mathematical methods in biology master of mathematics, second … · 2018-10-26 · mathematical...

92
Mathematical methods in biology Master of Mathematics, second year Benoˆ ıt Perthame Sorbonne Universit´ e, UMR 7598, Laboratoire J.-L. Lions, BC187 4, place Jussieu, F-75252 Paris cedex 5 October 26, 2018

Upload: others

Post on 29-Jul-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

Mathematical methods in biology

Master of Mathematics, second year

Benoıt PerthameSorbonne Universite,

UMR 7598, Laboratoire J.-L. Lions, BC187

4, place Jussieu, F-75252 Paris cedex 5

October 26, 2018

Page 2: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

2

Page 3: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

Contents

1 Lotka-Volterra equations 1

1.1 Movement and growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Nonnegativity principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.3 Operator monotonicity for competition or cooperative systems . . . . . . . . . . . . . . . 3

1.4 Challenges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 Reaction kinetics and entropy 7

2.1 Reaction rate equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2.2 The law of mass action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.3 Elementary examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.4 Small numbers of molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

3 Slow-fast dynamics and enzymatic reactions 13

3.1 Slow-fast dynamics (monostable) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

3.2 Slow-fast dynamics (general) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

3.3 Enzymatic reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

3.4 Belousov-Zhabotinskii reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

4 Some material for parabolic equations 23

4.1 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

4.2 The spectral decomposition of Laplace operators (Dirichlet) . . . . . . . . . . . . . . . . . 24

4.3 The spectral decomposition of Laplace operators (Neumann) . . . . . . . . . . . . . . . . 26

4.4 The Poincare and Poincare-Wirtinger inequalities . . . . . . . . . . . . . . . . . . . . . . . 26

4.5 Rectangles: explicit solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

4.6 Brownian motion and the heat equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

5 Relaxation, perturbation and entropy methods 31

5.1 Asymptotic stability by perturbation methods (Dirichlet) . . . . . . . . . . . . . . . . . . 31

5.2 Asymptotic stability by perturbation methods (Neuman) . . . . . . . . . . . . . . . . . . 33

5.3 Entropy and relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

5.4 Entropy: chemostat and SI system of epidemiology . . . . . . . . . . . . . . . . . . . . . . 36

5.5 The Lotka-Volterra prey-predator system with diffusion (Problem) . . . . . . . . . . . . . 37

3

Page 4: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

CONTENTS 1

5.6 Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

6 Blow-up and extinction of solutions 41

6.1 Semilinear equations; the method of the eigenfunction . . . . . . . . . . . . . . . . . . . . 42

6.2 Semilinear equations; the energy method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

6.3 Keller-Segel system; the method of moments . . . . . . . . . . . . . . . . . . . . . . . . . . 46

6.4 Non-extinction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

7 Linear instability, Turing instability and pattern formation 51

7.1 Turing instability in linear reaction-diffusion systems . . . . . . . . . . . . . . . . . . . . . 52

7.2 Spots on the body and stripes on the tail . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

7.3 The simplest nonlinear example: the non-local Fisher/KPP equation . . . . . . . . . . . . 56

7.4 Phase transition: what is NOT Turing instability . . . . . . . . . . . . . . . . . . . . . . . 58

7.5 Gallery of parabolic systems giving Turing patterns . . . . . . . . . . . . . . . . . . . . . . 59

7.5.1 A cell polarity system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

7.5.2 The CIMA reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

7.5.3 The diffusive Fisher/KPP system . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

7.5.4 The Brusselator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

7.5.5 The Gray-Scott system (2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

7.5.6 Schnakenberg system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

7.5.7 The FitzHugh-Nagumo system with diffusion . . . . . . . . . . . . . . . . . . . . . 68

7.5.8 The Gierer-Meinhardt system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

7.6 Models from ecology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

7.6.1 Competing species and Turing instability . . . . . . . . . . . . . . . . . . . . . . . 69

7.6.2 Prey-predator system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

7.6.3 Prey-predator system with Turing instability (problem) . . . . . . . . . . . . . . . 70

7.7 Keller-Segel with growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

8 Transport and advection equations 73

8.1 Transport equation; method of caracteristics . . . . . . . . . . . . . . . . . . . . . . . . . . 74

8.2 Advection and volumes transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

8.3 Advection (Dirac masses) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

8.4 Examples and related equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

8.4.1 Long time behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

8.4.2 Nonlinear advection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

8.4.3 The renewal equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

8.4.4 The Fokker-Planck equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

8.4.5 Other stochastic aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

Page 5: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

2 CONTENTS

Page 6: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

Chapter 1

Lotka-Volterra equations

As first examples, we present two general classes of equations that are used in biology: Lotka-Volterra

systems and chemical reactions. These are reaction-diffusion equations, or in a mathematical classifica-

tion, semilinear equations. Our goal is to explain what mathematical properties follow from the set-up

of the model: nonnegativity properties, monotonicity and entropy inequalities. These are very general

examples, for more material and more biologically oriented textbooks, the reader can also consult for

instance [33, 32, 38, 22, 9, 1].

1.1 Movement and growth

Class I are used in the area of population biology, and ecological interactions, and are characterized by

birth and death. For 1 ≤ i ≤ I, we denote by ni(t, x) the population densities of I interacting species

at the location x ∈ Rd (d = 2 for example). We assume that these species move randomly according

to Brownian motion with a bias according to a velocity Ui(t, x) ∈ Rd and that they have growth rates

Ri(t, x) (meaning birth minus death). Then, the system describing the dynamic of these population

densities is

∂tni − Di∆ni︸ ︷︷ ︸

random motion

+ div(Uini)︸ ︷︷ ︸oriented drift

= ni Ri ,︸ ︷︷ ︸growth and death

t ≥ 0, x ∈ Rd, i = 1, ..., I. (1.1)

In the simplest cases, the diffusion coefficients Di > 0 are constants depending on the species (they

represent active motion of individuals) and the bulk velocity Ui vanishes. Here we insist on the birth and

death rates Ri; they depend strongly on the interactions between species and we express this fact as a

nonlinearity

Ri(t, x) = Ri(n1(t, x), n2(t, x), ..., nI(t, x)

). (1.2)

A standard family of such nonlinearities results in quadratic interactions and is written

Ri(n1, n2, ..., nI) = ri +

I∑i=1

cijnj ,

Page 7: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

2 Relaxation and the energy method

with ri the intrinsic growth rate of the species i (it can be positive or negative) and cij the interaction

effect of species j on species i. The coefficients cij are usually neither symmetric nor nonnegative. One

can distinguish

• cij < 0, cji > 0: species i is a prey for j, species j is a predator for i. When several species eat no other

food, these are called trophic chains/interactions,

• cij > 0, cji > 0: a mutualistic interaction (both species help the other and benefit from it) to be

distinguished from symbiosis (association of two species whose survival depends on each other),

• cij < 0, cji < 0: a direct competition (both species compete for example for the same food),

• cii < 0 is usually assumed to represent intra-specific competition.

The quadratic aspect is related to the necessary binary encounters for the interaction to occur. Better

models include saturation effects: for instance the effect of too numerous predators is to decrease their

individual efficiency. This leads ecologists rather to use (with cii = −cii > 0)

Ri(n1, n2, ..., nI) = ri − ciini +

I∑j=1, j 6=i

cijnj

1 + nj,

the terminology Holing II is used in the case where j refers to a prey, i to a predator. Ecological networks

are described by such systems: diversity refers to the number of species I itself, connectance refers to the

proportion of interacting species, i.e. such that cij 6= 0. Connectance is low for trophic chain and higher

for a food web (species use several foods).

The original prey-predator system of Lotka and Volterra has two species, I = 2 (i = 1 the prey, i = 2

the predator). The prey (small fishe) can feed on abundant zooplankton and thus r1 > 0, while predators

(sharks) will die out without small fishe to eat (r2 < 0). The sharks eat small fishs in proportion to the

number of sharks (c12 < 0), while the shark population grows proportionally to the small fishe they can

consume c21 > 0). Therefore, we find the rule

r1 > 0, r2 < 0, c11 = c22 = 0, c12 < 0, c21 > 0.

1.2 Nonnegativity principle

In all generality, solutions of the Lotka-Volterra system (1.1) satisfy very few qualitative properties; there

are neither conservation laws (because they contain birth and death), nor entropy properties, a concept

which does not seem to be relevant for ecological systems. As we shall see the quadratic aspect may lead

to blow-up (solutions exist only for a finite time, see Chapter 6). Let us only mention here, that the

model is consistent with the property that population density be nonnegative.

Lemma 1.1 (Nonnegativity principle) Assume that the initial data n0i are nonnegative functions of

L2(Rd), that Ui ≡ 0 and that there is a locally bounded function Γ(t) such that |Ri(t, x)| ≤ Γ(t). Then,

the weak solutions in C(R+;L2(Rd)

)to the Lotka-Volterra system (1.1) satisfy ni(t, x) ≥ 0.

The definition and usual properties of weak solutions are given in Chapter 4, but we do not require

Page 8: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

1.3. OPERATOR MONOTONICITY FOR COMPETITION OR COOPERATIVE SYSTEMS 3

that theoretical background to show the formal manipulations leading to this result.

Proof. (Formal) Here, we use the method of Stampacchia. Set pi = −ni, then we have

∂tpi −Di∆pi = pi Ri.

Multiply by (pi)+ := max(0, pi) and integrate by parts. We find

1

2

d

dt

∫Rd

(pi(t, x)

)2+dx+Di

∫Rd

∣∣∇(pi)+

∣∣2 =

∫Rd

(pi(t, x)

)2+Ri (1.3)

and thus1

2

d

dt

∫Rd

(pi(t, x)

)2+dx ≤ Γ(t)

∫Rd

(pi(t, x)

)2+.

Therefore, we have ∫Rd

(pi(t, x)

)2+dx ≤ e2

∫ t0

Γ(s)ds

∫Rd

(p0i (x)

)2+dx.

But, the assumption n0i ≥ 0 implies that

∫Rd

(p0i (x)

)2+dx = 0, and thus, for all times we have

∫Rd

(pi(t)

)2+dx =

0, which means ni(t, x) ≥ 0.

(Rigorous) The difficulty here stems from the ‘linear’ definition of weak solutions of (1.1), which does not

allow for nonlinear manipulations such as multiplication by (pi)+ because this is not a smmoth function.

Section ?? explains how to resolve this difficulty.

Exercise. Perform the same formal proof with the drift terms Ui. Which assumption is needed on Ui?

Exercise. In the formal context of the Lemma 1.1, show that

d

dt

∫Rd

ni(t, x)dx =

∫Rd

Ri(t, x)ni(t, x)dx,

∫Rd

ni(t, x)2dx ≤∫Rd

(n0i (x)2dx e2

∫ t0

Γ(s)ds .

1.3 Operator monotonicity for competition or cooperative sys-

tems

The positivity property is general but a rather weak property. Comparison principle and monotonicity

principle are much stronger properties but require either competition or cooperative systems.

For simplicity we consider only two cooperative species, I = 2 in equation (1.1)–(1.2), and we make

the cooperative assumption

∂n2R1(n1, n2) ≥ 0,

∂n1R2(n1, n2) ≥ 0. (1.4)

Page 9: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

4 Relaxation and the energy method

Lemma 1.2 (Monotonicity principle for cooperative systems) Consider a cooperative system, that

is, (1.4) and two solutions (n1, n2), (m1,m2) such that for some constants 0 ≤ ni ≤ Γ1, 0 ≤ mi ≤ Γ1

and |Ri| ≤ Γ2, | ∂∂niRj | ≤ Γ3. Then, the Lotka-Volterra system (1.1)–(1.2) is ordered for all times

n0i ≤ m0

i , i = 1, 2 =⇒ ni(t) ≤ mi(t), ∀t ≥ 0, i = 1, 2.

Proof. Subtracting the two equations for n1 and m1, we find successively, by adding and subtracting

convenient terms,

∂t(n1 −m1)−D1∆(n1 −m1) = (n1 −m1)R1(n1, n2) +m1

(R1(n1, n2)−R1(m1,m2)

),

∂t

(n1 −m1)2+

2− D1(n1 −m1)+∆(n1 −m1) = (n1 −m1)2

+R1(n1, n2)

+m1(n1 −m1)+

(R1(n1, n2)−R1(m1, n2)

)+m1(n1 −m1)+

(R1(m1, n2)−R1(m1,m2)

).

The first and second lines are controlled by bounds on R

d

dt

∫Rd

(n1 −m1)2+

2+D1

∫Rd

|∇(n1 −m1)+|2 ≤ (Γ2 + Γ3)

∫Rd

(n1 −m1)2+

+

∫Rd

m1(n1 −m1)+

(R1(m1, n2)−R1(m1,m2)

),

We now use 0 ≤ m1 ≤ γ1, ∂∂n2R1(n1, n2) ≥ 0 and a Lipschitz constant of R1(n1, n2), to conclude

d

dt

∫Rd

(n1 −m1)2+

2≤ (Γ2 + Γ3)

∫Rd

(n1 −m1)2+ + Γ1Γ3

∫Rd

(n1 −m1)+(n2 −m2)+.

The same argument on n2 −m2 leads to the similar inequality

d

dt

∫Rd

(n2 −m2)2+

2≤ (Γ2 + Γ3)

∫Rd

(n2 −m2)2+ + Γ1Γ3

∫Rd

(n1 −m1)+(n2 −m2)+,

and adding these two inequalities, for u(t) =∫Rd [(n1−m1)2

+ + (n2−m2)2+] and 2ab ≤ a2 + b2, we obtain

for some Γ the inequalitydu(t)

dt≤ Γu(t).

Since u(0) = 0 from our initial order assumption, we conclude that u(t) = 0 for all t ≥ 0. That is the

conclusion of the lemma.

Exercise. Show that the result of Lemma 1.2 holds true for the system

∂tni − Di∆ni = Ri

(n1(t, x), n2(t, x), ..., nI(t, x)

).

Page 10: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

1.4. CHALLENGES 5

Exercise. For a cooperative system, assume that we are given an initial data such that (n01, n

02) is a

smooth subsolution of the steady state equation. Write the equation for ∂ni(t)∂t and prove that

∂ni(t)

∂t≥ 0.

Exercise. Write the proof that order is preserved for a general k × k cooperative system.

Exercise. In the case of 2 by 2 competitive systems, that means,

∂n2R1(n1, n2) ≤ 0,

∂n1R2(n1, n2) ≤ 0. (1.5)

Show that the order 0 ≤ n1 ≤ m1, 0 ≤ m2 ≤ n2 is preserved. There is no such general statement for

larger systems.

1.4 Challenges

Variability. There is always a large variability in living populations. For that reason, solutions of mod-

els with fixed parameters, for movement or growth, usually fit poorly with experiments or observations.

However, the models are useful for explaining qualitatively these observations but not for predicting in-

dividual behavior, a major problem when dealing with medical applications for example. A usual point

of view is that these parameters are distributed; one can use a statistical representation (see the software

Monolix at http://software.monolix.org/) in pharmacology. One can also use so-called structured popula-

tion models, see [32, 34, 38]. When modeling Darwinian evolution (see adaptive evolution in Section ??),

the parameters are part of the model solution and mutations are seen as changes in the model coefficients

selected by adaptation to the environment.

Small numbers. In physics and chemistry, 1023 is the normal order of magnitude for a number of

molecules. Populations that are studied with Lotka-Volterra equations are much smaller and 106 is al-

ready a large number. This means that exponentially decaying tails are very quickly meaningless because

they represent a number of individuals, less than one. The interpretation of such tails should be ques-

tioned carefully and several types of corrections can be included; demographic stochasticity is used in

Monte-Carlo simulations, which correspond to the survival threshold in PDEs1.

1Gauduchon, M. and Perthame, B. Survival thresholds and mortality rates in adaptive dynamic: conciliating determin-

istic and stochastic simulations. Mathematical Medicine and Biology 27 (2010), no. 3, 195–210

Page 11: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

6 Relaxation and the energy method

Page 12: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

Chapter 2

Reaction kinetics and entropy

When large numbers of molecules are involved in a chemical reaction, the kinetics are well described

by reaction rate equations. These are nonlinear equations describing the population densities (concen-

trations) of the reacting molecules and the specific form of these nonlinearities is usually prescribed by

the law of mass action. This section deals with this aspect of reaction kinetics. The derivation of these

equations is postponed to Section ?? where we introduce the chemical master equation, which better

describes a small number of reacting molecules, an important topic in cell biology.

2.1 Reaction rate equations

The general form of the equations leads to a particular structure for the right hand side of semi-linear

parabolic equations. They are written

∂tni − Di∆ni︸ ︷︷ ︸

molecular diffusion

+ ni Li = Gi︸ ︷︷ ︸reaction

, t ≥ 0, x ∈ Rd, i = 1, 2, ..., I. (2.1)

The quantities ni ≥ 0 are molecular concentrations, the loss terms Li ≥ 0 depend on all the molecules

nj (with which the molecule i can react) and the gain terms Gi ≥ 0, which also depends on the nj ’s,

denote the rates of production of ni from the other reacting molecules. The molecular diffusion rate of

these molecules is Di and can be computed according to the Einstein rule from the molecular size.

For a single reaction, the nonlinearities Li and Gi take the form

niLi =∑

reactions p

apikp

I∏j=1

napjj , Gi =

∑reactions p

bpi kp

I∏j=1

napjj . (2.2)

The powers apj , bpj ∈ N represent the number of molecules j necessary for the reaction p; this is the law

of mass action.

Nonnegativity property. We have factored the term ni in front of Li to ensure that this term van-

ishes at ni = 0. The loss term Li is not singular here because the product contains api ≥ 1 when

7

Page 13: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

8 CHAPTER 2. REACTION KINETICS AND ENTROPY

kpi > 0 (the reactant i really reacts, otherwise the corresponding Li vanishes). For that reason, as for

the Lotka-Volterra systems, the nonnegativity property holds true.

Lemma 2.1 A weak solution of (2.1) with Gi ≥ 0 and nonnegative initial data satisfies ni(t, x) ≥ 0,

∀i = 1, 2, ..., I.

Proof. Adapt the proof of Lemma 1.1.

Even though we do not know that Gi ≥ 0, this lemma is useful because one may argue as follows.

Solutions are first built using the positive part of Gi, then the lemma tells us that the ni’s are positive.

From formula (2.2), the positivity of the n′js implies that Gi is positive and thus that we have built the

solution to the correct problem.

Conservation of atoms. The second main property of these models comes from the conservation

of atoms, which asserts that some quantities should be constant in time. For each atom k = 1, ...,K, one

defines the number Nki of atoms k in the molecule i. Then all reactions should preserve the number of

atoms and, for all k ∈ 1, ...,K, the coefficients α, β, k and k′ should be such that

I∑i=1

Nki [ni Li −Gi] = 0, ∀(ni) ∈ RI . (2.3)

This implies that

∂t

I∑i=1

Nkini = ∆

I∑i=1

NkiDini, (2.4)

d

dt

∫Rd

I∑i=1

Nkini(t, x)dx = 0, 1 ≤ k ≤ K,

and thus, a priori, the weighted L1 bound holds∫Rd

I∑i=1

Nkini(t, x)dx =

∫Rd

I∑i=1

Nkin0i (x)dx, 1 ≤ k ≤ K. (2.5)

Except when all the diffusion rates Di’s are equal (then its main part is the heat equation), it is not easy

to extract conclusions from equation (2.4) except the conservation law (2.5). Very few general tools, such

as M. Pierre’s duality estimate, are available, see the survey1.

There is a third property, the entropy dissipation that we shall discuss later.

2.2 The law of mass action

Irreversible reaction. To begin with, consider molecular species Si undergoing the single irreversible

reaction

a1S1 + ...+ aISIk+−→ b1S1 + ...+ bISI ,

1Pierre, M. Global existence in reaction-diffusion systems with control of mass: a survey. Milan J. Math. 78 (2) (2010),

417–455

Page 14: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

2.2. THE LAW OF MASS ACTION 9

where ai and bi ∈ N, (ai)i=1,...,I 6= (bi)i=1,...,I , denote the number of molecules involved as reactants and

products. Then, using the law of mass action, the equation for the population number densities ni of

species i is written

∂tni −Di∆ni + aik+

I∏j=1

najj︸ ︷︷ ︸

loss of molecules by reaction

= bik+

I∏j=1

najj︸ ︷︷ ︸

gain of reaction product

, i = 1, 2, ..., I.

From the conservation of atoms, it is impossible that for all i we have bi ≥ ai (resp. ai ≤ bi). Therefore,

at least one reactant (ai < bi) should disappear and one product (bi > ai) be produced.

Reversible reaction. More interesting is when several reactions occur. Then they are represented by

a sum of such terms over each chemical reaction. For example, still with ai 6= bi ∈ N, the reversible

reaction

a1S1 + ...+ aISIk+−−k−

b1S1 + ...+ bISI ,

leads to reaction rate equations

∂∂tni −Di∆ni

loss of forward reacting molecules︷ ︸︸ ︷+ aik+

I∏j=1

najj +

loss of backward reacting molecules︷ ︸︸ ︷bik−

I∏j=1

nbjj

= bik+

I∏j=1

najj︸ ︷︷ ︸

gain of forward reaction product

+ aik−

I∏j=1

nbjj .︸ ︷︷ ︸

gain of backward reaction product

A more convenient form is

∂tni −Di∆ni = (bi − ai)

k+

I∏j=1

najj − k−

I∏j=1

nbjj

, i = 1, 2, ..., I. (2.6)

With this form, one can check the fundamental entropy property for reversible reactions. Define

S(t, x) :=

I∑i=1

ni[

ln(ni) + σi − 1], with

I∑i=1

σi(ai − bi) = ln k+ − ln k−. (2.7)

There are different possible choices for the constant σi, which can be more or less convenient depending

on the case. In all cases, one can check the

Lemma 2.2 (Entropy property for reversible reactions) The entropy dissipation equality holds

d

dt

∫S(t, x)dx = −

I∑i=1

Di

∫|∇ni|2

nidx−D(t, x) ≤ 0,

D(t, x) :=

∫ [ln(k+

I∏j=1

najj )− ln(k−

I∏j=1

nbjj )] [k+

I∏j=1

najj − k−

I∏j=1

nbjj

]dx ≥ 0.

