Toto je HTML verze souboru https://arxiv.org/abs/2407.02939.
Google automaticky vytváří HTML verze dokumentů při procházení webu.
arXiv:2407.02939v1 [math.NA] 3 Jul 2024
[go: nahoru, domu]

Page 1
INF-SUP STABLE DISCRETIZATION OF THE
QUASI-STATIC BIOT’S EQUATIONS IN POROELASTICITY
CHRISTIAN KREUZER AND PIETRO ZANOTTI
Abstract. We propose a new full discretization of the Biot’s equations in
poroelasticity. The construction is driven by the inf-sup theory, which we
recently developed. It builds upon the four-field formulation of the equations
obtained by introducing the total pressure and the total fluid content. We
discretize in space with Lagrange finite elements and in time with backward
Euler. We establish inf-sup stability and quasi-optimality of the proposed
discretization, with robust constants with respect to all material parameters.
We further construct an interpolant showing how the error decays for smooth
solutions.
1. Introduction
This paper is the second one in a series initiated by [24], regarding the analysis
and the discretization of the quasi-static Biot’s equations in poroelasticity. (See
(2.1) below for the statement of the problem). The series centers around the use of
the inf-sup theory for the stability and the error analysis, with the aim of highlight-
ing the possible advantages stemming from the proposed approach, which appears
to be new in this context.
The inf-sup theory is a framework for the analysis of general linear variational
problems. The main result therein is the so-called Banach-Ne˘cas theorem (see, e.g.,
[14, Section 25.3]), that characterizes the well-posedness of such problems. Suc-
cessful applications of the Banach-Ne˘cas theorem additionally provide a two-sided
stability estimate, entailing that the space of all possible solutions is isomorphic
to that of all possible data, cf. Theorem 2.4 below. Such an estimate is of special
interest, as it is the starting point for the derivation of sharp a posteriori error
estimates, since it establishes the equivalence of error and residual, cf. [43, Sec-
tion 4.1.4]. Moreover, if the same approach applies also to the discretization at
hand, then the resulting stability estimate is the starting point for the derivation
of sharp a priori error estimates, see [2] for conforming discretizations and [4] for
nonconforming ones. The inf-sup theory is useful also in the design of robust pre-
conditioners [20, 30] and in the convergence analysis of some adaptive procedures
[16].
The inf-sup theory is well-established for the analysis and the discretization of
stationary linear equations. We refer to [14] for an overview with several examples.
The situation is substantially different for evolution equations, that are usually an-
alyzed by other techniques. While the application of the inf-sup theory to the heat
equation has been recently considered by various authors (see, e.g., [15, Chapter 71]
2010 Mathematics Subject Classification. Primary 65M15, 65M60; Secondary 74F10, 76S05.
Key words and phrases. Inf-sup stability, quasi-static Biot’s equations, poroelasticity, quasi-
optimality, robustness, a priori analysis.
1
arXiv:2407.02939v1 [math.NA] 3 Jul 2024

Page 2
2
C. KREUZER AND P. ZANOTTI
and the references therein), we are aware of only a few results for other prototypical
problems, like the wave and the Schrödinger equation [39].
The quasi-static Biot’s equations in poroelasticity are not an exception in this
respect. In fact, on the one hand, some authors have used the inf-sup theory in the
analysis and the discretization of the stationary problem resulting from the semi-
discretization in time, see e.g. [21, 23, 27]. On the other hand, in the quasi-static
case, we are not aware of any result obtained by this approach. For the a posteriori
analysis, Li and Zikatanov [25] used, in a sense, an equivalent argument. For the a
priori analysis, all papers we are aware of build upon a different argument, namely
an energy estimate that seems to go back to a seminal contribution of Zenıšek [46].
A far by exhaustive list includes [3, 8, 9, 22, 26, 33, 34, 44].
To our best knowledge, our series is the first contribution making a systematic
use of the inf-sup theory in the design and in the error analysis of a discretization
of the quasi-static Biot’s equations. Our first paper [24] is devoted to the analysis
of the equations. It establishes the well-posedness as well as a two-sided stability
estimate, which is robust with respect to all material parameters. The latter con-
tributions and the functional setting distinguish our results from previous ones by
Zenıšek [46] and Showalter [38]. In particular, we consider an equivalent four-field
formulation of the equations, that is obtained from the original one by introducing
the so-called total pressure and total fluid content. In [24] we prove also that, in
certain circumstances, additional regularity in space of the data imply correspond-
ing additional regularity in space of the solution. Establishing a similar result in
time is more challenging, due to possible singularities at the initial time; compare
e.g. with [31, 38] and the discussion in Remark (4.10) below. Further previous
contributions to the regularity theory are in [6, 45].
In this paper we propose a backward Euler discretization in time and an abstract
discretization in space of the Biot’s equations. We establish the well-posedness and
a two-sided stability estimate via the inf-sup theory, by mimicking the argument in
[24]. Then we consider a simple realization of the abstract discretization in space,
making use of Lagrange finite elements for all variables. We prove a quasi-optimal
a priori error estimate, meaning that the error is equivalent to (i.e. bounded from
above and below by) the best error. To our best knowledge, it is the first time that
such a result is established for a discretization of the Biot’s equations.
The error notion we consider is motivated by the inf-sup theory and it is relatively
involved, as all variables are coupled in a nontrivial way. Therefore, we further
elaborate on our error estimate, by showing that the best error can be bounded
by a sum of decoupled best errors in standard norms, that are much easier to
investigate. All constants in our results are robust with respect to all material
parameters and do not require any additional regularity beyond the one guaranteed
by the well-posedness of the equations. Finally, we establish first-order convergence,
with respect to the space and time meshsize, for sufficiently smooth solutions.
Organization. In section 2 we state the equations and recall the main results
from [24]. Section 3 establishes the stability of an abstract discretization. Section 4
is devoted to the a priori error analysis for an exemplary concrete discretization,
which is also tested numerically in section 5.
Notation. Throughout the paper, we denote by ∥·∥X the norm of a Hilbert space
X. The dual space X
acts on X through the pairing ⟨·, ·⟩X
. We denote by L2(X),

Page 3
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
3
H1(X) and C0(X), respectively, the Bochner spaces of all L2, H1 and C0 functions
mapping the time interval [0,T] into X. For a measurable set Ω ⊆ Rd, we adopt
the simplified notation (·, ·)and ∥·∥for the scalar product and the norm in
L2(Ω). We write a ≲ b and a ≂ b when there are constants 0 < c ≤ c, possibly
different at different occurrences, such that a ≤ cb and cb ≤ a ≤ cb, respectively.
The hidden constants are independent of the material parameters involved in the
equations. The dependence on other relevant quantities is addressed case by case,
see e.g. Remark 4.1.
2. Inf-sup theory for the Biot’s equations
This section introduces the Biot’s equations and summarizes some results from
[24], that are useful for the construction and the analysis of the discretization in
the next sections.
2.1. Initial-boundary value problem. Let Ω ⊆ Rd, 1 ≤ d ≤ 3, be a bounded
domain with polyhedral and Lipschitz continuous boundary. The flow of a Newto-
nian fluid inside a linear elastic porous medium located in Ω, in the time interval
(0,T) with T > 0, is modeled by the quasi-static Biot’s equations as follows
(2.1a)
−div(2µ∇Su + (λdivu − αp)I) = fu
in Ω × (0,T)
t(αdivu + σp) − div(κ∇p) = fp
in Ω × (0,T).
The first equation states the momentum balance for the elastic porous medium,
whereas the second one is the mass balance for the fluid.
The unknown functions in the equations are the displacement u : Ω → Rd of the
elastic porous medium and the pressure p : Ω → R of the fluid. The differential
operator ∇S is the symmetric part of the gradient and I is d×d identity tensor. The
material parameters, denoted by Greek letters, are the Lamé constants µ,λ > 0,
the Biot-Willis constant α > 0, the constrained specific storage coefficient σ ≥ 0
and the hydraulic conductivity κ > 0. Consistently with [24], we assume that all
parameters are constant in Ω × [0,T] for simplicity. We refer to [24, Remark 2.1]
for a discussion on possible extensions.
We are interested in the initial-boundary value problem obtained by prescribing
also the initial condition
(2.1b)
(αdivu + σp)|t=0 = ℓ0
in Ω
as well as the boundary conditions
(2.1c)
u = 0
on Γu,E × (0,T)
(2µ∇Su + (λdivu − αp)I)n = gu
on Γu,N × (0,T)
p = 0
on Γp,E × (0,T)
κ∇p · n = gp
on Γp,N × (0,T)
where Γu,E ∪ Γu,N = ∂Ω=Γp,E ∪ Γp,N and Γu,E ∩ Γu,N = ∅ = Γp,E ∩ Γp,N . The
letter n denotes the outward unit normal vector on ∂Ω. The subscripts ‘E’ and
‘N’ are intended to assist the reader in distinguishing the portions of the boundary
with essential and natural conditions.
We point out that different statements of the initial and of the boundary con-
ditions are sometimes met in the literature. We refer to [24, Remark 2.2-2.3] for a
more extensive discussion on this point.

Page 4
4
C. KREUZER AND P. ZANOTTI
2.2. Weak Formulation and well-posedness. For convenience, we introduce a
compact notation for the differential operators in the Biot’s equations (2.1), namely
(2.2)
E := −div(2µ∇S)
D := div
L := −div(κ∇).
Notice that E and L act on u and p, respectively, in (2.1a). Comparing also with
the boundary conditions, it looks reasonable that the regularity of u in space in a
weak formulation can be described via
(2.3)
U :=
(
H1(Ω)d/RM if Γu,N = ∂Ω
H1
Γu,E
(Ω)d
otherwise
with RM denoting the rigid body motions. Analogously, for the regularity of p in
space, we consider
(2.4)
P :=
⎢
⎢k
H1(Ω) ∩ L2
0(Ω)
if Γp,N = ∂Ω
H1
Γp,E
(Ω) ∩ L2
0(Ω) if Γp,N ̸= ∂Ω, Γu,E = ∂Ω,σ = 0
H1
Γp,E
(Ω)
otherwise
where L2
0(Ω) = {q ∈ L2(Ω) | /
q = 0}. We refer to [24, Remark 2.5] for a
motivation of the nonstandard definition of P in the second case.
Remark 2.1 (Notation). We are aware of the fact, that the abstract notation in-
troduced here (and the subsequent one for all related spaces and operators) makes
the reading possibly harder. Still, it has the advantage that all combinations of the
boundary conditions (as well as the critical case σ = 0) can be treated at the same
time. In our view, this is quite important, because our approach to the analysis
and the discretization of the Biot’s equations is mainly the same in all cases, but
each case may require subtle minor modifications.
The action of the divergence on u in (2.1a) indicates that also the space
(2.5)
D := D(U) =
(
L2
0(Ω) if Γu,E = ∂Ω
L2(Ω) otherwise
plays a relevant role in the Biot’s equations. Similarly, since p is involved in (2.1a)
also without the action of any differential operator in space, we repeatedly make
use of
(2.6)
P =
(
L2
0(Ω) if Γp,N = ∂Ω or Γu,E = ∂Ω,σ = 0
L2(Ω) otherwise,
where the closure is taken with respect to the L2(Ω)-norm. An important point for
our analysis is that both D and P are subspaces of L2(Ω), but their mutual relation
depends on the boundary conditions and on the parameter σ. Therefore, it is useful
introducing
PD : L2(Ω) → D
and
P
P
: L2(Ω) → P,
the L2(Ω)-orthogonal projections onto D and P, respectively.
Remark 2.2 (Functional setting). The diagram in Figure 2.1 summarizes the rela-
tion among the spaces and the operators introduced up to this point. In addition,
we denote by i : P → P the inclusion operator and Dand iare the adjoint oper-
ators of D and i, respectively. The spaces D and P are identified with their duals

