Sei sulla pagina 1di 16

J. Math. Fluid Mech.

19 (2017), 709–724
c 2016 Springer International Publishing
1422-6928/17/040709-16 Journal of Mathematical
DOI 10.1007/s00021-016-0302-5 Fluid Mechanics

Analytic Semigroup Approach to Generalized Navier–Stokes Flows in Besov Spaces


Zhi-Min Chen

Communicated by H. Beirao da Veiga

Abstract. The energy dissipation of the Navier–Stokes equations is controlled by the viscous force defined by the Laplacian
−Δ, while that of the generalized Navier–Stokes equations is determined by the fractional Laplacian (−Δ)α . The existence
and uniqueness problem is always solvable in a strong dissipation situation in the sense of large α but it becomes complicated
when α is decreasing. In this paper, the well-posedness regarding to the unique existence of small time solutions and small
initial data solutions is examined in critical homogeneous Besov spaces for α ≥ 1/2. An analytic semigroup approach to
the understanding of the generalized Navier–Stokes equations is developed and thus the well-posedness on the equations is
examined in a manner different to earlier investigations.

Mathematics Subject Classification. 35B32, 35B35, 35Q35, 86A10.

Keywords. Generalized Navier–Stokes equations, well-posedness, analytic semigroup, Besov spaces.

1. Introduction

The Navier–Stokes equations, governing the motion of incompressible and viscous fluid substances through
the balance of viscous force, pressure force and fluid acceleration, are a fundamentally important system in
fluid mechanics. The three-dimensional Navier–Stokes equations also give rise to Leray’s question of global
existence of classical solutions [27]. This is of great interest to mathematicians.. In the attempt to answer
the question, a variety of dissipative equations were examined (see, for example [13,22,25,31,32,36]).
In the present paper, we are interested in the generalized Navier–Stokes equations
∂t u + (−Δ)α u + u · ∇u + ∇π = 0, ∇·u=0 (1)
in the domain (0, T ) × R for the dimension number d ≥ 2 and the time bound 0 < T ≤ ∞. Here
d

u and π represent respectively unknown fluid velocity and pressure. The fluid motion described by (1)
involves an energy dissipation which is controlled by the fractional Laplacian (−Δ)α . The existence,
uniqueness and regularity of the solution to (1) is essentially determined by the parameter α. When α = 1,
Eq. (1) is the Navier–Stokes equations and thus Leray’s question remains open for d ≥ 3. Equation (1)
with α > 1 is called a hyper-dissipative system. The global existence of classical solutions with d ≥ 3 can
only been obtained when α > d+2 4 [22,37]. However, the existence of classical global solution to the two-
dimensional Navier–Stokes equations can be solved readily due to the validity of the energy inequality in
the L2 space [20]. Actually, similar to the ideal flow governed by the two-dimensional Euler equations, the
maximum principle holds true for the vorticity formulation of the two-dimensional generalized Navier–
Stokes equations. This results in the validity of an additional L∞ estimate of the corresponding vorticity
field and hence the existence of the global classical solutions to the two-dimensional equations for α ≥ 0
(see, for example [6]). This global solution existence can be extended from that of the two-dimensional
Euler equations [5,19,41]. For the well-posedness with respect to the existence and uniqueness of small
global solutions or small time solutions to (1) with α > 12 , we refer to [39] for solutions in critical
homogeneous Besov spaces and [28] for solutions in critical homogeneous Q spaces introduced in [28,40].
For the limit case of α = 0 and d = 2, Eq. (1) represents actually the simplified atmospheric circulation
710 Z.-M. Chen JMFM

equations system of Charney and DeVore [7,9,12,35]. The vorticity formulation of (1) with d = 2 is
similar to the surface dissipative quasi-geostrophic equation [6,10,24,36]. However, the vorticity-velocity
relations ω = ∇ × u in the former and θ = ∇⊥ (−Δ)−1/2 u in the latter lead to the difference for the global
solution existence of the two systems. For the problem on the nonexistence of the global solutions to the
supercritical surface dissipative quasi-geostrophic equation, one may refer to [23].
A homogeneous Besov space is a collection of the functions f which can be expressed as the Littlewood–
Paley decomposition

f= Δj f
−∞<j<∞

for the dyadic blocks Δj . This unit decomposition helps define the norm of the homogeneous Besov space
and examine the function f by reducing to the examination of individual terms Δj f through Bernstein
inequalities (see, for example, [26]). The Littlewood–Paley decomposition approach is overwhelming in
the analytic study of fluid mechanics equations (see, for example, [14] for the compressible Navier–
Stokes equations, [15] for the density-dependent incompressible Navier–Stokes equations, [38,39] for the
generalized Navier–Stokes equations (1), [3] for the incompressible NavierStokes equations, and [8] for
the dissipative quasi-geostrophic equation).
Note that the fractional Laplacian (−Δ)α generates an analytic semigroup operator e−t(−Δ) , which
α

can be characterized in homogeneous Besov spaces in terms of the Littlewood–Paley decomposition


[33,34]. This characterization simplifies the derivation of a priori estimates of the system (1) and was
applied to the understanding of the well-posedness of the Navier–Stokes equations [11]. The purpose of
the present paper is to extend this approach to the generalized Navier–Stokes equations (1) and then to
provide a new approach to the understanding of a higher regularity problem initiated from [18].
< is adopted in the sense that a < a represents the inequality a ≤ Ca
For simplicity, the notation ∼ 1∼ 2 1 2
for a constant C independent of the quantities t ∈ [0, T ), x ∈ Rd , T > 0, functions u, v, f , g and a scalar
constant λ > 0. Moreover, we use the equivalence relation a1  a2 if both a1 ∼ < a and a < a hold true.
2 2∼ 1
As is well known, Eq. (1) is invariant with respect to the scale transformation
uλ (t, x) = λ2α−1 u(λ2α t, λx), πλ (t, x) = λ4α−2 π(λ2α t, λx).
That is, Eq. (1) is equivalent to the scaled equations
∂t uλ + (−Δ)α uλ + uλ · ∇uλ + ∇πλ = 0, ∇ · uλ = 0. (2)
According to this scale-invariant property, it is beneficial to examine Eq. (1) in homogeneous function
spaces. A spatial function space X s (Rd ) is said to be homogeneous of degree s if
f (λ ·) X s (Rd )  λs f X s (Rd ) for f ∈ X s ,
and a temporal-spacial function space Y s ([0, T ) × Rd ) for 0 < T ≤ ∞ is said to be homogeneous of degree
s if
f (λ2α ·, λ·) Y s ([0,T )×Rd )  λs f Y s ([0, λ2α T )×Rd ) for f ∈ Y s ([0, T ) × Rd )
for −∞ < s < ∞ and λ > 0.
For the integral formulation of (1) transformed as
u + (∂t + (−Δ)α )−1 (u · ∇u + ∇π) = 0, ∇·u=0 (3)
initially from the static state u(0) = 0, an analytic semigroup approach is applied to Eq. (3) on a
homogeneous space Y s involving homogeneous Besov spaces to obtain the estimate
< (∂ + (−Δ)α )−1 (u · ∇u) s < uu s+1−2α < u s u 1−2α .
u Y s ∼ t Y ∼ Y ∼ Y Y (4)
This together with a linearization technique implies that the homogeneous space Y s with s = 1 − 2α
is the critical space for ensuring the unique existence of the solution to (1). Moreover, it is obtained
that the estimates involve an unbounded exponential differential operator or an inverse of an analytic
Vol. 19 (2017) Analytic Semigroup Approach to Generalized Navier–Stokes Flows in Besov Spaces 711

