Parameter | Dimensional Value | Parameter | Dimensional Value |
The purpose of this work is to carry out investigations of a generalized two-phase model for porous media flow. The momentum balance equations account for fluid-rock resistance forces as well as fluid-fluid drag force effects, in addition, to internal viscosity through a Brinkmann type viscous term. We carry out detailed investigations of a one-dimensional version of the general model. Various a priori estimates are derived that give rise to an existence result. More precisely, we rely on the energy method and use compressibility in combination with the structure of the viscous term to obtain $ H^1 $-estimates as well upper and lower uniform bounds of mass variables. These a priori estimates imply existence of solutions in a suitable functional space for a global time $ T>0 $. We also derive discrete schemes both for the incompressible and compressible case to explore the role of the viscosity term (Brinkmann type) as well as the incompressible versus the compressible case. We demonstrate similarities and differences between a formulation that is based, respectively, on interstitial velocity and Darcy velocity in the viscous term. The investigations may suggest that interstitial velocity seems more natural to use in the formulation of momentum balance than Darcy velocity.
Citation: Yangyang Qiao, Huanyao Wen, Steinar Evje. Compressible and viscous two-phase flow in porous media based on mixture theory formulation[J]. Networks and Heterogeneous Media, 2019, 14(3): 489-536. doi: 10.3934/nhm.2019020
[1] | Yangyang Qiao, Huanyao Wen, Steinar Evje . Compressible and viscous two-phase flow in porous media based on mixture theory formulation. Networks and Heterogeneous Media, 2019, 14(3): 489-536. doi: 10.3934/nhm.2019020 |
[2] | Brahim Amaziane, Leonid Pankratov, Andrey Piatnitski . An improved homogenization result for immiscible compressible two-phase flow in porous media. Networks and Heterogeneous Media, 2017, 12(1): 147-171. doi: 10.3934/nhm.2017006 |
[3] | Steinar Evje, Kenneth H. Karlsen . Hyperbolic-elliptic models for well-reservoir flow. Networks and Heterogeneous Media, 2006, 1(4): 639-673. doi: 10.3934/nhm.2006.1.639 |
[4] | Alexei Heintz, Andrey Piatnitski . Osmosis for non-electrolyte solvents in permeable periodic porous media. Networks and Heterogeneous Media, 2016, 11(3): 471-499. doi: 10.3934/nhm.2016005 |
[5] | Leda Bucciantini, Angiolo Farina, Antonio Fasano . Flows in porous media with erosion of the solid matrix. Networks and Heterogeneous Media, 2010, 5(1): 63-95. doi: 10.3934/nhm.2010.5.63 |
[6] | Michael Eden, Michael Böhm . Homogenization of a poro-elasticity model coupled with diffusive transport and a first order reaction for concrete. Networks and Heterogeneous Media, 2014, 9(4): 599-615. doi: 10.3934/nhm.2014.9.599 |
[7] | Debora Amadori, Stefania Ferrari, Luca Formaggia . Derivation and analysis of a fluid-dynamical model in thin and long elastic vessels. Networks and Heterogeneous Media, 2007, 2(1): 99-125. doi: 10.3934/nhm.2007.2.99 |
[8] | Verónica Anaya, Mostafa Bendahmane, David Mora, Ricardo Ruiz Baier . On a vorticity-based formulation for reaction-diffusion-Brinkman systems. Networks and Heterogeneous Media, 2018, 13(1): 69-94. doi: 10.3934/nhm.2018004 |
[9] | Catherine Choquet, Ali Sili . Homogenization of a model of displacement with unbounded viscosity. Networks and Heterogeneous Media, 2009, 4(4): 649-666. doi: 10.3934/nhm.2009.4.649 |
[10] | Frederike Kissling, Christian Rohde . The computation of nonclassical shock waves with a heterogeneous multiscale method. Networks and Heterogeneous Media, 2010, 5(3): 661-674. doi: 10.3934/nhm.2010.5.661 |
The purpose of this work is to carry out investigations of a generalized two-phase model for porous media flow. The momentum balance equations account for fluid-rock resistance forces as well as fluid-fluid drag force effects, in addition, to internal viscosity through a Brinkmann type viscous term. We carry out detailed investigations of a one-dimensional version of the general model. Various a priori estimates are derived that give rise to an existence result. More precisely, we rely on the energy method and use compressibility in combination with the structure of the viscous term to obtain $ H^1 $-estimates as well upper and lower uniform bounds of mass variables. These a priori estimates imply existence of solutions in a suitable functional space for a global time $ T>0 $. We also derive discrete schemes both for the incompressible and compressible case to explore the role of the viscosity term (Brinkmann type) as well as the incompressible versus the compressible case. We demonstrate similarities and differences between a formulation that is based, respectively, on interstitial velocity and Darcy velocity in the viscous term. The investigations may suggest that interstitial velocity seems more natural to use in the formulation of momentum balance than Darcy velocity.
The importance of multiphase flow in porous media has long been recognized in many fields. Mathematical modelling of multiphase flow is essential in practical applications like enhanced oil recovery and geological CO
The model we are interested in describes flow of two compressible immiscible fluids, e.g., water (w), oil (o), or gas (g), moving in a porous media and takes the following form (we use "w" and "o" in the following as index):
$(ϕn)t+∇⋅(ϕnuo)=Qo,n=soρo(ϕm)t+∇⋅(ϕmuw)=Qw,m=swρwso∇Po+ng=−ˆkouo+ˆkow(uw−uo)+εo∇⋅(n∇uo)sw∇Pw+mg=−ˆkwuw−ˆkow(uw−uo)+εw∇⋅(m∇uw) $
|
(1.1) |
with capillary pressure
$ Pc=Po−Pw=Pc(sw),P′c(sw)<0. $
|
(1.2) |
Herein,
$ so+sw=1. $
|
(1.3) |
Furthermore,
Note that the viscous terms
In the following we will focus on nonlinear coupling mechanisms and we therefore assume physical parameters like porosity
The above model must be endowed with appropriate closure relations for densities
$ ρw−˜ρw0=PwCw,ρo−˜ρo0=PoCo, $
|
(1.4) |
where
Boundary conditions are prescribed as no-flux conditions:
$ ui⋅ν=0,x∈∂Ω,t>0,i=w,o $
|
(1.5) |
where
$ n(x,t=0)=n0(x),m(x,t=0)=m0(x),x∈Ω. $
|
(1.6) |
We may ignore the effects from the viscous terms in (1.1)
$ ˆkidef:=s2iϕμiKkri,i=w,o $
|
(1.7) |
where
$ Uidef:=ϕsiui=−Kkriμi(∇Pi+ρig)=−λi(∇Pi+ρig),λi:=Kkriμi, $
|
(1.8) |
for
● The interaction forces on the RHS of (1.1)
● The viscous terms
$ ∇p+ρg=−α0u+μ∇⋅(∇u), $
|
(1.9) |
where the quantities refer to the fluid phase, i.e.,
Some more precise remarks seem useful in order to set the model (1.1) into a broader context.
Remark 1.1. We may replace the viscous term in (1.1)
$ \nabla\cdot ( \varepsilon_{i, \text{eff}}\nabla \mathbf{u}_i), $ |
where the effective viscosity coefficient
Remark 1.2. An interesting study of a two-phase Brinkman type of model is found in [9]. A two-phase formulation based on Darcy's equations of the following form is studied with
$ st+∇⋅(Uw)=0,Uw=f(s)UT=−λw(s)∇P∇⋅UT=0UT=−λT(s)∇P, $
|
(1.10) |
where
$ −μΔUi+Ui=−λi∇P,i=w,o. $
|
(1.11) |
Summing the two momentum equations in (1.11) gives rise to a Brinkman type of momentum equation for the mixture of the two phases expressed in terms of the total Darcy velocity
$ −μΔUT+UT=−λT(s)∇P. $
|
(1.12) |
Taking the divergence of (1.12) gives in light of the total mass balance equation (1.10)
$ st+∇⋅(Uw)=0,Uw=f(s)UT−∇⋅(λT(s)∇P)=0−μΔUw+Uw=−λw(s)∇P,λw(s)=f(s)λT(s). $
|
(1.13) |
In [9] a notion of weak solutions to the Brinkman model (1.13) is introduced and convergence of an iterative approximation as well as full numerical scheme is demonstrated. Numerical experiments show that the numerical approximation is quite sensitive to the choice of
Remark 1.3. In the literature there seems to be an ongoing interesting discussion of various formulations of two-phase versions of porous media flow based on Brinkman's equation. See for example the work [41] for a discussion of this. Traditionally, the superficial phase velocity (Darcy velocity)
The aim of this paper is three-fold: (ⅰ) Present an example of stability analysis motivated by the study of compressible Navier-Stokes equation (and different from traditional two-phase porous media stability analysis) which exploits the structure of the viscous term in the momentum balance and accounts for compressibility; (ⅱ) Present an example of a numerical scheme both for the compressible and incompressible version of (1.1) in a one-dimensional setting and demonstrate similarities and differences through some specific simulations; (ⅲ) Gain some insight into the impact the viscosity terms have on the solution.
We end this section by giving a brief review of other works on the two-phase porous media model based on Darcy's law. Most works focus on the incompressible, immiscible two-phase flow case. For example, [2] studied the existence of weak solutions for the incompressible two-phase model in fractured porous media based on a dual-porosity formulation. Regularity and stability results were obtained in [8] when analysing a coupled system involving a saturation and a global pressure. In [6,5] the authors showed existence of a solution for an incompressible two-phase flow within a heterogeneous porous medium made of two rock types. Considering dynamic capillary pressure [24] for the incompressible two-phase flow in porous media, [7] proved the existence of a weak solution to a degenerate elliptic parabolic system whereas in [13,12] existence conditions for the traveling wave solution were derived. In particular, non-monotone weak solutions for the Buckley-Leverett equation were obtained. Interesting contributions have also been made concerning the compressible immiscible two-phase flow in porous media where phase densities are assumed to depend on their own pressure. Without using the feature of global pressure, Khalil and Saad [26] established an existence result for a three-dimensional model. In addition, the implicit finite volume scheme was studied in [33] to obtain convergence to a weak solution.
The purpose of this section is to derive a priori estimates of the solution of (1.1). The approach is quite different from the approach used for the incompressible model and formulation based on Darcy velocity
$ (n)t+(nuo)x=−nQp,n=soρo(m)t+(muw)x=−mQp+ρwQI,m=swρwso(Po)x=−ˆkouo+ˆk(uw−uo)+ng+εo(nuox)x,sw(Pw)x=−ˆkwuw−ˆk(uw−uo)+mg+εw(muwx)x,Pc=Po−Pw=Pc(sw), $
|
(2.14) |
subject to the boundary condition
$ uw(x=0,t)=uo(x=0,t)=0 $
|
$ uw(x=1,t)=uo(x=1,t)=0,t>0 $
|
(2.15) |
and initial condition
$ n(x,t=0)=n0(x),m(x,t=0)=m0(x),x∈[0,1]. $
|
(2.16) |
Note that the gravity constant
Definition of
$ mρo+nρw=ρwρo(i.e. ρo=nρwρw−m). $
|
(2.17) |
On the other hand, from (1.2) we have
$ Pc(sw)=Po−Pw=Coρo−Cwρw+Cw˜ρw0−Co˜ρo0. $
|
(2.18) |
Combining (2.18) with (2.17), we get
$ Po−Pw−Pc(sw)=Co(nρwρw−m)−Cwρw+Cw˜ρw0−Co˜ρo0−Pc(mρw)def:=F(ρw;m,n), $
|
(2.19) |
where we have introduced the function
By the definition of
$ F′u(u;m,n)=Co−mn(u−m)2−Cw+P′c(mu)(mu2). $
|
(2.20) |
Since
$ ρo(m,n)=nρwρw−m,sw=mρw,so=1−mρw=nρo. $
|
(2.21) |
For the limit case when
Notation. We first give some notation.