Page 5
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
5
via the L2(Ω)-scalar product. Thus, the square on the right side of the diagram
involves the Hilbert triplet
(2.7)
P ↩→ P ≡ P
↩→ P.
U
oo
E
OO
D
U
D
L2(Ω)
PD
~
PP
P
L
//
i
P
OO
i
D
D
P
P
Figure 2.1. Spaces and operators describing the regularity in
space for the weak formulation (2.8) of the Biot’s equations. The
triple lines on the bottom indicate identification via the L2(Ω)-
scalar product.
With this preparation, we are in position to recall the weak formulation of the
initial-boundary value problem (2.1) introduced in [24, section 2.3], namely
(2.8)
Eu + Dptot = ℓu
in L2(U)
λDu − ptot − αPDp = 0
in L2(D)
αP
P
Du + σp − m = 0
in L2(P)
tm + Lp = ℓp
in L2(P)
m(0) = ℓ0
in P.
The loads ℓu ∈ L2(U) and ℓp ∈ L2(P) result from the data fu and fp in the
equations (2.1a) as well as gu and gp in the boundary conditions (2.1c).
Remark 2.3 (Auxiliary variables). Compared to (2.1a), the weak formulation (2.8)
involves two additional unknown variables, namely the total pressure ptot and the
total fluid content m defined by the second and third equation, respectively. Intro-
ducing these variables is not strictly necessary for our analysis, but it substantially
simplifies the definition of the space Y1 in (2.9) below. The use of the total pressure
was observed in [27] to help also the design of robust linear solvers.
The main result in [24] states that (2.8) is uniquely solvable within the closure
Y1 of the space
(2.9)
Y1 := L2(U) × L2(D) × L2(P) × (L2(P) ∩ H1(P))
with respect to the norm
∥(Cu, Cptot, Cp, Cm)∥2
1 :=
T
0
(
∥Cu∥2
U
+
1
µ
∥Cptot2
+ ∥∂t
Cm + LCp∥
2
P
)
+ ∥ Cm(0)∥2
P
+
T
0
( 1
µ + λ
∥λDCu − Cptot − αPD Cp∥
2
+ γ∥αP
PDCu + σCp− Cm∥2
)
.
(2.10)

Page 6
6
C. KREUZER AND P. ZANOTTI
Here U and P are equipped with the energy norm
(2.11)
∥·∥2
U
= ⟨E·, ·⟩
U
and
∥·∥2
P
= ⟨L·, ·⟩
P
and the parameter γ is given by
γ :=
⎢⎢⎢⎢⎢⎢
⎢⎢⎢⎢⎢⎢k
min
{µ + λ
α2
,
1
σ
}
if σ > 0 and P ⊆ D
µ + λ
α2
+
1
σ
if σ > 0 and P⊈D
µ + λ
α2
if σ = 0.
(2.12)
Taking the closure is indeed necessary, because Y1 is not closed with respect to
∥·∥1, cf. [24, Proposition 4.4].
Theorem 2.4 (Well-posedness of the weak formulation). For all possible data
(ℓu,ℓp,ℓ0) ∈ L2(U)×L2(P)×P, the weak formulation (2.8) has a unique solution
y1 = (u, ptot, p, m) ∈ Y1, which satisfies the two-sided stability bound
∥y12
1
T
0
(∥ℓu2
U+ ∥ℓp2
P) + ∥ℓ02
P.
Moreover, we have m ∈ C0(P) as well as the norm equivalence
∥y12
1 ≂ ∥y12
1 + ∥m∥2
L(P) +
T
0
(λ∥Du∥2
+ γ−1∥p∥2
) .
All hidden constants depend only on Ω and T.
Proof. Combine [24, Theorem 3.5] with [24, Proposition 4.1].
Although we omit the proof of Theorem 2.4, it is worth roughly summarizing
how it is obtained. Indeed, we shall make use of a similar argument in order to
verify the well-posedness of the discretization introduced in the next section, cf.
Theorem 3.9 below.
The weak formulation (2.8) can be viewed as a special instance of the following
linear variational problem: given ℓ ∈ Y
2, find y1 ∈ Y1 such that
(2.13)
b(y1,y2) = ⟨ℓ, y2
Y2
∀y2 ∈ Y2.
The test space is obtained by collecting all possible test functions for (2.8), namely
Y2 := L2(U) × L2(D) × L2(P) × L2(P) × P
and it is equipped with the norm
∥(v, qtot, q, n, n0)∥2
2 :=
T
0
(
∥v∥2
U
+ ∥n∥2
P
)
+ ∥n02
P
+
T
0
((µ + λ)∥qtot2
+ γ−1∥q∥2
) .
(2.14)
The bilinear form b : Y1 × Y2 → R is defined according to the left-hand side of
(2.8) by
b(Cy1,y2) :=
T
0
(
⟨ECu + D
Cptot,v⟩
U
+ ⟨∂t
Cm + LCp, n⟩P
)
+ ⟨ Cm(0),n0
P
+
T
0
(
(λDCu − Cptot − αPD Cp, qtot)+ (αP
PDCu + σCp− Cm, q)
)
(2.15)

Page 7
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
7
for Cy1 = (Cu, Cptot, Cp, Cm) ∈ Y1 and (v, qtot, q, n, n0) ∈ Y2.
The so-called Banach-Ne˘cas theorem characterizes the well-posedness of the lin-
ear variational problem in terms of boundedness, inf-sup stability and nondegen-
eracy of the form b, see e.g. [14, theorem 25.9]. These properties are verified in
[24, section 3]. Their combination implies the well-posedness of (2.8) as well as the
two-sided stability bound in Theorem 2.4.
Remark 2.5 (Trial functions). As in (2.10) and (2.15), we use hereafter the super-
script ‘∼’ to distinguish a general trial function in Y1 from the solution of the weak
formulation (2.8).
Remark 2.6 (Functions and functionals). Owing to Remark 2.2, we often identify
the functions in D and P with their Riesz representatives in Dand P
and vice
versa. For instance, the term Dptot from the first equation in (2.8) and even the
space L2(P) ∩ H1(P) proposed for the total fluid content must be interpreted in
this way. As usual, we omit to explicitly indicate the Riesz isometries, accepting
some ambiguity, so as to alleviate the notation. We apply the same principles to
the discretization in the next sections.
3. Abstract inf-sup stable discretization
In this section we design a discretization of the weak formulation (2.8) of the
initial-boundary value problem for the Biot’s equations (2.1). The space discretiza-
tion is challenging, due to the nontrivial coupling of the variables and the various
differential operators acting on them. Therefore, we initially work with a general
discretization, so as to allow for the maximal flexibility. We discuss a concrete real-
ization in section 4 below. The time discretization seems less problematic, because
(2.8) involves only one time derivative. Hence, we directly make a concrete choice,
namely the backward Euler scheme.
The space discretization in this section builds upon a number of assumptions
that must be verified case by case. The set of our assumptions is identified by
special tags with the dedicated enumeration (H1), (H2), etc. With a small abuse,
we actually include among the assumptions also the definition of some relevant
constants. In those cases, the size (and not just the existence) of the constants is
the property to be verified in each concrete example.
The main result in this section is the well-posedness established in Theorem 3.9.
We do not attempt at analyzing the error at this level of generality. Indeed, we
believe that the estimates we would obtain either require too many assumptions
or are too much abstract to be of interest. Thus, we prefer to discuss the error
analysis for the concrete example in section 4.
Remark 3.1 (Notation for the discretization). In general, we mark all spaces and
operators related to the discretization in space by the subscript ‘s’ and those related
to the discretization in time by the subscript ‘t’. The combination ‘st’ of the
two subscripts identifies the full discretization in space and time. To alleviate the
notation, we use capitol letters (in place of subscripts) to distinguish the functions
involved in the discretization from those related to the original Biot’s equations.
3.1. Abstract discretisation in space. The general concept of our discretiza-
tion in space consists in replacing all spaces and operators in Figure 2.1 by finite-
dimensional counterparts, while preserving the structure of the diagram therein, cf.

Page 8
8
C. KREUZER AND P. ZANOTTI
Figure 3.1. In order to discretize the displacement and the pressure, we consider
two finite-dimensional linear spaces
(3.1)
Us
and
Ps
i.e. discrete counterparts of the spaces U in (2.3) and P in (2.4). We replace the
operators E and L in (2.2) by positive definite and self-adjoint linear operators
(3.2)
Es : Us → U
s
and
Ls : Ps → P
s .
In analogy with (2.11), we equip the two spaces with the energy norms
(3.3)
∥·∥2
Us := ⟨Es·, ·⟩Us
and
∥·∥2
Ps := ⟨Ls·, ·⟩Ps .
Then, the dual norms on U
s and P
s are given by
(3.4)
∥·∥2
U
s
:= sup
V ∈Us
⟨·,V ⟩
Us
∥V ∥Us
= ⟨·, E−1
s
·⟩Us
∥·∥2
P
s
:= sup
N∈Ps
⟨·,N⟩Ps
∥N∥Ps
= ⟨·, L−1
s ·⟩Ps .
Notice that Us and Ps are not required to be conforming, i.e. subspaces of U and
P, respectively. Therefore, in order to define an error notion on the sums U + Us
and P + Ps, we assume the following.
(H1)
The norms ∥·∥U and ∥·∥P in (2.11) can be extended to U + Us and
P + Ps with ∥·∥U ≂ ∥·∥Us in Us and ∥·∥P ≂ ∥·∥Ps in Ps.
In order to discretize the space D in (2.5), we consider a linear operator
(3.5)
Ds : Us → L2(Ω)
i.e. a discrete counterpart of the divergence D in (2.2). Then, we set
(3.6)
Ds := Ds(Us).
The proof of Theorem 2.4 given in [24] exploits the norm equivalence µ∥D· ∥2
U
∥·∥2
in D, that is nothing else than boundedness and surjectivity of D, cf. [14,
Lemma C40]. (Note that here D is identified with its dual via the L2(Ω)-scalar
product, cf. Remark 2.2.) In order to reproduce this property at the discrete level,
we assume the following.
(H2)
There are constants c = c(Us) and C = C(Us) with 0 < c ≤ C
and such that c∥·∥2
≤ µ∥D
s · ∥2
U
s
≤ C∥·∥2
in Ds.
Actually, this prescribes that Us/Ds is a stable pair for the discretization of the
Stokes equations. Note that also Ds is identified here with its dual space.
The proof of Theorem 2.4 given in [24] exploits also the density of P in P, that
gives rise to the Hilbert triplet in (2.7). Since Ps is finite-dimensional, we are led
to discretize P by Ps itself, giving rise to the triplet
(3.7)
Ps = Ps ≡ P
s = P
s .
Also in this case, the identification of Ps with its dual space is made via the L2(Ω)-
scalar product. Hence, the pairing ⟨·, ·⟩
Ps
coincides with (·, ·)upon identifying the
functionals in P
s with their Riesz representative.

Page 9
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
9
Remark 3.2 (Discretization of the Hilbert triplet). As Ps coincides with its closure,
the above Hilbert triplet is trivial from the algebraic viewpoint. Still, the spaces
in it play different roles and are equipped with different norms, depending on their
position, in analogy with (2.7). To alleviate the notation, we omit hereafter the
symbol of the closure.
We denote by PDs
and PPs
the L2(Ω)-orthogonal projections onto Ds and Ps,
respectively. In particular, the adjoint P
Ps
of the second projection maps functionals
on Ps into functionals on P. This observation is important for the definition of the
error notion in section 3.4 below, because the trial norm ∥·∥1 in (2.10) involves the
dual norm ∥·∥P. Thus we assume the following, in order to keep the norm of P
Ps
under control.
(H3)
There are constants c = c(Ps) and C = C(Ps) with 0 < c ≤ C and
such that c∥·∥P
s
≤ ∥P
Ps
· ∥P≤ C∥·∥P
s
in P
s .
The upper bound here is equivalent to the P-stability of the projection PPs . The
lower bound can be formulated as an inf-sup condition and it is equivalent to the
existence of a bounded right inverse of PPs , cf. [23, Proposition 3.5].
Finally, the proof of the inf-sup stability in Lemma 3.8 below for vanishing σ
makes use of the following assumption.
(H4)
The inclusion Ps ⊆ Ds holds true if σ = 0.
Notice that the spaces P and D in (2.6) and (2.5), respectively, satisfy the same
inclusion.
Remark 3.3 (Spurious pressure oscillations). The combination of the assumptions
(H2) and (H4) prescribes that Us/Ps is a stable pair for the discretization of the
Stokes equations if σ = 0. This property was observed both numerically [18] and
theoretically [29] to be important to prevent from spurious pressure oscillations
in certain regimes. Indeed, forgetting the time derivative for a moment (or, more
precisely, discretizing it in time) the Biot’s equations (2.1a) are close to the Stokes
equations for vanishing σ and small κ.
Figure 3.1 summarizes the relation among the spaces and the operators intro-
duced in this section. As announced, the structure is exactly as in Figure 2.1. We
refer to Remark 2.2 for the details of the notation.
U
s
oo
Es
OO
D
s
Us
D
L2(Ω)
PDs
}}
PPs
!!
Ps
Ls
//
i
P
sOO
i
D
s
Ds
Ps
P
s
Figure 3.1. Spaces and operators describing the regularity in
space for the discretization of the Biot’s equations. The triple lines
on the bottom indicate identification via the L2(Ω)-scalar product.