semigroup operator. This is known as the Gevrey regularity problem examined in [18] for Navier–Stokes
equations in Hilbert spaces and is further developed in [4,26]. For the recent investigations of this problem
in dissipative quasi-geostrophic equations and some other related dissipative equations, one may refer to
[1,2,16,17].
The significance of the present paper is the development of an analytic semigroup approach in Besov
spaces for the well-posedness and Gevrey regularity problem of the generalized Navier–Stokes equations
(1) and provides an alternative way for the understanding of the problem different to the Bernstein
estimates approach based on the dyadic Littlewood–Paley blocks in earlier studies (see, for example [39]).
This paper is organized as follows. Firstly, we display basic properties of Besov spaces and its con-
nection with the analytic semigroup e−t(−Δ) in Sect. 2. Next, we derive a priori estimates of linearized
α

equation of (1) in homogeneous Besov spaces by using the semigroup approach. Finally, the result ob-
tained for the linearized system is extended to the nonlinear equation (1) to show the well-posedness and
Gevrey regularity properties.

2. Besov Spaces and Basic Properties

For the readers’ convenience, we collect basic properties of Besov spaces and their homogeneous counter-
parts from [33,34].
Standard notion is used. S(Rd ) represents the Schwartz space of rapidly decreasing infinitely differen-
tial functions on Rd . Its dual space S  (Rd ) is the collection of all tempered distributions on Rd . Lp (Rd )
denotes the Lebesgue space of functions on Rd for 1 ≤ p ≤ ∞. F represents the Fourier transform in
S  (Rd ) satisfying F f (ξ) = Rd e−ix·ξ f (x)dx and the inverse Fourier transform is defined as F −1 .
As in [33, Definitions 2.3.1/1, 2.3.1/2 and 5.1.3/2], let φ be a smooth function on Rd in the following
sense φ(ξ) = 0 for |ξ| ≥ 2 and φ(ξ) = 1 for |ξ| ≤ 1 which produces the dyadic block symbols
ψj (ξ) = φ(2−j ξ) − φ(2−j+1 ξ)
and the dyadic blocks
Δj = F −1 ψj F
for j ∈ Z, the integer set. Therefore, for 1 ≤ p, q ≤ ∞ and −∞ < s < ∞, we define the Besov space
⎧ ⎛ ⎞ q1 ⎫

⎨ ∞ ⎪


s
Bp,q (Rd ) = f ∈ S  (Rd ); f Bp,q
s = ⎝ F −1 φF f qLp + 2jsq Δj f qLp ⎠ < ∞ ,

⎩ ⎪

j=1

and the homogeneous Besov space


⎧ ⎛ ⎞ q1 ⎫

⎨ ∞ ∞
 ⎪

Ḃp,q (R ) = f ∈ S  (Rd ); f =
s d
Δj f, f Ḃ s = ⎝ 2jsq Δj f qLp ⎠ < ∞ .

⎩ p,q ⎪

j=−∞ j=−∞

Here the lq norm reduces to the l∞ norm for the sequences { F −1 φF f Lp }∪{2js Δj f Lp }j∈N and
{2js Δj f Lp }j∈Z when q = ∞.
s
It is readily seen that the space Ḃp,q (Rd ) is homogeneous of degree s− dp and its norm can be expressed
as
f (λ ·) Bp,q
s
f Ḃ s = lim , λ = 2k (5)
p,q k→∞ s− d
λ p

for s > 0 and f ∈ Bp,q


s
(Rd ).
712 Z.-M. Chen JMFM

Lemma 1 ([33, Theorem 5.2.3/1]). Let −∞ < σ, s < ∞ and 1 ≤ p, q ≤ ∞. Then the fractional operator
σ
s
(−Δ) 2 maps Ḃp,q (Rd ) isomorphically onto Ḃp,q
s+σ
(Rd ) or there holds the norm equivalence
σ
(−Δ) 2 f Ḃ s  f Ḃp,q
s+σ for f ∈ Ḃp,q
s+σ
(Rd ). (6)
p,q

The real interpolation property of Besov spaces [33] gives the following interpolation inequality.
Lemma 2 ([33, Theorem 2.4.2 and Subsection 5.2.5]). Let 0 < θ < 1, 1 ≤ p, q0 , q1 , q ≤ ∞, −∞ < s <
τ < ∞ and σ = (1 − θ)s + θτ . Then there holds the real interpolation property
σ
Ḃp,q (Rd ) = (Ḃp,q
s
0
(Rd ), Ḃp,q
τ
1
(Rd ))θ,q
and
f Ḃ σ ∼< f 1−θ f θ (7)
s τ
Ḃp,q
p,q Ḃp,q 1
0

for f ∈ Ḃp,q
s
0
(Rd ) ∩ Ḃp,q
τ
1
(Rd ).
s
The combination of Eq. (5) and the characterization of the Besov space Bp,∞ (Rd ) (see [33, Eqs. 2.3.5/3,
Theorem 2.5.12]) by the difference operators
0h f (x) = f (x), 1h f (x) = f (x + h) − f (x), kh f (x) = 1h ( k−1
h f (x)) for h ∈ Rd
produces the following characterization of its homogeneous counterpart.
Lemma 3. Let s > 0, 1 ≤ p ≤ ∞ and an integer k > s. Then there holds the equivalence relation
kh f Lp
f Ḃ s  ess sup for f ∈ Ḃp,∞
s
(Rd ). (8)
p,∞
0=h∈Rd |h|s

The operator e−t(−Δ) = F −1 e−t|ξ| F is an analytic semigroup in Lp (Rd ) generated by the frac-
α

tional Laplacian (−Δ)α = F −1 |ξ|2α F , which maps the fractional Sobolev space Hp2α (Rd ) = (1 +
(−Δ)α )−1 Lp (Rd ) onto Lp (Rd ), due to the validity of the property

(−Δ)α e−t(−Δ) f Lp ∼
< t−1 f
α
Lp for t > 0. (9)

Therefore, for 1 < p < ∞ and k ≥ 1, the characterization property [34, Proposition 1.8.1/1]
 
 
Lp (Rd ), Hp2αk (Rd ) θ,∞ = f ∈ Lp (Rd ); sup tk−kθ (−Δ)αk e−t(−Δ) f Lp < ∞
α

t>0

and the real interpolation property [33, Theorems 2.4.2 and 2.5.6]
 
