Seediscussions,stats,andauthorprofilesforthispublicationat:https://www.researchgate.net/publication/277559098 LongitudinalDimensionsofPolygon-shaped PlanetaryWaves ArticleinJournalofAppliedNonlinearDynamics·June2015 DOI:10.5890/JAND.2015.06.005 READS 18 1author: RanisIbragimov UniversityofTexasatBrownsvilleandTexas… 77PUBLICATIONS864CITATIONS SEEPROFILE Availablefrom:RanisIbragimov Retrievedon:29July2016 Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 Journal of Applied Nonlinear Dynamics https://lhscientificpublishing.com/Journals/JAND-Default.aspx Longitudinal Dimensions of Polygon-shaped Planetary Waves Ranis N. Ibragimov1†and Guang Lin2 1 GE Global Research 1 Research Circle, Niskayuna, NY 12309, USA Mathematics, and School of Mechanical Engineering, Purdue University, West Lafayette, IN 47907, USA 2 Department of Submission Info Communicated by Lev Osctrovsky Received 9 July 2014 Accepted 1 March 2015 Available online 1 July 2015 Keywords Shallow water approximation Planetary hexagon-shaped waves Free boundary problem Atmospheric modeling Abstract Polygon-shaped longitudinal large-scale waves are described by means of higher-order shallow water approximation corresponding to the Cauchy–Poisson free boundary problem on the stationary motion of a perfect incompressible fluid circulating around a circle. It is shown that there are four basic physical parameters, which exert an influence on a wave number (or wave length), which is one of the basic values used to characterize the planetary flow pattern in mid-troposphere. Some analogy with the jet-stream following hexagon-shaped path at Saturn’s north pole is observed. ©2015 L&H Scientific Publishing, LLC. All rights reserved. 1 Introduction Planetary waves are large-scale perturbations of the atmospheric dynamical structure that extend coherently around a full longitude circle. They are important because they have significant influence on the wind speeds, temperature, distribution of ozone, and other characteristics of the middle atmosphere structure. They also play an important role in global climate control and weather prediction [1–3]. In oceanographic applications, understanding of the atmospheric processes mechanisms have greatly increased due to microstructures measurements over the past two decades. In terms of mathematical modeling, the large-scale atmospheric dynamics is usually described by moving air masses on a sphere or circle by means of three and two dimensional Navier–Stokes or Euler equations a thin rotating spherical shell (see e.g. [4–12] ) or within the theory of shallow water approximation [13–20]. Particularly, a largescale two-dimensional modeling with the inclusion of a spherical shape can be associated e.g. with the eastward moving, wave of warm water, known as a Kelvin wave that can be seen traveling eastward along the equator as shown in the left Panel of Figure 1 (see also [21–23]). Another spectacular example of circulating waves is demonstrated on the right panel of Figure 1 showing a jet stream that follows a hexagon-shaped path at the north pole of Saturn. The hexagon was hidden in darkness † Corresponding author. Email address: [email protected] ISSN 2164 − 6457, eISSN 2164 − 6473/$-see front materials © 2015 L&H Scientific Publishing, LLC. All rights reserved. DOI : 10.5890/JAND.2015.06.005 154 Ranis N. Ibragimov, Guang Lin /Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 Fig. 1 Left: Sea-level height data from November 2009 showing the dynamics of warm water known as Kelvin waves that can be seen traveling eastward along the equator (black line) in Nov. 01, 2009 image. El Ninos form when trade winds in the equatorial western Pacific relax over a period of months, sending Kelvin waves eastward across the Pacific like a conveyor belt. Image credit: NASA/JPL. Right: Image from Cassini, made possible only as Saturn’s north pole emerged from winter darkness, shows new details of a jet stream that follows a hexagon-shaped path and has long puzzled scientists (source: http://saturn.jpl.nasa.gov/video/videodetails/?videoID=200) during the winter of Saturn’s long year, a year that is equal to about 29 Earth years. But as the planet approached its August 2009 equinox and signaled the start of northern spring, the hexagon was revealed to Cassini’s cameras. This is the first time the whole hexagonal shape has been mapped out in visible light by Cassini, and these images show unprecedented details of Saturn’s high northern latitudes. The hexagon was originally discovered in images taken by Voyager spacecraft in the early 1980s. Since 2006, the Cassini Visual and Infrared Mapping Spectrometer (VIMS) instrument has been observing the hexagon at infrared wavelengths, but at lower spatial resolution than these visible light images. This image also shows another unexplained phenomena such as waves that can be seen traveling along hexagon. Scientists think the hexagon is a meandering jet stream at 77 degrees