• Keine Ergebnisse gefunden

Non-symmetric Hypercontractivity and the p-Entropy

In this section we will show how to deduce the rate of convergence to equilibrium for the family ofp-entropies, with 1<p<2, frome2. The main thing that will make the above possible isa non-symmetric hypercontractivityproperty of our Fokker–Planck equation

- namely, that any solution to the equation with (initially only) a finitep-entropy will eventually be “pushed” intoL2¡

Rd,f−1¢

, at which point we can use the information we gained one2.

Before we show this result, and see how it implies our main theorem, we explain why and how this non-symmetric hypercontractivity helps.

Lemma 1.4.1. Let fL1+¡ Rd¢

with unit mass. Then (i)

ep(f|f)= 1 p(p−1)

à kfkp

Lp³

Rd,f1p´−1

! . (ii) for any1<p1<p2≤2there exists a constant Cp1,p2>0such that

ep1(f|f)≤Cp1,p2ep2(f|f).

In particular, for any1<p<2

ep(f|f)≤Cpe2(f|f), for a fixed geometric constant.

Proof. (i) is trivial. To prove (i i) we consider the function g(y) :=

(p2(p21)

p1(p1−1)

yp1p1(y1)1

yp2p2(y−1)−1, y≥0,y6=1

1 , y=1.

Clearly g ≥ 0 on R+, and it is easy to check that it is continuous. Since we have limy→∞g(y)=0, we can conclude the result using (1.1.4).

It is worth to note that the second point of part (i i) of Lemma 1.4.1 can be extended to general generating function for an admissible relative entropy. The following is taken from [3]:

Lemma 1.4.2. Letψbe a generating function for an admissible relative entropy. Then one has that

ψ(y)≤2ψ00(1)ψ2(y), y≥0.

In particular ep≤2e2for any1<p<2whenever e2is finite.

Lemma 1.4.1 assures us that, if we start with initial data inL2¡

Rd,f1¢

, thenepwill be finite. Moreover, due to Theorem 1.1.4 forp=2, and the fact that the solution to (1.1.2) remains inL2¡

Rd,f−1¢

, we have that

ep(f(t)|f)≤2e2(f(t)|f)≤C e2(f0|f

1+t2n¢ e−2µt.

However, one can easily find initial data f06∈L2¡

Rd,f1¢

with finitep-entropies. If one can show that the flow of the Fokker–Planck equation eventually forces the solution to enterL2¡

Rd,f1¢

, we would be able to utilise the idea we just presented, at least from that time on.

Thisexplicit non-symmetrichypercontractivity result we desire, is the main new theo-rem we present in this section.

Theorem 1.4.3. Consider the Fokker–Planck equation (1.1.2)with diffusion and drift matricesDandCsatisfying Conditions (A)–(C). Let f0L1+¡

Rd¢

be a function with unit mass and assume there existsε>0such that

Z

Rdeε|x|2f0(x)d x< ∞. (1.4.1) (i) Then, for any q>1, there exists an explicit t0>0that depends only on geometric

constants of the problem such that the solution to(1.1.2)satisfies Z

Rd f(t,x)qf−1(x)d x≤

µ q

π(q+1)

qd2 µ 8π2 q−1

d 2µZ

Rdeε|x|2f0(x)d x

q

(1.4.2) for all tt0.

(ii) In particular, if f0satisfies ep(f0|f)< ∞for some1<p<2we have that

e2(f(t)|f)≤1 2

 Ã8p

2 3·21p

!d

¡p(p−1)ep(f0|f)+1¢p2

−1

, (1.4.3) for tt˜0(p)>0, which can be given explicitly.

Remark 1.4.4. As we consider ep in our hypercontractivity, which is, up to a constant, the Lp norm of g:= ff

with the measure f(x)d x, one can view our result as a hypercon-tractivity property of the Ornstein-Uhlenbeck operator, P (for an appropriate choice of the diffusion matrixQand drift matrixB), discussed in §1.3. With this notation,(1.4.3) is equivalent to

kg(t)kL2(f)Cp,dkg0kLp(f), tt˜0(p) (1.4.4) for1<p<2, where Cp,d :=

µ

8p 2 3·2

1p

d2

. Since e2decreases along the flow of our equation, (1.4.4)is valid for p=2with C2,d =1. Thus, by using the Riesz-Thorin theorem one can improve inequality(1.4.4)to the same inequality with the constant C

2 p−1

p,d . We would like to point out at this point that a simple limit process shows that (1.4.4)is also valid for p=1, but there is no connection between the L1norm of g and the Boltzmann entropy, e1, of f0.