Lp (Rd ), Hp2αk (Rd ) θ,∞ = Bp,∞
2αkθ
(Rd )
s
yield the characterization of the Besov space Bp,∞ (Rd )

 f Lp + sup tk− 2α (−Δ)αk e−t(−Δ) f Lp


s α
f Bp,∞
s (10)
t>0

for s > 0, 1 < p < ∞, and k > 2α


s
. Using (5) to remove the lower order part f Lp of (10), we obtain the
characterization of the homogeneous Besov space by the analytic semigroup e−t(−Δ) in the following.
α

Lemma 4. Let α > 0, s > 0, 1 < p < ∞ and an integer k > s/(2α). Then we have
 sup tk− 2α (−Δ)αk e−t(−Δ) f Lp for f ∈ Ḃp,∞
s α
f Ḃ s s
(Rd ). (11)
p,∞
t>0

Thus the homogeneous Besov space defined by the Littlewood–Paley decomposition reduces to that
characterized by the analytic semigroup e−t(−Δ) generated by (−Δ)α .
α
Vol. 19 (2017) Analytic Semigroup Approach to Generalized Navier–Stokes Flows in Besov Spaces 713

1 1
Lemma 5 ([26, Lemma 24.8]) Let d ≥ 1, 1 < p, q < ∞, r = p + 1q , f ∈ Lp (Rd ) and g ∈ Lq (Rd ). Then
there holds the inequality
et|∇|1 ((e−t|∇|1 f ) e−t|∇|1 g) Lr ∼
< f g ,
Lp Lq (12)
1 1 d
where the operator et 2α |∇|1 = F −1 et 2α |ξ|1 F by using the l1 norm |ξ|1 = i=1 |ξi |.
This Hölder-like inequality given in [26] is essential in the investigation of Gevrey regularity problem.
For the readers’ convenience, it is demonstrated as follows.

Proof. Let begin with the case d = 1. For |∇| = |∇|1 , fˆ = F f and real variables x, y, η, ξ ∈ R in the
one-dimensional situation, we have
et|∇| (e−t|∇| f e−t|∇| g)(x)

= eix(ξ+η) et|ξ+η|−t|η|−t|ξ| fˆ(η)ĝ(ξ)dηdξ
R2
 
ix·(ξ+η) ˆ
= e f (η)ĝ(ξ)dηdξ + eix(ξ+η) et|ξ+η|−t|η|−t|ξ| fˆ(η)ĝ(ξ)dηdξ
ηξ≥0 ηξ<0

= eix(ξ+η) (1 − e−2t|ξ| )fˆ(η)ĝ(ξ)dηdξ
ηξ≥0
 ∞  −η  0  ∞ 
+ dη dξ + dη dξ eix(ξ+η) e−2t|η| fˆ(η)ĝ(ξ)
0 −∞ −∞ −η
 ∞ ∞  0  −η 
+ dη dξ + dξ eix(ξ+η) e−2t|ξ| fˆ(η)ĝ(ξ)

0 −η −∞ −∞
 
ix(ξ+η) −2t|ξ| ˆ
= e (1 − e )f (η)ĝ(ξ)dηdξ + eixξ e−2t|η| fˆ(η)ĝ(ξ − η)dηdξ
ηξ≥0 ξη<0

.
+ eixξ e−2t|ξ−η| fˆ(η)ĝ(ξ − η)dηdξ = I1 + I2 + I3 .
ξη>0

Defining the half space Fourier transforms


 ∞  0
S+ f = eixξ fˆdx and S− f = eixξ fˆdx,
0 −∞

we have
 ∞  ∞  0  0
I1 = eixη fˆ(η)dη eixξ (1 − e−2t|ξ| )ĝ(ξ)dξ + eixη fˆ(η)dη eixξ (1 − e−2t|ξ| )ĝ(ξ)dξ
0 0 −∞ −∞
−2t|∇| −2t|∇|
= (S+ f )S+ (1 − e )g + (S− f )S− (1 − e )g,
 0   ∞ 
I2 = eixξ F g(y) eiyη e−2t|η| fˆ(η)dη (ξ)dξ
−∞ 0
 ∞    0
ixξ ˆ iyη −2t|η|
+ e F g(y) e e f (η)dη (ξ)dξ
0 −∞
   
= S− gS+ e−2t|∇| f + S+ gS− e−2t|∇| f

and, similarly,
I3 = S+ (f S+ e−2t|∇| g) + S− (f S− e−2t|∇| g).
Note that
< f
S± f ∼ Lτ and e−2t|∇|1 f Lτ ∼
< f
Lτ (13)
714 Z.-M. Chen JMFM

for τ = r, p, q since the operators S± are Lτ Fourier multiplier operators [30, Theorem 4.1.4] and e−2t|∇|1
is obviously an Lτ Fourier multiplier operator. We thus use Hölder inequality and estimates (13) to obtain
the validity of (12) in the case d = 1. When d > 1, the observation et|∇|1 = et|∂x1 | et|∂x2 | · · · et|∂xd | ensures
that the d-dimensional operator is decomposed into 1-dimensional operators. Thus there exist Lτ Fourier
multiplier operators Li , Mi and Ni given by S± and e−2t|∇|1 so that
3d

et|∇|1 ((e−t|∇|1 f ) e−t|∇|1 g) = Li ((Ni g) Mi f ).
i=1

This together with the estimates (13) and the Hölder inequality implies
3d

et|∇|1 ((e−t|∇|1 f ) e−t|∇|1 g) Lr ∼
< < g f
(Ni g) Mi f Lr ∼ Lp Lq
i=1

and completes the proof of (12) for d > 1. 

3. Analytic Semigroup Approach to Generalized Stokes Equations

In this paper, the existence of local solutions and small solutions of the nonlinear system (1) is examined
and can be obtained by a linearization technique. Therefore, it is beneficial to consider the linearized
system of (1) or the generalized Stokes equations
∂t u + ∇π + (−Δ)α u = ∇ · f, ∇ · u = 0, (14)
coupled with zero initial condition
u(0) = 0 in Rd . (15)
2 d
Here f : Rd → Rd , (∇ · f )j = i=1 ∂i fij and the pressure force term ∇π can be written as
∇π = ∇Δ−1 ∇ · (∇ · f )
due to divergence property of u. Hence (14) is written as
∂t u + (−Δ)α u = P ∇ · f in (0, T ) × Rd , (16)
for the bounded projection operator
P = δij − ∂i ∂j Δ−1
in Lp (Rd ) for 1 < p < ∞, where δij denotes the Kronecker symbol. Now we rewrite the differential
equation (16) in the form of the integral equation
 t
u(t) = (∂t + (−Δ)α )−1 P ∇ · f = e−(t−s)(−Δ) P ∇ · f (s)ds.
α
(17)
0
With the use of the characterization by the analytic semigroup, the purpose of this section is to show
that the operators (∂t + (−Δ)α )−1 P ∇ behaviours like the operator |∇|1−2α mapping a homogeneous
space of degree s into a homogeneous space of degree s + 1 − 2α.
Theorem 1. Let d ≥ 2, 0 < T ≤ ∞, 0 ≤ β < α, −2β < σ and 1 < p < ∞. Define the operators
1
S(t) = F −1 m(t, ξ)F and S −1 (t) = F −1 m(t,ξ) F for a symbol m(t, ·). Assume that the inequality
1
e− 4 (t−s)(−Δ) S(t)S −1 (s)g Lp ∼
α
< g (18)
Lp

holds true for g ∈ Lp (Rd ). Then u(t) presented in Eq. (17) is subject to the estimates
β β
S(t)u(t) Ḃ σ + t α S(t)u(t) Ḃp,∞
σ+2β ∼ ess sup t α S(t)f (t) σ+2β+1−2α
<
Ḃp,∞ (19)
p,∞ 0<t<T
Vol. 19 (2017) Analytic Semigroup Approach to Generalized Navier–Stokes Flows in Besov Spaces 715

and
β
< ess sup t α S(t)f (t)
S(t)u(t) L∞ ∼ 2β+1−2α+ d (20)
0<t<T Ḃp,∞ p

for 0 < t < T , provided that the right-hand sides are finite.
γ
A typical example of the operator S(t) is et α (−Δ) for 0 ≤ γ < α.
γ

