t}}}}{ \displaystyle {{\displaystyle {\partial f_{1} \over \partial t}}}}+{\displaystyle {{\displaystyle { \partial ^{2}f_{2} \over \partial y\partial t}}}}{\displaystyle {{\displaystyle {\partial f_{2} \over \partial t}}}}\right) \\ &\displaystyle &\displaystyle \quad =-{\displaystyle {1 \over \varepsilon ^{3}}}{\displaystyle { {\displaystyle {\partial P \over \partial y}}}} +{\displaystyle {\mathsf{Re} \over \lambda ^{2}}}\left( {\displaystyle {{\displaystyle {1 \over \varepsilon ^{2}}}}}{\displaystyle {{\displaystyle {\partial ^{2}f_{1} \over \partial x\partial y}}}}+{\displaystyle {{\displaystyle {\partial ^{2}f_{2} \over \partial x^{2}}}}}+ {\displaystyle {2 \over \varepsilon ^{2}}}{\displaystyle {{\displaystyle {\partial ^{2}f_{2} \over \partial y^{2}}}}}\right). {} \end{array} \end{aligned} $$
(27)

In the viscous domain y ∈ [σ, 1]

$$\displaystyle \begin{aligned} \mathbf{S}=\left( 1+{\displaystyle {\mathsf{Bn} \over 2II_{\mathbf{D}}}}\right) \left( \begin{array}{cc} 2\varepsilon {\displaystyle{{\displaystyle {\partial v_{1} \over \partial x}}}} & { \displaystyle{{\displaystyle {\partial v_{1} \over \partial y}}}}+\varepsilon ^{2}{ \displaystyle{{\displaystyle {\partial v_{2} \over \partial x}}}} \\ & \\ {\displaystyle{{\displaystyle {\partial v_{1} \over \partial y}}}}+\varepsilon ^{2}{ \displaystyle{{\displaystyle {\partial v_{2} \over \partial x}}}} & 2\varepsilon { \displaystyle{{\displaystyle {\partial v_{2} \over \partial y}}}} \end{array} \right) , \end{aligned}$$

The momentum equations are

$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle &\displaystyle {\displaystyle {\mathsf{Re} \over \varepsilon }}\left( {\displaystyle{{\displaystyle {\partial v_{1} \over \partial t}}}}+{\displaystyle{{\displaystyle {\partial v_{1} \over \partial x}}}}v_{1}+{ \displaystyle{{\displaystyle {\partial v_{1} \over \partial y}}}}v_{2}\right) \\ &\displaystyle &\displaystyle \quad =-{\displaystyle {1 \over \varepsilon ^{2}}}{\displaystyle{{\displaystyle {\partial P \over \partial x}}}}+2{ \displaystyle{{\displaystyle {\partial \over \partial x}}}}\left[ \left( 1+{\displaystyle {\mathsf{Bn} \over 2II_{\mathbf{D}}}}\right) {\displaystyle{{\displaystyle {\partial v_{1} \over \partial x}}}} \right] \\ &\displaystyle &\displaystyle \qquad +{\displaystyle {1 \over \varepsilon ^{2}}}{\displaystyle{{\displaystyle {\partial \over \partial y}}}} \left[ \left( 1+{\displaystyle {\mathsf{Bn} \over 2II_{\mathbf{D}}}}\right) \left( { \displaystyle{{\displaystyle {\partial v_{1} \over \partial y}}}}\right. \right. \left. \left. \varepsilon ^{2}{\displaystyle{{\displaystyle {\partial v_{2} \over \partial x}}}}\right) \right] , {} \end{array} \end{aligned} $$
(28)
$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle &\displaystyle {\displaystyle {\mathsf{Re} \over \varepsilon }}\left( {\displaystyle{{\displaystyle {\partial v_{2} \over \partial t}}}}+{\displaystyle{{\displaystyle {\partial v_{2} \over \partial x}}}}v_{1}+{ \displaystyle{{\displaystyle {\partial v_{2} \over \partial y}}}}v_{2}\right) =-{\displaystyle {1 \over \varepsilon ^{4}}}{\displaystyle{{\displaystyle {\partial P \over \partial y}}}} \\ &\displaystyle &\displaystyle \quad +{\displaystyle {1 \over \varepsilon ^{2}}}\left\{ {\displaystyle{{\displaystyle {\partial \over \partial x}}}}\left[ \left( 1+{\displaystyle {\mathsf{Bn} \over 2II_{\mathbf{D}}}}\right) \left( {\displaystyle{{\displaystyle {\partial v_{1} \over \partial y}}}}\right. \right. \right. +\left. \left. \varepsilon ^{2}{\displaystyle{{\displaystyle {\partial v_{2} \over \partial x}}}}\right) \right] +\left. 2{\displaystyle{{\displaystyle {\partial \over \partial y}}}}\left[ \left( 1+ {\displaystyle {\mathsf{Bn} \over 2II_{\mathbf{D}}}}\right) {\displaystyle{{\displaystyle {\partial v_{2} \over \partial y}}}}\right] \right\} . \\ {} \end{array} \end{aligned} $$
(29)

The model is consistent if II S < B n for y ∈ [0, σ], i.e.

$$\displaystyle \begin{aligned} \sqrt{\left( {\displaystyle {{\displaystyle {\partial f_{1} \over \partial y}}}} +\varepsilon ^{2}{\displaystyle {{\displaystyle {\partial f_{2} \over \partial x}}}}\right) ^{2}+2\varepsilon ^{2}\left[ \left( {\displaystyle {{\displaystyle {\partial f_{1} \over \partial x}}}}\right) ^{2}+\left( {\displaystyle {{\displaystyle {\partial f_{2} \over \partial y}}}}\right) ^{2}\right] }\leq {\displaystyle {\mathsf{Bn}\lambda ^{2}\varepsilon \over \mathsf{Re}}}. {} \end{aligned} $$
(30)

and if II S ≥B n for y ∈ [σ, 1], i.e. II D ≥ 0. Finally we notice that

$$\displaystyle \begin{aligned}- {\displaystyle {\mathsf{Re} \over \lambda ^{2} }}\left[ {\displaystyle {{\displaystyle {\partial f_{1} \over \partial y}}}}+\varepsilon ^{2} {\displaystyle {{\displaystyle {\partial f_{2} \over \partial x}}}}\right] _{y=\sigma ^{-}}. {} \end{aligned} $$
(31)

4.3.3 Asymptotic Expansion

The assumption ε ≪ 1, allows to seek the unknown fields (i.e. \(v_{1}^{\,}\), v 2, P, etc.) in the following form

$$\displaystyle \begin{aligned} \phi=\sum_{j=0}^{\infty}\phi^{(j)}\varepsilon^{j} \end{aligned}$$

Substituting into the governing equations we get a hierarchy of problems that must be matched together. The matching procedure requires the specification of Re, Bi, and λ 2. We consider: \(\mathsf {Re}\leq \mathcal {O}\left ( 1\right ) ,\) i.e. laminar flow and \(\mathsf {Bn}=\mathcal {O}\left ( 1\right ) \). We introduce the dimensionless parameter

$$\displaystyle \begin{aligned} \varGamma ={\displaystyle {{\displaystyle {\eta H \over \mu U}}}}={\displaystyle {{\displaystyle { \mathsf{Re} \over \lambda ^{2}}}}}, {} \end{aligned} $$
(32)

and consider two different cases which correspond to different behaviors of the inner core:

  1. (1)

    \(\varGamma =\mathcal {O}(1).\)

  2. (2)

    \(\varGamma =\mathcal {O}(\varepsilon ).\)

When \(\varGamma =\mathcal {O}\left ( 1\right ) \), namely \(\mathsf {Re}= \mathcal {O}\left ( \lambda ^{2}\right ) \), we have an almost rigid inner core, while the case \(\varGamma =\mathcal {O}\left ( \varepsilon \right ) \), i.e. \( \mathsf {Re}=\mathcal {O}\left ( \varepsilon \lambda ^{2}\right ) \), corresponds to a “ soft” inner core where deformations are non-negligible. Let us show that the first case leads to the classical Bingham model. Indeed, considering \(\mathsf {Re}=\mathcal {O}\left ( \lambda ^{2}\right ) \)

$$\displaystyle \begin{aligned} \mathcal{O}\left( 1\right) =\varGamma =\varepsilon {\displaystyle {t_{c}\eta \over \mu }},\ \ \ \Rightarrow \ \ \ {\displaystyle {t_{c}\eta \over \mu }}=\mathcal{O}\left( \varepsilon ^{-1}\right) . {} \end{aligned} $$
(33)

Evaluating the order of magnitude of the elastic stress S el and of the viscous stress S vis we have

$$\displaystyle \begin{aligned}S_{vis}=\mu {\displaystyle {U \over H}} \ \ \ \ \ \ \ \ \ S_{el}=\delta \eta {\displaystyle {L \over H}},\end{aligned}$$

where δ is the dimensionless order of magnitude of the longitudinal displacement. The continuity of the shear stress at the interface yields

$$\displaystyle \begin{aligned} {\displaystyle {S_{el} \over S_{vis}}}=\mathcal{O}\left( 1\right) \ \ \ \ \ \ \ \ \Rightarrow \ \ \ \ \ \ \ \delta {\displaystyle {t_{c}\eta \over \mu }}=\mathcal{O}\left( 1\right) . {} \end{aligned} $$
(34)

Hence we find \(\delta =\mathcal {O}\left ( \varepsilon \right )\), meaning an almost uniform longitudinal displacement as in the classical Bingham model. In the second case

$$\displaystyle \begin{aligned}\mathsf{Re}=\mathcal{O}(\varepsilon \lambda ^{2}) \ \ \ \ \ \ \ \ \Rightarrow \ \ \ \ \ \ \ {\displaystyle {t_{c}\eta \over \mu }}=\mathcal{O}\left( 1\right) \end{aligned}$$

which entails \(\delta =\mathcal {O}\left ( 1\right ) \). As a consequence the longitudinal displacement is not uniform.

4.3.4 First Case: \(\varGamma =\mathcal {O}(1)\)

We start considering the elastic domain, and we focus on the zero order terms. From (27) we obtain \(P^{\left ( 0\right ) }=P^{\left ( 0\right ) }\left ( x,t\right ) \). Next, since on the interface, we have that in the whole elastic region \(P^{\left ( 0\right ) }=P^{\left ( 0\right ) }\left ( x,\sigma ^{+},t\right ) \). From (26) we have

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} {\displaystyle {{\displaystyle {\partial ^{2}f_{1}^{~\left( 0\right) } \over \partial y^{2} }}}}=0, \\ \\ \left. {\displaystyle {{\displaystyle {\partial f_{1}^{~\left( 0\right) } \over \partial y}} }}\right\vert _{y=0}=0, \end{array} \right. \ \ \Longrightarrow \ \ f_{1}^{~\left( 0\right) }=f_{1}^{~\left( 0\right) }\left( x,t\right) . \end{aligned}$$

Mass conservation in the elastic domain becomes

$$\displaystyle \begin{aligned} {\displaystyle {\partial ^{2}f_{1}^{~\left( 0\right) } \over \partial x\partial t}}+{\displaystyle { \partial ^{2}f_{2}^{~\left( 0\right) } \over \partial y\partial t}}=0. {} \end{aligned} $$
(35)

Recalling (24) it is easy to show that

$$\displaystyle \begin{aligned} {\displaystyle {\partial \over \partial x}}\left( {\displaystyle {\partial f_{1}^{~\left( 0\right) } \over \partial t}}\right) =0. {} \end{aligned} $$
(36)

Therefore setting

$$\displaystyle \begin{aligned} \kappa ={\displaystyle {\partial f_{1}^{~\left( 0\right) } \over \partial t}}, \end{aligned}$$

we have \(\kappa =\kappa \left ( t\right ) \). From (21) we find

$$\displaystyle \begin{aligned} \kappa =v_{1}^{\left( 0\right) }\left( x,\sigma ^{+},t\right) . {} \end{aligned} $$
(37)

From (35)

$$\displaystyle \begin{aligned} {\displaystyle {\partial f_{2}^{~\left( 0\right) } \over \partial t}}\ \ \text{depends}\ \text{only}\ \text{on}\ t. \end{aligned}$$

and the transversal velocity vanishes everywhere in the elastic region, so that

$$\displaystyle \begin{aligned} v_{2}^{\left( 0\right) }\left( x,\sigma ^{+},t\right) =0. {} \end{aligned} $$
(38)

Let us now consider relation (31) at the zero order approximation. We get

$$\displaystyle \begin{aligned} {\displaystyle {\mathsf{Re} \over \lambda ^{2}}}\left. {\displaystyle {{\displaystyle {\partial f_{1}^{~\left( 1\right) } \over \partial y}}}}\right\vert _{y=\sigma ^{-}}=\left[ \left( 1+{\displaystyle {\mathsf{Bn} \over 2II_{\mathbf{D}}^{\left( 0\right) }}}\right) { \displaystyle {{\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial y}}}}\right] _{y=\sigma ^{+}}. {} \end{aligned} $$
(39)

The latter forces to analyze the first order term in (26), (27), namely

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} {\displaystyle {{\displaystyle {\partial ^{2}f_{1}^{~\left( 1\right) } \over \partial y^{2} }}}}={\displaystyle {{\displaystyle {\lambda ^{2} \over \mathsf{Re}}}}}{\displaystyle {{\displaystyle { \partial P^{\left( 0\right) } \over \partial x}}}}, \\ \\ \left. {\displaystyle {{\displaystyle {\partial f_{1}^{~\left( 1\right) } \over \partial y}} }}\right\vert _{y=0}=0, \end{array} \right. \ \ \Longrightarrow \ \ f_{1}^{~\left( 1\right) }={\displaystyle { {\displaystyle {\lambda ^{2} \over 2~\mathsf{Re}}}}}{\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}y^{2}+B\left( x,t\right) . \end{aligned}$$

Hence (39) becomes

$$\displaystyle \begin{aligned} {\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}\sigma = \left[ \left( 1+{\displaystyle {\mathsf{Bn} \over 2II_{\mathbf{D}}^{\left( 0\right) }}} \right) {\displaystyle {{\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial y}} }}\right] _{y=\sigma ^{+}}. {} \end{aligned} $$
(40)

It is easy to show that

$$\displaystyle \begin{aligned}II_{\mathbf{S}}=\left\vert {\displaystyle {{\displaystyle { \partial P^{\left( 0\right) } \over \partial x}}}}\right\vert y+\mathcal{O}\left( \varepsilon \right)\end{aligned}$$

in the elastic region so that condition (30) is fulfilled when

$$\displaystyle \begin{aligned} \left\vert {\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}} }\right\vert \sigma \leq \mathsf{Bn}. {} \end{aligned} $$
(41)

We now focus on the viscous part σ < y < 1. Here (29) entails \(P^{\left ( 0\right ) }=P^{\left ( 0\right ) }\left ( x,t\right ) \), so that the pressure is uniform on any channel section. Next, we observe that

$$\displaystyle \begin{aligned}II_{\mathbf{S}}=\left\vert {\displaystyle {{\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial y}}}}+\mathsf{Bi~}\text{sign}\left( {\displaystyle { {\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial y}}}}\right) \right\vert + \mathcal{O}\left( \varepsilon \right) \end{aligned}$$

and (39) can be rewritten as

$$\displaystyle \begin{aligned} \underset{II_{\mathbf{S}}\leq \mathsf{Bn}}{\underbrace{\left\vert { \displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}} \right\vert \sigma }}=\underset{II_{\mathbf{S}}\geq \mathsf{Bn}}{\underbrace{ \left\vert \left. {\displaystyle {{\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial y}}}}\right\vert _{y=\sigma ^{+}}+\mathsf{Bi~}\mathrm{sign}\left( \left. {\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial y}}\right\vert _{y=\sigma ^{+}}\right) \right\vert }}. \end{aligned}$$

We conclude that

$$\displaystyle \begin{aligned} \left. {\displaystyle {{\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial y}}} }\right\vert _{y=\sigma ^{+}}=0\ \ \ \ \ \ \ \ \ {\displaystyle { {\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}\sigma =\mathsf{Bi~} \mathrm{sign}\left( \left. {\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial y}}\right\vert _{y=\sigma ^{+}}\right) . {} \end{aligned} $$
(42)

