March 1, 2005
.
Global Well–posedness of the Three-dimensional Viscous Primitive Equations of Large Scale Ocean and Atmosphere Dynamics
Chongsheng Cao
Edriss S. Titi
(C. Cao) Department of Mathematics, Florida International University, University Park, Miami, FL 33199, USA. E-mail address : caoc@fiu.edu (E.S. Titi) Department of Mathematics, and Department of Mechanical and Aerospace Engineering, University of California, Irvine, CA 92697-3875, USA. Also: Department of Computer Science and Applied Mathematics, Weizmann Institute of Science, Rehovot 76100, Israel. E-mail address : etiti@math.uci.edu E-mail address : edriss.titi@weizmann.ac.il
-
Abstract.
In this paper we prove the global existence and uniqueness (regularity) of strong solutions to the three-dimensional viscous primitive equations, which model large scale ocean and atmosphere dynamics.
MSC Subject Classifications: 35Q35, 65M70, 86-08,86A10.
1 Introduction
Large scale dynamics of oceans and atmosphere is governed by the primitive equations which are derived from the Navier–Stokes equations, with rotation, coupled to thermodynamics and salinity diffusion-transport equations, which account for the buoyancy forces and stratification effects under the Boussinesq approximation.
Moreover, and due to the shallowness of the oceans and the atmosphere, i.e., the depth of the fluid layer is very small in comparison to the radius of the earth, the vertical large scale motion in the oceans and the atmosphere is much smaller than the horizontal one, which in turn leads to modeling the vertical motion by the hydrostatic balance. As a result one obtains the system ( 1 )–( 4 ), which is known as the primitive equations for ocean and atmosphere dynamics (see, e.g., [17] ,[18] ,[19] , [20] and references therein). We observe that in the case of ocean dynamics one has to add the diffusion-transport equation of the salinity to the system ( 1 )–( 4 ). We omitted it here in order to simplify our mathematical presentation. However, we emphasize that our results are equally valid when the salinity effects are taking into account.
Let us remark the horizontal motion can be further approximated by the geostrophic balance when the Rossby number (the ratio of the horizontal acceleration to the Coriolis force) is very small. By taking advantage of these assumptions and other geophysical considerations several intermediate models have been developed and used in numerical studies of weather prediction and long-time climate dynamics (see, e.g., [4] , [7] , [8] , [19] , [20] , [21] , [24] , [25] , [26] , [27] and references therein). Some of these models have also been the subject of analytical mathematical study (see, e.g., [2] , [3] , [5] , [6] , [10] , [11] , [12] , [14] , [22] , [23] , [29] and references therein).
In this paper we will focus on the
primitive equations in a cylindrical domain
where
is a smooth bounded domain in
:
|
(1)
|
|
(4)
|
where the horizontal velocity field
, the velocity field
, the temperature
and the pressure
are the unknowns.
is the Coriolis parameter,
is a given heat source. The viscosity and the heat diffusion operators
and
are given by
|
(5)
|
|
(6)
|
where
are positive constants representing the horizontal and vertical Reynolds numbers, respectively, and
are positive constants which stand for the horizontal and vertical heat diffusivity, respectively. We set
to be the horizontal gradient operator and
to be the horizontal Laplacian. We partition the boundary of
into:
|
(7)
|
|
(8)
|
|
(9)
|
We equip the system ( 1 )–( 4 ) with the following boundary conditions – with wind–driven on the top surface and non-slip and non-heat flux on the side walls and bottom (see, e.g., [17] , [18] , [19] , [21] , [24] ,[25] , [26] ):
|
(10)
|
|
(11)
|
|
(12)
|
where
is the wind stress on ocean surface,
is the normal vector of
, and
is typical temperature distribution of the top surface of the ocean. For simplicity we assume here that
and
are time independent. However, the results presented here are equally valid when these quantities are time dependent and satisfy certain bounds in space and time.