Proof. For an integer k > σ+2β


2α , it follows from Lemmas 1 and 4, Eqs. (9) and (18) and the boundedness
of the projection operator P in the space Lp (Rd ) that
S(t)u(t) Ḃp,∞
σ+2β

 t
k− σ+2β αk −τ (−Δ)α
P S(t)S −1 (s)e−(t−s)(−Δ) S(s)∇ · f (s)ds Lp
α
 sup τ 2α (−Δ) e
τ >0 0
 t
k− σ+2β − 34 (t−s+τ )(−Δ)α
< sup τ
∼ 2α (−Δ)αk e S(s)∇ · f (s) Lp ds
τ >0 0
 t
σ+2β 1
(t − s + τ )−1 (−Δ)αk e− 2 (t−s+τ )(−Δ) (−Δ)−α S(s)∇ · f (s) Lp ds
α
< k−
∼ sup τ 2α
τ >0 0
 t
σ+2β σ+2β β β
< sup τ k−
∼ 2α (t − s + τ )−k−1+ 2α s− α s α (−Δ)−α S(s)∇ · f (s) Ḃp,∞
σ+2β ds.
τ >0 0

After the calculation of the integral over the intervals [0, 2t ) and [ 2t , t), the previous inequality reduces to
σ+2β σ+2β β σ+2β β
S(t)u Ḃp,∞ <
σ+2β ∼ sup τ
k− 2α [(t + τ )−k+ 2α t− α + (t + τ )−k−1+ 2α t1− α ]
τ >0
β
× ess sup t α S(t)f (t) Ḃp,∞
σ+2β+1−2α
0<t<T
β β
< t −α ess sup t α S(t)f (t) Ḃ σ+2β+1−2α .
∼ 0<t<T p,∞
(21)

Furthermore, Eq. (9) and the interpolation inequality


1− γ γ
(−Δ)γ e−t(−Δ) g Lp ∼
< e−t(−Δ) g α −t(−Δ)
α α α

Lp (−Δ) e g Lαp
α

imply that
γ
(−Δ)γ e−t(−Δ) g Lp ∼
< t− α g ,
α
Lp (22)

and hence
γ
(−Δ)γ e−t(−Δ) g Ḃ s −α
α
<t
∼ g Ḃ s . (23)
p,∞ p,∞

Using Lemma 1, Eqs. (18) and (23) and the boundedness of the operator P , we have
 t
3
e− 4 (t−s)(−Δ) S(s)∇ · f (s) Ḃ σ ds
α
S(t)u(t) Ḃ σ ∼<
p,∞ p,∞
0
 t
β
<
∼ (t − s)−1+ α (−Δ)−(α−β) S(s)∇ · f (s) Ḃ σ ds
p,∞
0
 t
β β β
<
∼ (t − s)−1+ α s− α ds ess sup t α S(t)f (t) Ḃp,∞
σ+2β+1−2α
0 0<t<T
β
< ess sup t S(t)f (t) σ+2β+1−2α .
∼ 0<t<T
α
Ḃp,∞
716 Z.-M. Chen JMFM

To derive the estimate in the L∞ norm, we employ the Gagaliardo–Nirenberg inequality [29] to obtain
that
1− d d
e−t(−Δ) g L∞ ∼
< e−t(−Δ) g −t(−Δ) < t− 2αq g .
α α α d

Lq ∇e
q
g Lq q ∼ Lq (24)

for a real number q > d. If the number q > p satisfies additionally the condition 2β + dq < 2α, Eqs. (6),
(11), (18), (24) and (22) yield
 t
3
(t − s)− 2αq e− 4 (t−s)(−Δ) S(t)P ∇ · f (s) Lq ds
d α
S(t)u(t) L∞ ∼ <
0
 t
1
(t − s)− 2αq e− 2 (t−s)(−Δ) S(s)∇ · f (s) Lq ds
d α
<

0
 t 2β+ d
(t − s)− 2αq (t − s)−1+ 2α S(s)(−Δ)−α ∇ · f (s) 2β+ dq ds
d q
<

0 Ḃq,∞
 t
β β β
−1+ α −α
<
∼ (t − s) s ds ess sup t S(t)f (t)
α
2β+1−2α+ d
0 0<t<T Ḃq,∞ q

β
< ess sup t α S(t)f (t)
∼ 0<t<T
2β+1−2α+ d
p
,
Ḃp,∞

where we have used the imbedding theorem [33, Theorem 2.7.1 and Subsection 5.2.5]
γ+ d γ+ d
Ḃp,∞p (Rd ) → Ḃq,∞q (Rd ). (25)
Collecting terms, we have the desired estimate and thus complete the proof of Theorem 1. 

2α 1 1
1
The symbol m = e− 4 (t−s)|ξ| +t 2α |ξ|1 −s 2α |ξ|1 satisfies the conditions of the Marcinkiewicz multiplier
theorem [30, Theorem 6.1.6 ] and thus there holds the Lp estimate
1 1
1 1
e− 4 (t−s)(−Δ) et 2α |∇|1 e−s 2α |∇|1 g Lp ∼
α
< g
Lp for α > , (26)
2
1
which gives an example of S(t) = et 2α |∇|1 in (18). For this special choice, Theorem 1 can be stated as

Corollary 1. Let d ≥ 2, 12 < α, 0 ≤ β < α, 1 < p < ∞ and 0 < σ + 2β. Then u = (∂t − (−Δ)α )−1 P ∇ · f
is subject to the estimate
1 β 1 β 1
et 2α |∇|1 u(t) Ḃ σ + t α et 2α |∇|1 u(t) Ḃp,∞
σ+2β ∼ ess sup t α e
< t 2α |∇|1
f (t) Ḃp,∞
σ+2β+1−2α
p,∞ 0<t<T

for 0 < t < T ≤ ∞, provided that the right-hand side term is finite.

4. Well-Posedness of the Nonlinear Problem

The well-posedness of the nonlinear equations (1) can be obtained by combining the estimates of the
linear Stokes equations derived in Theorem 1 and the Banach contraction mapping principle. Therefore
it is necessary to examine the nonlinear term of (1) though the pointwise multiplication f g of functions
f and g in suitable homogeneous spaces in the form
< f 1−2α g 1−2α .
f g Y 2−4α ∼ Y Y (27)

To do so, we need the following function multiplication estimate.