Page 10
10
C. KREUZER AND P. ZANOTTI
3.2. Discretization in time. As announced, we consider a simple first-order dis-
cretization in time, namely backward Euler. To this end, we introduce a partition
(3.8)
0 = t0 < t1 < ··· < tJ = T
of the time interval [0,T] with J ≥ 1. We denote the local time intervals and their
length by
(3.9)
Ij := [tj−1,tj]
and
|Ij| := tj − tj−1
respectively, for j = 1,...,J.
For X ∈ {Us, Ds, Ps}, we consider the space of piecewise time-constant functions
on the above partition with values in X
S
0
t (X) := {X ∈ L(X) | X|Ij =: Xj ∈ X, j = 1,...,J}.
Whenever useful, we identify a function X ∈ S0
t (X) with the sequence of its values
(Xj)J
j=1 ⊆ X.
The space S0
t (X) is suitable for a first-order discretization in time of L2(X), i.e.
of the first three components in the trial (2.9) and of the first four components in
the test (2.2). The last component in the trial space (the fluid content) is different,
as it involves H1-regularity in time. We discretize it in time by
S
0
t (Ps) × Ps
where the second component plays the role of the initial value. Then we introduce
a discrete counterpart dt : S0
t (Ps) × Ps → S0
t (P
s ) of the time derivative ∂t, in the
vein of the backward Euler scheme, i.e.
(3.10)
dt(b
M, b
M0)Ij :=
b
Mj − b
Mj−1
|Ij|
for (b
M, b
M0) ∈ S0
t (Ps) × Ps and for all j = 1,...,J.
The proof of Theorem 2.4 given in [24] and, in particular, the control of the
point values of the total fluid content hinge on the following integration by parts
rule 2 /
T
0
(∂t
Cm, L
−1
Cm
P= ∥ Cm(T)∥2
P−∥ Cm(0)∥2
P, which holds true for Cm ∈ H1(P),
cf. [15, Lemma 64.40]. The operator dt defined above satisfies a similar relation.
Lemma 3.4 (Time discrete integration by parts rule). We have
2
tj
0
⟨dt(b
M, b
M0), L−1
s
b
M⟩Ps ≥ ∥b
Mj2
P
s − ∥b
M02
P
s
for all (b
M, b
M0) ∈ S0
t (Ps) × Ps and j = 1,...,J.
Proof. We exploit (3.10), rearrange terms and recall the second part of (3.4). It
results
2
tj
0
⟨dt(b
M, b
M0), CL−1
s M⟩Ps = 2
j
k=1
⟨b
Mk − b
Mk−1, L−1
s
b
MkPs
= ∥b
Mj2
P
s
+
j
k=1
∥b
Mk − b
Mk−12
P
s − ∥b
M02
P
s
cf. [15, eq. (67.9)]. This readily implies the claimed inequality.

Page 11
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
11
3.3. Full discretization. Combining the discretizations in space and time from
the previous sections, we are in position to propose an abstract full discretization
of the initial-boundary value problem (2.1) for the Biot’s equations.
We consider the trial space
(3.11)
Y1,st := S0
t (Us) × S0
t (Ds) × S0
t (Ps) × S0
t (Ps) × Ps.
Inspired by (2.10) and taking the assumption (H1) in section 3.1 into account, we
equip Y1,st with the norm
∥(CU, CPtot, CP, b
M, b
M0)∥2
1,st :=
T
0
(
∥CU∥2
U
+
1
µ
∥ CPtot2
+ ∥dt(b
M, b
M0) + Ls CP∥2
P
s
)
+ ∥b
M02
P
s
+
T
0
( 1
µ + λ
∥λDs CU − CPtot − αPDs CP∥2
+ γ∥αPPs Ds CU + σ CP − b
M∥2
)
.
(3.12)
Remark 3.5 (Equivalent trial norm). According to the assumption (H3) in sec-
tion 3.1, we could equivalently replace the discrete dual norm ∥·∥P
s
in the definition
of ∥·∥1,st by ∥P
Ps
· ∥P. All the results stated in section 3.4 hold true also in this
case, with the only difference that the hidden constants additionally depend on the
constants in (H3). This observation is important in view of the definition of the
error notion in section 4.2.
For the test space, we proceed similarly and set
(3.13)
Y2,st := S0
t (Us) × S0
t (Ds) × S0
t (Ps) × S0
t (Ps) × Ps.
Recalling again the assumption (H1) in section 3.1, we equip Y2,st with the norm
∥·∥2 in (2.14).
Let (Lu,Lp,L0) ∈ S0
t (U
s ) × S0
t (P
s ) × P
s be a discretization of the correspond-
ing data (ℓu,ℓp,ℓ0) ∈ L2(U) × L2(P) × Pin the weak formulation (2.8) of
the Biot’s equations. We consider the following full discretization of (2.8): find
(U, Ptot,P,M,M0) ∈ Y1,st such that
(3.14)
EsU + D
s Ptot = Lu
in S0
t (U
s )
λDsU − Ptot − αPDs P = 0
in S0
t (Ds)
αPPs DsU + σP − M = 0
in S0
t (Ps)
dt(M,M0) + LsP = Lp
in S0
t (P
s )
M0 = L0
in P
s .
Note that, also in this case, there are some ambiguities between functions and
functionals, that can be clarified in the vein of Remark 2.6.
Remark 3.6 (Two- and four-field formulation). The weak formulation 2.8 involves
four unknown variables but it can be equivalently rewritten as a two-field weak
formulation of the initial-boundary value problem (2.1), by eliminating the total
pressure ptot and the total fluid content m, cf. [24, Remark 2.7]. Analogously,
we could eliminate the discrete total pressure Ptot and the discrete fluid content
M from (3.14). In this way, we would equivalently obtain a discretization of the
two-field weak formulation, with trial and test spaces given by S0
t (Us)×S0
t (Ps)×Ps.

Page 12
12
C. KREUZER AND P. ZANOTTI
Since we aim at establishing the well-posedness of (3.14) via the inf-sup theory,
it is convenient viewing it as an instance of the following linear variational problem:
given L ∈ Y
2,st, find Y1 ∈ Y1,st such that
(3.15)
bst(Y1,Y2) = ⟨L, Y2
Y2,st
∀Y2 ∈ Y2,st.
Of course, this can be seen as a discretization of (2.13). Here, the bilinear form
bst : Y1,st × Y2,st → R is defined by
bst(CY1,Y2) :=
T
0
(
⟨Es CU + D
s CPtot,V ⟩Us + ⟨dt(b
M, b
M0) + Ls CP,N⟩Ps
)
+ ⟨b
M0,N0Ps
+
T
0
(
(λDs CU − CPtot − αPDs CP,Qtot)+ (αPPs Ds CU + σ CP − b
M,Q)
)
(3.16)
for CY1 = (CU, CPtot, CP, b
M, b
M0) ∈ Y1,st and Y2 = (V,Qtot,Q,N,N0) ∈ Y2,st.
A remarkable difference between (2.13) and (3.15) is that, in the latter one, we
do not have to explicitly take the closure of the trial space. Indeed, Y1,st is certainly
closed, being finite-dimensional.
3.4. Well-posedness. The goal of this section is to establish the well-posedness
of the discretization (3.14) by means of the inf-sup theory. Since the trial space
Y1,st and the test space Y2,st are finite-dimensional with equal dimension, we can
use a simplified version of the Banach-Ne˘cas theorem, which does not require to
verify the nondegeneracy of the form bst, see e.g. [14, Theorem 26.6]. Therefore,
the well-posedness is equivalent to the properties verified in the next two lemmas.
Lemma 3.7 (Boundedness). The bilinear form bst in (3.16) satisfies
sup
Y2∈Y2,st
bst(CY1,Y2)
∥Y22
≲ ∥CY11,st
for all CY1 ∈ Y1,st. The hidden constant depends only on the constants in the as-
sumptions (H1) and (H2) in section 3.1.
Proof. The claimed bound follows from the Cauchy-Schwarz inequality applied to
each term in the definition of bst, in combination with (3.3) and with the norm
equivalences in the assumptions (H1) and (H2).
Lemma 3.8 (Inf-sup stability). The bilinear form bst in (3.16) satisfies
sup
Y2∈Y2,st
bst(CY1,Y2)
∥Y22
(1 + T)1
2
(
∥CY12
1,st + ∥b
M∥2
L(P
s ) +
T
0
(
λ∥Ds CU∥2
+ γ−1∥ CP∥2
)
\1
2
.
for all CY1 = (CU, CPtot, CP, b
M, b
M0) ∈ Y1,st. The hidden constant depends only on the
constants in the assumptions (H1) and (H2) in section 3.1.
Proof. See section 3.5.
The combination of the two above lemmas implies the main result of this section.

Page 13
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
13
Theorem 3.9 (Well-posedness of the full discretization). For all possible loads
(Lu,Lp,L0) ∈ S0
t (U
s ) × S0
t (P
s ) × P
s , the equations (3.14) have a unique solution
Y1 = (U, Ptot,P,M,M0) ∈ Y1,st, which satisfies the two-sided stability bound
∥Y12
1,st
T
0
(
∥Lu2
U
s
+ ∥Lp2
P
s
)
+ ∥L02
P
s
.
(3.17)
Moreover, we have the norm equivalence
(3.18)
∥Y12
1,st ≂ ∥Y12
1,st + ∥M∥2
L(P
s ) +
T
0
(λ∥DsU∥2
+ γ−1∥P∥2
) .
The hidden constants depend only on the final time T and on the constants in the
assumptions (H1) and (H2) in section 3.1.
Proof. The equations (3.14) are equivalent to the linear variational problem (3.15)
with load L = (Lu, 0, 0,Lp,L0) ∈ Y
2,st. Then, the simplified version of the Banach-
Ne˘cas theorem for finite-dimensional linear variational problems [14, Theorem 26.6]
implies existence and uniqueness of the solution, in combination with Lemmas 3.7
and 3.8. The identity bst(Y1, ·) = L in Y
2,st further implies
∥Y12
1,st + ∥M∥2
L(P
s ) +
T
0
(
λ∥DsU∥2
+ γ−1∥P∥2
)
≲ ∥(Lu, 0, 0,Lp,L0)∥2
2,∗ ≲ ∥Y12
1,st
according to estimates in Lemmas 3.7 and 3.8. The hidden constants depend only on
the ones therein. Here ∥·∥2,∗ denotes the dual of the test norm. This readily implies
the second claimed equivalence. The first one follows by recalling the definition of
the norm ∥·∥2 in (2.14).
Remark 3.10 (Jumps of the discrete fluid content). Recall that the component M
of the solution Y1 in Theorem 3.9 represents the discretization of the total fluid
content m in Theorem 2.4. While the latter one is guaranteed to be continuous
in time, the former one is not, (indeed it is piecewise constant). Still, a careful
inspection at the proofs of Lemmas 3.4 and 3.8 reveals that the left-hand side in
the stability bound (3.17) controls the jumps of M in time, namely
J
j=1
∥Mj − Mj−12
P
s
≲ ∥Y12
1,st.
This can be viewed as a discrete counterpart of the continuity of m.
3.5. Inf-sup stability. This section is devoted to the proof of Lemma 3.8. The
argument is similar to the one in [24, section 3.1], which establishes the inf-sup
stability of the form b in (2.15). Nevertheless, some subtleties exist with regard
to the discretization and we therefore include a detailed proof so as to keep the
discussion as much complete and self-contained as possible.
Let CY1 = (CU, CPtot, CP, b
M, b
M0) ∈ Y1,st be given and set
CHDs := λDs CU − CPtot − αPDs CP
and
CHPs := αPPs Ds CU + σ CP − b
M

