Next Article in Journal
On Implicit Time–Fractal–Fractional Differential Equation
Next Article in Special Issue
On (k,ψ)-Hilfer Fractional Differential Equations and Inclusions with Mixed (k,ψ)-Derivative and Integral Boundary Conditions
Previous Article in Journal
Dynamic Physical Activity Recommendation Delivered through a Mobile Fitness App: A Deep Learning Approach
Previous Article in Special Issue
Controllability of a Class of Impulsive ψ-Caputo Fractional Evolution Equations of Sobolev Type
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Existence of Weak Solutions for the Vorticity Equation Related to the Stratosphere in a Rotating Spherical Coordinate System

1
School of Mathematics and Statistics, Liupanshui Normal University, Liupanshui 553004, China
2
Department of Mathematics, Guizhou University, Guiyang 550025, China
3
Department of Mathematical Analysis and Numerical Mathematics, Faculty of Mathematics, Physics and Informatics, Comenius University in Bratislava, Mlynská dolina, 842 48 Bratislava, Slovakia
4
Mathematical Institute, Slovak Academy of Sciences, Štefánikova 49, 814 73 Bratislava, Slovakia
*
Author to whom correspondence should be addressed.
Axioms 2022, 11(7), 347; https://doi.org/10.3390/axioms11070347
Submission received: 15 June 2022 / Revised: 12 July 2022 / Accepted: 16 July 2022 / Published: 20 July 2022
(This article belongs to the Special Issue Impulsive, Delay and Fractional Order Systems)

Abstract

:
In this paper, based on the Euler equation and mass conservation equation in spherical coordinates, the ratio of the stratospheric average width to the planetary radius and the ratio of the vertical velocity to the horizontal velocity are selected as parameters under appropriate boundary conditions. We establish the approximate system using these two small parameters. In addition, we consider the time dependence of the system and establish the governing equations describing the atmospheric flow. By introducing a flow function to code the system, a nonlinear vorticity equation describing the planetary flow in the stratosphere is obtained. The governing equations describing the atmospheric flow are transformed into a second-order homogeneous linear ordinary differential equation and a Legendre’s differential equation by applying the method of separating variables based on the concepts of spherical harmonic functions and weak solutions. The Gronwall inequality and the Cauchy–Schwartz inequality are applied to priori estimates for the vorticity equation describing the stratospheric planetary flow under the appropriate initial and boundary conditions. The existence and non-uniqueness of weak solutions to the vorticity equation are obtained by using the functional analysis technique.

1. Introduction

With the high complexity caused by factors such as the rotation of the Earth and heat input, the observed air flows in the atmosphere can be regarded as a description of large-scale motions (see [1]). For large-scale atmospheric flows, consideration of the full nature of the geometry of the Earth’s curved space is essential (see [2]). In order to transcend the limitations of the plane geometry of the f-plane approximation, the classical approach is to invoke the weak contribution of the curvature by using the β -plane approximation (see [3]). However, in contrast to the f-plane approximation, the β -plane approximation does not represent a consistent approximation of the geophysical flow governing equations in the mid-latitude and polar regions (see [4]). Therefore, it is necessary to establish the governing equation of the fluid in the spherical coordinate system.
Constantin and Johnson gave the motion control equation and the mass conservation equation in the rotating spherical coordinate system. The thin-layer asymptotic approximation was established based on the ratio of the ocean average depth to the Earth’s radius (see [5]). Martin established the exact solution of the governing equations of geophysical fluid dynamics involving discontinuous stratification in spherical coordinates (see [6,7]). The general problem of the ocean on a rotating sphere was studied. The exact solution of incompressible (constant density) inviscid fluid with velocity distribution below and along the surface was established. This can be regarded as a model of the Antarctic Circumpolar Current (see [8]). The spherical coordinate governing the equations of inviscid incompressible fluid fixed at a point on the rotating Earth and the free surface and rigid bottom boundary conditions were introduced. The exact solution of the system was given, which described a steady flow that moves only in the azimuth direction (see [9]). Using the fixed point method, Chu established the existence of strictly monotone bounded solutions for a given continuous vorticity. These results were related to the behavior of ocean flows in arctic gyres (see [10]). Wang et al. proposed a non-local formula for simulating the Antarctic Circumpolar Current without considering the vertical motion, which coded the horizontal flow components by introducing flow functions. Using the topological degree, zero exponent theory and fixed point technique, the existence of positive solutions for nonlinear vorticity, nonlocal boundary value problems was proven (see [11]). Haziot used the Mercator projection to map the circulation model from the sphere to the plane and obtained the boundary value problem of semilinear elliptic partial differential equations. For the constant and linear ocean vorticity, he studied the existence, regularity and uniqueness of solutions to this elliptic problem. The physical correlation of these results was also investigated (see [12]). Using spherical coordinates, Martin and Quirchmayr derived a new exact solution to the governing equations of geophysical fluid dynamics for an inviscid and incompressible fluid with a general density distribution and a forced term. Their explicit solution represents a steady purely azimuthal stratified flow with a free surface (see [13,14]). In addition, Constantin and Johnson showed that a consistent shallow-water approximation of the incompressible Navier–Stokes equation written in a spherical, rotating coordinate system produces, at the leading order in a suitable limiting process, a general linear theory for wind-induced ocean currents which reaches beyond the limitations of the classical Ekman spiral (see [15,16]).
In this paper, we consider the spherical coordinates from the Euler equation and mass conservation equation, combined with the appropriate boundary conditions, and choose the ratio of the average width of the stratosphere and planetary radius and the ratio between the vertical velocity and horizontal velocity as parameters. We utilize the two small parameters to establish the approximate system. Furthermore, we consider adding time dependence into the system and establish the governing equations to describe the atmospheric flow. A nonlinear vorticity equation describing stratospheric planetary flow is obtained by introducing a flow function to code the system. Based on the idea of spherical harmonic function and the concept of a weak solution, the existence and non-uniqueness of a weak solution to the vorticity equation are obtained under suitable boundaries.