Vol. 19 (2017) Analytic Semigroup Approach to Generalized Navier–Stokes Flows in Besov Spaces 717

Lemma 6. Let 1 < p < q < ∞, s + d


p > 0, an integer k > s + 2 dp and (k − 1)p < q < ∞. Then there holds
the estimate
1 
k−1 1 1
et 2α |∇|1 f g s+ d
<
∼ et 2α |∇|1 f js
+d et 2α |∇|1 g (k−j)s
+d
Ḃp,∞p k
Ḃp,∞ p
Ḃp,∞k p
j=1
1 1 1 1
+ et 2α |∇|1 f d−d
p q
et 2α |∇|1 g s+ d + d + et 2α |∇|1 g d−d et 2α |∇|1 f s+ d + d
Ḃp,∞ Ḃp,∞p q p
Ḃp,∞ q
Ḃp,∞p q

(28)
for functions f and g on Rd , provided the right-hand side terms are finite.
Proof. Let us begin with the proof of the inequality

k−1
f g s+ d
<
∼ f js
+d g (k−j)s
+d
+ f d−d g s+ d + d + g d−d f s+ d + d . (29)
Ḃp,∞p k
Ḃp,∞ p
Ḃp,∞k p p
Ḃp,∞ q
Ḃp,∞p q p
Ḃp,∞ q
Ḃp,∞p q
j=1

For the integer k ≥ s + dp + dq , it is readily seen that the function product under the difference operator
is subject to the binomial formula
k  
k
h (f (x)g(x)) =
k
jh f (x) k−j
h g(x + jh).
j
j=0

1 1 1
For the real number r so that r = p + q, the selection of q and k implies that
1 1 1 kr
< + or > p.
r p (k − 1)p k−1
By the imbedding property (25), the characterization property (8) and the Hölder inequality, we find
that
f g s+ d
< f g
∼ s+ d
Ḃp,∞p Ḃ r r,∞
 
k k
j=0 j jh f k−j
h g(· + jh) Lr
< ess sup
∼ d
0=h∈R
d |h|s+ r
k−1
j=1 jh f L kr k−j
h g L kr
+ f Lq kh g Lp + g Lq kh f Lp
< j k−j
∼ ess supd d
0=h∈R |h|s+ r

k−1
<
∼ f j
(s+ d ) g (k−j)
(s+ d )
+ f Lq g s+ d + g Lq f s+ d . (30)
k
Ḃ kr r
Ḃ k
kr ,∞
r Ḃp,∞r Ḃp,∞r
j=1 j
,∞
k−j

Since kr
j ≥ kr
k−1 > p, the imbedding property (25) implies that
j
0
d
−d j
(s+ d ) s+ d
Lq (Rd ) → Ḃq,∞ (Rd ) → Ḃp,∞
p q
(Rd ) and Ḃ kr
k
,∞
r
(Rd ) → Ḃp,∞
k p
(Rd ).
j

Thus Eq. (30) becomes (29).


1 1
Since et 2α |∇|1 jh f = jh et 2α |∇|1 f , following the derivation of (29) by applying the inequality (12)
instead of the Hölder inequality, we obtain (28) and hence complete the proof of Lemma 6.
Now we consider the nonlinear system described by (1). Let u solve Eq. (1) associated with the initial
condition
u(0) = u0 on Rd . (31)
718 Z.-M. Chen JMFM

Taking the integral form of this system, we have the mild solution of Eqs. (1) and (31) expressed as

u(t) = e−t(−Δ) u0 + (∂t + (−Δ)α )−1 P u · ∇u(t).


α
(32)

The well-posed result for Eqs. (1) and (31) is stated in the following.
1−2α+ d
Theorem 2. Let d ≥ 2, α > 12 , 1 < p < ∞, 2 − 2α + d
p > 0 and u0 ∈ Ḃp,∞ p
(Rd ). Then there exists
1−2α+ d
a constant T > 0 such that Eq. (32) admits a unique solution u ∈ L∞ (0, T ; Ḃp,∞ p
(Rd )) subject to the
estimate
 1 1

β
ess sup t α et 2α |∇|1 u(t) 2β+1−2α+ d + et 2α |∇|1 u(t) 1−2α+ d ∼ < u
0 1−2α+ d + 1, (33)
0<t<T Ḃp,∞ p p
Ḃp,∞ Ḃp,∞ p

where β is a constant such that


d
α > β > 0, 2β + 1 − 2α > 0, 2β + 2 − 4α + > 0.
p
1−2α+ d
Additionally, if u0 is sufficiently small in the space Ḃp,∞ p
(Rd ), the solution u exists globally in
time and the bound (33) remains true for T → ∞.

Proof. For a constant T > 0, we define the operator

Mu0 u(t) = e−t(−Δ) u0 + (∂t + (−Δ)α )−1 u(t)


α
(34)

and the space


 
1−2α+ d
XT = u∈ L∞ (0, T ; Ḃp,∞ p (Rd )); u XT ≤ C u0 1−2α+ d
p
Ḃp,∞

for a suitable constant C and the norm


 1 1

β
u XT = ess sup t α et 2α |∇|1 u(t) 2β+1−2α+ d + e t 2α |∇|1
u(t) 1−2α+ d . (35)
0<t<T Ḃp,∞ p
Ḃp,∞ p

It suffices to show that Mu0 is a contraction operator mapping XT into itself.


Indeed, it is readily seen that
1
et 2α |∇|1 e−t(−Δ) u0 XT ∼
α
< u
0 1−2α+ d
p
Ḃp,∞

1
3
by taking into account the inequality (23) and the fact that et 2α |∇|1 e− 4 t(−Δ) is a bounded multiplier in
α

1−2α+ d
Lp (Rd ) and hence in Ḃp,∞ (Rd ).
p

Let u, v ∈ XT and let the integer k and the real number q > p be given in Lemma 6. It follows from
Theorem 1 with σ = 1 − 2α + dp and the inequality (28) with s = 2β + 2 − 4α that

β 1
Mu0 u XT ∼
< u
0 1−2α+ d
p
+ ess sup t α et 2α |∇|1 uu(t) 2β+2−4α+ d
p
Ḃp,∞ t>0 Ḃp,∞
β 1 1
< u
∼ 0 1−2α+ d
p
+ ess sup t α et 2α |∇|1 u d−d
p q
et 2α |∇|1 u 2β+2−4α+ d + d
p q
Ḃp,∞ t>0 Ḃp,∞ Ḃp,∞


k−1 1 1
+ ess sup et 2α |∇|1 u j(2β+2−4α)
+d
et 2α |∇|1 u (k−j)(2β+2−4α) d
+
0<t<T Ḃp,∞ k p
Ḃp,∞ k p
j=1
Vol. 19 (2017) Analytic Semigroup Approach to Generalized Navier–Stokes Flows in Besov Spaces 719

and
Mu0 u − Mu0 v XT
 1 1

β
< ess sup t α
∼ et 2α |∇|1 (u − v)u 2β+2−4α+ d + et 2α |∇|1 (u − v)v 2β+2−4α+ d
0<t<T Ḃp,∞ p
Ḃp,∞ p

β 1 1
< ess sup t α et 2α |∇|1 (u − v) et 2α |∇|1 u
∼ 0<t<T
d−d
p q
2β+2−4α+ d + d
p q
Ḃp,∞ Ḃp,∞
β 1 1
+ ess sup t α et 2α |∇|1 v d−d et 2α |∇|1 (u − v) 2β+2−4α+ d + d
0<t<T p
Ḃp,∞ q
Ḃp,∞ p q

β 
k−1 1 1
+ ess sup t α et 2α |∇|1 (u − v) j(2β+2−4α)
+d
et 2α |∇|1 u (k−j)(2β+2−4α) d
+
0<t<T Ḃp,∞ k p
Ḃp,∞ k p
j=1

β 
k−1 1 1
+ ess sup t α et 2α |∇|1 (u − v) j(2β+2−4α)
+d
et 2α |∇|1 v (k−j)(2β+2−4α)
+d
.
0<t<T Ḃp,∞ k p
Ḃp,∞ k p
j=1

The number q > p is selected to be sufficiently large so that −2β < 1 − 2α + d


q < 0. Moreover we may
assume the integer k ≥ 2 so that
j
1−α<(2β + 2 − 4α) < 2β + 1 − 2α for j = 1, . . . , k.
k
By Lemma 2, we have the real interpolation relations
 
2β+2−4α+ d
p+q
d
d 1−2α+ dp d 2β+1−2α+ d
p d
Ḃp,∞ (R ) = Ḃp,∞ (R ), Ḃp,∞ (R ) , (36)
1−θ0 ,∞
 
p−q 1−2α+ d 2β+1−2α+ d
d d
Ḃp,∞ (R ) = Ḃp,∞ p (Rd ),
d
Ḃp,∞ p d
(R ) , (37)
θ0 ,∞
(k−j)(2β+2−4α)
 
+d 1−2α+ d 2β+1−2α+ d
Ḃp,∞ k p
(Rd ) = Ḃp,∞ p (Rd ), Ḃp,∞ p
(Rd ) , (38)
1−θj ,∞
j(2β+2−4α)
 
+d 1−2α+ d 2β+1−2α+ d
Ḃp,∞ k p
(Rd ) = Ḃp,∞ p (Rd ), Ḃp,∞ p
(Rd ) (39)
θj ,∞

for
2α − 1 − d
q
j
k (2β + 2 − 4α) − 1 + 2α
θ0 = , θj = , j = 1, . . . , k − 1.
2β 2β
We thus have
β 1 1
Mu0 u XT ∼
< u
0 1−2α+ d + ess sup t α et 2α |∇|1 u(t) 2β+1−2α+ d et 2α |∇|1 u(t) 1−2α+ d
Ḃp,∞ p 0<t<T Ḃp,∞ p
Ḃp,∞ p

β 1
< u
∼ 0 1−2α+ d + ess sup t α et 2α |∇|1 u(t) 2β+1−2α+ d u XT (40)
Ḃp,∞ p 0<t<T Ḃp,∞ p

and
Mu0 u − Mu0 v XT

k−1
β 1
1−θj
<
∼ ess sup t α et 2α |∇|1 u(t) 2β+1−2α+ d
θ
u XjT u − v XT
0<t<T Ḃp,∞ p
j=0


k−1
β 1
θ j
1−θ
+ ess sup t α et 2α |∇|1 v(t) 2β+1−2α+ d v XT j u − v XT . (41)
0<t<T Ḃp,∞ p
j=0
720 Z.-M. Chen JMFM

Since
β 1 β 1
lim t α et 2α |∇|1 u(t) 2β+1−2α+ d
p
+ lim t α et 2α |∇|1 v(t) 2β+1−2α+ d
p
= 0,
t→0 Ḃp,∞ t→0 Ḃp,∞

the estimates
Mu0 u XT ≤ C u0 1−2α+ d
p
Ḃp,∞

and
1
Mu0 u − Mu0 v XT ≤
u − v XT
2
hold true for a small constant T > 0 and a suitable constant C. Hence there exists a unique solution
u ∈ XT satisfying u = Mu0 u due to the Banach contraction principle.
For the existence of small global solution with T → ∞, we adopt the space
 
1−2α+ d 1
X∞ = u ∈ L∞ (0, ∞; Ḃp,∞ p
(Rd )); u X∞ ≤ u0 2 1−2α+ d
p
Ḃp,∞

for u X∞ expressed in (35) as T → ∞. By (40) and (41), we have the following bounds
1
Mu0 u X∞ ∼
< u
0 1−2α+ d
p
+ u 2X∞ ≤ u0 2 1−2α+ d , u ∈ X∞ ,
Ḃp,∞ p
Ḃp,∞
1
Mu0 u − Mu0 v X∞ ≤ u − v X∞ , u, v ∈ X∞
2
provided that u0 1−2α+ d
p
is sufficiently small. Therefore the global solution exists due to Banach con-
Ḃp,∞
traction principle. The proof of Theorem 2 is complete. 
1−2α+ d
Remark 1. Note that S(Rd ) is not dense in L∞ (Rd ) and hence is not dense in Ḃp,∞ p
(Rd ). That is,
e−t(−Δ) u0 − u0
α
1−2α+ d
p
→ 0 as t → 0
Ḃp,∞

1−2α+ d 1−2α+ d
for some u0 ∈ Ḃp,∞ p
(Rd ). Thus the continuity of the solution in the space Ḃp,∞ p
(Rd ) is generally
1−2α+ d 1−2α+ d
not obtainable. However, if we define the space B̊p,∞ p (Rd ) ⊂ Ḃp,∞ p (Rd ) to be the closure of S(Rd )
1−2α+ d
under the norm of Ḃp,∞ p (Rd ), then the solution u given in Theorem 2 satisfies the continuity property
1−2α+ d 1−2α+ d
u ∈ C([0, T ); B̊p,∞ p
(Rd )) whenever u0 ∈ B̊p,∞ p
(Rd ).
and hence solves the initial value problem defined by (1) and (31).
To show the continuity, we see that the proof of Theorem 1 yields the estimate
β
(∂t + (−Δ)α )−1 P u · ∇u(t) Ḃ σ + t α (∂t + (−Δ)α )−1 P u · ∇u(t) Ḃp,∞
σ+2β
p,∞

β
< ess sup t α uu σ+2β+1−2α .
∼ 0<t<T Ḃp,∞

This together with the proof of (40) using (29) instead of (28) implies that
u(t) − u0 1−2α+ d
p
Ḃp,∞

< e −t(−Δ)α β
∼ u0 − u0 1−2α+ d + ess sup t α u(t) 2β+1−2α+ d u(t) 1−2α+ d
Ḃp,∞ p 0<t<T Ḃp,∞ p
Ḃp,∞ p

for 0 < t < T . Hence we have


< lim e−t(−Δ) u − u
α
lim u(t) − u0 1−2α+ d
p ∼ 0 0 1−2α+ d
p
. (42)
t→0 Ḃp,∞ t→0 Ḃp,∞
Vol. 19 (2017) Analytic Semigroup Approach to Generalized Navier–Stokes Flows in Besov Spaces 721