Page 15: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

10 CHAPTER 2. REACTION KINETICS AND ENTROPY

The entropy dissipation property is very important because it dictates the long term behavior of the

reaction. Mathematically it is also useful to prove a priori estimates for the quantity |∇ni|2ni

but also for

integrability properties of powers of ni, which are necessary to define solutions of the reaction-diffusion

equation. This has been used recently in a series of papers by L. Desvillettes and K. Fellner2 for estimates

and relaxation properties, and by T. Goudon and A. Vasseur3 for regularity properties.

2.3 Elementary examples

Isomerization. The reversible isomerization reaction is the simplest chemical reaction. The atoms

within the molecule are not changed but only their spatial arrangement. The reaction is represented as

S1

k+−−k−

S2

and corresponds, in (2.6), to a1 = 1, b1 = 0, a2 = 0, b2 = 1, which meads to∂∂tn1 −D1∆n1 + k+n1 = k−n2,

∂∂tn2 −D2∆n2 + k−n2 = k+n1.

(2.8)

The conserved quantity is simply

d

dt

∫[n1(t, x) + n2(t, x)]dx = 0,

∫[n1(t, x) + n2(t, x)]dx =

∫[n0

1(x) + n02(x)]dx.

The formula (2.7) for the entropy (with a1 = b2 = 1, b1 = a2 = 0) gives

S(t, x) = n1

[ln(k+n1)− 1

]+ n2

[ln(k−n2)− 1

], (2.9)

and one can check the entropy dissipation relation

ddt

∫Rd S(t, x)dx = −

∫Rd

[D1|∇n1|2n1

+D2|∇n2|2n2

]dx

−∫Rd

[ln(k+ n2)− ln(k− n1)

][k−n2 − k+n1

]dx ≤ 0.

Dioxygen dissociation. The standard degradation reaction of dioxygen to monoxygen is usually asso-

ciated with hyperbolic models for fluid flows rather than diffusion. This is because it is a very energetic

reaction occurring at very high temperature (for atmosphere re-entry vehicles for example) and with

reaction rates depending critically on this temperature, see [17]. But for our purpose here, we forget

these limitations and consider the dissociation rate k+ > 0 of n1 = [O2] in n2 = [O], and conversely its

recombination with rate k− > 0

O2

k+−−k−

O +O.

2L. Desvillettes and K. Fellner, Entropy methods for reaction-diffusion equations: Slowly growing a-priori bounds. Rev.

Mat. Iberoamericana, 24 (2) (2008), 407–431.3T. Goudon and A. Vasseur, regularity analysis for systems of reaction-diffusion equations, Annales de l ’ENS, 43 (1)

(2010), 117–141.

Page 16: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

2.3. ELEMENTARY EXAMPLES 11

This leads to a1 = 1, b1 = 0, a2 = 0, b2 = 2, and thus we obtain∂∂tn1 −D1∆n1 + k+n1 = k−(n2)2,

∂∂tn2 −D2∆n2 + 2k−(n2)2 = 2k+n1,

(2.10)

with initial data n01 ≥ 0, n0

2 ≥ 0. According to the law of mass action, the term (n2)2 arises because the

encounter of two atoms of monoxygen is required for the reaction.

We derive the conservation law (number of atoms is constant) by a combination of the equations

∂t[2n1 + n2]−∆[2D1n1 +D2n2] = 0,

which implies that for all t ≥ 0∫Rd

[2n1(t, x) + n2(t, x)

]dx = M :=

∫Rd

[2n0

1(x) + n02(x)

]dx. (2.11)

For the simple case of the reaction (2.10), the formula (2.7) for the entropy (with a1 = 1, b2 = 2,

b1 = a2 = 0) gives

S(t, x) = n1

[ln(k+n1)− 1

]+ n2

[ln(k

1/2− n2)− 1

]. (2.12)

One can readily check that

Lemma 2.3 (Entropy inequality)

ddt

∫Rd S(t, x)dx = −

∫Rd

[D1|∇n1|2n1

+D2|∇n2|2n2

]dx

−∫Rd

[ln(k− (n2)2)− ln(k+ n1)

][k−(n2)2 − k+n1

]dx ≤ 0.

Exercise. Deduce from (2.11) and n1 ≥ 0, n2 ≥ 0 the a priori bound

2k−

∫ T

0

∫Rd

(n2)2 dxdt ≤M(1 + 2k+T ).

Hint. In (2.10), integrate the equation for n1.

Hemoglobin oxidation. Hemoglobin Hb can bind to dioxygen O2 to form HbO2 according to the

reaction

Hb+O2

k+−−k−

HbO2.

This reaction is important in brain imaging because the magnetic properties of Hb and HbO2 are dif-

ferent; the consumption of oxygen by neurons generates desoxyhemoglobin that can be detected by MRI

(magnetic resonance imaging) and thus, indirectly, indicates the location of neural activity.

The resulting system of PDEs, for n1 = [Hb], n2 = [O2] and n3 = [HbO2], is∂∂tn1 −D1∆n1 + k+n1n2 = k−n3,

∂∂tn2 −D2∆n2 + k+n1n2 = k−n3,

∂∂tn3 −D3∆n3 + k−n3 = k+n1n2.

Page 17: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

12 CHAPTER 2. REACTION KINETICS AND ENTROPY

From this system, we can derive two conservation laws for the total number of molecules Hb and O2,

∂t[n1 + n3]−∆[D1n1 +D3n3] = 0,

∂t[n2 + n3]−∆[D2n2 +D3n3] = 0.

these imply an L1 control

M :=

∫[n1(t) + n2(t) + n3(t) + n4(t)] =

∫[n0

1 + n02 + n0

3 + n04].

As in the case of dioxygene dissociation, integrating the third equation, one concludes the quadratic

estimate

k+

∫ ∞0

∫Rd

n1n2dx dt ≤M(1 + k−T ).

From (2.7) (with a1 = 1, b2 = 2, b1 = a2 = 0), this also comes with an entropy

S(t, x) = n1[ln(k1/2+ n1 − 1] + n2[ln(k

1/2+ n2 − 1] + n3[ln(k−n3 − 1].

Mathematical references and studies on this system can be found in B. Andreianov and H. Labani4.

Exercise. Another simple and generic example is the reversible reaction A+Bk+−−k−

C +D

∂∂tn1 −D1∆n1 + k+n1n2 = k−n3n4,

∂∂tn2 −D2∆n2 + k+n1n2 = k−n3n4,

∂∂tn3 −D3∆n3 + k−n3n4 = k+n1n2,

∂∂tn4 −D4∆n4 + k−n3n4 = k+n1n2.

(2.13)

with the ni ≥ 0 and the single atom conservation law∫Rd

[n1(t, x) + n2(t, x) + n3(t, x) + n4(t, x)

]dx = M :=

∫Rd

[n0

1(x) + n02(x) + n0

3(x) + n04(x)

]dx.

Choose ki so that S =

4∑i=1

ni ln(kini) is a convex entropy and write the entropy inequality for the chemical

reaction (2.13).

2.4 Small numbers of molecules

It may happen that in biological systems, e.g. intracellular bio-molecular interactions, the number of

molecules is not large so that the reaction rate equations are not applicable. The alternative is to use the

so-called chemical master equations that describe the dynamic of individual molecules.

These models are close to the jump processes described in Chapter ?? and with a brief presentation in

Section ??. They allow for a number of asymptotic limits (from discrete to continuous, from continuous

jumps to drift-diffusion). For these reasons, there is a large literature on this subject, on both the

theoretical and numerical aspects5,6,7.4B. Andreianov and H. Labani, Preconditioning operators and L∞ attractor for a class of reaction-diffusion systems.

Comm. Pure Appl. Anal. Vol. 11, No. 6 (2012) 2179–21995A. De Masi, E. Presutti, Mathematical methods for hydrodynamic limits. Springer-Verlag, Berlin, 19916D. T. Gillespie, Exact stochastic simulation of coupled chemical reactions. J. Phys. Chem. 81(25) 2340–2361 (1977)7J.-L. Doob, Markoff chains- Denumerable case. Trans. Amer. Math. Soc. 58(3), 455–473 (1945)

Page 18: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

Chapter 3

Slow-fast dynamics and enzymatic

reactions

In many situations, one can encounter very different time scales for different reactions, and more generally

phenomena. When a time scale is very short compared to the other, one wishes to consider that the fast

phenomenon is at equilibrium. This chapter gives circumstances where this is legitimate. A typical

example is the case of enzymatic reactions, where a low level of enzymes regulates a reaction. Other

examples arise in the spiking dynamic of neurons.

Throughout this chapter we ignore diffusion and work with ordinary differential equations. For a

parameter ε > 0, we study the system of two equations with a fast and a slow time scale εduε(t)dt = f

(uε(t), vε(t)

), uε(t = 0) = u0,

dvε(t)dt = g

(uε(t), vε(t)

), vε(t = 0) = v0.

(3.1)

Formally, the limiting system should be written as 0 = f(u(t), v(t)

),

dv(t)dt = g

(u(t), v(t)

), v(t = 0) = v0.

(3.2)

There are however several objections to such a general aymptotic result. Firstly, to pass to the limit in

nonlinear expressions as f(uε(t), vε(t)

), one needs compactness and for uε(t) the time derivative is not

under control. Secondly, the loss of an initial condition for u(t) should be explained. Thirdly, the relation

0 = f(u(t), v(t)

)does not necessarily define u(t) as a function of v(t) so as to insert it in the differential

equation for v(t). This last observation leads us to distinguish two cases

• ∂∂uf(u, v) < 0, then we can hope to find a smooth mapping F (v) such that f(F (v), v) = 0. We call it

the ‘monostable’ case.

• Otherwise, some organizing principle should be found to decide which branch of solution of f(u, v) = 0

is selected and we present some general results in section 3.2.

Page 19: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

14 Slow-fast dynamics and enzymatic reactions

The sign of ∂∂uf(u, v) also appears when one studies the stability of a steady state. The linearized

matrix is 1ε∂∂uf(u, v) 1

ε∂∂vf(u, v)

∂∂u (gu, v) ∂

∂v g(u, v)

=1

ε

∂∂uf(u, v) ∂

∂vf(u, v)

0 0

+O(1).

Again stability is related to the property ∂∂uf(u, v) < 0.

3.1 Slow-fast dynamics (monostable)

We begin with the simplest situation where a fast scale can be considered at equilibrium, meaning that

we can invert the relation

0 = f(u, v)⇐⇒ u = F (v),

with F a smooth function from R to R. Then, we end up with the single equation

dv(t)

dt= g(F (v(t)), v(t)

), v(t = 0) = v0.

In particular the initial data for u is lost, a phenomena called initial layer.

We are going to assume

f, g are Lipschitzian, (3.3)

∂f

∂u(u, v) ≤ −α < 0 for some constant α > 0. (3.4)

The second assumption implies that the equation on uε is monostable, as one can see it in the

Exercise. With the assumption (3.4), for the differential equation with v fixed

du(t)

dt= f

(u(t), v

), u(t = 0) = u0,

1. prove that |du(t)dt | ≤ |f(u0, v)|e−αt,

2. show that there is a limit u(t)→ U as t→∞.

Hint. Write the equation on z(t) = du(t)dt .

Theorem 3.1 We assume (3.3)–(3.4) and that for some T > 0 there is a constant M such that |uε(t)|+|vε(t)| ≤M . Then,

(i) vε(t) is Lipschitzian, uniformly in ε, for t ∈ [0, T ],

(ii) uε(t) is Lipschitzian, uniformly in ε, for t ∈ [τ,∞), for all τ ∈]0, T ],

(iii) there are u(t), v(t) ∈ C([0, T ]) such that, as ε→ 0, uε(t)→ u(t) uniformly in [τ, T ] and vε(t)→ v(t),

uniformly in [0, T ] and (3.2) holds.

Proof. The statement (i) is obvious since g(uε(t), vε(t)) is bounded. For statement (ii), we define

zε :=duε(t)

dt.

Differentiating the equation on uε, we compute using the chain rule

dzε(t)

dt= zε

∂f

∂u(uε, vε) +

∂f

∂v(uε, vε)

dvε(t)

dt,

Page 20: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

3.1. SLOW-FAST DYNAMICS (MONOSTABLE) 15

1

2

dzε(t)2

dt= z2

ε

∂f

∂u(uε, vε) + zε

∂f

∂v(uε, vε) g(uε, vε).

Thanks to assumption (3.3) and (3.4), we find

dzε(t)2

dt≤ −2αz2

ε + αz2ε +

1

α[∂f

∂v(uε, vε) g(uε, vε)]

2 ≤ −αz2ε +K

and thus

zε(t)2 ≤ z2

0e−αt/ε +

K

α=f(u0, v0)2

ε2e−αt/ε +

K

α.

This proves (ii).

For (iii) , we use the Ascoli-Arzela theorem and conclude that for a subsequence, we have uniform

convergence. Then we may pass to the limit in the integral form. The full family converges because of

uniqueness of solutions of (3.2).

Exercise. if the intital data is well-prepared, that means f(u0, v0) = 0, prove that uε(t) is uniformly

lipschitzian on [0,∞).

Exercise. Consider the system, with Lipschitz continous f and g, εduε(t)dt = f

(uε(t), vε(t)

), uε(t = 0) = u0 ∈ (0, 1),

dvε(t)dt = g

(uε(t), vε(t)

), vε(t = 0) = v0,

assuming that f(0, v) = f(1, v) = 0, f(u, v) ≥ 0 for u ∈ (0, 1).

1. Show that 0 < uε(t) < 1 for all times.

2. Show that∫∞

0|duε(t)

dt |dt ≤ 1 and conclude that, after extracting a subsequence, uε(t)→ u(t) as ε→ 0,

almost everywhere and in all Lp(0,∞), 1 ≤ p <∞.

3. Show that u = 1 and dv(t)dt = g

(1, v(t)

), v(t = 0) = v0.

We may also prove uniform bounds on uε(t) and vε(t).

Theorem 3.2 With the assumptions (3.3)–(3.4), for all T > 0, there is a constant M(T ) such that

|vε(t)| ≤M(T ), 0 ≤ t ≤ T,

Proof. We can write,

εduε(t)

dt= f

(0, vε(t)

)+

∫ uε(t),

0

∂f(σ, vε(t)

)∂u

dσ,

ε2d uε(t)2

dt ≤ uε(t)f(0, vε(t)

)− αuε(t)2

≤ α2 uε(t)

2 + C[1 + vε(t)2]− αuε(t)2

because f(·, ·) is Lipschitz continuous. In the same way, we find

ε2d vε(t)2

dt ≤ C|vε(t)|[1 + |uε(t)|+ |vε(t)|]

≤ α2 uε(t)

2 + C[1 + vε(t)2]

Page 21: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

16 Slow-fast dynamics and enzymatic reactions

Adding up these two inequalities rises

1

2

d

dt[vε(t)

2 + ε uε(t)2] ≤ C[1 + vε(t)

2] ≤ C[1 + vε(t)2 + ε uε(t)

2]

and the Gronwall inequality gives us that vε(t)2 + ε uε(t)

2 is locally bounded.

To bound uε(t)2 itself, we may now write the inequation for uε(t)

2 as

ε

2

d uε(t)2

dt≤ −α

2uε(t)

2 + C[1 +M(T )]

which proves the result because max((u0)2. 2C[1+M(T )

α

)is a solution with a larger initial data.

3.2 Slow-fast dynamics (general)

The assumption (3.4) is very strong and one can extend the result to more general systems but with

weaker conclusions. We keep the system εduε(t)dt = f

(uε(t), vε(t)

), uε(t = 0) = u0,

dvε(t)dt = g

(uε(t), vε(t)

), vε(t = 0) = v0,

(3.5)

f, g are Lipschitz continuous and C1. (3.6)

Theorem 3.3 With the assumption (3.6) and assuming that uε(t), vε(t) are bounded on [0, T ], we have∫ T

0

|f(uε(t), vε(t))|2dt ≤ C(T,K)ε.

If, for all v, f(u, v) does not vanish on an interval in u, then after extaction of a subsequence

uε(n) → u(t) a.e. vε(n) → v(t) uniformly on [0, T ]

and the relations holdsdv(t)

dt= g(u(t), v(t)

), f

(u(t), v(t)

)= 0.

Proof. 1. We define

F (u, v) =

∫ u

0

f(σ, v)dσ

and write ∫ T

0

|f(uε(t), vε(t))|2dt = ε

∫ T

0

f(uε(t), vε(t))duε(t)

dtdt.

SincedF (u(t), v(t))

dt= f(u(t), v(t))

du(t)

dt+∂F (u, v)

∂v

dv(t)

dt,

we conclude that∫ T

0

|f(uε(t), vε(t))|2dt = ε

∫ T

0

[dF (uε(t), vε(t))

dt− ∂F (uε, vε)

∂vg(uε, vε)

]dt

= ε[F (uε(T ), vε(T ))− F (u0, v0) + TK

].

Page 22: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

3.2. SLOW-FAST DYNAMICS (GENERAL) 17

Because uε(t), vε(t) are locally uniformly bounded, we conclude.

2. We define

G(u, v) =

∫ u

0

f(σ, v)2dσ.

We have

d

dtG(uε(t), vε(t)) = f(uε(t), vε(t))

2 duε(t)

dt+ 2g(uε(t), vε(t))

∫ uε(t)

0

f(σ, vε(t)) ∂vf(σ, vε(t))dσ

=f(uε(t), vε(t))

2

εf(uε(t), vε(t)) + 2g(uε(t), vε(t))

∫ uε(t)

0

f(σ, vε(t)) ∂vf(σ, vε(t))dσ

and thus G(uε(t), vε(t)) is locally BV. As a consequence for a subsequence, there is G(t) such that

G(uε(n)(t), vε(n)(t))→ G(t) a.e. as n→∞.

But because vε is uniformly Lipschitzian, by the Ascoli theorem, we also have vε(n)(t)→ v(t) and thus

G(uε(n)(t), v(t))→ G(t) a.e. as n→∞.

3. Because G(u, v) is strictly increasing in u from our assumption, the above convergence implies that

uε(n)(t) converges a.e. to some u(t) ∈ L∞(0, T ). The end follows immediately.

Exercise. (Morris-Lecar system) Consider the systemεdgε(t)

dt = G(vε(t))− gε(t),dvε(t)dt = g

(vε(t)

)+ gε(t)(V − vε(t)).

Identify the limiting system and show convergence.

Exercise. (Bistable case) Consider the case when, for a smooth function a : R→ (0, 1) we have

f(0, v) = f(1, v) = f(a(v), v) = 0,

f(u, v) =

< 0 for 0 < u < a(v),

> 0 for a(v) < u < 1.

1. Prove that∫ T

0|f(uε(t), vε(t))|dt ≤ C(T )ε.

2. Show that uε is uniformly with bounded total variation on [0, T ].

3. Show that, for a subsequence, as k →∞, vεk(t)→ v (uniformly), and uεk(t)→ u(t) (in Lp(0, T ), with

1 ≤ p <∞), with u, v a solution of (3.2) (almost everywhere).

[Hint.] Use the explicit formula for ϕ(u, v) :=∫ u

0sgn(f(σ, v)dσ

Exercise. (The FitzHugh-Nagumo system). This system, used in the neuroscience, is written εduε(t)dt = F

(uε(t)

)− vε(t), uε(t = 0) = u0,

dvε(t)dt = g

(uε(t), vε(t)

), vε(t = 0) = v0,

(3.7)

and F (u) = −u(u− a)(u+ a) is increasing (and thus unstable, see section 3.1) in the interval (− a√3, a√

3).

1. Show that∫ T

0|f(uε(t), vε(t))|dt ≤ C(T )ε.

Page 23: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

18 Slow-fast dynamics and enzymatic reactions

2. Show that, for a subsequence, as k →∞, vεk(t)→ v (uniformly), and uεk(t)→ u(t) (in Lp(0, T ), with

1 ≤ p <∞), with u, v a solution (almost everywhere) of F(u(t)

)= v,

dv(t)dt = g

(u(t), v(t)

), vε(t = 0) = v0,

(3.8)

Figure 3.1: Solution of the FitzHugh-Nagumo system (3.8) with F (u) = u(3 − u)(3 + u) and g(u, v) = u − .5.

Upper left u component, upper right v component. Lower red curve the nullcline v = F (u) and black curve

the solution u(t), v(t) in the phase plane (u, v).

3.3 Enzymatic reactions

More representative of biology are the enzymatic reactions, which are usually associated with the names

of Michaelis and Menten1. A substrate S can be transformed into a product P , without molecular change

as in the isomerization reaction, but this reaction occurs only if an enzyme E is present. The process

consists first of the formation of a complex SE, that can be itself dissociated into P +E. The formation

1Michaelis, L. and Menten, M. I., Die Kinetic der Invertinwirkung. Biochem. Z., 49, 333–369 (1913)

Page 24: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

3.3. ENZYMATIC REACTIONS 19

of the complex is a reversible reaction, but the conversion to P +E is an irreversible reaction, leading to

the representation

S + Ek+−−k−

C = SEkp−→ P + E.

Still using the law of mass action, this leads to the equations (we do not consider the space variable and

molecular diffusion, this is not the point here)ddtnS = k−nC − k+nSnE ,

ddtnE = (k− + kp)nC − k+nSnE ,

ddtnC = k+nSnE − (k− + kp)nC ,

ddtnP = kpnC .

This reaction comes with the initial data n0S > 0, n0

E > 0 and n0C = n0

P = 0. One can easily verify that

the substrate is entirely converted into the product

limt→∞

nP (t) = n0S , lim

t→∞nS(t) = 0, lim

t→∞nE(t) = n0

E , limt→∞

nC(t) = 0.

A conclusion that is incorrect if n0E = 0. This is a fundamental observation in enzymatic reactions that

a very small amount of enzyme is sufficient to convert the substrate into the product.

Two conservation laws hold true in these equations and this helps us to understand the above limit,nE(t) + nC(t) = n0

E + n0C = n0

E ,

nS(t) + nC(t) + nP (t) = n0S + n0

C + n0P = n0

S .(3.9)

Because the properties of the dynamic do not depend on nP , the system is equivalent to a 2×2 systemddtnS = k−nC − k+nS(n0

E − nC),

ddtnC = k+nS(n0

E − nC)− (k− + kp)nC .(3.10)

Briggs and Haldane(1925) arrived to the conclusion that this system can be further simplified with the

quasi-static approximation on nC . That means that nC(t) can be computed algebraically from nS(t),

thus leading to a single ODE for nS(t)ddtnS = k−nC − k+nS(n0

E − nC),

0 = k+nS(n0E − nC)− (k− + kp)nC ,

(3.11)

still with the initial data nS(t = 0) = n0S (but no initial condition on nC). We can of course write this in

a more condensed way. Since nC =k+nSn

0E

k−+k++k+nS, and after adding the two equations, we find the typical

enzymatic reaction kinetic equation

d

dtnS = −n0

E

kpk+nSk− + kp + k+nS

(3.12)

In other words, the law of mass action does not apply in this approximation and one finds the so-called

Michaelis-Menten law for the irreversible reactions, a Hill function rather than a polynomial.