Due to the boundary conditions ( 10 )–( 12 ), it is natural to assume that
and
satisfy the compatibility boundary conditions:
In addition, we supply the system with the initial condition:
|
(15)
|
|
(16)
|
In [17] and [18] the authors set up the mathematical framework to study the viscous primitive equations for the atmosphere and ocean circulation. Moreover, similar to the
Navier–Stokes equations, they have shown the global existence of weak solutions (without uniqueness) and the short time existence and uniqueness of strong solutions of the viscous primitive equations model. In [14] the authors proved the global existence and uniqueness of strong solutions to the viscous primitive equations in thin domains for a large set of initial data whose size depends inversely on the thickness of the domain. In this paper we show the global existence and uniqueness (regularity) of the strong solutions to the
viscous primitive equation model ( 1 )–( 16 ) in general cylindrical domain,
, and initial data. As in the case of
Navier–Stokes equations the question of uniqueness of the weak solutions to this model is still open.
2 Preliminaries
2.1 New Formulation
First, let us reformulate the system ( 1 )–( 16 ). We integrate the equation ( 3 ) in the
direction to obtain
By virtue of ( 10 ) and ( 11 ) we have
|
(17)
|
and
|
(18)
|
We denote by
|
(19)
|
In particular,
|
(20)
|
We will denote the fluctuation by
Notice that
From the above and ( 12 ) we obtain
|
(23)
|
and
|
(24)
|
By integrating equation ( 2 ) we obtain
Substitute ( 17 ) and the above relation into equation ( 1 ) we reach
| |
|
(25)
|
Remark 1.
Notice that due to the compatibility boundary conditions ( 13 ) and ( 14 ) one can convert the boundary condition ( 10 )–( 12 ) to be homogeneous by replacing
by
while ( 23 ) still true. For simplicity and without loss generality we will assume that
However, we emphasize that our results are still valid for general
and
provided they are smooth enough. In a forthcoming paper we will study the long-time dynamics and global attractors to the primitive equations with general
and
.
Therefore, under the assumption that
, we have the following new formulation for system ( 1 )–( 16 ):
| |
|
(26)
|
|
(27)
|
|
(28)
|
|
(29)
|
|
(30)
|
|
(31)
|
2.2 Properties of
and
By taking the average of equations ( 26 ) in the
direction, over the interval
, and using the boundary conditions ( 28 ), we obtain
| |
|
(32)
|
As a result of ( 22 ), ( 23 ) and integration by parts we have
|
(33)
|
By subtracting ( 32 ) from ( 26 ) and using ( 33 ) we get
| |
|
(34)
|
Therefore,
satisfies the following equations and boundary conditions:
| |
|
(35)
|
|
(37)
|
and
satisfies the following equations and boundary conditions:
| |
|
(38)
|
|
(39)
|
Remark 2.
We recall that by virtue of the maximum principle one is able to show the global well-posedness of the
viscous Burgers equations (see, for instance, [
16]
and references therein). Such an argument, however, is not valid for the
Navier–Stokes equations because of the pressure term. Remarkably, the pressure term is absent from equation ( 38 ). This fact allows us to obtain a bound for the
norm of
, which is a key estimate in our proof of the global regularity for the system ( 1 )–( 16 ).
2.3 Functional spaces and Inequalities
Let us denote by
and
the usual
Lebesgue and Sobolev spaces, respectively ([1] ). We denote by
|
(40)
|
Let
| |
| |
Let
and
be the closure spaces of
in
, and
in
under
topology, respectively.
Definition 1.
Let
and let
be a fixed positive time.
is called a weak solution of ( 26 )–( 31 ) on the time interval
if it satisfies ( 26 ) and ( 27 ) in weak sense, and also
| |
| |
| |
| |
where
and
are the dual spaces of
and
, respectively. Moreover, if
and
, a weak solution is called strong solution of ( 26 )–( 31 ) on
if, in addition, it satisfies
| |
| |
For convenience, we recall the following Sobolev and Ladyzhenskaya's inequalities in
([1] , [9] , [13] , [15] )
|
(41)
|
|
(42)
|
for every
and the following Sobolev and Ladyzhenskaya's inequalities in
([1] , [9] , [13] , [15] )
|
(43)
|
|
(44)
|
for every
Here
is a positive constant which might depend on the shape of
and
but not on their size. And we recall the Poincaré inequalities
|
(45)
|
|
(46)
|
| |
|
(47)
|
Here
is the area of
. Moreover, by ( 41 ) we get
| |
|
(48)
|
for every
Also, we recall the integral version of Minkowsky inequality for the
spaces,
. Let
and
be two measurable sets, where
and
are two positive integers.