●
● We define
$ ˜m(t)=∫10m(x,t)dx;˜n(t)=∫10n(x,t)dx. $
|
(2.22) |
Assumptions. The following assumptions are made:
● Capillary pressure
We assume that for
$ Φ(sw)≤Pc(˜sw)sw,0≤sw≤1 $
|
(2.23) |
where
$ supsw∈(0,1)|Pc(sw)|<∞,infsw∈(0,1)[−P′c(sw)]≥0 $
|
(2.24) |
and that
$ Cw˜ρw0−Co˜ρo0−supsw∈(0,1)Pc(sw)≥0. $
|
(2.25) |
Note that these constraints on the capillary pressure
● Source terms in (2.14)
● Interaction term
$ ˆkw=Iwm2m+n,ˆko=Ion2m+n,ˆk=Iwomnm+n. $
|
(2.26) |
Remark 2.1. Clearly, in view of (2.14)
$ ˜m(t)=∫10m0(x)dx=˜m0,˜n(t)=∫10n0(x)dx=˜n0 $
|
(2.27) |
where
Remark 2.2. As far as the condition on capillary pressure
$ Φ(sw)=−P∗c∫sw0ln(xa+δ)dx−C1−C2=−P∗ca∫sw/a+δδln(u)du−C1−C2=P∗ca(u−uln(u))|sw/a+δδ−C1−C2=P∗csw+P∗ca[δln(δ)−(sw/a+δ)ln(sw/a+δ)]−C1−C2. $
|
(2.28) |
Since
$ P_c^*a\Bigl[\delta\ln(\delta)-(s_w/a+\delta)\ln(s_w/a + \delta)\Bigr] - C_1 \leq 0, \qquad s_w \in[0, 1] $ |
for an appropriate choice of
$ \Phi(s_w) \leq P_c^*s_w - C_2. $ |
What remains to show then is that
$ P_c^*s_w - C_2 \leq P_c(\tilde{s}_w)s_w, \qquad 0\leq s_w\leq 1. $ |
Clearly,
First, we present a local (in time) existence result whose proof is presented in Appendix A. Then, we state an (almost) global in time existence result which relies on the local existence result combined with certain a priori estimates, see (2.29). Section 2.2 is devoted to these estimates.
Theorem 2.1. (Local existence) Assume that
$
\left\{ Iwok1εwk0+Iwok1εok0≤14,max{Iwok1k0εo+Eεo,1,Iwok1k0εw+Eεw,1}≤12, \right.
$
|
where
$
\left\{ Eεw,1=1εw[10C(k0)2(1+10Ck0)Iwok12k0+10IwIwok1εwk0+20IwIwok1εwk0+20IwIwok1εwk0],Eεo,1=1εo[10C(k0)2(1+10Ck0)Iwok12k0+20(Iwo)2k1εok0+10IoIwok1εok0+20(Iwo)2k1εok0], \right.
$
|
where
$
(m,n)∈C([0,T0];H1)∩C1([0,T0];L2),(uw,uo)∈C([0,T0];H2∩H10),infm>0,infn>0.
$
|
Now we are in the position to state our second result on the almost global existence.
Theorem 2.2. (Almost global existence) In addition to the assumptions of Theorem 2.1, for any given
$ (m, n)\in C([0, T];H^1)\cap C^1([0, T];L^2), (u_w, u_o)\in C([0, T];H^2\cap H_0^1), $ |
where
Moreover, we have the following estimates:
$
∫10[(sw)2x+(so)2x+(ρw)2x+(ρo)2x]dx≤C(T),∫10[(sw)2t+(so)2t+(ρw)2t+(ρo)2t]dx≤C(T),
$
|
for any
Remark 2.3. The constraint
(ⅰ)
Equipped with Theorem 2.1, we are going to prove Theorem 2.2. More precisely, let
1 It means that the solution exists on
$
{‖(m,n,sw,so,ρw,ρo)(t)‖H1+‖(uw,uo)(t)‖H2≤C(T),‖(mt,nt,(sw)t,(so)t,(ρw)t,(ρo)t)(t)‖L2≤C(T),inf(x,t)∈QT∗m(x,t)>0,inf(x,t)∈QT∗n(x,t)>0,
$
|
(2.29) |
for any
To get (2.29), we need the following lemmas. To simplify the proof, we let
(a) Energy estimate.
Lemma 2.1. For any
$ E(t)+∫t0∫10(εwmu2wx+εonu2ox)dxdt+∫t0∫10ˆk(uw−uo)2dxdt+∫t0∫10ˆkwu2wdxdt+∫t0∫10ˆkou2odxdt=E(0), $
|
(2.30) |
where
$ E(t)=Cw∫10m∫ρw˜ρws−˜ρws2dsdx+Co∫10n∫ρo˜ρos−˜ρos2dsdx+∫10[Pc(˜sw)sw−Φ(sw)]dx+∫10∫x0g(n+m)dydx, $
|
(2.31) |
where
Proof. From the two momentum equations of (2.14)
$∫10(εwmu2wx+εonu2ox)dx+∫10ˆk(uw−uo)2dx+∫10ˆkwu2wdx+∫10ˆkou2odx=(∫10nguodx+∫10mguwdx)−∫10(soPoxuo+swPwxuw)dx:=I0+I1. $
|
(2.32) |
For
$
\left\{ ddt∫x0n(y,t)dy=−nuo(x,t),ddt∫x0m(y,t)dy=−muw(x,t). \right.
$
|
Thus we have
$ {I_0} = - \frac{d}{{dt}}\int_0^1 {\int_0^x g } (n + m){\mkern 1mu} dy{\mkern 1mu} dx. $ | (2.33) |
As to
$ ∫10souoPoxdx=Co∫10souo(ρo)xdx=Co∫10nuo(ln(ρo))xdx=−Co∫10(nuo)xln(ρo)dx=Co∫10ntln(ρo)dx=Coddt∫10nln(ρo)dx−Co∫10so(ρo)tdx=Coddt∫10nln(ρo)dx−Coddt∫10ndx+Co∫10(so)tρodx=Coddt∫10nln(ρo)dx+∫10(so)tPodx+Co˜ρo0∫10(so)tdx, $
|
(2.34) |
and, by similar calculations
$∫10swuwPwxdx=Cw∫10swuw(ρw)xdx=Cw∫10muw(ln(ρw))xdx=−Cw∫10(muw)xln(ρw)dx=Cw∫10mtln(ρw)dx=Cwddt∫10mln(ρw)dx−Cw∫10sw(ρw)tdx=Cwddt∫10mln(ρw)dx−Cwddt∫10mdx+Cw∫10(sw)tρwdx=Cwddt∫10mln(ρw)dx+∫10(sw)tPwdx+Cw˜ρw0∫10(sw)tdx. $
|
(2.35) |
Note that here we have also used (2.27). Consequently, using that
$ −I1=Coddt∫10nln(ρo)dx+Cwddt∫10mln(ρw)dx−∫10swtPc(sw)dx+Co˜ρo0ddt∫10sodx+Cw˜ρw0ddt∫10swdx. $
|
That is,
$ −I1=Coddt∫10nln(ρo)dx+Cwddt∫10mln(ρw)dx−∫10Φ(sw)tdx+Co˜ρo0ddt∫10sodx+Cw˜ρw0ddt∫10swdx,Φ′(sw)=Pc(sw). $
|
Moreover, we see that
$ ∫10m∫ρw˜ρws−˜ρws2dsdx=∫10m[ln(s)+˜ρws]|ρw˜ρwdx=∫10m[ln(ρw)−ln(˜ρw)+˜ρwρw−˜ρw˜ρw]dx=∫10mln(ρw)dx−ln(˜ρw)∫10mdx+˜ρw∫10swdx−∫10mdx, $
|
(2.36) |
for some reference density
$ Cwddt∫10mln(ρw)dx=Cwddt∫10m∫ρw˜ρws−˜ρws2dsdx−Cw˜ρwddt∫10swdx $
|
(2.37) |
and
$ Coddt∫10nln(ρo)dx=Coddt∫10n∫ρo˜ρos−˜ρos2dsdx−Co˜ρoddt∫10sodx. $
|
(2.38) |
Thus, it follows that
$ −I1+ddt∫10Φ(sw)dx=Coddt∫10nln(ρo)dx+Cwddt∫10mln(ρw)dx+Co˜ρo0ddt∫10sodx+Cw˜ρw0ddt∫10swdx=Cwddt∫10m∫ρw˜ρws−˜ρws2dsdx+Coddt∫10n∫ρo˜ρos−˜ρos2dsdx+Co[˜ρo0−˜ρo]ddt∫10sodx+Cw[˜ρw0−˜ρw]ddt∫10swdx. $
|
(2.39) |
Note that in view of (1.4), the last line of (2.39) gives us
$ Co[˜ρo0−˜ρo]ddt∫10sodx+Cw[˜ρw0−˜ρw]ddt∫10swdx=−Po(˜ρo)ddt∫10sodx−Pw(˜ρw)ddt∫10swdx=−Po(˜ρo)ddt∫10sodx−Po(˜ρo)ddt∫10swdx+Pc(˜sw)ddt∫10swdx=Pc(˜sw)ddt∫10swdx, $
|
where
(ⅰ)
(ⅱ)
Hence, it follows from (2.39) that
$ −I1=Cwddt∫10m∫ρw˜ρws−˜ρws2dsdx+Coddt∫10n∫ρo˜ρos−˜ρos2dsdx+ddt∫10[Pc(˜sw)sw−Φ(sw)]dx. $
|
(2.40) |
Inserting (2.40) and (2.33) in (2.32) we get
$ Cwddt∫10m∫ρw˜ρws−˜ρws2dsdx+Coddt∫10n∫ρo˜ρos−˜ρos2dsdx+ddt∫10[Pc(˜sw)sw−Φ(sw)]dx+ddt∫10∫x0g(n+m)dydx+∫10(εwmu2wx+εonu2ox)dx+∫10ˆk(uw−uo)2dx+∫10ˆkwu2wdx+∫10ˆkou2odx=0. $
|
(2.41) |
We can rewrite (2.41) to be
$ddtE(t)+∫10(εwmu2wx+εonu2ox)dx+∫10ˆk(uw−uo)2dx+∫10ˆkwu2wdx+∫10ˆkou2odx=0 $
|
(2.42) |
with
Lemma 2.1 implies the following corollary.
Corollary 2.1. For any
$ \int_0^t {\int_0^1 {({\varepsilon _w}mu_{wx}^2 + {\varepsilon _o}nu_{ox}^2)} } {\mkern 1mu} dx{\mkern 1mu} ds + \int_0^t {\int_0^1 [ } \hat k{({u_w} - {u_o})^2} + {\hat k_w}u_w^2 + {\hat k_o}u_o^2]{\mkern 1mu} dx{\mkern 1mu} ds \le {K_0} $ |
where
$
K0=g(˜n0+˜m0)+Cw∫10m0∫ρw0˜ρws−˜ρws2dsdx+Co∫10n0∫ρo0˜ρos−˜ρos2dsdx+∫10[Pc(˜sw)sw0−Φ(sw0)]dx,
$
|
where
Remark 2.4. The energy equality as expressed by (2.30) contains several dissipation terms on its left-hand-side. The three last terms reflect that there is a loss of energy (i.e., an energy transformation) through the three different friction terms, respectively, with coefficients
Remark 2.5. The energy
(b) More regularity estimates.