Remark 1.4.5. Since its original definition for the Ornstein-Uhlenbeck semigroup in the work of Nelson, [16], the notion of hypercontractivity has been studied extensively for Markov diffusive operators (implying selfadjointness). A contemporary review of this topic can be found in [4]. For such selfadjoint generators, hypercontractivity is equiv-alent to the validity of a logarithmic Sobolev inequality, as proved by Gross [10]. For non-symmetric generators, however, this equivalence does not hold: While a log Sobolev inequality still implies hypercontractvity of related semigroups (cf. the proof of Theorem 5.2.3 in [4]), the reverse implication is not true in general (cf. Remark 5.1.1 in [22]). In particular, hypocoercive degenerate parabolic equations cannot give rise to a log Sobolev inequality, but they may exhibit hypercontractivity (as just stated above).

The last 20 years have seen the emergence of the, more delicate, study of hypercontractiv-ity for non-symmetric and even degenerate semigroups. Notable works in the field are the paper of Fuhrman, [9], and more recently the work of Wang et al., [6, 7, 21]. Most of these works consider an abstract Hilbert space as an underlying domain for the semigroup, and to our knowledge none of them give an explicit time after which one can observe the hy-percontractivity phenomena (Fuhrman gives a condition on the time in [9]).

Our hypercontractivity theorem, which we will prove shortly, gives not only an explicit and quantitative inequality, but also provides an estimation on the time one needs to wait before the hypercontractivity occurs. To keep the formulation of Theorem 1.4.3 sim-ple we did not include this “waiting time” there, but we emphasised it in its proof. More-over, the hypercontractivity estimate from Theorem 1.4.3(i) only requires(1.4.1), a weighted L1norm of f0. This is weaker than in usual hypercontractivity estimates, which use Lp norms as on the r.h.s. of (1.4.4).

It is worth to note that we prove our theorem under the setting of theep entropies, which can be thought of asLp spaces with a weight function that depends onp.

In order to be able to prove Theorem 1.4.3 we will need a few technical lemmas.

Lemma 1.4.6. Given f0L1+¡ Rd¢

with unit mass, the solution to the Fokker–Planck equa-tion(1.1.2)with diffusion and drift matricesDandCthat satisfy Conditions (A)–(C) is given by

f(t,x)= 1 (2π)d2 p

detW(t) Z

Rde12

¡xeCty¢T

W(t)1¡

xeCty¢

f0(y)d y, (1.4.5) where

W(t) :=2 Z t

0

eCsDeCTsd s.

This is a well known result, see for instance §1 in [12] or §6.5 in [19].

Lemma 1.4.7. Assume that the diffusion and drift matrices,DandC, satisfy Conditions (A)–(C), and letKbe the unique positive definite matrix that satisfies

2D=CK+KCT.

Then (in any matrix norm)

kW(t)−Kk ≤c(1+t2n)e2µt, t≥0,

where c>0is a geometric constant depending on n andµ, with n being the maximal defect of the eigenvalues ofCwith real partµ, defined in(1.1.5).

Proof. We start the proof by noticing thatK is given by K=2

Z

0

eCsDeCTsd s (see for instance [18]). As such

kW(t)−Kk ≤2 Z

t keCsDeCTskd s≤2kDk Z

t keCskkeCTskd s.

Using the fact that

AeCtA1=eACA1t

for any regular matrixA, we conclude that, ifJis the Jordan form ofC, then

keCtk ≤ kAJkkA−1J kkeJtk, (1.4.6) whereAJis the similarity matrix betweenCand its Jordan form.

For a single Jordan block of sizen+1 (corresponding to a defect ofnin the eigenvalue λ),Je, we find that

eJet=

eλt t eλt . . . tn!neλt eλt . .. tn−1

(n−1)!eλt . .. ...

0 eλt

where Je=

λ 1 0

. .. ...

1

0 λ

 .

Thus, we conclude that keJetxk1

n+1X

i=1 n+1X

j=i

tji

(ji)!eRe(λ)t¯

¯xj¯

¯≤ Ãn+1

X

i=1

¡1+tn¢ eRe(λ)t

! kxk1

=(n+1)¡ 1+tn¢

eRe(λ)tkxk1, t≥0.