Page 14
14
C. KREUZER AND P. ZANOTTI
for shortness. Since the projections PDs and PPs map onto Ds and Ps, respectively,
we have CHDs ∈ S0
t (Ds) as well as CHPs ∈ S0
t (Ps). We consider the test function
Y2,j :=
(
(CU + E−1
s
D
s CPtotj,
4 max{1,C}
µ + λ
CHDs χj,
min{1,c}
CHPs χj,
L−1
s (2b
M + dt(b
M, b
M0) + Ls CP)χj, L−1
s
b
M0
)
with j = 1,...,J, where χj : [0,T] → R denotes the indicator function of the inter-
val [0,tj]. Here the constants c and C are as in the assumption (H2) in section 3.1.
We refer to [24, Remark 3.9] for a motivation of the test function.
First of all, notice that we indeed have that Y2,j ∈ Y2,st, i.e. it is an admissible
test function. Indeed, recall the definition of the spaces Y1,st and Y2,st in (3.11)
and (3.13). The operator E−1
s
D
s maps Ds into Us and Ls is an isometry between Ps
and P
s . Moreover, the indicator function χj is piecewise constant on the partition
(3.8) of [0,T]. Hence, the multiplication by χj preserves the inclusion in Y2,st.
The proof consists of two main steps. First, we establish the lower bound
bst(CY1,Y2,j + Y2,J ) ≳∥CY12
1,st + ∥b
M∥2
L(P
s ) +
T
0
(
λ∥Ds CU∥2
+ γ−1∥ CP∥2
)
(3.19)
for a suitable index 1 ≤ j ≤ J. Then, we estimate the norm of the test function
(3.20)
∥Y2,j2
2 ≲ ∥CY12
1,st + (1 + T)∥b
M∥2
L(P
s )
for some hidden constant independent of j. The combination of these inequalities
readily implies the inf-sup stability claimed in Lemma 3.8.
We start with the proof of (3.19). Using the test function Y2,j in the definition
(3.16) of the form bst yields
(3.21)
bst(CY1,Y2,j) =
tj
0
⟨Es CU + D
s CPtot, CU + E−1
s
D
s CPtotUs
(=: I1)
+ 2
tj
0
⟨dt(b
M, b
M0) + Ls CP, L−1
s
b
M⟩Ps
(=: I2)
+ ∥b
M02
P
s
+
tj
0
∥dt(b
M, b
M0) + Ls CP∥2
P
s
+
tj
0
4 max{1,C}
µ + λ
∥ CHDs 2
+
tj
0
min{1,c}
∥ CHPs 2
.
Notice that the dual norms on the third line are obtained thanks to the second part
of (3.4). We are led to analyze the terms I1 and I2 on the right-hand side.
Regarding the first term, we have
I1 =
tj
0
(
⟨Es CU, CU⟩Us + 2⟨D
s CPtot, CU⟩Us + ⟨D
s CP, E−1
s
D
s CPtotUs
)
.
We rewrite the first summand with the help of the first identity in (3.3). Analo-
gously, we estimate the third summand from below by means of the first identity
in (3.4) and the assumption (H2) in section 3.1
⟨Es CU, CU⟩Us = ∥CU∥2
Us
and ⟨D
s CPtot, E−1
s
D
s CPtotUs
c
µ
∥ CPtot2
.

Page 15
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
15
We estimate the remaining term by observing that the assumption (H2) implies
also µ∥Ds · ∥2
≤ C∥·∥2
Usin Us. This bound and a Young’s inequality imply
⟨D
s CPtot, CU⟩Us = −(Ds CU, CHDs )+ λ∥Ds CU∥2
− α(Ds CU, PDs CP)
4
∥Ds CU∥2
1
4
∥CU∥2
Us
max{1,C}
µ + λ
∥ CHDs 2
− α(Ds CU, CP).
Notice that we were able to omit the projection PDs in the last term thanks to the
inclusion Ds CU ∈ S0
t (Ds). Combining this bound with the identities above reveals
J1
tj
0
(1
2
∥CU∥2
Us +
c
µ
∥ CPtot2
+
2
∥Ds CU∥2
2 max{1,C}
µ + λ
∥ CHDs 2
− 2α(Ds CU, CP)
)
.
Regarding the other critical term in (3.21), we have
I2 = 2
tj
0
(
⟨dt(b
M, b
M0), L−1
s
b
M⟩Ps + (b
M, CP)
)
.
The use of the L2(Ω)-scalar product in the second summand is justified by the
discussion after (3.7). We estimate the first summand from below by invoking
Lemma 3.4
2
tj
0
⟨dt(b
M, b
M0), L−1
s
b
M⟩Ps ≥ ∥b
Mj2
P
s − ∥b
M02
P
s
.
To investigate the second summand, assume first σ > 0. A Young’s inequality
reveals
(b
M, CP)= ( CHPs , CP)+ α(PPs Ds CU, CP)+ σ∥ CP∥2
≥ α(Ds CU, CP)+
σ
2
∥ CP∥2
1
∥ CHPs 2
.
As before, we omit the projection PPs thanks to the inclusion CP ∈ S0
t (Ps). Alterna-
tively, for general σ ≥ 0, it holds that
(b
M, CP)= ( CHPs , CP)+ α(PPs Ds CU, CP)+ σ∥ CP∥2
= −
1
α
( CHPs , CHDs )+
λ
α
( CHPs , Ds CU)
1
α
( CHPs , CPtot)
+ ( CHPs , CP − PDs CP)+ α(Ds CU, CP)+ σ∥ CP∥2
.
Bounding the term ( CHPs , CP − PDs CP)conveniently is subtle. Whenever Ps ⊆ Ds,
we have PDs CP = CP, hence ( CHPs , CP − PDs CP)= 0. When the above inclusion fails,
we have σ > 0 due to the assumption (H4) in section 3.1. Then, we apply Young’s
inequality to obtain ( CHPs , CP − PDs CP)≤ ∥ CHPs 2
/(2σ) + σ∥ CP∥2
/2. Combining this
observation with other applications of Young’s inequality, the previous lower bound
and the definition (2.12) of γ, we arrive at
(b
M, CP)≥ −
max{1,C}
2(µ + λ)
∥ CHDs 2
c
∥ CPtot2
λ
2
∥Ds CU∥2
+
σ
2
∥ CP∥2
2 min{1,c}
∥ CHPs 2
+ α(Ds CU, CP).

Page 16
16
C. KREUZER AND P. ZANOTTI
By combining the last two bounds with the above identity for I2, we obtain
I2 ≥ ∥b
Mj2
P
s − ∥b
M02
P
s
+
tj
0
(
max{1,C}
µ + λ
∥ CHDs 2
c
∥ CPtot2
λ
2
∥Ds CU∥2
+ σ∥ CP∥2
min{1,c}
∥ CHPs 2
+ 2α(Ds CU, CP)
)
.
After this preparation, we are in position to choose a specific test function in
order to establish (3.19). We let j ∈ {1,...,J} be such that ∥b
MjP
s
= ∥b
M∥L(P
s ).
Then, we insert the previous lower bounds of I1 and I2 into (3.21), resulting in
bst(CY1,Y2,j + Y2,J ) ≥
T
0
(1
2
∥CU∥2
Us +
c
∥ CPtot2
+ ∥dt(b
M, b
M0) + Ls CP∥2
P
s
)
+
T
0
( 1
µ + λ
∥ CHDs 2
+ γ∥ CHPs 2
+
λ
2
∥Ds CU∥2
+ σ∥ CP∥2
)
+ ∥b
M∥2
L(P
s ).
In combination with the definition (3.12) of the norm ∥·∥1,st and the assumption
(H1) in section 3.1, this almost establishes (3.19). Indeed, we have
bst(CY1,Y2,j + Y2,J ) ≳∥CY12
1,st + ∥b
M∥2
L(P
s ) +
T
0
(
λ∥Ds CU∥2
+ σ∥ CP∥2
)
.
The proof that we can indeed replace σ by γ−1 in the last summand follows verbatim
the argument in the proof of [24, Proposition 4.1].
The last step of the proof consists in establishing (3.20). According to the
definition (2.14) of the test norm ∥·∥2, we have for all j = 1,...,J, that
∥Y2,j2
2 =
tj
0
(
∥CU + E−1
s
D
s CPtot2
U
+ ∥L−1
s (2b
M + dt(b
M, b
M0) + Ls CP)∥2
P
)
+
tj
0
(16 max{1,C2}
µ + λ
∥ CHDs 2
+
16γ
min{1,c2}
∥ CHPs 2
)
+ ∥L−1
s
b
M02
P
.
We exploit the identities (3.3) and (3.4) and the norm equivalences in the as-
sumptions (H1) and (H2) from section 3.1. We extend also the integrals from the
interval [0,tj] to [0,T]. Hence, we obtain
∥Y2,j2
2
T
0
(
∥CU∥2
U
+
1
µ
∥ CPtot2
+ ∥b
M∥2
P
s
+ ∥dt(b
M, b
M0) + Ls CP∥2
P
s
)
+
T
0
( 1
µ + λ
∥ CHDs 2
+ γ∥ CHPs 2
)
+ ∥b
M02
P
s
.
Finally, we recall the definition (3.12) of the norm ∥·∥1,st and exploit the upper
bound /
T
0
∥·∥2
P
s
≤ T∥·∥2
L(P
s ). This establishes (3.20) for some hidden constant
independent of j and concludes the proof.
Remark 3.11 (Time-dependent space discretization). In our framework the space
discretization does not change in time and this is important for the proof of the
inf-sup stability in this section. Indeed, a change in the space discretisation would
imply the change of the operator Ls in time and this would have an effect on our
use of the integration by parts formula from Lemma 3.4. For the same reason,

Page 17
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
17
we argued in [24, Remark 2.1] that the parameter κ in the Biot’s equations (2.1a)
could be allowed to vary in space but not in time.
4. Discretisation with Lagrange elements in space
In this section we propose an exemplary concrete space discretization, based on
H1-conforming Lagrange finite elements for all variables. Hence, we first verify the
assumptions (H1)-(H4) from section 3.1, in order to infer the well-posedness. Then,
we discuss the a priori error analysis.
Our notation and assumptions for the finite elements are as follows. We denote
by T a face-to-face simplicial mesh of Ω. The shape constant of T is given by
(4.1)
max
T∈T
hT
rT
where hT is the diameter of a d-simplex T ∈ T and rT is the diameter of the largest
ball inscribed in T.
We denote by F the set of all faces of T and by Fi the interior faces. We assume
that the mesh is compatible with the boundary conditions (2.1c). This means that
each face F ∈ F \ Fi (i.e. each boundary face) satisfies
either F ⊆ Γu,N or F ⊆ Γu,E
and
either F ⊆ Γp,N or F ⊆ Γp,E.
Hence, we have two partitions of the boundary faces Fu,E ∪ Fu,N and Fp,E ∪ Fp,N ,
where each set Fis defined as
(4.2)
F:= {F ∈ F | F ⊆ Γ}.
We let the meshsize h and the normal n be the piecewise constant functions on
T and on the skeleton of T (i.e. the union of all faces) defined as
h|T := hT
and
n|F := nF
for all T ∈ T and F ∈ F, respectively. Here nF is a fixed unit normal vector of F,
pointing outside Ω if F is a boundary face.
The space Pk(S), k ≥ 0, consist of all polynomials of total degree ≤ k on an
n-simplex S ⊆ Rd with 1 ≤ n ≤ d. The corresponding space of (possibly discontin-
uous) piecewise polynomials on the mesh T is denoted by Pk(T).
4.1. Concrete space discretisation. A concrete realization of the abstract space
discretization from section 3.1 requires a specific choice of the spaces Us and Ps and
of the operators Es and Ls in (3.1) and (3.2), respectively. We have to prescribe
also the operator Ds in (3.5) and to characterize its range Ds in (3.6). Finally, we
need to verify the assumptions (H1)-(H4).
For the discretization of the displacement, we consider H1-conforming Lagrange
finite elements of degree k + 1 with k ≥ 1, i.e., we set
(4.3)
Us := Pk+1(T)d ∩ U.
The intersection with the space U from (2.3) enforces the global continuity (hence
the H1-conformity) as well as the boundary conditions. As usual in conforming
finite elements, we discretize the operator E in (2.2) by Es : Us → U
s defined as
(4.4)
⟨Es CU,V ⟩Us := ⟨E CU,V ⟩U
for all CU,V ∈ Us.