Lemma 2.2. For any
$ \int_0^1 [\varepsilon_w\frac{m_x^2}{m}+\varepsilon_o\frac{n_x^2}{n}]\, dx\le K_1, $ | (2.43) |
and
$
∫t0∫10so[(ρ1/2o)x]2dxds+∫t0∫10sw[(ρ1/2w)x]2dxds−∫t0∫10Pc′(sw)s2wxdxds≤C(T),
$
|
(2.44) |
where
$
K1=[∫10[εw(m0)2xm0+εo(n0)2xn0]dx+g2(˜m+˜n)T+K0]exp{ˉamax{1εw,1εo}T},
$
|
(2.45) |
and
Proof. Note that from (2.14)
$ (mx)t+(mxuw)x=−(muwx)x. $
|
(2.46) |
Note the appearance of the viscosity term on the RHS of (2.46). Combining (2.46) with (2.14)
$ (εwmx)t+(εwmxuw)x=−εw(muwx)x=−swPwx−ˆkwuw−ˆk(uw−uc)+mg. $
|
This is the same as
$ [m(εwmxm)]t+[m(εwmxm)uw]x=−swPwx−ˆkwuw−ˆk(uw−uc)+mg $
|
or
$ [εwmw]t+[εwmwuw]x=−swPwx−ˆkwuw−ˆk(uw−uc)+mg $
|
for
$ w = \frac{m_x}{m} $ |
which clearly, by using (2.14)
$ εwmwt+εwmuwwx=−swPwx−ˆkwuw−ˆk(uw−uc)+mg. $
|
(2.47) |
Now, we test (2.47) with
$ εw2ddt∫10mw2dx=−∫10swPwxwdx−∫10ˆkwuwwdx−∫10ˆk(uw−uc)wdx+∫10mgwdx $
|
(2.48) |
Similarly, for the oil phase we obtain
$ εo2ddt∫10nv2dx=−∫10soPoxvdx−∫10ˆkouovdx+∫10ˆk(uw−uo)vdx+∫10ngvdx $
|
(2.49) |
with
$ v = \frac{n_x}{n}. $ |
Next, we focus on the terms appearing on the RHS of (2.48):
$ −∫10swPwxwdx=−∫10swPwx(mxm)dx:=Jw,1. $
|
We note that
$ Jw,1=−∫10swPwxmxmdx=−∫10swxPwxdx−∫10swPwxswρwxmdx=−∫10swxPwxdx−4Cw∫10sw[(ρ1/2w)x]2dx. $
|
(2.50) |
Similarly, for
$ Jo,1=−∫10soPoxnxndx=−∫10soxPoxdx−∫10soPoxαoρoxndx=−∫10soxPoxdx−4Co∫10so[(ρ1/2o)x]2dx. $
|
(2.51) |
To conclude, we see that by summing (2.48) and (2.49), using (2.50), and (2.51), we get
$ 12ddt∫10[εwmw2+εonv2]dx+4Co∫10so[(ρ1/2o)x]2dx+4Cw∫10sw[(ρ1/2w)x]2dx=−∫10swxPwxdx−∫10soxPoxdx+∫10mgwdx+∫10ngvdx $
|
$ −∫10ˆk(uw−uo)wdx+∫10ˆk(uw−uo)vdx−∫10ˆkwuwwdx−∫10ˆkouovdx=∫10s2wxP′c(sw)dx+∫10mgwdx+∫10ngvdx−∫10ˆk(uw−uo)wdx+∫10ˆk(uw−uo)vdx−∫10ˆkwuwwdx−∫10ˆkouovdx $
|
(2.52) |
where we again have used
$ 12ddt∫10[εwmw2+εonv2]dx+4Co∫10so[(ρ1/2o)x]2dx+4Cw∫10sw[(ρ1/2w)x]2dx−∫10s2wxP′c(sw)dx=∫10g(mx+nx)dx−∫10ˆk(uw−uo)(mxm)dx+∫10ˆk(uw−uo)(nxn)dx−∫10ˆkwuw(mxm)dx−∫10ˆkouo(nxn)dx:=Kow0+Kw1+Ko1+Kw2+Ko2. $
|
(2.53) |
For
$ Kow0=g∫10mx√m√mdx+g∫10nx√n√ndx≤∫10(m2xm+n2xn)dx+14g2∫10(m+n)dx=∫10(mw2+nv2)dx+14g2(˜m+˜n). $
|
(2.54) |
In the following we make use of the functional form of the interaction coefficients
$ Kw1=−∫10ˆk(uw−uo)(mxm)dx≤14∫10ˆk(uw−uo)2dx+∫10ˆkm2xm2dx≤14∫10ˆk(uw−uo)2dx+Iwo∫10m2xmdx=14∫10ˆk(uw−uo)2dx+Iwo∫10mw2dx $
|
(2.55) |
and
$Ko1=∫10ˆk(uw−uo)(nxn)dx≤14∫10ˆk(uw−uo)2dx+∫10ˆkn2xn2dx≤14∫10ˆk(uw−uo)2dx+Iwo∫10n2xndx=14∫10ˆk(uw−uo)2dx+Iwo∫10nv2dx. $
|
(2.56) |
Furthermore,
$Kw2=−∫10ˆkwuw(mxm)dx≤14∫10ˆkwu2wdx+∫10ˆkwm2xm2dx=14∫10ˆkwu2wdx+Iw∫10m2xmdx=14∫10ˆkwu2wdx+Iw∫10mw2dx $
|
(2.57) |
and
$Ko2=−∫10ˆkouo(nxn)dx≤14∫10ˆkou2odx+∫10ˆkon2xn2dx=14∫10ˆkou2odx+Io∫10n2xndx=14∫10ˆkou2odx+Io∫10nv2dx. $
|
(2.58) |
Putting the estimates (2.54)–(2.58) into (2.53), integrating the result over
$ ∫10[εwmw2+εonv2]dx+4Co∫t0∫10so[(ρ1/2o)x]2dxds+4Cw∫t0∫10sw[(ρ1/2w)x]2dxds−∫t0∫10P′c(sw)s2wxdxds≤∫10[εwm0w20+εon0v20]dx+¯a∫t0∫10(mw2+nv2)dxds+14g2(˜m+˜n)t+12∫t0∫10ˆk(uw−uo)2dxds+14∫t0∫10ˆkwu2wdxds+14∫t0∫10ˆkou2odxds≤∫10[εwm0w20+εon0v20]dx+¯amax{1εw,1εo}∫t0∫10(εwmw2+εonv2)dxds+g2(˜m+˜n)T+K0, $
|
(2.59) |
where
Using Gronwall's inequality and (2.59), we get (2.43). (2.44) is given by (2.43) and (2.59).
In view of Lemma 2.2 it follows that
Corollary 2.2. It holds that
$
\left\{ m(x,t)+n(x,t)≤C(T),∫10(m2x+n2x)dx≤C(T), \right.
$
|
(2.60) |
for any
(c) Upper and lower bounds of density and related quantities.
Lemma 2.3. It holds that
$n≤ρo≤C(T),m≤ρw≤C(T), $
|
(2.61) |
for any
Proof. Since
${\rho _o} = \frac{{{C_w}}}{{{C_o}}}{\rho _w} - \frac{{{C_w}}}{{{C_o}}}{\tilde \rho _{w0}} + {\tilde \rho _{o0}} + \frac{1}{{{C_o}}}{P_c}({s_w}), $ | (2.62) |
we have
$ {\rho _o} = ({s_o} + {s_w}){\rho _o} = n + {s_w}[\frac{{{C_w}}}{{{C_o}}}{\rho _w} - \frac{{{C_w}}}{{{C_o}}}{{\tilde \rho }_{w0}} + {{\tilde \rho }_{o0}} + \frac{1}{{{C_o}}}{P_c}({s_w})], $ | (2.63) |
$ {\rho _w} = ({s_w} + {s_o}){\rho _w} = m + {s_o}[\frac{{{C_o}}}{{{C_w}}}{\rho _o} + {{\tilde \rho }_{w0}} - \frac{{{C_o}}}{{{C_w}}}{{\tilde \rho }_{o0}} - \frac{1}{{{C_w}}}{P_c}({s_w})]. $ | (2.64) |
Armed with the upper bounds of
Corollary 2.3. The following uniform lower bound holds
$ m, n \ge \frac{1}{C}, \qquad for\;\;any\;(x, t) \in [0, 1] \times [0, {T^*}), $ | (2.65) |
subject to the constraint that
Proof.
$ |m(x, t)-\tilde{m}| = |\int_0^1[m(x, t)-m(y, t)]\, dy| = |\int_0^1\int_y^xm_\xi(\xi, t)\, d\xi\, dy|\\ \le \int_0^1|m_x(x, t)|\, dx = \int_0^1\sqrt{m(x, t)}|\frac{m_x(x, t)}{\sqrt{m(x, t)}}|\, dx\\ \le \left(\int_0^1m(x, t)\, dx\right)^\frac{1}{2} \left(\int_0^1\frac{|m_x(x, t)|^2}{m}\, dx\right)^\frac{1}{2} \le \left(\frac{\tilde{m}K_1}{\varepsilon_w}\right)^\frac{1}{2}. $ |
Letting
$ m\ge\frac{1}{C} $ |
on
(d) Higher-order estimates.