Due to the equivalence of norms on finite dimensional spaces, there exists a geometric constantc1>0, that depends onn, such that

keJetk ≤c1¡ 1+tn¢

eRe(λ)t. (1.4.7)

Coming back to C, we see that the above inequality together with (1.4.6) imply that keCtkis controlled by the norm ofC’s largest (measured by the defect number) Jordan block of the eigenvalue with smallest real part. From this, and (1.4.7), we conclude that keCtk ≤c2(1+tn)e−µt, t≥0. (1.4.8) The same estimation forkeCTtkimplies that

kW(t)−Kk ≤c3 Z

t

¡1+s2n¢

e−2µsd s, for some geometric constantc3>0 that depends onn. Since

Z

t

s2ne−2µsd s=

· 1

t2n+ 2n

(2µ)2t2n−1+2n(2n−1)

(2µ)3 t2n−2+...+ (2n)!

(2µ)2n+1

¸ e−2µt we conclude the desired result.

While we can continue with a general matrix K, it will simplify our computations greatly ifK would have beenI. Since we are working under the assumption thatD= CS, the normalization from Theorem 1.2.5 implies exactly that. Thus, from this point onwards we will assume thatK isI.

Lemma 1.4.8. For anyε>0there exists an explicit t1>0such that for all tt1 kW1(t)−Ik ≤ε,

whereW(t)is as in Lemma 1.4.7. An explicit, but not optimal choice for t1is given by

t1(ε) := 1

2(µα)log

c(1+ε)³ 1+¡ n

αe

¢2n´ ε

, (1.4.9)

where0<α<µis arbitrary and c>0is given by Lemma 1.4.7.

Proof. We have that for any invertible matrixA

kA−1Ik = k(AI)A−1k ≤ kAIkkA−1k. In addition, ifkAIk <1, then

kA1k = k(I(IA))1k ≤ 1 1− kAIk. Thus, for anyt>0 such thatkW(t)−Ik <1 we have that

kW−1(t)−Ik ≤ kW(t)−Ik

1− kW(t)−Ik. (1.4.10)

Defining ˜t1(ε) as

t˜1(ε) :=min

½ s≥0

¯

¯

¯

¡1+t2n¢

e−2µtε

c(1+ε), ∀ts

¾

, (1.4.11)

with the constantcgiven by Lemma 1.4.7, we see from Lemma 1.4.7 that for anytt˜1(ε) kW(t)−Ik ≤ ε

1+ε.

Combining the above with (1.4.10), shows the first result fort1=t˜1(ε).

To prove the second claim we will show that t1(ε)≥t˜1(ε).

For this elementary proof we use the fact that maxt≥0 eattb=

µ b ae

b

for anya,b>0. Thus, choosinga=2α, where 0<α<µis arbitrary, andb=2nwe have that

¡1+t2n¢

e−2µt≤ µ

1+

³ n αe

´2n

e−2(µ−α)t, t≥0.

As a consequence, if µ

1+

³ n αe

´2n

e2(µ−α)tε

c(1+ε), ∀ts, (1.4.12) thenst˜1(ε) due to (1.4.11). The smallest possibles in (1.4.12) is obtained by solving the corresponding equality fort, and yields (1.4.9), concluding the proof.

We now have all the tools to prove Theorem 1.4.3

Proof of Theorem 1.4.3. To show (i) we recall Minkowski’s integral inequality, which will play an important role in estimating theLp norms off(t).

Minkowski’s Integral Inequality:For any non-negative measurable function F on(X1× X2,µ1×µ2), and any q≥1one has that

µZ

X2

¯

¯

¯

¯ Z

X1

F(x1,x2)dµ1(x1)

¯

¯

¯

¯

q

2(x2)

q1

≤ Z

X1

µZ

X2

|F(x1,x2)|q2(x2)

q1

1(x1).