Focussing now on (28) we find

$$\displaystyle \begin{aligned} -{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}+{\displaystyle {\partial \over \partial y }}\left[ {\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial y}}+\mathsf{Bi~} \mathrm{sign}\left( {\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial y}} \right) \right] =0, \end{aligned}$$

so that

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} -{\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}+{ \displaystyle {{\displaystyle {\partial ^{2}v_{1}^{\left( 0\right) } \over \partial y^{2}}}} }=0, \\ \\ v_{1}^{\left( 0\right) }\left( x,h,t\right) =0, \\ \\ \left. {\displaystyle {{\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial y}}} }\right\vert _{y=\sigma ^{+}}=0, \end{array} \right. {} \end{aligned} $$
(43)

Solving (43) we get

$$\displaystyle \begin{aligned} v_{1}^{\left( 0\right) }\left( x,y,t\right) =-{\displaystyle {1 \over 2}}{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}\left( h^{2}-y^{2}\right) +\mathsf{Bi~}\mathrm{sign}\left( \left. {\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial y}}\right\vert _{y=\sigma ^{+}}\right) \left( h-y\right) . {} \end{aligned} $$
(44)

In particular, (37) yields

$$\displaystyle \begin{aligned} \kappa =-{\displaystyle {1 \over 2}}{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}\left( h-\sigma \right) ^{2},\ \ \ \text{with}\ \ \ \omega =\omega \left( t\right) . {} \end{aligned} $$
(45)

Remark 2

We remark that

$$\displaystyle \begin{aligned} \kappa =-{\displaystyle {\mathsf{Bn} \over 2\sigma }}\mathrm{sign}\left( \left. {\displaystyle { \partial v_{1}^{\left( 0\right) } \over \partial y}}\right\vert _{y=\sigma ^{+}}\right) \left( h-\sigma \right) ^{2}. {} \end{aligned} $$
(46)

Therefore the derivative w.r.t. x of the r.h.s. of (46) must vanish and

$$\displaystyle \begin{aligned} {\displaystyle {\partial h \over \partial x}}={\displaystyle {h+\sigma \over 2\sigma }}{\displaystyle {\partial \sigma \over \partial x}}. {} \end{aligned} $$
(47)

From (18)

$$\displaystyle \begin{aligned} \int_{\sigma }^{h}{\displaystyle {{\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial x}}}}dy=0, {} \end{aligned} $$
(48)

since \(v_{2}^{\left ( 0\right ) }\left ( x,h,t\right ) =v_{2}^{\left ( 0\right ) }\left ( x,\sigma ^{+},t\right ) =0\). Hence, exploiting (44)

$$\displaystyle \begin{aligned} {\displaystyle {{\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial x}}}}= -{\displaystyle {1 \over 2}}{\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial x^{2}}} \left( h^{2}-y^{2}\right) -{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}} {\displaystyle {\partial h \over \partial x}}h+{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}\sigma {\displaystyle {\partial h \over \partial x}}. \end{aligned}$$

Thus (48) yields

$$\displaystyle \begin{aligned} \left( h-\sigma \right) ^{2}\left[ {\displaystyle {1 \over 6}}{\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial x^{2}}}\left( \sigma +2h\right) +{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}{\displaystyle {\partial h \over \partial x}}\right] =0. {} \end{aligned} $$
(49)

In case h is uniform we get

$$\displaystyle \begin{aligned} {\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial x^{2}}}=0 \ \ \ \ \ \Rightarrow \ \ \ \ \ \ \ P^{\left( 0\right) }\left( x,t\right) =P_{in}\left( t\right) -x\varDelta P\left( t\right) , \end{aligned}$$

and

$$\displaystyle \begin{aligned} \sigma \left( t\right) ={\displaystyle {\mathsf{Bn} \over \varDelta P\left( t\right) }}, \end{aligned}$$

from which we derive the classical Bingham flow condition we find in [4]

$$\displaystyle \begin{aligned} {\displaystyle {\mathsf{Bn} \over \varDelta P\left( t\right) }}<h, \end{aligned}$$

ensuring the flow within the channel.

Remark 3

In case h does depend on x (49) does not give rise to any solution consistent with (47) (a well known paradox of the Bingham model). Indeed, form (42)2

$$\displaystyle \begin{aligned} {\displaystyle {{\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial x^{2}}}}}=- {\displaystyle {1 \over \sigma }}{\displaystyle {{\displaystyle {\partial \sigma \over \partial x}}}}{ \displaystyle {{\displaystyle {\partial P^{(0)} \over \partial x}}}}, \end{aligned}$$

that inserted into (49) yields

$$\displaystyle \begin{aligned} {\displaystyle {{\displaystyle {\partial P^{(0)} \over \partial x}}}}(h-\sigma )^{2}\left[ {\displaystyle {\partial h \over \partial x}}-{\displaystyle {\partial \sigma \over \partial x}}{\displaystyle { \sigma +2h \over 6\sigma }}\right] =0. \end{aligned}$$

The above implies

$$\displaystyle \begin{aligned} {\displaystyle {\partial \sigma \over \partial x}}{\displaystyle {\sigma +2h \over 6\sigma }}={\displaystyle { \partial h \over \partial x}} {} \end{aligned} $$
(50)

Inserting (50) into (47) we get h = −2σ, that is a contradiction (lubrication paradox, see [6, 17, 21, 27]).

4.3.5 Second Case \(\varGamma =\mathcal {O}( \varepsilon )\)

Set \(\varepsilon \hat {\varGamma }=\varGamma \) with \( \hat {\varGamma }=\mathcal {O}\left ( 1\right ) \). From (31)

$$\displaystyle \begin{aligned} \left[ \left( 1+{\displaystyle {\mathsf{Bn} \over 2II_{\mathbf{D}}}}\right) \left( { \displaystyle {{\displaystyle {\partial v_{1} \over \partial y}}}}+\varepsilon ^{2}{ \displaystyle {{\displaystyle {\partial v_{2} \over \partial x}}}}\right) \right] _{y=\sigma ^{+}}=\hat{\varGamma}\left[ {\displaystyle {{\displaystyle {\partial f_{1} \over \partial y}}}}+\varepsilon ^{2}{\displaystyle {{\displaystyle {\partial f_{2} \over \partial x}}}}\right] _{y=\sigma ^{-}}, {} \end{aligned} $$
(51)

Once again \(P^{\left ( 0\right ) }=P^{\left ( 0\right ) }\left ( x,t\right ) \) in the whole domain. In the elastic part

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} {\displaystyle {{\displaystyle {\partial ^{2}f_{1}^{~\left( 0\right) } \over \partial y^{2} }}}}={\displaystyle {{\displaystyle {1 \over \hat{\varGamma}}}}}{\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}, \\ \\ \left. {\displaystyle {{\displaystyle {\partial f_{1}^{~\left( 0\right) } \over \partial y}} }}\right\vert _{y=0}=0, \end{array} \right. \ \ \Rightarrow \ \ f_{1}^{~\left( 0\right) }={\displaystyle {{\displaystyle { 1 \over 2~\hat{\varGamma}}}}}{\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}y^{2}+a\left( x,t\right) , {} \end{aligned} $$
(52)

so that (51) implies

$$\displaystyle \begin{aligned} {\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}\sigma = \left[ \left( 1+{\displaystyle {\mathsf{Bn} \over 2II_{\mathbf{D}}^{(0)}}}\right) { \displaystyle {{\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial y}}}}\right] _{y=\sigma ^{+}}. \end{aligned}$$

The term \(a\left ( x,t\right ) \) is unknown at this stage while \(v_{1}^{\left ( 0\right ) }\) is

$$\displaystyle \begin{aligned} v_{1}^{\left( 0\right) }={\displaystyle {{\displaystyle {1 \over 2~\hat{\varGamma}}}}}{ \displaystyle {{\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial t\partial x }}}}y^{2}+\kappa \left( x,t\right) , \end{aligned}$$

where now

$$\displaystyle \begin{aligned} \kappa \left( x,t\right) ={\displaystyle {\partial a\left( x,t\right) \over \partial t}}, {} \end{aligned} $$
(53)

can be interpreted as the uniform part of the longitudinal velocity. Checking condition (30) we find again (41).

Remark 4

At the leading order the longitudinal displacement is a superposition of a uniform displacement \(a\left ( x,t\right )\) and a non uniform displacement modulated by the pressure gradient. The latter becomes negligible for large values of \(\hat {\varGamma }\) and the first case is recovered for \( \hat {\varGamma }\gg 1\).

In the fluid region

$$\displaystyle \begin{aligned} {\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}\sigma =- \mathsf{Bn}. \end{aligned}$$

Moreover

$$\displaystyle \begin{aligned} v_{1}^{\left( 0\right) }\left( x,y,t\right) =-{\displaystyle {1 \over 2}}{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}\left( h^{2}-y^{2}\right) +{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}\sigma \left( h-y\right) . \end{aligned}$$

The jump condition (21) yields

$$\displaystyle \begin{aligned} {\displaystyle {{\displaystyle {1 \over 2~\hat{\varGamma}}}}}{\displaystyle {{\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial t\partial x}}}}\sigma ^{2}+\omega \left( x,t\right) =-{\displaystyle {1 \over 2}}{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}} \left( h-\sigma \right) ^{2}, \end{aligned}$$

with ω given by (53). From mass balance and boundary conditions

$$\displaystyle \begin{aligned} 0=\int_{0}^{h}{\displaystyle {\partial v_{1}^{\left( 0\right) } \over \partial x}} dy=\int_{0}^{\sigma }{\displaystyle {\partial \over \partial x}}\left( {\displaystyle {\partial f_{1}^{~\left( 0\right) } \over \partial t}}\right) dy+\int_{\sigma }^{h}{\displaystyle { \partial v_{1}^{\left( 0\right) } \over \partial x}}dy, \end{aligned}$$

which, after some of algebra, gives

$$\displaystyle \begin{aligned} \left( h-\sigma \right) ^{2}\left[ {\displaystyle {{\displaystyle {1 \over 6}}}}{ \displaystyle {{\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial x^{2}}}}} \left( 2h+\sigma \right) +{\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}{\displaystyle {{\displaystyle {\partial h \over \partial x}}}}\right] -{ \displaystyle {{\displaystyle {\sigma ^{3} \over 6~\hat{\varGamma}}}}}{\displaystyle {{\displaystyle { \partial ^{3}P^{\left( 0\right) } \over \partial x^{2}\partial t}}}}-\sigma { \displaystyle {{\displaystyle {\partial \omega \over \partial x}}}}=0. \end{aligned}$$

Therefore the mathematical problem at the zero order approximation is the following

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} {\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}\sigma =- \mathsf{Bn}, \\ \\ {\displaystyle {{\displaystyle {1 \over 2~\hat{\varGamma}}}}}{\displaystyle {{\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial t\partial x}}}}\sigma ^{2}+\kappa \left( x,t\right) =-{\displaystyle {{\displaystyle {1 \over 2}}}}{\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}\left( h-\sigma \right) ^{2}, \\ \\ -\left( h-\sigma \right) ^{2}\left[ {\displaystyle {{\displaystyle {1 \over 6}}}}{ \displaystyle {{\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial x^{2}}}}} \left( 2h+\sigma \right) +{\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}{\displaystyle {{\displaystyle {\partial h \over \partial x}}}}\right] \\ \ \ \ \ \ \ \ \ +{\displaystyle {{\displaystyle {\sigma ^{3} \over 6~\hat{\varGamma}}}}}{ \displaystyle {{\displaystyle {\partial \over \partial t}}}}\left( {\displaystyle {{\displaystyle { \partial ^{2}P^{\left( 0\right) } \over \partial x^{2}}}}}\right) +\sigma { \displaystyle {{\displaystyle {\partial \kappa \over \partial x}}}}=0. \end{array} \right. {} \end{aligned} $$
(54)

Example 1

Consider the stationary problem when h ≡ 1

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} {\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}\sigma =- \mathsf{Bn},\ \ \Rightarrow \ \ {\displaystyle {{\displaystyle {\partial \sigma \over \partial x}}}}{\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x }}}}=-{\displaystyle {{\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial x^{2} }}}}\sigma , \\ \\ \kappa \left( x,t\right) =-{\displaystyle {{\displaystyle {1 \over 2}}}}{\displaystyle { {\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}\left( 1-\sigma \right) ^{2},\ \ \Rightarrow \ \ {\displaystyle {{\displaystyle {\partial \kappa \over \partial x}}} }=-{\displaystyle {{\displaystyle {1 \over 2}}}}{\displaystyle {{\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial x^{2}}}}}\left( 1-\sigma ^{2}\right) , \\ \\ -{\displaystyle {{\displaystyle {\left( 1-\sigma \right) ^{2} \over 6}}}}{\displaystyle { {\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial x^{2}}}}}\left( 2+\sigma \right) +\sigma {\displaystyle {{\displaystyle {\partial \kappa \over \partial x}}}}=0. \end{array} \right. {} \end{aligned} $$
(55)

We get

$$\displaystyle \begin{aligned} {\displaystyle {1 \over 3}}{\displaystyle {{\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial x^{2}}}}}\left( 1-\sigma ^{3}\right) =0, \end{aligned}$$

yielding σ = 1 (which we do not consider) and \(P^{\left ( 0\right ) }=P_{in}-\varDelta Px\), with ΔP known. Therefore

$$\displaystyle \begin{aligned} \begin{array}{rcl} \sigma &\displaystyle =&\displaystyle ={\displaystyle {\mathsf{Bn} \over \varDelta P}}, {} \end{array} \end{aligned} $$
(56)
$$\displaystyle \begin{aligned} \begin{array}{rcl} \kappa &\displaystyle =&\displaystyle {\displaystyle {\varDelta P \over 2}} \left( 1-{\displaystyle {\mathsf{Bn} \over \varDelta P}}\right) ^{2}. {} \end{array} \end{aligned} $$
(57)

Requiring σ < 1 we get the usual Bingham flow condition.

4.3.6 Stationary Version of (54)

We show that the stationary version of system (54), namely

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} {\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}\sigma =- \mathsf{Bn}, \\ \\ \omega \left( x,t\right) =-{\displaystyle {{\displaystyle {1 \over 2}}}}{\displaystyle { {\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}\left( h-\sigma \right) ^{2}, \\ \\ -\left( h-\sigma \right) ^{2}\left[ {\displaystyle {{\displaystyle {1 \over 6}}}}{ \displaystyle {{\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial x^{2}}}}} \left( 2h+\sigma \right) +{\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}{\displaystyle {{\displaystyle {\partial h \over \partial x}}}}\right] +\sigma {\displaystyle {{\displaystyle {\partial \omega \over \partial x}}}}=0, \end{array} \right. {} \end{aligned} $$
(58)

admits a unique solution when h is a smooth bounded function of x with h ∈ [h m, h M]. From (58)1

$$\displaystyle \begin{aligned} {\displaystyle {{\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial x^{2}}}}}= {\displaystyle {\mathsf{Bn} \over \sigma ^{2}}}{\displaystyle {{\displaystyle {\partial \sigma \over \partial x}}}}, {} \end{aligned} $$
(59)

and from (58)2

$$\displaystyle \begin{aligned} \kappa \left( x,t\right) ={\displaystyle {{\displaystyle {\mathsf{Bn} \over 2~\sigma }}}} \left( h-\sigma \right) ^{2},\ \ \mathrm{and}\ \ {\displaystyle {\partial \kappa \over \partial x}}={\displaystyle {\mathsf{Bn}\ \left( h-\sigma \right) \over 2\sigma ^{2}}}\left[ 2\sigma {\displaystyle {\partial h \over \partial x}}-{\displaystyle {\partial \sigma \over \partial x}} \left( h+\sigma \right) \right] . \end{aligned}$$

Hence, system (58) can be rewritten as

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} {\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}\sigma =- \mathsf{Bn}, \\ \\ -\left( h-\sigma \right) ^{2}\left[ {\displaystyle {{\displaystyle {1 \over 6}}}}{ \displaystyle {{\displaystyle {\partial ^{2}P^{\left( 0\right) } \over \partial x^{2}}}}} \left( 2h+\sigma \right) +{\displaystyle {{\displaystyle {\partial P^{\left( 0\right) } \over \partial x}}}}{\displaystyle {{\displaystyle {\partial h \over \partial x}}}}\right] + \\ {\displaystyle {{\displaystyle {\mathsf{Bn}\ \left( h-\sigma \right) \over 2\sigma }}}} \left[ 2\sigma {\displaystyle {{\displaystyle {\partial h \over \partial x}}}}-{ \displaystyle {{\displaystyle {\partial \sigma \over \partial x}}}}\left( h+\sigma \right) \right] =0. \end{array} \right. {} \end{aligned} $$
(60)

In [10] it has been proved that there exists a unique pair of sufficiently regular functions \(\left ( \sigma \left ( x\right ) ,P^{\left ( 0\right ) }\left ( x\right ) \right ) \) such that:

  1. (a)

    \(0<\sigma \left ( x\right ) <h\left ( x\right ) \), for all \(x\in \left [ 0,1\right ] \).

  2. (b)

    \(P^{\left ( 0\right ) }\left ( 0\right ) =P_{in}\), and \(P^{\left ( 0\right ) }\left ( 1\right ) =P_{in}-\varDelta P\).

  3. (c)

    \(\sigma \left ( x\right ) \) and \(P^{\left ( 0\right ) }\left ( x\right ) \) fulfill the equations of (60).

Remark 5

Assuming that (60) is solvable according to the above definition, we find immediately a new “flow condition”. Indeed, integrating (60)1 between 0 and 1, we obtain

$$\displaystyle \begin{aligned} {\displaystyle {\varDelta P \over \mathsf{Bn}}}=\int_{0}^{1}{\displaystyle {dx \over \sigma \left( x\right) }}. \end{aligned}$$

Since

$$\displaystyle \begin{aligned} \int_{0}^{1}{\displaystyle {dx \over \sigma \left( x\right) }}>\int_{0}^{1}{\displaystyle {dx \over h\left( x\right) }}, {} \end{aligned} $$
(61)

if (60) admits a solution, the following inequality holds true

$$\displaystyle \begin{aligned} {\displaystyle {\varDelta P \over \mathsf{Bn}}}>\int_{0}^{1}{\displaystyle {dx \over h\left( x\right) }}. {} \end{aligned} $$
(62)

Going back to dimension variables we get

$$\displaystyle \begin{aligned} {\displaystyle {\varDelta P \over \tau _{o}}}>\int_{0}^{L}{\displaystyle {dx \over H\left( x\right) }}, \end{aligned}$$

which generalizes the classical Bingham flow condition.

4.4 Numerical Simulations

We present some numerical simulations to investigate the stationary behavior of σ(x) when \(\varGamma =\mathcal {O}(1)\). Suppose following function

$$\displaystyle \begin{aligned} h(x)={\displaystyle{{\displaystyle {1 \over 2}}}}-{\displaystyle{{\displaystyle {1 \over 10}}}}\arctan \left[ 200\left( x-{\displaystyle {1 \over 2}}\right) \right] , \end{aligned}$$

The channel profile and the free boundary separating the elastic and the viscous phase are shown in Figs. 11, 12, 13, 14 for different values of ΔPB n satisfying condition (62). We see that

$$\displaystyle \begin{aligned} \int_{0}^{1}{\displaystyle {{\displaystyle {dx^{{}^{\prime }} \over h(x^{{}^{\prime }})}}}}=2.2, \end{aligned}$$

and consequently we perform numerical simulation with ΔPB n > 2.2.

Fig. 11
figure 11

Plot of the free boundary (dashed line) y = σ(x) with ΔPB n = 2.5 and with condition (62) fulfilled

Fig. 12
figure 12

Plot of the free boundary (dashed line) y = σ(x) with ΔPB n = 3.5 and with condition (62) fulfilled

Fig. 13
figure 13

Plot of the free boundary (dashed line) y = σ(x) with ΔPB n = 5.5 and with condition (62) fulfilled

Fig. 14
figure 14

Plot of the free boundary (dashed line) y = σ(x) with ΔPB n = 9.5 and with condition (62) fulfilled

The amplitude of the inner core decreases as ΔPB n increases. When ΔPB n ≫ 1 (Fig. 14), the inner core approximately disappears and the system becomes almost purely viscous. In the same way, we observe that when ΔPB n is close to 2.2 then σ → h (Fig. 11 ) and the viscous part tends to disappear. Also in this case we speak of a limit, since condition (62) must be fulfilled.

5 Two Dimensional Channel Flow: A New Approach

To model the flow of a Bingham fluid one considers the balance of linear momentum written in the differential form

$$\displaystyle \begin{aligned} \rho {\displaystyle {D\mathbf{v} \over Dt}} =-\nabla P+\nabla \cdot \mathbf{S}, {} \end{aligned} $$
(63)

where ρ is density, v is velocity, P is pressure and S is the deviatoric part of the stress. Equation (63) is typically used in the whole domain, assuming that the velocity and the stress are continuous across the fluid/rigid interface. Within the liquid domain the fluid is assumed to behave as a viscous incompressible fluid, whereas in the rigid part the stress is indetermined. Indeed in the unyielded part we only know that the strain rate vanishes, i.e. D = 0. Assuming that Eq. (63) holds in every part of the domain may lead to paradoxes, as the one that occurs in lubrication regimes, [6, 17].

To avoid this occurrence we propose a novel approach which essentially consists in using a integral formulation for the balance of linear momentum within the unyielded part. We apply this approach to study the flow in a bidimensional channel of varying amplitude, with the driving force being an applied pressure gradient (Poiseuille flow). We assume that the aspect ratio of the channel is small, so that the lubrication approximation is suitable. In this case, Eq. (63) can be simplified introducing the ratio ε ≪ 1 between the length and the maximum amplitude and rescaling the problem in a nondimensional form. The solution can be sought as power series of ε, where the leading order is the one we are interested in. With this procedure one tacitly assumes that the nondimensional variables and their derivatives are \(\mathcal {O} (1)\) in the liquid and solid domain. In particular the stress components S ij are assumed to be everywhere \(\mathcal {O}(1)\), but this latter hypothesis can be checked “a posteriori” only in the liquid part, since in the rigid domain the stress is not determined.

This point is the central importance for our procedure. Indeed, the assumption \(S_{ij}= \mathcal {O}(1)\) and the use of (63) to derive the motion in the rigid part, leads to the well known “lubrication paradox” , which consists in a plug velocity that depends on the longitudinal coordinate. Note that the paradox disappears when one considers a deformable core, as shown in the previous sections. If one does not use Eq. (63) in the unyielded part and write the balance of linear momentum using an integral global approach similar to the one presented in [29] and in [28], the paradox is no longer present. Therefore, the unyielded part is treated as an evolving non material volume Ω t and its dynamics is modelled writing the balance of linear momentum as

$$\displaystyle \begin{aligned} \int_{\varOmega _{t}}{\displaystyle {\partial \over \partial t}}\left( \rho \mathbf{v}\right) dV+\int_{\partial \varOmega _{t}}\rho \mathbf{v}\left( \mathbf{v} \cdot \mathbf{n}\right) dS=\int_{\partial \varOmega _{t}}(\mathbf{T}\mathbf{n})dS, {} \end{aligned} $$
(64)

where T = −P I + S, is the Cauchy stress tensor and w the velocity of the boundary ∂Ω t. The advantage of this approach lies in the fact that the knowledge of the stress inside the rigid part is no longer needed and no guess has to be made on the order of magnitude of the stress components. Only the stress acting on the boundary of Ω t is required.

Therefore we need to know: (1) the forces acting on the yield surface σ (see Fig. 15); (2) the forces acting on the inlet and outlet of the channel. On σ the viscous stress is given once the problem in the viscous domain is solved. On the channel inlet and outlet the applied pressure, assumed to be a given datum of the problem, is required.

Fig. 15
figure 15

Sketch of the domain of the problem

When dealing with the leading order approximation in the channel flow, Eq. (64) becomes an integro-differential equation for the pressure P, whose solution allows to determine explicit expressions for the velocity field v and the yield surface σ. We prove that the longitudinal velocity is spatially uniform, while the transversal velocity vanishes (no paradox). We also show that these results can be also extended to the case of fluids with constant density and pressure dependent viscosity.

5.1 The Physical Model

Let us consider the flow of an incompressible Bingham fluid in a channel of length L and amplitude \(2h\left ( x\right ) \). Because of symmetry, we may limit our analysis to the upper part of the layer [0, h(x)]. The velocity field is given by

$$\displaystyle \begin{aligned} \mathbf{v}=v_{1}(x,y,t)\mathbf{i} +v_{2}(x,y,t)\mathbf{j}, \end{aligned}$$

The Cauchy stress is T = −P I + S, where the deviatoric part is the one of a Bingham fluid

$$\displaystyle \begin{aligned} \mathbf{S}=\left( 2\mu +{\displaystyle{{\displaystyle {\tau _{o} \over II_{\mathbf{D}}}}}}\right) \mathbf{D}. {} \end{aligned} $$
(65)

In the above μ is the viscosity, τ o is the yield stress. If D ≠ 0 we get

$$\displaystyle \begin{aligned}II_{\mathbf{S}}=2\mu II_{\mathbf{D}}+\tau _{o}\end{aligned}$$

which holds with II S ≥ τ o. Therefore, whenever D = 0, we have II S ≤ τ o and the stress is not determined. We assume that the viscous and the rigid regions are separated by a sharp interface y = ±σ(x, t). Assuming incompressibility we write

$$\displaystyle \begin{aligned} \mathrm{tr}\;\mathbf{D}={\displaystyle{{\displaystyle {\partial v_{1} \over \partial x}}}}+{\displaystyle{{\displaystyle {\partial v_{2} \over \partial y}}}}=0. {} \end{aligned} $$
(66)

5.2 The Viscous Domain

We write the governing equations in the viscous region neglecting body forces. These are the incompressibility condition (66) and

$$\displaystyle \begin{aligned} \rho \left( {\displaystyle{{\displaystyle {\partial v_{1} \over \partial t}}}}+v_{1}{\displaystyle{{\displaystyle {\partial v_{1} \over \partial x}}}}+v_{2}{\displaystyle{{\displaystyle {\partial v_{1} \over \partial y}}}}\right) =-{\displaystyle{{\displaystyle {\partial P \over \partial x}}}}+{\displaystyle{{\displaystyle {\partial S_{11} \over \partial x}}}}+{\displaystyle{{\displaystyle {\partial S_{12} \over \partial y }}}}, {} \end{aligned} $$
(67)
$$\displaystyle \begin{aligned} \rho \left( {\displaystyle{{\displaystyle {\partial v_{2} \over \partial t}}}}+v_{1}{\displaystyle{{\displaystyle {\partial v_{2} \over \partial x}}}}+v_{2}{\displaystyle{{\displaystyle {\partial v_{2}^{\ast } \over \partial y}}}}\right) =-{\displaystyle{{\displaystyle {\partial P \over \partial y}}}}+{\displaystyle{{\displaystyle {\partial S_{12} \over \partial x}}}}+{\displaystyle{{\displaystyle {\partial S_{22} \over \partial y }}}}, {} \end{aligned} $$
(68)

5.3 The Rigid Domain

The rigid domain Ω t at some time t > 0 is given by

$$\displaystyle \begin{aligned} \varOmega _{t}=\left\{ (x,y):\;\;x\in \lbrack 0,L],\;\;\;y\in \lbrack -\sigma ,\sigma ]\right\} . \end{aligned}$$

The integral momentum balance for the whole domain Ω t in the absence of body forces is given by (64). Focussing on the upper part of the domain (y > 0) we find that (64) can be rewritten as

$$\displaystyle \begin{aligned} 2{\displaystyle{{\displaystyle {\partial \over \partial t}}}}\left( \rho \mathbf{v}\right) \int_{0}^{L}\sigma (x,t)dx+2\rho \mathbf{v}\int_{0}^{L}{\displaystyle{{\displaystyle {\partial \sigma \over \partial t}}}} (x,t)dx=\int_{\partial \varOmega _{t}} \mathbf{T}\mathbf{n}\;dS. \end{aligned}$$

The external forces acting on the boundary ∂Ω t are expressed by the surface integral on the r.h.s. Assuming that P in, P out are the (uniform) pressures acting on the inlet and outlet of the channel, we find

$$\displaystyle \begin{aligned} \begin{array}{rcl} \int_{\partial \varOmega _{t}}(\mathbf{T}\mathbf{n})dS&\displaystyle =&\displaystyle 2\int_{0}^{L}\left[ \begin{array}{c} (-\sigma _{x}T_{11}+T_{12})_{\sigma } \\ \\ 0 \end{array} \right] dx \\ &\displaystyle &\displaystyle +2\int_{0}^{\sigma _{out}}\left( \begin{array}{c} -P_{out} \\ \\ 0 \end{array} \right) dy+2\int_{0}^{\sigma _{in}}\left( \begin{array}{c} P_{in} \\ \\ 0 \end{array} \right) dy, \end{array} \end{aligned} $$

where σ in = σ(0, t), σ out = σ(L, t). Recalling that in the rigid plug velocity is

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} v_{1}=k_{1}(t), \\ \\ v_{2}=k_{2}(t)=0\ \ \ \text{(by}\ \ \text{symmetry),} \end{array} \right. {} \end{aligned} $$
(69)

the dynamics of the whole rigid region is expressed only by the first component of (64), that is

$$\displaystyle \begin{aligned} \int_{0}^{L}{{{\displaystyle {\partial \over \partial t}}}}\left( \rho k_{1}\sigma \right) dx=\int_{0}^{L} \left[ -\sigma _{x}T_{11}+T_{12}\right] _{\sigma }dx+P_{in}\sigma _{in}-P_{out}\sigma _{out}, {} \end{aligned} $$
(70)

Hence the prescribed pressure difference driving the flow is

$$\displaystyle \begin{aligned} \varDelta P =P_{in} -P_{out}. {} \end{aligned} $$
(71)

The boundary condition on the channel wall is

$$\displaystyle \begin{aligned} \mathbf{v}(x,h,t)=0 {} \end{aligned} $$
(72)

i.e. the no-slip condition. On σ we impose

$$\displaystyle \begin{aligned} \left. \llbracket\mathbf{v}\cdot \mathbf{t}\rrbracket\right. _{y=\sigma }=0,\;\;\;\;\;\;\;\;\;\left. \llbracket\mathbf{v} \cdot \mathbf{n}\rrbracket\right. _{y=\sigma }=0, {} \end{aligned} $$
(73)
$$\displaystyle \begin{aligned} \left. \llbracket\mathbf{T}\mathbf{n}\cdot \mathbf{t}\rrbracket \right. _{y=\sigma }=0,\;\;\;\;\;\;\;\;\;\left. \llbracket \mathbf{T}\mathbf{n}\cdot \mathbf{n}\rrbracket\right. _{y=\sigma }=0, {} \end{aligned} $$
(74)

which express the continuity of the velocity and of the stress across the yield surface y = σ ( t and n are the tangent and normal unit vector to σ).

Remark 6

In Sect. 5.10 we will extend our model the case in which the viscosity depends on pressure, namely \(\mu =\mu \mu \left ( P\right )\).