Page 18
18
C. KREUZER AND P. ZANOTTI
Recall that the operator Ds should be a discretization of the divergence and that
the pair Us/Ds with Ds = Ds(Us) should enjoy a discrete Stokes inf-sup condition
in order to satisfy assumption (H2). Given the above definition of Us, one option
for Ds is suggested by the so-called Hood-Taylor pair, cf. [14, section 54.4]. Hence,
we consider H1-conforming Lagrange finite elements of degree k
Ds := Pk(T) ∩ H1(Ω) ∩ D
(4.5)
and we define Ds : Us → Ds by
(4.6)
(Ds CU,Qtot)= (D CU,Qtot)
for all CU ∈ Us and Qtot ∈ Ds. According to (2.5), the intersection with D in (4.5)
simply enforces the vanishing mean value when Γu,E = ∂Ω. Note also that Ds CU is
nothing else than the H-orthogonal projection of D CU onto Ds.
Finally, the assumption (H4) suggests that also the space Ps, for the discretiza-
tion of the pressure and of the total fluid content, should consist of H1-conforming
Lagrange finite elements of degree k. Therefore, we set
(4.7)
Ps := Pk(T) ∩ P.
The intersection with the space P from (2.4) enforces global continuity, the bound-
ary conditions and, possibly, the vanishing mean value. In this case, we define
Ls : Ps → P
s in terms of L in (2.2) by
(4.8)
⟨Ls CP,Q⟩Ps := ⟨L CP,Q⟩P
for all CP,Q ∈ Ps.
Remark 4.1 (Hidden constants). In order to simplify the statement of the next
results, it is implicitly understood hereafter that all hidden constants in our esti-
mates potentially depend on the final time T, the shape constant (4.1) of T and the
polynomial degree k. In particular, the latter is arbitrary but fixed in our setting.
The possible dependence on other relevant quantities is addressed case by case.
Having introduced all the spaces and the operators required for the space dis-
cretization in section 3.1, we can verify the validity of the assumptions (H1)-(H4).
Proposition 4.2 (Verification of the assumptions). Let the spaces Us, Ds and Ps
and the operators Es, Ds and Ls be defined by (4.3)-(4.8).
(1) The assumption (H1) holds true and the equivalences therein are actually
identities.
(2) The assumption (H2) holds true and the constants therein depend only on
the quantities mentioned in Remark 4.1, provided that each d-simplex in T
has at least one vertex in the interior of Ω.
(3) The assumption (H3) holds true and the constants therein depend only on
the quantities mentioned in Remark 4.1, provided that the grading of T,
defined as in [12], is strictly less than (
2k + d +
k)/(
2k + d −
k).
(4) The assumption (H4) holds true.
Proof. (1) Owing to the inclusions Us ⊆ U and Ps ⊆ P, there is no need to extend
the norms ∥·∥U and ∥·∥P. Moreover, the combination of (3.3) with (4.4) and (4.8)
readily implies ∥·∥Us = ∥·∥U in Us as well as ∥·∥Ps = ∥·∥P in Ps.

Page 19
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
19
(2) The upper bound in (H2) follows from the boundedness of Ds, which satisfies
∥Ds ·∥2
≤∥D·∥2
≲ µ−1∥·∥2
Uin Us, according to (2.2), (2.11) and (4.6). The lower
bound can be rephrased as
c∥·∥2
≤ µ
(
sup
V ∈Us
(DV, ·)
∥V ∥U
)2
in Ds. The definitions (2.2) and (2.11) reveal that this is equivalent to the discrete
Stokes inf-sup stability of the pair Us/Ds, i.e. of the Hood-Taylor pair. The latter
condition is known to hold true under the above assumption on the mesh; see [14,
sections 54.3 and 54.4].
(3) We have
∥P
Ps · ∥P= sup
w∈P
⟨·, PPs w⟩
Ps
∥w∥P
= sup
w∈P
(·, PPs w)
∥w∥P
in P
s . The inclusion Ps ⊆ P implies the lower bound in (H3) with c = 1, because
PPs
is a projection onto Ps. The upper bound is equivalent to the P-stability of
PPs ; cf. [41]. Owing to (2.11) and (4.7), this is the H1(Ω)-stability of the L2(Ω)-
orthogonal projection onto H1-conforming Lagrange finite element spaces. Such a
condition is known to hold under the asserted grading assumption thanks to [11,
Theorem 4.14(ii)] together with [11, Theorem 4.4] and [11, Remark 4.4].
(4) Let σ = 0. The combination of (4.5) and (4.7) with (2.4) and (2.5) implies,
for Γu,E = ∂Ω that
Ps = Pk(T) ∩ H1
Γp,E (Ω) ∩ L2
0(Ω) and Ds = Pk(T) ∩ H1(Ω) ∩ L2
0(Ω).
For Γu,E ̸= ∂Ω we have instead,
Ps ⊆ Pk(T) ∩ H1
Γp,E (Ω) and Ds = Pk(T) ∩ H1(Ω).
In both cases we have Ps ⊆ Ds, i.e., assumption (H4) is verified.
Remark 4.3 (Assumptions for (H3)). We refer to [12, Definitions 1.1] for the exact
definition of mesh grading. Clearly the grading of quasi-uniform meshes equals
1 and thus satisfies the grading condition in Proposition 4.2(3). It follows from
[12, Theorem 1.3], that the grading condition is also satisfied for all polynomial
degrees k ≥ 1 and dimensions d ≤ 6 provided T is obtained by adaptive bisection
of a colored initial mesh; see [12, Assumption 3.1] for the notion of colored mesh.
Note that assumption (H3) can be verified under different assumptions. Indeed,
the H1(Ω)-stability of the L2(Ω)-orthogonal projection onto finite element spaces
has been extensively analyzed by various authors; we refer [11] for an overview of
the existing results.
Having verified all assumptions in section 3.1, we deduce the well-posedness of
the discretization (3.14) with the spaces and the operators proposed in this section.
In particular, the inclusions Us ⊆ U and Ps ⊆ P suggest to define the data in the
discretization by restriction of the data in the weak formulation (2.8)
(4.9)
Lu = ℓ
u|S0
t (Us),
Lp = ℓ
p|S0
t (Ps)
and
L0 = ℓ0|Ps .
Theorem 4.4 (Well-posedness with Lagrange elements). Let the spaces Us, Ds and
Ps and the operators Es, Ds and Ls be defined by (4.3)-(4.8). Assume that T is ob-
tained from a colored initial mesh by newest vertex bisection and that each d-simplex

Page 20
20
C. KREUZER AND P. ZANOTTI
in T has at least one vertex in the interior of Ω. Then, the discretization (3.14)
with the data (4.9) has a unique solution Y1 = (U, Ptot,P,M,M0) ∈ Y1,st with
∥Y12
1,st ≂ ∥Y12
1,st + ∥M∥2
L(P
s ) +
T
0
(λ∥DsU∥2
+ γ−1∥P∥2
)
T
0
(
∥Lu2
U
s
+ ∥Lp2
P
s
)
+ ∥L02
P
s
T
0
(∥ℓu2
U+ ∥ℓp2
P) + ∥ℓ02
P.
(4.10)
All hidden constants depend only on the quantities mentioned in Remark 4.1.
Proof. The combination of Theorem 3.9 with Proposition 4.2 implies the existence
and the uniqueness of the solution and guarantees that the two equivalences in
(4.10) hold true. Then, the upper bound follows from the discretization (4.9) of the
data and the inclusions Us ⊆ U and Ps ⊆ P.
4.2. Quasi-optimality. In order to quantify the accuracy of the proposed dis-
cretization, we need an error notion that is related to both the trial norm ∥·∥1 in
(2.10) and to the discrete trial norm ∥·∥1,st in (3.12), cf. (4.12) below. A major
issue is that the former one involves the dual norm in P, whereas the latter one
involves the dual norm in Ps.
In order to compare functionals defined on the two spaces, we can use the adjoint
P: P
s → Pof a bounded projection P : P → Ps. The specific choice of P is critical
because of the term ∂t
Cm + LCp in the norm ∥·∥1. Since this term acts in space
via the pairing ⟨·, ·⟩
P
, one may want to employ the L2(Ω)-orthogonal projection,
because the pivot space in the Hilbert triplet (2.7) is equipped with the L2(Ω)-norm.
Still, according to (2.11), the action of L induces the P-norm, thus suggesting to
make use of the P-orthogonal projection. Of course, similar comments apply to the
counterpart of ∂t
Cm + LCp in ∥·∥1,st.
Both options have advantages and disadvantages. The latter one, being related
to the P-norm, suggests the use of piecewise polynomials of degree k + 1 for the
discretization of P, but this would be critical for the assumption (H4), as Ds consists
of piecewise polynomials of degree k, cf. (4.5)-(4.7). The former option does not
have this issue, therefore we go for it. Still, the derivation of higher-order decay
rates in space appears critical in this case, cf. Remark 4.12.
In accordance with the above discussion and with the inclusions Us ⊆ U and
Ps ⊆ P, we consider the error notion Err : Y1 × Y1,st → [0, +∞) defined as
ERR(Cy1, CY1)2 :=
T
0
(
∥u − CU∥2
U
+
1
µ
∥ptot − CPtot2
)
+
T
0
∥∂t
Cm + LCp− P
Ps (dt(b
M, b
M0) + Ls CP)∥2
P
+ ∥ Cm(0) − P
Ps b
M02
P
+
T
0
1
λ + µ
∥λDCu − Cptot − αPD Cp− (λDs CU − CPtot − αPDs CP)∥2
+
T
0
γ∥αP
PDCu + σCp− Cm − (αPPs Ds CU + σ CP − b
M)∥2
(4.11)