(1.4.13)

Next, we fix anε1=ε1(ε,q)∈(0, 1), to be chosen later. From Lemma 1.4.7 and 1.4.8 we see that, fortt1(ε1) with

t1(ε1) := 1

2(µ−α)log

c(1+ε1)³ 1+¡n

αe

¢2n´ ε1

 for some fixed 0<α<µ, we have that

kW(t)−Ik ≤ ε1

1+ε1 <ε1, kW−1(t)−Ik ≤ε1, and hence

W(t)>(1−ε1)I, W(t)−1≥(1−ε1)I. As such, fortt11)

¯

¯

¯

¯ e12

¡xeCty¢T

W(t)1¡

xeCty¢

f0(y)

¯

¯

¯

¯

q

eq2(1−ε1)

¯

¯xeCty¯

¯

2¯

¯f0(y)¯

¯

q (1.4.14)

and

detW(t)≥(1−ε1)d. (1.4.15)

We conclude, using (1.4.13), the exact solution formula (1.4.5), (1.4.14) and (1.4.15) that fortt11) it holds:

Z

Rd

¯

¯f(t,x)¯

¯

qf−1(x)d x

≤ (2π)d2 (2π(1−ε1))qd2

ÃZ

Rd

µZ

Rdeq2(1−ε1)

¯

¯xeCty¯

¯

2¯

¯f0(y)¯

¯

qe|x|

2 2 d x

1q d y

!q

= (2π)d2 (2π(1−ε1))qd2

ÃZ

Rd

µZ

Rdeq2(1−ε1)

¯¯xeCty¯

¯

2

e|x|

2 2 d x

q1

¯

¯f0(y)¯

¯d y

!q

.

(1.4.16)

We proceed by choosing ε1>0 such that q(1ε1)>1 (or equivalentlyε1< q−1q ) and denoting

η:=q(1ε1)−1>0.

Shifting thexvariable by12eCtyand completing the square, we find that Z

Rdeq2(1−ε1)

¯

¯xeCty¯

¯

2

e|x|

2 2 d x=

Z

Rdeη+21

¯

¯x12eCty¯

¯

2

e

¯

¯

¯x+1 2eCt y¯¯

¯ 2

2 d x

= Z

RdexeCtyeη2

¯

¯x12eCty¯

¯

2

d x= Z

Rdeη2

¯

¯

¯x−12

³ 1+2η

´ eCty

¯

¯

¯

2

e

³1 2+21η

´¯

¯eCty¯

¯

2

d x

= µ2π

η

d2 e

³1 2+21η´¯

¯eCty¯

¯

2

.

(1.4.17)

Using (1.4.8) we can find a uniform geometric constantc2such that keCtk2c22¡

1+tn¢2

e−2µt ≤2c22¡

1+t2n¢ e−2µt. Following the proof of Lemma 1.4.8 we recall that if

t≥ 1

2(µ−α)log

Ãc˜(1+ε2

1+αne¢2n

ε2

! , where 0<α<µis arbitrary and for any ˜c,ε2>0, then

¡1+t2n¢

e−2µtε2

˜

c(1+ε2). Thus, choosing

˜

c=c22(1+η)

= c22(1−ε1)

q(1−ε1)−1 and ε2= ε1

1−ε1

we get that if

tt21) := 1

2(µα)log

Ãc22(1−ε1

1+αen ¢2n

¡q(1−ε1)−1¢ ε1

! , where 0<α<µis arbitrary and for any ˜c,ε2>0, then

µ1 2+ 1

keCtk2c22(1+η) q¡

1+t2n¢

e2µt1.

Combining this with our previous computations ((1.4.16) and (1.4.17)), we find that for anytt01) :=max (t11),t21))

Z

Rd

¯

¯f(t,x)¯

¯

qf1(x)d x≤ (2π)d(1−q2) (1−ε1)qd2 ηd2

µZ

Rdeε1|y|2f0(y)d y

q

.

Ifε1is chosen more restrictively than before, namelyε1q−12q , then we have q−1

2 ≤η<q−1 and 1−ε1q+1 2q ,

which implies the first statement of the theorem by choosingε1:=min³ ε,q−12q ´

. For the proof of (ii) we note that (1.4.3) is equivalent to

kf(t)k2L2(Rd,f−1)≤ Ã8p

2 3·21p

!d

kf0k2

Lp

³Rd,f1−p

´. (1.4.18)

With the Hölder inequality we obtain Z

Rde

p−1

4p |x|2f0(x)d x≤ µZ

Rde|x|

2 4 d x

pp1µZ

Rdep−12 |x|2f0p(x)d x

p1

=2d2

p−1

p kf0kLp³

Rd,f1−p´.