Although not obvious at first glance, this result makes rigorous sense and we have the

Page 25: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

20 Slow-fast dynamics and enzymatic reactions

Proposition 3.4 (Validity of Michaelis-Menten law) Assume n0E is sufficiently small. The solu-

tions of (3.10) and of (3.12) satisfy

supt≥0|nS(t)− nS(t)| ≤ C(n0

S) n0E

for some constant C(n0S) independant of n0

E.

We recall that nC(t = 0) = uεC(t = 0) = 0 and nS(t = 0) = uεS(t = 0) > 0.

Proof. We reduce the system to slow-fast dynamic (see [13] for an extended presentation of the subject).

To simplify the notations, we set ε := n0E and we define the slow time and new unknowns as

ε := n0E , s = εt, uεS(s) = nS(t) ≥ 0, uεC(s) =

nC(t)

n0E

≥ 0.

Then the two systems are respectively written asddsu

εS = k−u

εC − k+u

εS(1− uεC), ε ddsu

εC = k+u

εS(1− uεC)− (k− + kp)u

εC ,

ddsuS = k−uC − k+uS(1− uC), 0 = k+uS(1− uC)− (k− + kp)uC .

(3.13)

With these notations the result in Proposition 3.4 is equivalent to

sups≥0|uεS(s)− uS(s)| ≤ Cε.

The situation is the favorable case of Section 3.1 with the addition that the convergence is uniform

globally and not only locally.

The details of the proof are left to the reader because it is a general conclusion from the theory of

slow-fast dynamic. We shall make use of the two consequences of (3.13)

d

ds[uεS + εuεC ] = −kpuεC ,

d

dsuS = −kpuC .

We have from (3.9) and uεC(t = 0) = 0,

(i) 0 ≤ uεC ≤ 1, 0 ≤ uC ≤ 1,

(ii) 0 ≤ uεS ≤ n0S , 0 ≤ uS ≤ n0

S ,

(iii) 0 ≤ uεC ≤ uM < 1, 0 ≤ uC ≤ uM < 1 with uM :=k+n

0S

k+n0S+k−+kp

independent of ε,

(iv) | ddsuεC(s)| ≤ K for some K(n0

S),

(v) From step (iv), introduce the bounded quantity rε(s) := − ddsu

εC(s) + k+u

εC(s)2, and

Rε(s) = uεS(s) + εuεC(s)− uS(s).

Compute that

uεC(s) =εrε(s)− k+u

εCR

ε + k+uεS

k+uS + k− + kp.

uεC − uC =εrε + k+(1− uεC)Rε − εk+u

εC

k+uS + k− + kp.

Page 26: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

3.4. BELOUSOV-ZHABOTINSKII REACTION 21

Conclude thatd

dsRε(s) + kp

k+(1− uεC)

k+uS + k− + kpRε(s) = εkp

−rε + k+uεC

k+uS + k− + kp.

(vi) Using steps (iii) and (v), conclude that |Rε(s)| ≤ Cε for a constant C.

Hints. (ii) uεS + uεC decreases.

(iii) Write ddsu

εC ≤ k+n

0S(1− uεC)− (k− + kp)u

εC .

(iv) Write the equation for z(s) = ddsu

εC(s) as ε ddsz+ λ(s)z(s) = µ(s) with λ > k−+ kp and µ bounded.

(vi) Write 12ddsR

ε(s)2 +ARε(s)2 ≤ εB|Rε(s)|, with A > 0, B > 0 two constants.

3.4 Belousov-Zhabotinskii reaction

Figure 3.2: Periodic regime of the system of ODEs (3.14) that mimics the Belousov-Zhabotinskii reaction.

The parameters are kA = kB = 1 and kC = 2 (Left), kC = 20 (Right). The solution nB is the top (red) curve,

nA is in blue and nC in green. In abscissae is time.

For all complex chemical reactions, the detailed description of all elementary reactions is not realistic.

Then, as for the enzymatic reaction, one simplifies the system by assuming that some reactions are much

faster than others, or that some components are in higher concentrations than others. These manipu-

lations may violate mass conservation and entropy inequality may be lost; this is a condition to obtain

pattern formation as in the example of the CIMA reaction in Section 7.5.2.

The famous Belousov-Zhabotinskii reaction is known as the first historical example to produce peri-

odic patterns. This was discovered in 1951 by Belousov, it remained unpublished because no respectable

chemist at that time could accept this idea. Belousov and Zhabotinskii received the Lenin Prize in 1980,

a decade after Belousov’s death and the discovery in the USA of other periodic reactions, simpler to

reproduce.

Page 27: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

22 Slow-fast dynamics and enzymatic reactions

We borrow a simple example from A. Turner2, this avoids entering the details of a real chemical reaction

(refer to [32] for a complete treatment). The simple example consists of three reactants denoted as A,

B, C, and three irreversible reactions (because of this feature, entropy dissipation and relaxation do not

hold).

A+BkA−→ 2A, B + C

kB−→ 2B, C +AkC−→ 2C.

Therefore, following the general formalism in (2.6), the system is written asddtnA = nA(kAnB − kCnC),

ddtnB = nB(kBnC − kAnA),

ddtnC = nC(kCnA − kBnB).

(3.14)

Notice that

nA(t) > 0, nB(t) > 0, nC(t) > 0, nA(t) + nB(t) + nC(t) = n0A + n0

B + n0C .

Numerical simulations are presented in Figure 3.2. For kc = 20, which is very large compared to kA

and kB one can estimate that the equations imply that kCnAnC = O(1) than thus nAnC ≈ 0. This is

oberved on the simulation on the right.

2A. Turner, A Simple Model of the Belousov-Zhabotinsky Reaction from First Principles, 2009,

http://www.aac.bartlett.ucl.ac.uk/processing/samples/bzr.pdf

Page 28: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

Chapter 4

Some material for parabolic

equations

4.1 Boundary conditions

When working in a domain Ω (a connected open set with sufficiently smooth boundary), we encounter

two elementary types of boundary conditions. The reaction-diffusion systems (1.1) or (2.1) are completed

either by the Dirichlet or the Neumann boundary conditions.

• Dirichlet boundary condition on ∂Ω.

ni = 0 on ∂Ω.

This boundary condition means that individuals or molecules, when they reach the boundary are defini-

tively removed from the population Ω, which therefore diminishes. This interpretation stems from the

Brownian motion underlying the diffusion equation (see below). But we can see that indeed, if we consider

the conservative chemical reactions (2.1) with (2.3), then

d

dt

∫Ω

I∑i=1

ni(t, x)dx =

I∑i=1

Di

∫∂Ω

∂νni,

with ν the outward normal to the boundary. But with ni(t, x) ≥ 0 in Ω and ni(t, x) = 0 in ∂Ω, we have∂∂νni ≤ 0, therefore the total mass diminishes, for all t ≥ 0,

d

dt

∫Ω

I∑i=1

ni(t, x)dx ≤ 0,

∫Ω

I∑i=1

ni(t, x)dx ≤∫

Ω

I∑i=1

n0i (x)dx.

• Neumann boundary condition on ∂Ω.

∂νni = 0 on ∂Ω,

still with ν the outward normal to the boundary. This means that individuals or molecules are reflected

when they hit the boundary. In the computation above for (2.1) with (2.3), the normal derivative vanishes

Page 29: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

24 Some material for parabolic equations

and we find directly mass conservation

d

dt

∫Ω

I∑i=1

ni(t, x)dx = 0,

∫Ω

I∑i=1

ni(t, x)dx =

∫Ω

I∑i=1

n0i (x)dx.

• Mixed or Robin boundary condition on ∂Ω.

αni +∂

∂νni = 0 on ∂Ω, α ≥ 0.

When reaching the boundary, there is either reflexion or removal with a propability related to α (α = 0

no removal, α =∞ no reflexion). The mass balance reads

d

dt

∫Ω

I∑i=1

ni(t, x)dx =

I∑i=1

Di

∫∂Ω

∂νni = −α

I∑i=1

Di

∫∂Ω

ni < 0.

• Full space. There is a large difference between the case of the full space Rd and the case of a bounded

domain. This can be seen by the results of spectral analysis in Section 4.2, which do not hold in the same

form in Rd. The countable spectral basis wk(x) is replaced by e−ix.ξ, and leads to the Fourier transform.

4.2 The spectral decomposition of Laplace operators (Dirichlet)

The spectral decomposition of the Laplace operator is going to play an important role in subsequent

chapters, therefore we introduce some material now. We begin with the Dirichlet boundary condition −∆u = f in Ω,

u = 0 on ∂Ω.(4.1)

Theorem 4.1 (Dirichlet) Consider a bounded connected open set Ω, then there is a spectral basis

(λk, wk)k≥1 for (4.1), that is,

(i) λk is a non-decreasing sequence with 0 < λ1 < λ2 ≤ λ3 ≤ ... ≤ λk ≤ ... and λk −→∞k→∞

,

(ii) (λk, wk) are eigenelements, i.e., for all k ≥ 1 we have −∆wk = λkwk in Ω,

wk = 0 on ∂Ω,

(iii) (wk)k≥1 is an orthonormal basis of L2(Ω),

(iv) we have w1(x) > 0 in Ω and the first eigenvalue λ1 is simple, and for k ≥ 2, the eigenfunction wk

changes sign and can be multiple.

Remark 1. The hypothesis that Ω is connected is simply used to guarantee that the first eigenvalue

is simple and the corresponding eigenfunction is positive in Ω. Otherwise, we have several additional

nonnegative eigenfunctions with a first eigenfunction in one component and 0 in the others.

Remark 2. The sequence wk is also orthogonal in H10 (Ω) (for Dirichlet boundary conditions) and H1(Ω)

Page 30: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

4.2. THE SPECTRAL DECOMPOSITION OF LAPLACE OPERATORS (DIRICHLET) 25

for Neumann boundary conditions (see below). Indeed, if wk is orthogonal to wj in L2(Ω), then from the

Laplace equation for wk and the Stokes formula∫Ω

∇wj .∇wk = λk

∫Ω

wjwk = 0.

Therefore, orthogonality in L2 implies orthogonality in H10 or H1.

Remark 3. Note that for k ≥ 2, the eigenfunction wk changes sign because∫

Ωw1wk = 0 and w1 has a

sign.

Remark 4. Notice the formula∞∑i=1

wi(x)wi(x′) = δ(x− x′).

Indeed, this means that for all test function in ϕ ∈ C(Ω) ∩ L2(Ω),

∞∑i=1

wi(x)

∫Ω

wi(x′)ϕ(x′)dx′ =

∫Ω

ϕ(x′)δ(x− x′)dx′ = ϕ(x),

which is the decomposition of ϕ on the orthonormal basis (wi)i≥1

Proof of Theorem 4.1. We only sketch the proof. For additional matters see [15] Ch. 5, [6] p. 96. The

result is based on two ingredients. (i) The spectral decomposition of self-adjoint compact linear mappings

on Hilbert spaces is a general theory that extends the case of symmetric matrices. (ii) The simplicity

of the first eigenvalue with a positive eigenfunction is also a consequence of the Krein-Rutman theorem

(infinite dimension version of the Perron-Froebenius theorem).

First step. 1st eigenelements. In the Hilbert space H = L2(Ω), we consider the linear subspace V =

H10 (Ω). Then we define the minimum on V

λ1 = min∫Ω|u|2=1

∫Ω

|∇u|2dx. (4.2)

This minimum is achieved because a minimizing sequence (un) will converge strongly inH and weakly in V

to w1 ∈ V with∫

Ω|w1|2 = 1, by the Rellich compactness theorem (see [10, 6]). Therefore,

∫Ω|∇w1|2dx ≤

lim infn→∞∫

Ω|∇un|2dx = λ1. This implies equality and that λ1 > 0. The variational principle associated

to this minimization problem says that

−∆w1 = λ1w1,

which implies that w1 is smooth in Ω (by elliptic regularity)

Second step. Positivity. Because in V ,∣∣∇|u|∣∣2 = |∇u|2 a.e., the construction above tells us that |w1| is

also a first eigenfunction and we may assume that w1 is nonnegative. By the strong maximum principle

applied to the Laplace equation, we obtain that w1 is positive inside Ω (because it is connected). This

also proves that all the eigenfunctions associated with λ1 have a sign in Ω because, in a connected open

set, w1 cannot satisfy the three properties (i) be smooth, (ii) change sign and (iii) |w1| be positive also.

Page 31: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

26 Some material for parabolic equations

Third step. Simplicity. Finally, we can deduce the simplicity of this eigenfunction because if there were

two independent eigenfunction, a linear combination would allow us to build one which changes sign (by

orthogonality to w1 for example) and this is impossible by the above positivity argument.

Fourth step. Other eigenelements. We may iterate the construction. Denote Ek the finite dimensional

subspace generated by the k-th first eigenspaces. We work on the closed subspace E⊥k of H, and we may

define

λk+1 = minu∈E⊥k ∩V,

∫Ω|u|2=1

∫Ω

|∇u|2dx.

This minimum is achieved by the same reason as before. The variational form gives that the minimizers

are solutions of the k + 1-th eigenproblem. They can form a multidimensional space but always finite

dimensional; otherwise we would have an infinite dimensional subspace of L2(Ω), whose unit ball is

compact thanks to the Rellich compactness theorem, since∫

Ω|∇u|2dx ≤ λk+1 in this ball

Also λk → ∞ as k → ∞ because one can easily build (with oscillations or sharp gradients) functions

satisfying∫

Ω|un|2 = 1 and

∫Ω|∇un|2dx ≥ n.

4.3 The spectral decomposition of Laplace operators (Neumann)

The same method applies to the Neumann boundary condition −∆u = f in Ω,

∂u∂ν = 0 on ∂Ω.

(4.3)

We state without proof the

Theorem 4.2 (Neumann) Consider a C1 bounded connected open set Ω, then there is a spectral basis

(λk, wk)k≥1 for (4.3), i.e.,

(i) λk is a non-decreasing sequence with 0 = λ1 < λ2 ≤ λ3 ≤ ... ≤ λk ≤ ... and λk −→∞k→∞

,

(ii) (λk, wk) are eigenelements, i.e., for all k ≥ 1 we have −∆wk = λkwk in Ω,

∂wk

∂ν = 0 on ∂Ω,

(iii) (wk)k≥1 is an orthonormal basis of L2(Ω)

(iv) w1(x) = 1|Ω|1/2 > 0, and for k ≥ 2, the eigenfunction wk changes sign.

4.4 The Poincare and Poincare-Wirtinger inequalities

As an outcome of the proof in Section 4.2, we recover the Poincare inequality for ω a bounded domain

λD1

∫Ω

|u|2 ≤∫

Ω

|∇u|2dx, ∀u ∈ H10 (Ω), (4.4)

Page 32: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

4.5. RECTANGLES: EXPLICIT SOLUTIONS 27

where λD1 denotes the first eigenvalue of the Laplace-Dirichlet equation (4.1). This inequality is just

another statement equivalent to (4.2).

The Poincare-Wirtinger inequality states that

λN2

∫Ω

|u− 〈u〉|2 ≤∫

Ω

|∇u|2dx, ∀u ∈ H1(Ω), (4.5)

where λN2 is the first positive eigenvalue for the Laplace-Neumann equation (4.3) and

〈u〉 =1

Ω

∫Ω

u(x)dx,

is the L2 projection on the first eigenfunction with Neumann boundary conditions (a constant). Therefore

this statement is nothing but the characterization of the second eigenvalue given in the proof in Section 4.2.

4.5 Rectangles: explicit solutions

In one dimension, one can compute explicitly the spectral basis because the solutions of −u′′ = λu are

all known. On Ω = (0, 1) we have

wk = ak sin(kπx), λk = (kπ)2, (Dirichlet)

wk = bk cos((k − 1)πx

), λk =

((k − 1)π

)2, (Neuman)

and ak and bk are normalization constants that ensure∫ 1

0|wk|2 = 1.

In two dimensions, on a rectangle (0, L1) × (0, L2) we see that the family is better described by two

indices k ≥ 1 and l ≥ 1 and we have

wkl = akl sin(kπx1

L1) sin(lπ

x2

L2), λkl =

((k

L1)2 + (

l

L2)2)π2, (Dirichlet)

wkl = bkl cos((k − 1)π

x

L1

)cos((l − 1)π

x

L2

), λkl =

((k − 1

L1)2 + (

l − 1

L2)2)π2, (Neuman)

These examples indicate that

• The first positive eigenvalue is of the order of 1/max(L1, L2)2 = 1diam(Ω)2 . For a large domain (even

with a large aspect ratio) we can expect that the first eigenvalues are close to zero and that the eigenvalues

are close to each other.

• Except the first one, eigenvalues can be multiple (take L1/L2 ∈ N).

• Large eigenvalues are associated with highly oscillating eigenfunctions.

Page 33: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

28 Some material for parabolic equations

Figure 4.1: Two sample paths of one dimensional Brownian motion according to the approximation (4.6).

Abscissae tk, ordinates Xk.

Figure 4.2: Two sample paths of two dimensional Brownian motion simulated according to (4.6).

4.6 Brownian motion and the heat equation

The relationship between Brownian motion and the heat equation explains, in a simple framework, why

a random walk of individuals leads to the terms −D∆n for the population density n(t, x). It is rather

difficult to construct rigorously Brownian motion. However, it is easy to give an intuitive approximation,

which is sufficient to build the probability law of Brownian trajectories.

To do so, we follow the Euler numerical method for an ODE with a time step ∆t that uses discrete

times tk = k∆t.

In a probability space with a measure denoted dP (ω), the initial position X0 ∈ Rd being given with a

probability law n0(x), we define iteratively a discrete trajectory Xk(ω) ∈ Rd as follows. We set

Xk+1(ω) = Xk(ω) +√

∆t Y k(ω) (4.6)

where Y k(ω) denotes a d-dimensional random variable independent of Xk (one speaks of independent

Page 34: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

4.6. BROWNIAN MOTION AND THE HEAT EQUATION 29

increments) and with a normal law N(y). We recall that ‘independent’ means

Ef(Xk, Y k) :=

∫f(Xk(ω), Y k(ω)

)dP (ω) =

∫Rd×Rd

f(x, y)dnk(x) N(y)dy (4.7)

with

• N(y) = 1(2π)d/2 e

−|y|2/2, the normal law,

• dnk(x) the law of the process Xk, defined as P(Φ(Xk(ω))

)=∫

Φ(x)dnk(x).

Two simulations are presented. Figure 4.1 depicts, in the one dimensional case (tk, Xk), two iterates

Xk (for two different ω). Figure 4.2 shows two iterates Xk in two dimensions.

Our purpose is to compute the law dnk(x) at the limit ∆t → 0. To do so, we first use a C3 function

u, with u, Du, D2u and D3u bounded. We use the Taylor expansion of u to compute

u(Xk+1) = u(Xk) +√

∆tDu(Xk).Y k +∆t

2D2u(Xk).(Y k, Y k) +O

(∆t|Y k|

)3/2,

and then

Eu(Xk+1) = Eu(Xk) +∆t

2ED2u(Xk).(Y k, Y k) +O

(∆t)3/2

. (4.8)

Indeed, because N(·) is radially symmetric, the formula (4.7) used with f(x, y) = Du(Xk).Y k yields

EDu(Xk).Y k = EDu(Xk).

∫Rd

yN(y)dy = 0.

Because∫Rd yiyjN(y)dy = δij , a further use of the formulas (4.8), (4.7) and the definition of the

probability law mentioned earlier, give∫Rd

u(x)dnk+1(x) =

∫Rd

u(x)dnk(x) +∆t

2

∫Rd

∆u(x)dnk(x) +O(∆t)3/2

.

After dividing by ∆t and integration by parts of the term ∆u, we may rewrite this as∫Rd

u[dnk+1 − dnk

∆t− 1

2∆dnk

]= O

(∆t)1/2

.

This holds true for any smooth function u and this means that, in the weak sense,

dnk+1 − dnk

∆t− 1

2∆dnk = O

(∆t)1/2

.

At the limit in the sense of distributions, as ∆t → 0, we obtain a probability law with a density n(t, x)

that satisfies ∂∂tn(t, x)− 1

2∆n(t, x) = 0,

n(0, x) = n0(x).

In particular, even though n0 is a probability measure, it follows from the regularizing effects of the heat

equation that n(t, x) is a (smooth) function and not only a measure as soon as t > 0.

Page 35: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

30 Some material for parabolic equations

The equation for n(t, x) (the heat equation here) is generically called the Kolmogorov equation for the

limiting process (Brownian motion).

Exercise. Prove that the density of the probability law nk satisfies the integral equation

nk+1(x)− nk(x) =

∫ [nk(x+

√∆t y)− nk(x)

]N(y)dy. (4.9)

Derive the heat equation for the limit n(t, x), as ∆t→ 0, if it exists. Show that this derivation uses only

two y-moments of N and not the full normal law.

See also the scattering equation in Chapter ??.

Page 36: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

Chapter 5

Relaxation, perturbation and

entropy methods

Solutions of nonlinear parabolic equations and systems presented in Chapter 1 can exhibit various and

sometimes complex behavior, a phenomena usually called pattern formation. In which circumstances can

such complex behavior happen? A first answer is given in this chapter by indicating some conditions for

relaxation to trivial steady states; then nothing interesting can happen! We present relaxation results by

perturbation methods (small nonlinearity) and entropy methods. Indeed, before we can understand how

patterns occur in parabolic systems, a necessary step is to understand why patterns should not appear

in principle! Solutions of parabolic equations undergo regularization effects that lead them to constant

or simple states. Several asymptotic results can be stated in this direction and we present some of them

in this chapter. Sections 5.1 and 5.2 have been very much influenced by the book [21].

5.1 Asymptotic stability by perturbation methods (Dirichlet)

The simplest possible long term behavior for a semilinear parabolic equation is simply the relaxation

towards a stable steady state (that we choose to be 0 here). This is possible when the following two

features are combined

• the nonlinear part is a small perturbation of a main (linear differential) operator,

• this main linear operator has a positive dominant eigenvalue.