Page 21
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
21
for Cy1 = (Cu, Cptot, Cp, Cm) ∈ Y1 and CY1 = (CU, CPtot, CP, b
M, b
M0) ∈ Y1,st. Notice that this
is not a norm on the sum Y1 + Y1,st. For instance, in general, we have D CU ̸= Ds CU
for CU ∈ S0
t (Us) ⊆ L2(U). Still, the above error notion measures the accuracy in the
approximation of all functions and functionals involved in the trial norm ∥·∥1 and
it holds that
(4.12)
ERR(Cy1, 0) = ∥Cy11
and
ERR(0, CY1) ≂ ∥CY11,st.
Indeed, the second equivalence follows from (1) and (3) in Proposition 4.2.
The equations (3.14) with the spaces and the operators defined by (4.3)-(4.8) are
not a conforming Petrov-Galerkin discretization of the weak formulation (2.8). In-
deed, the trial space Y1,st in the former problem is not a subspace of its counterpart
Y1 in the latter one. Nevertheless, we have for the test spaces the inclusion
Y2,st ⊆ Y2.
This property and the definition (4.9) of the load in (3.14) by restriction of the one in
(2.8) ensure that we can still guarantee the fundamental quasi-optimality property
of inf-sup stable conforming Petrov-Galerkin methods by a standard argument; cf.
[2].
Theorem 4.5 (Quasi-optimality). Let all assumptions in Theorem 4.4 be verified.
Denote by y1 ∈ Y1 and Y1 ∈ Y1,st, respectively, the solutions of (2.8) and (3.14)
with (4.9). Then, we have
ERR(y1,Y1) ≲ inf
˜Y1∈Y1,st
ERR(y1, CY1).
(4.13)
The hidden constant depends only on the quantities mentioned in Remark 4.1.
Proof. For CY1 ∈ Y1,st, the triangle inequality and (4.12) imply
ERR(y1,Y1) ≤ ERR(y1, CY1) + ERR(0,Y1 − CY1) ≲ ERR(y1, CY1) + ∥Y1 − CY11,st.
According to the inf-sup stability in Lemma 3.8, we have
∥Y1 − CY11,st ≲ sup
Y2∈Y2,h
bst(Y1 − CY1,Y2)
∥Y22
= sup
Y2∈Yh
2
b(y1,Y2) − bst(CY1,Y2)
∥Y22
,
where the identity follows by comparing problems (2.8) and (3.14). We exploit the
definitions (2.15), (3.16), and (4.4)-(4.8), as well as the invariance of the projection
PPs onto Ps. For y1 = (u, ptot, p, m) and CY1 = (CU, CPtot, CP, b
M, b
M0), this results in
b(y1,Y2) − bst(CY1,Y2) :=
T
0
((
E(u − CU),V 〉
U
+ (ptot − CPtot, DV )
)
+
T
0
(∂tm + L CP − P
Ps (dt(b
M, b
M0) + Ls CP),N
P
+ (m − P
Ps b
M0,N0
P
+
T
0
(λDu − ptot − αPDp − (λDs CU − CPtot − αPDs CP),Qtot
)
+
T
0
(αP
P
Du + σp − m − (αPPs Ds CU + σ CP − b
M),Q)
.
(4.14)

Page 22
22
C. KREUZER AND P. ZANOTTI
Then Cauchy-Schwarz inequalities, the bound ∥DV ∥2
≲ µ−1∥V ∥2
U
and the defini-
tion (2.14) of the test norm ∥·∥2 yield
∥Y1 − CY11,st ≲ ERR(y1, CY1).
Inserting this estimate into the first inequality above concludes the proof.
Remark 4.6 (Augmented error notion). Our definition of the error notion is in a
sense minimal, because we have included only the terms that arise by applying the
Cauchy-Schwarz inequality in (4.14). Still, the first equivalence in (4.10) clarifies
that that the statement and the proof of Theorem 4.5 will remain unchanged if we
augment ERR(·, ·) by adding any of the following terms
∥ Cm − P
Ps b
M∥2
L(P),
T
0
λ∥DCu − Ds CU∥2
,
and
T
0
γ−1∥Cp− CP∥2
.
4.3. A priori error estimates. The quasi-optimality stated above has two re-
markable properties and (at least) one clear disadvantage. One the one hand, the
estimate (4.13) holds true for any solution of the weak formulation (2.8) (i.e. no ad-
ditional regularity is required) and the hidden constant is robust with respect to all
material parameters. On the other hand, it is not immediate how (and even if) the
error decays to zero, because the definition (4.11) of ERR(·, ·) involves a nontrivial
coupling of the various components of the solution.
In this section, we investigate the latter aspect. Roughly speaking, we aim at
showing that the best error in the right-hand side of (4.13) is equivalent to a sum
of best errors. In order to preserve the two above-mentioned properties, we make
sure also that the constants in the equivalence are independent of the material
parameters and that no additional regularity of the solution of (2.8) is required
beyond the minimal one, namely y1 ∈ Y1. When such a result is available, the
decay of the error to zero can be easily discussed in terms of classical results from
approximation theory.
The precise statement of our main result is as follows. Notice that the last term
on the right-hand side of (4.15) does not appear in our definition of the error notion,
but we could equivalently include it according to Remark 4.6.
Theorem 4.7 (Best error decoupling). Let all assumptions in Theorem 4.4 be
verified and denote by y1 = (u, ptot, p, m) ∈ Y1 the solution of (2.8). Then we have
that
inf
˜Y1∈Y1,st
ERR(y1, CY1)2
inf
̂U∈S0
t (Us)
T
0
∥u − CU∥2
U
+
inf
̂Ptot∈S0
t (D)
T
0
1
µ
∥ptot − CPtot2
+
inf
̂
W ∈S0
t (Ps)
T
0
∥∂tm + Lp − P
Ps ̂
W∥2
P+ inf
̂
M0∈Ps
∥m(0) − P
Ps ̂
M02
P
+ ε−1
s
inf
̂P∈S0
t (Ps)
T
0
1
γ
∥p − CP∥2
.
(4.15)
The hidden constant depends on the quantities mentioned in Remark 4.1 and on
the constant in (4.20) and εs is defined in (4.24) below.
Proof. See sections 4.4-4.6.

Page 23
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
23
A first consequence of Theorem 4.7 is plain convergence: the error converges to
zero as the mesh-size converges to zero, both in space and time. This holds true
irrespective of the regularity of the solution. Moreover, first-order convergence can
be established under additional regularity assumptions.
Corollary 4.8 (First-order convergence). Let all assumptions in Theorem 4.4 be
verified. Denote by y1 = (u, ptot, p, m) ∈ Y1 and Y1 ∈ Y1,st, respectively, the
solutions of (2.8) and (3.14) with (4.9). Assume additionally
(4.16)
u ∈ H1(U) ∩ L2(H2(Ω)d)
ptot ∈ H1(D) ∩ L2(H1(Ω))
p ∈ H1(P) ∩ L2(H1(Ω))
(∂tm + Lp) ∈ H1(P) ∩ L2(L2(Ω))
m(0) ∈ L2(Ω).
Then, the error can be bounded from above as follows
ERR(y1,Y1)2
(
max
h
)2
(
1
κ
∥m(0)∥2
+
T
0
(
µ∥∇2u∥2
+
1
µ
∥∇ptot2
+
1
εsγ
∥∇p∥2
+
1
κ
∥∂tm + Lp∥2
)}
+
(
max
j=1,...,J
|Ij|
)2 T
0
(
∥∂tu∥2
U
+
1
µ
∥∂tptot2
+
1
εsγ
∥∂tp∥2
+ ∥∂t(∂tm + Lp)∥2
P
)
.
The hidden constant depends on the quantities mentioned in Remark 4.1 and on
the constant in (4.20) and εs is defined in (4.24) below.
Proof. Combine Theorem 4.5 with Theorem 4.7. Then, use standard Bramble-
Hilbert-like estimates (see, e.g., [40, Chapter 7]) in order to bound each term on
the right-hand side of (4.15).
Remark 4.9 (Regularity in space). According to [24, Theorem 5.4], the solution y1
of (2.8) satisfies the space regularity in (4.16) at least in some circumstances. To
be more precise, we have
u ∈ L2(H2(Ω)2)
ptot ∈ L2(H1(Ω))
p ∈ L2(H1(Ω))
(∂tm + Lp) ∈ L2(L2(Ω))
when the data are such that
u ∈ L2(L2(Ω)2),
p ∈ L2(L2(Ω)),
0 ∈ L2(Ω),
upon additionally assuming that Ω ⊆ R2 is convex, the boundary conditions (2.1c)
are posed on Γu,E = ∂Ω=Γp,N , and λ ≳ µ.
Remark 4.10 (Regularity in time). According to [31], the time regularity in (4.16)
may fail to hold even for smooth data. For a rough explanation, assume the data are
smooth in time, the solution enjoys the time regularity in (4.16) and, in addition,
we have m ∈ C1(P), i.e. we have one more time derivative than in Theorem 2.4.
Then, on the one-hand, the fourth equation in (2.8) implies p ∈ C0(P). On the
other hand, the evaluation of the other equations at t = 0 implies that the pair
(u(0),p(0)) solves the Stokes-like problem
(E + λDD)u(0) + αDPDp(0) = fu(0) in U
αP
P
divu(0) + σp(0) = ℓ0
in P.

Page 24
24
C. KREUZER AND P. ZANOTTI
In general, the solution of this problem is in U×P. Hence the condition p(0) ∈ P can
hold true only for compatible data, otherwise some singularity must be expected
at the initial time. We refer to the Terzaghi test case discussed in section 5 below
for a numerical illustration.
Remark 4.11 (Grading of the time partition). The constant in Theorem 4.7 depends
on the one in (4.20), which measures the grading of the partition employed for the
discretization in time. The latter constants enters into play because Theorem 4.7
does not assume additional regularity in time of the solution. This calls for a Scott-
Zhang-like interpolation in time, see section 4.4 below for the details. When all
components of the solution are continuous in time, a Lagrange-like interpolation
is possible and the constant in (4.20) does not enter into the error estimation, cf.
[40, Theorem 4.5]. Still, according to Remark 4.10, the continuity at t = 0 is not
obvious.
Remark 4.12 (Decay rate). The space discretization considered in this section is
actually of higher-order, because we make use of H1-conforming Lagrange finite
elements of degree k + 1 for the displacement and of degree k for the other com-
ponents of the solution with k ≥ 1, cf. (4.3)-(4.7). Therefore, the question arises
if a higher-order decay rate with respect to h can be obtained under higher regu-
larity assumptions on the solution. On the one hand, the P-norm errors on the
right-hand side of (4.15) appear to be critical in this respect, because the space
Ps, possibly incorporating boundary conditions, is used to approximate functional
from P, which have no prescribed boundary conditions. On the other hand, it is
unclear to us if the above-mentioned regularity of the solution can be expected in
general. Indeed, the regularity result mentioned in Remark 4.9, assumes Γp,E = ∅.
Due to the subtlety of this matter, we postpone any further investigation on this
point to future work.
The remaining part of this section is devoted to the proof of Theorem 4.7. For
this purpose, we aim at constructing a bounded interpolant I : Y1 → Y1,st. The
operator I cannot be obtained by just approximating each component of the solu-
tion irrespective of the others, because of the nontrivial coupling of the components
in the error notion (4.11). Thus, to make sure that I is bounded and robust with
respect to the material parameters, we must guarantee that it is compatible with
the coupling. Roughly speaking this means that, when the error notion involves
some combination of the components of the solution, it must be possible to ac-
curately approximate it by the corresponding combination of the components of
the interpolant. We achieve this goal with the help of a number of commutative
diagrams, cf. Figures 4.1 and 4.2.
We divide our construction into three parts. In Section 4.4 we first discuss the
interpolation in time. In Section 4.5 we discuss the interpolation in space. Finally,
in Section 4.6, we combine the interpolation in time and space in order to prove
Theorem 4.7.
4.4. Time interpolation. The error notion ERR(·, ·) involves several terms with
L2 regularity in time, so we initially consider the problem of approximating function
from L2(X) into S0
t (X) for some Hilber space X. Since we cannot control the point
values of a function in L2(X), we cannot use Lagrange interpolation; cf. remark 4.11.
Thus, we resort to Scott-Zhang-like [37] interpolation in the vein of [40, Section 4].