Suppose that
is measurable over
. Then,
|
(49)
|
3 A Priori estimates
In previous subsections we have reformulated the system ( 1 )–( 16 ) to be equivalent to the system ( 26 )–( 31 ). In this section we will get some useful a priori estimates about the solutions of ( 26 )–( 31 ).
The estimates presented here are formal. One can justify them rigorously by establishing them first for a Galerkin approximating system and by using the appropriate “Compactness Theorems” (see, for example, [9] ,[17] ,[18] ,[28] ) to accomplish them for the existence of solutions and the corresponding a priori estimates.
3.1
estimates
By taking the inner product of equation ( 27 ) with
, in
, we get
| |
| |
By integration by parts we get
|
(50)
|
As a result of the above we conclude
| |
| |
By ( 47 ) and Cauchy–Schwarz inequality we obtain
|
(51)
|
|
(52)
|
By ( 47 ) and thanks to Gronwall inequality we get
| |
|
(53)
|
By taking the inner product of equation ( 26 ) with
, in
, we reach
| |
| |
| |
By integration by parts we get
|
(54)
|
By ( 36 ) we have
|
(55)
|
By direct calculation we obtain
Thus, by ( 54 )–( 56 ) we have
| |
| |
By Cauchy–Schwarz and ( 53 ) we obtain
| |
| |
By ( 45 ) and ( 46 ) and thanks to Gronwall inequality we get
| |
| |
|
(57)
|
Therefore, by ( 53 ) and ( 57 ) we have
| |
|
(58)
|
where
| |
|
(59)
|
3.2
estimates
Taking the inner product of the equation ( 38 ) with
in
, we get
| |
| |
| |
By integration by parts we get
|
(60)
|
By direct calculation we obtain
|
(61)
|
By ( 36 ) and boundary condition ( 28 ) we have
|
(62)
|
Thus, by ( 60 )–( 62 ) we have
| |
| |
| |
Notice that by integration by parts and boundary condition ( 28 ) we have
| |
| |
| |
| |
Therefore, by Cauchy–Schwarz inequality and Hölder inequality we obtain
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
By using Minkowsky inequality ( 49 ), we get
| |
By ( 48 ),
| |
Thus, by Cauchy–Schwarz inequality we obtain
|
(63)
|
Similarly, by ( 49 ) and ( 42 ), we also get
| |
|
(64)
|
and
| |
|
(65)
|
Therefore, by ( 63 )–( 65 ) and ( 41 ), we reach
| |
| |
| |
| |
By Young's inequality and Cauchy–Schwarz inequality we have
| |
| |
By ( 58 ) and Gronwall inequality, we get
|
(66)
|
where
|
(67)
|
Taking the inner product of the equation ( 27 ) with
in
, and by ( 27 ), we get
| |
| |
By integration by parts and ( 36 we get
|
(68)
|
As a result of the above we conclude
| |
| |
By Gronwall, again, we get
|
(69)
|
3.3
estimates
3.3.1
estimates
By taking the inner product of equation ( 35 ) with
in
, and applying ( 36 ), we reach
| |
Following the proof to
Navier–Stokes equations (cf. e.g., [9] ) one would get
| |
By Cauchy–Schwarz inequality and Hölder inequality, we get
| |
| |
| |
| |
Thus, by Young's inequality and Cauchy–Schwarz inequality, we have
| |
By ( 58 ), ( 66 ) and thanks to Gronwall inequality and we obtain
|
(70)
|
where
|
(71)
|
3.3.2
estimates
Denote by
It is clear that
satisfies
| |
|
(72)
|
Taking the inner product of the equation ( 72 ) with
in
and using the boundary condition ( 28 ), we get
| |
| |
| |
By integration by parts we get
|
(73)
|
By direct calculation we obtain
Thus, by ( 73 ) and ( 74 ) we have
| |
| |
| |
| |
| |
By Young's inequality and Cauchy–Schwarz inequality, we have
| |
| |
| |
By ( 58 ), ( 66 ), ( 70 ), and Gronwall inequality, we get
|
(75)
|
where
|
(76)
|
3.3.3
estimates
By taking the inner product of equation ( 26 ) with
in
, we reach
| |
| |
| |
| |
| |
Notice that by applying the Proposition 2.2 in [5] with
and
, we get
| |
As a result and by ( 43 ) and ( 44 ), we obtain
| |
| |
Thus, by Young's inequality and Cauchy–Schwarz inequality, we have
| |
| |
By ( 58 ), ( 66 ), ( 70 ), ( 75 ) and thanks to Gronwall inequality, we obtain
|
(77)
|
where
|
(78)
|
3.3.4
estimates
Taking the inner product of the equation ( 27 ) with
in
, we get
| |
| |
| |
| |
| |
| |
| |
| |
By Young's inequality and Cauchy–Schwarz inequality we have
| |
| |
| |
By ( 66 ), ( 77 ), and Gronwall inequality, we get
| |
|
(79)
|
where
|
(80)
|
4 Existence and Uniqueness of the Strong Solutions
In previous sections we have reformulated the system ( 1 )–( 16 ) and got a priori estimates ( 58 )–( 79 ). In this section we show the global existence and uniqueness of the strong solution of the system ( 26 )–( 31 ).