Of course this simple relaxation behavior is rather boring and appears as the opposite of pattern forma-

tion when Turing instability occurs, and this will be described later, see Chapter 7.

To illustrate this, we consider, in a bounded domain Ω, the semi-linear heat equation with a Dirichlet

Page 37: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

32 Relaxation and the energy method

boundary condition∂∂tui(t, x)−Di∆ui(t, x) = Fi(t, x;u1, ..., uI), 1 ≤ i ≤ I, x ∈ Ω,

ui(t, x) = 0, x ∈ ∂Ω,

ui(t = 0, x) = u0i (x) ∈ L2(Ω).

(5.1)

We assume that F (t, x; 0) = 0, so that u ≡ 0 is a steady state solution. Is it a stable and attractive state?

Is the number of individuals leaving the domain through ∂Ω compensated by the birth term?

We shall use here a technical result. The Laplace operator (with Dirichlet boundary condition) has a

first eigenvalue λ1 > 0, associated with a positive eigenfunction, w1(x), which is unique except multipli-

cation by a constant,

−∆w1 = λ1 w1, w1 ∈ H10 (Ω). (5.2)

This eigenvalue is characterized as being the best constant in Poincare inequality (see Section 4.2 or the

book [10])

λ1

∫Ω

|v(x)|2 ≤∫

Ω

|∇v|2, ∀v ∈ H10 (Ω),

with equality only for v = µw1, µ ∈ R.

Theorem 5.1 (Asymptotic stability) Assume miniDi = D > 0 and that there is a (small) constant

L > 0 such that ∀u ∈ RI , t ≥ 0, x ∈ Ω,

|F (t, x;u)| ≤ L|u|, or more generally, F (t, x;u) · u ≤ L |u|2 , (5.3)

δ := Dλ1 − L > 0, (5.4)

then, ui(t, x) vanishes exponentially as t→∞, namely,∫Ω

|u(t, x)|2 ≤ e−2δt

∫Ω

|u0(x)|2. (5.5)

Proof. We multiply the parabolic equation (5.1) by ui and integrate by parts

1

2

d

dt

∫Ω

ui(t)2 +Di

∫Ω

∣∣∇ui(t)∣∣2 =

∫Ω

ui(t)Fi(t, x;u(t)

),

and using the characterization (5.2) of λ1, we conclude

1

2

d

dt

∫Ω

I∑i=1

ui(t)2 +Dλ1

∫Ω

I∑i=1

ui(t)2 ≤ L

∫Ω

I∑i=1

ui(t)2.

The result follows from the Gronwall lemma.

Exercise. Consider a smooth bounded domain Ω ⊂ Rd, a real number λ > 0 and two smooth and

Lipschitz continuous functions R(u, v), Q(u, v), such that R(0, 0) = Q(0, 0) = 0. Let(u(t, x), v(t, x)

)be

solutions of the system ∂∂tu−∆u = R(u(t, x), v(t, x)), t ≥ 0, x ∈ Ω,

u(t, x) = 0 sur ∂Ω,

∂∂tv + λv = Q(u(t, x), v(t, x)).

Page 38: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

5.2. ASYMPTOTIC STABILITY BY PERTURBATION METHODS (NEUMAN) 33

1. Recall the Poincare inequality for u.

2. Assume |R(u, v)| ≤ L(|u| + |v|) and |Q(u, v)| ≤ L(|u| + |v|), give a size condition for L such that for

all initial data, the solution (u, v) converges exponentially to (0, 0) for t→∞.

Solution. A simple answer is min(λ1, µ) − 2L =: δ > 0. A more elaborate condition is based on the

positive real number such that λ1 − µ = a− a−1 and is λ1 − L− a−1 = µ− L− a =: δ > 0.

5.2 Asymptotic stability by perturbation methods (Neuman)

The next simplest possible long term behavior for a parabolic equation is relaxation towards an homoge-

neous (i.e., independent of x) solution, which is not constant in time. This is possible when two features

are combined

• the nonlinear part is a small perturbation of a main (differential) operator,

• this main operator has a non-empty kernel (0 is the first eigenvalue).

Consider again, in a bounded domain Ω with outward unit normal ν, the semi-linear parabolic equation

with a Neumann boundary condition∂∂tui(t, x)−Di∆ui(t, x) = Fi(t;u1, ..., uI), 1 ≤ i ≤ I, x ∈ Ω,

∂∂ν(x)ui(t, x) = 0, x ∈ ∂Ω,

ui(t = 0, x) = u0i (x) ∈ L2(Ω).

(5.6)

The Laplace operator (with a Neumann boundary condition) has λ1 = 0 as a first eigenvalue, associated

with the constants w1(x) = 1/√|Ω| as eigenfunction. We shall use its second eigenvalue λ2, characterized

by the Poincare-Wirtinger inequality (see [10] and Section 4.2)

λ2

∫Ω

|v(x)− 〈v〉|2 ≤∫

Ω

|∇v|2, ∀v ∈ H1(Ω), (5.7)

with the average of v(x) defined as

〈v〉 =1

|Ω|

∫Ω

v.

Note that this is also the L2 projection on the eigenspace spanned by w1.

Theorem 5.2 (Relaxation to a homogeneous solution) Assume miniDi = D > 0 and(F (u)− F (v)

)· (u− v) ≤ L |u− v|2 , ∀u, v ∈ RI , (5.8)

δ := Dλ2 − L > 0, (5.9)

then, ui(t, x) tends to become homogeneous with an exponential rate, namely,∫Ω

|u(t, x)− 〈u(t)〉|2 ≤ e−2δt

∫Ω

|u0(x)− 〈u0〉|2. (5.10)

Page 39: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

34 Relaxation and the energy method

Proof. Integrating in x equation (5.6), we find

d

dt〈ui〉 = 〈Fi(t;u)〉,

therefore,d

dt[ui − 〈ui〉]−Di∆[ui − 〈ui〉] = Fi(t;u)− 〈Fi(t;u)〉.

Thus, using assumption (5.8), we find

12ddt

∫Ω|u− 〈u〉|2 +D

∫Ω|∇(u− 〈u〉

)|2 =

∫Ω

(F (t;u)− 〈F (t;u)〉

)· (u− 〈u〉)

=∫

ΩF (t;u) · (u− 〈u〉)

=∫

Ω

(F (t;u)− F (t; 〈u〉)

)· (u− 〈u〉)

≤ L∫

Ω|u− 〈u〉|2.

Therefore, with notation (5.9)

d

dt

∫Ω

|u− 〈u〉|2 ≤ −2δ

∫Ω

|u− 〈u〉|2.

The result (5.10) follows directly.

Exercise. Explain why we cannot allow a dependency on x in the nonlinearity Fi(t;u).

Exercise. Let v ∈ H2(Ω) satisfy ∂v∂ν = 0 on ∂Ω (Neumanncondition).

1. Prove, using the Poincare-Wirtinger ineqality (5.7), that

λ2

∫Ω

|∇v|2 ≤∫

Ω

|∆v|2. (5.11)

2. In the context of Theorem5.2, assume that

I∑i,j=1

DjFi(t;u)ξiξj ≤ L|ξ|2, ∀ξ ∈ RI . Using the above

inequality, prove that ∫Ω

∑i

|∇ui(t, x)|2 ≤ e−2δt

∫Ω

∑i

|∇u0i (x)|2. (5.12)

3. Deduce a variant of Theorem 5.2.

Hints. 1. Integrate by parts the expression∫

Ω|∇v|2. 2. Use the equation for d

dt∂ui

∂xl.

5.3 Entropy and relaxation

We have seen in Lemma 2.3 that reaction kinetic equations as in (2.10) are endowed with an entropy

(2.12). This originates from the microscopic N -particle stochastic systems from which reaction kinetics

are derived (see Section ??).

Page 40: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

5.3. ENTROPY AND RELAXATION 35

This entropy inequality is also very useful because it can be used to show relaxation to the steady

state, independently of the size of the constants k1, k2. To do that we consider a Neumann boundary

condition in a bounded domain Ω∂∂tn1 −D1∆n1 + k1n1 = k2(n2)2, t ≥, x ∈ Ω,

∂∂tn2 −D2∆n2 + 2k2(n2)2 = 2k1n1,

∂∂νn1 = ∂

∂νn2 = 0 on ∂Ω.

(5.13)

Theorem 5.3 The solutions of (5.13), with n0i ≥ 0, n0

i ∈ L1(Ω), n0i ln(n0

i ) ∈ L1(Ω), satisfy that

ni(t, x)→ Ni, as t→∞

with Ni the constants defined uniquely by

2N1 +N2 =

∫Ω

[2n01(x) + n0

2(x)]dx, k2(N2)2 = k1N1.

Proof. Then S(t, x) = n1

[ln(k1n1)− 1

]+ n2

[ln(k

1/22 n2)− 1

]satisfies, following Lemma 2.3,

ddt

∫ΩS(t, x)dx = −

∫Ω

[D1|∇n1|2n1

+D2|∇n2|2n2

]dx

−∫

Ω

[ln(k2 n

22)− ln(k1 n1)

][k2(n2)2 − k1n1

]dx.

And, because S is bounded from below, it also gives a bound on the entropy dissipation∫∞0

∫Ω

[D1|∇n1|2n1

+D2|∇n2|2n2

]dxdt

+∫∞

0

∫Ω

[ln(k2 n

22)− ln(k1 n1)

][k2(n2)2 − k1n1

]dxdt ≤ C(n0

1, n02).

(5.14)

This is again a better integrability estimate in x than the L1log estimate (derived from mass conservation)

because of the quadratic term (n2)2.

From a qualitative point of view, this says that the chemical reaction should lead the system to a

spatially homogeneous equilibrium state. Indeed, formally at least, the integral (5.14) can be bounded

only if

∇n1 = ∇n2 ≈ 0 as t→∞,

k2(n2)2 ≈ k1n1, ∇n1 = ∇n2 ≈ 0 as t→∞.

The first conclusion says that the dynamic becomes homogeneous in x (but may depend on t). The second

conclusion, combined with the mass conservation relation (2.11) shows that there is a unique possible

asymptotic homogeneous state because the constant state satisfies k2(N2)2 = k1N1 = k1( M|Ω|−N2), which

has a unique positive root.

Exercise. Consider (2.10) with a Neumann boundary condition and set

S(t, x) =1

k1Σ1

(k1n1(t, x)

)+

1

2k1/22

Σ2

(k

1/22 n2(t, x)

).

Page 41: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

36 Relaxation and the energy method

1. We assume that Σ1, Σ2 satisfy Σ′1(u) = Σ′2(u1/2). Show that Σ1 is convex if, and only if, Σ2 is convex.

2. Under the conditions in question 1, show that the equation (2.10) dissipates entropy.

3. Adapt Stampacchia’s method to prove L∞ bounds on n1, n2 (and which is the appropriate quantity

for the maximum principle). What are the intuitive Lp bounds.

Exercise. Consider (2.10) with Dirichlet boundary conditions and n0i ≥ 0.

1. Using the inequality∂nj

∂ν ≤ 0, which holds because nj ≥ 0, show that M(t) decreases where

M(t) =

∫Ω

[2n1(t, x) + n2(t, x)

]dx.

2. Consider the entropies of the previous exercise with the additional condition Σ′i(0) = 0. Show that

the equation (2.10) dissipates entropy.

(iii) Show that solutions of (2.10) with Dirichlet boundary conditions tend to 0 as t→∞.

5.4 Entropy: chemostat and SI system of epidemiology

Examples of entropy also appear in biological models. We treat here an example that arises as a first

modeling stage in two applications: 1. ecology and the model of the chemostat (the u represents a

nutrient, v a population that consumes the nutrient), 2. epidemiology with the celebrated Suceptible-

Infected model (SI system). In both cases the model isddtu = B − µuu− ruv,ddtv = ruv − µvv,

(5.15)

with B > 0, r > 0, µu > 0 and µv > 0 parameters.

In the chemostat B represents the renewal of nutrients u, µ the removal or degradation of nutrients

and r the consumption rate by the population and µv is the mortality and removal rate of the population.

In epidemiology, B represents the newborn, µu the mortality rate, r the encounter rate between sus-

ceptible and infected individuals (these encounters are responsible of new infected), µv is the mortality

of the infected (and recovery rate in the SIR model).

There is always a trivial (healthy) steady state

u0 = B/µu, v0 = 0,

and a positive steady state

u = µv/r, v =rB − µuµv

rµvif rB > µuµv. (5.16)

Depending on the sign of v we may have two different Lyapunov functionals (entropies)

Lemma 5.4 A ssume rB > µuµv and define

S(u, v) = −u ln(u)− v ln(v) + u+ v,

Page 42: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

5.5. THE LOTKA-VOLTERRA PREY-PREDATOR SYSTEM WITH DIFFUSION (PROBLEM) 37

we haved

dtS = − 1

u(√uB − u

√vr + µu)2. (5.17)

Lemma 5.5 We assume rB ≤ µuµv and define

S(u, v) = −u0 ln(u) + u+ v,

we haved

dtS = − v

µu(µvµu − rB)− 1

µuu(B − µuu)2. (5.18)

We leave the proofs of these lemmas to the reader and go directly to the conclusion

Proposition 5.6 Solutions of the system (5.15) behave as follows

If Br > µuµv, then the entropy S is convex and all solutions with v0 > 0 converge as t → ∞ to the

positive steady state.

If Br ≤ µuµv, then solutions become extinct (they converge to the trivial steady state as t → ∞).

The proof is standard and left to the reader. The steps are (i) u(t) is bounded, (ii) S(t) decreases,

in the case v < 0, the limit can be −∞ and this means that v(t) vanishes and the result (ii) follows,

otherwise S stays bounded and converges to a finite value, (iii) we conclude thanks to the right hand side

of (5.17).

5.5 The Lotka-Volterra prey-predator system with diffusion (Prob-

lem)

In the case of the Lotka-Volterra prey-predator system we can show relaxation towards a homogeneous

solution. The coefficients of the model need not be small as is required in Theorem 5.2. This is because

the model comes with a physically relevant quantity (such as entropy), which gives a global control.

Exercise. Consider the prey-predator Lotka-Volterra system without diffusion∂∂tn1 = n1[ r1 − an2],

∂∂tn2 = n2[−r2 + bn1],

where r1, r2, a and b are positive constants and the initial data n0i are positive.

1. Show that there are local solutions and that they remain positive.

2. Show that the entropy (Lyapunov functional)

E(t) = −r1 lnn2 + an2 − r2 lnn1 + bn1,

Page 43: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

38 Relaxation and the energy method

is constant. Show that E is bounded from below and that E → ∞ as n1 + n2 → ∞. Conclude that

solutions are global.

3. What is the unique steady state solution?

4. Show, using question 2, that the solutions are periodic (trajectories are closed).

Solution 1. By local Lipschitz regularity of the right hand side, the Cauchy-Lipschitz theorem asserts

that there is a local solution, i.e., defined on a maximal interval [0, T ∗]. We can write

n1(t) = n01e

∫ t0

[r1−an2(s)]ds > 0,

and, same thing for n2.

2. Set ϕi = lnni and write ∂∂tϕ1 = r1 − an2 = r1 − aeϕ2 ,

∂∂tϕ2 = −r2 − bn1 = −r2 + beϕ1 .

This is a Hamiltonian system and the Hamiltonian is constant along trajectories

H(t) = −r1ϕ2(t) + aeϕ2(t) − r2ϕ1(t) + beϕ1(t)] = H(0).

This proves that the ϕi(t) remain bounded from above and below, and thus that solutions are global.

3. n1 = r2/b, n2 = r1/a.

4. In each quadrant defined by the origin (n1, n2), we can write n2 as a function of n1 (or vice versa)

and conclude from that.

Exercise. Let Ω be a smooth bounded domain. Consider smooth positive solutions of the Lotka-Volterra

equation with diffusion and a Neumann boundary condition∂∂tn1 − d1∆n1 = n1[ r1 − an2],

∂∂tn2 − d2∆n2 = n2[−r2 + bn1],

∂ni

∂ν = 0 on ∂Ω, i = 1, 2,

where d1, d2, r1, r2, a and b are positive constants and the initial data n0i are positive.

1. Consider the quantity m(t) =∫

Ω[bn1(t, x) + an2(t, x)]dx. Show that m(t) ≤ m(0)ert and find the

value r.

2. Show that the convex entropy

E(t) =

∫Ω

[−r1 lnn2 + an2 − r2 lnn1 + bn1]dx,

a) is bounded from below, b) is decreasing.

Conclude that m(t) is bounded.

3. What finite integral do we obtain from the entropy dissipation?

3. Assume that the quantities ∇ lnni(t, x) converge, as t→∞,

a. What are their limits?

b. What can you conclude about the behavior of ni(t, x) as t→∞?

Page 44: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

5.6. PROBLEM 39

5.6 Problem

The goal of this problem is to show that for rank-1 nonlinearities, the long term behaviour is determinded

by simple states in x without a size condition.

Let Ω a smooth bounded domain. Consider a smooth positive solution of the Lotka-Volterra system

with diffusion and Neumann boundary condition∂∂tni −Di∆ni + ai(x)ni = ni ri(t), t ≥ 0, x ∈ Ω, i = 1, 2, ..., I,

∂ni

∂ν = 0 on ∂Ω, i = 1, 2, ..., I,

ni(t = 0, x) = n0i (x),

with the nonlinearity defined, for some given positive and smooth functions(ψi(x)

)i=1,...,I

, by

ri(t) = Ri( ∫

Ω

ψ1(x)n1(t, x)dx, ...,

∫Ω

ψI(x)nI(t, x)dx).

1. We also define the first eigenfunctions Ni(x) > 0 for the eigenvalue λi, defined by −Di∆Ni + ai(x)Ni = λiNi, x ∈ Ω,

∂Ni

∂ν = 0 on ∂Ω,∫

ΩNi(x)2dx = 1, i = 1, 2, ..., I.

Explain why the pair (Ni, λi) exists and give the corresponding Poincare-Wirtinger inequality.

2. We consider ni the solution of∂∂t ni −Di∆ni + ai(x)ni = λini, t ≥ 0, x ∈ Ω, i = 1, 2, ..., I,

∂ni

∂ν = 0 on ∂Ω, i = 1, 2, ..., I,

ni(t = 0, x) = n0i (x).

Prove that∫

Ωni(t, x)Ni(x)dx =

∫Ωn0i (x)Ni(x)dx. Find the constant αi such that, as t→∞,

‖ni(t, x)− αiNi(x)‖L2(Ω) ≤ ‖n0i (x)− αiNi(x)‖L2(Ω)e

−µit

and identify µi.

3. We write the solution of (5.6) as ni(t, x) = ρi(t)ni(t, x). Identify the evolution equation giving ddtρi(t)

in terms of the ρj and nj .

4. Assume that for some M > 0 we have Ri(Y1, ...YI

)< λi whenever Yi > M , i = 1, ..., I. Show that the

ρi(t) are bounded and that for some µ > 0 we have

‖ni(t, x)− ρi(t)αiNi(x)‖L2(Ω) ≤ C0e−µt.

5. In dimension 1, assuming R(+∞) < λ, R(−∞) > λ and R′ < 0. Show that the long term dynamic is

given by that of the equation

%(t) = %(t) R(%(t)α

∫ψ(x)N(x)dx

),

and that n(t, x) converges generically to a steady state ρN(x) as t→∞.

[Hint.] ρi(t) = ρi(t)(ri − λi)

Page 45: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

40 Relaxation and the energy method

Page 46: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

Chapter 6

Blow-up and extinction of solutions

We know from Chapter 5 that for ‘small’ nonlinearities, the solutions of parabolic systems relax to an

elementary state. When the nonlinearity is too large, it is possible that solutions of nonlinear parabolic

equations do not exist for all times or simply vanish, two scenarios that are the first signs of visible

nonlinear effects.

The mechanisms can be seen in simple ordinary differential equations. Blow-up means that the solution

becomes pointwise larger and larger and eventually becomes infinite in finite time. To see this, consider

the equation

z(t) = z(t)q, z(0) > 0, q > 1.

Its solution z(t) > 0 is given by

z(t)q−1 = z(0)q−1/(1− (q − 1)tz(0)q−1

)and thus it tends to infinity at t = T ∗ := 1

(q−1)z(0)q−1 . For q = 2, it is a typical illustration of the

alternative arising in the Cauchy-Lipschitz Theorem; solutions can only tend to infinity in finite time

(case z(0) > 0), or they are globally defined (case z(0) < 0).

The mechanism of extinction is illustrated by the equation with the opposite sign

z(t) = −z(t)q, z(0) > 0, q > 1.

Its solution

z(t)q−1 = z(0)q−1/(1 + (q − 1)tz(0)q−1

)(6.1)

vanishes as t→∞.

The purpose of this Chapter is to study in which respect these phenomena can occur or not when

diffusion is added.

We begin with the semilinear parabolic equation with Dirichlet boundary conditions

∂tu−∆u = u2.

We present several methods for proving blow-up in finite time. The first two methods are for semi-linear

parabolic equations, the third method is illustrated by the Keller-Segel system for chemotaxis; this is

41

Page 47: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

42 Blow-up and extinction of solutions

more interesting because it blows-up in all Lp norms, for all p > 1, but the L1 norm is conserved (this

represents the total number of cells in a system where cell multiplication is ignored).

The last section is devoted to a counter-intuitive result with respect to extinction.

6.1 Semilinear equations; the method of the eigenfunction

To study the case of nonlinear parabolic equations, we consider the model∂∂tu−∆u = u2, t ≥ 0, x ∈ Ω,

u(x) = 0, x ∈ ∂Ω,

u(t = 0, x) = u0(x) ≥ 0.

(6.2)

Here we treat the case when Ω is a bounded domain. To define a distributional solution is not completely

obvious because the right hand side u2 should be well defined, that is why we require that u belongs to

L2; then it remains to solve the heat equation with a right hand side in L1. That is why we call solution

of (6.2), a function satisfying for some T > 0,

u ∈ L2((0, T )× Ω

), u ∈ C

([0, T );L1(Ω)

). (6.3)

The question then is to know if effects of diffusion are able to overcome the effects of quadratic non-

linearity and (6.2) could have global solutions. The answer is given by the next theorems

Theorem 6.1 Assume that Ω is a bounded domain, u0 ≥ 0, u0 is sufficiently large (in a weighted L1

space introduced below), then there is a time T ∗ for which the solution of (6.2) satisfies

‖u‖L2((0,T )×Rd) −→T→T∗

∞.