2. Stratospheric Planetary Flows Equation

In this section, we introduce the rotating sphere coordinate system ( β , α , r ¯ ) , with β [ 0 , 2 π ] and α [ π 2 , π 2 ] representing the longitude and latitude angles, respectively, and r ¯ representing the distance from the primordial center to the center of the planet. ( e β , e α , e r ¯ ) represent the unit vectors of ( β , α , r ¯ ) in the three directions, respectively, and the corresponding velocity components are ( u ¯ , v ¯ , w ¯ ) , where e β points from west to east, e α points from south to north, and e r ¯ points out from the origin.
In the stratosphere, atmospheric currents can be considered inviscid (see [17]). Under the influence of the Coriolis force, the Euler equation can be determined by the following components (see [18]):
u ¯ t ¯   +   u ¯ r ¯ cos α u ¯ β   +   v ¯ r ¯ u ¯ α   +   w ¯ u ¯ r ¯   +   u ¯ w ¯ u ¯ v ¯ tan α r ¯     2 Ω ¯ ( v ¯ sin α w ¯ cos α )   =   1 ρ ¯ r ¯ cos α p ¯ β , v ¯ t ¯   +   u ¯ r ¯ cos α v ¯ β   +   v ¯ r ¯ v ¯ α   +   w ¯ v ¯ r ¯   +   v ¯ w ¯ + u ¯ 2 tan α r ¯   +   2 Ω ¯ u ¯ sin α   +   Ω ¯ 2 r ¯ sin α cos α   =   1 ρ ¯ r ¯ p ¯ α , w ¯ t ¯   +   u ¯ r ¯ cos α w ¯ β   +   v ¯ r ¯ w ¯ α   +   w ¯ w ¯ r ¯     u ¯ 2 + v ¯ 2 r ¯     2 Ω ¯ u ¯ cos α     Ω ¯ 2 r ¯ cos 2 α   =   1 ρ ¯ p ¯ r ¯     g ¯ ,
where p ¯ and ρ ¯ represent the atmospheric pressure and density, respectively, Ω ¯ is the constant rotation rate of the planet, and g ¯ stands for the acceleration of gravity.
The mass conservation equation in spherical coordinates can be expressed as
ρ ¯ t ¯   +   u ¯ r ¯ cos α ρ ¯ β   +   v ¯ r ¯ ρ ¯ α   +   w ¯ ρ ¯ r ¯   +   ρ ¯ 1 r ¯ cos α u ¯ β + α ( v ¯ cos α ) + 1 r ¯ 2 r ¯ ( r ¯ 2 w ¯ )   =   0 .
The inverse of the Rossby number is expressed as shown below:
μ   =   Ω ¯ R ¯ U ¯ ,
where R ¯ is the radius of the planet and U ¯ is the horizontal velocity scale.
The shallowness parameter κ and the ratio ϱ between the vertical scale W ¯ and horizontal velocity scale U ¯ can be given as follows:
κ   =   H ¯ R ¯ and ϱ   =   W ¯ U ¯ ,
Here, H ¯ is the mean width of the stratosphere. Based on the persistent large-scale circulation model of the stratosphere, we can give corresponding reference values for different dimensional scales (see [19,20,21]). The reference values are shown in Table 1.
We now transform the original variables to receive the dimensionless version of all the variables as follows:
( u ¯ , v ¯ , w ¯ )   =   ( U ¯ u , U ¯ v , W ¯ w ) , t ¯   =   R ¯ U ¯ t , r ¯   =   R ¯   +   H ¯ z , ρ ¯   =   ρ ˜ ρ , p ¯   =   ρ ˜ U ¯ 2 p .
Therefore, the dimensionless Euler Equation (1) is transformed into
u t + u ( 1 + κ z ) cos α u β   +   v 1 + κ z u α   +   ϱ κ w u z   +   ϱ u w u v tan α 1 + κ z 2 μ ( v sin α ϱ w cos α )   =   1 ρ ( 1 + κ z ) cos α p β , v t + u ( 1 + κ z ) cos α v β   +   v 1 + κ z v α   +   ϱ κ w v z   +   ϱ v w + u 2 tan α 1 + κ z + 2 μ u sin α   +   μ 2 ( 1 + κ z ) sin α cos α   =   1 ρ ( 1 + κ z ) p α , κ ϱ ( w t + u ( 1 + κ z ) cos α w β + v 1 + κ z w α + ϱ κ w v z )     κ u 2 + v 2 1 + κ z 2 μ κ u cos α     κ μ 2 ( 1 + κ z ) 2 cos 2 α   =   1 ρ p z     g ,
where g   =   g ¯ H ¯ U ¯ 2 , and the dimensionless mass conservation Equation (2) can be rewritten as
ρ t + u ( 1 + κ z ) cos α ρ β   +   v 1 + κ z ρ α   +   ϱ κ w ρ z   +   ρ ( 1 ( 1 + κ z ) cos α u β + 1 ( 1 + κ z ) cos α α ( v cos α ) + ϱ κ 1 ( 1 + κ z ) 2 z ( 1 + κ z ) 2 w )   =   0 .
According to the main characteristics of the physical correlation state of the thin shell stratosphere (see [19]), let the shallowness parameter κ 0 . Then, the governing Equation (3) is reduziertreduced as
u t   +   u cos α u β   +   v u α     u v tan α     2 μ v sin α   =   1 ρ cos α p β , v t   +   u cos α v β   +   v v α   +   u 2 tan α   +   2 μ u sin α   +   μ 2 sin α cos α   =   1 ρ p α , 0   =   1 ρ p z     g ,
and
u β   +   α ( v cos α )   =   0 .
In Equation (5), the stream function is introduced as follows:
u   =   Ψ α and v   =   1 cos α Ψ β .
Substituting Equation (6) into Equation (4) eliminates the pressure term p, and by a straightforward calculation, the governing equation of the flow pattern of the atmosphere in the rotating spherical coordinate system can be briefly expressed as
Δ Ψ t   +   1 cos α Ψ β Δ Ψ α Ψ α Δ Ψ β   +   2 μ Ψ β   =   0 ,
where Δ is called the Laplace–Beltrami operator on the surface of the unit sphere such that
Δ   =   1 cos α α cos α α   +   1 cos 2 α 2 β 2   =   2 α 2     tan α α   +   1 cos 2 α 2 β 2 .
Equation (7) is called the stratospheric planetary flows equation or simply the vorticity equation in [19]. In [19], the rigidity result of Equation (7) is established, and the Arnold’s stability criterion is given under the condition that Δ Ψ meets the appropriate conditions. The stability of the critical stationary solution is studied in the limit case where the stream function belongs to the sum of the first two eigenspaces of the Laplace–Beltrami operator. The local and global bifurcation results for the nonzonal stationary solutions of classical Rossby–Haurwitz waves are also obtained. However, it is not difficult to notice that there are no relevant results in the discussion on the existence of weak solutions for general vorticity equations. In this paper, we try to solve this problem.
It is worth mentioning that if the factor of time in Equation (7) is not taken into account, the governing Equation (7) can be regarded as a mathematical model of ocean circulation in which the vertical velocity is relatively weak compared with the horizontal velocity. This was originally developed by Constantin in [5]. After that, many scholars applied functional analysis technology and differential equation theory and found a lot of meaningful results (see [22,23,24,25,26,27,28,29,30,31,32,33]).