1−2α+ d 1+ d
Moreover, for small  > 0, the definition of the space B̊p,∞ p
(Rd ) ensures the existence of v0 ∈ Bp,∞p (Rd )
satisfying
u0 − v0 1−2α+ d
p
< .
Ḃp,∞

This yields that


e−t(−Δ) u0 − u0
α
1−2α+ d
p
Ḃp,∞

+ e−t(−Δ) (u0 − v0 ) + e−t(−Δ) v0 − v0


α α
< u − v
∼ 0 0 1−2α+ d
p
1−2α+ d
p
1−2α+ d
p
Ḃp,∞ Ḃp,∞ Ḃp,∞
 t
(−Δ)α e−s(−Δ) v0
α
∼  + u0 − v0
<
1−2α+ d
p
+ 1−2α+ d
p
ds
Ḃp,∞ 0 Ḃp,∞

<  + t v
∼ 0 1+ d .
Ḃp,∞p

The combination of the previous equation and (42) gives the estimate

lim u(t) − u0 1−2α+ d


p
<

t→0 Ḃp,∞

for any  > 0. We thus obtain the continuity of the solution u at the initial state t = 0 and therefore the
continuity of u for t > 0 due to the solution semigroup property of (1).
The Gevrey regularity result for α > 12 obtained in Theorem 2 is not extendable to the critical state
α = 12 as the proof of generalized Hölder estimate (12) is derived from Lp estimate, which is only available
for 1 < p < ∞. When α = 12 , it is necessary to consider the solution in the L∞ (Rd ) space. Therefore we
provide an application of Theorem 1 for α = 12 without involvement of Gevrey regularity.
1
Theorem 3. Let α = 2 and 1 < p < ∞. Then there exists a constant T > 0 such that initial value problem
d
defined by (1) and (31) admits a unique solution u ∈ C([0, T ); B̊p,∞
p
(Rd )) subject to the estimate
 
ess sup t2β u(t) 2β+ d + u(t) d + u(t) L∞ ∼ < u d
0 + u0 L∞ (43)
0<t<T p p
Ḃp,∞ p
Ḃp,∞ Ḃp,∞
d
for 0 < β < 12 , provided that the initial function u0 ∈ B̊p,∞
p
(Rd ) ∩ L∞ (Rd ).

Proof. The proof of Theorem 1 shows that


(∂t + (−Δ)α )−1 P u · ∇u(t) d
p
+ t2β (∂t + (−Δ)α )−1 P u · ∇u(t) 2β+ d
Ḃp,∞ Ḃp,∞ p

+ (∂t + (−Δ)α )−1 P u · ∇u(t) L∞ ∼


< ess sup t2β uu(t)
2β+ d . (44)
0<t<T Ḃp,∞ p

Defining the space


 d

XT = u ∈ C([0, T ); B̊p,∞
p
(Rd )); u XT ≤ C u0 p
d + C u0 L∞
Ḃp,∞

for a constant C and the norm


 
u XT = ess sup t2β u(t) 2β+ d + u(t) d + u(t) L∞
0<t<T Ḃp,∞ p p
Ḃp,∞

and recalling Mu0 u given in (34), we use Eq. (44) to obtain

Mu0 u − e−t(−Δ) u0 XT ∼
< ess sup t2β uu(t)
α
2β+ d . (45)
0<t<T Ḃp,∞ p
722 Z.-M. Chen JMFM

To estimate the right-hand side term of this equation, we employ (30), (38) and (39) to obtain

k−1
uu 2β+ d
<
∼ u j
(2β+ pd ) u Ḃ k−j + u L∞ u 2β+ pd
k (
2β+ d )
Ḃp,∞ p k
Ḃ kp p
Ḃp,∞
j=1 ,∞
kp
,∞
j k−j


k−1
<
∼ u 2βj
+d
u 2β(k−j)
+d
+ u L∞ u 2β+ d
k
Ḃp,∞ p
Ḃp,∞k p Ḃp,∞ p
j=1

< u u + u L∞ u
∼ d
p
2β+ d 2β+ d
Ḃp,∞ Ḃp,∞ p Ḃp,∞ p

for an integer k > 2β + dp . Therefore (45) becomes

Mu0 u − e−t(−Δ) u0 XT ∼
< ess sup t2β u(t)
α
2β+ d u XT . (46)
0<t<T Ḃp,∞ p

This combined with Remark 1 gives the continuity


< lim e−t(−Δ) u − u
α
lim Mu0 u(t) − u0 d
p ∼ 0 0 d
p
=0
t→0 Ḃp,∞ t→0 Ḃp,∞

and the boundedness


< e−t(−Δ) u 2β
α
Mu0 u XT ∼ 0 XT + ess sup t u(t) 2β+ d u XT
0<t<T Ḃp,∞ p
≤ C( u0 p
d + u0 L∞ )
Ḃp,∞

for a suitable choice of the constant C, since limt→0 t2β u(t) 2β+ d = 0. This shows that the operator
Ḃp,∞ p
Mu0 maps the space XT into itself when T > 0 is sufficiently small. Similarly, we also have the contraction
property
1
Mu0 u − Mu0 v XT ≤ u − v XT for u, v ∈ XT .
2
Since XT is a complete metric space, the operator Mu0 has a unique fixed point in XT as a result of the
Banach contraction principle. This completes the proof of Theorem 3. 
It should be mentioned that the well-posedness result described by Theorem 2 is essentially based
d
−d
p
on Theorem 1 and Lemma 6, by which the norm of the negative degree homogeneous space Ḃp,∞q
(Rd )
2β+1−2α+ d
in Lemma 6 can be controlled by those of the positive degree homogeneous space Ḃp,∞ p
(Rd ) and
1−2α+ d 1
the negative degree homogeneous space Ḃp,∞ p (Rd ) in Theorem 1 for 2 < α. For the weak dissipation
p−q
d d
1 d
situation 0 < α < 2, the negative degree homogeneous space Ḃp,∞ (R ) is no longer an interpolation
2β+1−2α+ d 1−2α+ d
space between the two positive degree homogeneous spaces Ḃp,∞ p
(Rd ) and Ḃp,∞ p (Rd ). Thus for
controlling the nonlinear term of (1), the Hölder-like estimate in Lemma 6 should be improved. To do so,
it is useful to consider the Hölder-like estimate involving derivatives as related to commutator estimates
[8,14,15,21].

Acknowledgements. The research is partially supported by the NSFC (Grant No. 11571240) of China.