Of course, this result means that u(t) also blows up in all Lp norms, 2 ≤ p ≤ ∞ because we work in a

bounded domain. It is also possible to prove that the blow-up time is the same for all these norms [35].

Proof of Theorem 6.1. We are going to arrive at a contradiction in the hypothesis that a function

u ∈ L2((0, T )× Ω

)can be a solution of (6.2) when T exceeds an explicit value computed later.

First we, notice that u(t, x) ≥ 0 because u0 ≥ 0.

Because we are working in a bounded domain, the smallest eigenvalue of the operator −∆ exists and

is associated with a positive eigenfunction (see Section 4.2) −∆w1 = λ1w1, w1 > 0 in Ω,

w1(x) = 0, x ∈ ∂Ω,∫

Ω(w1)2 = 1.

(6.4)

Page 48: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

6.1. SEMILINEAR EQUATIONS; THE METHOD OF THE EIGENFUNCTION 43

For a solution, we can multiply equation (6.2) by w1 and integrate by parts. We arrive at

ddt

∫Ωu(t, x) w1(x) dx =

∫Ω

∆u(t, x) w1(x) dx+∫

Ωu(t, x)2 w1(x) dx

=∫

Ωu(t, x) ∆w1(x) dx+

∫Ωu(t, x)2 w1(x) dx

= −λ1

∫Ωu(t, x) w1(x) dx+

∫Ωu(t, x)2 w1(x) dx

≥ −λ1

∫Ωu(t, x) w1(x) dx+

(∫Ωu(t, x) w1(x) dx

)2 (∫Ωw1(x) dx

)−1

(after using the Cauchy-Schwarz inequality). We set z(t) = eλ1t∫

Ωu(t, x) w1(x) and, with a =

(∫Ωw1(x) dx

)−1,

the above inequality readsd

dtz(t) ≥ ae−λ1t z(t)2,

and we obtain:d

dt

1

z(t)≤ −ae−λ1t,

1

z(t)≤ 1

z0− a1− e−λ1t

λ1.

Assume now the size condition

z0 >λ1

a. (6.5)

The above inequality contradicts that z(t) > 0 for e−λ1t ≤ 1− λ1

az0 . Therefore, the computation, and thus

the assumption that u ∈ L2((0, T )× Rd), fails before that finite time.

The size condition is necessary. There are various ways to understand this. For d ≤ 5 it follows from

a general elliptic equation

Theorem 6.2 There is a steady state solution u > 0 in Ω to −∆u = up, x ∈ Ω,

u(x) = 0, x ∈ ∂Ω,

when p satisfies

1 < p <d+ 2

d− 2.

We refer to [10] for a proof of this theorem and related results (as non-existence for p > d+2d−2 ).

One can see more directly that the size condition is needed. We choose µ = minΩλ1

w1(x) and set

w = µw1. This is a supersolution of (6.2) because

∂tw −∆w = λ1w ≥ w2.

One concludes that, when u0 ≤ w, solutions of (6.2) satisfy u(t) ≤ w for all times t, as long as the

solution exists (and this is sufficient to prove that the solution is global). Therefore, we have the

Page 49: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

44 Blow-up and extinction of solutions

Lemma 6.3 Under the smallness condition u0 ≤ minΩλ1

w1(·) w1, there is a global solution of (6.2) and

u(t) ≤ w, ∀t ≥ 0.

Proof. We subtract a solution u to w and find

∂t[u− w]−∆[u− w] ≤ u2 − w2 = [u− w][u+ w],

with u− w(t = 0) ≤ 0. From the comparison principle, we conclude that u− w(t) ≤ 0 for all times where

the solution exists. Therefore, by continuation methods, there is a global solution.

Exercise. Prove blow-up in finite time for the case of a general nonlinearity∂∂tu−∆u = f(u), t ≥ 0, x ∈ Ω,

u(x) = 0, x ∈ ∂Ω,

u(t = 0, x) = u0(x) > 0 large enough,

(6.6)

and f(u) ≥ cuα with α > 1.

Exercise. (Neumann boundary conditions) A solution in (0, T ) of the equation∂∂tu−∆u = u2, t ≥ 0, x ∈ Ω,

∂∂νu(x) = 0, x ∈ ∂Ω,

u(t = 0, x) = u0(x) ≥ 0,∫

Ωu0 > 0,

(6.7)

is a distributional solution satisfying u ∈ L2((0, T )× Ω).

Prove that there is no such solution after some time T ∗ and that ‖u(T )‖L1(Rd) −−−−−→T→T∗ ∞ (no size condi-

tion is required here).

[Hint] Because∫

Ωu(T ) =

∫Ωu0 +

∫ T0

∫Ωu2(t, x)dx dt, both u ∈ L2((0, T )×Ω) and ‖u(T )‖L1(Rd) blow-up

simultaneously. Conclude using the Cauchy-Schwarz inequality.

Exercise. For d = 1 and Ω =]− 1, 1[, construct a unique, even and non-zero solution of

−∆v = v2, v(±1) = 0.

Hint. Reduce it to − 12 (v′)2 = 1

3v3 + c0 and find a positive real number c1 such that

v′ = −√c1 −

2

3u3.

6.2 Semilinear equations; the energy method

Still considering semilinear parabolic equations, we present another method leading to a different size

condition. This uses the intrinsic properties of the equation better and does not use the sign condition.

Page 50: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

6.2. SEMILINEAR EQUATIONS; THE ENERGY METHOD 45

We consider a more general case with p > 1∂∂tu−∆u = |u|p−1u, t ≥ 0, x ∈ Ω,

u(x) = 0, x ∈ ∂Ω,

u(t = 0, x) = u0(x) 6= 0.

(6.8)

Before we state our result, it is useful to recall the energy principle underlying this equation. We define

the energy

E(u) =1

2

∫Ω

|∇u|2 − 1

p+ 1

∫Ω

|u|p+1.

Note that this has a mechanical interpretation: the term 12

∫Ω|∇u|2 is the kinetic energy and the term

− 1p+1

∫Ω|u|p+1 is the potential energy.

One can easily see that this energy decreases with time (in fact this comes from the structure of a

gradient flow for (6.8)). Indeed, using the chain rule and integration by parts, we have

ddtE

(u(t)

)=

∫Ω

[∇u.∇∂u

∂t − |u|p−1u∂u∂t

]= −

∫Ω∂u∂t

[∆u+ |u|p−1u

]= −

∫Ω

[∆u+ |u|p−1u

]2 ≤ 0.

Consequently, we have

E(t) ≤ E0 := E(u0). (6.9)

This explains the central role of the energy in

Theorem 6.4 For u0 ∈ H10 (Ω) satisfying E(u0) ≤ 0, there are no global solution of (6.8) with bounded

energy.

Proof. We define α = 12 −

1p+1 > 0. We combine the L2-estimate with the energy decay and obtain

14ddt

∫Ωu2dx = − 1

2

∫Ω

[ |∇u|2 − |u|p+1]dx

= −E(u) + α∫

Ω|u|p+1

≥ −E(u0) + α|Ω|(1−p)/2(∫

Ωu2)(p+1)/2

.

The last inequality uses Jensen’s inequality for the probability measure dx/|Ω|(∫Ω

u2 dx

|Ω|

)(p+1)/2

≤∫

Ω

up+1 dx

|Ω|.

This proves blow-up for E(u0) ≤ 0 because the positive function z(t) :=∫

Ωu2 satisfies

dz(t)

dt≥ βz(t)(p+1)/2, β := α|Ω|(1−p)/2 > 0,

and thus it blows-up in finite time, as shown in the introduction.

Page 51: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

46 Blow-up and extinction of solutions

Exercise. Find functions u0 ∈ H10 (Ω) that satisfy the condition E(u0) ≤ 0.

Hint. Use the fact that the two terms in the definition of the energy have different scales in u and consider

expressions as rV (x).

Exercise. Prove blow-up for(∫

Ω(u0)2

)(p+1)/2> 2

βE(u0).

The topic of blow-up is very rich and many modalities of blow-up, different blow-up rates, blow-up for

different norms and regularizing effects that prevent blow-up are also possible. See [35].

6.3 Keller-Segel system; the method of moments

We come back to the Keller-Segel model used to describe chemotaxis as mentioned in Section ??. We

recall that it consists of a system which describes the evolution of the population density of cells (bacteria,

amoebia,...) n(t, x), t ≥ 0, x ∈ Rd and the concentration c(t, x) of the attracting molecules released by

the cells themselves, ∂∂tn−∆n+ div(nχ∇c) = 0, t ≥ 0, x ∈ Rd,

−∆c+ τ c = n,

n(t = 0) = n0 ∈ L∞ ∩ L1+(Rd).

(6.10)

The first equation expresses the random (Brownian) diffusion of the cells with a bias directed by the

chemoattractant concentration with a sensitivity χ. The case χ = 0 means the cells do not react.

The chemoattractant c is directly released by the cell, diffuses on the substrate and is degraded with a

coefficient τ that scales in a such a way that τ−1/2 represents the activation length.

The notation L1+ means nonnegative integrable functions, and the parabolic equation for n gives non-

negative solutions (as expected for the cell density, see Chapter ??)

n(t, x) ≥ 0, c(t, x) ≥ 0. (6.11)

Another property we use is the conservation of the total number of cells

m0 :=

∫Rd

n0(x) dx =

∫Rd

n(t, x) dx. (6.12)

In particular, solutions cannot blow-up in L1. But we have the

Theorem 6.5 In R2, take τ = 0 and assume∫R2 |x|2n0(x)dx <∞.

(i) (Blow-up) When the initial mass satisfies

m0 :=

∫R2

n0(x)dx > mcrit := 8π/χ, (6.13)

then any solution of (6.10) becomes a singular measure in finite time.

(ii) When the initial data satisfies∫R2 n

0(x)| log(n0(x))|dx <∞ and

m0 :=

∫R2

n0(x)dx < mcrit := 8π/χ, (6.14)

Page 52: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

6.4. NON-EXTINCTION 47

there are weak solutions of (6.10) satisfying the a priori estimates∫R2

n[| ln(n(t))|+ |x|2] dx ≤ C(t),

‖n(t)‖Lp(R2) ≤ C(p, t, n0) for ‖n0‖Lp(R2) <∞, 1 < p <∞.

Here we only explain the argument for blow-up. We refer to [34, 7] for the complete proof of Theorem 6.5.

Proof. We follow Nagai’s argument based on the method of moments, assuming sufficient decay in x at

infinity. This is based on the formula

∇c(t, x) = −λ2

∫R2

x− y|x− y|2

n(t, y)dy, λ2 =1

2π.

Then, we consider the second x moment

m2(t) :=

∫R2

|x|2

2n(t, x)dx.

We have, from (6.10),

ddtm2(t) =

∫R2

|x|22 [∆n− div(nχ∇c)]dx

=∫R2 [2n+ χnx · ∇c]dx

= 2m0 − χλ2

∫R2×R2 n(t, x)n(t, y)x·(x−y)

|x−y|2

= 2m0 − χλ2

2

∫R2×R2 n(t, x)n(t, y) (x−y)·(x−y)

|x−y|2

(this last equality simply follows by a symmetry argument, interchanging x and y in the integral). This

yields finally,d

dtm2(t) = 2m0(1−

χ

8πm0).

Therefore, if we have m0 > 8π/χ, we arrive at the conclusion that m2(t) should become negative in finite

time, which is impossible since n is nonnegative. Consequently, the solution cannot be smooth until that

time.

6.4 Non-extinction

We consider now a nonlinearity with the negative sign, still with p > 1,∂∂tu−∆u = −up, t ≥ 0, x ∈ Rd,

u(t = 0, x) = u0(x) > 0.(6.15)

We recall that the solution remains positive u(t, x) > 0.

Page 53: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

48 Blow-up and extinction of solutions

As pointed out in the introduction, in the absence of diffusion, the solution (6.1) vanishes in infinite

time. Does diffusion change this effect?

To analyze this issue, we define the total mass

M(t) =

∫Rd

u(t, x)dx

which clearly decreasesd

dtM(t) = −

∫Rd

up(t, x)dx < 0. (6.16)

Following [24], we are going to prove non-extinction when diffusion is present

Theorem 6.6 Assume p > 1 + 2d and u0 ∈ L1 ∩ L∞(Rd), then the solution of (6.17) satisfies

limt→∞

M(t) > 0.

There is a nice biological interpretation in [24] related to the phenomena of so-called broadcast-spawning.

This is an external fertilization strategy used by various benthic invertebrates (sea urchins, anemones,

corals, jellyfish) whereby males and females release at the same time, short lived sperm and egg gametes

into the surrounding flow. The gametes are positively buoyant, and rise to the surface of the ocean. The

fertilized gametes form larva and are negatively buoyant and sink to the bottom of the ocean floor to start

a new colony. For the coral spawning problem, field measurements of the fertilization rates are often as

high as 90%. To arrive at such high rates, rapid dispersion on the ocean surface is required for successful

encounters and, as shown above, diffusion is clearly not sufficient, additional chemotactic attraction is

certainly involved.

The model at hand supposes that sperm and eggs have the same density (but this assumption can be re-

leased in a system of two parabolic equations, arriving at the same conclusion, see below). Chemotaxis is

omitted. The parameter p = 2 represents binary interactions leading to fertilized eggs that are withdrawn

from the balance equation. The theorem shows that, in the absence of chemotaxis, not all the eggs are

fertilized. In [24] it is proved that chemoattraction increases this rate but additional flow mixing does not.

Proof. Our assumption for p is equivalent to δ = (p− 1)d/2− 1 > 0.

First step. A lower bound. For all t > τ > 0, we can estimate, thanks to (??) with initial time τ ,∫Rd

up(t, x)dx ≤ C(d, p) M(τ)p (t− τ)−1−δ.

Therefore, we can write for T ≥ τ + a, and a > 0,

M(T ) = M(τ + a)−∫ T

τ+a

∫Rd

up(t, x)dxdt ≥M(τ + a)− C(d, p)M(τ)p∫ T

τ+a

(t− τ)−1−δdt,

and thus,

M(T ) ≥M(τ + a)− C(d, p)M(τ)p a−δ.

Page 54: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

6.4. NON-EXTINCTION 49

Second step. A upper bound. Let u be the solution to the linear equation∂∂t u−∆u = 0, t ≥ 0, x ∈ Rd,

u(t = 0, x) = u0(x) > 0.(6.17)

We have u ≤ u and thus, thanks to (??),

‖u(t)‖Lp(Rd) ≤ C(d, p)M(0)p t−1−δ.

Therefore, we find

M(τ + a) ≥M(τ)−∫ τ+a

τ

∫Rd

up(t, x)dxdt.

Because M(·) is non-increasing, we finally obtain

M(τ + a) ≥M(τ)− a‖u0‖p−1L∞(Rd)

M(τ).

Third step. Conclusion. Combining these two step, and choosing a with the rule a‖u0‖p−1L∞(Rd)

= 1/2, we

obtain

M(T ) ≥ 1

2M(τ)− C(d, p)M(τ)p a−δ.

If we had M(T )→ 0 as T →∞, we would conclude that

C(d, p)M(τ)p−1 a−δ ≥ 1

2,

which contradicts that M(τ) can vanish for large values of τ .

This proves the result.

Exercise. From the above argument derive an explicit bound for the minimum mass (depending on

M(0), p, d).

Exercise. Assume u0 ∈ L1 ∩ L∞(Rd) and consider the equation without diffusion

∂tu = −up, t ≥ 0, x ∈ Rd.

Show that∫Rd u(t, x)dx→ 0 as t→∞.

Exercise. Consider the system for eggs and other gametes (sperm)∂∂te− de∆e = −(eg)p/2, t ≥ 0, x ∈ Rd,∂∂tg − dg∆g = −(eg)p/2,

e(t = 0, x) = e0(x) > 0. g(t = 0, x) = g0(x) > 0

Show that both limt→∞∫Rd e(t, x)dx and limt→∞

∫Rd g(t, x)dx are positive.

See [24] again.

Page 55: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

50 Blow-up and extinction of solutions

Page 56: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

Chapter 7

Linear instability, Turing instability

and pattern formation

In his seminal paper1 A. Turing “suggests that a system of chemical substances reacting together and

diffusing through a tissue, is adequate to account for the main phenomena of morphogenesis”. He intro-

duces several concepts associated with the chemical basis of morphogenesis (and the name ‘morphogen’

itself), spatial chemical patterns, and what is now called ‘Diffusion Driven Instability’. The concept of

Turing instability has become standard and the aim of this chapter is to describe what it is (and what it

is not!).

The first numerical simulations of a system exhibiting Turing patterns was published in 1972 involving

the celebrated system of Gierer and Meinhardt [16] (see also Section 7.5.8).

It was only 20 years later that the first experimental evidence for a chemical reaction exhibiting spatial

patterns explained by these principles was obtained. This reaction is named the CIMA reaction after the

name of the reactants used by P. De Kepper et al2,3. See also Section 7.5.2.

In 1995, S. Kondo and R. Asai4 found an explanation of the patterns arising during the development

of animals, proposing that the model shoud be set in a growing domain, thus opening up a larger class of

possible patterns. Spots, which are a usual steady state for Turing systems in a fixed domain, can more

easily give way to bands in a growing domain.

Meanwhile, several nonlinear parabolic systems exhibiting Turing Patterns have been studied. Some

have been aimed at modeling particular examples of morphogenesis such as the models developed in

[29, 30]. Some have been derived as the simplest possible models exhibiting Turing instability conditions.

Nowadays, the biological interest in morphogenesis has evolved towards molecular cascades, pathways

and genetic networks. Some biologists doubt that, within cells or tissues, diffusion provides an adequate

description of molecular spreading. However, Turing’s mechanism remains both the simplest explanation

1A. M. Turing, The chemical basis of morphogenesis, Phil. Trans. Roy. Soc. London, B237, 37–7 (1952).2P. De Kepper, V. Castets, E. Dulos and J. Boissonade., Turing-type chemical patterns in the chlorite-iodide-malonic

acid reaction, Physica D 49 (1991), 161–1693V. Castets, E. Dulos, J. Boissonade and P. De Kepper, Experimental evidence of a sustained standing Turing-type

nonequilibrium chemical pattern. Phys. Rev. Letters 64(24) 2953–2956 (1990).4S. Kondo and R. Asai, A reaction-diffusion wave on the skin of the marine anglefish Pomocanthus, Nature (1995)

51

Page 57: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

52 Linear instability, Turing instability and pattern formation

for pattern formation, and one of the most counter-intuitive results in the field of Partial Differential

Equations.

This chapter presents Turing’s theory and several examples of diffusion driven instabilities. We begin

with the historical example of reaction-diffusion systems where the linear theory shows its exceptional

originality. Then we present nonlinear examples. The simplest of the latter is the non-local Fisher/KPP

equation, while some more standard parabolic systems are also presented.

7.1 Turing instability in linear reaction-diffusion systems

An amazingly counter-intuitive observation is the instability mechanism proposed by A. Turing [39].

Consider a linear 2× 2 O.D.E. system dudt = au+ bv,

dvdt = cu+ dv,

(7.1)

with real constant coefficients a, b, c and d. We assume that

T := a+ d < 0, D := ad− bc > 0. (7.2)

Consequently, we have

(u, v) = (0, 0) is a stable attractive point for the system (7.1). (7.3)

In other words, the matrix

A =

a b

c d

has two eigenvalues with negative real parts (or a single negative eigenvalue and a Jordan form). Indeed,

its characteristic polynomial is

(a−X)(d−X)− bc = X2 −XT +D,

and the two complex roots are X± = 12 [T ±

√T 2 − 4D].

Now consider a bounded domain Ω of Rd and add diffusion to the system (7.1),∂u∂t − σu∆u = au+ bv, x ∈ Ω,

∂v∂t − σv∆v = cu+ dv,

(7.4)

with either a Neumann or a Dirichlet boundary condition. In both cases, the state (u, v) = (0, 0) is still

a steady solution, of (7.4).

In principle, adding diffusion to the differential system (7.1) should increase stability. But surprisingly

we have

Page 58: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

7.1. TURING INSTABILITY IN LINEAR REACTION-DIFFUSION SYSTEMS 53

Theorem 7.1 (Turing instability theorem) Consider the system (7.4) where we fix the domain Ω

and the matrix A and σv > 0. We assume (7.2) with a > 0, d < 0. Then, for σu sufficiently small,

the steady state (u, v) = (0, 0) is linearly unstable. Moreover, only a finite number of eigenmodes are

unstable.

The usual interpretation of this result is as follows. Because a > 0 and d < 0, the quantity u is called

an activator and v an inhibitor. On the other hand, fixing a unit of time, the σ’s are scaled as the square

of length. The result can be extended as the

Turing instability alternative.

• Turing instability ⇐⇒ short range activator, long range inhibitor.

• Traveling waves ⇐⇒ long range activator, short range inhibitor.

There is no general proof of this statement, which can only be applied to nonlinear systems. But it is

a general observation that can be verified case by case. Also, the statement should be written with the

notion of ‘stable traveling waves’, because traveling waves can also connect an unstable state to a Turing

unstable state5. However, they are unstable in the sense that the dynamic creates periodic patterns.

Proof of Theorem 7.1. We consider the Laplace operator, with a Dirichlet or a Neumann condition

according to those considered for the system (7.4). It has an orthonormal basis of eigenfunctions (wk)k≥1

associated with positive eigenvalues λk,

−∆wk = λkwk.

We recall that we know that λk −−−−→k→∞ ∞. We use this basis to decompose u(t) and v(t), i.e.,

u(t) =

∞∑k=1

αk(t) wk, v(t) =

∞∑k=1

βk(t) wk.

We can project the system (7.4) on these eigenfunctions and arrive todαk(t)dt + σuλkαk(t) = aαk(t) + bβk(t),

dβk(t)dt + σvλkβk(t) = cαk(t) + dβk(t).

(7.5)

Now, we look for solutions with exponential growth in time, i.e., αk(t) = eλtαk, βk(t) = eλtβk with λ > 0

(in fact a complex number with Re(λ) > 0 is sufficient, but this does not change the conditions we find

below). The system is again reduced to λ αk + σuλkαk = aαk + bβk,

λ βk + σvλkβk = cαk + dβk.(7.6)

5G. Nadin, B. Perthame and M. Tang, Can a traveling wave connect two unstable states?The case of the nonlocal Fisher

equation. C. R. Acad. Sci. Paris, Ser. I 349 (2011) 553–557

Page 59: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

54 Linear instability, Turing instability and pattern formation

This is a 2 ∗ 2 linear system for αk, βk and it has a nonzero solution if, and only if, its determinant

vanishes

0 = det

λ+ σuλk − a − b

−c λ+ σvλk − d

Hence, there is a solution with exponential growth for those eigenvalues λk for which

there is a root λ > 0 to (λ+ σuλk − a)(λ+ σvλk − d)− bc = 0. (7.7)

This condition can be further reduced to the dispersion relation

λ2 + λ[(σu + σv)λk − T

]+ σu σv(λk)2 − λk(d σu + a σv) +D = 0.