3. Main Results

Consider Equation (7) with the associated initial condition
Ψ ( α , β , 0 )   =   Ψ 0 ( α , β ) ,
and boundary conditions
Ψ ( π / 2 , β , t ) and Ψ ( π / 2 , β , t ) is bounded , Ψ ( α , 0 , t )   =   Ψ ( α , 2 π , t ) .
We give the following assumptions:
Assumption 1.
Let Ψ 0 ( α , β ) , Ψ 0 α , 1 cos α Ψ 0 β , Δ Ψ 0 L 2 ( π 2 , π 2 ) × ( 0 , 2 π ) , further, let us assume that
Ψ 0 n Ψ 0 strongly in L 2 ( π 2 , π 2 ) × ( 0 , 2 π ) as n , Ψ 0 n α Ψ 0 α strongly in L 2 ( π 2 , π 2 ) × ( 0 , 2 π ) as n , 1 cos α Ψ 0 n β 1 cos α Ψ 0 β strongly in L 2 ( π 2 , π 2 ) × ( 0 , 2 π ) as n , Δ Ψ 0 n Δ Ψ 0 strongly in L 2 ( π 2 , π 2 ) × ( 0 , 2 π ) as n ,
where Ψ 0 n , Ψ 0 n α , 1 cos α Ψ 0 n β and Δ Ψ 0 n are respectively approximate functions of Ψ 0 , Ψ 0 α , 1 cos α Ψ 0 β and Δ Ψ 0 in L 2 ( π 2 , π 2 ) × ( 0 , 2 π ) .
We define the norm as follows:
ψ 2   =   0 2 π π 2 π 2 ψ 2 cos α d α d β for ψ L 2 ( 0 , 2 π ) × ( π 2 , π 2 ) .
Now, we turn our consideration to the following boundary value problem:
Δ Ψ + λ Ψ   =   0 , Ψ ( α , 0 , t )   =   Ψ ( α , 2 π , t ) , Ψ ( π / 2 , β , t ) and Ψ ( π / 2 , β , t ) is bounded .
By applying the separation of variables technique, we set
Ψ   =   Θ ( α ) · Φ ( β ) ,
By substituting Equation (11) into Equation (10), we have
cos 2 α Θ Θ     sin α cos α Θ Θ   +   λ cos 2 α   =   Φ Φ   =   m 2 ,
where m is a non-negative integer. Equation (12) can be divided into the following two differential equations:
Φ   +   m 2 Φ   =   0 ,
and
Θ     sin α cos α Θ   +   λ m 2 cos 2 α Θ   =   0 .
Equation (13) is a second-order homogeneous linear differential equation with constant coefficients, and the general solution of Equation (13) can be expressed as
Φ   =   C ¯ m cos m β   +   C ˜ m sin m β .
Let x   =   sin α , α ( π 2 , π 2 ) , and Θ ( α )   =   y ( x ) . Then, Equation (14) can be equivalently converted to
( 1 x 2 ) y ( x )     2 x y ( x )   +   λ m 2 1 x 2 y ( x )   =   0 ,
Equation (15) is Legendre’s differential equation if m   =   0 , and its solution can be expressed as
P l ( sin α ) = 1 π π 2 π 2 ( sin α + i cos α cos α ^ ) l d α ^ = 1 2 l ι = 0 L ( 1 ) ι ( 2 l 2 ι ) ! ι ! ( l ι ) ! ( l 2 ι ) ! ( sin α ) l 2 ι ,
where l is a non-negative integer, λ   =   l ( l + 1 ) is the proper value, and
L = l 2 , l is even , l 1 2 , l is odd .
Let P l m ( sin α ) be the adjoint function of the Legendre polynomial P l ( sin α ) . Then, we have
P l m ( sin α )   =   ( l + 1 ) ( l + 2 ) ( l + m ) π π 2 π 2 ( sin α + i cos α cos α ^ ) l sin m α ^ d α ^ .
By simple calculation, we can find
P l ( sin α ) | α = π 2   =   ( 1 ) l , P l ( sin α ) | α = π 2   =   1 , P l m ( sin α ) | α = π 2   =   0 , P l m ( sin α ) | α = π 2   =   0 .
Let the approximate solution of Equation (7) with the initial condition in Equation (8) and boundary conditions in Equation (9) be
Ψ n ( α , β , t )   =   l = 0 n m = 0 l C ¯ l m n ( t ) cos m β + C ˜ l m n ( t ) sin m β P l m ( sin α ) ,
By taking the partial derivative, we have
Ψ n β   =   l = 1 n m = 1 l m C ¯ l m n ( t ) sin m β + C ˜ l m n ( t ) cos m β P l m ( sin α ) , Δ Ψ n   =   l = 1 n ( l ( l + 1 ) ) m = 0 l C ¯ l m n ( t ) cos m β + C ˜ l m n ( t ) sin m β P l m ( sin α ) ,