north latitude, but they don’t know what controls the path the stream takes. Multiple images acquired by the VIMS instrument over a 12-day period showed that the feature is nearly stationary and is likely an unusually strong pole-encircling planetary wave that extends deep into the atmosphere. Scientists had speculated that a large vortex seen outside the hexagon during the Voyager observations exerted forces on the jet stream making it adopt a hexagonal pattern in a manner similar to how jet streams on Earth divert around high-pressure systems. However, in these new images, the vortex is notably absent while the hexagon persists almost 30 years after it was first seen. The images were taken in visible light with the Cassini spacecraft wide-angle camera on Jan. 3, 2009. The images were obtained at a distance of approximately 764,000 kilometers (475,000 miles) from Saturn. The smallest resolved features at the latitude of the hexagon have a horizontal scale of approximately 100 kilometers. Recent laboratory experiments in [24] suggest that the observed Saturn’s North Polar Hexagon might result from the stabilization of a standing waves caused by the difference in angular velocity. However, because of the complex atmospheric structure in Saturn, the provided experiments do not provide the clear answers and these waves and the six-sided shape of the Ranis N. Ibragimov, Guang Lin /Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 155 jet stream remain a mystery up to the date. The information about this phonemenon can be fouind at http://saturn.jpl.nasa.gov/video/videodetails/?videoID=200. The primary focus of this paper is to a show that longitudinal large-scale waves within a central gravity field might also provide a similar polygon-shaped structure when looking from above the North Pole, as shown in Figure 1 by means of two-dimensional free boundary large-scale shallow water approximation describing a simple atmospheric motion around an equatorial plane (see also [25]). As has been discussed in [26] and [27], the mathematical model can be derived from the assumption that the atmosphere is approximated by a perfect fluid and its motion is irrotational and pressure on a free boundary is constant. It is also postulated that the fluid depth is small compared to the radius of the circle and the gravity vector is directed to the center of the circle. In the first approximation, shallow water equations represent the mathematical theory that can be used to investigate the fluid flows in channels (see e.g. [28–30]). However, as has been discussed in [27], this theory does not reveal the role the role of an undisturbed level of the fluid surface which is needed to determine the precision of the first approximation. A higher order approximation is derived in this work. 2 The model We introduce polar coordinates x = r cos θ , y = r sin θ and use the following notation: R is the radius of the Earth, θ is a polar angle, r is the distance from the origin, h = h0 + η (t, θ ) , where h0 is undisturbed level of atmosphere above the Earth and η (t, θ ) is the level of disturbance of a free boundary, as shown schematically in Figure 2. It is supposed in what follows that θ ∈ [0, 2π ] while r ∈ [R, h (t, θ )] . The → homogeneous gravity field − g is assumed to be a constant and directed to the center of the Earth. The restriction θ ∈ [0, 2π ] appears for the following reason: the velocity potential ϕ (ς ) can be introduced by the analyticity of the complex potential ϑ (ς ) = ϕ + iψ , where ς = reiθ is the independent complex variable and ψ (ς ) is the stream function. Correspondingly, the complex velocity d ϑ /d ς is a singlevalued analytic function of ς , although ϑ is not single-valued. In fact, when we turn around the bottom 2´π r = R once, ϕ increases by − ∂∂ψr (R, θ ) d θ which has a positive sign by the maximum principle (Hopf’s 0 lemma). Hence, if we remove the width of annulus region θ = 0, r ∈ [R, R + h0 ] , then at every point (r, θ ) , the complex potential ϑ (ς ) is a single-valued analytic function. → We start with the usual assumption that the velocity field − v = (vr , vθ ) satisfies the Euler’s equations and the no-leak condition vr = 0 on a solid bottom r = R. We also assume the kinematic condition on the free boundary. Namely, the velocity on the free boundary r = R + h (t, θ ) is tangential to the free boundary. We define the free boundary by equation f = r − h (θ ,t) = 0 so that the kinematic condition is written as ∂f − df = +→ v ∇ f = 0, (1) dt ∂t where ∇= ∂ 1 ∂ , ∂r r ∂θ . (2) In what follows, it is assumed that the fluid motion is potential in the domain of the motion which allows to introduce the stream function ψ (t, r, θ ) via vr = − 1 ∂ψ , r ∂θ vθ = ∂ψ . ∂r (3) So that the no-leak condition on the solid boundary can be written as ψ (R, θ ,t) = 0 whereas the kinematic condition (1) takes the form 156 Ranis N. Ibragimov, Guang Lin /Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 