5.4 Scaling

Set

$$\displaystyle \begin{aligned} H=\sup_{x\in \lbrack 0,L]}h(x), \end{aligned}$$

and introduce

$$\displaystyle \begin{aligned} \varepsilon ={\displaystyle{{\displaystyle {H \over L}}}}\ll 1, \end{aligned}$$

which is crucial for applying the classical thin film approach. We rescale the problem using the following non dimensional variables

$$\displaystyle \begin{aligned} \tilde{x}={\displaystyle{{\displaystyle {x \over L}}}},\;\;\;\;\;\;\;\;\;\tilde{y}={ \displaystyle{{\displaystyle {y \over \varepsilon L}}}},\;\;\;\;\;\;\;\;\; \tilde{\sigma}={\displaystyle{{\displaystyle {\sigma \over \varepsilon L}}}} ,\;\;\;\;\;\;\;\;\;\tilde{h}={\displaystyle{{\displaystyle {h \over \varepsilon L}}}} ,\;\;\;\;\;\;\;\;\;\tilde{t}={\displaystyle{{\displaystyle {t \over (L/U)}},} } \end{aligned}$$
$$\displaystyle \begin{aligned} \tilde{v}_{1}={{{\displaystyle {v_{1} \over U}}}},\;\;\;\;v_{2}={{{\displaystyle {v_{2} \over \varepsilon U}}}},\;\;\;\;\tilde{P}={{{\displaystyle {P-P_{out} \over P_{c}}}}},\;\;\;\;\varDelta \tilde{P}={\displaystyle {\varDelta P \over P_{c}}}, {} \end{aligned} $$
(75)
$$\displaystyle \begin{aligned} \tilde{\mathbf{S}}={\displaystyle{{\displaystyle {\mathbf{S} \over (\mu U/H)}}}},\;\;\;\;\;\tilde{\mathbf{D}}={\displaystyle{{\displaystyle {\mathbf{D} \over (U/H)}}}},\;\;\;\;\;\tilde{II}_{D}={\displaystyle{{\displaystyle { II_{D} \over (U/H)}}}},\;\;\;\;\;\tilde{II}_{S}={\displaystyle{ {\displaystyle {II_{S} \over (\mu U/H)}},}} \end{aligned}$$

where we select the reference pressure using the Poiseuille formula

$$\displaystyle \begin{aligned} P_c={\displaystyle {\mu L U \over H^2}} \end{aligned} $$
(76)

After some algebra (and neglecting the tildas) we find

$$\displaystyle \begin{aligned} \mathbf{S}=\left( 2+{\displaystyle{{\displaystyle {\mathsf{Bn} \over II_{D}}}}}\right) \mathbf{D}, \end{aligned}$$

where

$$\displaystyle \begin{aligned} \mathsf{Bn}={\displaystyle{{\displaystyle {\tau _{o}H \over \mu U}}~}} \end{aligned}$$

is the Bingham number. Moreover

$$\displaystyle \begin{aligned} II_{D}=\sqrt{\varepsilon ^{2}\left( {\displaystyle{{\displaystyle {\partial v_{1} \over \partial x}}}}\right) ^{2}+{\displaystyle{{\displaystyle {1 \over 4}}}}\left( {\displaystyle{ {\displaystyle {\partial v_{1} \over \partial y}}}}+\varepsilon ^{2}{\displaystyle{{\displaystyle { \partial v_{2} \over \partial x}}}}\right) ^{2}}. \end{aligned}$$

Equations (66)–(68) become

$$\displaystyle \begin{aligned} {\displaystyle{{\displaystyle {\partial v_{1} \over \partial x}}}}+{\displaystyle{{\displaystyle { \partial v_{2} \over \partial y}}}}=0, {} \end{aligned} $$
(77)
$$\displaystyle \begin{aligned} \varepsilon \mathsf{Re}\left( {\displaystyle{{\displaystyle {\partial v_{1} \over \partial t }}}}+v_{1}{\displaystyle{{\displaystyle {\partial v_{1} \over \partial x}}}}+v_{2}{ \displaystyle{{\displaystyle {\partial v_{1} \over \partial y}}}}\right) =-{\displaystyle{ {\displaystyle {\partial P \over \partial x}}}}+\varepsilon {\displaystyle{{\displaystyle {\partial S_{11} \over \partial x}}}}+{\displaystyle{{\displaystyle {\partial S_{12} \over \partial y}}}}, {} \end{aligned} $$
(78)
$$\displaystyle \begin{aligned} \varepsilon ^{3}\mathsf{Re}\left( {\displaystyle{{\displaystyle {\partial v_{2} \over \partial t}}}}+v_{1}{\displaystyle{{\displaystyle {\partial v_{2} \over \partial x}}}}+v_{2}{ \displaystyle{{\displaystyle {\partial v_{2} \over \partial y}}}}\right) =-{\displaystyle{ {\displaystyle {\partial P \over \partial y}}}}+\varepsilon ^{2}{\displaystyle{{\displaystyle { \partial S_{12} \over \partial x}}}}+\varepsilon {\displaystyle{{\displaystyle {\partial S_{22} \over \partial y}}}}, {} \end{aligned} $$
(79)

where

$$\displaystyle \begin{aligned}\mathsf{Re}=\left( {\displaystyle{{\displaystyle {\rho UH \over \mu}}}}\right) \end{aligned}$$

is the Reynolds number. The inner core equation (70) becomes

$$\displaystyle \begin{aligned} \varepsilon \mathsf{Re}\int_{0}^{1}{\displaystyle{{\displaystyle {\partial \over \partial t }}}}(k_{1}\sigma )dx=\int_{0}^{1}\left[\,P\sigma _{x}-\varepsilon \sigma _{x}S_{11}+S_{12}\right] _{\sigma ^{+}}~dx+\varDelta P\sigma _{in}. {} \end{aligned} $$
(80)

The boundary conditions (72)–(74) become

$$\displaystyle \begin{aligned} \mathbf{v}(x,h,t)=0, {} \end{aligned} $$
(81)
$$\displaystyle \begin{aligned} \llbracket v_{1}\rrbracket_{y=\sigma }=\llbracket v_{2}\rrbracket_{y=\sigma }=0, {} \end{aligned} $$
(82)
$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} \llbracket P\rrbracket\left[ 1+\varepsilon ^{2}\left( {\displaystyle{{\displaystyle { \partial \sigma \over \partial x}}}}\right) ^{2}\right] _{y=\sigma }+\left[ \varepsilon ^{3}S_{11}\left( {\displaystyle{{\displaystyle {\partial \sigma \over \partial x}}}}\right) ^{2}-2\varepsilon ^{2}S_{12}\left( {\displaystyle{{\displaystyle {\partial \sigma \over \partial x}}}}\right) +\varepsilon S_{22}\right] _{y=\sigma }=0, \\ \\ \llbracket S_{12}\rrbracket_{y=\sigma }+\varepsilon \left( {\displaystyle{ {\displaystyle {\partial \sigma \over \partial x}}}}\right) \left[ S_{22}-S_{11}- \varepsilon S_{12}{\displaystyle{{\displaystyle {\partial \sigma \over \partial x}}}}\right] _{y=\sigma }=0. \end{array} \right. {} \end{aligned} $$
(83)

In the rigid domain the non dimensional velocity field is

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} v_{1}=k_{1}(t), \\ \\ v_{2}=0, \end{array} \right. {} \end{aligned} $$
(84)

with k 1 = k 1U.

5.5 The Leading Order Approximation

We look for a solution in which the main variables of the problem can be expressed as power series of ε

$$\displaystyle \begin{aligned} \mathbf{v}=\sum_{j=0}^{\infty}{\mathbf{v}}^{(j)}\varepsilon^{j} \ \ \ \ \ \ \ \ P=\sum_{j=0}^{\infty}P^{(j)}\varepsilon^{j}\ \ \ \ \ \ \ \ \ \sigma=\sum_{j=0}^{\infty}\sigma^{(j)}\varepsilon^{j} \end{aligned}$$

We further assume that h(x) is such that

$$\displaystyle \begin{aligned}{\displaystyle{{\displaystyle {\partial h \over \partial x}}}}= {O}\left( 1\right) \end{aligned}$$

and we limit our analysis to the leading order, assuming that \(\mathsf {Bi=}\mathcal {O}\left ( 1\right ) \) and \(\mathsf { Re\leqslant }\mathcal {O}\left ( 1\right ) \). We begin by observing that

$$\displaystyle \begin{aligned} S_{12}^{(0)}=\left[ 1+{\displaystyle{{\displaystyle {\mathsf{Bn} \over |v_{1y}^{(0)}|}}}} \right] v_{1y}^{(0)}, \end{aligned}$$

and, since we are looking for a solution with \(v_{1y}^{(0)}<0\) in the upper part of the channel, we write

$$\displaystyle \begin{aligned} S_{12}^{(0)}=v_{1y}^{(0)}-\mathsf{Bn}. \end{aligned}$$

The problem reduces to

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} {\displaystyle{{\displaystyle {\partial v_{1}^{(0)} \over \partial x}}}}+{\displaystyle{ {\displaystyle {\partial v_{2}^{(0)} \over \partial y}}}}=0, \\ \\ -{\displaystyle{{\displaystyle {\partial P^{(0)} \over \partial x}}}}+{\displaystyle{{\displaystyle { \partial \over \partial y}}}}\left( {\displaystyle{{\displaystyle {\partial v_{1}^{(0)} \over \partial y}}}}\right) =0, \\ \\ -{\displaystyle{{\displaystyle {\partial P^{(0)} \over \partial y}}}}=0, \end{array} \right. {} \end{aligned} $$
(85)

with boundary conditions

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} \left. {\displaystyle{{\displaystyle {\partial v_{1}^{(0)} \over \partial y}}}}\right\vert _{y=\sigma ^{(0)}}=0, \\ \\ v_{1}^{(0)}(x,h,t)=0. \end{array} \right. {} \end{aligned} $$
(86)

Integrating we find

$$\displaystyle \begin{aligned} v_{1}^{(0)}=-P_{x}^{(0)}{{{\displaystyle {(h-y)(y-2\sigma ^{(0)}+h) \over 6}}}}. {} \end{aligned} $$
(87)

Exploiting the continuity equation we get

$$\displaystyle \begin{aligned} v_{2}^{(0)}=-{\displaystyle{{\displaystyle {\partial \over \partial x}}}}\left[\,P_{x}^{(0)}{ {{\displaystyle {(y-h)^{2}(y-3\sigma ^{(0)}+2h) \over 6}}}}\right] . {} \end{aligned} $$
(88)

Evaluating the velocity components on the yield surface we get

$$\displaystyle \begin{aligned} \left. v_{1}^{(0)}\right\vert _{y=\sigma ^{(0)}}=k_{1}^{(0)}(t)=-P_{x}^{(0)}{ {{\displaystyle {(h-\sigma ^{(0)})^{2} \over 2}}}}, {} \end{aligned} $$
(89)
$$\displaystyle \begin{aligned} \left. v_{2}^{(0)}\right\vert _{y=\sigma ^{(0)}}={{{\displaystyle {\partial \over \partial x}}}}\left[ -P_{x}^{(0)}{\displaystyle{{\displaystyle {(h-\sigma ^{(0)})^{3} \over 3}}}} \right] -\sigma _{x}^{(0)}P_{x}^{(0)}{\displaystyle{{\displaystyle {(h-\sigma ^{(0)})^{2} \over 2}}}}=0, \end{aligned}$$

which entails

$$\displaystyle \begin{aligned} \underset{k_{1}^{(0)}}{\underbrace{\left( -{P_{x}^{(0)}\displaystyle{{\displaystyle { (h-\sigma ^{(0)})^{2} \over 2}}}}\right) }}{\displaystyle\cdot {{\displaystyle {\partial \over \partial x}}}}\left[ {\displaystyle{{\displaystyle {2 \over 3}}}}(h-\sigma ^{(0)})\right] =-\sigma _{x}^{(0)}\underset{k_{1}^{(0)}}{\underbrace{\left( -P_{x}^{(0)}{ \displaystyle{{\displaystyle {(h-\sigma ^{(0)})^{2} \over 2}}}}\right) }}~. \end{aligned}$$

Supposing \(k_{1}^{(0)}\neq 0\), we find

$$\displaystyle \begin{aligned} \sigma ^{(0)}\left( x,t\right) =-2h\left( x\right) -C, {} \end{aligned} $$
(90)

where C is unknown. Let us now consider the rigid core equation (80) at the zero order

$$\displaystyle \begin{aligned} \operatorname*{\displaystyle \int}_{0}^{1}P^{(0)}\sigma _{x}^{(0)}dx-\mathsf{Bn} +\varDelta P\sigma _{in}^{(0)}=0, \end{aligned}$$

which, after an integration by parts, reduces to

$$\displaystyle \begin{aligned} -\int_{0}^{1}P_{x}^{(0)}\sigma ^{(0)}dx=\mathsf{Bn}. {} \end{aligned} $$
(91)

Substituting (90) into (91), we obtain

$$\displaystyle \begin{aligned} C ={{{\displaystyle {2\operatorname*{\displaystyle \int}_{0}^{1}P_{x}^{(0)}hdx- \mathsf{Bn} \over \varDelta P}}}}. \end{aligned}$$

We thus have

$$\displaystyle \begin{aligned} \sigma ^{(0)} =-2h\left( x\right) +{{{\displaystyle {\mathsf{Bi-}2 \operatorname*{\displaystyle \int}_{0}^{1}P_{x}^{(0)}h~dx \over \varDelta P }} }}, {} \end{aligned} $$
(92)

or equivalently

$$\displaystyle \begin{aligned} \sigma ^{(0)} =2(h_{in}-h\left( x\right) )+{{{\displaystyle {\mathsf{ Bi} \over \varDelta P }}}}+{{{\displaystyle {2 \over \varDelta P }}}} \int_{0}^{1}P^{(0)}h_{x}dx, {} \end{aligned} $$
(93)

where h in = h(0). Defining the viscous region width as

$$\displaystyle \begin{aligned} \ell ^{\left( 0\right) } =h\left( x\right) -\sigma ^{(0)}\left( x,t\right) , {} \end{aligned} $$
(94)

formula (92) entails

$$\displaystyle \begin{aligned} \ell ^{\left( 0\right) } =3h\left( x\right) +{{{\displaystyle {2 \operatorname*{\displaystyle \int}_{0}^{1}P_{x}^{(0)}h~dx-\mathsf{Bn} \over \varDelta P }}}}. {} \end{aligned} $$
(95)

Hence

$$\displaystyle \begin{aligned} k_{1}^{(0)}=-P_{x}^{(0)}{{{\displaystyle {\ell ^{\left( 0\right) }{}^{2} \over 2}}.}} {} \end{aligned} $$
(96)

Now, differentiating(~96) with respect to x, we obtain

$$\displaystyle \begin{aligned} P_{xx}^{\left( 0\right) }+6{\displaystyle {h_{x} \over \ell ^{\left( 0\right) }}}P_{x}^{\left( 0\right) }=0, \end{aligned}$$

that is the integro-differential equation

$$\displaystyle \begin{aligned} P_{xx}^{\left( 0\right) }+{\displaystyle {6h_{x} \over \left[ 3h+{{{{{\displaystyle {2 \operatorname*{\displaystyle \int}_{0}^{1}P_{x}^{(0)}h~dx-\mathsf{Bn} \over \varDelta P\left( t\right) }}}}}}\right] }P_{x}^{\left( 0\right) }=0.} {} \end{aligned} $$
(97)

The boundary conditions are \(\left . P^{\left ( 0\right ) }\right \vert _{x=0}=\varDelta P \), and \(\left . P^{\left ( 0\right ) }\right \vert _{x=1}=0\). The solution \(P^{\left ( 0\right ) }\) of (97) is then used to evaluate the \(v_{1}^{\left ( 0\right ) }\) via ( 87), \(v_{2}^{\left ( 0\right ) }\) via (88) and the yield surface σ (0) via (93).

Remark 7

From (92) we observe that \(\sigma _{x}^{(0)}=-2h_{x}\), meaning that the amplitude of the rigid core becomes larger as the channel narrows, whereas it shrinks as the channel becomes wider. This is in accordance with what found in [21].