Lemma 2.4. The following estimate holds
$ \int_0^1 [ ({s_w})_x^2 + ({s_o})_x^2 + ({\rho _w})_x^2 + ({\rho _o})_x^2]{\mkern 1mu} dx \le C(T) $ | (2.66) |
for any
Proof. Differentiating (2.64) with respect to
$ (\rho_w)_x = m_x+ \frac{C_o}{C_w}(s_o\rho_o)_x + \tilde{\rho}_{w0}(s_o)_x - \frac{C_o}{C_w}\tilde{\rho}_{o0}(s_o)_x \\ - \frac{1}{C_w}(s_o)_xP_c(s_w) - \frac{1}{C_w}s_oP_c^\prime(s_w)(s_w)_x\\ = m_x+ \frac{C_o}{C_w}n_x- (\tilde{\rho}_{w0} - \frac{C_o}{C_w}\tilde{\rho}_{o0})(s_w)_x \\ + \big[\frac{1}{C_w}P_c(s_w) - \frac{1}{C_w}s_oP_c^\prime(s_w)\big](s_w)_x, $ | (2.67) |
where we have used
With (2.67), we proceed to estimate
$ (s_w)_x = (\frac{m}{\rho_w})_x = \frac{m_x}{\rho_w}-\frac{s_w(\rho_w)_x}{\rho_w}\\ = \frac{m_x}{\rho_w}-\frac{s_w}{\rho_w}[m_x+\frac{C_o}{C_w} n_x]+\frac{s_w}{\rho_w} (\tilde{\rho}_{w0} - \frac{C_o}{C_w}\tilde{\rho}_{o0})(s_w)_x\\ -\frac{s_w}{\rho_w}\big[\frac{1}{C_w}P_c(s_w) - \frac{1}{C_w}s_oP_c^\prime(s_w)\big](s_w)_x, $ |
which implies that
$ (s_w)_x\Big[\rho_w-s_w(\tilde{\rho}_{w0} - \frac{C_o}{C_w}\tilde{\rho}_{o0})+s_w\big[\frac{1}{C_w}P_c(s_w) - \frac{1}{C_w}s_oP_c^\prime(s_w)\big]\Big]\\ = m_x-s_w(m_x+\frac{C_o}{C_w} n_x). $ | (2.68) |
By (2.62), we have
$ \rho_w = \frac{C_o}{C_w}\rho_o+\tilde{\rho}_{w0} - \frac{C_o}{C_w}\tilde{\rho}_{o0} - \frac{1}{C_w} P_c(s_w). $ |
Substituting this identity and that
$ (s_w)_x = (s_om_x-s_w\frac{C_o}{C_w} n_x)\Big[\rho_w-s_w(\tilde{\rho}_{w0} - \frac{C_o}{C_w}\tilde{\rho}_{o0})\\ +s_w\big[\frac{1}{C_w}P_c(s_w) - \frac{1}{C_w}s_oP_c^\prime(s_w)\big]\Big]^{-1}. $ | (2.69) |
Since
$ \rho_w = \frac{C_o}{C_w}\rho_o + \tilde{\rho}_{w0} - \frac{C_o}{C_w}\tilde{\rho}_{o0}-\frac{1}{C_w}P_c(s_w), $ | (2.70) |
we have
$ \rho_w-s_w(\tilde{\rho}_{w0} - \frac{C_o}{C_w}\tilde{\rho}_{o0})+s_w\big[\frac{1}{C_w}P_c(s_w) - \frac{1}{C_w}s_oP_c^\prime(s_w)\big]\\ = \frac{C_o}{C_w}\rho_o +s_o\Big[\tilde{\rho}_{w0} - \frac{C_o}{C_w}\tilde{\rho}_{o0}-\frac{1}{C_w}P_c(s_w)\Big]-\frac{1}{C_w}s_ws_oP_c^\prime(s_w). $ | (2.71) |
Combining (2.71) with (2.69), (2.61), (2.65), and the assumptions (2.24) and (2.25), allow us to conclude that
$ \int_0^1(s_w)_x^2\, dx\le C\int_0^1\frac{1}{\rho_o^2}(m_x^2+n_x^2)\, dx \\ \le C\int_0^1\frac{s_o^2}{n^2}(m_x^2+n_x^2)\, dx\le C(T). $ | (2.72) |
This combined with (2.67) and (2.60) gives
$ \int_0^1(\rho_w)_x^2\, dx\le C(T). $ | (2.73) |
By virtue of (2.70), the fact that
Lemma 2.5. For any
$ \int_0^1\big[(u_w)_x^2+(u_o)_x^2\big]\, dx\le C(T). $ | (2.74) |
Proof. By virtue of (2.32), we have
$∫10(εwmu2wx+εonu2ox)dx+∫10[ˆk(uw−uo)2+ˆkwu2w+ˆkou2o]dx=−(∫10nguodx+∫10mguwdx)−∫10(soPoxuo+swPwxuw)dx≤C(‖uo‖L∞+‖uw‖L∞)+C‖(ρo)x‖L2‖uo‖L2+C‖(ρw)x‖L2‖uw‖L2≤C(T)(‖(uo)x‖L2+‖(uw)x‖L2)≤C(T)(‖√n(uo)x‖L2+‖√m(uw)x‖L2)≤12∫10(εwmu2wx+εonu2ox)dx+C(T), $
|
(2.75) |
where we use
Using (2.65) again, and (2.75), we get
$ \int_0^1(u_{wx}^2 + u_{ox}^2)\, dx\le C(T). $ | (2.76) |
Corollary 2.4. For any
$ \int_0^1\big[(u_w)_{xx}^2+(u_o)_{xx}^2\big]\, dx\le C(T). $ | (2.77) |
Proof. From the equation of
$ \varepsilon_wm(u_{w})_{xx} = -\varepsilon_wm_xu_{wx}+s_w P_{wx} + \hat{k}_wu_w +\hat{k}(u_{w}-u_o) + mg, $ |
which together with (2.65), (2.60), (2.66), and (2.74) gives
$ \int_0^1(u_{w})_{xx}^2\, dx\le C(T)\int_0^1m_x^2u_{wx}^2\, dx+C(T)\\ \le C(T)\|u_{wx}^2\|_{L^\infty}\int_0^1m_x^2\, dx+C(T)\\ \le C(T)\Big(\|(u_{w})_x\|_{L^2}^2+\int_0^1|(u_{w})_x(u_{w})_{xx}|\, dx\Big)+C(T) \\ \le \frac{1}{2}\int_0^1(u_{w})_{xx}^2\, dx+C(T), $ |
where we use
$ \int_0^1(u_{w})_{xx}^2\, dx\le C(T). $ |
Similarly, we get the estimate of
Corollary 2.5. For any
$ \int_0^1\Big[m_t^2+n_t^2+(s_w)_t^2+(s_o)_t^2+(\rho_w)_t^2+(\rho_o)_t^2\Big]\, dx\le C(T). $ | (2.78) |
Proof. By using the equation of
$ \int_0^1m_t^2\, dx\le 2\int_0^1\Big[m^2(u_w)_x^2+m_x^2u_w^2\Big]\, dx\\ \le 2\|m\|_{L^\infty}^2\int_0^1(u_w)_x^2\, dx+2\|u_w\|_{L^\infty}^2\int_0^1m_x^2\, dx\\ \le C(T). $ | (2.79) |
Similarly we get
$ \int_0^1n_t^2\, dx\le C(T). $ | (2.80) |
We consider (2.69) with
$ \int_0^1(s_w)_t^2\, dx\le C\int_0^1\frac{1}{\rho_o^2}(m_t^2+n_t^2)\, dx \\ \le C\int_0^1\frac{s_o^2}{n^2}(m_t^2+n_t^2)\, dx\le C(T), $ | (2.81) |
where we use (2.65), (2.79), and (2.80). Since
$ \int_0^1(s_o)_t^2\, dx\le C(T). $ |
By virtue of (2.67) where
$ \int_0^1(\rho_w)_t^2\, dx\le C(T). $ |
This combined with (2.70) and (2.81) gives
$ \int_0^1(\rho_o)_t^2\, dx\le C(T). $ |
With the above estimates, we get (2.29). Thus the proof of Theorem 2.2 is complete.
The main objective of this section is to carry out some testing of the numerical schemes presented in Appendix D, respectively, for the compressible and incompressible version of (1.1). First, we want to test general stability properties. Second, we seek some insight into the role played by using Darcy velocity
$ so(Pw+Pc)x=−(ˆko+ˆk)uo+ˆkuw+ng+εo(ρoUox)xsw(Pw)x=−(ˆkw+ˆk)uw+ˆkuo+mg+εw(ρwUwx)x, $
|
(3.82) |
and modify (4.165)
$ so(Pw+Pc)x=−(ˆko+ˆk)uo+ˆkuw+ng+(εoρo)Uoxxsw(Pw)x=−(ˆkw+ˆk)uw+ˆkuo+mg+(εwρw)Uwxx. $
|
(3.83) |
Third, we also test the behavior of the scheme as the coefficients
We choose parameters as specified in Table 1. In particular, when combined with the relations (4.132) it gives rise to a fractional flow function given by
$ f_w = \frac{k_w/\mu_w}{k_w/\mu_w+k_o/\mu_o} = \frac{s_w^2}{s_w^2 + (1-s_w)^2} = f_w(s_w). $ |
Input parameters of reservoir and fluid properties used for for the below simulations. Note that
Parameter | Dimensional Value | Parameter | Dimensional Value |
The function is illustrated in Fig. 1 (left figure). The initial data of water saturation is set to be as in [9]:
$ sw(x,t=0)={0.80≤x≤2,0.8exp(−150(x−2)2)2<x≤100. $
|
(3.84) |
A horizontal reservoir layer is considered in this case and porosity is also assumed to be 1. The whole test is a 10-day flooding process with a constant interstitial water injection rate at left boundary,
Case 1. First, we want to compare the numerical results obtained by using the scheme from Appendix D (incompressible variant) and compare with similar results presented in [9], which are based on the model (1.13). We also mimic their scheme by slightly modifying the scheme prescribed in Appendix D (incompressible variant) where
Upper row: Results produced by the discrete scheme described in Appendix D (incompressible model). Three kinds of curves are plotted including the case without viscous effect, i.e.,
(ⅰ) A new "water bank" is formed behind the front of the water as a result of the viscous terms. This is a local effect restricted to the region right behind the water front where large gradients in velocity are present;
(ⅱ) Internal viscous forces slow down the transport effect, especially at the saturation front where velocity involves dramatic changes. Right behind the water bank, the model with Darcy velocity involved in the viscous term tends to develop oscillatory behavior;
(ⅲ) The solution shows a clear sensitivity to the magnitude of
The classical Buckley-Leverett model solution (i.e.,
$ \partial_x (\partial_x (s_i u_i)) = \partial_x (s_i \partial_x (u_i)) + \partial_x (u_i \partial_x (s_i)). $ |
Clearly, the viscous term based on
Simulation results with smaller viscous parameters after 10 days of water flooding. Three kinds of curves are compared: zero viscous effect, Darcy velocity
Case 2. In Fig. 4 we show simulation results when we vary the internal relation between
The results after 10 days with initial data are shown in Fig. 1 with interstitial velocity in viscous term. Four curves are compared: the one with large values of
Case 3. Now, we move to another case which has a different initial condition but still with the same injection rate of water,
Initial water saturation profile from Coclite et al. [9]
.Left: The results from Coclite et al. [9] based on Darcy velocity in viscous term. Right: Numerical scheme (after 8 days) which uses interstitial velocity in viscous term with different viscous values
Next, we use the numerical scheme described in Appendix D (compressible variant) to compute and illustrate the numerical behavior for the compressible model. Comparison is made with the cases shown in Fig. 2 for the incompressible model. The compressible fluids are assumed to be given by the pressure-density relation (1.4). The initial water saturation is also the same as shown in Fig. 1. The water saturation
Comparison between the compressible model and the incompressible model for water-oil flow with
In order to shed light on the difference observed in Fig. 7, we explore the pressure profiles at various times (shown in Fig. 8). It is clear from these plots that water pressure keeps increasing, especially for the water displacing part. Water can be compressed in the compressible model therefore the water density will also increase, which leads to a larger viscous effect since density is included in the coefficient of the viscous term. Water will feel more resistance forces and it is more difficult to displace oil resulting in a delay effect.
The water pressure evolution in the compressible model for the case with Darcy velocity in viscous term (left figure) and the case with interstitial velocity in viscous term (right figure). Water pressure increases with time in the water displacing part of the reservoir layer which leads to a compression effect where the magnitude of the viscous terms increase and thereby slows down the displacement of the water front
.Injection of water versus gas. Finally, a numerical example is shown with gas injection to displace oil in the compressible model instead of water. The parameters of gas are the same as for water, as described in Table 1, except using the pressure-density relation:
Left: Comparison of saturation profiles for water injection and gas injection, respectively, after the same time period (10 days) in the compressible model using interstitial velocity in viscous term (
Some main observations from the investigations of the present manuscript are:
● We have found that exploiting the fact that the viscous term depends on the interstitial fluid flow velocity
● We formulated finite difference schemes, both for the incompressible and compressible version of the model. These schemes allow us to systematically gain some insight into the effect of compressibility as well as the effect from the viscous terms that account for the frictional resistance within the fluid. We also observe that by using Darcy velocity in the viscous term, the resulting scheme tends to give more oscillatory behavior similar to that reported in [9].
● In particular, when the viscous coefficients
We are grateful for instructive comments from the anonymous reviewers that helped improving certain parts of a first version of this manuscript.
We apply a method similar to the one used in our previous work [16] to prove the local existence and uniqueness. Hence, in order to make the proof more compact we will heavily refer to that paper for details and highlight what is different.
First, we consider the solution space
$ S: = S_{T_0, A_1} = \Big\{v\in C([0, T_0]; H_0^{1}\cap H^{2})\Big| \|v\|_{C([0, T_0]; H^2)}\le AA_1\Big\} $ |
where
Step 1. Construct an iteration sequence.
We define an iteration sequence to approximate (2.14) which takes the following form
$ (nk)t+(nkuk−1o)x=0(mk)t+(mkuk−1w)x=0skoPkox=−ˆkkouko+ˆkk(uk−1w−uko)+εo(nkukox)x−nkgskwPkwx=−ˆkkwukw−ˆkk(ukw−uko)+εw(mkukwx)x−mkg $
|
(4.85) |
with the initial-boundary value conditions:
$ (ukw,uko)(0,t)=(ukw,uko)(1,t)=0,t≥0, $
|
and
$ (mk,nk)(x,0)=(m0,n0)(x),x∈[0,1], $
|
for
$ (m^k, n^k)\in C([0, T_0];H^1)\cap C^1([0, T_0];L^2), \ (u^k_w, u^k_o)\in C([0, T_0];H_0^{1}\cap H^{2}). $ |
Step 2. Boundedness of the sequence.