Theorem 2.
Suppose that
. Then for every
,
and
there is a unique strong solution
of the system ( 26 )–( 31 ), which depends continuously on the initial data.
Remark: The steps of the following proof are formal in the sense that they can be made more rigorous by proving the corresponding estimates first for the Galerkin approximation system. Then the estimates for the exact solution can be established by passing to the limit in the Galerkin procedure by using the appropriate “Compactness Theorems” (see, for example, [9] ,[17] , [18] , and [28] for details).
-
Proof.
One can show the existence of the strong solution based on the a priori estimates ( 70 ), ( 75 ), ( 77 ) and ( 79 ). Here we only show the continuous dependence on the initial data and the the uniqueness of the strong solutions.
Let
and
be two strong solutions of the system ( 26 )–( 31 ) with initial data
and
, respectively. Denote by
It is clear that
| |
|
(81)
|
| |
|
(82)
|
|
(83)
|
|
(84)
|
By taking the inner product of equation ( 81 ) with with
in
, and equation ( 82 ) with
, in
we get
| |
| |
| |
and
| |
| |
By integration by parts, and the boundary conditions ( 28 ) and ( 29 ), we get
|
(85)
|
|
(86)
|
By direct calculation we obtain
|
(87)
|
Thus, by ( 85 ), ( 86 ) and ( 87 ) we have
| |
| |
and
| |
| |
Notice that
|
(88)
|
|
(89)
|
Moreover,
| |
| |
| |
By Cauchy–Schwarz inequality, we get
|
(90)
|
By using Minkowsky inequality ( 49 ) and ( 41 ), we obtain
| |
|
(91)
|
and
| |
|
(92)
|
Similarly, we have
|
(93)
|
Therefore, by estimates ( 88 )–( 93 ), we reach
| |
| |
| |
By Young's inequality, we get
| |
Thanks to Gronwall inequality, we obtain
| |
| |
Since
is a strong solution, we have
| |
The above inequality proves the continuous dependence of the solutions on the initial data, and in particular, when
, we have
for all
. Therefore, the strong solution is unique.
□
Acknowledgments
This work was supported in part by the NSF grant No. DMS–0204794, the MAOF Fellowship of the Israeli Council of Higher Education, and by the USA Department of Energy, under contract number W–7405–ENG–36 and ASCR Program in Applied Mathematical Sciences. References
-
R.A. Adams, Sobolev Spaces, Academic Press, New York, 1975.
-
A. Babin, A. Mahalov and B. Nicolaenko, Fast singular oscillating limits and global regularity for the
primitive equations of geophysics, Math. Model. Numer. Anal. 34 (2000), 201–222.
-
V. Barcilon, P. Constantin and E.S. Titi, Existence of solutions to the Stommel-Charney model of the Gulf Stream, SIAM J. Math. Anal. 19 (1988), 1355–1364.
-
G. Browning, A. Kasahara, H.-O. Kreiss, Initialization of the primitive equations by the bounded derivative method, J. Atmospheric Sci. 37 (1980), 1424–136.
-
C. Cao and E.S. Titi, Global Well–posedness and Finite Dimensional Global Attractor for a 3–D Planetary Geostrophic Viscous Model, Comm. Pure Appl. Math. 56, 198-233, 2003
-
C. Cao, E.S. Titi and M. Ziane, A “horizontal” hyper–diffusion
thermocline planetary geostrophic model: well-posedness and long time behavior, Nonlinearity 17, 1749–1776, 2004.