Δ Ψ n β   =   l = 1 n ( l ( l + 1 ) ) m = 0 l m C ¯ l m n ( t ) sin m β + C ˜ l m n ( t ) cos m β P l m ( sin α ) .
Suppose that the approximate solution Ψ n ( α , β , t ) satisfies the following two equations:
1 Π l m 0 2 π π 2 π 2 [ Δ Ψ n t + 1 cos α Ψ n β Δ Ψ n α Ψ n α Δ Ψ n β + 2 μ Ψ n β ] cos m β P l m ( sin α ) cos α d α d β   =   0 ,
and
1 Π l m 0 2 π π 2 π 2 [ Δ Ψ n t + 1 cos α Ψ n β Δ Ψ n α Ψ n α Δ Ψ n β + 2 μ Ψ n β ] sin m β P l m ( sin α ) cos α d α d β   =   0 ,
where
Π l m = 0 2 π π 2 π 2 P l m ( sin α ) 2 cos α sin 2 m β d α d β = 2 π δ m ( l + m ) ! ( 2 l + 1 ) ( l m ) !
and
δ m = 2 , m   =   0 , 1 , m     0 .
By substituting Equation (16) into Equations (17) and (18), the following ordinary differential equations can be obtained:
l ( l + 1 ) d C ¯ l m n d t   +   F ¯ l m ( C ¯ k i n , C ˜ q j n )   =   0 , l ( l + 1 ) d C ˜ l m n d t   +   F ˜ l m ( C ¯ k i n , C ˜ q j n )   =   0 .
Remark 1.
F ¯ l m and F ˜ l m are analytic functions of C ¯ k i n and C ˜ q j n ( i = 0 , 1 , , k ; k = 0 , 1 , , n ; j = 0 , 1 , , q ; q = 0 , 1 , , n ) . When C ¯ k i n and C ˜ q j n are uniformly bounded for any t [ 0 , T ] , F ¯ l m and F ˜ l m are also bounded, and the Lipschitz conditions are satisfied. According to ordinary differential equation theory, solutions C ¯ l m n ( t ) and C ˜ l m n ( t ) of Equation (19) with respect to t [ 0 , T ] are unique, which also shows that d C ¯ l m n d t and d C ˜ l m n d t are uniformly bounded. Therefore, C ¯ l m n ( t ) and C ˜ l m n ( t ) are equicontinuous and uniformly bounded with respect to t [ 0 , T ] . By applying the Arzela–Ascoli theorem, C ¯ l m n ( · ) and C ˜ l m n ( · ) are compact with respect to n.
Lemma 1.
There is a value M 0   >   0 such that
Ψ n β 2     M 0 Δ Ψ n 2 .
Proof. 
On the one hand, we have
Ψ n β 2 = 0 2 π π 2 π 2 Ψ n β 2 cos α d α d β = π 2 π 2 0 2 π β Ψ n Ψ n β cos α d β d α 0 2 π π 2 π 2 Ψ n 2 Ψ n β 2 cos α d α d β 1 2 0 2 π π 2 π 2 1 cos α 2 Ψ n β 2 2 cos α d α d β + 0 2 π π 2 π 2 ( cos α Ψ n ) 2 cos α d α d β 1 2 1 cos α 2 Ψ n β 2 2 + Ψ n 2 .
On the other hand, we have
Δ Ψ n 2 = 0 2 π π 2 π 2 Δ Ψ n 2 cos α d α d β = l = 0 n m = 0 l ( l ( l + 1 ) ) 2 C ¯ l m n 2 + C ˜ l m n 2 Π l m λ 0 2 l = 0 n m = 0 l C ¯ l m n 2 + C ˜ l m n 2 Π l m = λ 0 2 0 2 π π 2 π 2 Ψ n 2 cos α d α d β = λ 0 2 Ψ n 2 ,
where λ 0 is the smallest positive proper value. Furthermore, we have
Δ Ψ n cos α 2 = 0 2 π π 2 π 2 cos α Δ Ψ n 2 cos α d α d β = 0 2 π π 2 π 2 [ α cos α Ψ n α 2 + 1 cos 2 α 2 Ψ n β 2 2 + 2 cos α α cos α Ψ n α 2 Ψ n β 2 ] cos α d α d β = 0 2 π π 2 π 2 α cos α Ψ n α 2 cos α d α d β + 0 2 π π 2 π 2 1 cos α 2 Ψ n β 2 2 cos α d α d β + 2 0 2 π π 2 π 2 2 Ψ n α β 2 cos α d α d β = α cos α Ψ n α 2 + 1 cos α 2 Ψ n β 2 2 + 2 2 Ψ n α β 2 Δ Ψ n 2 .
Hence, we find
Ψ n β 2     1 2 Δ Ψ n 2 + Ψ n 2     1 2 Δ Ψ n 2 + 1 λ 0 2 Δ Ψ n 2     M 0 Δ Ψ n 2 .
This completes the proof. □
By multiplying both ends of Equations (17) and (18) by l ( l + 1 ) C ¯ l m n ( t ) and l ( l + 1 ) C ˜ l m n ( t ) , respectively, and adding them in turn, we have
0 2 π π 2 π 2 Δ Ψ n t + 1 cos α Ψ n β Δ Ψ n α Ψ n α Δ Ψ n β + 2 μ Ψ n β Δ Ψ n cos α d α d β   =   0 .
By using the Cauchy–Schwartz inequality and the properties of the derivatives, we obtain