5.6 Flow Condition

Let us investigate the conditions on ΔP that prevent the system from coming to a halt. Let \(h\left ( x\right ) \equiv h_{in}\). From (93) we get:

  • \(\varDelta P>{\displaystyle {{\displaystyle {\mathsf {Bn} \over h_{in}}}}}\), ⇒ σ (0) < h in (the fluid is flowing)

  • \(\varDelta P<{\displaystyle {{\displaystyle {\mathsf {Bn} \over h_{in}}}}}\), ⇒ σ (0) > h in (no flow)

When h(x) is not uniform we have to ensure that \(\sigma ^{\left ( 0\right ) }<h\left ( x\right ) \), in order to prevent the flow from stopping. Recalling (95) we must impose

$$\displaystyle \begin{aligned} \ell ^{\left( 0\right) } \geq 3h_{min}-2h_{in}-{{{\displaystyle {\mathsf{Bn} \over \varDelta P}}}}-{{{\displaystyle {2 \over \varDelta P}}}} \int_{0}^{1}P^{(0)}h_{x}dx, {} \end{aligned} $$
(98)

where \(h_{\min }=\min _{x\in \lbrack 0,1]}h\). To estimate the integral in the r.h.s. we remark that \(P^{\left ( 0\right ) }\) fulfils Eq. (97), that is an equation of elliptic type. Maximum principle entails \(0\leq P^{\left ( 0\right ) }\leq \varDelta P\). So, writing

$$\displaystyle \begin{aligned} \int_{0}^{1}P^{(0)}h_{x}dx=\underset{\leq 0}{~\underbrace{\int_{\left\{ h_{x}\leq 0\right\} }P^{(0)}h_{x}dx}}~+~\underset{\geq 0}{\underbrace{ \int_{\left\{ h_{x}\geq 0\right\} }P^{(0)}h_{x}dx}}~, \end{aligned}$$

we have

$$\displaystyle \begin{aligned} \varDelta P\min \left\{ \underline{h}_{~x};0\right\} \leq \varDelta P\max \left\{ \overline{h}_{x};0\right\}, \end{aligned}$$

where

$$\displaystyle \begin{aligned} \underline{h}_{~x}=\min_{x\in \lbrack 0,1]}h_{x}(x),\;\;\;\;\mathrm{and} \;\;\;\;\overline{h}_{x}=\min_{x\in \lbrack 0,1]}h_{x}(x). \end{aligned}$$

In conclusion

$$\displaystyle \begin{aligned} 2\min \left\{ \underline{h}_{~x};0\right\} \leq {\displaystyle{{\displaystyle {2 \over \varDelta P}}}}\int_{0}^{1}P^{(0)}h_{x}dx\leq 2\max \left\{ \overline{h} _{x};0\right\} . {} \end{aligned} $$
(99)

Therefore, recalling (98), we have

$$\displaystyle \begin{aligned}\ell ^{\left( 0\right) } \geq 3h_{\min }-2h_{in}-{{{\displaystyle { \mathsf{Bn} \over \varDelta P}}}}-{{{\displaystyle {2 \over \varDelta P}}}}\int_{0}^{1}P^{(0)}h_{x}dx \end{aligned}$$
$$\displaystyle \begin{aligned}\geq 3h_{\min }-2\max \left\{ \overline{h}_{x};0\right\} -2h_{in}-{\displaystyle { \mathsf{Bn} \over \varDelta P}}, {} \end{aligned} $$
(100)

If we assume

$$\displaystyle \begin{aligned}\left( 3h_{\min }-2\max \left\{ \overline{h}_{x};0\right\} -2h_{in}\right) >0\end{aligned}$$

and require that

$$\displaystyle \begin{aligned} 3h_{\min }-2\max \left\{ \overline{h}_{x};0\right\} -2h_{in}-{\displaystyle {\mathsf{Bi } \over \varDelta P}}>0, \end{aligned}$$

which implies

$$\displaystyle \begin{aligned} \varDelta P>{\displaystyle {\mathsf{Bn} \over 3h_{\min }-2\max \left\{ \overline{h}_{x};0\right\} -2h_{in}}}, {} \end{aligned} $$
(101)

we are sure that the flow never comes to a stop.

Example 2

If we consider a linear wall profile

$$\displaystyle \begin{aligned} h=h_{in}+~\underset{\varDelta h}{\underbrace{\left( h_{out}-h_{in}\right) }}~x, \end{aligned}$$

where h out > 0, there are two possibilities:

  • Δh > 0, ⇒ \(h_{\min }=h_{in}\), and \(\max \left \{ \overline {h}_{x};0\right \} =\varDelta h\). Condition (101) yields

    $$\displaystyle \begin{aligned} {{{\displaystyle {\mathsf{Bn} \over \varDelta P}}}}<~\underset{2h_{out}-3h_{in}}{\underbrace{ h_{in}-2\varDelta h}}~,\ \ \ \ \ \Leftrightarrow \ \ \ \ \varDelta P>{{{\displaystyle { \mathsf{Bn} \over 2h_{out}-3h_{in}}},}} \end{aligned}$$

    where, of course, we assume 2h out − 3h in > 0

  • Δh < 0, ⇒ \(h_{\min }=h_{out}\), and \(\max \left \{ \overline {h}_{x};0\right \} =0\). Inequality (101) entails

    $$\displaystyle \begin{aligned} {{{\displaystyle {\mathsf{Bn} \over \varDelta P}}}}<~\underset{3h_{out}-2h_{in}}{\underbrace{ 2\varDelta h+h_{out}}}~,\ \ \ \ \Leftrightarrow \ \ \ \ \varDelta P>{{{\displaystyle { \mathsf{Bn} \over 3h_{out}-2h_{in}}},}} \end{aligned}$$

    where now we require 3h out − 2h in > 0.

5.7 Inner Core Appearance or Disappearance

A non uniform channel profile may cause the appearance/disappearance of the rigid plug. These phenomena (highlighted also in [6] and [21]) are not possible when the channel profile is uniform, namely when \(h\left ( x\right ) \equiv h_{in}\). Recalling (93 ), we set

$$\displaystyle \begin{aligned} \sigma ^{\left( 0\right) } =\max \left\{ 0;~2(h_{in}-h )+{{{\displaystyle {\mathsf{Bn} \over \varDelta P}}}}+{{{\displaystyle {2 \over \varDelta P}}}}\int_{0}^{1}P^{(0)}h_{x}dx\right\} , \end{aligned}$$

in order to avoid physical inconsistencies. Hence, \(\sigma ^{\left ( 0\right ) }\) vanishes when

$$\displaystyle \begin{aligned} h \geq h_{in}+{{{\displaystyle {\mathsf{Bn} \over 2\varDelta P}}}}+{{{\displaystyle {1 \over \varDelta P}}}}\int_{0}^{1}P^{(0)}h_{x}dx. {} \end{aligned} $$
(102)

The r.h.s. of (102) is a critical value, that we denote as h crt , such that, whenever h ≥ h crt the core disappears.

Example 3

Let us consider the channel profile

$$\displaystyle \begin{aligned} h(x)={{{\displaystyle {\arctan \left[ 5\left( {\displaystyle{{\displaystyle {1 \over 2}}}}-x\right) \right] \over 4\arctan \left( {\displaystyle{{\displaystyle {5 \over 2}}}}\right) }}}}+{ \displaystyle{{\displaystyle {3 \over 4}}}}. {} \end{aligned} $$
(103)

depicted with the dashed line in Fig. 20. We now estimate h crt exploiting (102), when ΔP = 10, and B i = 5,

$$\displaystyle \begin{aligned} \begin{array}{rcl} h\left( x\right) &\displaystyle \geq &\displaystyle 1+{{{\displaystyle {\mathsf{Bn} \over 2\varDelta P}}}}-{{{\displaystyle {1 \over \varDelta P}}}}\int_{0}^{1}P^{(0)}\left\vert h_{x}\right\vert dx \\[0.05in] &\displaystyle \geq &\displaystyle 1+{{{\displaystyle {\mathsf{Bn} \over 2\varDelta P}}}}-\left\Vert h_{x}\right\Vert _{L^{2}}\geq 1+{{{\displaystyle {\mathsf{Bn} \over 2\varDelta P}}-0.58\approx 0.67.}} \end{array} \end{aligned} $$

The core-free region is thus obtained solving h ≥ h crt, which we approximate with h ≥ 0.67, whose solution is the interval 1 ≤ x ≤ 0.58. Looking at Fig. 20 the actual core-free region is 1 ≤ x ≤ 0.55, which substantially agrees with the above estimate.

5.8 Solution for an Almost Flat Channel

When \(h\left ( x\right ) \equiv h_{in}\) (i.e. uniform channel) equation (93) gives

$$\displaystyle \begin{aligned} \sigma ^{(0)} ={{{\displaystyle {\mathsf{Bn} \over \varDelta P }}} }. {} \end{aligned} $$
(104)

Equation (97) yields

$$\displaystyle \begin{aligned}P^{\left( 0\right) } =\left( 1-x\right) \varDelta P.\end{aligned}$$

The velocity field becomesFootnote 7

$$\displaystyle \begin{aligned} \left\{ \begin{array}{lll} v_{1}^{(0)} & = & -\varDelta P \left[ {\displaystyle{{\displaystyle { (y-\sigma ^{(0)})^{2} \over 2}}}}-{\displaystyle{{\displaystyle {(1-\sigma ^{(0)})^{2} \over 2}}}} \right] , \\ & & \\ v_{2}^{(0)} & = & 0, \end{array} \right. {} \end{aligned} $$
(105)

and we also find

$$\displaystyle \begin{aligned}k_{1}^{(0)}={\displaystyle{{\displaystyle {\varDelta P \over 2}}}}(1-\sigma ^{(0)})^{2}.\end{aligned}$$

Let us now consider a non-uniform channel profile \(h\left ( x\right ) \), assuming that amplitude width variation is small. We thus set

$$\displaystyle \begin{aligned} h\left( x\right) =\left\langle h\right\rangle +\phi \left( x\right) , {} \end{aligned} $$
(106)

where \(\left \langle h\right \rangle \) denotes the spatial average along the channel, i.e.

$$\displaystyle \begin{aligned}\left\langle ~h~\right\rangle =\operatorname*{\displaystyle \int}_{0}^{1}h\left( x\right) dx\end{aligned}$$

and assume that \(\max \left \vert \phi \left ( x\right ) \right \vert \) is small, that is we consider an almost flat channel. We look for \(P^{\left ( 0\right ) }\) in the form

$$\displaystyle \begin{aligned} P^{\left( 0\right) }=\left( 1-x\right) \varDelta P +\varPi, {} \end{aligned} $$
(107)

where \(\left . \varPi \right \vert _{x=0}=\left . \varPi \right \vert _{x=1}=0\), and where we expect that both \(\max \left \vert \varPi \right \vert \), \(\max \left \vert \varPi _{x}\right \vert \) are small. Inserting (107) into (92) we obtain

$$\displaystyle \begin{aligned} \sigma ^{\left( 0\right) } ={\displaystyle {\mathsf{Bn} \over \varDelta P}} -2\phi \left( x\right) -{\displaystyle {2 \over \varDelta P}}\int_{0}^{1}\varPi _{x}\phi ~dx\approx {\displaystyle {\mathsf{Bn} \over \varDelta P }}-2\phi \left( x\right) . {} \end{aligned} $$
(108)

Concerning \(\ell ^{\left ( 0\right ) }\) we have

$$\displaystyle \begin{aligned} \ell ^{\left( 0\right) }\left( x,t\right) \approx \left\langle h\right\rangle -{\displaystyle {\mathsf{Bn} \over \varDelta P\left( t\right) }}+3\phi \left( x\right) . {} \end{aligned} $$
(109)

Exploiting then (97) we compute the pressure field solving

$$\displaystyle \begin{aligned} \varPi _{xx}+{\displaystyle {6\phi _{x} \over \ell ^{\left( 0\right) }}}\left( -\varDelta P+\varPi _{x}\right) =0. \end{aligned}$$

Neglecting ϕ x Π x we end up with the following problem

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} \varPi _{xx}-2\varDelta P\left[ {\displaystyle{{\displaystyle {\phi _{x} \over \phi +\mathcal{A}}}} }\right] =0,\;\;\mathrm{where}\;\;\;\mathcal{A}={\displaystyle{{\displaystyle {<h>\over 3}}} }-{\displaystyle{{\displaystyle {\mathsf{Bn} \over 3\varDelta P}},}} \\ \\ \left. \varPi \right\vert _{x=0}=\left. \varPi \right\vert _{x=1}=0, \end{array} \right. \end{aligned}$$

so that

$$\displaystyle \begin{aligned} \varPi _{x}=\left( const\right) +2\varDelta P\ln \left[ 1+{{{\displaystyle {\phi \left( x\right) \over \mathcal{A}}}}}\right] \approx \left( const\right) +2\varDelta P{{ {\displaystyle {\phi \left( x\right) \over \mathcal{A}}}}}. \end{aligned}$$

In conclusion

$$\displaystyle \begin{aligned} \varPi ={\displaystyle{{\displaystyle {2\varDelta P\left( t\right) \over \mathcal{A}}}}}\int_{0}^{x}\phi (x^{{}^{\prime }})dx^{{}^{\prime }}, \end{aligned}$$

which yields

$$\displaystyle \begin{aligned} P^{\left( 0\right) } =\varDelta P \left( 1-x\right) +{\displaystyle{{\displaystyle {6\varDelta P^{2} \over \left\langle h\right\rangle \varDelta P-\mathsf{Bn}}}}}\int_{0}^{x}\phi \left( x^{\prime }\right) dx^{\prime }. {} \end{aligned} $$
(110)

Example 4

Let us consider \(h\left ( x\right ) =1+mx\), with m small. We write

$$\displaystyle \begin{aligned} h\left( x\right) =\underset{\left\langle h\right\rangle }{\underbrace{1+ {\displaystyle {m \over 2}}}}+\underset{\phi \left( x\right) }{\underbrace{m\left( x-{\displaystyle {1 \over 2}}\right) }}~. \end{aligned}$$

In this case

$$\displaystyle \begin{aligned} \sigma ^{\left( 0\right) }={\displaystyle {\mathsf{Bn} \over \varDelta P}}-2m\left( x-{\displaystyle {1 \over 2 }}\right) , \end{aligned}$$

and

$$\displaystyle \begin{aligned} P^{\left( 0\right) }=\varDelta P \left( 1-x\right) +{{{\displaystyle {3m\varDelta P^{2} \over \left\langle h\right\rangle \varDelta P- \mathsf{Bn}}}}}x\left( x-1\right). \end{aligned}$$

We see that \(\sigma _{x}^{(0)}=-2m\), i.e. the core amplitude widens for m < 0 and shrinks for m > 0.

Example 5

Let us consider a wavy channel

$$\displaystyle \begin{aligned} h(x)=1-\theta \cos \left[ 2\pi \delta \left( x-{\displaystyle{{\displaystyle {1 \over 2}}}} \right) \right] , {} \end{aligned} $$
(111)

where δ > 0, and θ ≪ 1. We write

$$\displaystyle \begin{aligned} h\left( x\right) =\underset{\left\langle h\right\rangle }{\underbrace{\left[ 1-{\displaystyle {\theta \over \pi \delta }}\sin \left( \pi \delta \right) \right] }}+ \underset{\phi \left( x\right) }{\underbrace{\theta \left[ {\displaystyle {\sin \left( \pi \delta \right) \over \pi \delta }}-\cos \left( 2\pi \delta \left( x-{\displaystyle {1 \over 2 }}\right) \right) \right] }}~, \end{aligned}$$

with \(\max \left \vert \phi \right \vert =\mathcal {O}\left ( \theta \right ) \ll 1\). Exploiting (108) we obtain

$$\displaystyle \begin{aligned} \sigma ^{\left( 0\right) }\approx {\displaystyle {\mathsf{Bn} \over \varDelta P}}-2\theta \left[ {\displaystyle {\sin \left( \pi \delta \right) \over \pi \delta }}-\cos \left( 2\pi \delta \left( x-{\displaystyle {1 \over 2}}\right) \right) \right] , {} \end{aligned} $$
(112)

The behavior for θ = 0.1, and δ = 1∕5 is shown in Figs. 16 and 17. In particular in Fig. 17 a close-up showing the difference between the approximated solution (112) and the computed one (see next section) is displayed.