-
J.G. Charney, The use of the primitive equations of motion in numerical prediction, Tellus 7 (1955), 22–26.
-
J.G. Charney, The gulf stream as an inertial boundary layer, Proc. Nat. Acad. Sci. U.S.A., 41(1955), 731–740.
-
P. Constantin and C. Foias, Navier-Stokes Equations, The University of Chicago Press, 1988.
-
P. Constantin, A. Majda and E. Tabak, Formation of strong fronts in the
-D quasigeostrophic thermal active scalar, Nonlinearity 7 (1994), 1495–1533.
-
P. Constantin, A. Majda and E. Tabak, Singular front formation in a model for quasigeostrophic flow, Phys. Fluids 6 (1994), 9–11.
-
P.F. Embid and A.J. Majda, Low Froude number limiting dynamics for stably stratified flow with small or finite Rossby numbers, Geophys. Astrophys. Fluid Dynam. 87 (1998), 1–50.
-
G.P. Galdi, An Introduction to the Mathematical Theory of the Navier-Stokes Equations, Vol. I & II, Springer-Verlag, 1994.
-
C. Hu, R. Temam and M. Ziane, The primitive equations on the large scale ocean under the small depth hypothesis, Discrete Contin. Dyn. Syst. 9 (2003), 97–131.
-
O.A. Ladyzhenskaya, The Boundary Value Problems of Mathematical Physics, Springer-Verlag, 1985.
-
Ladyženskaja, O. A., Solonnikov, V. A. and Ural'ceva, N. N., Linear and Quasilinear Equations of Parabolic Type, American Mathematical Society, Providence, R.I., (1967).
-
J.L. Lions, R. Temam and S. Wang, New formulations of the primitive equations of atmosphere and applications, Nonlinearity 5 (1992), 237–288.
-
J.L. Lions, R. Temam and S. Wang, On the equations of the large scale Ocean, Nonlinearity 5 (1992), 1007–1053.
-
J. Pedlosky, Geophysical Fluid Dynamics, Springer-Verlag, New York, 1987.
-
L.F. Richardson, Weather Prediction by Numerical Process, Cambridge University Press, Cambridge 1922 (reprint, Dover, New York, 1988).
-
R. Samelson, Coastal boundary conditions and the baroclinic structure of wind–driven continental shelf currents, Journal of Physical Oceanography, 27(1997), 2645-2662.
-
R. Samelson, R. Temam and S. Wang, Some mathematical properties of the planetary geostrophic equations for large scale ocean circulation, Applicable analysis, 70(1998), 147-173.
-
R. Samelson, R. Temam and S. Wang, Remarks on the planetary geostrophic model of gyre scale ocean circulation, Differential and Integral Equation, 13 (2000), 1–14.
-
R. Samelson and G. Vallis, A simple friction and diffusion scheme for planetary geostrophic basin models, J. Phys. Oceanogr., 27 (1997), 186–194
-
R. Samelson and G. Vallis, Large-scale circulation with small diapycnal diffusion: The two-thermocline limit, J. Marine Res., 55 (1997), 223–275
-
D. Seidov, An intermediate model for large–scale ocean circulation studies, Dynamics of Atmospheres and Oceans, 25(1996), 25-55.
-
H. Stommel, The westward intensification of wind–driven ocean currents, Trans. Amer. Geophys. union, 29(1948), 291–304.
-
R. Temam, Navier-Stokes Equations, Theory and Numerical Analysis, North-Holland, 1984.
-
R. Temam and M. Ziane, Some mathematical problems in geophysical fluid dynamics, Handbook of mathematical fluid dynamics, 2003
(C. Cao) Department of Mathematics, Florida International University, University Park, Miami, FL 33199, USA. E-mail address : caoc@fiu.edu (E.S. Titi) Department of Mathematics, and Department of Mechanical and Aerospace Engineering, University of California, Irvine, CA 92697-3875, USA. Also: Department of Computer Science and Applied Mathematics, Weizmann Institute of Science, Rehovot 76100, Israel. E-mail address : etiti@math.uci.edu E-mail address : edriss.titi@weizmann.ac.il