Page 25
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
25
Recall (3.8)-(3.9). For j ∈ {1,...,J}, the polynomial ψj ∈ P1(Ij) defined as
(4.17)
ψj(t) := 6
t − tj−1
|Ij|2
2
|Ij|
is such that,
(4.18)
Ij
j = Q(tj)
∀Q ∈ P1(Ij).
We define J : L2(X) → S0
t (X) by
(J x)|Ij :=
Ij
j
for x ∈ L2(X) and j = 1,...J.
Since the error notion in (4.11) involves the discrete time derivative dt, it is
useful to introduce another interpolant CJ : L2(X) → S0
t (X), defined as
( CJx)|Ij :=
Ij
(∫ t
tj−1
x(s)ds
\
ψj(t)dt +
Ij−1
(∫ tj−1
t
x(s)ds
)
ψj−1(t)dt
|Ij|
for j = 2,...J and
( CJx)|I1 :=
I1
(∫ t
0
x(s)ds
)
ψ1(t)
|I1|
.
The first part of Lemma 4.13 below ensures that both J x and CJx are near best
approximations of x in the L2(X)-norm. The second part additionally clarifies that
the two interpolants are related, through the weak and the discrete time derivative,
as summarized by the commuting diagram in Figure 4.1.
Lemma 4.13 (Time interpolation). The operators J and CJ defined above are such
that
T
0
∥x − J x∥2
X
≤ 4 inf
X∈S0
t (X)
T
0
∥x − X∥2
X
(4.19a)
T
0
∥x − CJx∥2
X
inf
X∈S0
t (X)
T
0
∥x − X∥2
X
(4.19b)
for all x ∈ L2(X). Moreover, for x ∈ H1(X), we have
(4.19c)
dt(J x, x(0)) = CJ∂tx.
The hidden constant in (4.19b) is an increasing function of
(4.20)
max
j=2,...,J
|Ij−1|
|Ij|
.
Proof. According to [40, Remark 4.1], the operator J is invariant on S0
t (X) and it
is bounded with
T
0
∥J x∥2
X
≤ 4
T
0
∥x∥2
X
for all x ∈ L2(X). The combination of the two properties implies (4.19a).

Page 26
26
C. KREUZER AND P. ZANOTTI
Invariance of CJ on S0
t (X) follows by elementary calculations employing (4.18).
In order to prove stability for CJ, we observe for x ∈ L2(X) and j ≥ 2 that
∥( CJx)|Ij X
1
|Ij|
Ij
∥x∥X
Ij
j| +
1
|Ij|
Ij−1
∥x∥X
Ij−1
j−1|
(∫
Ij
∥x∥2
X
\1
2 (∫
Ij
j|2
\1
2
+
|Ij−1|
|Ij|
(∫
Ij−1
∥x∥2
X
\1
2 (∫
Ij−1
j−1|2
\1
2
.
By recalling (4.17)-(4.18), we arrive at
Ij
∥ CJx∥2
X
≤ 4
(
1 +
|Ij−1|
|Ij|
)∫
Ij−1∪Ij
∥x∥2
X
.
The same argument applies for j = 1. Summing over j = 1,...,J proves (4.19b).
Finally (4.19c) follows from the fundamental theorem of calculus and (4.18). □
H1(X)
( J , (·)(0) )
t
// L2(X)
˜
J
S
0
t (X) × X
dt
// S0
t (X)
Figure 4.1. Commutative diagram representing the relation be-
tween the time interpolants. The second component in the opera-
tor ( J , (·)(0) ) on the left is the evaluation at t = 0.
4.5. Space interpolation. Regarding the approximation in space, the error notion
ERR(·, ·) leads to the problem of interpolating functions from U into Us, from D into
Ds as well as from P and Pinto Ps. Moreover, the spaces are related via the
operators D and L and their discrete counterparts and the error notion involves
various L2(Ω)-orthogonal projections. Therefore, commutative diagrams like the
one in Figure 4.1 are of interest also in this context.
In order to deal with all these tasks, we invoke the existence of an operator
which maps H1-conforming piecewise polynomials into H2-conforming ones and
enjoys several other properties, whose importance is made clear along the proof of
the next results in this section. The operator is defined on the Lagrange space of
degree k without boundary conditions, namely
S1
k := Pk(T) ∩ H1(Ω).
Moreover, we denote the jumps and the averages across faces by the usual symbols
{{·}} and [·], cf. [14, Section 38.2.1].

Page 27
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
27
Lemma 4.14 (Smoothing operator). There is a linear operator S : S1
k → H2(Ω)
which satisfies the following properties
S(Ps) ⊆ P
(4.21a)
∇SQ · n = 0
on Γp,N
(4.21b)
T
(SQ)N =
T
QN
∀N ∈ Pk(T), T ∈ T
(4.21c)
F
(SQ)N =
F
QN
∀N ∈ Pk−1(F), F ∈ Fi ∪ Fp,N
(4.21d)
as well as the stability estimate
∥Dj(Q − SQ)∥2
T
F∈Fi∪Fp,N , F∩T̸=∅
F
{{h}}3−2j|[∇Q] · n|2
(4.21e)
for all Q ∈ S1
k and j ∈ {0, 1, 2}. The hidden constant depends only on the quantities
mentioned in Remark 4.1.
Proof. See Appendix A.
Having the operator S at hand, we define I: L2(Ω) → S1
k via the problem
(4.22a)
(I
Cptot,Q)= (Cptot, SQ)
∀Q ∈ S1
k
for Cptot ∈ L2(Ω). Analogously, we define IP: P→ Ps via the problem
(4.22b)
(IPCm, N)= ⟨ Cm, SN⟩P
∀N ∈ Ps
for Cm ∈ P. Note that the main difference between Iand IPis in the range. We
introduce also IU : U → Us by
(4.22c)
IUCu := CU + Rs(IDCu − Ds CU)
for Cu ∈ U, where CU is the U-orthogonal projection of u onto Us and Rs : Ds → Us is
a right inverse of Ds. The existence and the boundedness of the latter operator are
equivalent to Proposition 4.2(2), cf. [14, Lemma C.42]. Notice also that we have
IDCu ∈ Ds in view of Lemma 4.15(2) below.
Let us clarify the logic behind the definitions in (4.22). The interpolant Iis
intended for the approximation of the total pressure, IPfor the total fluid content
and the pressure, and IU for the displacement. It is important for our purpose that
IPis P-stable and this requires the use of an operator S in the right-hand side of
(4.22b), mapping the test functions (at least) into P. (We mention, in passing, that
S is actually required to map into even more regular functions, in order to enforce
the L2(Ω)-stability of the interpolant IP introduced below, cf. Lemma 4.17.) In
principle, the use of S is not needed in (4.22a), when only stability in the L2(Ω)-
norm is required. Still, it is important at some point relating Iand IP
via a
commutative diagram, cf. Figure 4.2 below. Therefore, we use S also in (4.22a).
Finally, the interpolant IU is nothing else than the U-orthogonal projection plus a
correction, that is necessary for another commutative diagram.
Lemma 4.15 (Space interpolation – Part 1). The operators I, IPand IU defined
in (4.22) enjoy the following properties.
(1) Imaps D into Ds and, for all Cptot ∈ L2(Ω), we have
∥Cptot − I
Cptot≲ inf
̂Ptot∈S1
k
∥Cptot − CPtot.

Page 28
28
C. KREUZER AND P. ZANOTTI
(2) For all Cm ∈ Pand Cp ∈ P, we have, respectively,
∥ Cm − P
Ps IPCm∥P≲ inf
̂
M∈Ps
∥ Cm − P
Ps ̂
M∥P
and
∥Cp− IPCp∥≲ inf
̂P∈Ps
∥Cp− CP∥.
(3) For all Cu ∈ U, we have DsIUCu = IDCu as well as
∥Cu − IUCu∥U ≲ inf
̂U∈Us
∥Cu − CU∥U.
(4) The following identities hold true in L2(Ω)
PDs I= IPD
and
PPs I= IPP
P
.
The hidden constants depend only on the quantities mentioned in Remark 4.1.
Proof. (1) We first check that Iindeed maps D into Ds. Recall (2.5) and (4.5).
For D = L2(Ω) there is nothing to prove. If D = L2
0(Ω), then (4.22a) implies, for
Cptot ∈ D, that
(I
Cptot, 1)= (Cptot, S1)= (Cptot, 1)= 0.
Note that the identity S1 = 1 follows from (4.21e) with j = 1. This confirms the
inclusion I
Cptot ∈ Ds. A similar argument reveals that Iis invariant on S1
k . In
fact, owing to (4.21c), we have, for Cptot ∈ S1
k , that
(I
Cptot,Q)= (Cptot, SQ)= (Cptot,Q)
∀Q ∈ S1
k .
Finally, for general Cptot ∈ H, the boundedness (4.21e) of S with j = 0, combined
with a discrete trace inequality [13, Lemma 12.8] and an inverse estimate [13,
Lemma 12.1], yields
∥I
Cptot= sup
Q∈S1
k
(Cptot, SQ)
∥Q∥
≲ ∥Cptot.
The claimed estimate follows by the invariance and the boundedness of I.
(2) For Cm ∈ Pand ̂
M ∈ Ps, we have
Cm − P
Ps IPCm = Cm − P
Ps ̂
M + P
Ps
M − IPCm).
We bound the norm of the second summand on the right-hand side with the help of
Proposition 4.2(3), the inclusion Ps ⊆ L2(Ω) and (4.21e) with j = 1 and a discrete
trace inequality
∥P
Ps
M − IPCm)∥P≂ ∥̂
M − IPCm∥P
s
= sup
N∈Ps
⟨̂
M − IPCm, N⟩Ps
∥N∥P
= sup
N∈Ps
⟨P
Ps ̂
M − Cm, SN⟩P
∥N∥P
≲ ∥P
Ps ̂
M − Cm∥P.
The first claimed estimate follows by combining this bound with the above identity.
The other estimate follows by establishing invariance on Ps as well as boundedness
in the L2(Ω)-norm exactly as in (1).
(3) Recall that the operator Rs in (4.22c) is a right inverse of Ds, i.e. DsRs
is the identity on Ds. Then, the first part of the statement directly follows from
the definition of IU. We verify the second part as in (1), i.e. we prove that IU is

Page 29
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
29
invariant on Us and bounded in the U-norm. For Cu ∈ Us, we have CU = Cu in (4.22c).
Then, according to (4.21c) and (4.6), we infer
(IDCu, Qtot)= (DCu, SQtot)= (DCu, Qtot)= (Ds
Cu, Qtot)
∀Qtot ∈ Ds.
This implies IDCu = Ds
Cu and, in turn, IUCu = Cu. Next, for general Cu ∈ U, we have
∥IUCu∥U ≲ ∥CU∥U + ∥IDCu − Ds CU∥.
Indeed, the boundedness of Rs is equivalent to 4.2(3); see [14, Lemma C.42]. We
conclude the proof of (3) with the help of the boundedness of I, D and Ds in the
respective norms and by the definition of CU as the U-orthogonal projection of Cu.
(4) Recall that PD and PDs are the L2(Ω)-orthogonal projections onto the spaces
D and Ds, defined in (2.5) and (4.5), respectively. For Cptot ∈ L2(Ω), we have
IPD Cptot ∈ Ds, in view of (1) above. Then, for all Qtot ∈ Ds, it holds that
(IPD Cptot,Qtot)= (PD Cptot, SQtot)= (Cptot, SQtot)= (I
Cptot,Qtot).
In particular, we can remove PD after the second identity, because S maps Ds into
D, thanks to (4.21c). Hence, the first claimed identity is verified. The proof of the
other one is similar. Recall that P
P
and PPs are the L2(Ω)-orthogonal projections
onto the spaces P and Ps, defined in (2.4) and (4.7), respectively. For Cp ∈ L2(Ω)
and N ∈ Ps, we have
(IPP
P Cp, N)= (P
P Cp, SN)= (Cp, SN)= (I
Cp, N).
This time we could remove P
P
after the second identity thanks to (4.21a). Thus,
also the second claimed identity is verified.
U
IU
D
// D
I
Us
Ds
// Ds
P
IP
L
// P
IP∗
Ps
Ls
// P
s
L2(Ω)
I
PD
// D
I
S1
k
PDs
// Ds
L2(Ω)
I
PP
// P
IP∗
S1
k
PPs
// Ps
Figure 4.2. Commutative diagrams representing the relations
among the space interpolants.
The interpolants defined up to this point are compatible with the divergence, i.e.
with the operators D and Ds, and with the various L2(Ω)-orthogonal projections,
as summarized in Figure 4.2. Still, the compatibility with the Laplacian, i.e. with

Page 30
30
C. KREUZER AND P. ZANOTTI
the operators L and Ls, is not guaranteed. Since L and Ls are one-to-one, the
diagram on the upper right corner of Figure 4.2 uniquely defines IP : P → Ps as
(4.23)
IP := L−1
s IPL.
The next lemma reveals that also IP has favorable approximation properties. To
this end, we introduce the following broken H2-norm
∥·∥2
H2(T) := κ
T∈T
∥D2 · ∥2
L2(T) +
F∈Fi∪Fp,N
∥{{h}}−1/2
[∇·]∥
2
L2(F)
as well as the constant
(4.24)
εs := inf
N∈Ps
∥LsN∥
∥N∥H2(T)
.
Remark 4.16 (Discrete elliptic regularity). The constant εs in (4.24) is certainly
finite, because Ps is finite-dimensional. The size of the constant is related to the
control of a H2-like norm by the L2-norm of Ls, i.e. our discretization of the
Laplacian. Therefore, the constant is a discrete measure of the elliptic regularity.
Note also that εs can be equivalently interpreted as an inf-sup constant
εs = inf
N∈Ps
sup
̂P∈Ps
⟨LN, CP⟩P
∥N∥H2(T)∥ CP∥
,
where the stability of the standard weak formulation of the Laplacian is analyzed
in a nonstandard (i.e. nonsymmetric) setting, cf. (4.8). We are aware of only a few
results regarding the size of εs. For H1-conforming Lagrange finite elements and
convex Ω, Makridakis established a lower bound in terms of the shape parameter
(4.1) of T, provided that the mesh is not too much graded, see [28]. It is unclear
to us whether the latter condition is necessary or not. When Ω is not convex,
the connection with the elliptic regularity suggests that a lower bound of εs only
in terms of shape regularity might be not possible. As for the question raised in
Remark 4.12, we postpone further investigation on this point to future work.
Lemma 4.17 (Space interpolation – Part 2). The operator IP defined above has a
unique bounded extension (denoted by the same symbol) to P which satisfies for all
Cp ∈ P the estimate
∥Cp− IP Cp∥≲ ε−1
s
inf
̂P∈Ps
∥Cp− CP∥.
The hidden constant depends only on the quantities mentioned in Remark 4.1.
Proof. As for Lemma 4.15(2)-(3), we verify the claimed estimate by showing that
IP is invariant on Ps and that it is bounded in the L2(Ω)-norm. Note that the
latter property implies also the existence of a unique bounded extension of IP to P,
because P is a dense subspace thereof. Assume first Cp ∈ Ps. Owing to (4.23) and
(4.21a), we have
⟨IP Cp, LsN⟩Ps = ⟨LCp, SN⟩P = ⟨Cp, LSN⟩P
N ∈ Ps.
We recall (2.2) and integrate by parts element-wise [1, eq. (3.6)]
⟨IP Cp, LsN⟩Ps = κ
⎝−
T∈T
T
(∆Cp)SN +
F∈Fi∪Fp,N
F([∇Cp] · n)SN
 .

Page 31
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
31
Then, we make use of (4.21c)-(4.21d) and we integrate by parts back. It results
⟨IP Cp, LsN⟩Ps = ⟨Cp, LN⟩P= ⟨Cp, LsN⟩
Ps
where the last identity is obtained with the help of (4.8). Since Ls is one-to-one,
we conclude IP Cp = Cp.
Next, for general Cp ∈ P, a similar argument as before yields
∥IP Cp∥= sup
N∈Ps
⟨IP Cp, LsN⟩Ps
∥LsN∥
= sup
N∈Ps
⟨Cp, LSN⟩P
∥LsN∥
.
Recall the inclusion SN ∈ H2(Ω) and (4.21a)-(4.21b). We integrate by parts, then
we invoke (4.21e) with j = 2 and (4.24)
∥IP Cp∥H = κ(Cp, −∆SN)≲ ∥Cp∥∥N∥H2(T) ≤ ε−1
s ∥Cp∥∥LsN∥.
The combination of this bound with the previous identity implies the announced
boundedness of IP.
4.6. Space-time interpolation. We are in position to combine the interpolants
from the two previous sections in order to conclude the proof of Theorem 4.7. To
this end, let us preliminary recall that the operators acting in time commute with
those acting in space. Our main device is the space-time interpolant I : Y1 → Y1,st
defined as
ICy1 := (IUJ Cu, IJ Cptot, IP CJC
p, IPJ Cm, IPCm(0)
)
for Cy1 = (Cu, Cptot, Cp, Cm) ∈ Y1.
Our strategy to verify Theorem 4.7 is as follows. We first establish (4.15) for
y1 ∈ Y1. Then we extend the result by density, because the right-hand side of
(4.15) is bounded with respect to the norm ∥·∥1, cf. Theorem 2.4.
First of all, we recall the definitions of Y1 and Y1,st in (2.9) and (3.11), respec-
tively, and notice that I is well-defined. In fact, the discussion in the previous
sections shows that the interpolation of each component acts as follows
L2(U)
J
−→ S0
t (U)
IU
−→ S0
t (Us)
for Cu
L2(D)
J
−→ S0
t (D)
I
−→ S0
t (Ds)
for Cptot
L2(P) ˜
J
−→ S0
t (P)
IP
−→ S0
t (Ps)
for Cp
H1(P)
J
−→ S0
t (P)
IP∗
−→ S0
t (Ps)
for Cm
P
∗ IP∗
−→ Ps
for Cm(0).
In particular, the second line makes use of Lemma 4.15(1) and the last one exploits
the inclusion Cm ∈ H1(P) ⊆ C0(P).
In order to verify Theorem 4.7, we bound the left-hand side of (4.15) by
(4.25)
inf
˜Y1∈Y1,st
ERR(Cy1, CY1)2 ≤ ERR(Cy1, ICy1)2.
Then, we recall the definition (4.11) of the error notion and we bound the six
terms therein (denoted by (i), (ii), ... for brevity) one by one. Lemma 4.13 and
Lemma 4.15(3) state that J and IU are bounded linear projections. Therefore,
their combination J IU is a L2(U)-bounded linear projection onto S0
t (Us). This
implies
(i) ≲
inf
̂U∈S0
t (Us)
T
0
∥Cu − CU∥2
U
.

Page 32
32
C. KREUZER AND P. ZANOTTI
The same reasoning with Lemma 4.15(1) implies
(ii) ≲
inf
̂Ptot∈S0
t (Ds)
T
0
1
µ
∥Cptot − CPtot2
.
Regarding the third term, we first recall (4.19c)
dt(IPJ Cm, IPCm(0)) = IPdt(J Cm, Cm(0)) = IPCJ∂t
Cm,
then, we make use of (4.23)
LsIP CJC
p = IPCJ LC
p
and, combining the two identities, we arrive at
dt(IPJ Cm, IPCm(0)) + LsIP CJC
p = IPCJ(∂t
Cm + LCp).
The same argument used in the above bounds for (i) and (ii), with Lemma 4.15(2),
implies
(iii) ≲
inf
̂
W ∈S0
t (Ps)
T
0
∥∂t
Cm + LCp− P
Ps ̂
W∥2
P.
For the fourth term, we directly use Lemma 4.15(2) to obtain
(iv) ≲ inf
̂
M0∈Ps
∥ Cm(0) − P
Ps ̂
M02
P.
Concerning the fifth term, we invoke Lemma 4.15(3)
λDsIUJ Cu − IJ Cptot − αPDs IP CJC
p =
IJ (DCu − Cptot − αPDCp) + αIJ PD Cp− αPDs IP CJC
p
The reasoning from the bound of (ii) shows that the first summand in the right-
hand side is a near best approximation of (DCu− Cptot −αPDCp) in S
0
t (Ds). The other
two summands can be rewritten as αPDs (IJ −IP CJ)C
p, thanks to Lemma 4.15(4).
Arguing as before, we see that both IJ Cp and IP CJC
p are near best approximations
of Cp in S0
t (Ps), the latter one in view of Lemma 4.17. These observations, the
triangle inequality and the definition (2.12) of γ reveal
(v) ≲
inf
̂Qtot∈S0
t (Ds)
T
0
1
λ + µ
∥(DCu−Cptot−αPD Cp)− CQtot2
−1
s
inf
̂P∈S0
t (Ps)
T
0
1
γ
∥Cp− CP∥2
.
The sixth and last term can be treated similarly. We invoke again Lemma 4.15(3)
and Lemma 4.15(4), so as to obtain
αPPs DsIUJ Cu + σIP CJC
p − IPJ Cm
= IPJ (αP
PDCu + σCp− Cm) + σ(IP CJ −IPJ )Cp.
Again, all operators yield near best approximation in the respective spaces. In
particular, for IPJ , we make use of the second part of Lemma 4.15(2). Thus we
arrive at the following estimate
(vi) ≲
inf
̂Q∈S0
t (Ps)
T
0
γ∥(P
PDCu + σCp− Cm) − CQ∥2
+ ε−1
s
inf
̂P∈S0
t (Ps)
T
0
1
γ
∥Cp− CP∥2
with the help of the definition (2.12) of γ. We insert the above bounds of (i)-(vi)
into (4.25). This establishes (4.15) for Cy1 ∈ Y1 and the right-hand side in the
resulting estimate is bounded in terms of the norm ∥·∥1, cf. Theorem 2.4. Thus,

Page 33
INF-SUP STABLE DISCRETIZATION OF THE BIOT’S EQUATIONS
33
we can extend the estimate by density from Y1 to Y1. We conclude by noticing
that the bounds of (v) and (vi) simplify to
(v)+(vi) ≲ ε−1
s
inf
̂P∈S0
t (Ps)
T
0
1
γ
∥Cp− CP∥2
if Cy1 = y1 is the solution of (2.8).
5. Numerical results
In this section we test the performance of the concrete space discretization pro-
posed in Section 4 on two well-established benchmarks for the Biot’s equations.
Owing to Corollary 4.8 and Remark 4.12, we restrict ourselves to the lowest order
case k = 1, corresponding to quadratic (resp. linear) H1-conforming Lagrange fi-
nite elements for the displacement (resp. for the other variables). All experiments
have been implemented with the help of the library ALBERTA 3.0, see [19, 35].
5.1. Terzaghi’s problem. The consolidation of a vertical soil column of total
depth H > 0 is a classical one-dimensional test case in poroelasticity. The column
is subject to compression and it is drained on top, whereas it is impermeable and
no displacement occurs at the bottom. Moreover, there is no other force acting in
the interior of the column, the initial fluid content equals zero and there are no
sources nor sinks. Therefore, in this case, the initial-boundary value problem (2.1)
for the Biot’s equations reads
(5.1)
−(2µ + λ)uzz + αpz = 0,
t(αuz + σp) − κpzz = 0, in (0,H) × (0,T)
−(2µ + λ)uz + αp = F,
p = 0, on {0} × (0,T)
u = 0,
pz = 0, on {H} × (0,T)
αuz + σp = 0, in (0,H) × {0}
with the subscripts (·)z and (·)zz denoting the partial derivatives with respect to
the depth z ∈ (0,H) and F the modulus of the force acting on top of the column.
Remarkably, the exact solution of (5.1) is explicitly known, cf. [32, Section 4.1.1].
In particular, the pressure equals
(5.2)
p(z, t) = p0
+∞
m=0
4
(2m + 1)π
sin
((2m + 1)πz
2H
)
exp
(
(2m + 1)2π2
Cγkt
4H2
)
with the auxiliary parameters
Cγ :=
( α2
2µ + λ
+ σ
)−1
and
p0 :=
αCγF
2µ + λ
.
The displacement can be derived via the relation
uz =
αp − F
2µ + λ
and the other variables are obtained through their definition, cf. Remark 2.3.
In analogy with [32, Section 4.1.1], we set
H = 1,
T = 1,
F = 103

Page 34
34
C. KREUZER AND P. ZANOTTI
as well as 1
µ = 41667, λ = 27778, α = 1, σ = 0.1, κ = 10−6.
The explicit knowledge of the exact solution makes this test case particularly
suited to observe the error decay rate. Owing to (2.12), (4.11), Remark 4.6 and
Theorem 4.7, we consider the following error notion
(5.3)
T
0
(
∥u − U∥2
U