Fig. 16
figure 16

The channel profile \(h \left ( x \right ) \) is (111) and of \( \sigma ^{ \left ( 0 \right ) }\) given by (92 ), (112), with B n = 5, ΔP = 10.5 , δ = 0.2, θ = 0.1

Fig. 17
figure 17

Close up for the difference between σ (0) given by (112) and σ (0) given by ( 92)

Fig. 18
figure 18

Plot of x-component of the velocity, h given by (111), δ = 0.1, θ = 0.02, and B n = 5 , ΔP = 10.5

5.9 Numerical Simulations

Setting \(F=P_{x}^{\left ( 0\right ) }\), the elliptic problem (97) can be transformed in the following integral equation

$$\displaystyle \begin{aligned} F=-\varDelta P\dfrac{\exp \left\{ -\operatorname*{\displaystyle \int}_{0}^{x}{ \displaystyle{{\displaystyle {6h_{x^{\prime }} \over \ell _{F}}}}}dx^{\prime }\right\} }{ \operatorname*{\displaystyle \int}_{0}^{1}\exp \left\{ -\mathop{\displaystyle \int} _{0}^{x}{\displaystyle{{\displaystyle {6h_{x^{\prime }} \over \ell _{F}}}}}dx^{\prime }\right\} dx}, {} \end{aligned} $$
(113)

where, recalling (95),

$$\displaystyle \begin{aligned} \ell _{F}=\min \left\{ h\left( x\right) ,~3h\left( x\right) +{{{\displaystyle {2 \operatorname*{\displaystyle \int}_{0}^{1}Fh~dx-\mathsf{Bn} \over \varDelta P}}}}\right\} {.} \end{aligned}$$

Now, if the conditions ensuring that \(\ell ^{\left ( 0\right ) }\) is strictly positive (Sect. 5.6) are fulfilled, we can solve (113) through the following iterative procedure:

Step j = 0. Set F 0 = −ΔP, and \(\ell _{F,~0}=\min \left \{ h\left ( x\right ) ,~3h\left ( x\right ) -{{{\displaystyle { {\displaystyle {\mathsf {Bn} \over \varDelta P}}}}-}}2 \mathop {\displaystyle \int } _{0}^{1}h~dx\right \} .\)

Step j = 1. \(F_{1}=-\varDelta P{\displaystyle {\dfrac {\exp \left \{ - \operatorname *{\displaystyle \int }_{0}^{x}{\displaystyle {{\displaystyle { 6h_{x^{\prime }} \over \ell _{F,~0}}}}}dx^{\prime }\right \} }{ \operatorname *{ \displaystyle \int }_{0}^{1}\exp \left \{ - \operatorname *{\displaystyle \int }_{0}^{x}{ \displaystyle {{\displaystyle {6h_{x^{\prime }} \over \ell _{F,~0}}}}}dx^{\prime }\right \} dx} }}\).

……

Step j > 1. \(F_{j}=-\varDelta P{\displaystyle {\dfrac {\exp \left \{ - \operatorname *{\displaystyle \int }_{0}^{x}{\displaystyle {{\displaystyle { 6h_{x^{\prime }} \over \ell _{F,~j-1}}}}}dx^{\prime }\right \} }{ \operatorname *{\displaystyle \int }_{0}^{1}\exp \left \{ - \mathop {\displaystyle \int } _{0}^{x}{\displaystyle {{\displaystyle {6h_{x^{\prime }} \over \ell _{F,~j-1}}}}}dx^{\prime }\right \} dx}}}\), with

$$\displaystyle \begin{aligned} \ell _{F,\ j-1}=\min \left\{ h\left( x\right) ,~3h\left( x\right) +{{{ \displaystyle{{\displaystyle {2\operatorname*{\displaystyle \int}_{0}^{1}F_{j-1}~h~dx-\mathsf{ Bi} \over \varDelta P}}}}}}\right\} . \end{aligned}$$

Iterating the procedure until the desired tolerance is reached, we determine the solution \(F=P^{(0)}_{x}\). Integration then provides the pressure field P (0). We can show that, under suitable hypotheses, the solution of (113) exists and is unique.

In Figs. 16, 17 we have plotted h(x) and σ (0)(x) for the wavy channel profile given by (111). In Figs. 18, 19 we have reported the contour plots of \(v_{1}^{(0)}\) , and \(v_{2}^{(0)}\), when h(x) is given by (111), with δ = 0.1, θ = 0.02, and B n = 5.

Fig. 19
figure 19

Plot of y-component of the velocity, h given by (111), δ = 0.1, θ = 0.02, and B n = 5 , ΔP = 10.5

Fig. 20
figure 20

Plot of σ (0) and h, when h is given by ( 103). B n = 5, ΔP = 10.5

The solid colored regions of Figs. 18 and 19 denote the core, with vanishing transversal velocity and uniform longitudinal velocity. Notice also the symmetry of the transversal velocity shown in Fig. 19. In Figs. 20, 21, 22 we have considered the profile (103). The yield surface σ (0) and the velocities \(v_{1}^{\left ( 0\right ) }\), \(v_{2}^{\left ( 0\right ) }\) are reported respectively.

Fig. 21
figure 21

Plot of x-component of the velocity, h given by (103). B n = 5, ΔP = 10.5

Fig. 22
figure 22

Plot of y-component of the velocity, h given by (103). B n = 5, ΔP = 10.5

5.10 Model with Pressure Dependent Viscosity

In this section we extend our model to the case of a pressure-dependent viscosity. Going back to dimensional variables equation (65) rewrites in this way

$$\displaystyle \begin{aligned} \mathbf{D}={\displaystyle {II_{\mathbf{D}} \over 2\mu \left(P\right) II_{\mathbf{D}}+{\tau _{o}}}}\mathbf{S}. \end{aligned}$$

The viscosity is expanded considering

$$\displaystyle \begin{aligned} \mu (P)=\mu \Big(P^{(0)}+\varepsilon P^{(1)}+\varepsilon ^{2}P^{(2)}+\ldots\ldots.\Big), \end{aligned}$$

so that, around ε = 0, we get μ = μ (0) + εμ (1) + ε 2 μ (2) + …, where

$$\displaystyle \begin{aligned} \mu ^{(0)}=\mu (P^{(0)}),\;\;\;\;\;\;\mu ^{(1)}={\displaystyle{{\displaystyle { d\mu \over dP}}}}(P^{(0)})\;P^{(1)}. {} \end{aligned} $$
(114)

Following the same procedure described in Sect. 5.5, the non-dimensional leading order problem becomes

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} {\displaystyle{{\displaystyle {\partial v_{1}^{(0)} \over \partial x}}}}+{\displaystyle{ {\displaystyle {\partial v_{2}^{(0)} \over \partial y}}}}=0, \\ \\ -{\displaystyle{{\displaystyle {\partial P^{(0)} \over \partial x}}}}+{\displaystyle{{\displaystyle { \partial \over \partial y}}}}\left( \mu ^{(0)}\left(P^{\left( 0\right) }\right) {\displaystyle{{\displaystyle {\partial v_{1}^{(0)} \over \partial y}}}}\right) =0, \\ \\ -{\displaystyle{{\displaystyle {\partial P^{(0)} \over \partial y}}}}=0, \end{array} \right. \end{aligned}$$

whose boundary conditions are still given by (86). We get

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} v_{1}^{(0)}={{{\displaystyle{{\displaystyle {P_{x}^{(0)} \over \mu ^{(0)}\left(P^{\left( 0\right) }\right) }}}}{\displaystyle{{\displaystyle {(y-h^{(0)})(y-2\sigma ^{(0)}+h^{(0)}) \over 6}}}}}}, \\ \\ v_{2}^{(0)}={{{\displaystyle{{\displaystyle {\partial \over \partial x}}}}}}\left[ { \displaystyle{{\displaystyle {P_{x}^{(0)} \over \mu ^{(0)}\left(P^{\left( 0\right) }\right) }}}}{{{\displaystyle{{\displaystyle {(y-h^{(0)})^{2}(y-3\sigma ^{(0)}+2h^{(0)}) \over 6}}}}}}\right] , \end{array} \right. \end{aligned}$$

and

$$\displaystyle \begin{aligned} k_{1}^{(0)}=-{{{\displaystyle {P_{x}^{(0)} \over \mu (P^{(0)})}}}}{{{\displaystyle { (h^{(0)}-\sigma ^{(0)})^{2} \over 2}}.}} \end{aligned}$$

The interface \(\sigma ^{\left ( 0\right ) }\) is still given by (92 ), while Eq. (97) modifies in this way

$$\displaystyle \begin{aligned} \left( {\displaystyle {P_{x}^{(0)} \over \mu (P^{(0)})}}\right) _{x}+6{\displaystyle {h_{x} \over \ell ^{\left( 0\right) }}}{\displaystyle {P_{x}^{(0)} \over \mu (P^{(0)})}}=0, {} \end{aligned} $$
(115)

where \(\ell ^{\left ( 0\right ) }\) is given by (95).

Example 6

In case \(\mu \left ( P\right ) =e^{\gamma P}\), and h ≡ 1, we get

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} v_{1}^{(0)}={\displaystyle{{\displaystyle {[e^{-\gamma P_{in}}-e^{-\gamma P_{out}}] \over \gamma }}}}\left[ {\displaystyle{{\displaystyle {(y-\sigma ^{(0)})^{2} \over 2}}}}-{ \displaystyle{{\displaystyle {(1-\sigma ^{(0)})^{2} \over 2}}}}\right] , \\ \\ v_{2}^{(0)}=0, \\ \\ \sigma ^{(0)}={\displaystyle{{\displaystyle {\mathsf{Bn} \over \varDelta P}}}}, \\ \\ P=P_{in}-{\displaystyle{{\displaystyle {1 \over \gamma }}}}\ln \left[ 1+\left( e^{\gamma \varDelta P}-1\right) x\right] . \end{array} \right. {} \end{aligned} $$
(116)

We now consider \(h^{\left ( 0\right ) }=1+mf\left ( x\right ) \), with m small perturbation. We look for a solution of (115) of the form

$$\displaystyle \begin{aligned} P^{(0)} =P_{in}-{\displaystyle{{\displaystyle {1 \over \gamma }}}}\ln \left[ 1+\left( e^{\gamma \varDelta P}-1\right) x\right] +m\varPi , {} \end{aligned} $$
(117)

with Π = 0 on x = 0, 1. After inserting ( 117) into (115) and neglecting the m 2, we find

$$\displaystyle \begin{aligned} \varPi =-{{{\displaystyle {6 \over \gamma }}{\displaystyle {\left( e^{\gamma \varDelta P}-1\right) \over 1+\left( e^{\gamma \varDelta P}-1\right) x}}}}\left[ x\int_{0}^{1}f(\xi )d\xi -\int_{0}^{x}f(\xi )d\xi \right] , \end{aligned} $$

and

$$\displaystyle \begin{aligned} \sigma ^{(0)}={{{\displaystyle {\mathsf{Bn} \over \varDelta P}}}}-m\left[ 2f(x)-{{{\displaystyle {2 \over \gamma \varDelta P}}}}\int_{0}^{1}{{{\displaystyle {f(\xi )\left( e^{\gamma \varDelta P}-1\right) \over 1+\left( e^{\gamma \varDelta P}-1\right) \xi }}}}d\xi \right] . \end{aligned} $$

6 Planar Squeeze

We consider the flow of an incompressible Bingham fluid placed between parallel plates of length in a channel of length L. The gap between the plates occupied by the fluid has amplitude 2h(t), as depicted in Fig. 23 (see also [19]). Because of symmetry, we confine our analysis to the upper part of the layer, namely \([0,h\left ( t\right ) ]\). The velocity field is v = u(x, y, t)i + v(x, y, t)j, where x, y are the longitudinal and transversal coordinate respectively.

Fig. 23
figure 23

A schematic representation of the squeezing channel

We assume that the region where II S ≥ τ o (yielded) and the region where II S ≤ τ o (unyielded) are separated by a sharp interface y = ±Y (x, t) representing the yield surface. We also define the inner plug

$$\displaystyle \begin{aligned} \varOmega _{p}=\left\{ (x,y):\;\;x\in \lbrack 0,L],\;\;\;y\in \lbrack -Y ,Y ]\right\} . \end{aligned}$$

We may have Y (x, t) = 0 for some \(x\in \left ( 0,L\right ) \) and for some t, so that Ω p becomes a segment of zero measure. The rigid plug Ω p moves uniformly with velocity

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} u=u_{p}(t), \\ \\ v=0,\ \ \ \text{(by}\ \ \text{symmetry).} \end{array} \right. {} \end{aligned} $$
(118)

Neglecting inertia and body forces, the governing equations in the viscous phase are

$$\displaystyle \begin{aligned} \mathrm{tr}\mathbf{D}=0, \end{aligned}$$

and

$$\displaystyle \begin{aligned} -{\displaystyle{{\displaystyle {\partial P \over \partial x}}}}+{ \displaystyle{{\displaystyle {\partial S_{11} \over \partial x}}}}+{ \displaystyle{{\displaystyle {\partial S_{12} \over \partial y}}=0}}, {} \end{aligned} $$
(119)
$$\displaystyle \begin{aligned} -{\displaystyle{{\displaystyle {\partial P \over \partial y}}}}+{ \displaystyle{{\displaystyle {\partial S_{12} \over \partial x}}}}+{ \displaystyle{{\displaystyle {\partial S_{22} \over \partial y}}=0}}, {} \end{aligned} $$
(120)

The integral momentum balance for the domain Ω p is given by

$$\displaystyle \begin{aligned} \int_{\varOmega _{p}}{\displaystyle{{\displaystyle {\partial \over \partial t}}}}\left( \rho \mathbf{v}\right) dV+\int_{\partial \varOmega _{p}}\rho \mathbf{v} \left( \mathbf{v}\cdot \mathbf{n}\right) dS=\int_{\partial \varOmega _{p}}(\mathbf{T}\mathbf{n })dS, {} \end{aligned} $$
(121)

where ρ is the material density. Neglecting the inertial terms, we get following equationFootnote 8

$$\displaystyle \begin{aligned} \int_{0}^{L}\left[ -Y _{x}T_{11}+T_{12}\right] _{Y ^{+}~}dx+P_{Y _{0}}Y _{0}-P_{Y _{1}}Y _{1}=0. {} \end{aligned} $$
(122)

where \(P_{Y_{o}}\), \(P_{Y _{1}}\) represent the normal stresses on x = 0 and x = L. As usual we impose

$$\displaystyle \begin{aligned} \left. \boldsymbol{v}\right\vert _{y=h}\cdot \mathbf{t}=0,\;\;\;\;\;\left( \left. \mathbf{v}\right\vert _{y=h}-\mathbf{w}\right) \cdot \mathbf{n}=0, {} \end{aligned} $$
(123)

where w ⋅n is the wall normal velocity and t is the wall tangent vector. On Y  we write

$$\displaystyle \begin{aligned} \left. \llbracket\mathbf{v}\rrbracket\right. _{y=Y }=0, {} \end{aligned} $$
(124)
$$\displaystyle \begin{aligned} \left. \llbracket\mathbf{T}\mathbf{n}\cdot \mathbf{t} \rrbracket\right. _{y=Y }=0,\;\;\;\;\;\;\;\;\;\left. \llbracket\mathbf{T}\mathbf{n}\cdot \mathbf{n}\rrbracket \right. _{y=Y }=0, {} \end{aligned} $$
(125)

while at x = 0

$$\displaystyle \begin{aligned} \left\{\begin{array}{l} u=0,\\ \\ S_{12}=0, \end{array} \right. \end{aligned} $$
(126)

6.1 Squeezing Between Parallel Plates

We assume \(h=h\left ( t\right ) \) and we set

$$\displaystyle \begin{aligned} H=\max_{t^\geq 0}h(t). \end{aligned}$$

We define the aspect ratio ε = HL , assuming ε ≪ 1. Then we rescale the problem using the following non dimensional variables

$$\displaystyle \begin{aligned} \tilde{x}={\displaystyle{{\displaystyle {x \over L}}}},\;\;\;\;\;\;\;\;\;\tilde{y}={ \displaystyle{{\displaystyle {y \over \varepsilon L}}}},\;\;\;\;\;\;\;\;\; \tilde{Y} ={\displaystyle{{\displaystyle {Y \over \varepsilon L}}}} ,\;\;\;\;\;\;\;\;\;\tilde{h}={\displaystyle{{\displaystyle {h \over H}}}} ,\;\;\;\;\;\;\;\;\;\tilde{t}={\displaystyle{{\displaystyle {t \over T}},}} \end{aligned}$$

where T is the characteristic time scale, i.e. the “squeezing time”. We define the characteristic transversal velocity as V = HT, and the longitudinal velocity as U = Vε, so that u = uU, v = vV = v ∕(εU). Pressure is again rescaled exploiting the Poiseuille formula P c = (μLU)∕H 2 and we set P = PP c, P out = P outP c, where P out is the (given) pressure field applied at the channel outlet. We suppose that P out is constant in time and space. Next we introduce

$$\displaystyle \begin{aligned} \mathbf{S}={\displaystyle{{\displaystyle {\mathbf{S} \over (\mu U/H)}}}},\;\;\;\;\;\mathbf{D}={\displaystyle{{\displaystyle {\mathbf{D} \over (U/H)}}}},\;\;\;\;\;II_{\mathbf{D}}={\displaystyle{ {\displaystyle {II_{\mathbf{D}} \over (U/H)}}}},\;\;\;\;\;II_{ \mathbf{S}}={\displaystyle{{\displaystyle {II_{\mathbf{S}} \over (\mu U/H)}},}} \end{aligned}$$

so that

$$\displaystyle \begin{aligned} \mathbf{S}=\left( 2+{\displaystyle{{\displaystyle {\mathsf{Bn} \over II_{D}}}}}\right) \mathbf{D}, \end{aligned}$$

where

$$\displaystyle \begin{aligned} \mathsf{Bn}={\displaystyle{{\displaystyle {\tau _{o}H \over \mu U}}~}} \end{aligned}$$

is again the Bingham number. The mechanical incompressibility constraint and momentum balance become (neglect the tildas)

$$\displaystyle \begin{aligned} {\displaystyle{{\displaystyle {\partial u \over \partial x}}}}+{\displaystyle{{\displaystyle { \partial v \over \partial y}}}}=0, {} \end{aligned} $$
(127)
$$\displaystyle \begin{aligned} -{\displaystyle{{\displaystyle {\partial P \over \partial x}}}}+\varepsilon {\displaystyle{ {\displaystyle {\partial S_{11} \over \partial x}}}}+{\displaystyle{{\displaystyle {\partial S_{12} \over \partial y}}=0}}, {} \end{aligned} $$
(128)
$$\displaystyle \begin{aligned} -{\displaystyle{{\displaystyle {\partial P \over \partial y}}}}+\varepsilon ^{2}{ \displaystyle{{\displaystyle {\partial S_{12} \over \partial x}}}}+\varepsilon {\displaystyle {{\displaystyle {\partial S_{22} \over \partial y}}=0}}. {} \end{aligned} $$
(129)

Equation (122) can be rewritten as

$$\displaystyle \begin{aligned} \int_{0}^{1}\left[\,PY _{x}-\varepsilon Y _{x}S_{11}+S_{12}\right] _{Y ^{+}}~dx+P_{Y _{0}}Y _{0}-P_{out}Y _{1}=0, {} \end{aligned} $$
(130)

where \(P_{out}=P_{Y _{1}}\). Boundary conditions (123) become

$$\displaystyle \begin{aligned} \left. u\right\vert _{h}=0,\ \ \ \ \ \ \left. v\right\vert _{h}= \overset{\cdot }{h}, {} \end{aligned} $$
(131)

since the squeezing velocity is

$$\displaystyle \begin{aligned}\overset{\cdot }{h}={\displaystyle {\partial h \over \partial t}}<0\end{aligned}$$

Jump conditions on Y  become

$$\displaystyle \begin{aligned} \llbracket u\rrbracket_{y=Y }=\llbracket v\rrbracket_{y=Y }=0, {} \end{aligned} $$
(132)
$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} \llbracket P\rrbracket\left[ 1+\varepsilon ^{2}\left( {\displaystyle{{\displaystyle { \partial Y \over \partial x}}}}\right) ^{2}\right] _{y=Y }+\left[ \varepsilon ^{3}S_{11}\left( {\displaystyle{{\displaystyle {\partial Y \over \partial x}}}}\right) ^{2}-2\varepsilon ^{2}S_{12}\left( {\displaystyle{{\displaystyle {\partial Y \over \partial x}}}}\right) +\varepsilon S_{22}\right] _{y=Y }=0, \\ \\ \llbracket S_{12}\rrbracket_{y=Y }+\varepsilon \left( {\displaystyle{ {\displaystyle {\partial Y \over \partial x}}}}\right) \left[ S_{22}-S_{11}- \varepsilon S_{12}{\displaystyle{{\displaystyle {\partial Y \over \partial x}}}}\right] _{y=Y }=0, \end{array} \right. {} \end{aligned} $$
(133)

while conditions (126) become

$$\displaystyle \begin{aligned} \left\{\begin{array}{l} u=0,\\ \\ S_{12}=0. \end{array} \right. \end{aligned} $$
(134)

6.2 Problem at the Leading Order

As for the channel, we look for a solution expressed as power series of ε, assuming \(\mathsf {Bi=}\mathcal {O }\left ( 1\right ) \). We get

$$\displaystyle \begin{aligned} S_{12}^{(0)}=u_{y}^{(0)}-\mathsf{Bn} \end{aligned} $$

since we are looking for a solution with \(u_{y}^{(0)}<0\) in the upper part of the channel. Equations (127)–(129) reduces to

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} {\displaystyle{{\displaystyle {\partial u^{(0)} \over \partial x}}}}+{\displaystyle{ {\displaystyle {\partial v^{(0)} \over \partial y}}}}=0, \\ \\ -{\displaystyle{{\displaystyle {\partial P^{(0)} \over \partial x}}}}+{\displaystyle{{\displaystyle { \partial \over \partial y}}}}\left( {\displaystyle{{\displaystyle {\partial u^{(0)} \over \partial y}}}}\right) =0, \\ \\ -{\displaystyle{{\displaystyle {\partial P^{(0)} \over \partial y}}}}=0, \end{array} \right. {} \end{aligned} $$
(135)

with boundary conditions

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} \left. {\displaystyle{{\displaystyle {\partial u^{(0)} \over \partial y}}}}\right\vert _{y=Y }=0\\ \\ u^{(0)}(x,h,t)=0 \end{array} \right. {} \end{aligned} $$
(136)

For the sake of simplicity we suppress the superscript (0). Since P = P(x, t) we get

$$\displaystyle \begin{aligned} u=-P_{x}{{{\displaystyle {(h-y)(y-2Y +h) \over 2}}}}. {} \end{aligned} $$
(137)

Moreover exploiting mass conservation

$$\displaystyle \begin{aligned} \overset{\cdot }{h}-v\left(y,t\right) =\int_{y}^{h}{{{\displaystyle {\partial \over \partial x}}}}\left[\,P_{x}{{{\displaystyle {(h-y^{\prime })(y^{\prime }-2Y +h) \over 2}}} }\right] dy^{\prime }~. \end{aligned}$$

Evaluating u, v on Y  and recalling conditions (132), we obtain

$$\displaystyle \begin{aligned} u_p(t)=-P_{x}{{{\displaystyle {(Y -h)^{2} \over 2}}}}, {} \end{aligned} $$
(138)
$$\displaystyle \begin{aligned} \left. v\right\vert _{y=Y }=\overset{\cdot }{h}+{{{\displaystyle {\partial \over \partial x}}}}\left[\,P_{x}{{{\displaystyle {(Y -h)^{3} \over 3}}}}\right] -{\displaystyle {Y _{x} \over 2}}P_{x}(Y -h)^{2}=0. {} \end{aligned} $$
(139)

The plug equation (130) becomes

$$\displaystyle \begin{aligned} -\int_{0}^{1}P_{x}Y dx=\mathsf{Bn}. {} \end{aligned} $$
(140)

Recalling (134) we have u y = 0 in x = 0 implying \(\left . P_{x}\right \vert _{x=0}=0\). The solid region must be detached from x = 0, since otherwise u p ≡ 0, i.e. no rigid domain motion. Accordingly there must be some \(s\left ( t\right ) \in \lbrack 0,1]\), not a priori known, such that \(Y \left ( x,t\right ) \equiv 0\), for \(0\leq x\leq s\left ( t\right ) \). Hence the spatial domain \(\left [ 0,1\right ] \) can be split in two sub-domains (see Fig. 23):

  • \(0\leq x\leq s\left ( t\right ) \), where Y ≡ 0;

  • \(s\left ( t\right ) <x\leq 1\), where Y  does not vanish.

Assuming that the longitudinal velocity is continuous across \( s\left ( t\right ) \), we have

$$\displaystyle \begin{aligned} u_p(t)=-{\displaystyle {P_{x}\left( s,t\right) \over 2}}(\underset{0}{\underbrace{ Y \left( s,t\right) }}-h)^{2}=-{\displaystyle {P_{x}\left( s,t\right) \over 2}}h^{2}, {} \end{aligned} $$
(141)

where \(P_{x}\left ( s,t\right ) \) is unknown at this stage. From (139) we get

$$\displaystyle \begin{aligned} -\overset{\cdot }{h}+{\displaystyle {2 \over 3}}{{{\displaystyle {\partial \over \partial x}}}}\Big[- \underset{u_p(t)}{\underbrace{{\displaystyle {P_{x} \over 2}}(Y -h)^{2}}} ~(Y -h)\Big]-Y _{x}\underset{u_p(t)}{\underbrace{\left[ - {\displaystyle {P_{x} \over 2}}(Y -h)^{2}\right] }}~=0, \end{aligned}$$

that is

$$\displaystyle \begin{aligned} \overset{\cdot }{h}+{{{\displaystyle {1 \over 3}}}}u_p(t)Y _{x}=0, {} \end{aligned} $$
(142)

In order to avoid physical inconsistencies, we set

$$\displaystyle \begin{aligned} Y \left( x,t\right) =\max \left\{ 0,-{\displaystyle{{\displaystyle {3\dot{h}s(t) \over u_p(t)}}}}\left( {\displaystyle {x \over s\left( t\right) }}-1\right) \right\}. {} \end{aligned} $$
(143)

The local instantaneous discharge is given by

$$\displaystyle \begin{aligned} Q\left( x,t\right) =\int_{0}^{Y }u_p dy+\int_{Y }^{h}udy=u_pY -{\displaystyle {P_{x}\left( x,t\right) \over 3}}\left( h-Y \right) ^{3}. {} \end{aligned} $$
(144)

Mass conservation then requires \(Q\left ( x,t\right ) =-\overset {\cdot }{h}x\), so that \(Q\left ( s,t\right ) =-\overset {\cdot }{h}s\) so that

$$\displaystyle \begin{aligned} u_p(t)=-{\displaystyle {3 \over 2}}{\displaystyle {\overset{\cdot }{h} \over h}}s, {} \end{aligned} $$
(145)

which is positive since \(\overset {\cdot }{h}<0\). As a consequence

$$\displaystyle \begin{aligned} Y \left( x,t\right) =\max \left\{ 0,~2h\left( t\right) \left( {\displaystyle {x \over s\left( t\right) }}-1\right) \right\} . {} \end{aligned} $$
(146)

Therefore the fluid squeezes out of the channel only if \(Y \left ( 1,t\right ) <h\left ( t\right ) \), namely when \(s\left ( t\right ) >2/3\). In \(x\in \left [ 0,s \right ]\) we have Y = 0 and the pressure fulfills Eq. (139) with the boundary condition \(P_{x}\left ( 0,t\right ) =0\)

$$\displaystyle \begin{aligned} \left\{ \begin{array}{ll} P_{xx}={\displaystyle{{\displaystyle {3 \over h^{3}}}}}\overset{\cdot }{h},\ & 0<x<s,\ \ t\geq 0 \\ & \\ P_{x}\left( 0,t\right) =0 & t\geq 0. \end{array} \right. \end{aligned}$$

Therefore

$$\displaystyle \begin{aligned} P\left( x,t\right) ={\displaystyle{{\displaystyle {3\overset{\cdot }{h} \over 2h^{3}}}}} x^{2}+A\left( t\right) ,\ \end{aligned}$$

with \(A\left ( t\right ) \) still unknown at this stage. Recalling that Y  is linear in x we integrate (138) between x and 1 getting

$$\displaystyle \begin{aligned} P\left( x,t\right) =P_{out}+{\displaystyle{{\displaystyle {3 \over 2}}}}\overset{\cdot }{h} \left( {\displaystyle{{\displaystyle {s \over h}}}}\right) ^{3}\left[ {\displaystyle{{\displaystyle {1 \over 2-3s}}}}-{\displaystyle{{\displaystyle {1 \over 2x-3s}}}}\right] , \;\;\;\;\;\;\;\; s\left( t\right) <x\leq 1. \end{aligned} $$
(147)

Then imposing the continuity of P across x = s we get

$$\displaystyle \begin{aligned} P\left( x,t\right) ={\displaystyle{{\displaystyle{{\displaystyle {3\overset{\cdot }{h} \over 2h^{3}}}}}}}\left( x^{2}-s^{2}\right) +P_{out}-3\overset{\cdot }{h}{ \displaystyle{{\displaystyle {s^{2} \over h^{3}}}}}\left( {\displaystyle{{\displaystyle {s-1 \over 2-3s}}}} \right),\;\;\;\;\;\;\;\; 0\leq x\leq s\left( t\right). \end{aligned} $$
(148)

Finally rewriting (140) as

$$\displaystyle \begin{aligned} \operatorname*{\displaystyle \int}_{s}^{1}P_{x}Y dx=-\mathsf{Bn}, \end{aligned}$$

we get

$$\displaystyle \begin{aligned} f(s)=s^{2}\left[ {\displaystyle {2\left( 1-s\right) \over 3s-2}}+\ln \left( {\displaystyle {3s-2 \over s}} \right) \right] =-{\displaystyle {2 \over 3}}{\displaystyle {\mathsf{Bn}h^{2} \over \overset{\cdot }{h}}}. {} \end{aligned} $$
(149)

Hence, solving (149) we find \(s\left ( t\right ) \) and we are able to determine the pressure field in the whole channel and the rigid domain as well. We observe that s(t) is not a material point so that, in principle, s(t) can also be still (i.e. \(\dot {s}(t)=0\)), while the rigid plug is moving with velocity u p(t). Figure 24 shows the behavior of the function f(s) in the l.h.s. of (149) with s ∈ (2∕3, 1). We easily realize that f(s) is monotonically decreasing for 2∕3 < s ≤ 1 and that its range is \(\left [ 0,+\infty \right ) \). So, given any \(-(2\mathsf {Bn} h^{2})/(3\dot {h})>0\), there exists one and only one s fulfilling (149).

Fig. 24
figure 24

Behavior of \(f \left ( s \right )\) for 2∕3 < s ≤ 1

The force acting on the unit surface of upper plate is

$$\displaystyle \begin{aligned} P\left( t\right) =\int_{0}^{1}P\left( x,t\right) dx=P_{out}+{{{\displaystyle {\overset{ \cdot }{h}s^{3} \over 2h^{3}}}}}{\displaystyle {5-3s \over 2-3s}}-{{{\displaystyle {3\overset{\cdot }{h}s^{3} \over 4h^{3}}}\ln }}\left( {\displaystyle {3s-2 \over s}}\right) . \end{aligned}$$

Exploiting (149) we get

$$\displaystyle \begin{aligned} P\left( t\right) -P_{out}={\displaystyle {\mathsf{Bn} \over 2}}{\displaystyle {s \over h}}-\dot{h}\left( {{ {\displaystyle {s \over h}}}}\right) ^{3}{{{\displaystyle {1 \over (3s-2)}}.}} {} \end{aligned} $$
(150)

Remark 8

When B n → 0, the solution of (149) is simply s = 1, i.e. the solid region does not exists at all (as physically expected for a Newtonian fluid). Furthermore formula (150) reduces to

$$\displaystyle \begin{aligned} P\left( t\right) -P_{out}=-{{{\displaystyle {\dot{h} \over h^{3}}},}} \end{aligned}$$

corresponding to the Newtonian fluid planar squeezing, [17]. These results confirms the physical consistency of our model.