1 2 t Δ Ψ n 2   +   1 2 0 2 π π 2 π 2 Ψ n β · ( Δ Ψ n 2 ) α Ψ n α · ( Δ Ψ n 2 ) β d α d β μ 0 2 π π 2 π 2 Ψ n β 2 + Δ Ψ n 2 cos α d α d β = μ Ψ n β 2 + Δ Ψ n 2 .
As with Lemma 1 and Equation (21), it is not hard to verify the following:
t Δ Ψ n 2     M 1 Δ Ψ n 2 .
By integrating from 0 to t on both sides of Equation (22) and applying the Gronwall inequality, we obtain
Δ Ψ n 2     Δ Ψ 0 n 2 e M 1 t     ( Δ Ψ 0 n Δ Ψ 0 2 + Δ Ψ 0 2 ) e M 1 t     C ^ 1 .
By multiplying both sides of Equations (17) and (18) by C ¯ l m n ( t ) and C ˜ l m n ( t ) , respectively, and adding them in turn, we have
0 2 π π 2 π 2 Δ Ψ n t + 1 cos α Ψ n β Δ Ψ n α Ψ n α Δ Ψ n β + 2 μ Ψ n β Ψ n cos α d α d β   =   0 .
Hence, we have
| 0 2 π π 2 π 2 t α cos α Ψ n α + 1 cos α 2 Ψ n β 2 Ψ n cos α d α d β + 1 2 0 2 π π 2 π 2 Ψ n 2 β · Δ Ψ n α Ψ n 2 α · Δ Ψ n β d α d β | μ 0 2 π π 2 π 2 Ψ n β 2 + Ψ n 2 cos α d α d β ,
By applying the integration by parts formula, we have
t Ψ n α 2 + 1 cos α Ψ n β 2 2 μ Ψ n 2 + Ψ n β 2 2 μ 1 λ 0 2 Δ Ψ n 2 + Ψ n α 2 + 1 cos α Ψ n β 2 M 2 + 2 μ Ψ n α 2 + 1 cos α Ψ n β 2 .
When integrating from 0 to t at both sides of Equation (25) and applying the Gronwall inequality again, there exists a C ^ 2 > 0 such that
Ψ n α 2   +   1 cos α Ψ n β 2 [ M 2 T + Ψ 0 α 2 + 1 cos α Ψ 0 β 2 + Ψ 0 n α Ψ 0 α 2 + 1 cos α Ψ 0 n β 1 cos α Ψ 0 β 2 ] e 2 μ t     C ^ 2 .
Theorem 1.
Assume that ( H ) holds. Then, Equation (7) with the initial condition in Equation (8) and the boundary condition in Equation (9) has at least a weak solution.
Proof. 
It is easy to verify that { Δ Ψ n } , Ψ n α , and 1 cos α Ψ n β are uniformly weakly compact with respect to t [ 0 , T ] in L 2 ( 0 , 2 π ) × ( π 2 , π 2 ) .
We are now going to show that cos α Ψ n α is strongly compact with respect to t [ 0 , T ] in L 2 ( 0 , 2 π ) × ( π 2 , π 2 ) . In fact, it is easy to know from the above statement that cos α Ψ n α is uniformly bounded with respect to t [ 0 , T ] .
Let us show that cos α Ψ n α is equicontinuous in L 2 ( 0 , 2 π ) × ( π 2 , π 2 ) . In fact, by using the mean value theorem and combining Equations (20) and (23), we obtain
| 0 2 π π 2 π 2 cos ( α + Δ α ) Ψ n α ( α + Δ α , β + Δ β , t ) cos α Ψ n α ( α , β , t ) 2 cos α d α d β | = | 0 2 π π 2 π 2 [ α cos ( α + ξ 1 Δ α ) Ψ n α ( α + ξ 1 Δ α , β + Δ β , t ) · Δ α + cos α 2 Ψ n α β ( α , β + ξ 2 Δ β , t ) · Δ β ] 2 cos α d α d β | 2 α cos α Ψ n α 2 · ( Δ α ) 2 + cos α 2 Ψ n α β 2 · ( Δ β ) 2 2 Δ Ψ n 2 ( Δ α ) 2 + ( Δ β ) 2 2 C ^ 1 ( Δ α ) 2 + ( Δ β ) 2 ,
where 0   <   ξ 1   <   ξ 2   <   1 , which shows that cos α Ψ n α is strongly compact in L 2 ( 0 , 2 π ) × ( π 2 , π 2 ) . Analogously, Ψ n β is strongly compact in L 2 ( 0 , 2 π ) × ( π 2 , π 2 ) .
Assume that
Ψ n Ψ weakly in L 2 ( 0 , 2 π ) × ( π 2 , π 2 ) as n ,
Then, we have
Ψ n α Ψ α weakly in L 2 ( 0 , 2 π ) × ( π 2 , π 2 ) as n , 1 cos α Ψ n β 1 cos α Ψ β weakly in L 2 ( 0 , 2 π ) × ( π 2 , π 2 ) as n , Δ Ψ n Δ Ψ weakly in L 2 ( 0 , 2 π ) × ( π 2 , π 2 ) as n .
When setting Ψ n α ζ 1 ( α , β , t ) , 1 cos α Ψ n β ζ 2 ( α , β , t ) , and Δ Ψ n ζ 3 ( α , β , t ) , for ξ ( α , β , t ) C 1 π 2 , π 2 × ( 0 , 2 π ) × [ 0 , T ] , by the definition of the generalized derivative and Equation (26), we have