References

[1] Bae, H., Biswas, A.: Gevrey regularity for a class of dissipative equations with analytic nonlinearity. arXiv:1403.1603v1
[math]
[2] Biswas, A.: Gevrey regularity for the supercritical quasi-geostrophic equation. J. Differ. Equ. 257, 1753–1772 (2014)
Vol. 19 (2017) Analytic Semigroup Approach to Generalized Navier–Stokes Flows in Besov Spaces 723

[3] Cannone, M.: Harmonic analysis tools for solving the incompressible Navier–Stokes equations. In: Handbook Mathe-
matical Fluid Dynamics, Chap. 3, pp. 161–244 (2004)
[4] Cao, C., Rammaha, M.A., Titi, E.S.: The Navier–Stokes equations on the rotating 2-D sphere: Gevrey regularity and
asymptotic degrees of freedom. Z. Angew. Math. Phys. 50, 341–360 (1999)
[5] Chae, D., Constantin, P., Wu, J.: Inviscid models generalizing the 2D Euler and the surface quasi-geostrophic equations.
Arch. Rat. Mech. Anal. 202, 35–62 (2011)
[6] Chae, D., Constantin, P., Wu, J.: Dissipative models generalizing the 2D Navier–Stokes and the surface quasi-geostrophic
equations. Indiana Univ. Math. J. 61, 1997–2018 (2012)
[7] Charney, J.G., DeVore, J.G.: Multiple flow equilibria in the atmosphere and blocking. J. Atmos. Sci. 36, 1205–1216
(1979)
[8] Chen, Q., Miao, C., Zhang, Z.: A new Bernsteins inequality and the 2D dissipative quasi-geostrophic equation. Commun.
Math. Phys. 271, 821–838 (2007)
[9] Chen, Z.-M.: Steady-state bifurcation analysis of a strong nonlinear atmospheric vorticity equation. J. Math. Anal.
Appl. 431, 1–21 (2015)
[10] Chen, Z.-M.: Bifurcating steady-state solutions of the dissipative quasi-geostrophic equation in Lagrangian formulation.
Nonlinearity 29, 3132–3147 (2016)
[11] Chen, Z.-M., Xin, Z.: Homogeneity criterion for the Navier–Stokes equations in the whole spaces. J. Math. Fluid Mech.
3, 152–182 (2001)
[12] Chen, Z.-M., Xiong, X.: Equilibrium states of the Charney–DeVore quasi-geostrophic equation in mid-latitude
atmosphere. J. Math. Anal. Appl. 444, 1403–1416 (2016)
[13] Constantin, P., Wu, J.: Behavior of solutions of 2D quasi-geostrophic equations. SIAM J. Math. Anal. 30, 937–948
(1999)
[14] Danchin, R.: Global existence in critical spaces for compressible Navier–Stokes equations. Invent. Math. 141, 579–614
(2000)
[15] Danchin, R.: Density-dependent incompressible viscous fluids in critical spaces. Proc. R. Soc. Edinburgh A 133, 1311–
1334 (2003)
[16] Dong, H.: Dissipative quasi-geostrophic equations in critical Sobolev spaces: smoothing effect and global well-posedness.
Discrete Contin. Dyn. Syst. 26, 1197–1211 (2010)
[17] Dong, H., Li, D.: Spatial analyticity of the solutions to the subcritical dissipative quasi-geostrophic equations. Arch.
Rat. Mech. Anal. 189, 131–158 (2008)
[18] Foias, C., Temam, R.: Gevrey class regularity for the solutions of the Navier–Stokes equations. J. Funct. Anal. 87,
359–369 (1989)
[19] Kato, T.: On classical solutions of the two-dimensional non-stationary Euler equation. Arch. Rat. Mech. Anal. 25,
188–200 (1967)
[20] Kato, T.: Strong Lp -solutions of the Navier–Stokes equation in Rm , with applications to weak solutions. Math. Z. 187,
471–480 (1984)
[21] Kato, T., Ponce, G.: Commutator estimates and the Euler and Navier–Stokes equations. Commun. Pure Appl. Math.
XLI, 891–907 (1988)
[22] Katz, N.H., Pavlovi, N.: A cheap Caffarelli–Kohn–Nirenberg inequality for the Navier–Stokes equation with hyper-
dissipation. Geom. Funct. Anal. 355–379 (2002)
[23] Kiselev, A.: Some recent results on the critical surface quasi-geostrophic equation: a review. Hyperbolic problems: theory,
numerics and applications. In: Proc. Sympos. Appl. Math., vol. 67, Part 1. AMS, Providence, pp. 105–122 (2009)
[24] Kiselev, A., Nazarov, F., Volberg, A.: Global well-posedness for the critical 2D dissipative quasi- geostrophic equation.
Invent. Math. 167, 445–453 (2007)
[25] Ladyzhenskaya, O.: The Mathematical Theory of Viscous Incompressible Flow. Gordon & Breach, New York (1969)
[26] Lemarié-Rieusset, G.: Recent Developments in the Navier–Stokes Problem. Chapman & Hall/CRC, London (2002)
[27] Leray, J.: Sur le mouvement d’un liquide visqueux emplissant l’espace. Acta Math. 63, 193–248 (1934)
[28] Li, P., Zhai, Z.: Well-posedness and regularity of generalized Navier–Stokes equations in some critical Q-spaces. J. Funct.
Anal. 259, 2457–2519 (2010)
[29] Nirenberg, L.: On elliptic partial differential equations. Ann. Sc. Norm. Sup. Pisa 13, 123–131 (1959)
[30] Stein, E.: Singular Integrals and Differentiability Properties of Functions. Princeton University Press, Princeton (1970)
[31] Tao, T.: Global regularity for a logarithmically supercritical hyper-dissipative Navier–Stokes equation. Anal. PDE 2,
361–366 (2009)
[32] Tao, T.: Finite time blowup for an averaged three-dimensional Navier–Stokes equation. J. Am. Math. Soc. 29, 601–674
(2016)
[33] Triebel, H.: Theory of Function Spaces. Birkhäuser Verlag, Basel (1983)
[34] Triebel, H.: Theory of Function Spaces II. Birkhäuser Verlag, Basel (1992)
[35] Wolansky, G.: Existence uniqueness and stability of stationary barotropic flow with forcing and dissipation. Commun.
Pure Appl. Math. 41, 19–46 (1988)
[36] Wu, J.: Quasi-geostrophic-type equations with initial data in Morrey spaces. Nonlinearity 10, 1409–1420 (1997)
[37] Wu, J.: Generalized MHD equations. J. Differ. Equ. 195, 284–312 (2003)
[38] Wu, J.: The generalized incompressible Navier–Stokes equations in Besov spaces. Dyn. Partial Differ. Equ. 1, 381–400
(2004)
724 Z.-M. Chen JMFM

[39] Wu, J.: Lower bounds for an integral involving fractional Laplacians and the generalized Navier–Stokes equations in
Besov spaces. Commun. Math. Phys. 263, 803–831 (2005)
[40] Xiao, J.: Homothetic variant of fractional Sobolev space with application to Navier–Stokes system. Dyn. Partial Differ.
Equ. 2, 227–245 (2007)
[41] Yudovich, V.: Nonstationary flow of an ideal incompressible liquid. Zhurn. Vych. Mat. 3, 1032–1066 (1963)

Zhi-Min Chen
Ship Science
University of Southampton
Southampton SO16 7QF
UK
email: zhimin@soton.ac.uk
and
School of Mathematics and Statistics
Shenzhen University
Shenzhen 518052
China

(accepted: October 18, 2016; published online: November 2, 2016)

Potrebbero piacerti anche