6.3 Numerical Simulation

We perform here some numerical simulations to investigate the behavior of our asymptotic solution at the leading order. To illustrate the dependence of the solution on the Bingham number we consider the cases: B n = 1 and B n = 25. We plot the yield surface Y , the pressure field P and the axial velocity u, assuming that the plates have constant velocity so that

$$\displaystyle \begin{aligned} h(t)=1-t,\;\;\;\;\;\;\;\;\;\;\dot{h}(t)=-1. \end{aligned} $$
(151)

We consider t ∈ [0, 0.6], which guarantees that the plates do not come in touch in the select time interval. We set h f = h(0.6), representing the half gap width at time t f = 0.6 and s f = s(0.6) representing the onset of the rigid plug at time t = 0.6. The yield surface Y  and pressure field P are plotted for different times t belonging to the selected time interval and for x ∈ [0, 1]. The axial velocity u is plotted at time t = 0.6 (i.e. when h = h f) for a finite number of x ∈ [s f, 1] and for y ranging in [0, h f].

In Figs. 25, 26 we have plotted the yield surface Y (x, t) and the upper plate y = h(t) at different times in the time interval [0, t f]. We have plotted the upper plate only for x ∈ [s(t), 1] so that the evolution of the onset of the plug x = s(t) is visible. We notice that the slope of the unyielded plug becomes smaller as s(t) increases, as expected. In Figs. 27, 28 we have plotted the pressure field at different times in the time interval [0, t f] in the whole domain x ∈ [0, 1]. Also for this case the position x = s(t) has been put in evidence. We notice that the pressure within the gap increases as B n increases.

Fig. 25
figure 25

Y  for B n = 1 and h given by (151)

Fig. 26
figure 26

Y  for B n = 25 and h given by (151)

Fig. 27
figure 27

P for B n = 1 and h given by (151)

Fig. 28
figure 28

P for B n = 25 and h given by (151)

In Figs. 29, 30 we have plot the axial velocity profile at time t = 0.6 for some fixed x ∈ [s f, 1]. In particular velocity is plotted for x = 0.69, x = 0.73, x = 0.77, x = 0.81, x = 0.85. As one can easily observe the velocity of the plug is the same for each (x, y) belonging to the plug.

Fig. 29
figure 29

u for B n = 1 and h given by (151)

Fig. 30
figure 30

u for B n = 25 and h given by (151)

Finally in Figs. 31, 32 we have plotted the squeeze force given in (150) for different values of the Bingham number, B n. We have plotted (150) for the linear squeezing (151) and for the exponential squeezing

$$\displaystyle \begin{aligned} h(t)=\exp(-t),\;\;\;\;\;\;\;\;\;\;\dot{h}(t)=-\exp(-t). \end{aligned} $$
(152)

We observe that the linear squeezing requires a grater squeezing force than the exponential squeezing. This is physically consistent, since in the linear case the plates move faster than in the exponential case.

Fig. 31
figure 31

Squeeze force for linear h(t), (151)

Fig. 32
figure 32

Squeeze force for exponential h(t), (152)

6.4 Squeezing Between Surfaces

In this section we generalize the problem to the case in which the parallel plates are surfaces y = ±h(x, t) that are approaching the channel centerline, as shown in Fig. 33. In this case

$$\displaystyle \begin{aligned} H=\max_{\substack{ x\in \left[ 0,L\right] \\ t>0}}h(x,t), \end{aligned}$$

and we again assume HL = ε ≪ 1. The theory develops exactly as in Sect. 6.1, so that (140) still holds. We split \(\left [ 0,1\right ] \) into \(\left [ 0,s\right ] \) and \(\left [ s ,1\right ] \), so that continuity of u across \(s\left ( t\right ) \) yields

Fig. 33
figure 33

A schematic representation of the squeezing channel

$$\displaystyle \begin{aligned} u_p(t)=-{\displaystyle {P_{x}\left( s,t\right) \over 2}}h(s,t)^{2}. \end{aligned}$$

Recalling (139) we find

$$\displaystyle \begin{aligned} h_{t}+{{{\displaystyle {1 \over 3}}}}u_p(t)Y _{x}+{{{\displaystyle {2 \over 3}}}}u_p(t)h_{x}=0, \end{aligned}$$

which generalizes (142). We thus get the following Cauchy problem

$$\displaystyle \begin{aligned} \left\{ \begin{array}{l} Y _{x}={\displaystyle{{\displaystyle {3 \over u_p}}}}\left[ -h_{t}-{ \displaystyle{{\displaystyle {2 \over 3}}}}u_p h_{x}\right] , \\ \\ Y (s,t)=0, \end{array} \right. \end{aligned}$$

whose solution is

$$\displaystyle \begin{aligned} Y (x,t)={\displaystyle{{\displaystyle {3 \over u_p}}}}\left[ - \int_{s}^{x}h_{t}d\xi -{\displaystyle{{\displaystyle {2 \over 3}}}}u_p(h(x,t)-h(s,t)) \right] , {} \end{aligned} $$
(153)

where s is still unknown. Following (143) we set

$$\displaystyle \begin{aligned} Y (x,t)=\max \left\{ 0,~-{{{\displaystyle {3 \over u_p}}}}\int_{s}^{x}h_{t}d\xi -2(h(x,t)-h(s,t))\right\} . {} \end{aligned} $$
(154)

The local discharge is

$$\displaystyle \begin{aligned} Q\left( x,t\right) =\underset{u_pY }{\underbrace{ \int_{0}^{Y }u_p dy}}~+~\underset{-{\displaystyle {P_{x}\left( x,t\right) \over 3}}\left( h-Y \right) ^{3}}{\underbrace{\int_{Y }^{h}\left[ -P_{x} {{{\displaystyle {(h-y)(y-2Y +h) \over 2}}}}\right] dy}}~, \end{aligned}$$

while mass conservation h t + Q x = 0 implies

$$\displaystyle \begin{aligned} Q(x,t)=-\int_{0}^{x}h_{t}d\xi , \end{aligned}$$

since Q(0, t) = 0. We find

$$\displaystyle \begin{aligned} Q(s,t)=-\int_{0}^{s}h_{t}d\xi =\underset{u_p}{\underbrace{-{ \displaystyle{{\displaystyle {P_{x}(s,t)h^{2}(s,t) \over 2}}}}}}{\displaystyle{{\displaystyle {2h(s,t) \over 3}}}}, \end{aligned}$$

implying

$$\displaystyle \begin{aligned} u_p(t)=-{\displaystyle{{\displaystyle {3 \over 2}}}}{\displaystyle{{\displaystyle {1 \over h(s,t)}}}} \int_{0}^{s}h_{t}d\xi , {} \end{aligned} $$
(155)

which is the generalization of (145). In conclusion, substituting (155) into (154), we find

$$\displaystyle \begin{aligned} Y (x,t)=\max \left\{ 0,~-2h(x,t)+2h(s,t){\displaystyle{{\displaystyle { \int_{0}^{x}h_{t}d\xi \over \int_{0}^{s}h_{t}d\xi }}}}\right\} . {} \end{aligned} $$
(156)

In \(x\in \left [ 0,s\right ] \) the pressure fulfils

$$\displaystyle \begin{aligned} \left\{ \begin{array}{ll} -h_{t}+{\displaystyle{{\displaystyle {\partial \over \partial x}}}}\left[\,P_{x}{ \displaystyle{{\displaystyle {h^{3} \over 3}}}}\right] =0,\ & \begin{array}{l} 0<x<s\left( t\right) , \\ t\geq 0, \end{array} \\ & \\ P_{x}\left( 0,t\right) =0, & ~t\geq 0, \end{array} \right. {} \end{aligned} $$
(157)

so that

$$\displaystyle \begin{aligned} P(x,t)=\int_{0}^{x}\left[ {\displaystyle{{\displaystyle {3 \over h(\tilde{x},t)^{3}}}}} \int_{0}^{\tilde{x}}h_{t}d\xi \right] d\tilde{x}+A(t), {} \end{aligned} $$
(158)

with \(A\left ( t\right ) \) to be determined. In \(x\in \left [ s,1\right ] \) we have

$$\displaystyle \begin{aligned} P_x(x,t)=-{\displaystyle {{\displaystyle {2u_p(t) \over (h-Y)^2}}}}, \end{aligned}$$

so that

$$\displaystyle \begin{aligned} P_{x}={{{\displaystyle {{3} \over h(s,t)\left[ h(x,t)-Y (x,t)\right] ^{2}}}}} \int_{0}^{s}h_{t}d\xi , {} \end{aligned} $$
(159)

with Y  given by (156). We observe that (157) and ( 159) yield \(\ \left . P_{x}\right \vert _{s^{-}}=\left . P_{x}\right \vert _{s^{+}}\). Let us now integrate (159) between x and 1 with the boundary condition \(P\left ( 1,t\right ) =P_{out}\). We find

$$\displaystyle \begin{aligned} P_{out}-P\left( x,t\right) ={{{\displaystyle {3 \over h(s,t)}}}}\int_{0}^{s}h_{t}d\xi \left[ \int_{x}^{1}{{{\displaystyle {d~\tilde{x} \over \left( h(\tilde{x},t)-Y (\tilde{x} ,t)\right) ^{2}}}}}\right] {} \end{aligned} $$
(160)

Imposing \(\left . P\right \vert _{s^{-}}=\left . P\right \vert _{s^{+}}\), from ( 158), (160) we find A(t), so that the pressure can be written in terms of s throughout the whole domain. Substituting (156) and (159) into (140) we get

$$\displaystyle \begin{aligned} \int_{s}^{1}{\displaystyle{{\displaystyle {\left[ -2h(\tilde{x},t)+2h(s,t){ \displaystyle{{\displaystyle {\int_{0}^{\tilde{x}}h_{t}d\xi \over \int_{0}^{s}h_{t}d\xi }}}} \right] _{+} \over \left[ h(\tilde{x},t)-\left[ -2h(\tilde{x},t)+2h(s,t){ \displaystyle{{\displaystyle {\int_{0}^{\tilde{x}}h_{t}d\xi \over \int_{0}^{s}h_{t}d\xi }}}} \right] _{+}\right] ^{2}}}}}d\tilde{x}=-{\displaystyle{{\displaystyle {\mathsf{Bn} \over \left[ {\displaystyle{{\displaystyle {3 \over h(s,t)}}}}\int_{0}^{s}h_{t}d\xi \right] }}}}, {}\end{aligned} $$
(161)

which provides an integral equation for the unknown s(t). Equation (161) can be solved once we know the explicit form of the function h(x, t).

When h(x, t) = f(x)g(t), with f ⋅ g > 0, then (161) can be rewritten as

$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle &\displaystyle \left( \int_{0}^{s}fd\xi \right) ^{2}\int_{s}^{1}{\displaystyle{{\displaystyle {\left[ -f(\tilde{x})\left( \int_{0}^{s}fd\xi \right) +f(s)\left( \int_{0}^{\tilde{x} }fd\xi \right) \right] _{+}d\tilde{x} \over \left[ {\displaystyle{{\displaystyle {f(\tilde{x }) \over 2}}}}\left( \int_{0}^{s}fd\xi \right) -\left[ -f(\tilde{x})\left( \int_{0}^{s}fd\xi \right) +f(s)\left( \int_{0}^{\tilde{x}}fd\xi \right) \right] _{+}\right] ^{2}}}}} \\ &\displaystyle &\displaystyle \quad =-{\displaystyle{{\displaystyle {2\mathsf{Bn}f(s)g(t)^{2} \over 3 \dot{g}(t)}}}}. {} \end{array} \end{aligned} $$
(162)

Example 7

Let us consider

$$\displaystyle \begin{aligned} h(x,t)=f(x)g\left( t\right) ,\ \ \text{with}\ \ \ f\left( x\right) =e^{-\beta x},\ \ \;g(t)=e^{-\alpha t}, \end{aligned}$$

where α, and β both positive. Exploiting (162) we find

$$\displaystyle \begin{aligned} {{{\displaystyle {4(1-e^{-\beta s})^{2} \over \beta }}}}\int_{s}^{1}{{{\displaystyle {(e^{-\beta s}-e^{-\beta x}) \over \left[ e^{-\beta x}(3-e^{-\beta s})-2e^{-\beta s}\right] ^{2}}}}}dx=\left( {{{\displaystyle {2\mathsf{Bn} \over 3\alpha }}}}\right) e^{-\beta s}e^{-\alpha t}. {} \end{aligned} $$
(163)

or equivalently

$$\displaystyle \begin{aligned} \underset{\mathcal{G}(s)}{\underbrace{{{{\displaystyle {(1-e^{-\beta s})^{2} \over \beta ^{2}e^{-2\beta s}}}}}\left[ \ln {{{\displaystyle {|2+e^{-\beta }-3e^{\beta(s-1) }| \over |e^{-\beta}-e^{\beta (s-1)}|}}}}+{{{\displaystyle {2(e^{\beta (s-1)}-1) \over 2+e^{-\beta }-3e^{\beta(s-1) }}}}}\right] }}=\left( {\displaystyle{{\displaystyle {2\mathsf{Bn} \over 3\alpha }}}}\right) e^{-\alpha t}, {} \end{aligned} $$
(164)

which is an implicit equation for s(t). Notice that taking the limit β → 0 of the l.h.s. of (164) we recover the l.h.s. of ( 149), as expected. We immediately realize that (164) admits a unique solution \(s(t)\in ( \hat {s}(\beta ),1]\), with

$$\displaystyle \begin{aligned} \hat{s}(\beta )=1+{\displaystyle{{\displaystyle {1 \over \beta }}}}\ln \left[ {\displaystyle{ {\displaystyle {e^{-\beta }+2 \over 3}}}}\right] , \end{aligned}$$

for each value of \(\left ( 2\mathsf {Bn}/3\alpha \right ) e^{-\alpha t}\). In particular it is easy to show that

$$\displaystyle \begin{aligned} {\displaystyle{{\displaystyle {2 \over 3}}}}<\hat{s}(\beta ) <1,\;\;\;\;\;\;\;\;\;\;\forall \beta >0, \end{aligned}$$

so that s ∈ (2∕3, 1) for all t > 0. Recalling (156) we get

$$\displaystyle \begin{aligned} Y (x,t)=2e^{-\alpha t}\left[ {\displaystyle{{\displaystyle {e^{-\beta s}-e^{-\beta x} \over (1-e^{-\beta s})}}}}\right] . \end{aligned} $$
(165)

Clearly Y > 0 for every x > s, and Y = 0 at x = s. Actually we can show that Y  and h never meets. Indeed, suppose that Y (x, t) < h(x, t), then

$$\displaystyle \begin{aligned} 2\left[ {\displaystyle{{\displaystyle {e^{-\beta s}-e^{-\beta x} \over (1-e^{-\beta s})}}}} \right] <e^{-\beta x}, \end{aligned}$$

or analogously

$$\displaystyle \begin{aligned} 2e^{\beta x}<3e^{\beta s}-1. {} \end{aligned} $$
(166)

Hence Y < h if and only if (166) holds true for each x ≥ s . Now recall that \(s\geq \hat {s}(\beta )>2/3\), for every finite time t > 0 and β > 0. Therefore

$$\displaystyle \begin{aligned} 2e^{\beta x}\leq \max_{x\in \lbrack s,1]}\{2e^{\beta x}\}=2e^{\beta }=3e^{\beta \hat{s}}-1<3e^{\beta s}-1, \end{aligned}$$

which proves that (166) holds true. As a consequence we get

$$\displaystyle \begin{aligned} 0\leq Y (x,t)<h(x,t),\;\ \ \;\forall \ \ x\geq s,\;\;\;\;t>0. \end{aligned}$$

We observe that Y → h, which in turn tends to 0, only in the limit t →. In Fig. 34 we have plot the advancing front x = s(t) for different values of B n ranging from B n = 0.1 to B n = 100. The parameters used are α = 2 and β = 0.4.

Fig. 34
figure 34

The advance of the front x = s(t) for B n ranging between 0.1 and 100