Assume that
In fact, as a consequence of that
$
\left\{ ∥uk−1w,x(⋅,t)∥L∞≤∥uk−1w,x(⋅,t)∥W1,1≤AA1∥uk−1o,x(⋅,t)∥L∞≤∥uk−1o,x(⋅,t)∥W1,1≤AA1 \right.
$
|
(4.86) |
for
By virtue of (4.85)
$
\left\{ k0≤mk≤k1,andk0≤nk≤k1on[0,1]×[0,T0],∥∂Pw(mk,nk)∂nk∥L∞([0,1]×[0,T0])≤C,∥∂Pw(mk,nk)∂mk∥L∞([0,1]×[0,T0])≤C,∥∂Po(mk,nk)∂nk∥L∞([0,1]×[0,T0])≤C,∥∂Po(mk,nk)∂mk∥L∞([0,1]×[0,T0])≤C, \right.
$
|
(4.87) |
and
$
\left\{ ∫10|nkx(x,t)|2dx≤C,∫10|mkx(x,t)|2dx≤C. \right.
$
|
(4.88) |
In fact, (4.87)
Multiplying (4.85)
$ \varepsilon_o\int_0^1n^k|u_{ox}^k|^2\, dx = -\int_0^1s_o^k P^k_{ox}{u^k_{o}}\, dx-\int_0^1\hat{k}_o^k|u_o^k|^2\, dx\\ \qquad +\int_0^1\hat{k}^k(u_{w}^{k-1}-u_o^k)u_{o}^k\, dx-\int_0^1n^kgu_{o}^k\, dx \\ \le -\frac{1}{2}\int_0^1\hat{k}_o^k|u_o^k|^2\, dx+\frac{1}{2}\int_0^1\hat{k}^k|u_{w}^{k-1}|^2\, dx -\frac{1}{2}\int_0^1\hat{k}^k|u_{o}^k|^2\, dx+C \\ \le -\frac{1}{2}\int_0^1\hat{k}_o^k|u_o^k|^2\, dx+\frac{I_{wo}k_1}{2}(AA_1)^2 -\frac{1}{2}\int_0^1\hat{k}^k|u_{o}^k|^2\, dx+C, $ | (4.89) |
and
$ \varepsilon_o\int_0^1n^k|u_{oxx}^k|^2\, dx = -\varepsilon_o\int_0^1(n^k)_xu^k_{ox}u_{oxx}^k\, dx+\int_0^1s_o^k P^k_{ox}{u^k_{oxx}}\, dx\\ +\int_0^1\hat{k}_o^ku_o^ku^k_{oxx}\, dx-\int_0^1\hat{k}^k(u_{w}^{k-1}-u_o^k)u_{oxx}^k\, dx+\int_0^1n^kgu_{oxx}^k\, dx\\ \le \frac{\varepsilon_o}{2}\int_0^1n^k|u_{oxx}^k|^2\, dx+\frac{5\varepsilon_o}{2}\int_0^1\frac{1}{n^k}|(n^k)_x|^2|u^k_{ox}|^2\, dx +\frac{C}{\varepsilon_o}\\ +\frac{5}{2\varepsilon_o}\int_0^1\frac{1}{n^k}|\hat{k}_o^k|^2|u_o^k|^2\, dx +\frac{5}{2\varepsilon_o}\int_0^1\frac{1}{n^k}|\hat{k}^k|^2|u_{w}^{k-1}-u_o^k|^2\, dx\\ \le \frac{\varepsilon_o}{2}\int_0^1n^k|u_{oxx}^k|^2\, dx+\frac{5\varepsilon_o}{2k_0}\int_0^1|(n^k)_x|^2|u^k_{ox}|^2\, dx +\frac{C}{\varepsilon_o}+\frac{5I_o}{2\varepsilon_o}\int_0^1\hat{k}_o^k|u_o^k|^2\, dx\\ +\frac{5(I_{wo})^2k_1(AA_1)^2}{\varepsilon_o}+\frac{5I_{wo}}{\varepsilon_o}\int_0^1\hat{k}^k|u_o^k|^2\, dx\\ \le \frac{\varepsilon_o}{2}\int_0^1n^k|u_{oxx}^k|^2\, dx+\frac{5C\varepsilon_o}{2k_0}\Big\||u^k_{ox}|^2\Big\|_{L^\infty} +\frac{C}{\varepsilon_o}+\frac{5I_o}{2\varepsilon_o}\int_0^1\hat{k}_o^k|u_o^k|^2\, dx\\ +\frac{5(I_{wo})^2k_1(AA_1)^2}{\varepsilon_o}+\frac{5I_{wo}}{\varepsilon_o}\int_0^1\hat{k}^k|u_o^k|^2\, dx, $ | (4.90) |
where we have used Cauchy inequality, (4.87), (4.88). Note that we obtain
$ \Big\||u^k_{ox}|^2\Big\|_{L^\infty}\le \int_0^1|u^k_{ox}|^2\, dx+2\|u^k_{ox}\|_{L^2}\|u^k_{oxx}\|_{L^2} \\ \le (1+\frac{1}{\delta})\int_0^1|u^k_{ox}|^2\, dx+\delta\int_0^1|u^k_{oxx}|^2\, dx, $ | (4.91) |
for any small
By virtue of (4.87), (4.89), (4.90), and (4.91), we have
$
\left\{ ∫10|ukox|2dx≤1εo[Iwok12k0(AA1)2+Ck0],∫10ˆkko|uko|2dx+∫10ˆkk|uko|2dx≤Iwok1(AA1)2+2C, \right.
$
|
(4.92) |
and
$ \int_0^1 |u_{oxx}^k|^2\, dx \le\frac{10C}{(k_0)^2}(1+\frac{10C}{k_0})\int_0^1|u^k_{ox}|^2\, dx +\frac{4C}{(\varepsilon_o)^2k_0}\\ +\frac{10I_o}{(\varepsilon_o)^2k_0}\int_0^1\hat{k}_o^k|u_o^k|^2\, dx +\frac{20(I_{wo})^2k_1(AA_1)^2}{(\varepsilon_o)^2k_0}+\frac{20I_{wo}}{(\varepsilon_o)^2k_0}\int_0^1\hat{k}^k|u_o^k|^2\, dx, $ | (4.93) |
where we take
Putting (4.92) into (4.93), we have
$ \int_0^1|u_{oxx}^k|^2\, dx \le\frac{10C}{(k_0)^2}(1+\frac{10C}{k_0})\Big(\frac{I_{wo}k_1}{2\varepsilon_ok_0}(AA_1)^2+\frac{C}{\varepsilon_ok_0}\Big) +\frac{4C}{(\varepsilon_o)^2k_0}\\+ \frac{20(I_{wo})^2k_1(AA_1)^2}{(\varepsilon_o)^2k_0}+\Big[\frac{10I_o}{(\varepsilon_o)^2k_0} +\frac{20I_{wo}}{(\varepsilon_o)^2k_0}\Big](I_{wo}k_1(AA_1)^2+2C)\\ = E_{\varepsilon_o, 1}(AA_1)^2 +E_{\varepsilon_o, 2}, $ | (4.94) |
where
In view of (4.92) and (4.94), we have
$ \|u^k_o(t)\|_{H^2}^2\le \Big[\frac{I_{wo}k_1}{k_0\varepsilon_o}+E_{\varepsilon_o, 1}\Big](AA_1)^2+\frac{2C}{\varepsilon_ok_0}+E_{\varepsilon_o, 2}. $ | (4.95) |
Similar to (4.89) and (4.90), we have
$ \varepsilon_w\int_0^1m^k|u_{wx}^k|^2\, dx+\frac{1}{2}\int_0^1\hat{k}_w^k|u_w^k|^2\, dx+\frac{1}{2}\int_0^1\hat{k}^k|u_w^k|^2\, dx \le \frac{1}{2}\int_0^1\hat{k}^k|u_{o}^k|^2\, dx+C, $ |
and
$ \frac{\varepsilon_w}{2}\int_0^1m^k|u_{wxx}^k|^2\, dx \le \frac{5C\varepsilon_w}{k_0}(1+\frac{10C}{k_0})\int_0^1|u^k_{wx}|^2\, dx +\frac{2C}{\varepsilon_w}+\frac{5I_w}{\varepsilon_w}\int_0^1\hat{k}_w^k|u_w^k|^2\, dx\\ +\frac{10I_w}{\varepsilon_w}\int_0^1\hat{k}^k|u_{w}^{k}|^2\, dx+\frac{10I_w}{\varepsilon_w}\int_0^1\hat{k}^k|u_o^k|^2\, dx, $ |
which yield
$ \int_0^1|u_{wx}^k|^2\, dx \le \frac{1}{\varepsilon_w}\big[\frac{I_{wo}k_1}{2k_0}(AA_1)^2+\frac{2C}{k_0}\big], \\ \int_0^1\hat{k}_w^k|u_w^k|^2\, dx+\int_0^1\hat{k}^k|u_w^k|^2\, dx \le I_{wo}k_1(AA_1)^2+4C, $ | (4.96) |
and
$ \int_0^1|u_{wxx}^k|^2\, dx \le \frac{10C}{(k_0)^2}(1+\frac{10C}{k_0})\frac{1}{\varepsilon_w}\big[\frac{I_{wo}k_1}{2k_0}(AA_1)^2+\frac{2C}{k_0}\big] +\frac{4C}{(\varepsilon_w)^2k_0}\\ +\Big[\frac{10I_w}{(\varepsilon_w)^2k_0}+\frac{20I_w}{(\varepsilon_w)^2k_0} +\frac{20I_w}{(\varepsilon_w)^2k_0}\Big](I_{wo}k_1(AA_1)^2+4C)\\ = E_{\varepsilon_w, 1}(AA_1)^2+E_{\varepsilon_w, 2}. $ | (4.97) |
where
In view of (4.96) and (4.97), we have
$ \|u^k_w(t)\|_{H^2}^2\le \Big[\frac{I_{wo}k_1}{k_0\varepsilon_w}+E_{\varepsilon_w, 1}\Big](AA_1)^2+\frac{4C}{\varepsilon_wk_0}+E_{\varepsilon_w, 2}. $ | (4.98) |
Letting
$ \max\Big\{\frac{I_{wo}k_1}{k_0\varepsilon_o}+E_{\varepsilon_o, 1}, \frac{I_{wo}k_1}{k_0\varepsilon_w}+E_{\varepsilon_w, 1}\Big\}\le\frac{1}{2} $ | (4.99) |
which can be done by assuming for instance that
$ \|u^k_o(t)\|_{H^2}\le AA_1, \, \, \|u^k_w(t)\|_{H^2}\le AA_1, $ | (4.100) |
where we choose that
$ A_1\ge\frac{1}{A}\max\Big\{\Big(\frac{4C}{\varepsilon_ok_0}+2E_{\varepsilon_o, 2}\Big)^\frac{1}{2}, \Big(\frac{8C}{\varepsilon_wk_0}+2E_{\varepsilon_w, 2}\Big)^\frac{1}{2}\Big\} $ | (4.101) |
Consequently, we have
Step 3. Compactness arguments.
By virtue of the estimates uniformly for
$(ukiw,ukio)⇀(uw,uo) weak-* in L∞([0,T0];H10∩H2),nki⇀n weak-* in L∞([0,T0];H1),mki⇀m weak-* in L∞([0,T0];H1),(nkit,mkit)⇀(nt,mt) weak-* in L∞([0,T0];L2) $
|
(4.102) |
as
$ (u_w, u_o)\in L^{\infty}([0, T_0];H^{2}\cap H^{1}_0), \ \text{and}\ (m, n)\in L^{\infty}([0, T_0];H^{1}), $ |
and
$nki→n in C([0,1]×[0,T0]),mki→m in C([0,1]×[0,T0]), $
|
(4.103) |
as
$ k_0\le m\le k_1, \ \text{and}\ k_0\le n\le k_1 \ \text{on}\ [0, 1]\times[0, T_0]. $ |
Step 4. Convergence of
We are going to investigate the convergence of the neighbor sequence of
To begin with, we need the estimates of the difference between
$
\left\{ ˉmk+1t+ˉmk+1xukw+mkx(ukw−uk−1w)+ˉmk+1[ukw]x+mk(ukw−uk−1w)x=0,ˉmk+1(x,0)=0 \right.