Hence,ep(f0|f)< ∞implies (1.4.1) withε=p4p1, and (1.4.18) follows from (1.4.2) with q=2 and ˜t0(p)=t0³p

1 4p

´ .

Remark 1.4.9. If the condition(1.4.1)holds forε=12we can give an explicit upper bound for the “waiting time” in the hypercontractivity estimate (1.4.2). For such εwe have ε1:=min³

ε,q2q1´

= q2q1, and by choosing α= µ2 we can see that t0(ε1)from the proof of Theorem 1.4.3 is

t0(q) := 1 µlog

 max

³

c(3q−1), 2c22q+1q−1

´µ 1+

³2n µe

´2nq−1

 ,

where c,c2are geometric constants found in the proof of Lemma 1.4.7.

With the non-symmetric hypercontractivity result at hand, we can finally complete the proof of our main theorem for 1<p<2.

Proof of Theorem 1.1.4 for1<p<2. Using Theorem 1.4.3 (i i) we find an explicit T0(p) such that for any tT0(p) the solution to the Fokker–Planck equation, f(t), is inL2¡

Rd,f−1¢

. Proceeding similarly to the previous remark (but now with q =2 and ε=p4p−1) we haveε1:=min

³p

−1 4p ,14

´

=p4p−1. This yields the following upper bound for the

“waiting time” in the hypercontractivity estimate (1.4.3):

T0(p) :=1 µlog

 max³

c(5p−1), 2c223pp2++1p´µ 1+

³2n µe

´2np−1

 .

Using Lemma 1.4.2, Theorem 1.1.4 for p=2 (which was already proven in §1.3), and inequality (1.4.3) we conclude that for anytT0(p)

ep(f(t)|f)≤2e2(f(t)|f)≤2 ˜c2e2¡

f(T0(p))|f¢³ 1+¡

tT0(p)¢2n´

e−2µ(t−T0(p))

≤2 ˜cpe2µT0(p)¡

p(p−1)ep(f0|f)+1¢p2¡

1+t2n¢ e−2µt.

(1.4.19)

To complete the proof we recall that any admissible relative entropy decreases along the flow of the Fokker–Planck equation (see [2] for instance). Thus, for anytT0(p) we have that

ep(f(t)|f)≤ep(f0|f)≤ep(f0|f)e2µT0(p)¡

1+t2n¢

e−2µt. (1.4.20) The theorem now follows from (1.4.19) and (1.4.20), together with the fact that for a 1<p<2

ep(f0|f)≤Cp

¡p(p−1)ep(f0|f)+1¢p2 , whereCp:=supx0 x

(p(p−1)x+1)

p2 < ∞.

We end this section with a slight generalization of our main theorem:

Theorem 1.4.10. Letψbe a generating function for an admissible relative entropy. As-sume in addition that there exists Cψ>0such that

ψp(y)≤Cψψ(y) (1.4.21)

for some1<p <2and all y∈R+. Then, under the same setting of Theorem 1.1.4 (but now with the assumption eψ(f0|f)< ∞) we have that

eψ(f(t)|f)≤cp,ψ¡

eψ(f0|f)+1¢2p¡

1+t2n¢

e2µt, t≥0, where cp,ψ>0is a fixed geometric constant.

Proof. The proof is almost identical to the proof of Theorem 1.1.4. Due to (1.4.21) we know thatep(f0|f)< ∞. As such, according to Theorem 1.4.3 (i i) there exists an ex-plicitT0(p) such that for alltT0(p) we have that f(t)∈L2¡

Rd,f1¢ and

e2(f(t)|f)≤1 2

 Ã8p

2 3·21p

!d

¡Cψp(p−1)eψ(f0|f)+1)¢2p

−1

.

The above, together with Lemma 1.4.2 gives the appropriate decay estimate oneψ for tT0(p). Sinceeψdecreases along the flow of our equation, we can deal with the inter-valtT0(p) like in the previous proof, yielding the desired result.

In the next, and last, section of this chapter we will mention another natural quantity in the theory of the Fokker–Planck equations - the Fisher information. We will briefly explain how the method we presented here is different to the usual technique one con-siders when dealing with the entropy. Moreover we describe how to infer from our main theorem an improved rate of convergence to equilibrium - in relative Fisher informa-tion.