h (θ,t) R g θ r h0 0 Fig. 2 Schematic showing a longitudinal atmospheric motion circulating around the Earth. ∂h 1 ∂ψ 1 ∂ψ ∂h + + = 0. ∂t r ∂θ r ∂r ∂θ Since 1 vr = − r R+h ˆ R ∂ vθ dr, ∂θ (4) (5) we can also write Eq. (4) at the free boundary r = R + h as the mass balance equation. i.e., 1 ∂ ∂h + ∂t R + h ∂θ R+h ˆ vθ dr = 0. (6) R Following [27], we next define the average velocity u (θ ,t) as 1 u (θ ,t) = h R+h ˆ R 1 vθ (r, θ ,t) dr = ψ (R + h, θ ,t) . h (7) In terms of the average velocity, the kinematic condition (4) is written as ∂h 1 + ∂t r R+h ˆ R 1 ∂ψ ∂h ∂ vθ dr + = 0. ∂θ r ∂r ∂θ Finally, the dynamic condition is obtained from the requirement that the pressure p is constant at the free boundary r = R + h (θ ,t). Thus, projecting of the impulse equation → 1 →2 1 ∂− v → + ∇( |− v | ) + ∇p = − g, ∂t 2 ρ (8) Ranis N. Ibragimov, Guang Lin /Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 on the tangential vector 157 1 ∂h − → τ =( , 1) r ∂θ to the free boundary yields 1 ∂ 1 2 1 1 ∂h ∂ p 1 ∂ p 1 ∂ h ∂ vr ∂ vθ 1 ∂ h ∂ 1 2 + + vr + v2θ + vr + v2θ + ( + ) = 0, r ∂θ ∂t ∂t r ∂θ ∂r 2 r ∂θ 2 ρ r ∂θ ∂r r ∂θ (9) where ρ is a constant fluid density. Thus, since p|r=R+h = const. and ψ is the harmonic function at the domain of the fluid motion, the model describing a longitudinal atmospheric motion around the Earth can be written as the following free boundary problem: 2 ∂ 2ψ ∂ψ 2∂ ψ = 0 (R < r < R + h), (10) + r +r 2 2 ∂θ ∂r ∂r ψ (R, θ ,t) = 0, (11) ψ (R + h, θ ,t) = u (θ ,t) h, (12) 1 ∂ h ∂ 2ψ 1 ∂ 1 ∂ψ 2 ∂ 2ψ ∂ψ 2 g ∂h − 2 ) ]+ + [ 2( ) +( = 0, (r = R + h) , (13) ∂ t ∂ r r ∂ θ ∂ t ∂ θ 2r ∂ θ r ∂ θ ∂r r ∂θ ∂h ∂ + (uh) = 0, (r = R + h). (14) r ∂t ∂θ One can check by direct differentiation that there exists an exact stationary solution to the model (10)–(14) given by Γ r (15) h0 = 0, ψ0 = − log( ), 2π R where Γ = const.is intensity of the vortex (source) localized at the center of the earth and is related with the the rotation rate of the earth (angular velocity Ω = 2π rad/day ≈ 0.73 × 10−4 s−1 ) by the equation Γ = 2π ΩR2 . The solution (15) corresponds to the singular constant flow with an undisturbed free surface with the vortex localized at the origin. However, the vortex is isolated since R represents a solid boundary. Thus the exact solution (15) can be visualized as a flow whose streamlines are concentric circles with the common center at the origin. Understanding of singular flows were conduced in [13,31,32] and [14]. As has been remarked in [33], the computational experiments in the latter papers provide a credible evidence to support the conclusion that singular solutions may exist on a stationary sphere in terms of shallow water approximation. We remark that, in terms of physical interpretation, the fluid particles at the North and South Poles spin around themselves at a rate Ω = 2π rad/ day, whereas fluid particles in the polar domain θ ∈ [θ0 , π − θ0 ] do not spin around themselves but simply translate provided π θ0 ∈ 0, 2 . Thus the physically possible atmospheric motion rotating around the poles correspond to the flows that are being translated along the equatorial plane (Ibragimov, [9, 10]). At certain extent, the above ansatz (15) can also be associated with such atmospheric phenomena as illustrated in Figure 3 which is used to show the NASA images of a polar vortex on Venus (Left panel) and clouds circling over Saturn’s north pole (Right panel). Particularly, it is believed that the polar vortex as shown in the left panel of Figure 3 is a very powerful whirlpool swirling steadily around the planet’s poles at all times. It might be caused by a gigantic hurricane with two calm, dark eyes. This double-eyed feature, dubbed the “dipole of Venus,” was thought to form when warm air from the planet’s equator rose and traveled toward the pole, where it cooled and sank to form a deep, swirling atmospheric pit. For decades, astronomers expected to find 158 Ranis N. Ibragimov, Guang Lin /Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 Fig. 3 Left: Recent pictures of a polar vortex on Venus which is attributed to cloud formations on the palnet leaving an unexplained dark hole. It has been iscovered at Venus’ north pole by the Pioneer Venus spacecraft in 1979. Credit: ESA/Virtis/INAF-IASF/Obs. de Paris-LESIA Right: Picture taken by the Cassini spacecraft of clouds circling over Saturn’s north pole (source: http://www.wired.com/2010/09/venus-polar-vortex/) a similar vortex at Venus’ south pole. While Venus itself rotates slowly, just once every 117 Earth days, its atmosphere whips around the planet once every four Earth days. This “super-rotating” atmosphere ought to form massive storms at both poles, astronomers reasoned [34, 35]. The image of the clouds circling over Saturn’s north pole shown in the right panel of Figure 3 are taken by the Cassini spacecraft from a distance of about 380,000 kilometers and represents the stunning detail in Saturn’s atmosphere. Clouds rise and sink and get stretched out, forming long valleys and ridges, streamers circling the planet’s pole. This vortex is over 2000 kilometers; that’s far bigger than a fully mature hurricane on Earth, but unlike a terrestrial cyclone, this may be a permanent feature in Saturn’s atmosphere (the source: http://www.wired.com/2010/09/venus-polar-vortex/; see also [36]). 3 Shallow water approximation It is useful to recast the model in nondimensional form by introducing the following dimensionless variables: R t θ = θ, r = R + h0 r, h = h0 , h, t = √ gh0 , u = gh0 u. ψ = h0 gh0 ψ We next introduce the parameter (16) h0 . (17) R Of course, water is shallow if the parameter ε is small. So, in the present model (10)–(14), the functions η (θ ,t) and ψ (r, θ ,t) are two unknown functions whereas the parameter ε is a given parameter. Although shallow water theory is usually related to the case when the water depth is small relative to the wavelengths of the waves, we find it more appropriate to choose the radius of the earth R as a ε= Ranis N. Ibragimov, Guang Lin /Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 159 natural physical scale since, in the frame of the present model, we consider waves with wavelengths of the order of the radius of the Earth. The the dynamic condition (13) is then nondimensionalized as follows: 1 1 ∂ 2ψ ε2 ∂ 2ψ ∂ h ∂ ε2 ∂ψ 2 ∂ψ 2 ∂h − ) )+ + ( ( ) +( = 0. 2 2 ∂ t ∂ r (1 + ε h) ∂ t ∂ θ ∂ θ 2 (1 + ε h) ∂ θ (1 + ε h) ∂ θ ∂r (1 + ε h) ∂ θ (18) Following the Lagrange’s method we represent the stream function ψ by the following series expansion: ψ = ∑ε n ψ (n) . Then the Laplace equation (10) takes the form n ∂ 2 ψ (0) ∂ 2 ψ (1) ∂ 2 ψ (0) ∂ ψ (0) ) + ε ( + 2r + ∂ r2 ∂ r2 ∂ r2 ∂r 2 (0) ∂ 2 ψ (0) ∂ 2 ψ (2) ∂ 2 ψ (1) ∂ ψ (1) ∂ ψ (0) 2∂ ψ + r ) + 0(ε 3 ) = 0. + + 2r + r + +ε 2 ( ∂θ2 ∂ r2 ∂ r2 ∂ r2 ∂r ∂r (19) A comparison of the terms with the same order ε in equation (19) yields a recurrent system of differential equations for the determination of all functions ψ (n) , i.e. the Lagrange method consists in presentation of ψ as the solution of the Cauchy problem with boundary conditions (11)–(12) for ψ (0) and zero boundary conditions for ψ (1) and ψ (2) . Thus, up to the order ε 2 , the function ψ is determined as follows: ψ = ur + ε (u r r2 r2 r3 r r − uh ) + ε 2 (uh − uθ θ + uθ θ h2 − uh2 ). 2 2 4 6 6 4 (20) Note that the unknowns u and h are related by the dynamic and conditions (18) and the kinematic condition (14) which is written in nondimensional form as follows: ∂ 1∂ 2 ε h + 2h + (uh) = 0. 2 ∂t ∂θ (21) (1 + ε h)−1 = 1 − ε h + ε 2h2 + 0 ε 3 , (22) Using the Taylor series expansion and keeping the terms 0 ε 2 , we write the dynamic condition (18) as ∂ 2ψ 1 ∂ 2 ∂ ψ 2 ∂ψ 2 ∂ 2ψ ∂ h + ) ) − ε2 (ε ( ) +( ∂ t∂ r 2 ∂ θ ∂θ ∂r ∂ t∂ θ ∂ θ εh ∂ ∂ ψ 2 ∂h ∂h ) + ε h )(ε h − 1) + ( = 0. (23) 2 ∂θ ∂r ∂θ ∂θ Substituting ψ given by (20) into equation (23), we arrive at the following system of nonlinear shallow water equations (higher-order analogue of the Su - Gardner equations [37]): +( 5h u u2 ut − ht + 2huuθ − hθ + hhθ − 2h [ut + uuθ + hθ ]) 2 2 2 2 2 h h 1 2 +ε 2 ( ut − uθ θ t + uθ θ hht − ut θ hhθ + h2 uθ uθ θ 4 3 3 3 3 2 1 h2 + h uuθ + hhθ uuθ θ − uuθ θ θ + h2 hθ ) + o ε 3 = 0, 4 3 3 ∂ 1∂ 2 ε h + 2h + (uh) = 0. 2 ∂t ∂θ ut + uuθ + hθ + ε ( (24) (25) 160 Ranis N. Ibragimov, Guang Lin /Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 Application of Lie Group Analysis ( [12, 38, 39]) shows that there exists an exact singular nonstationary one-parameter invariant solution (an invariant solution of a differential equation is a solution of the differential equation which is also an invariant curve (surface) of a group admitted by the differential equation. Such solutions can be found without determining its general solution.) of the model (24)–(25) at zeroth order epsilon, which can be written as 2θ + α, 3t u= h=( θ − α )2 , 3t (26) where α is an arbitrary constant. Sophus Lie proposed for the first time to study the symmetries of differential equations and use them for constructing solutions at the end of nineteenth century. By a symmetry we mean a continuous group of transformations acting on the dependent and independent variables of the system of differential equations so that the system stays unchanged ( [40, 41]). Since the effects of rotation are not included in this work, the invariant solution (26) is different from the set of invariant solutions obtained in [40]. These solutions for different values of time t are plotted in Figure 4 versus the polar angle θ , in which we set α = 10. For example, the exact solution for h at t = 0.1 can be associated with a single wave oscillating around the equatorial plane, as shown in Figure 5. Finding the invariant solutions of the complete shallow water system (24)– (25) with ε = 0 will be the task of the forthcoming project. Unfortunately, any small perturbation of an equation breaks the admissible group of transformations and reduces the applied value of these ”refined” equations and group theoretical methods in general. Therefore, development of methods of group analysis stable with respect to small perturbations of differential equations has become vital. As is discussed in the Conclusion section, the methods of Approximate Group Analysis will be applied to the given model (24)–(25). 120 Exact solution 80 100 Exact solution 80 θ=π U (t =0.1) H (t = 0.1) U (t = 0.3) H (t = 0.3) U 60 40 H 20 0 0.2 0.4 0.6 0.8 1 time 60 40 20 0 0 1 2 3 θ 4 5 6 Fig. 4 Exac solution of the shallow water model, that results in the limiting case of the system system(24)–(25) at ε = 0. Ranis N. Ibragimov, Guang Lin /Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 161 r h(θ,t) R θ Fig. 5 Schematic presentation of the invariant solution of the zeroth-order model (26) for h(θ ,t) at t = 0.1. 4 Stationary waves In order to investigate the polygon-shaped waves having the structure similar to the hexagon as shown in Figure 1, we analyze the model (10)–(14) in the reference frame moving with a wave so that the unknown functions do not depend on time. In this case, the two-dimensional model (10)–(14) for two unknowns ψ (r, θ ) and h (θ ) > 0 is written in non-dimensional variables (16) as ε2 2 ∂ 2ψ ∂ψ 2∂ ψ = 0 (0 < r < h) , + (1 + ε r) + ε (1 + ε r) 2 2 ∂θ ∂r ∂r ψ (0, θ ) = 0, ψ (h, θ ) = Q, (27) (28) ε2 ∂ψ 2 ∂ψ 2 ) + 2gh = 2b, (r = h), ( ) +( 2 (1 + ε r) ∂ θ ∂r (29) ∂h ∂ + (uh) = 0, (r = h) , ∂t ∂θ (30) r where b is the Bernoulli’s constant, Q = h0 u0 is the constant representing a flow rate and, as follows from (15), Γ ln (1 + ε ) . (31) u0 = 2π h0 We first observe that the model (27)–(30) can be reduced to the following relation evaluated at the free boundary: ∂ ∂θ ˆh [(1 + ε r)( 0 ∂ψ 2 ε2 ∂ ψ 2 ∂ψ 2 ε2 ∂ ψ 2 dh ) − ( ) + ) ] = −(1 + ε h)[( ( ) ] . 2 ∂r 1 + εr ∂ θ ∂r (1 + ε h) ∂ θ d θ (32) 162 Ranis N. Ibragimov, Guang Lin /Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 In the stationary case, for the equation (25), we have uh = Q = const. Using this relation, we can exclude u from (24), integrate (32) and substitute the expression for ψ given by (20) to arrive to the following first-order differential equation for the unknown h (θ ) : 2 1 2 2 dh 2 ε2 ε2 ε Q ( ) = − ε h4 + (ε b − 1)h3 + (2b + Q2 )h2 + ( Q2 − c)h + Q2 , 3 dθ 3 4 2 (33) where c is an integrating constant. We remark that in the limiting case ε → 0 (which corresponds to the case of the flat bottom when R → ∞), we obtain the cubic Bouusinesq-Rayleigh equation [42]. We denote by h j ( j = 1, 2, 3, 4) the roots of the polynomial (33). Then the Viète theorem [43] yields the following relation between the parameters Q, c and b are the roots h j ( j = 2, 3, 4) : Q2 = − 23ε + ε2 (h2 h3 + h2 h4 + h3 h4 ) − h2 − h3 − h4 2 ε − 23ε h2 h13 h4 + 34 H − 16 ; 1 1 1 2ε ε c = Q2 [ + + + ] − h2 h3 h4 ; 2 h2 h3 h4 3 ε 1 b = Q2 H + (h2 h3 + h2 h4 + h3 h4 ) , 2 3 where H= 1 1 1 + + . h3 h4 h2 h4 h2 h3 (34) (35) (36) (37) Additionally, a simple perturbation analysis shows that formulae (37) imply that h1 represents a nonphysical solution with the following asymptotic: h1 → ∞ as R → ∞. Namely, h1 = − 3 1 . 2ε h2 h3 h4 (38) and the roots h2 , h3 and h4 are also the roots of the Bouusinesq-Rayleigh equation in the limiting case when ε → 0. The existence of non-trivial wave-like solutions corresponds to the case when all the roots of the polynomial (33) are real and have the values in the interval 0 < h2 h3 < h < h4 . Particularly, we also remark that, since, according to its physical meaning, h is positive and continuous function, the domain of the admissible solution is the interval [h3 , h4 ]. Implicitly, the shape of the free boundary is given by the quadrature εQ θ=√ 3 ˆh h3 ds , (ε s + λ )(s − h2 ) (s − h3 ) (s − h4 ) (39) where λ > 0 is a constant. Namely, according to the relation (38) for h1 , λ= 3 1 . 2 h2 h3 h4 (40) We next introduce a small finite Jacobi amplitude a = h4 − h3 and assume that a is of order ε . In this approximation, the change of variable of integration s = h3 + ξ reduces the integral in (39) to the following asymptotic form:. ˆh h3 ds ˜ (ε s + λ )(s − h2 ) (s − h3 ) (s − h4 ) ˆξ 0 dξ , (ε h3 + λ ) (h3 − h2 ) ξ (1 − ξ ) (41) Ranis N. Ibragimov, Guang Lin /Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 163 which can now be evaluated in terms of elementary functions. As follows from (41), the polygon-shaped wave structure is described by a 2π −periodic nonlinear wave and the wavenumber n is determined by the relation √ 3 (ε h3 + λ )(h3 − h2 ) . (42) n= ε For example, in order to visualize the particular hexagon-shaped path as shown in in Figure 1, we set n = 6 and visualize the gravity waves by detecting the wave trougs and crests by locating the points of minimum and maximum of points of minimum of the integral in (39) as shown in Table 1. Table 1 trougs θ =0 crests π 6 θ= θ= θ= π 3 π 2 θ= θ= 2π 3 5π 6 θ =π θ= 7π 6 θ= θ= 4π 3 9π 6 θ= θ= 4π 3 11π 6 θ = 2π Schematically, the resulting waves for n = 6 can be visualized as shown in Figure 6, where the scales are chosen arbitrarily but the points of trougs and crests correspond to values of θ in Table 1. In terms of the mathematical modeling presented here, the value R should not represent necessarily the radius of the planet, it can be just a radial scale satisfying the relation R/h0 << 1 so that the waves can be observed rotating not only around the planet but also around the polar axis, in clockwise or anticlockwise sense (depending on the sing of Γ) looking above the North Pole. π/2 5π/6 π/6 R θ 11π/6 7π/6 9π/6 Fig. 6 Visualization of the hexagon-shaped shallow gravity longitudinal waves. As seen from the relation (42), the nature of the polygon-shaped free boundary (i.e., the wave number or wave length) is detected by three physical roots h j ( j = 2, 3, 4) of the polynomial equation (33). These roots are also the roots of the Bouusinesq-Rayleigh equation that is obtained in the limiting case when R → ∞. In turns, the four basic parameters, which exert an influence on the roots h j are 164 Ranis N. Ibragimov, Guang Lin /Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 the Bernoulli constant b, flow rate Q, constant of integration c and the parameter ε . The question naturally arises to whether these parameters can be explained physically. We remark that the values ε represents the altitude of the unperturbed atmospheric layer (e.g. for the Earth, it can be associated with troposphere), the value Q is determined by the unperturbed value h0 and the intensity of the vortex localized at the pole of the planet, the Bernoulli constant b can be detected by evaluating the specific atmospheric properties (such as the pressure, density and the velocity field) at certain altitude. Finally, since the parameter c was obtained after the integration of the relation (32), physically it represents the horizontal component of the impulse flux that is also dependent on the specific atmospheric properties. Understanding the structure of longitudinal planetary waves depending of the atmospheric properties have been investigated earlier in meteorological sciences ( [44, 45]). 5 Discussions In summary, we have identified four basic parameters that control the polygon-shaped stationary planetary longitudinal waves. The analysis is based on the higher-order shallow water approximation, which allows to reveal the role of the unperturbed level h0 , which is needed to determine the precision of the first approximation [28,37]. We did not include this analysis in the frame of the present work. However, the preliminary analysis shows that, at the second -order approximation, splitting phenomenon of the shallow water system (24)–(25) takes place. Namely, there are two different approaches in deriving the equation (23). Correspondingly, there are two different forms of the shallow water systems and the 2 difference is observed at 0 ε terms. This analysis will be presented in our forthcoming work and will be published elsewhere. Permanent water waves have been considered in a large number of papers. However, most researchers are concerned with fluid motion which is infinitely deep and extends infinitely both rightward and leftward (see e.g. [46–48] or [49] for the history). In this work, we focused on the deriving the model that uses a curved solid bottom and the circular shape of the unperturbed atmospheric layer in order to visualize a polygon-shaped structure of longitudinal planetary waves. For visualization purposes, Figure 7 presents the same Saturn’s jet stream that follows a hexagon-shaped path at the north pole but from the different perspective. We hope that a polygon-shaped structure and especially the detection of the wave number by the four physical parameters found is this work might also be of interest for scientists who are trying to figure out what causes the hexagon, where it gets and expels its energy and how it has stayed so organized for so long. Finally, we are interested to find approximately invariant solutions of the higher-order shallow water model (24)–(25). We found the exact invariant solution (26) for the limiting case when ε = 0. Methods of classical group analysis allow to single out symmetries with remarkable properties among all equations of mathematical physics. Unfortunately, any small perturbation of an equation breaks the admissible group of transformations and reduces the applied value of these ”refined” equations and group theoretical methods in general. Therefore, development of methods of group analysis stable with respect to small perturbations of differential equations has become vital. In order to find both the approximate symmetries and approximately invariant solutions of the model (24)–(25) with ε = 0, governing equations in question. In order to find approximately invariant solution, we will use the ASP symmetry packages that has been developed originally in [50] and employed in our earlier work [38], in which invariant and approximately invariant solutions of non-linear internal gravity waves forming a column of stratified fluid affected by the Earth’s rotation were found. The ASP symmetry packages is based on the analytic algorithm developed in [51] . The same package has also been incorporated in finding the new class of exact solutions of the Navier-Stokes equations in a thin spherical shell that were associated with Ranis N. Ibragimov, Guang Lin /Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 165 Fig. 7 This view is centered on Saturn’s north pole. North is up and rotated 33 degrees to the left. The image was taken with the Cassini spacecraft wide-angle camera on June 14, 2013 using a spectral filter. nonlinear atmospheric flows with west-to-east jets perturbations that can also be visualized as Kelvin wave that can be seen traveling eastward along the equator as shown in the left panel of Figure ?? (see also [52] ). The goal of these forthcoming studies is to get an explicit analytic expression for h (θ ) without a small-amplitude assumption as has been used in this work. Acknowledgments This research was supported in part by an appointment to the U.S. Department of Energy’s Visiting Faculty Program. References [1] Summerhayes, C.P. and Thorpe, S.A. (1996), Oceanography: An Illustrative Guide, Wiley, New York. [2] Anderson, R.F., Ali, S., Brandtmiller, L.L., Nielsen, S.H., and Fleisher, M.Q. (2009), Wind-driven upwelling in the Southern Ocean and the deglacial rise in atmospheric CO2 , Science, 323, 1443. [3] Yulaeva, E. and Wallace, J.M. (1994), The signature of ENSO in global temperatures and precipitation fields derived from the Microwave Sounding Unit. Journal of Climate, 7, 1719–1736. [4] Blinova, E.N. (1943), A hydrodynamical theory of pressure and temperature waves and of centers of atmospheric action, C. R. Dokl. Acad. Sci.URSS, 39, 257. [5] Blinova, E.N. (1956), A method of solution of the nonlinear problem of atmospheric motions on a planetary scale, Doklady Akademii Nauk SSSR N.S., 110, 975. [6] Boehm, M. T. and Lee, S. (2003), The implications of tropical Rossby waves for tropical tropopause cirrus formation and for the equatorial upwelling of the Brewer–Dobson circulation. Journal of the Atmospheric Sciences, 60, 247–261. [7] Lions, L., Teman, R. and Wang, S. (1992), On the equations of the large-scale ocean, Nonlinearity, 5, 1007. [8] Lions, L., Teman, R. and Wang, S. (1992), New formulations of the primitive equations of atmosphere and applications, Nonlinearity, 5, 237. [9] Ibragimov, N.H. and Ibragimov, R.N. (2011), Integration by quadratures of the nonlinear Euler quations 166 Ranis N. Ibragimov, Guang Lin /Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 modeling atmospheric flows in a thin rotating spherical shell, Physics Letters A, 375, 3858–3865. [10] Ibragimov, R.N. (2011), Nonlinear viscous fluid patterns in a thin rotating spherical domain and applications, Physics of Fluids, 23, 123102. [11] Ibragimov, R.N. and Pelinovsky, D.E. (2009), Incompressible viscous fluid flows in a thin spherical shell, Journal of Mathematical Fluid Mechanics, 11, 60. [12] Ibragimov, R.N., Ibragimov, N.H. and Galiakberova, L.R. (2014), Symmetries and conservation laws of a spectral nonlinear model for atmospheric baroclinic jets, Mathematical Modelling of Natural Phenomena, 9 (5), 32–39. [13] Iftimie, D. and Raugel, G. (2001), Some results on the NS equations in thin three-dimensional domains, Journal of Differential Equations, 169, 281. [14] Weijer, W., Vivier, F., Gille, S.T. and Dijkstra, H. (2007), Multiple oscillatory modes of the Argentine Basin. Part II: The spectral origin of basin modes, Journal of Physical Oceanography, 37, 2869. [15] Bachelor, G.K. (1967), An Introduction to Fluid Dynamics, Cambridge University Press, Cambridge. [16] Herrmann, E. (1896), The motions of the atmosphere and especially its waves, Bulletin of the American Mathematical Society, 2, 285. [17] Holton, J. R., Haynes, P.H., McIntyre, M.E. Douglass, A.R., Rood, C.D. and Pfister, L. (1995), Stratosphere– troposphere exchange. Reviews of Geophysics, 33, 403–439. [18] Toggweiler, R. and Russel, J.L. (2008), Ocean circulation on a warming climate, Nature London, 451, 286. [19] Philanders, S.G. (1978), Variability of the tropical oceans, Dynamics Oceans & Atmos., 3, 191. [20] Ibragimov, R.N. (2010), Free boundary effects on stability of two phase planar fluid motion in annulus: Migration of the stable mode, Communications in Nonlinear Science and Numerical Simulation, 15 (9), 2361–2374. [21] Romea, R.D. and Allen, J.S. (1983), On vertically propagating coastal Kelvin waves at low latitudes. Journal of Physical Oceanography, 13 (1), 241–254. [22] Simmons, A. J. and Hoskins, B.J. (1978), The life cycles of some nonlinear baroclinic waves. Journal of the Atmospheric Sciences, 35, 414–432. [23] Shindell, D.T. and Schmidt, G.A. (2004), Southern hemisphere climate response to ozone changes and greenhouse gas increases, Geophys. Res. Lett., 31, L18209, doi:10.1029/2004GL020724. [24] Aguiar, A.C., Read, P.L., Wordsworth, R.D., Salter, T., and Yamazaki, Y.H. (2010), A laboratory model of Saturn’s North Polar Hexagon, Icarus, 206 (2), 755–763. [25] Grise, K.M. and Thompso, D. W. (2012), Equatorial Planetary Waves and Their Signature in Atmospheric Variability, Journal of the Atmospheric Sciences, 69(3), 857–874. [26] Ibragimov, R.N. (2001), On the tidal motion around the earth complicated by the circular geometry of the ocean’s shape without Coriolis forces, Mathematical Physics, Analysis and Geometry , 4, 51–63. [27] Ibragimov, R.N. and Villasenor, H. (2014), Energy Balance Associated With a Mixing Process at the Interface of a Two-Layer Longitudinal Atmospheric Model, Journal of Fluids Engineering, 136(7), doi:10.1115/1.4026857 [28] Friedrichs, K.O. and Hyers, D.H. (1954), The existence of solitary waves, Communications on Pure and Applied Mathematics , 7. [29] Ibragimov, R.N. and Aitbayev, R. (2009), Three-dimensional non-linear rotating surface waves in channels of variable depth in the presence of formation of a small perturbation of atmospheric pressure across the channel, Communications in Nonlinear Science and Numerical Simulation, 14, 3811–3820. [30] Ibragimov, R.N. and Pelinovsky, D. (2007), Three-dimensional gravity waves in a channel of variable depth, Communications in Nonlinear Science and Numerical Simulation, 13, 2104–2133. [31] Ben-Yu, G. (1995), Spectral method for vorticity equations on spherical surface, Mathematics of Computation, 64, 1067. [32] Swarztrauber, P.N. (2004), Shallow water flow on the sphere, Monthly Weather Review, 132, 3010. [33] Ibragimov, R.N. and Pelinovsky, D.E. (2010), Effects of rotation on stability of viscous stationary flows on a spherical surface, Physics of Fluids, 22, 126602. [34] Astronomy, March 2013. [35] Saber, R. (2012), A Family Brief: the Science Is All In Xlibris Corporation, 2009. [36] Bell, J. et al. (2001), Hubble captures best view of mars ever obtained From Earth. HubbleSite. NASA. http://hubblesite.org/newscenter/archive/releases/2001/24. Retrieved 2010-02-27. [37] Su, C.H. and Gardner, C.S. (1968), Derivation of the Korteweg-de Vries equations and Burgers equations, Jornal of Mathematical Physics, 10, 3. [38] Ibragimov, R.N, Jefferson, G. and Carminati, J. (2013), Invariant and approximately invariant solutions of non-linear internal gravity waves forming a column of stratified fluid affected by the Earth’s rotation, International Journal of Non-Linear Mechanics, 51, 28–44. [39] Ibragimov, R.N, Jefferson, G. and Carminati, J. (2013), Explicit invariant solutions associated with nonlinear Ranis N. Ibragimov, Guang Lin /Journal of Applied Nonlinear Dynamics 4(2) (2015) 153–167 [40] [41] [42] [43] [44] [45] [46] [47] [48] [49] [50] [51] [52] 167 atmospheric fllows in a thin rotating spherical shell with and without west-to-east jets perturbations, Analysis and Mathematical Physics, 3 (3), 201-294. Chesnokov, A.A. (2010), Symmetries of shallow water equations on a rotating plane, Journal of Applied and Industrial Mathematics, 4(1), 24–34. Ovsyannikov, L.V. (1982), Group Analysis of Differential Equations, Nauka, Moscow, (1978), English transl., ed. W.F. Ames, Academic Press, New York, (1982). Spiegal, E.A. and Veronis, G. (1960), On the Boussinesq Approximation for a Compressible Fluid, The Astrophysical Journal, 131, 442. Hazewinkel, Michiel, ed. (2001), Viète theorem, Encyclopedia of Mathematics, Springer, ISBN 978-1-55608010-4. Winson, J.S. (1960), Some new data on the longitudinal dimensions of planetary waves, Journal of the Meteorological Society of Japan, 17, 522–531. Edmon, H.J., Hoskins, J., and McIntyre, M.E. (1980), Eliassen–Palm cross sections for troposphere. Journal of the Atmospheric Sciences, 37, 2600–2616. Crapper, G.D. (1984), Introduction to Water Waves,Ellis Horwood, London. Stoker, J.J. (1957), Water Waves, Interscience Publishers, Inc., New York. Okamoto, H. (1986), Nonstationary free boundary problem for perfect fluid with surface tension, Journal of the Mathematical Society of Japan, 38 (3). Stokes, G.G. (1847), On the theory of oscillatory waves, Translation Cambridge Philosophical Society, 8, 441–455. Vu, K.T., Jefferson, G.F., and Carminati, J. (2012), Finding generalised symmetries of differential equations using the MAPLE package DESOLVII, Computer Physics Communications, 183, 1044–1054. Baikov, V.A., Gazizov, R.K., and Ibragimov, N.H. (1988), Approximate symmetries of equations with a small parameter, Mat. Sb., 136, 435–450 (English translation: Math. USSR-Sb 64 (1989), 427–441). Ibragimov, R.N., Dameron, and Dannangoda, C.(2012), Sources and sinks of energy balance for nonlinear atmospheric motion perturbed by west-to-east winds progressing on a surface of rotating spherical shell, Discontinuity, Nonlinearity and Complexity, 1(1), 41–55.
© Copyright 2026 Paperzz