Because the first order coefficient of this polynomial is positive, it can have a positive root if, and only

if, the zeroth order term is negative

σu σv(λk)2 − λk(d σu + a σv) +D < 0,

and we arrive to the final condition

(λk)2 − λk(d

σv+

a

σu) +

Dσuσv

< 0. (7.8)

Because λk > 0 and Dσuσv

> 0, this polynomial can take negative values only for dσv

+ aσu

> 0 and

sufficiently large, with Dσuσv

sufficiently small. It is difficult to give an accurate general characterization

in terms of σu and σvwhen (a, b, c, d) are fixed because we do not know, in general, the distribution of

the eigenvalues.

To go further, we set

θ =σuσv,

and we write explicitly the roots of the above polynomial and we require that

λk ∈ [Λ−,Λ+], Λ± =1

2σvθ

[dθ + a±

√(dθ + a

)2 − 4Dθ]. (7.9)

We can restrict ourselves to the regime θ small, then the Taylor expansion gives,

Λ± =dθ + a

2σvθ

[1±

√1− 4Dθ

(dθ + a)2

],

Λ± ≈a

2σvθ

[1±

[1− 2Dθ

(dθ + a)2

]],

and thus

Λ− ≈Da σv

= O(1), Λ+ ≈a

σvθ 1.

In the regime σu small, σv of the order of 1, the interval [Λ−,Λ+] becomes very large, hence we know

that some eigenvalues λk will belong to this interval.

Page 60: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

7.2. SPOTS ON THE BODY AND STRIPES ON THE TAIL 55

Note however that, because limk→∞ λk = +∞, there are only a finite number of unstable modes λk.

In principle one will observe the mode wk0 corresponding to the largest possible λ in (7.7) among the

λk’s that satisfy the condition (7.9). This does not correspond necessarily to the largest value of λk that

satisfy the inequality (7.8).

Exercise. Compute the first two terms in the expansion of λ in λk →∞.

Solution. λ ≈ −σuλk + T .

Exercise. Check how the condition (7.8) is generalized if we only impose the more general instability

criteria that (7.7) holds with λ ∈ C and Re(λ) > 0.

7.2 Spots on the body and stripes on the tail

Figure 7.1: Examples of animals with spots and stripes. Left: Discus (http://animal-

world.com/encyclo/fresh/cichlid/discus.php). Right: My cat.

Several papers6,7 (see also [32]) give a striking explanation of how Turing instability provides us with

a possible explanation of why so many animals have spots on the body and stripes on the tail, see Fig-

ure 7.1. In short, in a long and narrow domain (a tail), typical eigenfunctions are ’bands’, and with a

more spherically shaped domain (a mathematical square body), the eigenfunctions are ’spots’ or ’chess-

boards’. A much better and more detailed discussion with precise biological cases demonstrated, and

also the original papers where this idea is first expressed, can be found in [32] Vol. II Chapter 3.

To explain this, we consider the Neumann boundary condition and use our computations of eigenvalues

from Section 4.5. In one dimension, in a domain [0, L], the eigenvalues and eigenfunctions are

λk =

(πk

L

)2

, wk(x) = cos

(πkx

L

), k ∈ N.

6Oster, G.F.; Shubin, N.; Murray, J. D. and Alberch P. Evolution and morphogenetic rules: the shape of the vertebrate

limb in ontogeny and phylogeny Evolution 45, 862-884 (1988)7Maini, P. K. How the mouse got its stripes. PNAS 100(17): 9656–9657 (2003). DOI: 10.1073/pnas.1734061100

Page 61: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

56 Linear instability, Turing instability and pattern formation

In a rectangle [0, L1]× [0, L2], we obtain the eigenelements, for k, l ∈ N

λkl =

(πk

L1

)2

+

(πl

L2

)2

, wkl(x, y) = cos

(πkx

L1

)cos

(πly

L2

).

Consider a narrow stripe, say L2 ≈ 0 and L1 1. The condition (7.9), namely λkl ∈ [Λ−,Λ+], will

impose l = 0 otherwise λkl will be very large and cannot fit the interval [Λ−,Λ+]. The corresponding

eigenfunctions are bands parallel to the y axis.

When L2 ≈ L1, the distribution of sums of squared integers generically determines that the distribution

of the λkl ∈ [Λ−,Λ+] will be for l ≈ k.

To conclude this section, we point out that growing domains during development also very strongly

influence the topic of pattern formation. Again, we refer to [32] for a detailed analysis of the Turing

patterns, and their interpretation in development biology.

7.3 The simplest nonlinear example: the non-local Fisher/KPP

equation

As a simple nonlinear example to explain what is Turing instability, we consider the non-local Fisher/KPP

equation∂

∂tu− ν ∂

2

∂x2u = r u(1−K ∗ u), t ≥ 0, x ∈ R, (7.10)

still with ν > 0, r > 0, given parameters. For the convolution kernel K, we choose a smooth probability

density function

K(·) ≥ 0,

∫R

K(x)dx = 1, K ∈ L∞(R) (at least).

Compared to the Fisher/KPP equation, this takes into account that competition for resources can be of

long range (the size of the support of K) and not only local.

This idea has been proposed in ecology as an improvement of the Fisher equation that takes into

account long range competition for resources (N. F. Britton8,9). In semi-arid regions the roots of trees,

in competition for water, can cover, up to ten times the external size of the tree itself (while in temperate

regions the ratio is roughly one to one). This leads to the so-called ‘tiger bush’ landscape [26].

The same equation has also been proposed as a simple model of adaptive evolution . Then x repre-

sents a phenotypical trait, see S. Genieys et al 10. The Laplace term represents mutations and the right

hand side, growth and competition. The convolution kernel is used to express that competition is higher

between individuals, whose traits are close to each other (see also Section ??).

8Britton, N. F. Spatial structures and periodic traveling waves in an integro-differential reaction- diffusion population

model. SIAM J. Appl. Math. 50(6), 1663–1688 (1990)9Gourley, S. A. Travelling front solutions of a non-local Fisher equation. J. Math. Biol. 41, 272–284 (2000)

10Genieys, S., Volpert, V. and Auger, P., Pattern and waves for a model in population dynamic with non-local consumption

of resources. Math. Model. Nat. Phenom. 1 (2006), no. 1, 65–82

Page 62: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

7.3. THE SIMPLEST NONLINEAR EXAMPLE: THE NON-LOCAL FISHER/KPP EQUATION 57

The convolution term has a drastic effect for solutions; it can induce solutions that exhibit a behavior

quite different from those of the Fisher/KPP equation. The reason is mainly that the maximum principle

is lost with the non-local term. Again, we notice that the steady state u ≡ 0 is unstable, that u 1 is

also unstable because it induces a strong decay. In one dimension, for a general reaction function f(u) the

conditions read f(0) = 0, f ′(0) > 0 and f(u) < 0 for u large. Consequently there is a point u0 satisfying

(generically) f(u0) = 0, f ′(u0) < 0, i.e. a stable steady state should be found between the unstable ones.

This is the case of the nonlinearities arising in the Fisher/KPP equation that we have already encountered.

In the infinite dimensional framework at hand, we shall see that under certain circumstances, the steady

state u ≡ 1 can be unstable in the sense of

Definition 7.2 The steady state u ≡ 1 is called linearly unstable if there are perturbations such that the

linearized system has exponential growth in time.

Then, the following conditions are satisfied

Definition 7.3 A steady state u0 is said to form Turing patterns if

(i) there is no blow-up, no extinction (it is between two unstable states as above),

(ii) it is linearly unstable,

(iii) the corresponding growth modes are bounded (no high frequency oscillations).

Obviously when Turing instability occurs, solutions should exhibit strange behavior because they re-

main bounded away from the two extreme steady states, they cannot converge to the steady state u0 or

oscillate rapidly. In other words, they should exhibit Turing patterns. See Figure 7.2 for a numerical

solution of (7.10).

In practice, to check linear instability we use a spectral basis. In compact domains the concept can be

handled using eigenfunctions of the Laplace operator as we did in Section 7.1. In the full line, we may

use the generalized eigenfunctions, these are the Fourier modes. We define the Fourier transform as

u(ξ) =

∫Ru(x)e−ix ξdx.

Theorem 7.4 Assume the condition

∃ξ0 such that K(ξ0) < 0, (7.11)

then, for ν/r sufficiently small (depending on ξ0 and K(ξ0)), the non-local Fisher/KPP equation (7.10)

is nonlinearly Turing unstable.

A practical consequence of this theorem is that solutions should create Turing patterns as mentioned

earlier. This can easily be observed in numerical simulations, see Figure 7.2.

The non-local Fisher equation also gives an example of the already mentioned Turing instability

alternative.

• Turing instability ⇐⇒ K(·) is long range.

• Traveling waves ⇐⇒ K(·) is short range.

Page 63: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

58 Linear instability, Turing instability and pattern formation

Figure 7.2: Two steady state solutions of the non-local Fisher/KPP equation (7.10) in 2 dimensions with

different diffusion coefficients.

Indeed the non-local term K ∗u is the inhibitor (negative) term. The diffusion represents the activator

(with a coefficient normalized to 1). The limit of very short range is the case of K = δ, a Dirac mass,

and we recover the Fisher/KPP equation. More on this is proved in [4].

Proof. (i) The state u ≡ 0 and u ≡ ∞ are indeed formally both unstable (to prove this rigorously is not

so easy for u ≡ ∞.)

(ii) The linearized equation around u ≡ 1 is obtained by setting u = 1 + u and keeping the first order

terms, we obtain

∂tu− ν ∂

2

∂x2u = −r K ∗ u.

And we look for solutions of the form u(t, x) = eλtv(x) with λ > 0. This means that we should find

eigenfunctions associated with the positive eigenvalue λ, that means solution v(x) of

λv − ν ∂2

∂x2v = −r K ∗ v.

We look for a possible Fourier mode v(x) = eix ξ1 that we insert in the previous equation. Then we obtain

the condition

λ+ νξ21 = −r K(ξ1), for some λ > 0. (7.12)

And it is indeed possible to obtain such a λ and a ξ1 = ξ0 under the conditions of the Theorem.

(iii) The possible unstable modes ξ0 are obviously bounded because K is bounded as the Fourier transform

of a probability density (|K| ≤ 1).

Note however that the mode ξ1 we observe in practice is that with the highest growth rate λ.

7.4 Phase transition: what is NOT Turing instability

What happens if the third condition in Definition 7.3 does not hold? The system remains bounded away

from zero and infinity by condition (i) and it is unstable by condition (ii). But it might ’blow-up’ by high

frequency oscillations.

As an example of such an unstable system, which is not Turing unstable, we consider the phase

Page 64: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

7.5. GALLERY OF PARABOLIC SYSTEMS GIVING TURING PATTERNS 59

transition model also used in Section ??,∂u∂t −∆A(u) = 0, x ∈ Ω,

∂∂νu = 0 on ∂Ω,

(7.13)

with

A(u) = u (3− u)2. (7.14)

Because A′(u) = 3(3−u)(1−u) < 0, the equation (7.13) is a backward-parabolic equation in the interval

u ∈ (1, 3). We expect that linear instability occurs in this interval. We take u = 2 and set

u = 2 + u,

Inserting this in the above equation we find the linearized equation for u(t, x)∂u∂t −A

′(2)∆u = 0, x ∈ Ω,

∂∂ν u = 0 on ∂Ω.

We set γ = −A′(2) > 0 and we look for solutions u(t, x) = eλtw(x), which are unstable, i.e., λ > 0. These

are given by

λw + γ∆w = 0,

and thus they stem from the Neumann eigenvalue problem in Theorem 4.1. We have

λ = λiγ, w = wi.

We can see that all the eigenvalues of the Laplace operator generate possible unstable modes and thus

they can be of very high frequency. And we expect to see the mode corresponding to the largest λ, i.e., to

the largest λi which of course does not exist because λi −→i→∞

∞. In the space variable these correspond

to highly oscillatory eigenfunctions wi that we can observe numerically.

Figure 7.3 gives numerical solutions of (7.13)–(7.14) corresponding to two different grids; high frequency

solutions are obtained that depend on the grid. This defect explains why we require bounded unstable

modes in the definition of Turing instability. It also explains why in Section ?? we have introduced a

relaxation system.

More on the subject of phase transitions can be found in Chapter ?? on the Stefan problem.

7.5 Gallery of parabolic systems giving Turing patterns

Many examples of nonlinear parabolic systems exhibiting Turing instabilities (and patterns) have been

widely studied. This section gives several examples but it is in no way complete; the theories are far too

complicated to be presented here and their use in biology and other applications are far too numerous.

Page 65: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

60 Linear instability, Turing instability and pattern formation

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

Figure 7.3: Two numerical solutions of the phase transition system (7.13)–(7.14) for (Right) 80 grid points

and (Left) 150 grid points. The oscillation frequencies depend on the grid and these are not Turing patterns.

7.5.1 A cell polarity system

We take the following system from Morita and Ogawa11∂∂tu− σ1∆u = −f(u) + v, t ≥ 0, x ∈ Ω,

∂∂tv − σ2∆v = f(u)− v,∂u(t,x)∂ν = ∂v(t,x)

∂ν = 0 on ∂Ω.

(7.15)

We take a function f ∈ C2(R+;R+) that, for a given value uc > 0, satisfies

f(0) = 0, f ′(u) > 0 for u ∈ [0, uc[, −1 < f ′(u) < 0 for u > uc, f(+∞) = f∞ ≥ 0.

This system is mass conservative since we have

u(t, x) > 0, v(t, x) > 0,d

dt

∫Ω

[u(t, x) + v(t, x)]dx = 0.

These properties give us the non extinction/non blow-up (in L1) conditions.

We analyze Turing instability.

The differential system is stable. The corresponding differential system isddtU = −f(U) + V, t ≥ 0,

ddtV = f(U)− V,

U(0) > 0, V (0) > 0.

It satisfies U(t) + V (t) = U(0) + V (0) =: M0 > 0 and thus can be also written as

d

dtU = −f(U) +M0 − U(t) := G(U(t)).

11Morita Y. and Ogawa T. Stability and bifurcation of nonconstant solutions of a reaction-diffusion system with conser-

vation of mass, Nonlinearity 23 (2010) 1387–1411.

Page 66: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

7.5. GALLERY OF PARABOLIC SYSTEMS GIVING TURING PATTERNS 61

Therefore, it preserves the positive cone

U(t) > 0, V (t) > 0,

indeed, at the first point t0 where u(t0) = 0 we would have ddtU(t0) = M0 > 0, which is a contradiction

(same argument to handle the function V ).

Since G′(u) = −f ′(u) − 1 < 0, G(0) > 0 and G(+∞) = −∞, there is a single steady state (u, v)

characterized by G(u) = 0 that is equivalent to

u+ v = M0, v = f(u). (7.16)

Because G(U) > 0 for U ≤ u, G(U) < 0 for U ≥ u, it is very clear that (monotonically)

U(t) −→t→∞

u, V (t) −→t→∞

v.

Turing instability. Consider a steady state (7.16). We compute the differential matrix for the right

hand side (−f ′(u) 1

f ′(u) −1

)To fit the assumptions of Theorem 7.1, we have to check T := −f ′(u)−1 < 0, D := 0 (a degenerate case,

corresponding to mass conservation). The only possibility is that u is the activator, which imposes

f ′(u) < 0 ⇐⇒ u > uc.

Then, we find that unstable modes eλt(αwk, βwk) exist if there is a positive root to the polynomial

λ2 + λ[λk(σ1 + σ2) + f ′(u) + 1] + (σ1λk + f ′(u))(σ2λk + 1)− f ′(u) = 0.

As in Section 7.1, this is equivalent to

(σ1λk + f ′(u))(σ2λk + 1)− f ′(u) = σ1σ2λ2k + σ1λk + σ2f

′(u)λk < 0,

λk +1

σ2≤ |f

′(u)|σ1

.

This is clearly satisfied for some eigenvalues λk when σ1 is sufficiently small; this is the same result as in

Theorem 7.1.

7.5.2 The CIMA reaction

As mentioned earlier, the first experimental evidence of Turing instability was obtained with the CIMA

(chlorite-iodide-malonic acid) chemical reaction. It was modeled by I. Lengyel and I. R. Epstein12 who

proposed the system (with c = 1) ∂u

∂t− σu∆u = a− u− 4uv

1 + u2,

∂v

∂t− σv∆v = bc u− cuv

1 + u2.

(7.17)

12Modeling of Turing structure in the chlorite-iodide-malonic acid-starch reaction system, Science 251 (1991) 650–652.

Page 67: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

62 Linear instability, Turing instability and pattern formation

Figure 7.4: Labyrinth and spot patterns in the CIMA reaction (7.17). The solutions have been computed

with the software FreeFEM++ [14, 19] by S. Kaber.

Here u (the activator) denotes the iodide (I−) concentration and v (the inhibitor) the chlorite (ClO−2 )

concentration. Existence of steady states was analyzed 13,14.

Here we consider this system with a > 0, b > 0, c > 0. There is a single homogeneous steady state

u =a

4b+ 1, v = b(1 + u2).

The simple invariant region for solutions (that means the bounds are satisfied for all times if initially

true) are given by the maximum principle

0 ≤ u ≤ a, b ≤ v ≤ b(1 + a2) := vM .

Also solutions cannot become extinct because we can use the upper bound on v to go further and find

u ≥ umin, umin(1 +4vM

1 + u2min

) = a,

(in fact we can go even further and find a more restrictive invariant region, iterating the argument).

These a priori bounds are consequences of the maximum principle and we skip the derivation.

Therefore, according to our theory of Turing patterns, solutions cannot become extinct or blow-up and

it remains to study the linearized operator around the steady state (u, v).

Lemma 7.5 The CIMA reaction system (7.17) is Turing unstable if

4b > 1, c > 2√

16b2 − 1,

√4b+ 1

4b− 1< u <

c+√c2 − 4(16b2 − 1)

2(4b− 1),

and u is the activator, v the inhibitor.

Consequently, for this range of parameters, σv of the order of 1, and σu sufficiently small, there will

be Turing patterns.

13W. Ni and M. Tang. Turing patterns in the Lengyel-Epstein system for the CIMA reaction. Trans. Amer. Math. Soc

357 (2005) 3953–396914F. Yi, J. Wei and J. Shi. Diifusion-driven instability and bifurcation in the Lengyel-Epstein reaction-diffusion system.

Nonl. Anal. Real World Appl. 9 (2008) 1038–1051

Page 68: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

7.5. GALLERY OF PARABOLIC SYSTEMS GIVING TURING PATTERNS 63

With the first two conditions for b and c, becausec+√c2−4(16b2−1)

2(4b−1) >√

4b+14b−1 , the third set of inequalities

is a condition to ensure that a is not too small, neither too large.

Proof. We compute the differential matrix for the right hand side −1− 4v 1−u2

(1+u2)2 − 4u1+u2

bc− cv 1−u2

(1+u2)2 − cu1+u2

=

−1− 4b2

v (1− u2) − 4buv

bc− cb2

v (1− u2) − cbuv

Using Theorem 7.1, and because the bottom right coefficient is negative, we have to check the conditions

−1− 4b2

v(1− u2) > 0, T r = −1− 4b2

v(1− u2)− cbu

v< 0,

Det = (1 +4b2

v(1− u2))

cbu

v+

4bu

v(bc− cb2

v(1− u2)) =

cbu

v+

4cb2u

v> 0.

The condition on the determinant is always satisfied. The only constraints on a, b, c come from the

first line. Replacing v by its value, we need first

4b2u2 − 4b2 − b(1 + u2) > 0⇐⇒ u2 >4b+ 1

4b− 1.

Secondly, the trace condition gives

(4b− 1)u2 − cu− (4b+ 1) < 0 =⇒c−

√c2 − 4(16b2 − 1)

2(4b− 1)< u <

c+√c2 − 4(16b2 − 1)

2(4b− 1),

which is the announced condition.

A subtle calculation that is left to the reader shows that√

4b+14b−1 >

c−√c2−4(16b2−1)

2(4b−1) and thus the state-

ment is proved.

Then one can assert that for σu σv, Turing patterns will be produced (in accordance with Theo-

rem 7.1.

7.5.3 The diffusive Fisher/KPP system

The simplest system we have encountered so far is the diffusive Fisher/KPP system already mentioned

in Section ??. It reads ∂u

∂t− du∆u = g(u)v,

∂v

∂t− dv∆v = −g(u)v.

(7.18)

For du = dv, the solution v = 1− u reduces (7.18) to the Fisher equation (??) with f(u) = g(u)(1− u).

Therefore, this system still exhibits traveling waves. It has been proved that even for dv 6= du, the system

has monotonic traveling waves [5, 27, 28] for a power law nonlinearity g(u) = un.

From the Turing instability alternative in Section 7.1, we do not expect Turing patterns at this stage.

This is why interesting examples are always a little more elaborate.

Exercise. Consider the steady states (U = γ, V = 0), γ > 0 of (7.18) with g(u) = un.

1. Compute the linearized equation around this steady state.

2. In the whole space or a bounded domain with a Neumann boundary condition, show that this steady

state is always stable.

Page 69: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

64 Linear instability, Turing instability and pattern formation

7.5.4 The Brusselator

Prigogine and Lefever15,16 proposed an example now called the Brusselator. This is certainly the simplest

system exhibiting Turing patterns; see also Section 7.5.5 and Section 7.5.6.

Let A > 0, B > 0 be given positive real numbers. In a finite domain Ω, we consider the following 2 ? 2

system with a Neumann boundary condition∂u

∂t− du∆u = A− (B + 1)u+ u2v,

∂v

∂t− dv∆v = Bu− u2v.

(7.19)

We check below that there is a single positive steady state that exhibits the linear conditions for Turing

instability. Also the solution cannot vanish, thanks to the term A > 0. But we are not aware of a proof

that the solutions remain bounded for t large.

Figure 7.5: The Turing instability region of the brusselator in the (A,B) plane.

Exercise. 1. Check that the only homogeneous steady state is (U = A, V = BA ).