0 2 π π 2 π 2 ζ 1 · ξ · cos α d α d β = lim n 0 2 π π 2 π 2 Ψ n α · ξ · cos α d α d β = lim n 0 2 π π 2 π 2 Ψ n α ξ cos α d α d β = 0 2 π π 2 π 2 Ψ α ξ cos α d α d β = 0 2 π π 2 π 2 Ψ α · ξ · cos α d α d β
which implies that Ψ α   =   ζ 1 ( α , β , t ) . Similarly, we can verify that 1 cos α Ψ β   =   ζ 2 ( α , β , t ) .
On the other hand, we have
0 2 π π 2 π 2 ζ 3 · ξ · cos α d α d β = lim n 0 2 π π 2 π 2 Δ Ψ n · ξ · cos α d α d β = lim n 0 2 π π 2 π 2 1 cos α α cos α Ψ n α + 1 cos 2 α 2 Ψ n β 2 ξ cos α d α d β = lim n 0 2 π π 2 π 2 cos α Ψ n α ξ α + 1 sin α Ψ n β ξ β d α d β = 0 2 π π 2 π 2 cos α Ψ α ξ α + 1 sin α Ψ β ξ β d α d β = 0 2 π π 2 π 2 Δ Ψ · ξ · cos α d α d β ,
which shows that Δ Ψ   =   ζ 3 .
We prove that the following conclusion is true:
lim n 0 2 π π 2 π 2 1 cos α Ψ n β ξ ˜ α Δ Ψ n Ψ n α ξ ˜ β Δ Ψ n cos α d α d β = 0 2 π π 2 π 2 1 cos α Ψ β ξ ˜ α Δ Ψ Ψ α ξ ˜ β Δ Ψ cos α d α d β
where ξ ˜ ( α , β , t )   =   cos 2 α ξ ¯ ( α , β , t ) , ξ ¯ ( α , β , t ) C 1 π 2 , π 2 × ( 0 , 2 π ) × [ 0 , T ] .
In fact, by simple calculation, we have
| 0 2 π π 2 π 2 [ 1 cos α ( Ψ n β Δ Ψ n 2 cos α sin α ξ ¯ + cos 2 α ξ ¯ α Ψ β Δ Ψ 2 cos α sin α ξ ¯ + cos 2 α ξ ¯ α Ψ n α Δ Ψ n ξ ¯ β cos 2 α + Ψ α Δ Ψ ξ ¯ β cos 2 α ) ] cos α d α d β | | 0 2 π π 2 π 2 cos α Ψ n α cos α Ψ α Δ Ψ n ξ ¯ β cos α d α d β |
+ | 0 2 π π 2 π 2 cos α Ψ α Δ Ψ n Δ Ψ ξ ¯ β cos α d α d β | + | 0 2 π π 2 π 2 Ψ n β Ψ β · ξ ¯ α cos α 2 ξ ¯ sin α Δ Ψ n cos α d α d β | + | 0 2 π π 2 π 2 ξ ¯ α cos α 2 ξ ¯ sin α · Δ Ψ n Δ Ψ Ψ β cos α d α d β | < ε .
Suppose ξ ˜ ( α , β , t ) has the following approximate solution:
ξ ˜ n ( α , β , t ) = cos 2 α ξ ¯ n ( α , β , t ) = l = 0 n m = 0 l c ¯ l m ( t ) cos m β + c ˜ l m ( t ) sin m β P l m ( sin α ) ,
where
c ¯ l m ( t )   =   1 Π l m 0 2 π π 2 π 2 cos 2 α ξ ¯ ( α , β , t ) cos m β P l m ( sin α ) cos α d α d β , c ˜ l m ( t )   =   1 Π l m 0 2 π π 2 π 2 cos 2 α ξ ¯ ( α , β , t ) sin m β P l m ( sin α ) cos α d α d β .
By multiplying Equations (17) and (18) by c ¯ l m ( t ) and c ˜ l m ( t ) , respectively, and then adding them together and integrating from 0 to t, we have
0 t 0 2 π π 2 π 2 Δ Ψ n t + 1 cos α Ψ n β Δ Ψ n α Ψ n α Δ Ψ n β + 2 μ Ψ n β ξ ˜ n cos α d α d β d t   =   0 ,
Then, by using the integration by parts formula, we obtain
0 t 0 2 π π 2 π 2 [ ( Δ Ψ n ξ ˜ n t 1 cos α Ψ n β ξ ˜ n α Ψ n α ξ ˜ n β Δ Ψ n + 2 μ ξ ˜ n Ψ n β ) cos α ] d α d β d t + 0 2 π π 2 π 2 Δ Ψ 0 n ( α , β ) cos α d α d β   =   0 .
Now let n in Equation (27). Then, the conclusion of Theorem 1 is proven. □
Next, we give the following non-uniqueness of weak solutions:
Theorem 2.
Suppose Ψ ( α , β , t ) is a weak solution to Equation (7) with the initial condition in Equation (8) and the boundary condition in Equation (9). For any function where Ψ ˜ ( t ) C 1 [ 0 , T ] and Ψ ˜ ( 0 )   =   0 , then Ψ ¯ ( α , β , t )   =   Ψ ( α , β , t )   +   Ψ ˜ ( t ) is also a weak solution to Equation (7) with the initial condition of Equation (8) and boundary condition of Equation (9).
Proof. 
Noting that Ψ is a weak solution to Equation (7), Equation (27) holds. Based on the assumption of the Theorem 2, we obtain
0 t 0 2 π π 2 π 2 [ ( Δ Ψ ¯ ξ ˜ t 1 cos α Ψ ¯ β ξ ˜ α Ψ ¯ α ξ ˜ β Δ Ψ ¯ + 2 μ ξ ˜ Ψ ¯ β ) cos α ] d α d β d t + 0 2 π π 2 π 2 Δ Ψ ¯ 0 ( α , β ) cos α d α d β = 0 ,
which implies that Theorem 2 is correct. □