$
|
(4.104) |
and
$
\left\{ ˉnk+1t+ˉnk+1xuko+nkx(uko−uk−1o)+ˉnk+1[uko]x+nk(uko−uk−1o)x=0,ˉnk+1(x,0)=0 \right.
$
|
(4.105) |
for
Using (4.104), we have
$ \frac{d}{dt}\int_0^1|\bar{m}^{k+1}|^2\, dx \le \bar{C}\|(u^{k}_{w}-u^{k-1}_{w})_x\|_{L^2}\|\bar{m}^{k+1}\|_{L^2}+\bar{C}A\|\bar{m}^{k+1}\|_{L^2}^2, $ | (4.106) |
where
$ \frac{d}{dt}\int_0^1|\bar{n}^{k+1}|^2\, dx \le \bar{C}\|(u^{k}_{o}-u^{k-1}_{o})_x\|_{L^2}\|\bar{n}^{k+1}\|_{L^2}+\bar{C}A\|\bar{n}^{k+1}\|_{L^2}^2. $ | (4.107) |
Using (4.85)
$ \varepsilon_o[n^k(u^{k}_{o}-u^{k-1}_{o})_x]_{x} = -\varepsilon_o[(n^k-n^{k-1})(u^{k-1}_{o})_x]_{x} +(s_o^{k}-s^{k-1}_o)P^{k}_{ox}\\ +s^{k-1}_o[P_o^{k}-P_o^{k-1}]_x+g(n^k-n^{k-1})+ [\hat{k}_o(m^k, n^k)-\hat{k}_o(m^{k-1}, n^{k-1})]u_o^{k-1}\\ +\hat{k}_o^k(u_o^k-u_o^{k-1})- \hat{k}^k[u_{w}^{k-1}-u_w^{k-2}-(u_o^k-u_o^{k-1})]\\ -[\hat{k}(m^k, n^k)-\hat{k}(m^{k-1}, n^{k-1})](u_{w}^{k-2}-u_o^{k-1}), $ | (4.108) |
where
Denote
$ \|[\bar{u_o}^{k}]_{x}\|_{L^2}^2\le\bar{C}\|\bar{n}^k(u^{k-1}_{o})_x\|_{L^2}^2+\frac{\bar{C}}{(\varepsilon_o)^2 }\|(s^{k}_o-s^{k-1}_o)P^{k}_{ox}\|_{L^1}^2 \\ +\frac{\bar{C}}{(\varepsilon_o)^2 }\|[s^{k-1}_o]_x(P_o^{k}-P_o^{k-1})\|_{L^1}^2+\frac{\bar{C}}{(\varepsilon_o)^2 }\|s^{k-1}_o(P_o^{k}-P_o^{k-1})\|_{L^2}^2 +\frac{\bar{C}}{(\varepsilon_o)^2}\|\bar{m}^k\|_{L^2}^2\\ -\frac{1}{\varepsilon_ok_0}\|\sqrt{\hat{k}_o^k}\bar{u_o}^k\|_{L^2}^2+ \frac{\bar{C}}{(\varepsilon_o)^2}\|\hat{k}_o(m^k, n^k)-\hat{k}_o(m^{k-1}, n^{k-1})\|_{L^2}^2\\ -\frac{1}{2\varepsilon_ok_0}\|\sqrt{\hat{k}^k}\bar{u_o}^k\|_{L^2}^2 +\frac{1}{2\varepsilon_ok_0}\|\sqrt{\hat{k}^k}\bar{u_w}^{k-1}\|_{L^2}^2\\ +\frac{\bar{C}}{(\varepsilon_o)^2} \|\hat{k}(m^k, n^k)-\hat{k}(m^{k-1}, n^{k-1})\|_{L^2}^2, $ | (4.109) |
where
$
\left\{ |sko−sk−1o|≤ˉC(|ˉmk|+|ˉnk|)and|Pko−Pk−1o|≤ˉC(|ˉmk|+|ˉnk|),|ˆko(mk,nk)−ˆko(mk−1,nk−1)|≤ˉC(|ˉmk|+|ˉnk|)and|ˆk(mk,nk)−ˆk(mk−1,nk−1)|≤ˉC(|ˉmk|+|ˉnk|). \right.
$
|
This, together with Hölder inequality, (4.109) and the boundedness of
$ \|[\bar{u_o}^{k}]_{x}\|_{L^2}^2 \le \bar{C}(1+A^2)(\|\bar{m}^k\|_{L^2}^2+\|\bar{n}^k\|_{L^2}^2)-\frac{1}{\varepsilon_ok_0}\|\sqrt{\hat{k}_o^k}\bar{u_o}^k\|_{L^2}^2 \\ -\frac{1}{2\varepsilon_ok_0}\|\sqrt{\hat{k}^k}\bar{u_o}^k\|_{L^2}^2 +\frac{1}{2\varepsilon_ok_0}\|\sqrt{\hat{k}^k}\bar{u_w}^{k-1}\|_{L^2}^2, $ | (4.110) |
where
Similarly, by virtue of (4.85)
$ \varepsilon_w[m^k(u^{k}_{w}-u^{k-1}_{w})_x]_{x}\\ = -\varepsilon_w[(m^k-m^{k-1})(u^{k-1}_{w})_x]_{x} +(s_w^{k}-s^{k-1}_w)P^{k}_{wx}+s^{k-1}_w[P_w^{k}-P_w^{k-1}]_x\\ +g(m^k-m^{k-1})+\hat{k}_w^k(u_w^k-u_w^{k-1})+ [\hat{k}_w(m^k, n^k)-\hat{k}_w(m^{k-1}, n^{k-1})]u_w^{k-1}\\ + \hat{k}^k[u_{w}^{k}-u_w^{k-1}-(u_o^k-u_o^{k-1})]+[\hat{k}(m^k, n^k)-\hat{k}(m^{k-1}, n^{k-1})](u_{w}^{k-1}-u_o^{k-1}), $ | (4.111) |
and
$ \|[\bar{u_w}^{k}]_{x}\|_{L^2}^2 \le \bar{C}(1+A^2)(\|\bar{m}^k\|_{L^2}^2+\|\bar{n}^k\|_{L^2}^2)-\frac{1}{\varepsilon_wk_0}\|\sqrt{\hat{k}_w^k}\bar{u_w}^k\|_{L^2}^2 \\ -\frac{1}{2\varepsilon_wk_0}\|\sqrt{\hat{k}^k}\bar{u_w}^k\|_{L^2}^2 +\frac{1}{2\varepsilon_wk_0}\|\sqrt{\hat{k}^k}\bar{u_o}^{k}\|_{L^2}^2, $ | (4.112) |
where
Putting (4.110) into (4.112), we have
$ \|[\bar{u_w}^{k}]_{x}\|_{L^2}^2 \le \bar{C}_1(\|\bar{m}^k\|_{L^2}^2+\|\bar{n}^k\|_{L^2}^2) +\frac{1}{2\varepsilon_w k_0}\|\sqrt{\hat{k}^k}\bar{u_w}^{k-1}\|_{L^2}^2\\ \le \bar{C}_1(\|\bar{m}^k\|_{L^2}^2+\|\bar{n}^k\|_{L^2}^2) +\frac{I_{wo}k_1}{2\varepsilon_w k_0}\|[\bar{u_w}^{k-1}]_x\|_{L^2}^2, $ | (4.113) |
for
Combining (4.106), (4.107), (4.110) and (4.113) with Cauchy inequality, we have
$ \frac{d}{dt}\int_0^1(|\bar{m}^{k+1}|^2+|\bar{n}^{k+1}|^2)\, dx +\|(\bar{u_w}^{k+1})_{x}\|_{L^2}^2 +\|(\bar{u_o}^{k+1})_{x}\|_{L^2}^2\\ \le \bar{C}_2(\|\bar{m}^{k+1}\|_{L^2}^2+\|\bar{n}^{k+1}\|_{L^2}^2) +\Big(\frac{I_{wo}k_1}{\varepsilon_w k_0} +\frac{I_{wo}k_1}{\varepsilon_ok_0}\Big)(\|(\bar{u_w}^{k})_x\|_{L^2}^2+\|(\bar{u_o}^{k})_x\|_{L^2}^2), $ |
which yields
$ \frac{d}{dt}f^{k+1}(t) + g^{k+1}(t) \le \bar{C}_2 f^{k+1}(t) +\Big(\frac{I_{wo}k_1}{\varepsilon_w k_0} +\frac{I_{wo}k_1}{\varepsilon_ok_0}\Big)g^k(t), $ | (4.114) |
where
$ f^k = \int_0^1(|\bar{m}^{k}|^2+|\bar{n}^{k}|^2)\, dx, \, \, g^k = \|(\bar{u_w}^{k})_{x}\|_{L^2}^2 +\|(\bar{u_o}^{k})_{x}\|_{L^2}^2 $ |
Then, (4.114) together with Gronwall inequality yields
$ f^{k+1}(t)\le \Big(\frac{I_{wo}k_1}{\varepsilon_w k_0} +\frac{I_{wo}k_1}{\varepsilon_ok_0}\Big)\int_0^te^{\bar{C}_2(t-s)}g^k(s)\, ds. $ | (4.115) |
Integrating (4.114) over
$ \int_0^{T_0}g^{k+1}(s)\, ds \le \Big(\bar{C}_2 T_0e^{\bar{C}_2T_0}+1\Big)\Big(\frac{I_{wo}k_1}{\varepsilon_w k_0} +\frac{I_{wo}k_1}{\varepsilon_ok_0}\Big)\int_0^{T_0}g^k(s)\, ds. $ | (4.116) |
Letting
$
\left\{ ˉC2T0eˉC2T0≤1,Iwok1εwk0+Iwok1εok0≤14 \right.
$
|
(4.117) |
for small
$
\left\{ Σ∞k=1∫T00gk+1(t)dt<∞,Σ∞k=1maxt∈[0,T0]fk+1(t)<∞. \right.
$
|
(4.118) |
Thus, (4.118) combined with (4.102)
$ (u_w^{{k_i} - 1}, u_o^{{k_i} - 1}) \rightharpoonup ({u_w}, {u_o})\;\;{\rm{weak - *in}}\;\;{L^\infty }([0, {T_0}];H_0^1 \cap {H^2}) $ | (4.119) |
as
Step 5. Conclusion.
Based on (4.102), (4.103), and (4.119), it can be verified that
The purpose of this section is to elaborate on some of the differences between the model (1.1) and other Brinkman-type of two-phase models like the one mentioned in Remark 1.2.
The incompressible, viscous version of the generalized two-phase model (1.1). It seems useful to impose some simplifying assumptions on the model (1.1) and derive a reduced version of it. We may impose the following assumptions:
● incompressible fluid, i.e.,
● porosity
We may rescale the time such that the porosity disappears in the time derivative terms. The mass and momentum equations in (1.1) now take the form
$(so)t+∇⋅(ϕsouo)=Qoρo,Uo=ϕsouo(sw)t+∇⋅(ϕswuw)=Qwρw,Uw=ϕswuwso[∇Pw+∇Pc+ρog]−εoρo∇⋅(so∇uo)=+ˆkuw−[ˆko+ˆk]uosw[∇Pw+ρwg]−εwρw∇⋅(sw∇uw)=−[ˆkw+ˆk]uw+ˆkuo, $
|
(4.120) |
where we have used that
(4.121) |
That is, we find for
(4.122) |
where
(4.123) |
Note that the resistance force coefficients
$UT=Uw+Uo=−ˆλT∇Pw−ˆλo∇Pc−[ˆλwρw+ˆλoρo]g+(εwϕρw)swˆko+ˆkˆkoˆkw+ˆk[ˆko+ˆkw]∇⋅(sw∇uw)+(εoϕρo)soˆkw+ˆkˆkoˆkw+ˆk[ˆko+ˆkw]∇⋅(so∇uo). $
|
(4.124) |
Remark 4.1. A fundamental difference between the above expression (4.124) for the total superficial velocity
$ Qoρo+Qwρw=−∇⋅(ˆλT∇Pw)−∇⋅(ˆλo∇Pc)−∇⋅([ˆλwρw+ˆλoρo]g)+∇⋅((εwϕρw)swˆko+ˆkˆkoˆkw+ˆk[ˆko+ˆkw]∇⋅(sw∇uw))+∇⋅((εoϕρo)soˆkw+ˆkˆkoˆkw+ˆk[ˆko+ˆkw]∇⋅(so∇uo)). $
|
(4.125) |
This gives an elliptic pressure equation for
We may conclude that it does not seem straightforward to analyse the incompressible, viscous model (4.120) by relying on an approach similar to the one used in [9]. The main reason is the use of interstitial fluid velocity
The incompressible, inviscid case
$ ∇Pw=−UTˆλT−ˆλoˆλT∇Pc−[ˆfwρw+ˆfoρo]g, $
|
(4.126) |
where
$ ˆfw=ˆλwˆλT=ˆfw(sw),ˆfo=ˆλoˆλT=ˆfo(sw) $
|
(4.127) |
are the standard fractional flow functions which satisfy that
(4.128) |
where
$ ˆhdef:=ˆfwˆλo−ϕsoswˆk(ˆkoˆkw+ˆk[ˆko+ˆkw])=s2w(1−sw)2s2wˆko+s2oˆkw+ˆkϕ=ˆh(sw). $
|
(4.129) |
From this, it also follows that
$ Uo=UT−Uw=UT(1−ˆfw(sw))+ˆh(sw)Δρg−ˆh(sw)∇Pc(sw). $
|
(4.130) |
Consequently, the model (4.120) has been reduced to solving the following conservation law
$ ∂tsw+∇⋅Uw=Qwρw, $
|
(4.131) |
where
Remark 4.2. The model (4.131) combined with (4.128) (last line), (4.127), (4.129), and the generalized mobility functions as defined by (4.123), recovers the classical incompressible immiscible model except that we now also include for an additional water-oil drag force effect through the term
$ ˆkw=IwμwKϕsαw,ˆko=IoμoKϕsβw,ˆk=IμwμoKϕsw(1−sw), $
|
(4.132) |
where
Appendix C: A pressure evolution equation. From the two mass balance equations we get after multiplying the
$ ρwsoCoPot+ρoswCwPwt+ρw(soρouo)x+ρo(swρwuw)x=−ρwρosoQp−ρwρoswQp+ρwρoQI $
|
(4.133) |
or
$ κPot−ρoswCwP′cswt+ρw(soρouo)x+ρo(swρwuw)x=ρwρo(QI−Qp), $
|
(4.134) |
with
$ κ=soρw1Co+swρo1Cw=soρw∂ρo∂Po+swρo∂ρw∂Pw. $
|
(4.135) |
Clearly,
$ s_{wt} = -s_{ot} = -\Bigl(\frac{n}{\rho_o}\Bigr)_t = -\frac{1}{\rho_o}n_t + \frac{n}{\rho_o^2}\rho_{ot} = -\frac{1}{\rho_o}n_t + \frac{n}{C_o\rho_o^2}P_{ot}. $ |
Consequently,
$ - \frac{\rho_o s_w}{C_w}P_{c}' s_{wt} = \frac{\rho_o s_w}{C_w}P_{c}' \Bigl[ \frac{1}{\rho_o}n_t - \frac{n}{C_o\rho_o^2}P_{ot} \Bigr] = \frac{s_wP_{c}'}{C_w}n_t - \frac{s_os_wP_{c}'}{C_wC_o}P_{ot}. $ |
Using this in (4.134) we get
$ κPot+swP′cCwnt−soswP′cCwCoPot+ρw(soρouo)x+ρo(swρwuw)x=ρwρo(QI−Qp), $
|
(4.136) |
that is,
(4.137) |
Note that
$ \Bigl[\kappa - \frac{s_os_wP_{c}'\;\;}{\;\;C_wC_o\;\;\;\;\;}\Bigr] = \frac{s_o\tilde{\rho}_w}{C_o} + \frac{s_w\rho_o}{C_w}: = \tilde{\kappa}, \qquad \tilde{\rho}_w = \rho_w-\frac{s_wP_c'}{C_w} $ |
so that
$ ˜κPot+˜ρw(nuo)x+ρo(muw)x=ρwρo(QI−Qp)+swP′cCwnQp $
|
(4.138) |
or
$Pot+˜η˜ρw(nuo)x+˜ηρo(muw)x=˜ηρwρo(QI−Qp)+˜ηswP′cCwnQp,˜η=1˜κ=CwCoso˜ρwCw+swρoCo. $
|
(4.139) |
Similarly, we have for
$ κPwt+ρwsoP′cCoswt+ρw(soρouo)x+ρo(swρwuw)x=ρwρo(QI−Qp). $
|
(4.140) |
Clearly,
$ s_{wt} = \Bigl(\frac{m}{\rho_w}\Bigr)_t = \frac{1}{\rho_w}m_t - \frac{m}{\rho_w^2}\rho_{wt} = \frac{1}{\rho_w}m_t - \frac{m}{C_w\rho_w^2}P_{wt}. $ |
Consequently,
$ \frac{\rho_w s_o}{C_o}P_{c}' s_{wt} = \frac{\rho_w s_o}{C_o}P_{c}' \Bigl[ \frac{1}{\rho_w}m_t - \frac{m}{C_w\rho_w^2}P_{wt} \Bigr] = \frac{s_oP_{c}'}{C_o}m_t - \frac{s_ws_oP_{c}'}{C_wC_o}P_{wt}. $ |
Using this in (4.140) we get
$ κPwt+soP′cComt−soswP′cCwCoPwt+ρw(soρouo)x+ρo(swρwuw)x=ρwρo(QI−Qp), $
|
(4.141) |
that is,
$ [κ−soswP′cCwCo]Pwt+ρw(soρouo)x+[ρo−soP′cCo](swρwuw)x=ρwρo(QI−Qp)−soP′cCo(ρwQI−mQp). $
|
(4.142) |
Note that
$ \Bigl[\kappa - \frac{s_os_wP_{c}'\;\;}{\;\;C_wC_o\;\;\;\;\;}\Bigr] = \frac{s_o\rho_w}{C_o} + \frac{s_w\tilde{\rho}_o}{C_w}: = \tilde{\kappa}, \qquad \tilde{\rho}_o = \rho_o-\frac{s_oP_c'}{C_o}, $ |
so that
$ ˜κPwt+ρw(nuo)x+˜ρo(muw)x=ρwρo(QI−Qp)−soP′cCo(ρwQI−mQp), $
|
(4.143) |
or
$Pwt+˜ηρw(nuo)x+˜η˜ρo(muw)x=˜ηρwρo(QI−Qp)−˜ηsoP′cCo(ρwQI−mQp),˜η=1˜κ=CwCosoρwCw+sw˜ρoCo. $
|
(4.144) |
Appendix D: Discrete schemes for the compressible/incompressible version of (1.1).
A semi-discrete scheme for the compressible model. We consider discrete schemes for both the compressible and incompressible version of (2.14). For that purpose we introduce a reformulation that brings the compressible model closer to the incompressible model. In particular, we solve explicitly only for the mass transport of
The starting point is the model (2.14) with
$ (m)t+(muw)x=−mQp+ρwQIPwt+˜ηρw(nuo)x+˜η˜ρo(muw)x=˜ηρwρo(QI−Qp)−˜ηsoP′cCo(ρwQI−mQp)so(Pw+Pc)x=−(ˆko+ˆk)uo+ˆkuw+ng+εo(nuox)xsw(Pw)x=−(ˆkw+ˆk)uw+ˆkuo+mg+εw(muwx)x $
|
(4.145) |
with
$ ˜η=CwCosoρwCw+sw˜ρoCo,˜ρo=ρo−soP′cCo. $
|
(4.146) |
We refer to Appendix C and (4.144) which gives the pressure evolution equation (4.145)
$ sw=mρw(Pw),so=1−swn=soρo(Po)=(1−mρw(Pw))ρo(Po)=n(m,Pw)Po=Pc(sw)+Pw=Po(m,Pw) $
|
(4.147) |
We solve (4.145) on our domain
$ uw|∂Ω=uo|∂Ω=0 $
|
(4.148) |
and initial condition
$ m(x,t=0)=m0(x),Pw(x,t=0)=Pw(m0(x),n0(x)). $
|
(4.149) |
System of ODEs. We consider the domain
$ x_1 = \frac{1}{2}\Delta x, \quad x_2 = (1+\frac{1}{2})\Delta x, \quad \ldots , \quad x_j = (j-\frac{1}{2})\Delta x, \quad \ldots , \quad x_{N_x} = (N_x-\frac{1}{2})\Delta x $ |
and cell interfaces
$ x_{1/2} = 0, \quad x_{3/2} = \Delta x, \quad \ldots , \quad x_{j+1/2} = j\Delta x, \quad \ldots , \quad x_{N_x+1/2} = N_x\Delta x = 1, $ |
where
Step 1: Mass transport. We solve for
$ dmjdt+1Δx([muw]j+1/2−[muw]j−1/2)=−mjQp,j+ρwjQI,j,m=swρw $
|
(4.150) |
where
$ [muw]j+1/2={mjuw,j+1/2,if uw,j+1/2≥0;mj+1uw,j+1/2,if uw,j+1/2<0. $
|
(4.151) |
This can also be expressed as
$ [m u_w]_{j+1/2} = \frac{m_j+m_{j+1}}{2}u_{w, j+1/2} - \frac{1}{2}(m_{j+1} - m_j)|u_{w, j+1/2}|. $ |
Step 2: Computation of velocities and pressure. Next, we solve for
$ dPwjdt+˜ηjρwj1Δx([nuo]j+1/2−[nuo]j−1/2)+˜ηj˜ρoj1Δx([muw]j+1/2−[muw]j−1/2)=[˜ηρwρo]j(QI,j−Qp,j)−[˜ηsoP′cCo]j(ρwjQI,j−mjQp,j), $
|
(4.152) |
which is combined with the momentum balance equations
$ so,j+1/21Δx(Pw,j+1−Pw,j)=−so,j+1/21Δx(Pc,j+1−Pc,j)−ˆko,j+1/2uo,j+1/2+ˆkj+1/2(uw,j+1/2−uo,j+1/2)+nj+1/2g+εo1Δx2(nj+1[uo,j+3/2−uo,j+1/2]−nj[uo,j+1/2−uo,j−1/2]),sw,j+1/21Δx(Pw,j+1−Pw,j)=−ˆkw,j+1/2uw,j+1/2−ˆkj+1/2(uw,j+1/2−uo,j+1/2)+mj+1/2g+εw1Δx2(mj+1[uw,j+3/2−uw,j+1/2]−mj[uw,j+1/2−uw,j−1/2]). $
|
(4.153) |
Here we note that the average
$ sw,j+1/2={sw,j,if uw,j+1/2>0;sw,j+sw,j+12,if uw,j+1/2=0;sw,j+1, if uw,j+1/2<0. $
|
(4.154) |
Similarly, for
$ ˆkj+1/2={ˆkj,if uw,j+1/2>0 & uo,j+1/2>0 ;ˆkj+ˆkj+12,if uw,j+1/2uo,j+1/2≤0;ˆkj+1, if uw,j+1/2<0 & uo,j+1/2<0. $
|
(4.155) |
Moreover,
A fully discrete scheme. We assume that we have given
Step 1: Mass transport.
$ mk+1j−mkjΔt+1Δx([muw]kj+1/2−[muw]kj−1/2)=−mkjQp,j+ρkwjQI,j, $
|
(4.156) |
where
$ [muw]kj+1/2={mkjukw,j+1/2,if ukw,j+1/2≥0;mkj+1ukw,j+1/2,if ukw,j+1/2<0. $
|
(4.157) |
Having computed
$ sk+1/2w,j=mk+1jρw(Pkw,j). $
|
(4.158) |
Similarly, we compute updated mass
Step 2: Computation of velocities and pressure. Next, we solve simultaneously for
$ Pk+1wj−PkwjΔt+[˜ηρw]k+1/2j1Δx([nk+1/2uk+1o]j+1/2−[nk+1/2uk+1o]j−1/2)+[˜η˜ρo]k+1/2j1Δx([mk+1uk+1w]j+1/2−[mk+1uk+1w]j−1/2)=[˜ηρwρo]k+1/2j(QI,j−Qp,j)−[˜ηsoP′cCo]k+1/2j(ρkwjQI,j−mk+1jQp,j), $
|
(4.159) |
which is combined with the momentum balance equations
$ sk+1/2o,j+1/21Δx(Pk+1w,j+1−Pk+1w,j)=−sk+1/2o,j+1/21Δx(Pk+1/2c,j+1−Pk+1/2c,j)−ˆkk+1/2o,j+1/2uk+1o,j+1/2+ˆkk+1/2j+1/2(uk+1w,j+1/2−uk+1o,j+1/2)+nk+1/2j+1/2g+εo1Δx2(nk+1/2j+1[uk+1o,j+3/2−uk+1o,j+1/2]−nk+1/2j[uk+1o,j+1/2−uk+1o,j−1/2]),sk+1/2w,j+1/21Δx(Pk+1w,j+1−Pk+1w,j)=−ˆkk+1/2w,j+1/2uk+1w,j+1/2−ˆkk+1/2j+1/2(uk+1w,j+1/2−uk+1o,j+1/2)+mk+1j+1/2g+εw1Δx2(mk+1j+1[uk+1w,j+3/2−uk+1w,j+1/2]−mk+1j[uk+1w,j+1/2−uk+1w,j−1/2]). $
|
(4.160) |
Equipped with
(4.161) |
from which we also compute the updated oil mass
Remark 4.3. The upwind discretization of
A semidiscrete scheme for the incompressible model. When fluids are incompressible the model (2.14) takes the following form with unknown variables
$ (sw)t+(swuw)x=−swQp+QI(so)t+(souo)x=−soQpsw(Pw)x=−ˆkwuw−ˆk(uw−uo)+swρwg+εwρw(swuwx)xso(Pw+Pc)x=−ˆkouo+ˆk(uw−uo)+soρog+εoρo(souox)x, $
|
(4.162) |
subject to the boundary condition
$ uw|∂Ω=uo|∂Ω=0 $
|
(4.163) |
and initial condition
$ sw(x,t=0)=sw0(x). $
|
(4.164) |
Note that we can only determine
$ (sw)t+(swuw)x=−swQp+QI(swuw+souo)x=−Qp+QIsw(Pw)x=−ˆkwuw−ˆk(uw−uo)+swρwg+εwρw(swuwx)xso(Pw+Pc)x=−ˆkouo+ˆk(uw−uo)+soρog+εoρo(souox)x. $
|
(4.165) |
This formulation is consistent with and follows directly from (4.145) by letting
Step 1: Mass transport. We solve the following ODE for
$ dsw,jdt+1Δx([swuw]j+1/2−[swuw]j−1/2)=−sw,jQp,j+QI,j $
|
(4.166) |
where
$ [swuw]j+1/2={sw,juw,j+1/2,if uw,j+1/2≥0;sw,j+1uw,j+1/2,if uw,j+1/2<0. $
|
(4.167) |
Step 2: Computation of velocities and pressure. Next, we solve for
$ 1Δx([swuw]j+1/2−[swuw]j−1/2)+1Δx([souo]j+1/2−[souo]j−1/2)=QI,j−Qp,j $
|
(4.168) |
which is combined with the momentum balance equations
$ sw,j+1/21Δx(Pw,j+1−Pw,j)=−ˆkw,j+1/2uw,j+1/2−ˆkj+1/2(uw,j+1/2−uo,j+1/2)+sw,j+1/2ρwg+εwρwΔx2(sw,j+1[uw,j+3/2−uw,j+1/2]−sw,j[uw,j+1/2−uw,j−1/2])so,j+1/21Δx(Pw,j+1−Pw,j)=−so,j+1/21Δx(Pc,j+1−Pc,j)−ˆko,j+1/2uo,j+1/2+ˆkj+1/2(uw,j+1/2−uo,j+1/2)+so,j+1/2ρog+εoρoΔx2(so,j+1[uo,j+3/2−uo,j+1/2]−so,j[uo,j+1/2−uo,j−1/2]). $
|
(4.169) |
Here we note that the average
$ sw,j+1/2={sw,j,if uw,j+1/2>0;sw,j+sw,j+12,if uw,j+1/2=0;sw,j+1, if uw,j+1/2<0. $
|
(4.170) |
Similarly, for
$ ˆkj+1/2={ˆkj,if uw,j+1/2>0 & uo,j+1/2>0 ;ˆkj+ˆkj+12,if uw,j+1/2uo,j+1/2≤0;ˆkj+1, if uw,j+1/2<0 & uo,j+1/2<0. $
|
(4.171) |
Moreover,
A fully discrete scheme for the incompressible model. We can now proceed with a description of a full discrete scheme for the incompressible case which bears clear similarity to the scheme for the compressible model.
Step 1: Mass transport.
$ sk+1w,j−skw,jΔt+1Δx([swuw]kj+1/2−[swuw]kj−1/2)=−skw,jQp,j+QI,j $
|
(4.172) |
where
$ [swuw]kj+1/2={skw,jukw,j+1/2,if ukw,j+1/2≥0;skw,j+1ukw,j+1/2,if ukw,j+1/2<0. $
|
(4.173) |
Having computed
Step 2: Computation of velocities and pressure. We solve for
$ 1Δx([sk+1wuk+1w]j+1/2−[sk+1wuk+1w]j−1/2)+1Δx([sk+1ouk+1o]j+1/2−[sk+1ouk+1o]j−1/2)=QI,j−Qp,j $
|
(4.174) |
which is combined with the momentum balance equations
$ sk+1w,j+1/21Δx(Pk+1w,j+1−Pk+1w,j)=−ˆkk+1w,j+1/2uk+1w,j+1/2−ˆkk+1j+1/2(uk+1w,j+1/2−uk+1o,j+1/2)+sk+1w,j+1/2ρwg+εwρwΔx2(sk+1w,j+1[uk+1w,j+3/2−uk+1w,j+1/2]−sk+1w,j[uk+1w,j+1/2−uk+1w,j−1/2]),sk+1o,j+1/21Δx(Pk+1w,j+1−Pk+1w,j)=−sk+1o,j+1/21Δx(Pk+1c,j+1−Pk+1c,j)−ˆkk+1o,j+1/2uk+1o,j+1/2+ˆkk+1j+1/2(uk+1w,j+1/2−uk+1o,j+1/2)+sk+1o,j+1/2ρog+εoρoΔx2(sk+1o,j+1[uk+1o,j+3/2−uk+1o,j+1/2]−sk+1o,j[uk+1o,j+1/2−uk+1o,j−1/2]). $
|
(4.175) |
Remark 4.4. The upwind discretization of
1. | Jahn O. Waldeland, William J. Polacheck, Steinar Evje, Collective tumor cell migration in the presence of fibroblasts, 2020, 100, 00219290, 109568, 10.1016/j.jbiomech.2019.109568 | |
2. | Yangyang Qiao, Steinar Evje, A compressible viscous three-phase model for porous media flow based on the theory of mixtures, 2020, 141, 03091708, 103599, 10.1016/j.advwatres.2020.103599 | |
3. | Yangyang Qiao, Huanyao Wen, Steinar Evje, Viscous Two-Phase Flow in Porous Media Driven by Source Terms: Analysis and Numerics, 2019, 51, 0036-1410, 5103, 10.1137/19M1252491 | |
4. | Sergei Tantciura, Yangyang Qiao, Pål Ø. Andersen, Simulation of Counter-Current Spontaneous Imbibition Based on Momentum Equations with Viscous Coupling, Brinkman Terms and Compressible Fluids, 2022, 141, 0169-3913, 49, 10.1007/s11242-021-01709-9 | |
5. | Yangyang Qiao, Steinar Evje, A general cell–fluid Navier–Stokes model with inclusion of chemotaxis, 2020, 30, 0218-2025, 1167, 10.1142/S0218202520400096 | |
6. | Gamal B. Abdelaziz, M. Abdelgaleel, Z.M. Omara, A. S. Abdullah, Emad M.S. El-Said, Swellam W. Sharshir, Ashraf M. Elsaid, Mohamed A. Dahab, A simplified model of low Re, immiscible, gas–liquid flow, and heat transfer in porous media numerical solution with experimental validation, 2022, 0891-6152, 1, 10.1080/08916152.2022.2113931 | |
7. | Yangyang Qiao, Pål Østebø Andersen, Sadegh Ahmadpour, 2023, Effective Relative Permeabilities Based on Momentum Equations with Brinkmann Terms and Viscous Coupling, 10.2118/214388-MS | |
8. | Yangyang Qiao, Hans Joakim Skadsem, Steinar Evje, An Integrated Modeling Approach for Vertical Gas Migration Along Leaking Wells Using a Compressible Two-Fluid Flow Model, 2023, 0169-3913, 10.1007/s11242-023-02005-4 | |
9. | Yangyang Qiao, Pål Østebø Andersen, Effective Relative Permeabilities Based on Momentum Equations with Brinkman Terms and Viscous Coupling, 2023, 1086-055X, 1, 10.2118/214388-PA | |
10. | Y. Qiao, P. Ø. Andersen, 2024, Mathematical Modelling of Co-Current and Counter-Current Spontaneous Imbibition into a 2D Matrix Block Accounting for Boundary Conditions, Gravity and Viscous Coupling, 10.2118/222505-MS |
Input parameters of reservoir and fluid properties used for for the below simulations. Note that
Parameter | Dimensional Value | Parameter | Dimensional Value |
Parameter | Dimensional Value | Parameter | Dimensional Value |
Water fractional flow function
Upper row: Results produced by the discrete scheme described in Appendix D (incompressible model). Three kinds of curves are plotted including the case without viscous effect, i.e.,
Simulation results with smaller viscous parameters after 10 days of water flooding. Three kinds of curves are compared: zero viscous effect, Darcy velocity
The results after 10 days with initial data are shown in Fig. 1 with interstitial velocity in viscous term. Four curves are compared: the one with large values of
Initial water saturation profile from Coclite et al. [9]
Left: The results from Coclite et al. [9] based on Darcy velocity in viscous term. Right: Numerical scheme (after 8 days) which uses interstitial velocity in viscous term with different viscous values
Comparison between the compressible model and the incompressible model for water-oil flow with
The water pressure evolution in the compressible model for the case with Darcy velocity in viscous term (left figure) and the case with interstitial velocity in viscous term (right figure). Water pressure increases with time in the water displacing part of the reservoir layer which leads to a compression effect where the magnitude of the viscous terms increase and thereby slows down the displacement of the water front
Left: Comparison of saturation profiles for water injection and gas injection, respectively, after the same time period (10 days) in the compressible model using interstitial velocity in viscous term (