2. Write the linearized system around this steady state.

3. Check that for B < 1 + A2, this steady state is attractive for the associated differential equation

(du = dv = 0).

4. Let du > 0, dv > 0 and consider an eigenpair (λi, wi) of the Laplace equation with a Neumann

boundary condition. Write the condition for A, B, du, dv, λi for which there is an unstable mode λ > 0.

5. Show that for θ := dudv< 1 the interval [Λ−,Λ+] for λi is not empty.

6. Show that for θ small we have Λ− ≈ A2

dv, Λ+ ≈ B−1

2dvθ. Conclude that for dv fixed and du small, the

steady state becomes unstable.

15Prigogine, I. and Lefever, R. Symmetry breaking instabilities in dissipative systems. II. J. Chem. Phys. 48, 1695–1700

(1968).16Nicolis G. and Prigogine, I. Self-organization in non-equilibrium systems. Wiley Interscience, New-York (1977).

Page 70: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

7.5. GALLERY OF PARABOLIC SYSTEMS GIVING TURING PATTERNS 65

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

0

1

2

3

4

5

6

7

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

0

1

2

3

4

5

6

7

Figure 7.6: Two numerical solutions of the brusselator system (7.19) with A = B = 2 and 200 grid points.

The choice of diffusion coefficients are (Left) dv = 1. and du = 0.005 (Right) dv = 0.1 and du = 0.001. The

first component u exhibits one or two strong peak(s) upward while v exhibits one or two minimum(s) and has

been magnified by a factor 5.

Solution. 2. ∂u∂t − du∆u = (B − 1)u+A2v,

∂v∂t − dv∆v = −Bu−A2v.

3. det = A2, tr = B − 1−A2 and see Section 7.1 for the condition tr < 0, which leads to B < 1 +A2.

4. As in the general theory, we arrive to a second order polynomial for λ, which implies that the constant

term should be negative, leading to the condition

dudvλ2i + λi(A

2du − (B − 1)dv) +A2 < 0.

5. To have two positive roots, we need a negative slope at the origin, i.e., B > 1 + dudvA2 and also

(B − 1)dv −A2du > 2A√du dv ⇐⇒ B > 1 + 2

√θA+ θA2.

This is compatible with the condition of question 3 if, and only if, θ < 1.

6. We have

Λ± =1

2dvθ[B − 1−A2θ ±

√(B − 1−A2θ)2 − 4A2θ ].

The Taylor expansion for θ small reads

Λ± =1

2dvθ(B − 1−A2θ)

[1±

√1− 4A2θ

(B − 1−A2θ)2

]

Λ± ≈1

2dvθ(B − 1−A2θ)

[1± (1− 2A2θ

(B − 1−A2θ)2)

]and thus

Λ− ≈A2

dv, Λ+ ≈

B − 1

dvθ.

For dv fixed and θ small this interval will contain eigenvalues λi.

The two parabolic regions in (A,B) are drawn in Figure 7.5.

Page 71: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

66 Linear instability, Turing instability and pattern formation

7.5.5 The Gray-Scott system (2)

Gray and Scott17 introduced this system as a model of a chemical reaction between two constituants. It

only differs from the Brusselator (7.19) by a change in the constant coefficients of the reaction term∂u

∂t− du∆u = unv −Au,

∂v

∂t− dv∆v = −unv +B(1− v).

(7.20)

Again A > 0, B ≥ 0 are constants (input, degradation of consituants) and n is an integer, which

determines the number of molecules u which react with a single molecule v.

The system (7.20) has the advantage over the Brusselator (7.19), in that it satisfies the Turing Instability

Principle in Section 7.1. To explain this, we only consider the case

n = 2, B > 4A2. (7.21)

We first notice that there are three steady states to the Gray-Scott system ; the trivial one (U0 = 0, V0 = 1)

and non-vanishing ones (U±, V±), given by

UV = A, AU = B(1− V ).

We can eliminate V or U and find AU2 −BU +AB = 0 and BV 2 −BV +A2 = 0, this means that

U± =B ±

√B2 − 4A2B

2A, V± =

B ∓√B2 − 4A2B

2B. (7.22)

It is easy (but tedious) to see that the following results hold:

Lemma 7.6 With the assumption (7.21), the steady state (U−, V−) is linearly unstable, the state (U0 =

0, V0 = 1) is linearly stable and (U+, V+) is linearly stable under the additional condition (7.23) below.

Proof. The linearized systems read∂u∂t − du∆u = (2UV −A)u+ U2v,

∂v∂t − dv∆v = −2UV u− (U2 +B)v.

We begin with the trivial steady state (U0, V0). Together with the general analysis of Section 7.1, we

compute, for the trivial steady state, the quantities

D0 = AB > 0, T = −A−B < 0.

This means that for the steady state (U0, V0), the corresponding O.D.E. system is linearly attractive.

For the non-vanishing ones, using the above rule UV = A, we also compute∂u∂t − du∆u = Au+ U2v,

∂v∂t − dv∆v = −2Au− (U2 +B)v,

17Gray, P. and Scott S. K. Autocatalytic reactions in the isothermal continuous stirred tank reactor: isolas and other

forms of multistability. Chem. Eng. Sci. 38(1) 29–43 (1983)

Page 72: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

7.5. GALLERY OF PARABOLIC SYSTEMS GIVING TURING PATTERNS 67

D = A(U2 −B) > 0, T = A−B − U2.

We deduce from (7.22)

U2± −B = B2

2A2 − 2B ± B2A2

√B2 − 4A2B

= B2A2

[B − 4A2 ±

√B2 − 4A2B

].

Therefore, obviously D+ > 0 and

D− =B

2A2

√B2 − 4A2B

[√B − 4A2 −

√B]< 0.

Therefore, for the steady state (U−, V−), the corresponding O.D.E. system is unstable.

It remains to check the trace condition for (U+, V+). We have

T = A− B2

2A2− B

2A2

√B2 − 4A2B < 0, (7.23)

which is an additional condition to be checked. Note that, for the limiting value B = 4A2, this inequality

means that A > 1/8. Therefore, it is clearly compatible with (7.21).

We can now come back to the traveling wave solutions. We consider the particular case du = dv, A = B

and choosing v(t, x) = 1− u(t, x). Then, the two equations of (7.20) reduce to the single equation

∂u

∂t− du∆u = u2(1− u)−Au = u(u− u2 −A).

This is the situation of the Allen-Cahn (bistable) equation where, for A sufficiently small, we have three

steady states U0 = 0 is stable, U− is unstable and U+ > U− > 0 is stable. Therefore, we have indeed a

unique traveling wave solution (see Section ??). C. B. Mratov and V. V. Ospinov18 develop an extended

study of traveling waves in Gray-Scott system.

7.5.6 Schnakenberg system

The Schnakenberg system19 is still another variant of the Brusselator (7.19) and the Gray-Scott system

(7.20) with a constant input in the reaction terms. It is written as∂u

∂t− du∆u = C + u2v −Au,

∂v

∂t− dv∆v = B − u2v.

(7.24)

With C = 0, it has been advocated by T. Kolokolnikov et al20 as a simple model for spot-pattern

formation and spot-splitting in the following asymptotic regime∂u

∂t− ε∆u = u2v −Au,

ε∂v

∂t− dv∆v = B − u2v

ε.

(7.25)

The scaling has the advantage to show explicitly the connection with Theorem 7.1, u is the activator, v

the inhibitor in the linearized system around u = BAε , v = A2ε

B .

18Muratov, C. B. and Osipov, V. V. Traveling spike autosolitons in the Gray-Scott model. Physica D 155 (2001) 112–13119Schnakenberg J., Simple chemical reactions with limit cycle behavior. J. Theor. Biol. 81, 389–400 (1979)20Kolokolnikov, T.; Ward, M. J. and Wei, J. Spot Self-Replication and dynamic for the Schnakenburg Model in a Two-

Dimensional Domain, J. Nonlinear Science 19 No 1 (2009), 1–56

Page 73: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

68 Linear instability, Turing instability and pattern formation

7.5.7 The FitzHugh-Nagumo system with diffusion

Consider again the FitzHugh-Nagumo system as already studied in Section ??, but with diffusion for

both components, ∂u

∂t− du∆u = u(1− u)(u− 1

2)− v,

∂v

∂t− dv∆v = µu− v,

(7.26)

with a Neumann boundary conditions.

Assume that

µ > (1− u)(u− 1

2) ∀u ∈ R.

Then the only homogeneous steady state is (0, 0) and this is stable for the associated differential equation.

Exercise. We fix dv > 0. Show that when du is sufficiently small, the steady state is unstable.

As already mentioned, there are many variants, one of them is∂u

∂t− du∆u = u(1− u)(u− α)− βuv,

∂v

∂t− dv∆v = γuv − δv(1− v).

(7.27)

This relates prey-predator systems and FitzHugh-Nagumo systems. See Section 7.6.3.

7.5.8 The Gierer-Meinhardt system

The Gierer-Meinhardt [16, 29, 30] system is one of the most famous examples exhibiting Turing instability

and Turing patterns. It can be considered as a model for chemical reactions with only two reactants

denoted by u(t, x) and v(t, x).∂∂tu(t, x)− d1∆u(t, x) +Au = up/vq, t ≥ 0, x ∈ Ω,

∂∂tv(t, x)− d2∆v(t, x) +Bv = ur/vs,

(7.28)

with a Neumann boundary condition.

Among this class, several authors have used the particular case with a single parameter∂∂tu(t, x)− d1∆u(t, x) +Au = u2/v, t ≥ 0, x ∈ Ω,

∂∂tv(t, x)− d2∆v(t, x) + v = u2.

(7.29)

The diffusion coefficients satisfy

d1 1 d2.

A possible limit is d2 →∞, v → constant and thus we arrive to the reduced system for a steady state−ε2∆u+ u = up, x ∈ Ω,

∂u∂ν = 0 on ∂Ω.

(7.30)

Page 74: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

7.6. MODELS FROM ECOLOGY 69

Following Berestycki and Lions, when p < d+2d−2 , there is a unique radial spike like solution u = u0(xε )

with

−∆u0 + u0 = up0, x ∈ Rd, u0 > 0. (7.31)

But, see21, there are solutions concentrating on the boundary ∂Ω without limitation on p.

Also for nonlinear systems of elliptic equations, there is a large literature, see J. Wei et al22 and the

references therein. They show that there are numerous types of solutions, and that they may undergo

concentration properties in the case−ε2∆u+ u = u2

v , x ∈ R,−∆v + v = u2.

(7.32)

7.6 Models from ecology

There are many standard models from ecology. The simplest versions never satisfy the conditions for

Turing instability and we review some such examples first. Then we conclude with a more elaborate

model which satisfies the Turing conditions for instability.

7.6.1 Competing species and Turing instability

Let us come back to models of competing species as already mentioned in Section ??. Let the coefficients

r1, r2, α1 and α2 be positive in the system∂∂tu1 − d1∆u1 = r1u1(1− u1 − α2u2),

∂∂tu2 − d2∆u2 = r2u2(1− α1u1 − u2).

(7.33)

We have seen in Section ?? that the positive steady state (U1, U2) is stable if, and only if, α1 < 1,

α2 < 1. As stated in Theorem 7.1, to be Turing unstable, we require that one of the diagonal coefficients

is positive in the linearized matrix

L =

−r1U1 −α2r1U1

−α1r2U2 −r2U2

We see this is not the case. There is no activator in such systems.

7.6.2 Prey-predator system

In system (7.33), we can also consider a prey-predator situation where α1 < 0 (u1 is the prey) and

0 < α2 < 1 (u2 is the predator). With these conditions, the positive steady state is given by

(U1, U2) =

(1− α2

1− α2α1,

1− α1

1− α2α1

).

21A. Malchiodi and M. Montenegro, Multidimensional boundary layers for a singularly perturbed Neumann problem,

Duke Math. J., 124 (2004) 105–14322J. Wei, M. Winter. Symmetric and asymmetric multiple clusters in a reaction-diffusion system. NoDEA 14 (2007)

No 5-6, 787–823

Page 75: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

70 Linear instability, Turing instability and pattern formation

Because

tr(L) = −(r1U1 + r2U2) < 0, det(L) = r1r2U1U2(1− α1α2) > 0,

this steady state is stable.

Again, according to Theorem 7.1, it cannot be Turing unstable because both diagonal coefficients are

negative.

7.6.3 Prey-predator system with Turing instability (problem)

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.00

2

4

6

8

10

12

14

16

18

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.011.0

11.5

12.0

12.5

13.0

13.5

14.0

Figure 7.7: The Turing patterns generated by the prey-predator system (7.34) with α = 1.8, β = .5, γ = .08,

du = .001 and dv = .6. Left: component u, Right: component v. We have used 300 grid points for the

computational domain 0 ≤ x ≤ 1.

Consider the prey-predator system (and see Section 7.5.7 for variants)∂∂tu− du∆u = u(1 + u− γ u

2

2 − βv),

∂∂tv − dv∆v = v(1− v + αu).

(7.34)

The purpose of the problem is to show there are parameters α > 0, β > 0, γ > 0 for which Turing

instability occurs. We assume

β < 1, αβ < 1.

1. Show there is a unique homogeneous stationary state (u > 0, v > 0). Show that

γu > 2(1− αβ).

2. Compute the linearized matrix A of the differential system around this stationary state.

3. Compute the trace of A and show that Tr(A) < 0 if, and only if, the following condition is satisfied :

u[1− α− γu] < 1.

4. Compute the determinant of A and show that Det(A) > 0.

5. Which sign condition is also required for one of the coefficients of this matrix? How is it written in

terms of γu?

Page 76: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

7.6. MODELS FROM ECOLOGY 71

6. Suppose also that αβ > 12 , α > 1. Show that the above conditions are satisfied for γ sufficiently small.

7. For coefficients as in 6., state a Turing instability result concerning du, dv.

Figure 7.7 shows a one dimensional numerical simulation of the system (7.34) in the Turing instability

regime.

Note also that the system has a priori bounds that follow from the maximum principle. If initially

true, we have

γu2

2≤ 1 + u =⇒ u ≤ uM :=

1 +√

1− 2γ

γ,

v ≤ vM := 1 + uM .

Solution

1. The non-zero homogeneous steady state is given by v = αu+ 1 and

0 = γu2

2− u+ βv − 1 = γ

u2

2+ u (αβ − 1) + β − 1.

Since β < 1, its positive solution is given by

γu = 1− αβ +√

(1− αβ)2 + 2γ(1− β) > 2(1− αβ).

2. The linearized matrix about this steady state is

A =

u(1− γu) −βu

αv −v

3. We have Tr(A) = u(1− γu− α)− 1 and Tr(A) < 0 if, and only if, u[1− α− γu] < 1.

4. We have

Det(A) = −u(1− γu)v + αβuv = uv (αβ + γu− 1)

and using question 1

Det(A) > uv (1− αβ) > 0.

5. The last condition to have Turing instability is that one of the diagonal coefficients of A is positive.

It can only be the upper left coefficient and this is satisfied if γu < 1.

6. As γ → 0, we have γu → 2(1 − αβ) and the condition αβ < 1/2 is sufficient to ensure γu < 1 for

γ sufficiently small. We also have to ensure u[1 − α − γu] < 1 (from question 3) and this converges to

u[−1−α+2αβ]. in addition with α > 1 we have in fact [−1−α+2αβ] < 0, which guarantees the desired

condition.

7. With these conditions we know from Theorem 7.1 that in a bounded domain, for du sufficiently small

and dv of the order of 1, the steady state is linearly unstable.

Page 77: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

72 Linear instability, Turing instability and pattern formation

7.7 Keller-Segel with growth

Exercise. Consider the one dimensional Keller-Segel system with growth ut − uxx + χ(uvx)x = u(1− u), x ∈ R,

−dvxx + v = u.(7.35)

1. Show that u = 1, v = 1 is a steady state.

2. Linearize the system around this steady state (1, 1).

3. In the Fourier variable, reduce the system to a single equation for U(t, k),

Ut + U Λ(k) = 0,

and compute Λ(k).

4. Show that it is linearly stable under the condition χ ≤ (1 +√d)2.

Solution 1. This is easy.

2. Ut − Uxx + χVxx = −U , −dVxx + V = U .

3. Ut + k2U + k2χV = −U , −dk2V + V = U .

We can eliminate V and find Λ(k) = [k2 + 1− χ k2

1+dk2 ].

4. The linear stability condition is that all solutions have the time decay e−λt with λ > 0, in other words

Λ(k) > 0. Using the shorter notation X = k2 ≥ 0, it reads 1 +X(d+ 1− χ) + dX2 ≥ 0. The analysis of

the roots of this polynomial leads to 4.

Page 78: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

Chapter 8

Transport and advection equations

We have seen how to describe the diffusion of a population under the effect of random (Brownian) motion

of its individuals. Another description of motion is when individuals or particles follow an underlying

fluid that transport them deterministically. We speak of transport when one follows the positions (trajec-

tories) of individuals/particles and of advection for the resulting effect on density distribution. But let us

mention that the word convection is also used, mostly when the fluid motion is generated by heat transfer.

We consider a given velocity field v(x, t) ∈ Rd. We call the transport equation, that means the equation

in the strong form, which is to find u : Rd × R→ R such that∂∂tu(x, t) + v(x, t).∇u = 0, t ≥ 0, x ∈ Rd,

u(x, t = 0) = u0(x).(8.1)

We call the advection equation, the equation in divergence form, which is to find n : Rd ×R→ R such

that ∂∂tn(x, t) + div

(n(x, t)v(x, t)

)= 0, t ≥ 0, x ∈ Rd,

n(x, t = 0) = n0(x).(8.2)

One can build solutions to these equations using the notion of forward characteristics, that are the

solutions of the system of differential equationsX(t; y) = v(X(t; y), t),

X(t = 0; y) = y ∈ Rd.(8.3)

We recall that these trajectories exist, and that y 7→ X(t; y) is one-to-one on Rd, under the Cauchy-

Lipschitz assumptions,∀R > 0, T > 0, ∃M1(R, T ), M2(T ) (two constants), such that, ∀|t| ≤ T

|v(x, t)− v(y, t)| ≤M1(R, T )|x− y|, ∀x, y with |x| ≤ R, |y| ≤ R,

|v(x, t)| ≤M2(T )(1 + |x|) ∀x ∈ Rd.

(8.4)

Page 79: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

74 Transport and advection equations

8.1 Transport equation; method of caracteristics

The solution of equation (8.1) is given by a simple formula using the forward characteristics

Lemma 8.1 (Method of characteristics) For C1 data v and u0, the solution of (8.1) are C1 and

given by

u(X(t; y), t) = u0(y) ∀t ∈ R, ∀y ∈ Rd. (8.5)

In other words, solutions are constant along the characteristics.

Proof. For a C1 fields v and initial data u0, using the regularity theory in the Cauchy-Lipschitz theory,

one can writeddtu(X(t; y), t) = ∂

∂tu(X(t; y), t) + X(t; y).∇u(X(t; y), t)

= ∂∂tu(X(t; y), t) + v

(X(t; y), t

).∇u(X(t; y), t).

This derivative vanishes if, and only if, the transport equation (8.1) is satisfied.

To define u(x, t) at a given point x ∈ Rd, one needs to invert the mapping y 7→ X(t; y) := x. This is

easy and given by the backward characteristics departing at time t from the position x,dYds (s;x, t) = v(Y (s;x, t), s), s ∈ R,Y (s;x, t) = x ∈ Rd.

(8.6)

The inversion formula reads

x = X (t, Y (0;x, t)) , (8.7)

which yields the variant

u(x, t) = u0(Y (0;x, t)

)∀x ∈ Rd, t ∈ R. (8.8)

From the representation formula (8.8), and because x 7→ Y (0;x, t) is an homeomorphism, we conclude

that

Theorem 8.2 (Weak solutions of the transport equation) For u0 ∈ L∞(Rd), there is a unique

bounded distributional solution of (8.1) given by (8.8) for almost every (x, t).

Proof. For existence, consider a sequence of initial data u0n ∈ C1(Rd) such that ‖u0

n‖∞ ≤ ‖u0‖∞ + 1/n

and u0n −→n→∞

u0 almost everywhere. Then, we define un(x, t) = u0n

(Y (0;x, t)

)and, passing to the limit

n → ∞, we recover the formula (8.1). To see that this defines indeed a distributional solution, we just

pass to the limit in the definition of distributional solutions: for any test function ϕ ∈ C1c

([0, T ] × Rd

)with ϕ(·, T ) = 0, we have

−∫ T

0

∫Rd

un(x, t)[∂tϕ+ divϕ

]dxdt =

∫Rd

u0n(x)ϕ(x, t = 0)dx.

Therefore, using the Lebesgue convergence theorem, we have also

−∫ T

0

∫Rd

u(x, t)[∂tϕ+ divϕ

]dxdt =

∫Rd

u0(x)ϕ(x, t = 0)dx. (8.9)

Page 80: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

8.2. ADVECTION AND VOLUMES TRANSFORMATION 75

Uniqueness is proved by Hilbert’s duality method. Substracting two solutions, (8.9) holds with u0 ≡ 0.

Using the next section, it remains to build, for ψ ∈ C1c

((0,∞)× Rd

), a solution of

∂tϕ+ div(ϕv) = ψ, t ∈ (0, T ), x ∈ Rd, ϕ(·, T ) = 0.

Notice that the C1 theory uses the additional assumption divv ∈ C1 which is unnecessary relaxing the

assumption ϕ ∈ C1 to Lipschitz continuity.

We also conclude from the representation formula (8.8) several properties

• u0 ≥ 0 =⇒ u(x, t) ≥ 0,

• ‖u(t)‖∞ ≤ ‖u0‖∞,

• the solution is defined for all times t ∈ R, in particular we can choose the ‘initial data’ at any time

T ∈ R• the solution has the same regularity as its initial data (no regularizing effects).

Another extension to velocity fields v which are merely W 1,1 or BV can be developed thanks to the

DiPerna-Lions theory and its extensions, see [?, ?].

A geometric interpretation of the formula in Lemma 8.1 is to take A0 a measurable subset of Rd.Then, we consider the transported set

A(t) := X(t, y); y ∈ A0. (8.10)

The set A(t) can be determined as the (weak) solution ot equation (8.1)

u(x, t) = 1A(t), for u0 = 1A0.

Exercise. For a ∈ C1, give the formulaalong the charateristics to solve the equation∂∂tu(x, t) + v(x, t).∇u+ a(x, t)u = 0, t ≥ 0, x ∈ Rd,

u(x, t = 0) = u0(x) ∈ C1.(8.11)

Exercise. For f ∈ C1, give the formulaalong the charateristics to solve the equation∂∂tu(x, t) + v(x, t).∇u+ a(x, t)u = f(x, t), t ≥ 0, x ∈ Rd,

u(x, t = 0) = u0(x) ∈ C1.(8.12)

8.2 Advection and volumes transformation

The geometry of transport by the differential system (8.3) yields another question. How are densities

evolved? In other words, what is the volume of the set A(t) compared to that of A0 in (8.10). This is

Page 81: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

76 Transport and advection equations

what equation (8.2) tells us.

The general formula for the solution of equation (8.2) is not as simple than for transport, and can be

derived from the expression

∂tn(x, t) + v(x, t).∇n(x, t) + n(x, t)div v = 0.

Using the proof of Lemma 8.1 one finds, along the characteristics and using the chain rule, that

d

dtn(X(t; y), t) + (div v)(t,X(t; y))n(X(t; y), t) = 0.

and thus

Proposition 8.3 The solution of equation (8.2) is

n(X(t, y), t)

[exp

∫ t

0

div v (X(s; y), s)ds

]= n0(y) ∀t ≥ 0, ∀y ∈ Rd. (8.13)

To give an interpretation of this expression, we recall some facts on the differential system (8.3). We

consider the d× d matrix Dxv(X(t; y), t) and the volume transformation

J(y, t) = det

(∂X(t; y)

∂y

). (8.14)

They satisfy respectively the equations

d

dt

∂X(t; y)

∂y= Dxv(t,X(t; y)).

∂X(t; y)

∂y,

and (this is not an obvious calculation)ddtJ(y, t) = div v(t,X(t; y)) J(y, t),

J(0, y) = 1.(8.15)

Therefore, we have

J(y, t) = e∫ t0

(div v)(s,X(s,y)) ds = det(∂X(t; y)

∂y

).

We can summarize these results and write

Theorem 8.4 (Strong solutions) With the Cauchy-Lipschitz assumptions (8.4) and with v ∈ C1(R×Rd), div v ∈ C1(R×Rd), and n0 ∈ C1, there is a unique solution n ∈ C1(Rd ×R) of equation (8.2) given

by the formula

n(X(t; y), t)J(y, t) = n0(y). (8.16)

It satisfies ∫Rd

n(x, t)dx =

∫Rd

n0(x)dx,

∫Rd

|n(x, t)|dx =

∫Rd

|n0(x)|dx. (8.17)

Page 82: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

8.2. ADVECTION AND VOLUMES TRANSFORMATION 77

The absolute value (and |n|+ and other related variants) are the only nonlinearities which are preservec

by the flow in general. Compressible flows (see below) are the exception with this respect.

Exercise. Using the solution of (8.2) along the characteritics and mass conservation, prove that I(y, t) =

exp∫ t

0div v (X(s; y), s)ds and J(y, t) defined by (8.14) are equal.

[Hint]. Use that n(x, t) can be used as an initial data and thus all reasonable functions can be attained

at time t.

As before, the theory can be extended to distributional solutions and the natural space is L1

Theorem 8.5 (Weak solutions) With the assumptions of Theorem 8.4 for v and for n0 ∈ L1(Rd),there is a unique distributional solution n ∈ C(R;L1(Rd)) of equation (8.2) and it satisfies the equali-

ties (8.17).

A consequence of the theorem is that equation (8.2) transport the densities; indeed coming back to the

transport of sets in (8.10), we have

Vol(A(t)) =∫n(x, t)dx =

∫n(X(t; y))det

(∂X(t;y)∂y

)dy

=∫n0(y)dy = Vol(A0).

In other words, the transport equation transports the values but transforms the volumes. The advection

equation rearrange the values so as to preserv the volumes (and thus the densities).

We conclude from the above the properties

• n0 ≥ 0 =⇒ n(x, t) ≥ 0,

• ‖n(t)‖1 = ‖n0‖1,

• the solution is defined for all times t ∈ R, in particular we can choose the ‘initial data’ at any time

T ∈ R,

• the solution has the same regularity as its initial data (no regularizing effects).

A flow that preserves volume is a flow for which J(y, t) ≡ 1 in (8.15), this called an incompressible flow

Definition 8.6 A flow v is incompressible if

div v(x, t) = 0, ∀x ∈ Rd, t ∈ R.

That also means J(y, t) ≡ 1.

The role of compressibility can be seen when computing nonlinear quantities S(n). We have

∂tS(n) + div

(S(n) v

)+ [nS′(n)− S(n)]divv = 0,

In particular, incompressible flows will preserv any Lp norm

d

dt

∫Rd

np(x, t)dx = 0.

Page 83: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

78 Transport and advection equations

The absolute value plays a special role among the choices of nonlinearities because for S(n) = |n| one

has S′(n) = sgn(n) and nS′(n) − S(n) = 0. Therefore, in distribution sense, for solutions of (8.2), we

also have∂

∂t|n(x, t)|+ div(|n(x, t)|v(x, t)) = 0

Exercise. For f ∈ C1, give the formula which solves the equation∂∂tn(x, t) + div(nv(x, t)) = f, t ≥ 0, x ∈ Rd,

u(x, t = 0) = u0(x) ∈ C1.(8.18)

Exercise. Determine the volume of the set A(t) defined by (8.10).

8.3 Advection (Dirac masses)

Another point of view gives the same conclusion that mass conservation is a fundamental property of the

advection equation. We have a simple representation formula for the solution is the

Lemma 8.7 For n0(x) :=

K∑k=1

ρ0kδ(x− y0

k), the solution of (8.2) is given by the formula

n(x, t) =

K∑k=1

ρ0kδ(x−X(t; y0

k)).

More generally, we may represent the initial data as n0(x) =∫Rd n

0(y)δ(x − y) (replacing the finite

sum by an integral) and we find

n(x, t) =

∫Rd

n0(y)δ(x−X(t; y)

)dy. (8.19)

For this reason we recover that the population density is transported by the flow field v according to

the equation (8.2).

Proof of Lemma 8.7. Because this is a linear equation, we only have to prove the formula for one

Dirac mass and the weight ρ0 = 1. Then, the definition of a weak solution means that for all smooth test

functions u(x, t) we have, for all T > 0,∫Rd

n(x, T )u(x, T )dx−∫ T

0

∫Rd

n(x, t)[ ∂∂tu(x, t) + v(x, t).∇u(x, t)

]=

∫Rd

n0(x)u(0, x).

We choose for u, a solution of the transport equation (8.1) and find∫Rd

n(x, T )u(x, T ) dx = u(y0, 0) = u(X(T ; y0), T )

by the methof of characteristics. see Lemma 8.1. Because for the transport equation we may choose the

initial data at time T , this equality holds true for all C1 functions v(·, T ), which means that n(x, T ) is a

Dirac mass as announced.

Page 84: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

8.4. EXAMPLES AND RELATED EQUATIONS 79

8.4 Examples and related equations

8.4.1 Long time behaviour

Theorem 8.8 Assume that there is α > 0 such that

(x− y).(v(x, t)− v(y, t)

)≤ −α|x− y|2, ∀x, y ∈ Rd, t ≥ 0, (8.20)

then, there is a position X(t) such that, whatever is the initial data n0, as t→∞,

n(x, t) ≈ ρδ(x−X(t)).

Proof. We recall that n(x, t) = ρδ(x−X(t)) is a solution for the initial data n0(x) = ρδ(x−X(0)).

Consider two solutions with different initial data n01(x), n0

2(x) ≥ 0 (this is not a restriction because of

one can always consider separately the positive and negative parts),

∂tn1(x, t)n2(y, t) + divx

(v(x, t)n1(x, t)n2(y, t)

)+ divy

(v(y, t)n1(x, t)n2(y, t)

)= 0, t ≥ 0, x, y ∈ Rd,

Therefore

d

dt

∫|x− y|2n1(x, t)n2(y, t)dxdy = 2

∫(x− y).

(v(x, t)− v(y, t)

)n1(x, t)n2(y, t)dxdy

≤ −2α

∫|x− y|2n1(x, t)n2(y, t)dxdy.

We find that ∫|x− y|2n1(x, t)n2(y, t)dxdy ≤ C0e−2αt.

This means that both ni concentrate at Dirac masses with the same location ρi(t)δ(x−X(t)). But the

mass conservation tells us that ρi(t) is a known constant.

Exercise. For v ∈ C1, show that condition (8.20) implies that the symmetric matrixDSv =(∂vi∂xj

+∂vj∂xi

)di,j=1

satisfies: DSv ≤ −2αI.

Exercise. For v ∈ C1 and div v(x, t) ≥ α > 0 show that, for all p > 1

‖n(t)‖p → 0 as t→∞.

8.4.2 Nonlinear advection

Several models in physics and biology lead to consider a the nonlinear drift. Examples are

• n(x, t) denotes the number of cells of size x and v(x, t) is the growth rate of cells which might depend

on environmental conditions (nutrients) which are changed by all the cells, whatever is their size,

• n(x, t) denotes the number of polymers of length x which can increase or decrease by addition of

monomers, with a rate v which may depend on the number of polymers.

Page 85: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

80 Transport and advection equations

We consider here a formalism which uses a given weight ψ ∈ C1(Rr;R+) and measures the growth rate

with

v(x, t) = V (x, I(t)), I(t) =

∫Rd

ψ(x)n(x, t)dx.

Here, we make the assumption that there is α > 0 such that

(x− y).(V (x, I)− V (y, I)

)≤ −α|x− y|2, ∀x, y ∈ Rd, I ≥ 0.

Theorem 8.9 Assume that∫n0 := M0 and

M0‖DIv‖∞‖Dψ‖∞ < α.

Then, there is a position X(t), independant of the initial data, such that, as t→∞,

n(x, t) ≈ ρδ(x−X(t)).

Proof. For two different initial data n0i , i = 1, 2 we obtain two functions Ii(t). We get (see before)

1

2

d

dt

∫|x− y|2n1(x, t)n2(y, t)dxdy =

∫(x− y).

(v(x, I1(t))− v(y, I2(t))

)n1(x, t)n2(y, t)dxdy

=

∫(x− y).

(v(x, I1(t))− v(y, I1(t))

)n1(x, t)n2(y, t)dxdy

+

∫(x− y).

(v(y, I1(t)))− v(y, I2(t))

)n1(x, t)n2(y, t)dxdy

≤ −α∫|x− y|2n1(x, t)n2(y, t)dxdy +M0‖DIv‖∞|I1(t)− I2(t)|

(∫|x− y|2n1(x, t)n2(y, t)dxdy

)1/2

.

But we have

M0(I1(t)− I2(t)) =

∫[ψ(x)− ψ(y)]n1(x, t)n2(y, t)dxdy

∣∣I1(t)− I2(t)∣∣ ≤ ‖Dψ‖∞

M0

∫|x− y|n1(x, t)n2(y, t)dxdy ≤ ‖Dψ‖∞

(∫|x− y|2n1(x, t)n2(y, t)dxdy

)1/2

.

Therefore, we control the above quantity as

12ddt

∫|x− y|2 n1(x, t)n2(y, t)dxdy

≤∫|x− y|2n1(x, t)n2(y, t)dxdy

[− α+M0‖DIv‖∞‖Dψ‖∞

].

And we conclude as before.

8.4.3 The renewal equation

In many different areas of biology, the age structured equation is used widely. It describes a population

of individuals who have age a at time t. The equation is∂∂tn(a, t) + ∂

∂an(a, t) + d(a)n(a, t) = 0, a ≥ 0, t ≥ 0,

n(a = 0, t) =∫∞

0b(a)n(a, t)da,

Page 86: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

8.4. EXAMPLES AND RELATED EQUATIONS 81

where d and b denote respectively the age dependent death and birth rates.

This equation expresses that age and time evolve with the same speed.

When a characteristic enters the domain where the equation is stated, which is the case here at a = 0,

a boundary condition is needed. Here it is used to express that new born are borned at age a = 0.

For a theory, see [?].

8.4.4 The Fokker-Planck equation

8.4.5 Other stochastic aspects

Page 87: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

82 Transport and advection equations

Page 88: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

Bibliography

[1] Baker, R. E.; Gaffney, E. A. and Maini, P. K. Partial differential equations for self-organization in

cellular and developmental biology. Nonlinearity 21 (2008) R251–R290.

[2] Barles, G. Solutions de viscosite des equations de Hamilton-Jacobi. Springer-Verlag Berlin Heidel-

berg, 1994.

[3] Berestycki, H. and Hamel, F. Reaction-Diffusion Equations and Propagation Phenomena. Book in

preparation.

[4] Berestycki, H.; Nadin, G.; Perthame, B. and Ryzhik, L. The non-local Fisher-KPP equation: trav-

eling waves and steady states. Nonlinearity (22), 2813–2844, 2009.

[5] Berestycki, H.; Nicolaenko, B. and Scheurer, B. Traveling wave solutions to combustion models and

their singular limits. SIAM J. Math. Anal. 16(6) (1985) 1207–1242.

[6] Brezis, H. Analyse fonctionnelle, theorie et applications. Masson, 1983.

[7] Blanchet, A.; Dolbeault, J. and Perthame, B. Two-dimensional Keller-Segel model: Optimal critical

mass and qualitative properties of the solutions. Electron. J. Diff. Eqns., Vol. 2006(2006), No. 44,

1–32.

[8] Calvez, V. and Carrillo, J. A. Volume effects in the Keller-Segel model: energy estimates preventing

blow-up. J. Math. Pures et Appl. (2006) 862–175.

[9] Cantrell, R. S. and Cosner, C. Spatial Ecology via Reaction–Diffusion Equations, John Wiley & Sons,

Sussex, 2003.

[10] Evans, L. C. Partial Differential Equations, Graduate Studies in Mathematics, Vol. 19, American

Mathematical Society (1998).

[11] Fife, P. C. Mathematical aspects of reacting and diffusing systems, Lecture Notes in biomathematics

28, Springer Verlag, 1979.

[12] Fisher, R. A. The genetical theory of natural selection. Clarendon Press, 1930. Deuxieme ed. : Dover,

1958. Troisieme edition, presentee et annotee par Henry Bennett : Oxford Univ. Press, 1999.

[13] Francoise, J.-P. Oscillations en biologie. Coll. Math. et Appl., SMAI, Springer, Paris (2005).

83

Page 89: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

84 BIBLIOGRAPHY

[14] FreeFEM++, Software available at http://www.freefem.org

[15] Gilbarg, D. and Trudinger, N. Elliptic partial differential equations of second order, second edition.

Springer (2003).

[16] Gierer, A. and Meinhardt, H. A theory of biological pattern formation. Kybernetik 12 (1972) 30–39.

[17] Godlewski E. and Raviart, P.-A. Numerical Approximation of Hyperbolic Systems of Conservation

Laws, Appl. Math. Series 118. Springer-Verlag (1996).

[18] Golding, I.; Kozlovski, Y.; Cohen, I. and BenJacob, E. Studies of bacterial branching growth using

reaction-diffusion models for colonial development. Physica A, 260:510–554 (1998).

[19] F. Hecht. New development in freefem++. J. Numer. Math., 20(3-4):251–265, 2012.

[20] Hillen, T. and Painter, K. A User’s Guide to PDE Models for Chemotaxis. J. Math. Biol., 58(1),

183–217, 2009.

[21] Jost, J. Mathematical Methods in Biology and Neurobiology. Available on author’s homepage.

[22] Keener, J. and Sneyd, J. Mathematical Physiology. Springer (2008).

[23] Keller, E. F. and Segel, L. A. Model for chemotaxis. J. Theor. Biol., 30:225–234, 1971.

[24] Kiselev, A. and Ryzhik, L. Biomixing by chemotaxis and enhancement of biological reactions. Comm.

in PDEs 37 (2012) 298–318.

[25] Kolmogorov, A. N.; Petrovski, I. G. and Piskunov, N. S. Etude de l’equation de la diffusion avec

croissance de la quantite de matiere et son application a un probleme biologique. Bulletin Universite

d’Etat de Moscou (Bjul. Moskowskogo Gos. Univ.), Serie internationale A1 (1937), 1–26.

[26] Lefever, R. and Lejeune, O. On the origin of tiger bush. Bull. Math. Biology. 59(2) (1997) 263–294.

[27] Logak, E. and Loubeau, V. Travelling wave solutions to a condensed phase combustion model.

Asymptotic Anal. 12 (1996), no. 4, 259–294.

[28] Marion, M. Qualitative properties of a nonlinear system for laminar flames without ignition temper-

ature. Nonlinear Anal. 9 (1985), no. 11, 1269–1292.

[29] Meinhardt, H. Models of biological pattern formation Academic Press, London (1982).

[30] Meinhardt, H. The algorithmic beauty of sea shells Springer-Verlag, Heidelberg, New-York (1995).

[31] Mimura, M.; Sakaguchi, H. and Matsuchita M. Reaction diffusion modeling of bacterial colony

patterns. Physica A, 282:283–303 (2000).

[32] Murray, J. D. Mathematical biology, Vol. 1 and 2, Second edition. Springer (2002).

[33] Okubo, A. and Levin, S. A. (Eds) Diffusion and ecological problems, modern perspectives, 2nd edition.

Springer-Verlag, New-York (2001).

Page 90: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

BIBLIOGRAPHY 85

[34] Perthame, B. Transport equations arising in biology. L. N. series ’Frontiers in mathematics’,

Birkhauser (2007).

[35] Quittner P. and Souplet P. Superlinear parabolic problems. Blow-up, global existence and steady state.

Birkhauser Advanced texts, 2007.

[36] Schwartz, L. Methodes mathematiques pour les sciences physiques. Hermann, 1983.

[37] Stevens, A. Derivation of chemotaxis-equations as limit dynamic of moderately interacting stochastic

many particle systems. SIAM J. Appl. Math. 61 (2000) 183–212.

[38] Thieme, H. R. Mathematics in population biology. Woodstock Princeton university press. Princeton

(2003).

[39] Turing, A. M., The chemical basis of morphogenesis. Phil. Trans. Roy. Soc. B, 237 (1952), 37–72.

[40] Vazquez, J. L. The porous medium equation. Mathematical theory. Oxford Mathematical Mono-

graphs. The Clarendon Press, Oxford University Press, Oxford, 2007. ISBN: 978-0-19-856903-9.

[41] Volpert, A. I.; Volpert, V. A. and Volpert V. A. Traveling wave solutions of parabolic systems.

Translation of Mathematical Monographs, Vol. 140, AMS, Providence, 1994.

Page 91: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

Index

Activator, 51, 56, 59, 60, 65, 67

Adaptive evolution, 5, 54

Allen-Cahn equation, 65

Asymptotic (analysis, limits, regime), 12, 65

Asymptotic stability, 29–31

Belousov-Zhabotinskii reaction, 19, 20

Blow-up, 2, 39, 40, 42–45, 55, 56, 58, 60

Brownian motion, 1, 21, 26, 28

Brusselator, 62–65

Cauchy-Lipschitz (theorem), 36, 39

Cell polarity, 58

Characteristics, 72

Chemical master equations, 7, 12

Chemostat, 34

Chemotaxis, 39, 44, 46

CIMA reaction, 19, 49, 59, 60

Competing species, 67

Competition systems, 3, 4

Connectance, 2

Conservation of atoms, 8

Cooperative species, 3

Cooperative systems, 3

Diffusion driven instability, 49, 50

Diffusive Fisher/KPP system, 61

Dispersion relation, 52

Distributions (weak solutions), 27, 40, 42

Ecological networks, 2

Entropy, 1, 2, 7–12, 19, 20, 29, 32–36

Enzymatic reactions, 13, 16, 17, 19

Epidemiology, 34

Fisher/KPP equation, 50, 54–56

FitzHugh-Nagumo, 66

FreeFEM++, 60

Gierer-Meinhardt system, 66

Gradient flow, 43

Gray-Scott system, 64, 65

Growing domains, 54

Hamiltonian (system), 36

Heat equation, 8, 26–29, 40

Hemoglobin, 11

Holing II, 2

Inhibitor, 51, 56, 60, 65

Initial layer, 13

Instability, 29, 49, 50, 53–55, 66

Intrinsic growth rate, 2

Invariant region, 60

Irreversible reaction, 8, 16, 18, 20

Isomerization, 10, 16

Keller-Segel system, 39, 44, 70

Kolmogorov equation, 28

Krein-Rutman theorem, 23, 37

Laplace operator, 22, 24, 30, 31, 51, 55, 57

Law of mass action, 7–9, 11, 16, 18

Lotka-Volterra, 1–3, 5, 8, 35–37

Lyapunov, 34, 35

Mass action, 8

Maximum principle, 23, 34, 55, 60, 69

Michaelis-Menten law, 16, 18

Mixed boundary condition, 22

Moments (method of), 44, 45

Monolix, 5

86

Page 92: Mathematical methods in biology Master of Mathematics, second … · 2018-10-26 · Mathematical methods in biology Master of Mathematics, second year Beno^ t Perthame Sorbonne Universit

INDEX 87

Monostable, 13, 14

Monotonicity (principle), 1, 3

Morphogenesis, 49

Mutualistic, 2

Non-extinction, 45, 46, 58

Non-local Fisher/KPP equation, 50, 54–56

Perturbation, 29, 31, 55

Phase transition, 56–58

Pierre’s duality estimate, 8

Poincare inequality, 30, 31

Poincare-Wirtinger inequality, 31, 32

Prey-predator system, 2, 35, 66–68

Reaction kinetics, 7, 18, 32

Reaction rate equations, 7, 9, 12

Reaction-diffusion (equation, system), 1, 8, 10, 12,

21, 50, 58

Relaxation, 29, 31, 32, 35

Rellich theorem, 23, 24

Reversible reaction, 9, 10, 12, 16

Robin boundary condition, 22

Scattering equation, 28

Schnakenberg system, 65

Sensitivity (coefficient, matrix), 44

SI system, 34

Slow-fast dynamic, 18, 19

Slow-fast dynamics, 13, 15

Spectral (analysis, basis, decomposition), 22–25, 55

Spike, 67

Spot, 53, 60

Stampacchia (method of), 3, 34

Stefan problem, 57

Symbiosis, 2

Tiger bush, 54

Transport equation, 72

Traveling waves, 51, 55, 61, 65

Trophic, 2

Turing (instability, pattern,...), 29, 49–51, 53–62,

64, 66–69

Turing instability alternative, 51, 55, 61

Turing instability theorem, 51

Variability, 5

Weak solutions, 2, 3, 8, 44, 76

well-prepared, 15