4. Conclusions

According to the concept of spherical harmonic function and weak solutions, the governing equation describing atmospheric flow is transformed into a second-order homogeneous linear ordinary differential equation and a Legendre’s differential equation by using the method of separating variables. In this paper, the existence and non-uniqueness of weak solutions of vorticity equations are obtained by using the Gronwall inequality and Cauchy–Schwartz inequality under the appropriate initial and boundary conditions. Compared with the vorticity equation established in [5], our conclusion considers the time dependence of the system. The authors of [19] studied the stability of the Rossby–Haurwitz stationary solution but did not provide the existence proof for the solution. In this paper, we give the existence results of weak solutions, which to a certain extent provides a theoretical basis for the stability study of [19].

Author Contributions

The contributions of all authors are equal. All authors have read and agreed to the published version of the manuscript.

Funding

This work is partially supported by Guizhou Provincial Science and Technology Projects (ZK[2022]535), Training Object of High Level and Innovative Talents of Guizhou Province ((2016)4006), Major Research Project of Innovative Group in Guizhou Education Department ([2018]012), the Slovak Research and Development Agency under the contract No. APVV-18-0308, and by the Slovak Grant Agency VEGA No. 1/0358/20 and No. 2/0127/20.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Constantin, A.; Johnson, R.S. On the modelling of large-scale atmospheric flow. J. Differ. Equ. 2021, 285, 751–798. [Google Scholar] [CrossRef]
  2. Constantin, A.; Johnson, R.S. On the propagation of waves in the atmosphere. Proc. R. Soc. Math. Phys. Eng. Sci. 2020, 477, 20200424. [Google Scholar] [CrossRef] [PubMed]
  3. Vallis, G.K. Atmospheric and Oceanic Fluid Dynamics; Cambridge University Press: Cambridge, UK, 2017. [Google Scholar]
  4. Dellar, P.J. Variations on a beta-plane: Derivation of non-traditional beta-plane equations from Hamilton’s principle on a sphere. J. Fluid Mech. 2011, 674, 174–195. [Google Scholar] [CrossRef]
  5. Constantin, A.; Johnson, R.S. Large gyres as a shallow-water asymptotic solution of Euler’s equation in spherical coordinates. Proc. R. Soc. Math. Phys. Eng. Sci. 2017, 473, 20170063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Martin, C.I. Azimuthal equatorial flows in spherical coordinates with discontinuous stratification. Phys. Fluids 2021, 33, 026602. [Google Scholar] [CrossRef]
  7. Martin, C.I.; Quirchmayr, R. Exact solutions and internal waves for the Antarctic Circumpolar Current in spherical coordinates. Stud. Appl. Math. 2022, 148, 1021–1039. [Google Scholar] [CrossRef]
  8. Constantin, A.; Johnson, R.S. An exact, steady, purely azimuthal flow as a model for the Antarctic Circumpolar Current. J. Phys. Oceanogr. 2016, 46, 3585–3594. [Google Scholar] [CrossRef]
  9. Constantin, A.; Johnson, R.S. An exact, steady, purely azimuthal equatorial flow with a free surface. J. Phys. Oceanogr. 2016, 46, 1935–1945. [Google Scholar] [CrossRef]
  10. Chu, J. Monotone solutions of a nonlinear differential equation for geophysical fluid flows. Nonlinear Anal. 2018, 166, 144–153. [Google Scholar] [CrossRef]
  11. Wang, J.; Fečkan, M.; Zhang, W. On the nonlocal boundary value problem of geophysical fluid flows. Z. Angew. Math. Phys. 2021, 72, 27. [Google Scholar] [CrossRef]
  12. Haziot, S.V. Study of an elliptic partial differential equation modelling the Antarctic Circumpolar Current. Discret. Contin. Dyn. Syst. 2019, 39, 4415–4427. [Google Scholar] [CrossRef] [Green Version]
  13. Martin, C.I.; Quirchmayr, R. A steady stratied purely azimuthal flow representing the Antarctic Circumpolar Current. Monatsh. Math. 2020, 192, 401–407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Martin, C.I.; Quirchmayr, R. Explicit and exact solutions concerning the Antarctic Circumpolar Current with variable density in spherical coordinate. J. Math. Phys. 2019, 60, 101505. [Google Scholar] [CrossRef]
  15. Constantin, A.; Johnson, R.S. Large-scale oceanic currents as shallow water asymptotic solutions of the Navier-Stokes equation in rotating spherical coordinates. Deep.-Sea Res. Part II Top. Stud. Oceanogr. 2019, 160, 32–40. [Google Scholar] [CrossRef]
  16. Constantin, A.; Johnson, R.S. Ekman-type solutions for shallow-water flows on a rotating sphere: A new perspective on a classical problem. Phys. Fluids 2019, 31, 021401. [Google Scholar] [CrossRef] [Green Version]
  17. Catling, D.C. Planetary atmospheres. In Treatise on Geophysics; Schubert, G., Ed.; Elsevier: Oxford, UK, 2015; pp. 429–472. [Google Scholar]
  18. Gill, A.E. Atmosphere-Ocean Dynamics; Academic Press: Cambridge, UK, 1982. [Google Scholar]
  19. Constantin, A.; Germain, P. Stratospheric planetary flows from the perspective of the Euler equation on a rotating sphere. Arch. Ration. Mech. Anal. 2022, 245, 587–644. [Google Scholar] [CrossRef]
  20. Dowling, T.E. Dynamics of Jovian atmospheres. Annu. Rev. Fluid Mech. 1995, 27, 293–334. [Google Scholar] [CrossRef]
  21. Lunine, J.I. The atmospheres of Uranus and Neptune. Annu. Rev. Astron. Astrophys. 1993, 31, 217–263. [Google Scholar] [CrossRef]
  22. Marynets, K. A nonlinear two-point boundary-value problem in geophysics. Monatsh. Math. 2019, 188, 287–295. [Google Scholar] [CrossRef]
  23. Marynets, K. On a two-point boundary-value problem in geophysics. Appl. Anal. 2019, 98, 553–560. [Google Scholar] [CrossRef]
  24. Marynets, K. A weighted Sturm-Liouville problem related to ocean flows. J. Math. Fluid Mech. 2018, 20, 929–935. [Google Scholar] [CrossRef]
  25. Chu, J. On a differential equation arising in geophysics. Monatsh. Math. 2018, 187, 499–508. [Google Scholar] [CrossRef]
  26. Chu, J. On a nonlinear model for arctic gyres. Ann. Mat. Pura Appl. 2018, 197, 651–659. [Google Scholar] [CrossRef]
  27. Chu, J. On an infinite-interval boundary-value problem in geophysics. Monatsh. Math. 2019, 188, 621–628. [Google Scholar] [CrossRef]
  28. Chu, J.; Marynets, K. Nonlinear differential equations modeling the Antarctic Circumpolar Current. J. Math. Fluid Mech. 2021, 23, 92. [Google Scholar] [CrossRef]
  29. Wang, J.; Fečkan, M.F.; Wen, Q.; O’Regan, D. Existence and uniqueness results for modeling jet flow of the Antarctic Circumpolar Current. Monatsh. Math. 2021, 194, 601–621. [Google Scholar] [CrossRef]
  30. Wang, J.; Zhang, W.; Fečkan, M. Periodic boundary value problem for second-order differential equations from geophysical fluid flows. Monatsh. Math. 2021, 195, 523–540. [Google Scholar] [CrossRef]
  31. Fečkan, M.; Li, Q.; Wang, J. Existence and Ulam-Hyers stability of positive solutions for a nonlinear model for the Antarctic Circumpolar Current. Monatsh. Math. 2022, 197, 419–434. [Google Scholar] [CrossRef]
  32. Zhang, W.; Wang, J.; Fečkan, M. Existence and uniqueness results for a second order differential equation for the ocean flow in arctic gyres. Monatsh. Math. 2020, 193, 177–192. [Google Scholar] [CrossRef]
  33. Zhang, W.; Fečkan, M.; Wang, J. Positive solutions to integral boundary value problems from geophysical fluid flows. Monatsh. Math. 2020, 193, 901–925. [Google Scholar] [CrossRef]
Table 1. Reference values of parameters corresponding to some planets.
Table 1. Reference values of parameters corresponding to some planets.
Planet H ¯ R ¯ g ¯ Ω ¯ W ¯ U ¯ μ κ ϱ
Earth40 km6371 km9.8 m/s27.27 × 10−5 rad/s 10 3 m/s50 m/s96 × 10−32 × 10−5
Saturn200 km58,232 km10.4 m/s2 1.62 × 10 4 rad/s 10 2 m/s150 m/s63 3 × 10 3 6 × 10 5
Uranus150 km25,362 km8.8 m/s2 1.04 × 10 4 rad/s 10 5 m/s150 m/s18 6 × 10 3 6 × 10 8
Jupiter270 km69,911 km24.8 m/s2 1.76 × 10 4 rad/s 10 2 m/s150 m/s82 4 × 10 3 6 × 10 5
Neptune200 km24,622 km11.1 m/s2 1.08 × 10 4 rad/s 10 3 m/s200 m/s13 8 × 10 3 5 × 10 6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, W.; Fečkan, M.; Wang, J. The Existence of Weak Solutions for the Vorticity Equation Related to the Stratosphere in a Rotating Spherical Coordinate System. Axioms 2022, 11, 347. https://doi.org/10.3390/axioms11070347

AMA Style

Zhang W, Fečkan M, Wang J. The Existence of Weak Solutions for the Vorticity Equation Related to the Stratosphere in a Rotating Spherical Coordinate System. Axioms. 2022; 11(7):347. https://doi.org/10.3390/axioms11070347

Chicago/Turabian Style

Zhang, Wenlin, Michal Fečkan, and Jinrong Wang. 2022. "The Existence of Weak Solutions for the Vorticity Equation Related to the Stratosphere in a Rotating Spherical Coordinate System" Axioms 11, no. 7: 347. https://doi.org/10.3390/axioms11070347

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop