<<

Chapter 4 Shallow Equations

he are a set of equations that describe, not surprisingly, a shallow layer of , and in particular one that T is in hydrostatic balance and has constant density. The equations are useful for two reasons: (i) They are a simpler set of equations than the full three-dimensional ones, and so allow for a much more straightforward analysis of sometimes complex problems. (ii) In spite of their simplicity, the equations provide a reasonably real- istic representation of a variety of phenomena in atmospheric and oceanic dynamics. Put simply, the shallow water equations are a very useful model for geo- physical . Let’s dive head first into the equations and see what they can do for us.

4.1 Shallow Water Equations of Motion

The shallow water equations apply, by definition, to a fluid layer of con- stant density in which the horizontal scale of the flow is much greater than the layer depth, and which have a at the top (or sometimes at the bottom). Because the fluid is of constant density the fluid motion is fully determined by the and mass continuity equations, and because of the assumed small aspect ratio the hydrostatic approximation is well satisfied, as we discussed in Section 3.2.2. Thus, consider a fluid above which is another fluid of negligible density, as illustrated in Fig. 4.1. Our notation is that 풗 = 푢퐢̂ + 푣퐣̂ + 푤퐤̂ is the three-dimensional and 풖 = 푢퐢̂ + 푣퐣̂ is the horizontal velocity, ℎ(푥, 푦) is the thickness of the column, 퐻 is its mean height, and 휂 is the height of the free surface. In a flat-bottomed container 휂 = ℎ, whereas in general ℎ = 휂 − 휂퐵, where 휂퐵 is the height of the floor of the container. 63 Chapter 4. Shallow Water Equations 64

Fig. 4.1: A shallow water system where ℎ is the thick- T ness of a , 퐻 its mean thickness, 휂 the height

of the free surface and 휂퐵 is the height of the lower, rigid, surface above some arbitrary origin, typically chosen such

that the average of 휂퐵 is zero. B The quantity 휂퐵 is the devia- tion free surface height so we

have 휂 = 휂퐵 + ℎ = 퐻 + 휂푇.

4.1.1 Momentum Equations The vertical momentum equation is just the hydrostatic equation,

∂푝 = −휌 푔, (4.1) ∂푧 0 and, because density is assumed constant, we may integrate this to

푝(푥, 푦, 푧, 푡) = −휌0푔푧 + 푝표. (4.2)

At the top of the fluid, 푧 = 휂, the is determined by the weight of The key assumption underly- the overlying fluid and this is negligible. Thus, 푝 = 0 at 푧 = 휂, giving ing the shallow water equa- tions is that of a small as- 푝(푥, 푦, 푧, 푡) = 휌0푔(휂(푥, 푦, 푡) − 푧). (4.3) pect ratio, so that H/L ≪ 1, where H is the fluid depth The consequence of this is that the horizontal gradient of pressure is indepen- and L the horizontal scale dent of height. That is of motion. This gives rise to the hydrostatic approxima- ∂ ∂ ∇ 푝 = 휌 푔∇ 휂, where ∇ = 퐢̂ + 퐣̂ . (4.4) tion, and this in turn leads 푧 0 푧 푧 ∂푥 ∂푦 to the z-independence of the velocity field and the (In the rest of this chapter we the subscript 푧 unless that causes ambi- ‘sloshing’ nature of the flow. guity; the three-dimensional gradient operator is denoted by ∇3. We also mostly use Cartesian coordinates, but the shallow water equations may certainly be applied over a spherical planet.) The horizontal momentum equations therefore become D풖 1 = − ∇푝 = −푔∇휂. (4.5) D푡 휌0 The right-hand side of this equation is independent of the vertical coordi- nate 푧. Thus, if the flow is initially independent of 푧, it must stay so. (This 푧-independence is unrelated to that arising from the rapid rotation neces- sary for the Taylor–Proudman effect.) The 푢 and 푣 are functions of 푥, 푦 and 푡 only, and the horizontal momentum equation is therefore

D풖 ∂풖 ∂풖 ∂풖 = + 푢 + 푣 = −푔∇휂. (4.6) D푡 ∂푡 ∂푥 ∂푦 65 4.1 Shallow Water Equations of Motion

Fig. 4.2: The mass budget for a column of 퐴 in a flat-bottomed shallow water system. The fluid leaving the column is ∮ 휌ℎ풖 ⋅ 풏 d푙 where 풏 is the unit vector normal to the boundary of the fluid column. There is a non-zero vertical velocity at the top of the column if the mass convergence into the column is non-zero.

In the presence of rotation, (4.6) easily generalizes to

D풖 + 풇 × 풖 = −푔∇휂, (4.7) D푡 where 풇 = 푓퐤̂ . Just as with the fully three-dimensional equations, 푓 may be constant or may vary with latitude, so that on a spherical planet 푓 = 2훺 sin 휗 and on the 훽-plane 푓 = 푓0 + 훽푦.

4.1.2 Mass The mass contained in a fluid column of height ℎ and cross-sectional area 퐴 is given by ∫ 휌 ℎ d푨 (see Fig. 4.2). If there is a net flux of fluid across 퐴 0 the column boundary (by ) then this must be balanced by a net increase in the mass in 퐴, and therefore a net increase in the height of the water column. The mass convergence into the column is given by

퐹푚 = mass flux in = − ∫ 휌0풖 ⋅ d푺, (4.8) 푆 where 푆 is the area of the vertical boundary of the column. The surface area of the column is composed of elements of area ℎ풏 훿푙, where 훿푙 is a line element circumscribing the column and 풏 is a unit vector perpendicular to the boundary, pointing outwards. Thus (4.8) becomes

퐹푚 = − ∮ 휌0ℎ풖 ⋅ 풏 d푙. (4.9)

Using the divergence theorem in two dimensions, (4.9) simplifies to

퐹푚 = − ∫ ∇ ⋅ (휌0풖ℎ) d퐴, (4.10) 퐴 where the integral is over the cross-sectional area of the fluid column (looking down from above). This is balanced by the local increase in height of the water column, given by d d ∂ℎ 퐹푚 = ∫ 휌0 d푉 = ∫ 휌0ℎ d퐴 = ∫ 휌0 d퐴. (4.11) d푡 d푡 퐴 퐴 ∂푡 Chapter 4. Shallow Water Equations 66

The Shallow Water Equations For a single-layer fluid, and including the term, the inviscid shallow water equations are: D풖 momentum: + 풇 × 풖 = −푔∇휂, (SW.1) D푡 Dℎ mass continuity: + ℎ∇ ⋅ 풖 = 0, (SW.2) D푡 ∂ℎ or + ∇ ⋅ (ℎ풖) = 0, (SW.3) ∂푡 where 풖 is the horizontal velocity, ℎ is the total fluid thickness, 휂 is the height of the upper free surface, and ℎ and 휂 are related by

ℎ(푥, 푦, 푡) = 휂(푥, 푦, 푡) − 휂퐵(푥, 푦), (SW.4)

where 휂퐵 is the height of the lower surface (the bottom topogra- phy). The material derivative is

D ∂ ∂ ∂ ∂ = + 풖 ⋅ ∇ = + 푢 + 푣 , (SW.5) D푡 ∂푡 ∂푡 ∂푥 ∂푦 with the rightmost expression holding in Cartesian coordinates.

Because 휌0 is constant, the balance between (4.10) and (4.11) leads to ∂ℎ ∫ [ + ∇ ⋅ (풖ℎ)] d퐴 = 0, (4.12) 퐴 ∂푡 and because the area is arbitrary the integrand itself must vanish, whence,

∂ℎ Dℎ + ∇ ⋅ (풖ℎ) = 0 or + ℎ∇ ⋅ 풖 = 0. (4.13a,b) ∂푡 D푡

This derivation holds whether or not the lower surface is flat. If it is, then ℎ = 휂, and if not ℎ = 휂 − 휂퐵. Equations (4.7) and (4.13) form a complete set, summarized in the shaded box above.

4.1.3 Reduced Equations Consider now a single shallow moving layer of fluid on top of a deep, qui- escent fluid layer (Fig. 4.3), and beneath a fluid of negligible inertia. This configuration is often used as a model of the upper : the upper layer represents flow in perhaps the upper few hundred metres of the ocean, the lower layer being the near-stagnant abyss. If we turn the model upside- down we have a perhaps slightly less realistic model of the atmosphere: the lower layer represents motion in the troposphere above which lies an 67 4.1 Shallow Water Equations of Motion

Fig. 4.3: The reduced gravity shallow water system. An active layer lies over a deep, denser, quiescent layer. In a common variation the upper surface is held flat by a rigid

lid, and 휂0 = 0.

inactive stratosphere. The equations of motion are virtually the same in both cases, but for definiteness we’ll think about the oceanic case. The pressure in the upper layer is given by integrating the hydrostatic equation down from the upper surface. Thus, at a height 푧 in the upper layer 푝1(푧) = 푔휌1(휂0 − 푧), (4.14) where 휂0 is the height of the upper surface. Hence, everywhere in the upper layer, 1 ∇푝1 = 푔∇휂0, (4.15) 휌1 and the momentum equation is D풖 + 풇 × 풖 = −푔∇휂 . (4.16) D푡 0 In the lower layer the pressure is also given by the weight of the fluid above it. Thus, at some level 푧 in the lower layer,

푝2(푧) = 휌1푔(휂0 − 휂1) + 휌2푔(휂1 − 푧). (4.17) But if this layer is motionless the horizontal in it is zero and therefore ′ 휌1푔휂0 = −휌1푔 휂1 + constant, (4.18) ′ where 푔 = 푔(휌2 − 휌1)/휌1 is the reduced gravity, and in the ocean 휌2 − ′ 휌1)/휌 ≪ 1 and 푔 ≪ 푔. The momentum equation becomes D풖 + 풇 × 풖 = 푔′∇휂 . (4.19) D푡 1 The equations are completed by the usual mass conservation equation, Dℎ + ℎ∇ ⋅ 풖 = 0, (4.20) D푡 ′ where ℎ = 휂0 − 휂1. Because 푔 ≫ 푔 ,(4.18) shows that surface displace- ments are much smaller than the displacements at the interior interface. We see this in the real ocean where the mean interior isopycnal displace- ments may be several tens of metres but variations in the mean height of the ocean surface are of the order of centimetres. Chapter 4. Shallow Water Equations 68

4.2 Conservation Properties

There are two common types of conservation property in : (i) ma- terial invariants; and (ii) integral invariants. Material invariance occurs when a property (휑 say) is conserved on each fluid element, and so obeys the equation D휑/D푡 = 0. An integral invariant is one that is conserved af- ter an integration over some, usually closed, volume; energy is an example. The simplicity of the shallow water equations allows us to transparently see how these arise.

4.2.1 Energy Conservation: an Integral Invariant Since we have made various simplifications in deriving the shallow water system, it is not self-evident that energy should be conserved, or indeed what form the energy takes. The density (KE), meaning the 2 kinetic energy per unit area, is 휌0ℎ풖 /2. The potential energy density (PE) of the fluid is ℎ 1 2 PE = ∫ 휌0푔푧 d푧 = 휌0푔ℎ . (4.21) 0 2

The factor 휌0 appears in both kinetic and potential energies and, because it is a constant, we will omit it. For algebraic simplicity we also assume the bottom is flat, at 푧 = 0. Energy is moved from place to place by pressure as Using the mass conservation equation (4.13b) we obtain an equation well as advection, and is not a for the evolution of potential energy density, namely material invariant. However, D 푔ℎ2 in total energy is conserved + 푔ℎ2∇ ⋅ 풖 = 0 (4.22a) and it is an integral invariant. D푡 2 or ∂ 푔ℎ2 푔ℎ2 푔ℎ2 + ∇ ⋅ (풖 ) + ∇ ⋅ 풖 = 0. (4.22b) ∂푡 2 2 2 From the momentum and mass continuity equations we obtain an equa- tion for the evolution of kinetic energy density, namely

D ℎ풖2 풖2ℎ ℎ2 + ∇ ⋅ 풖 = −푔풖 ⋅ ∇ (4.23a) D푡 2 2 2 or ∂ ℎ풖2 ℎ풖2 ℎ2 + ∇ ⋅ (풖 ) + 푔풖 ⋅ ∇ = 0. (4.23b) ∂푡 2 2 2 Adding (4.22b) and (4.23b) we obtain

∂ 1 1 (ℎ풖2 + 푔ℎ2) + ∇ ⋅ [ 풖 (푔ℎ2 + ℎ풖2 + 푔ℎ2)] = 0 (4.24) ∂푡 2 2 or ∂퐸 + ∇ ⋅ 푭 = 0, (4.25) ∂푡 where 퐸 = KE + PE = (ℎ풖2 + 푔ℎ2)/2 is the density of the total energy and 푭 = 풖(ℎ풖2/2 + 푔ℎ2) is the energy flux. If the fluid is confined to a domain 69 4.2 Conservation Properties bounded by rigid walls, on which the normal component of velocity van- ishes, then on integrating (4.24) over that area and using Gauss’s theorem, the total energy is seen to be conserved; that is d퐸̂ 1 d = ∫ (ℎ풖2 + 푔ℎ2) d퐴 = 0. (4.26) d푡 2 d푡 퐴 Such an energy principle also holds in the case with bottom topography. Note that, as we found in the case for a compressible fluid in Chapter 2, the energy flux in (4.25) is not just the energy density multiplied by the velocity; it contains an additional term 푔풖ℎ2/2, and this represents the energy transfer occurring when the fluid does work against the pressure .

4.2.2 Potential : a Material Invariant The vorticity of a fluid, denoted 흎, is defined to be the curl of the velocity field. Let us also define the shallow water vorticity, 흎∗, as the curl of the horizontal velocity. We therefore have: 흎 ≡ ∇ × 풗, 흎∗ ≡ ∇ × 풖. (4.27) Because ∂푢/∂푧 = ∂푣/∂푧 = 0, only the vertical component of 흎∗ is non- zero and ∂푣 ∂푢 흎∗ = 퐤̂ ( − ) = 퐤̂ 휁. (4.28) ∂푥 ∂푦 Considering first the non-rotating case, we use the vector identity 1 (풖 ⋅ ∇)풖 = ∇(풖 ⋅ 풖) − 풖 × (∇ × 풖), (4.29) 2 to write the momentum equation, (4.7) as ∂풖 1 + 흎∗ × 풖 + 풇 × 풖 = −∇ (푔휂 + 풖2). (4.30) ∂푡 2 To obtain an evolution equation for the vorticity we take the curl of (4.30), and make use of the vector identity

∇ × (흎∗ × 풖) = (풖 ⋅ ∇)흎∗ − (흎∗ ⋅ ∇)풖 + 흎∗∇ ⋅ 풖 − 풖∇ ⋅ 흎∗ = (풖 ⋅ ∇)흎∗ + 흎∗∇ ⋅ 풖, (4.31) using the fact that ∇⋅흎∗ is the divergence of a curl and therefore zero, and (흎∗ ⋅ ∇)풖 = 0 because 흎∗ is perpendicular to the surface in which 풖 varies. A similar expression holds for 풇 so that taking the curl of (4.30) gives ∂휁 + (풖 ⋅ ∇)(휁 + 푓) = −(휁 + 푓)∇ ⋅ 풖, (4.32) ∂푡 where 휁 = 퐤̂ ⋅ 흎∗ and 푓 varies with latitude (and on the beta-plane 푓 = 푓0 + 훽푦). Since 푓 does not vary with time we can write (4.32) as D (휁 + 푓) = −(휁 + 푓)∇ ⋅ 풖. (4.33) D푡 Chapter 4. Shallow Water Equations 70

The mass conservation equation, (4.13b) may be written as 휁 + 푓 Dℎ − (휁 + 푓)∇ ⋅ 풖 = , (4.34) ℎ D푡 and using this equation and (4.32) we obtain D 휁 + 푓 Dℎ (휁 + 푓) = , (4.35) D푡 ℎ D푡 which is equivalent to

D푄 휁 + 푓 = 0 where 푄 = ( ). (4.36) D푡 ℎ

The important quantity 푄 is known as the , and (4.36) is the potential vorticity equation.

4.3 Shallow Water

Let us now look at the gravity waves that occur in shallow water. To iso- late the essence we consider waves in a single fluid layer, with a flat bot- tom and a free upper surface, in which gravity provides the sole restoring force.

4.3.1 Non-Rotating Shallow Water Waves Given a flat bottom the fluid thickness is equal to the free surface displace- ment (Fig. 4.1), and taking the basic state of the fluid to be at rest we let

ℎ(푥, 푦, 푡) = 퐻 + ℎ′(푥, 푦, 푡) = 퐻 + 휂′(푥, 푦, 푡), (4.37a) 풖(푥, 푦, 푡) = 풖′(푥, 푦, 푡). (4.37b) The mass conservation equation, (4.13b), then becomes ∂휂′ + (퐻 + 휂′)∇ ⋅ 풖′ + 풖′ ⋅ ∇휂′ = 0, (4.38) ∂푡 and neglecting squares of small quantities this yields the linear equation ∂휂′ + 퐻∇ ⋅ 풖′ = 0. (4.39) ∂푡 Similarly, linearizing the momentum equation, (4.7) with 풇 = 0, yields ∂풖′ = −푔∇휂′. (4.40) ∂푡 Eliminating velocity by differentiating (4.39) with respect to time and taking the divergence of (4.40) leads to ∂2휂′ − 푔퐻∇2휂′ = 0, (4.41) ∂푡2 71 4.3 Shallow Water Waves which may be recognized as a equation. We can find the relationship for this by substituting the trial solution

휂′ = Re ̃휂ei(풌⋅풙−휔푡), (4.42) where ̃휂 is a complex constant, 풌 = 푘퐢̂ + 푙퐣̂ is the horizontal and Re indicates that the real part of the solution should be taken. If, for simplicity, we restrict attention to the one-dimensional problem, with no variation in the 푦-direction, then substituting into (4.41) leads to the dispersion relationship 휔 = ±푐푘, (4.43) where 푐 = √푔퐻; that is, the wave speed is proportional to the square root of the mean fluid depth and is independent of the wavenumber — the waves are dispersionless. The general solution is a superposition of all such waves, with the of each wave (or Fourier component) being determined by the Fourier decomposition of the initial conditions. Because the waves are dispersionless, the general solution can be writ- ten as 1 휂′(푥, 푡) = [퐹(푥 − 푐푡) + 퐹(푥 + 푐푡)] , (4.44) 2 where 퐹(푥) is the height field at 푡 = 0. From this, it is easy to see that the shape of an initial disturbance is preserved as it propagates both to the right and to the left at speed 푐.

4.3.2 Rotating Shallow Water (Poincaré) Waves We now consider the effects of rotation on shallow water waves. Lineariz- ing the rotating, flat-bottomed 푓-plane shallow water equations, (SW.1) and (SW.2) on page 66, about a state of rest we obtain

∂푢′ ∂휂′ ∂푣′ ∂휂′ − 푓 푣′ = −푔 , + 푓 푢′ = −푔 , (4.45a,b) ∂푡 0 ∂푥 ∂푡 0 ∂푦

∂휂′ ∂푢′ ∂푣′ + 퐻 ( + ) = 0. (4.45c) ∂푡 ∂푥 ∂푦 To obtain a dispersion relationship we let

(푢, 푣, 휂) = (̃푢,̃푣,̃휂)ei(풌⋅풙−휔푡), (4.46) and substitute into (4.45), giving

−i 휔 −푓0 i 푔푘 ̃푢 ( 푓0 −i 휔 i 푔푙 )(̃푣) = 0. (4.47) i 퐻푘 i 퐻푙 −i 휔 ̃휂

This homogeneous equation has non-trivial solutions only if the determi- nant of the matrix vanishes, and that condition gives

2 2 2 2 휔(휔 − 푓0 − 푐 퐾 ) = 0, (4.48) Chapter 4. Shallow Water Equations 72

Fig. 4.4: Dispersion rela- tion for Poincaré waves Rotating (Poincaré) and non-rotating shallow Non-rotating water waves. is 3 scaled by the Coriolis fre- quency 푓, and wavenumber by the inverse deformation 2 radius √푔퐻/푓. For small the frequency of the Poincaré waves is ap- proximately 푓, and for high 1 wavenumbers it asymptotes to that of non-rotating waves.

0 ‐3 ‐2 ‐1 0 1 2 3

where 퐾2 = 푘2 + 푙2 and 푐2 = 푔퐻. There are two classes of solution to (4.48). The first is simply 휔 = 0, i.e., time-independent flow correspond- ing to geostrophic balance in (4.45). Because geostrophic balance gives a divergence-free velocity field for a constant Coriolis parameter the equa- tions are satisfied by a time-independent solution. The second set of so- lutions gives the 2 2 2 2 2 휔 = 푓0 + 푐 (푘 + 푙 ), (4.49) or 2 2 2 2 휔 = 푓0 + 푔퐻(푘 + 푙 ). (4.50) The corresponding waves are known as Poincaré waves, and the disper- (Jules) Henri Poincaré (1854– sion relationship is illustrated in Fig. 4.4. Note that the frequency is al- 1912) was a remarkable ways greater than the Coriolis frequency 푓0. There are two interesting French mathematician, physi- limits: cist and philosopher. As well as the Poincaré waves of this (i) The short wave limit. If chapter he is also known as 푓2 one of the founders of what is 퐾2 ≫ 0 , (4.51) now known as chaos theory 푔퐻 — he knew that the weather where 퐾2 = 푘2 + 푙2, then the dispersion relationship reduces to was inherently, and not just that of the non-rotating case (4.43). This condition is equivalent to in practice, unpredictable. requiring that the be much shorter than the deformation radius, defined by √푔퐻 퐿 = . (4.52) 푑 푓 Specifically, if 푙 = 0 and 휆 = 2π/푘 is the wavelength, the condition is 2 2 2 2 휆 ≪ 퐿푑(2π) . The numerical factor of (2π) is more than an order of magnitude, so care must be taken when deciding if the condition is satisfied in particular cases. Furthermore, the wavelength must still be longer than the depth of the fluid, otherwise the shallow wa- ter condition is not met. 73 4.3 Shallow Water Waves

(ii) The long wave limit. If 푓2 퐾2 ≪ 0 , (4.53) 푔퐻 that is if the wavelength is much longer than the deformation radius 퐿푑, then the dispersion relationship is

휔 = 푓0. (4.54) These waves are known as inertial oscillations. The equations of mo- tion giving rise to them are

∂푢′ ∂푣′ − 푓 푣′ = 0, + 푓 푢′ = 0, (4.55) ∂푡 0 ∂푡 0 which are equivalent to material equations for free particles in a rotating frame, unconstrained by pressure forces.

4.3.3 Kelvin Waves The is a particular type of that exists in the pres- ence of both rotation and a lateral boundary. Suppose there is a solid boundary at 푦 = 0; general harmonic solutions in the 푦-direction are not allowable, as these would not satisfy the condition of no normal flow at the boundary. Do any wave-like solutions exist? The answer is yes, and to The affirmative answer to the show that we begin with the linearized shallow water equations, namely question in the text was pro- ′ ′ ′ ′ vided by William Thomson in ∂푢 ′ ∂휂 ∂푣 ′ ∂휂 the nineteenth century. Thom- − 푓0푣 = −푔 , + 푓0푢 = −푔 , (4.56a,b) ∂푡 ∂푥 ∂푡 ∂푦 son later became Lord Kelvin, taking the name of the river ∂휂′ ∂푢′ ∂푣′ + 퐻 ( + ) = 0. (4.56c) flowing near his university in ∂푡 ∂푥 ∂푦 Glasgow, and the waves are ′ now eponymously known as The fact that 푣 = 0 at 푦 = 0 suggests that we look for a solution with Kelvin waves. 푣′ = 0 everywhere, whence these equations become

∂푢′ ∂휂′ ∂휂′ ∂휂′ ∂푢′ = −푔 , 푓 푢′ = −푔 , + 퐻 = 0. (4.57a,b,c) ∂푡 ∂푥 0 ∂푦 ∂푡 ∂푥 Equations (4.57a) and (4.57c) lead to the standard wave equation

∂2푢′ ∂2푢′ = 푐2 , (4.58) ∂푡2 ∂푥2 where 푐 = √푔퐻, the usual wave speed of shallow water waves. The solu- tion of (4.58) is ′ 푢 = 퐹1(푥 + 푐푡, 푦) + 퐹2(푥 − 푐푡, 푦), (4.59) with corresponding surface displacement

′ 휂 = √퐻/푔 [−퐹1(푥 + 푐푡, 푦) + 퐹2(푥 − 푐푡, 푦)] . (4.60)

The solution represents the superposition of two waves, one (퐹1) travel- Chapter 4. Shallow Water Equations 74

ling in the negative 푥-direction, and the other in the positive 푥-direction. To obtain the 푦 dependence of these functions we use (4.57b) which gives ∂퐹 푓 ∂퐹 푓 1 = 0 퐹 , 2 = − 0 퐹 , (4.61) ∂푦 √푔퐻 1 ∂푦 √푔퐻 2 with solutions

푦/퐿푑 −푦/퐿푑 퐹1 = 퐹(푥 + 푐푡)e , 퐹2 = 퐺(푥 − 푐푡)e , (4.62)

where 퐿푑 = √푔퐻/푓0 is the radius of deformation. If we consider flow in the half-plane in which 푦 > 0, then for positive 푓0 the solution 퐹1 grows exponentially away from the wall, and so fails to satisfy the condition of boundedness at infinity. It thus must be eliminated, leaving the general solution

Wave magnitude ′ −푦/퐿 ′ 푢 = e 푑 퐺(푥 − 푐푡), 푣 = 0, (4.63a,b,c) 휂′ = √퐻/푔e−푦/퐿푑 퐺(푥 − 푐푡).

Equator These are Kelvin waves and they decay exponentially away from the boundary. In general, for 푓0 positive (negative) the boundary is to the right (left) of the wave direction. Given a constant Coriolis parameter, we could also have obtained a solution on a meridional wall, again with y the wave moving with the wall on its the right if 푓0 is positive. The equa- tor also acts like a wall, with Kelvin waves propagating eastward along it, x decaying to either side, as in Fig. 4.5. (The details of the solution for equa- torial Kelvin waves differ from the case with 푓 = 푓0, because now 푓 = 0 Fig. 4.5: Kelvin waves prop- at the equator itself, but the structure is similar to the solution above.) agating eastward at the equator and decaying rapidly away to either side. 4.4 Geostrophic Adjustment Geostrophic balance occurs in the shallow water equations when the 푈/푓퐿 is small and the Coriolis term dominates the advec- tive terms in the momentum equation. In the single-layer shallow water equations the geostrophic flow is:

풇 × 풖푔 = −푔∇휂. (4.64) Thus, the geostrophic velocity is proportional to the slope of the surface, as sketched in Fig. 4.6. The result arises because by hydrostatic balance the slope of an interfacial surface is directly related to the difference in pressure gradient on either side, so that in geostrophic balance the veloc- ity is related to the slope. But consider the more general question, why are the atmosphere and ocean close to geostrophic balance? Suppose that the initial state were not balanced, that it had small scale motion and sharp pressure gradients. How would the system evolve? It turns out that all of these sharp gradients would radiate away, leaving behind a geostrophically balanced state. The process is called geostrophic adjustment, and it occurs quite generally in rotating fluids, whether stratified or not. The shallow water equations are an ideal set of equations to explore this process, as we now see. 75 4.4 Geostrophic Adjustment

 

            

       

Fig. 4.6: Geostrophic flow 4.4.1 Posing the Problem in a shallow water system, with a positive value of the Let us consider the free evolution of a single shallow layer of fluid whose Coriolis parameter 푓, as in initial state is manifestly unbalanced, and we suppose that surface dis- the Northern Hemisphere. placements are small so that the evolution of the system is described by The pressure force is directed the linearized shallow equations of motion. These are down the gradient of the height field, and this can ∂풖 ∂휂 + 풇 × 풖 = −푔∇휂, + 퐻∇ ⋅ 풖 = 0, (4.65a,b) be balanced by the Coriolis ∂푡 ∂푡 force if the fluid velocity is at right angles to it. If were where 휂 is the free surface displacement and 퐻 is the mean fluid depth, 푓 negative, the geostrophic and we omit the primes on the linearized variables. flow would be reversed. 4.4.2 Non-Rotating Flow We consider first the non-rotating problem set, with little loss of gener- ality, in one dimension. We suppose that initially the fluid is at rest but with a simple discontinuity in the height field so that

+휂 푥 < 0 휂(푥, 푡 = 0) = { 0 (4.66) −휂0 푥 > 0, and 푢(푥, 푡 = 0) = 0 everywhere. We can realize these initial conditions physically by separating two fluid masses of different depths by a thin dividing wall, and then quickly removing the wall. What is the subsequent evolution of the fluid? The general solution to the linear problem is given by (4.44) where the functional form is determined by the initial conditions so that here 퐹(푥) = 휂(푥, 푡 = 0) = −휂0 sgn(푥). (4.67) Equation (4.44) states that this initial pattern is propagated to the right and to the left. That is, two discontinuities in fluid height move to the right and left at a speed 푐 = √푔퐻. Specifically, the solution is 1 휂(푥, 푡) = − 휂 [sgn(푥 + 푐푡) + sgn(푥 − 푐푡)]. (4.68) 2 0 The initial conditions may be much more complex than a simple front, but, because the waves are dispersionless, the solution is still simply a sum Chapter 4.(FPTUSPQIJD Shallow "EKVTUNFOU Water Equations 76

rotating non-rotating 1.0 1.0 0.5 0.5 0.0 0.0 height 0.5 0.5 1.0 1.0 8 4 0 4 8 8 4 0 4 8 0.5 0.5 0.0 0.0 0.5 0.5

v velocity v 1.0 1.0 1.5 1.5 8 4 0 4 8 8 4 0 4 8 1.0 1.0

0.5 0.5

0.0 0.0 u velocity u

0.5 0.5 8 4 0 4 8 8 4 0 4 8 x/L x/L

Fig.'JH 4.7: TheʾF solutions TPMVUJPOT of PG UIF TIBMMPXof the translationXBUFS FRVBUJPOT of PCUBJOFEthose initial CZ OVNFSJDBMMZ conditions JOUFHS toBUJOH the right UIF FRVBUJPOT and to the left at the shallowPG NPUJPO water XJUI equations BOE XJUIPVU SPUBUJPOspeed ʾF푐. The QBOFMT velocity TIPX TOBQTIPUT field in PG this UIF class TUBUF ofPG UIF problem nVJE TPMJE is obtained MJOFT TPPO from obtainedBGUFS CFJOH by numerically SFMFBTFE GSPN in- B TUBUJPOBSZ JOJUJBM TUBUF SFE EBTIFE MJOFT XJUI B IFJHIU EJTDPOUJOVJUZ ʾF SPUBU tegratingJOH nPX the JT FWPMWJOH equations UPXBSE of BO FOE TUBUF TJNJMBS UP 'JH  XIFSFBT∂푢 UIF OPOSP∂휂UBUJOH nPX XJMM FWFOUVBMMZ CFDPNF TUBUJPOBSZ *O UIF OPOSPUBUJOH DBTF JT EFmOFE VTJOH UIF SPUBUJOH= −푔 QBSBNFUFST,  (4.69) motion with and without ৴ਆ ∂푡 ∂푥 rotation. The panels show snapshots of the state of the which gives, using (4.44), TP UIBU FOFSHZ DPOTFSWBUJPO IPMET JO UIF GPSN fluid (solid lines) soon after 푔 being released from a station- 푢 = − E[퐹(푥 + 푐푡) − 퐹(푥 − 푐푡)]. (4.70)    E 2푐 ৭  ary initial state (red dashed ৭ DŽ ৰ੊ ਉ૙ ੍ E    ਕ lines) with a height discon- Consider the case with initial conditions given by (4.66). At a given tinuity.QSPWJEFE The rotating UIF JOUFHSBM flow PG is UIF EJWFSHFODF UFSN WBOJTIFT BT JU OPSNBMMZ XJMM JO B DMPTFE EPNBJO ćF ĘVJE IBT B OPO[FSP QPUFOUJBMlocation, FOFSHZ awayú from theE initial JG UIFSF disturbance, BSF WBSJBUJPOT the JO ĘVJE fluid IFJHIU remains BOE at rest and evolving toward an end state  þ÷ú ਉ૙ ਙ similarXJUI UIFto Fig. JOJUJBM4.8 whereas DPOEJUJPOT  undisturbed UIF JOJUJBM until QPUFOUJBM the front FOFSHZ arrives. JT After the front has passed, the fluid the non-rotating flow will surface is again undisturbed and the velocity is uniform and non-zero. ú eventually become station- Specifically:  E  ৸৭ৱ  DŽ ਉ૙ ਙ ary. In the non-rotating  case 퐿 is defined using −휂 sgn(푥) 0 |푥| > 푐푡 ćJT JT푑 OPNJOBMMZ JOĕOJUF JG UIF ĘVJE IBT OP휂 CPVOEBSJFT = { 0 BOE UIF JOJUJBM푢 QPUFOUJBM= { FOFSHZ EFOTJUZ JT (4.71) the rotatingFWFSZXIFSF parameters. 0 (휂 푔/푐) |푥| < 푐푡. ਉ૙ 0 *O UIF OPOSPUBUJOH DBTF BOE XJUI JOJUJBM DPOEJUJPOT  BęFS UIF GSPOU IBT QBTTFE UIF QP UFOUJBM FOFSHZ EFOTJUZ JT [FSP BOEThe UIF LJOFUJD solution FOFSHZ with EFOTJUZ a discontinuity JT  in the height VTJOH field,  and BOE zero initial ৰਖ   ਉ૙   ćVT BMM UIF QPUFOUJBM FOFSHZ JT MPDBMMZ DPOWFSUFE UP LJOFUJD FOFSHZ BT UIF GSPOU QBTTFT ਅ  ਉৰ velocity, is illustrated in the right-hand panels in Fig. 4.7. The front prop- BOE FWFOUVBMMZ UIF LJOFUJD FOFSHZagates JT in EJTUSJCVUFE either direction VOJGPSNMZ from BMPOH the UIF discontinuity MJOF *O UIF and, DBTF in JMMVTUSBUFE this case, the final JO 'JH  UIF QPUFOUJBM FOFSHZvelocity, BOE LJOFUJD as well FOFSHZ as the BSF fluid CPUI displacement, SBEJBUFE BXBZ is GSPN zero. UIF That JOJUJBM is, EJTUVSthe disturbance CBODF /PUF UIBU BMUIPVHI XFis radiated DBO TVQFSQPTF completely UIF TPMVUJPOT away. GSPN EJČFSFOU JOJUJBM DPOEJUJPOT XF DBOOPU TVQFSQPTF UIFJS QPUFOUJBM BOE LJOFUJD FOFSHJFT ćF HFOFSBM QPJOU JT UIBU UIF FWPMVUJPO PG UIF EJTUVSCBODF JT OPU DPOĕOFE UP JUT JOJUJBM MPDBUJPO 4.4.3 Rotating Flow *O DPOUSBTU JO UIF SPUBUJOH DBTF UIF DPOWFSTJPO GSPN QPUFOUJBM UP LJOFUJD FOFSHZ JT MBSHFMZ DPO ĕOFE UP XJUIJO B EFGPSNBUJPORotation SBEJVT PG UIF makes JOJUJBM a profound EJTUVSCBODF differenceBOE BU MPDBUJPOT to the GBSadjustment GSPN UIF JOJUJBMproblem of the EJTUVSCBODF UIF JOJUJBM TUBUF JTshallow FTTFOUJBMMZ water VOBMUFSFE system, ćF because DPOTFSWBUJPO a steady, PG QPUFOUJBM adjusted, WPSUJDJUZ solution IBT can QSF exist with WFOUFE UIF DPNQMFUF DPOWFSTJPOnon-zero PG QPUFOUJBM gradients FOFSHZ in UP the LJOFUJD height FOFSHZ field — B SFTVMU the associated UIBU JT OPU pressure TFOTJUJWF gradients UP UIF QSFDJTF GPSN PG UIF JOJUJBMbeing DPOEJUJPOT balanced by the — and potential vorticity conserva- tion provides a powerful constraint on the fluid evolution. In a rotating 77 4.4 Geostrophic Adjustment shallow fluid that conservation is represented by ∂푄 + 풖 ⋅ ∇푄 = 0, (4.72) ∂푡 where 푄 = (휁 + 푓)/ℎ. In the linear case with constant Coriolis parameter, (4.72) becomes ∂푞 휂 = 0, 푞 = (휁 − 푓 ). (4.73) ∂푡 0 퐻 This equation may be obtained either from the linearized velocity and mass conservation equations, (4.65), or from (4.72) directly. In the latter case we can write 휁 + 푓 1 휂 1 휂 푓 푞 푄 = 0 ≈ (휁+푓 ) (1 − ) ≈ (푓 + 휁 − 푓 ) = 0 + , (4.74) 퐻 + 휂 퐻 0 퐻 퐻 0 0 퐻 퐻 퐻 having used 푓0 ≫ |휁| and 퐻 ≫ |휂|. The term 푓0/퐻 is a constant and so dynamically unimportant, as is the 퐻−1 factor multiplying 푞. Further, the advective term 풖⋅∇푄 becomes 풖⋅∇푞, and this is second order in perturbed quantities and so is neglected. Thus, making these approximations, (4.72) reduces to (4.73). The potential vorticity field is therefore fixed in space! Of course, this was also true in the non-rotating case where the fluid is initially at rest. Then 푞 = 휁 = 0 and the fluid remains irrotational through- out the subsequent evolution of the flow. However, this is rather a weak constraint on the subsequent evolution of the fluid; it does nothing, for example, to prevent the conversion of all the potential energy to kinetic energy. In the rotating case the potential vorticity is non-zero, and poten- tial vorticity conservation and geostrophic balance are all we need to infer the final steady state, assuming it exists, without solving for the details of the flow evolution, as we now see. With an initial condition for the height field given by (4.66), the initial potential vorticity is given by If an unbalanced flow is left to freely evolve, gravity −푓 휂 /퐻 푥 < 0 waves will propagate quickly 푞(푥, 푦) = { 0 0 (4.75) 푓 휂 /퐻 푥 > 0, away from the disturbance. 0 0 However, the balanced com- and this remains unchanged throughout the adjustment process. The fi- ponent of the flow is con- nal steady state is then the solution of the equations strained by potential vorticity conservation and it can only 휂 ∂휂 ∂휂 휁 − 푓 = 푞(푥, 푦), 푓 푢 = −푔 , 푓 푣 = 푔 , (4.76a,b,c) evolve advectively, and so 0 퐻 0 ∂푦 0 ∂푥 much more slowly. Thus, after some time, only the where 휁 = ∂푣/∂푥−∂푢/∂푦. Because the Coriolis parameter is constant, the balanced part of the flow re- velocity field is horizontally non-divergent and we may define a stream- mains, with the unbalanced 휓 = 푔휂/푓0. Equations (4.76) then reduce to flow having been radiated away and dissipated. This pro- 2 1 cess is geostrophic adjustment. (∇ − 2 ) 휓 = 푞(푥, 푦), (4.77) 퐿푑 where 퐿푑 = √푔퐻/푓0 is the radius of deformation, as in (4.52), sometimes called the ‘Rossby radius’. It is a naturally occurring length scale in prob- lems involving both rotation and gravity, and arises in a slightly different form in stratified fluids.  Chapter 4. Shallow Water Equations $IBQUFS  4IBMMPX 8BUFS 4ZTUFNT78 Fig. 4.8: Solutions of a lin- ear geostrophic adjustment B problem. (a) Initial height field, given by (4.66) with

휂0 = 1. (b) Equilibrium (fi- - nal) height field, 휂 given by

(4.79) and 휂 = 푓0휓/푔. (c) C 'JH  4PMVUJPOTEquilibrium PG geostrophic B MJOFBS ve- HFP TUSPQIJD BEKVTUNFOUlocity, normal QSPCMFN to the gradient B *OJUJBM IFJHIU mFME HJWFO CZ XJUI of the height  field, given by -  C &RVJMJCSJVN(4.80). (d) mOBM Potential IFJHIU vorticity, mFME  HJWFO CZ  given byBOE (4.75), and this D does &RVJ૙  D MJCSJVN HFPTUSPQIJDnot evolve. WFMPDJUZ The distance, OPSNBM푥 is UP૙  UIF HSBEJFOUnondimensionalized PG IFJHIU૙ਈ mFME ૪ਉ HJWFOby the CZ   E deformation 1PUFOUJBM WPSUJDJUZ radius 퐿 HJWFO푑 and CZ -  BOE UIJTthe velocity EPFT OPU by FWPMWF휂0(푔/푓0퐿푑). Changes to the initial state ʾF EJTUBODF JT OPOEJNFOTJPOBMJ[FE E ீ CZ UIF EFGPSNBUJPOoccur SBEJVT within BOE(퐿푑 UIF) of WF the initial discontinuity. MPDJUZ CZ ਙ  $IBOHFT UP UIF JOJ ਆ UJBM TUBUF PDDVS XJUIJO ৴ PG UIF JOJUJBM   ਆ EJTDPOUJOVJUZ૙ ਉਈ ৴ � ਆ ৴

8F TPMWF UIJT TFQBSBUFMZ GPS TheBOE initial conditionsBOE UIFO NBUDI (4.75) UIF admit TPMVUJPOT of a nice BOE analytic UIFJS ĕSTU solution, EFSJWBUJWFT for the BU BMTP JNQPTJOH UIF DPOEJUJPOflow will UIBU remain UIF uniform WFMPDJUZ in EFDBZT푦, and UP (4.77 [FSP) reduces BT to  ćF TPMVUJPO JT ਙ  ਙ 2 ∂ 휓 1 푓0휂0 ਙ F 2 − 2 휓 = sgnਙ(푥). Ҿ kú (4.78) ∂푥 퐿푑 퐻  F÷ਙ৴ਆ   We solve÷ ਉ૙ thisਈ separately  ÷ forਆ ਙ푥 > 0 and 푥 < 0 and then match the solutions ૪ʱ ਙ৴ ćF WFMPDJUZ ĕFME BTTPDJBUFE XJUIand their UIJT ਉ૙ JT first PCUBJOFEਈ derivatives  ÷ GSPN at C D 푥 = 0, also ਙ  BOE imposing  JT the condition that the velocity decays to zero as 푥 → ±∞. The solution is F  −푥/퐿푑 −(푔휂0/푓0)(1 − e ) 푥 > 0 휓 = {  ÷]ਙ]৴ਆ (4.79) ਉ૙+(푔휂 /푓 )(1 − e푥/퐿푑 ) 푥 < 0. ਖ   ਗ  ÷ 0 0  ćF WFMPDJUZ JT QFSQFOEJDVMBS UP UIF TMPQF PG UIF GSFFਈ৴ TVSGBDF ਆ BOE B KFU GPSNT BMPOH UIF JOJUJBM EJT DPOUJOVJUZ BT JMMVTUSBUFE JO 'JHThe  velocity field associated with this is obtained from (4.76b,c), and is ćF JNQPSUBOU QPJOU PG UIJT QSPCMFN JT UIBU UIF WBSJBUJPOT JO푔휂 UIF IFJHIU BOE ĕFME BSF OPU SB 푢 = 0, 푣 = − 0 e−|푥|/퐿푑 . (4.80) EJBUFE BXBZ UP JOĕOJUZ BT JO UIF OPOSPUBUJOH QSPCMFN 3BUIFS QPUFOUJBM푓 퐿 WPSUJDJUZ DPOTFSWBUJPO DPOTUSBJOT UIF JOĘVFODF PG UIF BEKVTUNFOU UP XJUIJO B EFGPSNBUJPO0 SBEJVT푑 XF TFF OPX XIZ UIJT OBNF JT BQQSPQSJBUF PG UIF JOJUJBMThe velocity EJTUVSCBODF is perpendicular ćJT QSPQFSUZ to the JT slope B HFOFSBM of thePOF free JO surface, HFPTUSPQIJD and a BE jet KVTUNFOU ‰ JU BMTP BSJTFT JG UIFforms JOJUJBM along DPOEJUJPO the initial DPOTJTUT discontinuity, PG B WFMPDJUZ as illustrated KVNQ in Fig. 4.8. " TOBQTIPU PG UIF UJNF FWPMVUJPOThe important PG ĘPX PCUBJOFE point of this CZ problem B OVNFSJDBM is that JOUFHSBUJPO the variationsPG UIF in the TIBMMPX height XBUFS FRVBUJPOT GPS CPUI SPUBUJOHand field BOE are OPOSPUBUJOH not radiated ĘPX away JT to JMMVTUSBUF infinity,E as JO in 'JH the  non-rotating ćF JOJUJBM problem. DPO EJUJPOT BSF B KVNQ JO UIF IFJHIURather, ĕFME potential BT JO 'JH vorticity  'SPOUT conservation QSPQBHBUF constrains BXBZ BU the B TQ influenceFFE of the JO CPUI DBTFT CVU JO UIF SPUBUJOHadjustment ĘPX UIFZ to within MFBWFa CFIJOE deformation B HFPTUSPQIJDBMMZ radius (we see CBMBODFE now why TUBUF this XJUI name B is appropriate) of the initial disturbance. This property is a general one OPO[FSP NFSJEJPOBM WFMPDJUZ ਉৰ   in geostrophic adjustment — it also arises if the initial conditionƉ consists of a velocity jump. The time evolution of the rotating flow, obtained by  &OFSHFUJDT PG "EKVTUNFOUa numerical integration, is illustrated in the left-hand panels of Fig. 4.7.

)PX NVDIĦ PG UIF JOJUJBM QPUFOUJBM FOFSHZ PG UIF ĘPX JT MPTU UP JOĕOJUZ CZ HSBWJUZ XBWF SBEJBUJPO BOE IPX NVDI JT DPOWFSUFE UP LJOFUJD FOFSHZ ćF MJOFBS FRVBUJPOT  MFBE UP

  ૬   ৰ੊ ਉ૙ ਉৰò ď ੊૙    ૬ਕ 79 4.5 A Variational Perspective on Adjustment

Fronts propagate away at a speed √푔퐻 = 1, just as in the non-rotating case, but in the rotating flow they leave behind a geostrophically balanced state with a non-zero meridional velocity.

4.5 ♦ A Variational Perspective on Adjustment

In the non-rotating problem, all of the initial potential energy is eventu- ally radiated away to infinity. In the rotating problem, the final state con- tains both potential and kinetic energy, mostly trapped within a deforma- tion radius of the initial disturbance, because potential vorticity conser- vation on parcels prevents all of the energy being dispersed. This suggests that it may be informative to think of the geostrophic adjustment prob- lem as a variational problem: we seek to minimize the energy consistent with the conservation of potential vorticity. We stay in the linear approx- imation in which, because the advection of potential vorticity is neglected, potential vorticity remains constant at each point. The energy of the flow, 퐸̂, is given by the sum of potential and kinetic energies, namely

퐸̂ = ∫(퐻풖2 + 푔휂2) d퐴, (4.81)

(where d퐴 = d푥 d푦) and the potential vorticity field is

휂 휂 푞 = 휁 − 푓 = (푣 − 푢 ) − 푓 , (4.82) 0 퐻 푥 푦 0 퐻 where the subscripts 푥, 푦 denote derivatives. The problem is then to ex- tremize the energy subject to potential vorticity conservation. This is a constrained problem in the calculus of variations, sometimes called an isoperimetric problem because of its origins in maximizing the area of a surface for a given perimeter. The mathematical problem is to extremize the integral

2 2 2 퐼 = ∫ {퐻(푢 + 푣 ) + 푔휂 + 휆(푥, 푦)[(푣푥 − 푢푦) − 푓0휂/퐻]} d퐴, (4.83) where 휆(푥, 푦) is a Lagrange multiplier, undetermined at this stage. It is a function of space: if it were a constant, the integral would merely extrem- ize energy subject to a given integral of potential vorticity, and rearrange- ments of potential vorticity (which here we wish to disallow) would leave the integral unaltered. As there are three independent variables there are three Euler– Lagrange equations that must be solved in order to minimize 퐼. These are

∂퐿 ∂ ∂퐿 ∂ ∂퐿 − − = 0, ∂휂 ∂푥 ∂휂푥 ∂푦 ∂휂푦 (4.84) ∂퐿 ∂ ∂퐿 ∂ ∂퐿 ∂퐿 ∂ ∂퐿 ∂ ∂퐿 − − = 0, − − = 0, ∂푢 ∂푥 ∂푢푥 ∂푦 ∂푢푦 ∂푣 ∂푥 ∂푣푥 ∂푦 ∂푣푦 Chapter 4. Shallow Water Equations 80

where 퐿 is the integrand on the right-hand side of (4.83). Substituting the expression for 퐿 into (4.84) gives, after a little algebra,

휆푓 ∂휆 ∂휆 2푔휂 − 0 = 0, 2퐻푢 + = 0, 2퐻푣 − = 0, (4.85) 퐻 ∂푦 ∂푥

and then eliminating 휆 gives the simple relationships

푔 ∂휂 푔 ∂휂 푢 = − , 푣 = . (4.86) 푓0 ∂푦 푓0 ∂푥 These are the equations of geostrophic balance! Thus, in the linear ap- proximation, geostrophic balance is the minimum energy state for a given field of potential vorticity.

Notes and References The shallow water equations are sometimes called the Saint-Venant equations af- ter Adhémar Jean Claude Barré de Saint-Venant (1797–1886) who wrote down a form of the equations in 1871. Pierre-Simon Laplace (1749–1825) previously wrote down a linear version of the equations on the sphere, now known as Laplace’s tidal equations, in 1776, and one would not be surprised if Euler knew about the equations before that.

Problems 4.1 Using the shallow water equations: (a) A cylindrical column of air at 30° latitude with radius 100 km expands horizontally (shrinking in depth accordingly) to twice its original ra- dius. If the air is initially at rest, what is the mean tangential velocity at the perimeter after the expansion? (b) An air column at 60°N with zero relative vorticity (휁 = 0) stretches from the surface to the , which we assume is a rigid lid, at 10 km. The air column moves zonally on to a plateau 2.5 km high. What is its relative vorticity? Suppose it then moves southwards to 30°N, staying on the plateau. What is its vorticity? 4.2 Show that the vertical velocity within a shallow water system is given by

푧 − 휂 Dℎ D휂 푤 = 푏 + 푏 . (P4.1) ℎ D푡 D푡 Interpret this result, showing that it gives sensible answers at the top and bottom of the fluid layer. 4.3 In the long-wave limit of Poincaré waves, fluid parcels behave as free- agents; that is, like free solid particles moving in a rotating frame unencum- bered by pressure forces. Why then, is their frequency given by 휔 = 푓 = 2훺 where 훺 is the rotation rate of the coordinate system, and not by 훺 itself? Do particles that are stationary or move in a straight line in the inertial frame of reference satisfy the dispersion relationship for Poincaré waves in this limit? Explain. [See also Durran (1993) and Egger (1999).] 81 Problems

4.4 Linearize the 푓-plane shallow water system about a state of rest. Suppose that there is an initial disturbance given in the general form

i(푘푥+푙푦) 휂 = ∬ ̃휂푘,푙 e d푘 d푙, (P4.2)

where 휂 is the deviation surface height and the Fourier coefficients ̃휂푘,푙 are given, and that the initial velocity is zero. (a) Obtain the geopotential field at the completion of geostrophic adjust- ment, and show that the deformation scale is a natural length scale in the problem. (b) Show that the change in total energy during the adjustment is always less than or equal to zero. Neglect any initial divergence. N.B. Because the problem is linear, the Fourier modes do not interact. 4.5 If energy conservation is one of the most basic physical laws, how can en- ergy be lost in geostrophic adjustment? 4.6 Geostrophic adjustment of a velocity jump. In which we consider the evolu- tion of the linearized 푓-plane shallow water equations in an infinite do- main. (a) Show that the linearized potential vorticity, 푞′, for the shallow water system is given by ℎ′ 푞′ = 휁′ − 푓 , (P4.3) 0 퐻 using standard notation. (b) If the flow is in geostrophic balance show that the relative vorticity is given by 휁′ = ∇2휓, ′ where 휓 = 푔ℎ /푓0. Hence show that the potential vorticity is then given by ′ 2 1 푞 = ∇ 휓 − 2 휓, 퐿푑

and write down an expression for 퐿푑. (c) Suppose that initially the fluid surface is flat, the zonal velocity is zero and the meridional velocity is given by

푣(푥) = 푣0 sgn(푥). (P4.4) Find the equilibrium height and velocity fields at 푡 = ∞. (d) What are the initial and final kinetic and potential energies? Partial solution:

The potential vorticity is 푞 = 휁 − 푓0휂/퐻, so that the initial state and final state are both given by

푞 = 2푣0 훿(푥). (P4.5) 2 2 −2 (Why?) The final state streamfunction is thus given by (∂ /∂푥 − 퐿푑 ) 휓 = 푞, with solution 휓 = 휓0 exp(푥/퐿푑) and 휓 = 휓0 exp(−푥/퐿푑) for 푥 < 0 and 푥 > 0, 2 where 휓0 = −퐿푑푣0 (why?), and 휂 = 푓0휓/푔. The energy is 퐸 = ∫(퐻푣 + 푔휂2)/2 d풙. The initial KE is infinite, the initial PE is zero, and the final state 2 has 푃퐸 = 퐾퐸 = 푔퐿푑휂0/4; that is, the energy is equipartitioned between kinetic and potential energies. 4.7 In the shallow water equations show that, if the flow is approximately geostrophically balanced, the energy at large scales is predominantly poten- tial energy and the energy at small scales is predominantly kinetic energy. Define precisely what ‘large scale’ and ‘small scale’ mean in this context. Chapter 4. Shallow Water Equations 82

4.8 In the shallow water geostrophic adjustment problem, show that at large scales the velocity essentially adjusts to the height field, and that at small scales the height field essentially adjusts to the velocity field. Your deriva- tion may be detailed and mathematical, but explain the result at the end in words and in physical terms. Chapter 5 Geostrophic Theory

eostrophic and hydrostatic balance are the two dominant balances in meteorology and and in this chapter G we exploit these balances to derive various simplified sets of equa- tion. The ‘problem’ with the full equations is that they are too complete, and they contain motions that we don’t always care about — sound waves and gravity waves for example. If we can eliminate these modes from the outset then our path toward understanding is not littered with obstacles. Our specific goal is to derive various sets of ‘geostrophic equations’, in particular the planetary-geostrophic and quasi-geostrophic equations, by making use of the fact that geostrophic and hydrostatic balance are closely satisfied. We do this first for the shallow water equations and then for the stratified, three-dimensional equations. We will use the Boussinesq equa- tions, but a treatment in pressure coordinates would be very similar. The bottom topography, 휂퐵, can be an unneeded complication in the deriva- tions below and readers may wish to simplify by setting 휂퐵 = 0.

5.1 Scaling the Shallow Water Equations

In order to simplify the equations of motion we first scale them — we choose the scales we wish to describe, and then determine the approx- imate sizes of the terms in the equations. We then eliminate the small terms and derive a set of equations that is simpler than the original set but that consistently describes motion of the chosen scale. With the odd exception, we will denote the scales of variables by capital letters; thus, if 퐿 is a typical length scale of the motion we wish to describe, and 푈 is a typical velocity scale, then (푥, 푦) ∼ 퐿 or (푥, 푦) = ீ(퐿), (5.1) (푢, 푣) ∼ 푈 or (푢, 푣) = ீ(푈), and similarly for the other variables in the equations. 83 Chapter 5. Geostrophic Theory 84

We then write the equations of motion in a nondimensional form by writing the variables as

(푥, 푦) = 퐿(̂푥,̂푦), (푢, 푣) = 푈(̂푢,̂푣), (5.2) where the hatted variables are nondimensional and, by supposition, are ீ(1). The various terms in the momentum equation then scale as: ∂풖 + 풖 ⋅ ∇풖 + 풇 × 풖 = −푔∇휂, (5.3a) ∂푡 푈 푈2 ஹ 푓푈 ∼ 푔 , (5.3b) 푇 퐿 퐿 where the ∇ operator acts in the 푥–푦 plane and ஹ is the of the variations in the surface displacement. We choose an ‘advective scale’ for time, meaning that 푇 = 퐿/푈 and 푡 = 푡퐿/푈̂ , and the time derivative then scales the same way as the advection. The ratio of the advective term to the rotational term in the momentum equation (5.3) is (푈2/퐿)/(푓푈) = 푈/푓퐿; this is the Rossby number that we previously encountered. The geostrophic theory of We are interested in flows for which the Rossby number is small, in this chapter applies when which case the Coriolis term is largely balanced by the pressure gradient. the Rossby number is small. From (5.3b), variations in 휂 scale according to On Earth the theory is gen- erally appropriate for large- 푓푈퐿 푓2퐿2 퐿2 ஹ = = Ro = Ro퐻 , (5.4) scale flow in the mid- and 푔 푔 퐿2 high latitude atmosphere 푑 and ocean. On other plan- where 퐿푑 = √푔퐻/푓 is the deformation radius and 퐻 is the mean depth ets the applicability of of the fluid. The ratio of variations in fluid height to the total fluid height geostrophic theory depends thus scales as on how rapidly the planet ஹ 퐿2 rotates and how big it is. ∼ Ro 2 . (5.5) Venus has a rotation rate 퐻 퐿푑 some 200 times slower than Now, the thickness of the fluid, ℎ, may be written as the sum of its Earth and the Rossby num- mean and a deviation, ℎ퐷 ber of the large-scale circu- lation is quite large. Jupiter ℎ = 퐻 + ℎ퐷 = 퐻 + (휂푇 − 휂퐵), (5.6) rotates much faster than Earth (a Jupiter day is about where, referring to Fig. 4.1, 휂퐵 is the height of the bottom topography and 10 hours), and the planet is 휂푇 is the height of the fluid above its mean value. Given the scalings above, also much bigger, and the the deviation height of the fluid may be written as Rossby number remains small even close to the equator. 퐿2 퐿2 휂푇 = Ro 2 퐻̂휂푇 and 휂 = 퐻 + 휂푇 = 퐻 (1 + Ro 2 ̂휂푇), (5.7) 퐿푑 퐿푑

where ̂휂푇 is the ீ(1) nondimensional value of the surface height deviation. We apply the same scalings to ℎ itself and, if ℎ퐷 = ℎ − 퐻 = 휂푇 − 휂퐵 is the deviation of the thickness from its mean value, then

2 퐿 ̂ ℎ = 퐻 + ℎ퐷 = 퐻 (1 + Ro 2 ℎ퐷), (5.8) 퐿푑 ̂ where ℎ퐷 is the nondimensional deviation thickness of the fluid layer. 85 5.2 Geostrophic Shallow Water Equations

Nondimensional momentum equation If we use (5.7) and (5.8) to scale height variations, (5.2) to scale lengths and velocities, and an advective scaling for time, then, and since ∇̂휂= ∇̂휂푇, the momentum equation (5.3) becomes

∂풖̂ Ro [ + (풖̂ ⋅ ∇)풖]̂ + 풇̂ × 풖̂ = −∇̂휂, (5.9) ∂푡̂

̂ ̂ ̂ ̂ where 풇 = 퐤푓 = 퐤푓/푓0, where 푓0 is a representative value of the Coriolis parameter. (If 푓 is a constant, then 푓̂ = 1, but it is informative to explic- itly write 푓̂ in the equations. Also, where the operator ∇ operates on a nondimensional variable, the differentials are taken with respect to the nondimensional variables ̂푥,̂푦.) All the variables in (5.9) will now be as- sumed to be of order unity, and the Rossby number multiplying the local time derivative and the advective terms indicates the smallness of those terms. By construction, the dominant balance in (5.9) is the geostrophic balance between the last two terms.

Nondimensional mass continuity (height) equation The (dimensional) mass continuity equation can be written as Dℎ 1 Dℎ ℎ + ℎ∇ ⋅ 풖 = 0 or 퐷 + (1 + 퐷 ) ∇ ⋅ 풖 = 0, (5.10) D푡 퐻 D푡 퐻 since Dℎ/D푡 = Dℎ퐷/D푡. Using (5.2) and (5.8) the above equation may be written

2 ̂ 2 퐿 Dℎ퐷 퐿 ̂ Ro ( ) + [1 + Ro ( ) ℎ퐷] ∇ ⋅ 풖̂ = 0. (5.11) 퐿푑 D푡̂ 퐿푑

Equations (5.9) and (5.11) are the nondimensional versions of the full shal- low water equations of motion. Since the Rossby number is small we might expect that some terms in this equation can be eliminated with lit- tle loss of accuracy, depending on the size of the second nondimensional 2 parameter, (퐿/퐿푑) , as we now explore.

5.2 Geostrophic Shallow Water Equations 5.2.1 Planetary-Geostrophic Equations We now derive simplified equation sets that are appropriate in particular parameter regimes, beginning with an equation set appropriate for the very largest scales. Specifically, we take

퐿 퐿 2 Ro ≪ 1, ≫ 1 such that Ro ( ) = ீ(1). (5.12) 퐿푑 퐿푑 The first inequality implies we are considering flows in geostrophic bal- ance. The second inequality means we are considering flows much larger Chapter 5. Geostrophic Theory 86

than the deformation radius. The ratio of the deformation radius to scale of motion of the fluid is called the Burger number; that is, Bu ≡ 퐿푑/퐿, so here we are considering small Burger-number flows. The smallness of the Rossby number means that we can neglect the material derivative in the momentum equation, (5.9), leaving geostrophic balance. Thus, in dimensional form, the momentum equation may be written, in vectorial or component forms, as

풇 × 풖 = −∇휂, or (5.13) ∂휂 ∂휂 푓푣 = 푔 , 푓푢 = −푔 . ∂푥 ∂푦

The planetary-geostrophic equations are appropriate Looking now at the mass continuity equation, (5.11), we see that there for geostrophically balanced are no small terms that can be eliminated. Thus, we have simply the full flow at very large scales. In mass conservation equation, the shallow water version, they consist of the full mass ∂ℎ conservation equation along + ∇ ⋅ (ℎ풖) = 0, (5.14) with geostrophic balance. ∂푡

where ℎ and 휂 are related by 휂 = 휂퐵 + ℎ, where 휂퐵 is the height of the bot- tom topography. Equations (5.13) and (5.14) form the planetary geostrophic shallow water equations. There is only one time derivative in the equations, so there can be no gravity waves. The system is evolved purely through the mass continuity equation, and the flow field is diagnosed from the height field.

Planetary-geostrophic potential vorticity In the (full) shallow water equations potential vorticity is conserved, meaning that D 휁 + 푓 ( ) = 0. (5.15) D푡 ℎ In the planetary-geostrophic equations we can use (5.13) and (5.14) to show that this conservation law becomes

D 푓 ( ) = 0, (5.16) D푡 ℎ

as might be expected since 휁 is smaller than 푓 by a factor of the Rossby number. An alternate derivation of the planetary-geostrophic equations is to go directly from (5.15) to (5.16), by virtue of the smallness of the Rossby number, and then simply use (5.16) instead of (5.14) as the evolu- tion equation. 87 5.2 Geostrophic Shallow Water Equations

5.2.2 Quasi-Geostrophic Equations The quasi-geostrophic equations are appropriate for scales of the same order as the deformation radius, and so for

퐿 퐿 2 Ro ≪ 1, = ீ(1) so that Ro ( ) ≪ 1. (5.17) 퐿푑 퐿푑 Since the Rossby number is small the momentum equations again reduce The quasi-geostrophic equa- to geostrophic balance, namely (5.13). In the mass continuity equation, tions are appropriate for we now eliminate all terms involving Rossby number to give geostrophically balanced flow at so-called synoptic scales, or weather scales. This scale ∇ ⋅ 풖 = 0. (5.18) is mainly determined by the Rossby radius of deformation Neither geostrophic balance nor (5.18) are prognostic equations, and it which is about 1000 km in the appears we have derived an uninteresting, static, set of equations. In atmosphere and 100 km (and fact we haven’t gone far enough, since nothing in our derivation says less in high latitudes) in the that these quantities do not evolve. To see this, let us suppose that the ocean. Coriolis parameter is nearly constant, which is physically consistent with the idea that scales of motion are comparable to the deformation scale. Geostrophic balance with a constant Coriolis parameter gives

∂휂 ∂휂 푓 푢 = −푔 , −푓 푣 = −푔 , giving ∇ ⋅ 풖 = 0. (5.19) 0 ∂푦 0 ∂푧

That it to say, the geostrophic flow is divergence-free and we therefore should not suppose that ∇ ⋅ 풖 = 0 is the dominant term in the height equation. However, with a little more care we can in fact derive a set of equations that evolves under these conditions, and that furthermore is extraordinar- ily useful, for it describes the flow on the scales of motion corresponding to weather. We make three explicit assumptions: (i) The Rossby number is small and the flow is in near geostrophic balance. (ii) The scales of motion are similar to the deformation scales, so that 2 퐿 ∼ 퐿푑 and Ro(퐿/퐿푑) ≪ 1.

(iii) Variations of the Coriolis parameter are small, so that 푓 = 푓0 + 훽푦 where 훽푦 ≪ 푓0.

The velocity is then equal to a geostrophic component, 풖푔 plus an ageostrophic component, 풖푎 where |풖푔| ≫ |풖푎| and the geostrophic ve- locity satisfies

푓0 × 풖푔 = −푔∇휂, (5.20) which, because of the use of a constant Coriolis parameter (assumption (iii)), implies ∇ ⋅ 풖푔 = 0. We proceed from the shallow water vorticity equation which, as in (4.32), is ∂휁 + (풖 ⋅ ∇)(휁 + 푓) = − (휁 + 푓)∇ ⋅ 풖. (5.21) ∂푡 Chapter 5. Geostrophic Theory 88

A consequence of (5.20) is that the right-hand side contains only the ageostrophic velocity, which is small, and since 휁 is smaller than 푓 by a factor of the Rossby number we can ignore 휁∇ ⋅ 풖 and take 푓 to be equal to 푓0. On the left-hand side the velocities are well-approximated by using the geostrophic flow, so that we have

∂휁푔 + (풖 ⋅ ∇)(휁 + 푓) = −푓 ∇ ⋅ 풖 , (5.22) ∂푡 푔 푔 0 푎 where on the left-hand side 푓 can be replaced by 훽푦. We now use the mass continuity equation to obtain an expression for the divergence. From (5.10) the mass continuity equation is

Dℎ 퐷 + (퐻 + ℎ )∇ ⋅ 풖 = 0, (5.23) D푡 퐷

2 −1 and since 퐻 ≫ ℎ퐷 (using (5.11), 퐻 is bigger by a factor (퐿푑/퐿) 푅표 ), the equation becomes D (휂 − 휂 ) + 퐻∇ ⋅ 풖 = 0. (5.24) D푡 퐵 푎 Combining (5.22) and (5.23) gives

D 푓 (휂 − 휂 ) (휁 + 푓 − 0 퐵 ) = 0. (5.25) D푡 푔 퐻

It appears that we have two dynamical variables here, 휁푔 and 휂. How- ever, they are related through geostrophic balance, and the fact that the geostrophic flow is non-divergent. Thus, we may define a streamfunction 휓 such that 푢푔 = −∂휓/∂푦, 푣푔 = ∂휓/∂푥, whence ∂푢/∂푥 + ∂푣/∂푦 = 0. The vorticity and height fields are related to the streamfunction by Both the planetary- ∂푣 ∂푢 푓0휓 geostrophic and the quasi- 휁 = − = ∇2휓, and 휂 = , (5.26a,b) 푔 ∂푥 ∂푦 푔 geostrophic equations can be written in the form of where the second relation comes from geostrophic balance. Equation an evolution equation for potential vorticity, along (5.25) may then be written as with an ‘inversion’ to deter- mine the velocity fields and D푞 휓 푓 휂 = 0, 푞 = ∇2휓 + 훽푦 − + 0 퐵 , (5.27) the height field. The differ- D푡 퐿2 퐻 ence in the two equation 푑 sets lies in the approxima- tions made in the inversion. where 퐿푑 = √푔퐻/푓0 and the variable 푞 is the quasi-geostrophic potential In the planetary-geostrophic vorticity. system relative vorticity is ig- nored and 푄 = 푓/ℎ. In quasi- geostrophy the variations 5.2.3 Quasi-Geostrophic Potential Vorticity in the height field are small Connection to shallow water potential vorticity and 푞 = 훽푦 + 휁 − 푓0휂푇/퐻. Both equation sets as- The quantity 푞 given by (5.27) is an approximation (except for dynamically sume a low Rossby number. unimportant constant additive and multiplicative factors) to the shallow 89 5.3 Scaling in the Continuously-Stratified System water potential vorticity. To see the truth of this statement let us begin with the expression for the shallow water potential vorticity,

푓 + 휁 푄 = . (5.28) ℎ

For simplicity we set bottom topography to zero and then ℎ = 퐻(1+휂푇/퐻) and assume that 휂푇/퐻 is small to obtain

푓 + 휁 1 휂푇 1 휂푇 푄 = ≈ (푓 + 휁) (1 − ) ≈ (푓0 + 훽푦 + 휁 − 푓0 ). 퐻(1 + 휂푇/퐻) 퐻 퐻 퐻 퐻 (5.29) Because 푓0/퐻 is a constant it has no effect in the evolution equation, and the quantity given by 휂 푞 = 훽푦 + 휁 − 푓 푇 (5.30) 0 퐻 is materially conserved. Using geostrophic balance we have 휁 = ∇2휓 and 휂푇 = 푓0휓/푔 so that (5.30) is identical (except for 휂퐵) to (5.27). The approximations needed to go from (5.28) to (5.30) are the same as those used in our earlier, more long-winded, derivation of the quasi- geostrophic equations. That is, we assumed that 푓 itself is nearly con- stant, and that 푓0 is much larger than 휁, equivalent to a low Rossby num- ber assumption. It was also necessary to assume that 퐻 ≫ 휂푇 to enable the expansion of the height field and this approximation is equivalent to requiring that the scale of motion not be significantly larger than the de- formation scale. The derivation is completed by noting that the advection of the potential vorticity should be by the geostrophic velocity alone, and we recover (5.27).

5.3 Scaling in the Continuously-Stratified System

We now apply the same scaling ideas, mutatis mutandis, to the stratified . We use the hydrostatic Boussinesq equations, which we write as D풖 + 풇 × 풖 = −∇ 휙, (5.31a) D푡 푧 ∂휙 = 푏, (5.31b) ∂푧 D푏 = 0, (5.31c) D푡 ∇ ⋅ 풗 = 0. (5.31d)

We will consider only the case of a flat bottom, but topography is a rela- tively straightforward extension. Anticipating that the average stratifica- tion may not scale in the same way as the deviation from it, let us sepa- rate out the contribution of the advection of a reference stratification in (5.31c) by writing 푏 = 푏(푧)̃ + 푏′(푥, 푦, 푧, 푡). (5.32) Chapter 5. Geostrophic Theory 90

The thermodynamic equation then becomes

D푏′ + 푁2푤 = 0, (5.33) D푡 where 푁2 = ∂푏/∂푧̃ and the advective derivative is still three-dimensional. We then let 휙 = 휙(푧)̃ + 휙′, where 휙̃ is hydrostatically balanced by 푏̃, and the hydrostatic equation becomes

∂휙′ = 푏′. (5.34) ∂푧 Equations (5.33) and (5.34) replace (5.31c) and (5.31b), and 휙′ is used in (5.31a).

5.3.1 Scaling the Equations We scale the basic variables by supposing that 퐿 (푥, 푦) ∼ 퐿, (푢, 푣) ∼ 푈, 푡 ∼ , 푧 ∼ 퐻, 푓 ∼ 푓 , 푁 ∼ 푁 , (5.35) 푈 0 0

where the scaling variables (capitalized, except for 푓0) are chosen to be such that the nondimensional variables have magnitudes of the order of unity, and the constant 푁0 is a representative value of the stratification. We presume that the scales chosen are such that the Rossby number is small; that is Ro = 푈/(푓0퐿) ≪ 1. In the momentum equation the pressure term then balances the Coriolis force,

|풇 × 풖| ∼ |∇휙′|, (5.36)

and so the pressure scales as

′ 휙 ∼ 훷 = 푓표푈퐿. (5.37) Using the hydrostatic relation, (5.37) implies that the scales as 푓 푈퐿 푏′ ∼ 퐵 = 0 , (5.38) 퐻 and from this we obtain (∂푏′/∂푧) 퐿2 2 ∼ Ro 2 , (5.39) 푁 퐿푑 where 푁0퐻 퐿푑 = (5.40) 푓0 is the deformation radius in the continuously-stratified fluid, analogous to the quantity √푔퐻/푓0 in the shallow water system, and we use the same symbol for both. In the continuously-stratified system, if the scale of mo- tion is the same as or smaller than the deformation radius, and the Rossby number is small, then the variations in stratification are small. The choice of 91 5.3 Scaling in the Continuously-Stratified System scale is the key difference between the planetary-geostrophic and quasi- geostrophic equations. Finally, at least for now, we nondimensionalize the vertical velocity by using the mass conservation equation, ∂푤 ∂푢 ∂푣 = − ( + ), (5.41) ∂푧 ∂푥 ∂푦 with the scaling 푈퐻 푤 ∼ 푊 = . (5.42) 퐿 This scaling will not necessarily be correct if the flow is geostrophically balanced. In this case we can then estimate 푤 by cross-differentiating geostrophic balance (with ̃휌 constant for simplicity) to obtain the linear geostrophic vorticity equation and corresponding scaling: ∂푤 훽푈퐻 훽푣 ≈ 푓 , 푤 ∼ 푊 = . (5.43a,b) ∂푧 푓0

If variations in the Coriolis parameter are large and 훽 ∼ 푓0/퐿, then (5.43b) is the same as (5.42), but if 푓 is nearly constant then 푊 ≪ 푈퐻/퐿. Given the scalings above (using (5.42) for 푤) we nondimensionalize by setting 푈 (̂푥,̂푦)= 퐿−1(푥, 푦), 푧̂ = 퐻−1푧, (̂푢,̂푣)= 푈−1(푢, 푣), 푡̂ = 푡, 퐿 ′ (5.44) 퐿 ̂ −1 ̂ 휙 ̂ 퐻 ′ ̂푤= 푤, 푓 = 푓0 푓, 휙 = , 푏 = 푏 , 푈퐻 푓0푈퐿 푓0푈퐿 where the hatted variables are nondimensional. The horizontal momen- tum and hydrostatic equations then become D풖̂ Ro + 풇̂ × 풖̂ = −∇휙,̂ (5.45) D푡̂ and ∂휙̂ = 푏.̂ (5.46) ∂푧̂ The nondimensional mass conservation equation is simply ∂ ̂푢 ∂̂푣 ∂ ̂푤 ∇ ⋅ 풗̂ = ( + + ) = 0, (5.47) ∂̂푥 ∂ ̂푦 ∂푧̂ and the nondimensional thermodynamic equation is

푓 푈퐿 푈 D푏̂ 퐻푈 0 + 푁̂2푁2 ̂푤= 0, (5.48) 퐻 퐿 D푡̂ 0 퐿 or, re-arranging, D푏̂ 퐿 2 Ro + ( 푑 ) 푁̂2 ̂푤= 0. (5.49) D푡̂ 퐿 The nondimensional equations are summarized in the box on the follow- ing page. Chapter 5. Geostrophic Theory 92

Nondimensional Boussinesq Primitive Equations

The nondimensional, hydrostatic, Boussinesq equations in a ro- tating frame of reference are: D풖̂ Horizontal momentum: Ro + 풇̂ × 풖̂ = −∇휙,̂ (PE.1) D푡̂ ∂휙̂ Hydrostatic: = 푏,̂ (PE.2) ∂푧̂ ∂ ̂푢 ∂̂푣 ∂ ̂푤 Mass continuity: + + = 0, (PE.3) ∂̂푥 ∂ ̂푦 ∂푧̂ D푏̂ 퐿 2 Thermodynamic: Ro + ( 푑 ) 푁̂2 ̂푤= 0. (PE.4) D푡̂ 퐿

5.4 Planetary-Geostrophic Equations for Stratified Fluids

The planetary-geostrophic equations are appropriate for geostrophic flow at large horizontal scales. The specific assumptions we apply to the primi- tive equations are that: (i) Ro ≪ 1, 2 2 (ii) (퐿푑/퐿) ≪ 1. More specifically, Ro(퐿/퐿푑) = ீ(1) . We also allow 푓 to have its full variation with latitude, although we still use Cartesian coordinates. If we look at the equations in the shaded box above we see that, if we are to retain only the dominant terms, the momen- tum equation should simply be replaced by geostrophic balance, whereas hydrostatic and mass continuity equations are unaltered. In the thermodynamic equation both terms are of order Rossby num- ber, and therefore we retain both. This circumstance arises because, by assumption, we are dealing with large scales and on these scales the per- turbation buoyancy varies as much as the mean buoyancy itself. The buoy- ancy equation reverts to the evolution of total buoyancy, namely D푏 = 0, (5.50) D푡 where the material derivative is fully three-dimensional. This is the only evolution equation in the system. Thus, to summarize, the planetary- geostrophic equations of motion are, in dimensional form,

D푏 = 0, D푡 (5.51) ∂휙 풇 × 풖 = −∇휙, = 푏′, ∇ ⋅ 풗 = 0. ∂푧 93 5.5 Quasi-Geostrophic Equations for Stratified Fluids

Potential vorticity Manipulation of (5.51) reveals that we can equivalently write the equa- tions as an evolution equation for potential vorticity. Thus, the evolution equations may be written as

D푄 ∂푏 = 0, 푄 = 푓 . (5.52) D푡 ∂푧

The inversion — i.e., the diagnosis of velocity, pressure and buoyancy — is carried out using the hydrostatic, geostrophic and mass conservation equations, and in a real situation the right-hand sides of the buoyancy and potential vorticity equations would have terms representing heating.

5.4.1 Applicability to the Ocean and Atmosphere In the atmosphere a typical deformation radius 푁퐻/푓 is about 1000 km. The constraint that the scale of motion be much larger than the deforma- tion radius is thus quite hard to satisfy, since one quickly runs out of room on a planet whose equator-to-pole distance is 10 000 km. Thus, only the largest scales of motion can satisfy the planetary-geostrophic scaling in the atmosphere and we should then also properly write the equations in spherical coordinates. In the ocean the deformation radius is about 100 km, so there is lots of room for the planetary-geostrophic equations to hold, and indeed much of the theory of the large-scale structure of the ocean involves the planetary-geostrophic equations.

5.5 Quasi-Geostrophic Equations for Stratified Fluids

Let us now consider the appropriate equations for geostrophic flow at scales of order the deformation radius.

5.5.1 Scaling and Assumptions The nondimensionalization and scaling are as before and so the nondi- mensional equations are those in the shaded box on the facing page. The Coriolis parameter is given by ̂ 풇 = (푓0 + 훽푦) 퐤. (5.53) The variation of the Coriolis parameter is now assumed to be small (this is a key difference between the quasi-geostrophic system and the planetary- geostrophic system), and in particular we assume that 훽푦 is approximately the size of the relative vorticity, and so is much smaller than 푓0. The main assumptions needed to derive the QG system are then: (i) The Rossby number is small, Ro ≪ 1. (ii) Length scales are of the same order as the deformation radius, 퐿 ∼ 퐿푑 or 퐿/퐿푑 = ீ(1). (iii) Variations in Coriolis parameter are small, and specifically |훽푦| ∼ Ro푓0. Chapter 5. Geostrophic Theory 94

Given these assumptions, we can write the horizontal velocity as the sum of a geostrophic component and an ageostrophic one: ̂ 풖 = 풖푔 + 풖푎, where 푓0 퐤 × 풖푔 = −∇휙 and |풖푔| ≫ |풖푎|. (5.54) Since the Coriolis parameter is constant in our definition of the geostrophic velocity the geostrophic divergence is zero; that is

∂푢푔 ∂푣푔 + = 0. (5.55) ∂푥 ∂푦

The vertical velocity is thus given by the divergence of the ageostrophic velocity, ∂푤 ∂푢 ∂푣 = − 푎 + 푎 . (5.56) ∂푧 ∂푥 ∂푦 Since the ageostrophic velocity is small, the actual vertical velocity is smaller than the scaling suggested by the mass conservation equation in its original form. That is, 푈퐻 푊 ≪ . (5.57) 퐿 The quasi-geostrophic equa- tions are probably the most used set of equations in theo- 5.5.2 Derivation of Stratified QG Equations retical meteorology, certainly for midlatitude dynamics. We begin by cross-differentiating the horizontal momentum equation to give, after a few lines of algebra, the vorticity equation:

D ∂푢 ∂푣 ∂푢 ∂푤 ∂푣 ∂푤 (휁 + 푓) = −(휁 + 푓) ( + ) + ( − ). (5.58) D푡 ∂푥 ∂푦 ∂푧 ∂푦 ∂푧 ∂푥 We now apply the quasi-geostrophic assumption, namely: (i) The geostrophic velocity and vorticity are much larger than their ageostrophic counterparts, and therefore we use geostrophic values for the terms on the left-hand side. (ii) On the right-hand side we keep the horizontal divergence (which is small) only where it is multiplied by the big term 푓. Furthermore, because 푓 is nearly constant we replace it with 푓0. (iii) The second term in large parentheses on the right-hand side is smaller than the advection terms on the left-hand side by the ra- tio [푈푊/(퐻퐿)]/[푈2/퐿2] = [푊/퐻]/[푈/퐿] ≪ 1, because 푤 is small, as noted above. We thus neglect it. Given the above, (5.58) becomes

D푔 ∂푢 ∂푣 ∂푤 (휁 + 푓) = −푓 ( + ) = 푓 , (5.59) D푡 푔 0 ∂푥 ∂푦 0 ∂푧

where the second equality uses mass continuity and D푔/D푡 = ∂/∂푡 +풖푔 ⋅∇ — note that only the (horizontal) geostrophic velocity does any advecting, because 푤 is so small. 95 5.5 Quasi-Geostrophic Equations for Stratified Fluids

Now consider the three-dimensional thermodynamic equation. Since 푤 is small it only advects the basic state, and the perturbation buoyancy is advected only by the geostrophic velocity. Thus, (5.33) becomes,

′ D푔푏 + 푤푁2 = 0. (5.60) D푡 We now eliminate 푤 between (5.59) and (5.60), which, after a little algebra, yields

′ D푔푞 ∂ 푓 푏 = 0, 푞 = 휁 + 푓 + ( 0 ). (5.61) D푡 푔 ∂푧 푁2

Hydrostatic and geostrophic balance enable us to write the geostrophic velocity, vorticity, and buoyancy in terms of streamfunction 휓 [= 푝/(푓0휌0)]: ̂ 2 ′ 풖푔 = 퐤 × ∇휓, 휁푔 = ∇ 휓, 푏 = 푓0∂휓/∂푧. (5.62) Thus, we have, now omitting the subscript g,

D푞 ∂ 1 ∂휓 = 0, 푞 = ∇2휓 + 푓 + 푓2 ( ). (5.63a,b) D푡 0 ∂푧 푁2 ∂푧

Only the variable part of 푓 (i.e., 훽푦) is relevant in the second term on the right-hand side of the expression for 푞. The material derivative may be expressed as D푞 ∂푞 ∂푞 = + 풖 ⋅ ∇푞 = + 퐽(휓, 푞). (5.64) D푡 ∂푡 ∂푡 where 퐽(휓, 푞) = ∂휓/∂푥 ∂푞/∂푦 − ∂휓/∂푦 ∂푞/∂푥. The quantity 푞 is known as the quasi-geostrophic potential vorticity, and it is conserved when advected by the horizontal geostrophic flow. All the other dynamical variables may be obtained from potential vorticity as fol- lows: (i) Streamfunction, using (5.63b). (ii) Velocity: 풖 = 퐤̂ × ∇휓 [≡ ∇⊥휓 = −∇ × (퐤휓)]̂ . (iii) Relative vorticity: 휁 = ∇2휓 .

(iv) Perturbation pressure: 휙 = 푓0휓. ′ (v) Perturbation buoyancy: 푏 = 푓0∂휓/∂푧.

By inspection of (5.63b) we see that a length scale 퐿푑 = 푁퐻/푓0, emerges naturally from the quasi-geostrophic dynamics. It is the scale at which buoyancy and relative vorticity effects contribute equally to the potential vorticity, and is called the deformation radius; it is analogous to the quan- tity √푔퐻/푓0 arising in shallow water theory. In the upper ocean, with −2 −1 3 −4 −1 푁 ≈ 10 s , 퐻 ≈ 10 m and 푓0 ≈ 10 s , then 퐿푑 ≈ 100 km. At high latitudes the ocean is much less stratified and 푓 is somewhat larger, and the deformation radius may be as little as 20 km. In the atmosphere, with Chapter 5. Geostrophic Theory 96

−2 −1 4 푁 ≈ 10 s , 퐻 ≈ 10 m, then 퐿푑 ≈ 1000 km. It is this order of magni- tude difference in the deformation scales that accounts for a great deal of the quantitative difference in the dynamics of the ocean and the atmo- sphere. If we take the limit 퐿푑 → ∞ then the stratified quasi-geostrophic equations reduce to D푞 = 0, 푞 = ∇2휓 + 푓. (5.65) D푡 This is the two-dimensional vorticity equation; the high stratification of this limit has suppressed all vertical motion, and variations in the flow become confined to the horizontal plane. Finally, if we allow density to vary in the vertical and carry through the quasi-geostrophic derivation we find that the quasi-geostrophic potential vorticity is given by 푓2 ∂ ̃휌 ∂휓 푞 = ∇2휓 + 푓 + 0 ( ), (5.66) ̃휌 ∂푧 푁2 ∂푧 where ̃휌 is a reference profile of density and a function of 푧 only, typically decreasing approximately exponentially with height.

5.5.3 Upper and Lower Boundary Conditions and Buoyancy Ad- vection The solution of the elliptic equation in (5.63b) requires vertical bound- ary conditions on 휓 at the ground and at some upper boundary (e.g., the tropopause), and these are given by use of the thermodynamic equation. For a flat, slippery, rigid surface the vertical velocity is zero so that the thermodynamic equation may be written as D푏′ ∂휓 = 0, 푏′ = 푓 , 푧 = 0, 퐻. (5.67) D푡 0 ∂푧 This equation provides the values of ∂휓/∂푧 at the boundary that are then used when solving (5.63b). A heating term could be added as needed, and if there is no upper boundary (as in the real atmosphere) we would apply a condition that the motion decays to zero with height, but we will not deal with that condition further. If the bottom boundary is not flat then the vertical velocity is non-zero and the thermodynamic equation becomes, at the bottom boundary, D ∂휓 ( ) + 푤푁2 = 0, 푤 = 풖 ⋅ ∇휂 , (5.68) D푡 ∂푧 퐵

where 휂퐵 is the height of the topography. In practice, the bottom bound- ary condition is often incorporated into the definition of potential vortic- ity, as we shall see in the two-level model.

5.6 The Two-Level Quasi-Geostrophic Equations

The continuously-stratified quasi-geostrophic equations have, as their name suggests, a continuous variation in the vertical. It turns out to be 97 5.6 The Two-Level Quasi-Geostrophic Equations

Fig. 5.1: A two-level quasi- geostrophic fluid system with a flat bottom and a rigid lid, with 푤 = 0 at both. Here the levels are allowed to be of unequal thicknesses, but in

the text we take 퐻1 = 퐻2.

extraordinarily useful to constrain this variation to that of two levels, in which the velocity is defined at just two vertical levels, and the tempera- ture at just one level in the middle, but keeping the full horizontal vari- ation. The resulting equations are not only algebraically much simpler than the continuous equations but they also capture many of their impor- tant features. There are two ways to derive the equations. One is by con- sidering, ab initio, the motion of two immiscible shallow layers of fluid, one on top of the other, giving the ‘two-layer’ quasi-geostrophic equa- tions. The second is by finite-differencing the continuous equations, and this is the procedure we follow here. Both methods give identical answers and we use ‘two-layer’ and ‘two-level’ interchangeably throught the book.

5.6.1 Deriving the Equations The quasi-geostrophic potential vorticity may be written as ∂ 푓2 ∂휓 푞 = ∇2휓 + 훽푦 + ( 0 ). (5.69) ∂푧 푁2 ∂푧 We now apply this equation at the two levels, 1 and 2, in Fig. 5.1, taking 푁 to be constant and 퐻1 = 퐻2 = 퐻/2 for simplicity (these restrictions may be easily relaxed). Using simple expressions for vertical finite differences gives 2 2 푓0 1 ∂휓 휓1 − 휓2 푞1 = ∇ 휓1 + 훽푦 + 2 (( ) − ). (5.70) 푁 퐻/2 ∂푧 top 퐻/2

In this equation (∂휓/∂푧)top is the value of ∂휓/∂푧 at the top of the domain and 2(휓1 − 휓2)/퐻 is the value of ∂휓/∂푧 in the middle of the domain. The value at the top is given by the thermodynamic equation, (5.67), and if there is no heating we set D/D푡 (∂휓/∂푧)top = 0. The potential vorticity equation for the upper level is then given by D푞 푘2 1 = 0, 푞 = ∇2휓 + 훽푦 + 푑 (휓 − 휓 ), (5.71a,b) D푡 1 1 2 2 1 2 2 2 2 where 푘푑 = 8푓0 /푁 퐻 . In proceeding this way we have built the bound- ary conditions into the definition of potential vorticity. A thermody- Chapter 5. Geostrophic Theory 98

Quasi-Geostrophic Equations

Continuously-stratified The adiabatic quasi-geostrophic potential vorticity equation for a Boussinesq fluid on the 훽-plane, is

D푞 ∂ 푓2 ∂휓 = 0, 푞 = ∇2휓 + 푓훽푦 + ( 0 ), (QG.1) D푡 ∂푧 푁2 ∂푧 where 휓 is the streamfunction. The horizontal velocities are given by (푢, 푣) = (−∂휓/∂푦, ∂휓/∂푥). The boundary conditions at the top and bottom are given by the buoyancy equation,

D푏 ∂휓 = 0, 푏 = 푓 . (QG.2) D푡 0 ∂푧 Two-level The two-level or two-layer quasi-geostrophic equations are

D푞 푘2 푖 = 0, 푞 = ∇2휓 + 훽푦 + 푑 (휓 − 휓 ), (QG.3) D푡 푖 푖 2 푗 푖 where 푖 = 1, 2, denoting the top and bottom levels respectively, and 푗 = 3 − 푖.

namic source and/or frictional terms may readily be added to the right- hand of (5.71a) as needed. 2 Defining 푘푑 so that 푘푑/2 We apply an exactly analogous procedure to the lower level, and so ob- rather than just 2 is used 푘푑 tain an expressions for the evolution of the lower layer potential vorticity, in the two-level quasi- namely, geostrophic potential vor- 2 ticity is for later algebraic con- D푞2 2 푘푑 = 0, 푞2 = ∇ 휓2 + 훽푦 + (휓1 − 휓2), (5.72) venience rather than being D푡 2 of fundamental importance. Once again the boundary conditions on buoyancy — the finite difference analogue of the buoyancy equations at the top and bottom — are built in to the definition of potential vorticity. If topography is present then it has an effect through the buoyancy equation at the bottom and it is incorporated into the definition of the lower level potential vorticity, which becomes

2 2 푘푑 푓0휂푏 푞2 = ∇ 휓2 + 훽푦 + (휓1 − 휓2) + . (5.73) 2 퐻2 This has a similar form to the shallow water potential vorticity with topog- raphy, as in (5.27). The two-level equations have proven to be enormously useful in the development of dynamical oceanography and meteorology and we will come back to them in later chapters. The quasi-geostrophic equations are summarized in the box above. 99 5.7 Frictional Geostrophic Balance and Ekman Layers

5.7 Frictional Geostrophic Balance and Ekman Layers 5.7.1 Preliminaries Within a few hundred meters of the ground in the atmosphere, and in the upper and lower tens of metres in the ocean, frictional effects are impor- tant and geostrophic balance by itself is not the leading order balance in the momentum equations. Rather, a frictional term is important and the momentum equation becomes, for a Boussinesq fluid,

풇 × 풖 = −∇휙 + 푭, (5.74) where 푭 is a frictional term and ∇휙 is the horizontal gradient of pressure, at constant 푧. The ultimately comes from molecular and Fig. 5.2: An idealized bound- ary layer in which fields vary 푭 = 휈∇2풖 is of the form . However, in geophysical settings the vertical rapidly in order to satisfy the derivative dominates and it is common to write (5.74) in the form boundary conditions, here that U = 0. The scale is a 1 ∂흉 훿 풇 × 풖 = −∇푧휙 + , (5.75) measure of the thickness of 휌0 ∂푧 the and H is a typical scale of the variations where 흉 is the stress, and since density is assumed constant we shall ab- away from the boundary. sorb it into the definition of 흉. (The quantity 흉/휌0 is the ‘kinematic stress’, but we shall just refer to it as the stress and also denote it 흉. Also, in some other fluid dynamical contexts the stress is a tensor but here 흉 is a vector.) Commonly, the regions where the stress term is important are boundary layers (see Fig. 5.2) and they exist because the fluid takes on values at the boundary that differ from the values in the fluid interior. Boundary layers are ubiquitous in fluids in many circumstances, and if the Coriolis term is important then the region is called an . As noted, the stress (1861–1930), ultimately arises from molecular forces and 흉 = 휈∂풖/∂푧 where 휈 is the the polar explorer and states- coefficient of viscosity, again just including vertical derivatives. However, man, apparently wanted to on the scale of the Ekman layer molecular effects are greatly amplified by understand the motion of pack ice and of his ship, the the effects of small scale (as discussed more in Chapter 10) and Fram, embedded in the ice we commonly replace 휈 by a much larger ‘ viscosity’, 퐴. and which seemed to move In the atmosphere we imagine that the flow is nearly geostrophic with neither the wind nor the above an Ekman layer, with frictional effects coming from the need to surface current. He posed bring the speed of flow down from its high geostrophic value to zero at the problem to V. W. Ekman the ground. In the ocean, on the other hand, the stress largely arises from (1874–1954) who then calcu- the wind in the atmosphere blowing over the ocean surface. The stress lated the direction of the flow then decays to zero with depth, and in the deeper ocean the flow again is in the upper ocean and pub- geostrophic. The stress itself is continuous across the ocean–atmosphere lished a paper on the matter interface, usually with a maximum value at the interface, decaying with in 1905. height in the atmosphere and with depth in the ocean. The equations of motion of the Ekman layer are completed by the mass continuity equation, ∇⋅풗 = 0, and the hydrostatic equation, ∂휙/∂푧 = 푏. In order to treat the simplest case we take buoyancy to be constant, and in that case, and without any additional loss of generality, we take 푏 = 0. The horizontal gradient of pressure is independent of height, which implies that, even though velocity may vary rapidly in the boundary layer the pressure gradient force does not. Chapter 5. Geostrophic Theory 100

5.7.2 Properties of Ekman Layers We shall now determine three important properties of Ekman layers: (i) The transport in an Ekman layer. (ii) The depth of the Ekman layer. (iii) The vertical velocity at the edge of the Ekman layer. The importance of these will become clear as we go on.

Transport We may write the Ekman layer equation (5.74) as ∂흉 ∂흉 풇 × (풖 − 풖 ) = or equivalently 푓풖 = −퐤̂ × , (5.76a,b) 푔 ∂푧 푎 ∂푧

where 풖푔 is the such that 풇 × 풖푔 = −∇휙, 풖푎 = 풖 − 풖푔 is the ageostrophic wind and 퐤̂ is the unit vector in the vertical (so that 풇 = 푓퐤̂ ). Consider an ocean with a stress that diminishes with depth from its surface values, 흉0. If we integrate (5.76) over the depth of the Ekman layer (i.e., down from the surface until the stress vanishes) we obtain −f × Ua τ0 ̂ 푦 푥 ∫ 푓풖푎 d푧 = −퐤 × 흉0, or ∫ 푓푢푎 d푧 = 휏0 , ∫ 푓푣푎 d푧 = −휏0 , (5.77)

Ua where the superscripts denote the 푥 and 푦 components of the stress. That is to say, the integrated ageostrophic transport in the Ekman layer, 푼 = Fig. 5.3: Vertically inte- 푎 ∫ 풖푎 d푧, is at right angles to the stress at the surface, as in Fig. 5.3. This re- grated transport in a North- sult is independent of the detailed form of the stress, and it gives a partial ern Hemisphere ocean answer to Nansen’s question — if the surface stress is in the direction of Ekman layer. A stress 흉0 at the surface induces an the wind, the water beneath it flows at an angle to it, as a consequence of the Coriolis force. More generally, the is the most ageostrophic flow, 푼푎. In or- der that the Coriolis force immediate response of the ocean to the atmosphere and is thus of key im- on the ageostrophic flow portance in the ocean . In the atmosphere we live in the Ekman can balance the , and Ekman layer effects are partly responsible for determining the stress the ageostrophic flow surface wind. must be at right angles to the stress, as illustrated. Depth Although the transport in the Ekman layer is independent of the detailed form of the stress, the depth — namely the extent of the frictional influ- ence of the surface — is not and to calculate it we use the Ekman-layer equation in the form

∂2풖 풇 × 풖 = −∇휙 + 퐴 , (5.78) ∂푧2 where 퐴 is a coefficient of viscosity. Suppose for the moment that the vertical scale of the flow is 퐻; the ratio of size of the frictional term to the Coriolis term is then (퐴/푓퐻2), and this ratio defines the , Ek: 퐴 Ek ≡ ( ). (5.79) 푓퐻2 101 5.7 Frictional Geostrophic Balance and Ekman Layers

In the fluid interior the Ekman number is usually small. In the Ekman layer itself, however, the viscous terms must be large (otherwise we are not in the Ekman layer), and the vertical scale must therefore be smaller than 퐻. If the viscous and Coriolis terms in (5.78) are required to be the same magnitude we obtain an Ekman layer depth, 훿, of

퐴 1/2 훿 ∼ ( ) . (5.80) 푓

The Ekman layer depth decreases with viscosity, becoming a thin bound- ary layer as 퐴 → 0. We also see that the Ekman number can be recast as the ratio of the Ekman layer depth to the vertical scale in the interior; that is, from (5.79) and (5.80), Ek = (훿/퐻)2.

Vertical velocity The vertical velocity at the edge of the Ekman layer may be calculated using the mass continuity equation, ∂푤 ∂푢 ∂푣 = − ( + ). (5.81) ∂푧 ∂푥 ∂푦 The right-hand side may be calculated from (5.76) and, noting that diver- gence of the geostrophic velocity is given by ∇ ⋅ 풗푔 = −훽푣푔/푓, we find that ∂푢 ∂푣 훽푣푔 ∂ 흉 + = − + curl ( ), (5.82) ∂푥 ∂푦 푓 ∂푧 푧 푓 푦 푥 where curl푧(흉/푓) = ∂푥(휏 /푓)−∂푦(휏 /푓). Consider an ocean Ekman layer in which the vertical velocity is zero at the surface and the stress is zero at the bottom of the layer. If we integrate (5.81) over the depth of that layer and use (5.82) we obtain

흉 훽푉푔 푤 = curl 0 − . (5.83) 퐸 푧 푓 푓 where 푉푔 = ∫ 푣푔 d푧 is the integral of the meridional geostrophic velocity over the Ekman layer. Evidently, friction induces a vertical velocity at the edge of the Ekman layer, proportional to the curl of the stress at the surface, and this is perhaps the most used result in Ekman layer theory. This result is particularly useful for the top Ekman layer in the ocean, where the stress can be regarded as a function of the overlying wind and is largely independent of the flow in the ocean; an equivalent result ap- plies in the atmosphere, but here the surface stress is a function of the geostrophic wind in the free atmosphere. If the curl of the at the top of the ocean is anticyclonic (i.e., negative if 푓 > 0, positive if 푓 < 0) then a downward velocity is induced at the base of the Ekman layer, a phenomenon known as Ekman pumping, and this is particularly important in setting the structure of the subtropical gyres in the ocean, discussed in Chapter 14. Chapter 5. Geostrophic Theory 102

Notes and References The first systematic derivation of the quasi-geostrophic equations was given by Charney (1948), although aspects of the concept, and the words‘quasi-geostrophy’, appeared in Durst & Sutcliffe (1938) and Sutcliffe (1947). The various forms of geostrophic equations were brought together in a review article by Phillips (1963) and since then both the quasi-geostrophic and planetary-geostrophic equations have been staples of dynamical meteorology and dynamical oceanography.

Problems 5.1 Do either or both: (a) Carry through the derivation of the quasi-geostrophic system start- ing with the anelastic equations and obtain (5.66). (b) Carry through the derivation of the quasi-geostrophic system in pressure coordinates. In each case, state the differences between your results and the Boussinesq result. 5.2 (a) The shallow water planetary geostrophic equations may be derived by simply omitting 휁 in the equation D 휁 + 푓 = 0 (P5.1) D푡 ℎ by invoking a small Rossby number, so that 휁/푓 is small. We then relate the velocity field to the height field by hydrostatic balance and obtain: D 푓 ∂ℎ ∂ℎ ( ) = 0, 푓푢 = −푔 , 푓푣 = 푔 . (P5.2) D푡 ℎ ∂푦 ∂푥 The assumptions of hydrostatic balance and small Rossby number are the same as those used in deriving the quasi-geostrophic equa- tions. Explain nevertheless how some of the assumptions used for quasi-geostrophy are in fact different from those used for planetary- geostrophy, and how the derivations and resulting systems differ from each other. Use any or all of the momentum and mass conti- nuity equations, scaling, nondimensionalization and verbal explana- tions as needed. (b) Explain if and how your arguments in part (a) also apply to the strati- fied equations (using, for example, the Boussinesq equations or pres- sure coordinates). 5.3 In the derivation of the quasi-geostrophic equations, geostrophic balance leads to the lowest-order horizontal velocity being divergence-free — that

is, ∇ ⋅ 풖0 = 0. It seems that this can also be obtained from the mass con- servation equation at lowest order. Is this a coincidence? Suppose that the Coriolis parameter varied, and that the momentum equation yielded

∇ ⋅ 풖0 ≠ 0. Would there be an inconsistency? 5.4 Consider a wind stress imposed by a mesoscale cyclonic (in the at- mosphere) given by 2 흉 = −퐴e−(푟/휆) (푦 퐢̂ − 푥 퐣),̂ (P5.3) where 푟2 = 푥2 + 푦2, and 퐴 and 휆 are constants. Also assume constant Cori- olis gradient 훽 = ∂푓/∂푦 and constant ocean depth 퐻. In the ocean, find 103 Problems

(a) the Ekman transport, (b) the vertical velocity 푤퐸(푥, 푦, 푧) below the Ek- man layer, (c) the northward velocity 푣(푥, 푦, 푧) below the Ekman layer and (d) indicate how you would find the westward velocity 푢(푥, 푦, 푧) below the Ekman layer. 5.5 In an atmospheric Ekman layer on the 푓-plane let us write the momentum equation as 1 ∂흉 풇 × 풖 = −∇휙 + , (P5.4) 휌푎 ∂푧

where 흉 = 퐴휌푎∂풖/∂푧 and 퐴 is a constant eddy viscosity coefficient. An independent formula for the stress at the ground is 흉 = 퐶휌푎풖, where 퐶 is a constant. Let us take 휌푎 = 1, and assume that in the free atmosphere the ̂ wind is geostrophic and zonal, with 풖푔 = 푈퐢. (a) Find an expression for the wind vector at the ground. Discuss the lim- its 퐶 = 0 and 퐶 = ∞. Show that when 퐶 = 0 the frictionally-induced vertical velocity at the top of the Ekman layer is zero. (b) Find the vertically integrated horizontal mass flux caused by the boundary layer. (c) When the stress on the atmosphere is 흉, the stress on the ocean beneath is also 흉. Why? Show how this is consistent with Newton’s third law. (d) Determine the direction and strength of the surface current, and the mass flux in the oceanic Ekman layer, in terms of the geostrophic wind in the atmosphere, the oceanic Ekman depth and the ratio 휌a/휌o, where 휌o is the density of the . Include a figure showing the direc- tions of the various and currents. How does the boundary-layer mass flux in the ocean compare to that in the atmosphere? (Assume, as needed, that the stress in the ocean may be parameterized with an eddy viscosity.) Partial solution for (a): A useful trick in Ekman layer problems is to write

the velocity as a complex number, ̂푢= 푢 + i푣 and ̂푢푔 = 푢푔 + i푣푔. The fundamental Ekman layer equation may then be written as

∂2푈̂ 퐴 = i푓푈,̂ (P5.5) ∂푧2 ̂ where 푈 = ̂푢− ̂푢푔. The solution to this is

(1 + i)푧 ̂푢− ̂푢 = [̂푢(0)− ̂푢 ] exp[− ], (P5.6) 푔 푔 푑

where 푑 = √2퐴/푓 and the boundary condition of finiteness at infinity eliminates the exponentially growing solution. The boundary condition

at 푧 = 0 is ∂ ̂푢/∂푧= (퐶/퐴)̂푢; applying this gives [̂푢(0)− ̂푢푔] exp(iπ/4) = −퐶푑̂푢(0)/(√2퐴), from which we obtain ̂푢(0), and the rest of the solution follows. Chapter 6 Rossby Waves

aves are familiar to almost everyone. Gravity waves cover the ocean surface, sound waves allow us to talk and light waves W enable us to see. This chapter provides an introduction to their properties, paying particular attention to a wave that is especially impor- tant to the large scale flow in both ocean and atmosphere — the . We start with an elementary introduction to wave kinematics, dis- cussing such concepts as phase speed and . Then, beginning with Section 6.3, we discuss the dynamics of Rossby waves, and this part may be considered to be the natural follow-on from the geostrophic the- ory of the previous chapter. Rossby waves then reappear frequently in later chapters.

6.1 Fundamentals and Formalities

6.1.1 Definitions and Kinematics A wave is more easily recognized than defined. Loosely speaking, a wave is a propagating disturbance that has a characteristic relationship between its frequency and size, called a dispersion relation. To see what all this means, and what a dispersion relation is, suppose that a disturbance, 휓(풙, 푡) (where 휓 might be velocity, streamfunction, pressure, etc.), satis- fies the equation 퐿(휓) = 0, (6.1) where 퐿 is a linear operator, typically a polynomial in time and space derivatives; one example is 퐿(휓) = ∂∇2휓/∂푡 + 훽∂휓/∂푥. If (6.1) has con- stant coefficients (if 훽 is constant in this example) then harmonic solu- tions may often be found that are a superposition of plane waves, each of which satisfy 휓 = Re 휓̃ei휃(풙,푡) = Re 휓̃ei(풌⋅풙−휔푡), (6.2) 104 105 6.1 Fundamentals and Formalities

Fig. 6.1: The propagation of a two-dimensional wave. (a) Two lines of constant phase (e.g., two wavecrests) at a

time 푡1. The wave is propa- gating in the direction 풌 with wavelength 휆. (b) A line of constant phase at two succes- sive times. The phase speed is the speed of advancement of the wavecrest in the direction

of travel, and so 푐푝 = 푙/(푡2 − 푡1). The phase speed in the 푥- direction is the speed of propagation of the wave- where 휓̃ is a complex constant, 휃 is the phase, 휔 is the wave frequency and crest along the 푥-axis, and 푥 푦 푧 풌 is the vector wavenumber (푘, 푙, 푚) (also written as (푘 , 푘 , 푘 ) or, in sub- 푥 푥 푐푝 = 푙 /(푡2 − 푡1) = 푐푝/ cos 휙. script notation, 푘푖). The prefix Re denotes the real part of the expression, but we will drop it if there is no ambiguity. Waves are characterized by having a particular relationship between the frequency and wavevector known as the dispersion relation. This is an equation of the form 휔 = 훺(풌), (6.3) where 훺(풌), or 훺(푘푖), and meaning 훺(푘, 푙, 푚), is some function deter- mined by the form of 퐿 in (6.1) and which thus depends on the particu- lar type of wave — the function is different for sound waves, light waves and the Rossby waves and gravity waves we will encounter in this book. Unless it is necessary to explicitly distinguish the function 훺 from the frequency 휔, we often write 휔 = 휔(풌).

6.1.2 and Phase Speed A common property of waves is that they propagate through space with some velocity, which in special cases might be zero. Waves in fluids may carry energy and momentum but do not necessarily transport fluid parcels themselves. Further, it turns out that the speed at which proper- ties like energy are transported (the group speed) may be different from the speed at which the wave crests themselves move (the phase speed). Let’s try to understand this statement, beginning with the phase speed. A summary of key results is given on page 107.

Phase speed Consider the propagation of monochromatic plane waves, for that is all that is needed to introduce the phase speed. Given (6.2) a wave will propa- gate in the direction of 풌 (Fig. 6.1). At a given instant and location we can align our coordinate axis along this direction, and we write 풌 ⋅ 풙 = 퐾푥∗, where 푥∗ increases in the direction of 풌 and 퐾2 = |풌|2 is the magnitude of the wavenumber. With this, we can write (6.2) as

∗ ∗ 휓 = Re 휓̃ei(퐾푥 −휔푡) = Re 휓̃ei퐾(푥 −푐푡), (6.4) Chapter 6. Rossby Waves 106

where 푐 = 휔/퐾. From this equation it is evident that the phase of the wave propagates at the speed 푐 in the direction of 풌, and we define the phase speed by 휔 푐 ≡ . (6.5) 푝 퐾 The wavelength of the wave, 휆, is the distance between two wavecrests — that is, the distance between two locations along the line of travel whose phase differs by 2π — and evidently this is given by 2π 휆 = . (6.6) 퐾 In (for simplicity) a two-dimensional wave, and referring to Fig. 6.1, the wavelength and wave vectors in the 푥- and 푦-directions are given by, 휆 휆 휆푥 = , 휆푦 = , 푘푥 = 퐾 cos 휙, 푘푦 = 퐾 sin 휙. (6.7) cos 휙 sin 휙 In general, lines of constant phase intersect both the coordinate axes and propagate along them. The speed of propagation along these axes is given by 푥 푦 푥 푙 푐푝 퐾 휔 푦 푙 푐푝 퐾 휔 푐 = 푐 = = 푐 = , 푐푝 = 푐 = = 푐 = , 푝 푝 푙 cos 휙 푝 푘푥 푘푥 푝 푙 sin 휙 푝 푘푦 푘푦 (6.8) using (6.5) and (6.7), and again referring to Fig. 6.1 for notation. The speed of phase propagation along any one of the axes is in general larger than the phase speed in the primary direction of the wave. The phase speeds are 푥 clearly not components of a vector: for example, 푐푝 ≠ 푐푝 cos 휙. Analo- gously, the wavevector 풌 is a true vector, whereas the wavelength 휆 is not. To summarize, the phase speed and its components are given by

휔 푥 휔 푦 휔 푐 = , 푐 = , 푐푝 = . (6.9) 푝 퐾 푝 푘푥 푘푦

6.1.3 The Dispersion Relation The above description is kinematic, in that it applies to almost any distur- bance that has a wavevector and a frequency. The particular dynamics of a wave are determined by the relationship between the wavevector and the frequency; that is, by the dispersion relation. Once the dispersion re- lation is known a great many of the properties of the wave follow in a more-or-less straightforward manner. Picking up from (6.3), the disper- sion relation is a functional relationship between the frequency and the wavevector of the general form 휔 = 훺(풌). (6.10) Perhaps the simplest example of a linear operator that gives rise to waves is the one-dimensional equation ∂휓 ∂휓 + 푐 = 0. (6.11) ∂푡 ∂푥 107 6.1 Fundamentals and Formalities

Wave Fundamentals • A wave is a propagating disturbance that has a characteristic relationship between its fre- quency and size, known as the dispersion relation. Waves typically arise as solutions to a linear problem of the form 퐿(휓) = 0, where 퐿 is, commonly, a linear operator in space and time. Two examples are

∂2휓 ∂ ∂휓 − 푐2∇2휓 = 0 and ∇2휓 + 훽 = 0, (WF.1) ∂푡2 ∂푡 ∂푥 where the second example gives rise to Rossby waves. • Solutions to the governing equation are often sought in the form of plane waves that have the form 휓 = Re 퐴ei(풌⋅풙−휔푡), (WF.2) where 퐴 is the wave amplitude, 풌 = (푘, 푙, 푚) is the wavevector, and 휔 is the frequency. • The dispersion relation connects the frequency and wavevector through an equation of the form 휔 = 훺(풌) where 훺 is some function. The relation is normally derived by substituting a trial solution like (WF.2) into the governing equation. For the examples of (WF.1) we obtain 휔 = 푐2퐾2 and 휔 = −훽푘/퐾2 where 퐾2 = 푘2 + 푙2 + 푚2 or, in two dimensions, 퐾2 = 푘2 + 푙2. • The phase speed is the speed at which the wave crests move. In the direction of propagation and in the 푥, 푦 and 푧 directions the phase speeds are given by, respectively,

휔 푥 휔 푦 휔 푧 휔 푐 = , 푐 = , 푐푝 = , 푐 = , (WF.3) 푝 퐾 푝 푘 푙 푝 푚

where 퐾 = 2π/휆 and 휆 is the wavelength. The wave crests have both a speed (푐푝) and a direction of propagation (the direction of 풌), like a vector, but the components defined in (WF.3) are not the components of that vector. • The group velocity is the velocity at which a or wave group moves. It is a vector and is given by

∂휔 푥 ∂휔 푦 ∂휔 푧 ∂휔 풄 = with components 푐 = , 푐푔 = , 푐 = . (WF.4) 푔 ∂풌 푔 ∂푘 ∂푙 푔 ∂푚 Most physical quantities of interest are transported at the group velocity.

Substituting a trial solution of the form 휓 = Re 퐴ei(푘푥−휔푡) into (6.11) we obtain (−i휔 + 푐i푘)퐴 = 0, giving the dispersion relation

휔 = 푐푘. (6.12)

The phase speed of this wave is 푐푝 = 휔/푘 = 푐. A couple of other examples of governing equations, dispersion relations and phase speeds are:

2 ∂ 휓 2 2 2 2 2 푥 푐퐾 푦 푐퐾 − 푐 ∇ 휓 = 0, 휔 = 푐 퐾 , 푐 = ±푐, 푐 = ± , 푐푝 = ± , ∂푡2 푝 푝 푘 푙 (6.13a)

∂ 2 ∂휓 −훽푘 휔 푥 훽 푦 훽푘/푙 ∇ 휓 + 훽 = 0, 휔 = , 푐 = , 푐 = − , 푐푝 = − , ∂푡 ∂푥 퐾2 푝 퐾 푝 퐾2 퐾2 (6.13b) Chapter 6. Rossby Waves 108

Fig. 6.2: Superposition of two sinusoidal waves with wavenumbers 푘 and 푘 + 훿푘, producing a wave (solid line) that is modulated by a slowly varying wave en- velope or packet (dashed). The envelope moves at the

group velocity, 푐푔 = ∂휔/∂푘, and the phase moves at

the group speed, 푐푝 = 휔/푘. where 퐾2 = 푘2 + 푙2 and the examples are two-dimensional, with variation in 푥 and 푦 only. A wave is said to be nondispersive if the phase speed is independent of the wavelength. This condition is satisfied for the simple example (6.11) but is manifestly not satisfied for (6.13b), and these waves (Rossby waves, in fact) are dispersive. Waves of different then travel at differ- ent speeds so that a group of waves will spread out — disperse — even if the medium is homogeneous. When a wave is dispersive there is another characteristic speed at which the waves propagate, the group velocity, and we come to this shortly. Most media are inhomogeneous, but if the medium varies sufficiently slowly in space and time — and in particular if the variations are slow compared to the wavelength and period — we may still have a local dis- persion relation between frequency and wavevector,

휔 = 훺(풌; 풙, 푡), (6.14)

where 푥 and 푡 are slowly varying parameters. We resume our discussion of this topic in Section 6.5, but before that we introduce the group velocity.

6.2 Group Velocity Group velocity seems to have been first articulated in about Information and energy do not, in general, propagate at the phase speed. 1841 by the Irish mathemati- cian and physicist William Rather, most quantities of interest propagate at the group velocity, a quan- Rowan Hamilton (1806–1865), tity of enormous importance in wave theory. Roughly speaking, group who is also remembered for velocity is the velocity at which a packet or a group of waves will travel, his formulation of ‘Hamilto- whereas the individual wave crests travel at the phase speed. To introduce nian mechanics’. Hamilton the idea we will consider the superposition of plane waves, noting that a was largely motivated by truly monochromatic plane wave already fills all space uniformly so that optics, and it was George there can be no propagation of energy from place to place. Stokes, and John Strutt (also known as Lord Rayleigh) who fur- 6.2.1 Superposition of Two Waves ther developed and gener- alized the idea in fluid dy- Consider the linear superposition of two waves. Limiting attention to namics in the nineteenth the one-dimensional case, consider a disturbance that is the sum of two and early twentieth centuries. waves, 휓 = Re 휓(̃ ei(푘1푥−휔1푡) + ei(푘2푥−휔2푡)). (6.15) 109 6.3 Rossby Wave Essentials

Let us further suppose that the two waves have similar wavenumbers and frequency, and, in particular, that 푘1 = 푘 + 훥푘 and 푘2 = 푘 − 훥푘, and 휔1 = 휔 + 훥휔 and 휔2 = 휔 − 훥휔. With this, (6.15) becomes

휓 = Re 휓̃ei(푘푥−휔푡)[ei(훥푘푥−훥휔푡) + e−i(훥푘푥−훥휔푡)] (6.16) = 2 Re 휓̃ei(푘푥−휔푡) cos(훥푘 푥 − 훥휔 푡).

The resulting disturbance, illustrated in Fig. 6.2 has two aspects: a rapidly The group velocity, 풄푔 is varying component, with wavenumber 푘 and frequency 휔, and a more given by 풄푔 = ∂휔/∂풌 or, slowly varying envelope, with wavenumber 훥푘 and frequency 훥휔. The in Cartesian components, envelope modulates the fast oscillation, and moves with velocity 훥휔/훥푘; 풄푔 = (∂휔/∂푘, ∂휔/∂푙, ∂휔/∂푚). in the limit 훥푘 → 0 and 훥휔 → 0 this is the group velocity, 푐 = ∂휔/∂푘. It is a vector, and it gives the 푔 velocity at which wave pack- Group velocity is equal to the phase speed, 휔/푘, only when the frequency ets, and hence energy, are is a linear function of wavenumber. The energy in the disturbance moves propagated. The phase speed, at the group velocity — note that the node of the envelope moves at the 푐푝 = 휔/푘, is the speed at speed of the envelope and no energy can cross the node. These concepts which the phase of an indi- generalize to more than one dimension, and if the wavenumber is the vidual wave moves. For some three-dimensional vector 풌 = (푘, 푙, 푚) then the three-dimensional enve- waves the group velocity and lope propagates at the group velocity given by phase speed may be in op- posite directions, so that the ∂휔 ∂휔 ∂휔 ∂휔 wave crests appear to move 풄푔 = ≡ ( , , ). (6.17) backwards through the wave ∂풌 ∂푘 ∂푙 ∂푚 packets.

The group velocity is also written as 풄푔 = ∇풌휔 or, in subscript notation, 푐푔푖 = ∂훺/∂푘푖, with the subscript 푖 denoting the component of a vector. The above derivation can be generalized to apply to the superposi- tion of many waves. It also applies when the medium through which the waves propagate is not homogeneous, provided that changes occur on a longer space scale than the wavelength of the waves. Energy and most other physically meaningful properties of waves travel at the group veloc- ity, as explicitly shown for Rossby waves in Section 9.2.2.

6.3 Rossby Wave Essentials

Rossby waves are the most prominent wave in the atmosphere and ocean on large scales, although gravity waves sometimes rival them. They can best be described using the quasi-geostrophic equations, as follows. The Rossby wave, the most well-known wave in meteo- 6.3.1 The Linear Equation of Motion rology and oceanography, is The relevant equation of motion is the inviscid, adiabatic potential vor- named for Carl-Gustav Rossby ticity equation in the quasi-geostrophic system, as discussed in Chapter who described the essential dy- namics in Rossby (1939). The 5, namely waves are in fact present in ∂푞 + 풖 ⋅ ∇푞 = 0, (6.18) the solutions of Hough (1898) ∂푡 although in a rather oblique where 푞(푥, 푦, 푧, 푡) is the potential vorticity and 풖(푥, 푦, 푧, 푡) is the hori- form and Rossby is normally zontal velocity. The velocity is in turn related to a streamfunction by given credit for having discov- 푢 = −∂휓/∂푦, 푣 = ∂휓/∂푥, and the potential vorticity is some function ered them. Chapter 6. Rossby Waves 110

of the streamfunction, which might differ from system to system. Two examples, one applying to a continuously-stratified system and the sec- ond to a single layer, are

∂ 푓2 ∂휓 푞 = 푓 + 휁 + ( 0 ) and 푞 = 휁 + 푓 − 푘2 휓, (6.19a,b) ∂푧 푁2 ∂푧 푑

2 where 휁 = ∇ 휓 is the relative vorticity and 푘푑 = 1/퐿푑 is the inverse radius Definitions of 푘푑 and 퐿푑 dif- of deformation for a shallow water system. fer from source to source We now linearize (6.18); that is, we suppose that the flow consists of a (and book to book), of- time-independent component (the ‘basic state’) plus a perturbation, with ten by factors of 2 or π the perturbation being small compared with the mean flow. The basic or similar. Caveat lector. state must satisfy the time-independent equation of motion, and it is com- mon and useful to linearize about a zonal flow, 푢(푦, 푧). The basic state is then purely a function of 푦 and so we can write

푞 = 푞(푦, 푧) + 푞′(푥, 푦, 푡), 휓 = 휓(푦, 푧) + 휓′(푥, 푦, 푧, 푡), (6.20)

with a similar notation for the other variables, and 푢 = −∂휓/∂푦 and 푣 = 0. Substituting into (6.18) gives, without approximation,

∂푞′ + 풖 ⋅ ∇푞 + 풖 ⋅ ∇푞′ + 풖′ ⋅ ∇푞 + 풖′ ⋅ ∇푞′ = 0. (6.21) ∂푡 The primed quantities are presumptively small so we neglect terms in- volving their products. Further, we are assuming that we are linearizing about a state that is a solution of the equations of motion, so that 풖⋅∇푞 = 0. Finally, since 푣 = 0 (since ∫ ∂휓/∂푥 d푥 = 0) and ∂푞/∂푥 = 0 we obtain

∂푞′ ∂푞′ ∂푞 + 푈 + 푣′ = 0, (6.22) ∂푡 ∂푥 ∂푦

where 푈 ≡ 푢. This equation or one very similar appears very commonly in studies of Rossby waves. Let us first consider the simple example of waves in a single layer.

6.3.2 Waves in a Single Layer Consider a system obeying (6.18) and (6.19b). The dynamics are more easily illustrated on a Cartesian 훽-plane for which 푓 = 푓0 + 훽푦, and since 푓0 is a constant it does not appear in our subsequent derivations.

Infinite deformation radius If the scale of motion is much less than the deformation scale then we make the approximation that 푘푑 = 0 and the vorticity equation may be written as ∂휁 + 풖 ⋅ ∇휁 + 훽푣 = 0. (6.23) ∂푡 111 6.3 Rossby Wave Essentials

We linearize about a constant zonal flow, 푈, by writing

∂ ∂∇2휓′ ∂휓′ ∇2휓′ + 푈 + 훽 = 0. (6.24) ∂푡 ∂푥 ∂푥 This equation is just a single-layer version of (6.22), with ∂푞/∂푦 = 훽, 푞′ = ∇2휓′ and 푣′ = ∂휓′/∂푥. The coefficients in (6.24) are not functions of 푦 or 푧; this is not a re- quirement for wave motion to exist but it does enable solutions to be found more easily. Let us seek solutions in the form of a plane wave, namely 휓′ = Re 휓̃ei(푘푥+푙푦−휔푡), (6.25) where 휓̃ is a complex constant. Solutions of this form are valid in a do- main with doubly-periodic boundary conditions; solutions in a channel can be obtained using a meridional variation of sin 푙푦, with no essential changes to the dynamics. The amplitude of the oscillation is given by 휓̃ and the phase by 푘푥+푙푦−휔푡, where 푘 and 푙 are the 푥- and 푦-wavenumbers and 휔 is the frequency of the oscillation. Substituting (6.25) into (6.24) yields

[(−휔 + 푈푘)(−퐾2) + 훽푘]휓̃ = 0, (6.26) where 퐾2 = 푘2 + 푙2. For non-trivial solutions the above equation implies

훽푘 휔 = 푈푘 − , (6.27) 퐾2 and this is the dispersion relation for barotropic Rossby waves. Evidently the velocity 푈 Doppler shifts the frequency by the amount 푈푘. The com- ponents of the phase speed and group velocity are given by, respectively,

푥 휔 훽 푦 휔 푘 훽푘 푐 ≡ = 푈 − , 푐푝 ≡ = 푈 − , (6.28a,b) 푝 푘 퐾2 푙 푙 퐾2푙 and 2 2 푥 ∂휔 훽(푘 − 푙 ) 푦 ∂휔 2훽푘푙 푐 ≡ = 푈 + , 푐푔 ≡ = . (6.29a,b) 푔 ∂푘 (푘2 + 푙2)2 ∂푙 (푘2 + 푙2)2 The phase speed in the absence of a mean flow is westward, with waves of longer wavelengths travelling more quickly, and the eastward current speed required to hold the waves of a particular wavenumber stationary 푥 2 (i.e., 푐푝 = 0) is 푈 = 훽/퐾 . The background flow 푈 evidently just provides a uniform shift to the phase speed, and (in this case) can be transformed away by a change of coordinate. The 푥-component of the group velocity may also be written as the sum of the phase speed plus a positive quantity, namely 2훽푘2 푐푥 = 푐푥 + . (6.30) 푔 푝 (푘2 + 푙2)2 This means that the zonal group velocity for Rossby wave packets moves 푥 eastward relative to its zonal phase speed. A stationary wave (푐푝 = 0) has Chapter 6. Rossby Waves 112

Fig. 6.3: A two-dimensional y (푥–푦) Rossby wave. An ini- η(t = 0) ζ < 0 tial disturbance displaces a . material line at constant lati- tude (the straight horizontal line) to the solid line marked 휂(푡 = 0). Conservation of . potential vorticity, 훽푦 + 휁, η(t > 0) leads to the production of ζ > 0 relative vorticity, 휁, as shown. The associated velocity field (arrows on the circles) then an eastward group velocity, and this has implications for the ‘downstream advects the fluid parcels, development’ of wave packets, but we will not pursue that topic here. and the material line evolves into the dashed line with the We see from (6.29) the group velocity is negative (westward) if the 푥- phase propagating westward. wavenumber is sufficiently small compared to the 푦-wavenumber. Es- sentially, long waves move information westward and short waves move in- formation eastward, and this is a common property of Rossby waves. The 푥-component of the phase speed, on the other hand, is always westward relative to the mean flow.

Finite deformation radius For a finite deformation radius the basic state 훹 = −푈푦 is still a solu- tion of the original equations of motion, but the potential vorticity corre- 2 2 ̂ sponding to this state is 푞 = 푈푦푘푑 +훽푦 and its gradient is ∇푞 = (훽+푈푘푑)퐣. The linearized equation of motion is thus

∂ ∂ ∂휓′ ( + 푈 ) (∇2휓′ − 휓′푘2 ) + (훽 + 푈푘2 ) = 0. (6.31) ∂푡 ∂푥 푑 푑 ∂푥

Substituting 휓′ = 휓̃ei(푘푥+푙푦−휔푡) we obtain the dispersion relation,

2 2 푘(푈퐾 − 훽) 훽 + 푈푘푑 휔 = 2 2 = 푈푘 − 푘 2 2 . (6.32) 퐾 + 푘푑 퐾 + 푘푑 It is clear from the second form of the above equation that the uniform velocity field no longer provides just a simple Doppler shift of the fre- quency, nor does it provide a uniform addition to the phase speed. This is because the current does not just provide a uniform translation, but, if 푘푑 is non-zero, it also modifies the basic potential vorticity gradient, as explored further in Problem 6.1.

6.3.3 The Mechanism of Rossby Waves The fundamental mechanism underlying Rossby waves may be under- stood as follows. Consider a material line of stationary fluid parcels along a line of constant latitude, and suppose that some disturbance causes their displacement to the line marked 휂(푡 = 0) in Fig. 6.3. In the displacement, the potential vorticity of the fluid parcels is conserved, and in the simplest 113 6.3 Rossby Wave Essentials

Essentials of Rossby Waves

• Rossby waves owe their existence to a gradient of potential vorticity in the fluid. If a fluid parcel is displaced, it conserves its potential vorticity and so its relative vorticity will in general change. The relative vorticity creates a velocity field that displaces neighbouring parcels, whose relative vorticity changes and so on. • A common source of a potential vorticity gradient is differential rotation, or the 훽-effect, and in this case the associated Rossby waves are called planetary waves. In the presence of non-zero 훽 the ambient potential vorticity increases northward and the phase of the Rossby waves propagates westward. Topography is another source of potential vorticity gradients. In general, Rossby waves propagate to the left of the direction of increasing potential vortic- ity. • A common equation of motion for Rossby waves is

∂푞′ ∂푞′ ∂푞 + 푈 + 푣′ = 0, (RW.1) ∂푡 ∂푥 ∂푦 with an overbar denoting the basic state and a prime a perturbation. In the case of a single layer of fluid with no mean flow this equation becomes

∂ ∂휓′ (∇2 + 푘2 )휓′ + 훽 = 0, (RW.2) ∂푡 푑 ∂푥 with dispersion relation −훽푘 휔 = 2 2 2 . (RW.3) 푘 + 푙 + 푘푑

푥 • In the absence of a mean flow (i.e., 푈 = 0), the phase speed in the zonal direction (푐푝 = 휔/푘) is always negative, or westward, and is larger for large waves. For (RW.3) the components of the group velocity are given by

2 2 2 푥 훽(푘 − 푙 − 푘푑) 푦 2훽푘푙 푐푔 = , 푐푔 = . (RW.4) 2 2 2 2 2 2 2 2 (푘 + 푙 + 푘푑) (푘 + 푙 + 푘푑) The group velocity is westward if the zonal wavenumber is sufficiently small, and eastward if the zonal wavenumber is sufficiently large. case of barotropic flow on the 훽-plane the potential vorticity is the abso- lute vorticity, 훽푦 + 휁. Thus, in either hemisphere, a northward displace- ment leads to the production of negative relative vorticity and a south- ward displacement leads to the production of positive relative vorticity. The relative vorticity gives rise to a velocity field which, in turn, advects the parcels in the material line in the manner shown, and the wave prop- agates westward. In more complicated situations, such as flow in two layers, considered below, or in a continuously-stratified fluid, the mechanism is essentially the same. A displaced fluid parcel carries with it its potential vorticity and, Chapter 6. Rossby Waves 114

in the presence of a potential vorticity gradient in the basic state, a poten- tial vorticity anomaly is produced. The potential vorticity anomaly pro- duces a velocity field (an example of potential vorticity inversion) which further displaces the fluid parcels, leading to the formation of a Rossby wave. The vital ingredient is a basic state potential vorticity gradient, such as that provided by the change of the Coriolis parameter with latitude.

6.4 Rossby Waves in Stratified Quasi-Geostrophic Flow

We now consider the dynamics of linear waves in stratified quasi- geostrophic flow on a 훽-plane, with a resting basic state. The interior flow is governed by the potential vorticity equation, (5.63), and lineariz- ing this about a constant mean zonal flow, 푈, gives

∂ ∂ ∂ 푓2 ∂휓′ ∂휓′ ( + 푈 ) [∇2휓′ + ( 0 )] + 훽 = 0, (6.33) ∂푡 ∂푥 ∂푧 푁2 ∂푧 ∂푥

2 2 where, for simplicity, we suppose that 푓0 /푁 does not vary with 푧.

6.4.1 Dispersion Relation and Group Velocity Deferring the issues of boundary conditions for a moment, let us seek solutions of the form

휓′ = Re 휓̃ei(푘푥+푙푦+푚푧−휔푡), (6.34)

where 휓̃ is a constant. Substituting (6.34) into (6.33) gives, after a couple of lines of elementary algebra, the dispersion relation

훽푘 휔 = 푈푘 − 2 2 2 2 2 . (6.35) 푘 + 푙 + 푚 푓0 /푁

As in most wave problems the frequency is, by convention, a positive quantity. In reality we usually have to satisfy a boundary condition at the top and bottom of the fluid and these are determined by the thermodynamic equation, (5.67), which after linearization becomes ∂ ∂ ∂휓′ ( + 푈 )( ) + 푁2푤 = 0. (6.36) ∂푡 ∂푥 ∂푧 If the boundaries are flat, rigid, surfaces then 푤 = 0 at those boundaries suggesting that instead of (6.34) we choose a streamfunction of the form

푚푗π푧 휓′ = Re 휓̃ei(푘푥+푙푦−휔푡) cos ( ) , 푚 = 1, 2 … , (6.37) 퐻 푗 which satisfies ∂휓′/∂푧 = 0 at 푧 = 0, 퐻. The dispersion relation is the same as (6.35) with 푚 equal to 푚푗π/퐻. If the domain is finite in the horizontal direction also — as it nearly always is — then the horizontal wavenum- bers are also quantized. For example, if the extent of the domain in the 115 6.4 Rossby Waves in Stratified Quasi-Geostrophic Flow

Phase propagation Fig. 6.4: A schematic east- west section of an upwardly propagating Rossby wave. The slanting lines are lines of constant phase and ‘high’ and Warm, poleward ‘low’ refer to the pressure or streamfunction values. Both 푘 and 푚 are negative so the Cool, equatorward phase lines are oriented up and to the west. The phase propagates westward and High downward, but the group velocity is upward.

Low

High

푥-direction is 퐿푥 and the flow is periodic in that direction, then 푘 is re- stricted to the values 푘 = 2π푘푗/퐿푥 where 푘푗 is an integer, and similarly for 푙 and 푙푗. (It is common albeit sometimes confusing to refer to both 푘 and 푘푗 as the wavenumber and also to drop the subscript 푗 on 푘푗 with the distinction then made clear by context, and similarly for 푙 and 푙푗.) Using (6.35), the three components of the group velocity for these waves are:

2 2 2 2 2 푥 훽[푘 − (푙 + 푚 푓0 /푁 )] 푐푔 = 푈 + , (6.38a) 2 2 2 2 2 2 (푘 + 푙 + 푚 푓0 /푁 ) 2 2 푦 2훽푘푙 푧 2훽푘푚푓0 /푁 푐푔 = , 푐푔 = . (6.38b,c) 2 2 2 2 2 2 2 2 2 2 2 2 (푘 + 푙 + 푚 푓0 /푁 ) (푘 + 푙 + 푚 푓0 /푁 ) The propagation in the horizontal is similar to the propagation in a single- layer model. We see also that higher baroclinic modes (bigger 푚) will have a more westward group velocity. The vertical group velocity is propor- tional to 푚, and for waves that propagate signals upward we choose 푚 to 푧 have the same sign as 푘 so that 푐푔 is positive. If there is no mean flow then the zonal wavenumber 푘 is negative (in order that frequency is positive) and 푚 must then also be negative. Energy then propagates upward but the phase propagates downward! This case is illustrated in Fig. 6.4 and to understand it better let us look at the vertical propagation in more detail. Vertical propagation of Rossby waves was not discussed by Rossby, who was mainly in- 6.4.2 Vertical Propagation of Rossby Waves terested in their properties in Conditions for wave propagation a horizontal plane. Rather, it was Charney & Drazin (1961) With a little re-arrangement the dispersion relation, (6.35), may be writ- who first considered the verti- ten as cal propagation of the waves 푁2 훽 in a stratified fluid in any de- 푚2 = ( − (푘2 + 푙2)) , 2 (6.39) tail. 푓0 푈 − 푐 Chapter 6. Rossby Waves 116

Fig. 6.5: The boundary be- tween propagating waves and evanescent waves as a function of zonal wind and wavenumber, calculated using (6.41). The abscissa

is the 푥-wavenumber, 푘푗, and the results show the boundary with two values of

meridional wavenumber, 푙푗. Small wavenumbers (large scales) can more easily prop- agate into the stratosphere, and propagation is inhib- ited if the zonal wind is too strong or is negative.

where 푐 = 휔/푘. For waves to propagate upwards we require that 푚2 > 0: if 푚2 < 0 the wavenumber is imaginary and the wave amplitude either grows with height (which is unphysical) or is damped. This condition implies 훽 0 < 푈 − 푐 < . (6.40) 푘2 + 푙2 For waves of some given frequency (휔 = 푘푐) the above expression pro- vides a condition on 푈 for the vertical propagation of planetary waves. Stationary, vertically oscilla- For stationary waves 푐 = 0 and the criterion is tory modes can exist only for zonal flows that are eastward 훽 and that are less than some 0 < 푈 < 2 2 , (6.41) 2 2 푘 + 푙 critical value 푈푐 = 훽/(푘 + 푙 ). as illustrated in Fig. 6.5. That is to say, the vertical propagation of sta- tionary Rossby waves occurs only in eastward winds, and winds that are 2 2 weaker than some critical value, 푈푐 = 훽/(푘 +푙 ) that depends on the scale of the wave. The lower limit, where 푈 = 푐 (and so 푈 = 0 if stationary) is known as a critical level and the upper limit is a ‘turning level’. Since (6.39) is just a form of the dispersion relation, for any given frequency there is a criterion for propagation given by (6.40), and for any given 푈 there may be a frequency that allows propagation. However, the zero frequency case is often regarded as the most important. 2 2 The upper critical velocity, 푈푐 = 훽/(푘 + 푙 ), is a function of wavenum- ber and increases with horizontal wavelength. Thus, for a given eastward flow long waves may penetrate vertically when short waves are trapped, an effect sometimes referred to as ‘Charney–Drazin filtering’. An impor- tant consequence is that the stratospheric motion is typically of larger scales than that of the troposphere, because Rossby waves tend to be ex- cited first in the troposphere (by baroclinic instability and by flow over topography, among other things), but the shorter waves are trapped and 117 6.4 Rossby Waves in Stratified Quasi-Geostrophic Flow only the longer ones reach the stratosphere. In the summer, the strato- spheric winds are often westwards (because polar regions in the strato- sphere are warmer than equatorial regions) and all waves are trapped in the troposphere; the eastward stratospheric winds that favour vertical penetration occur in the other three seasons, although very strong east- ward winds can suppress penetration in mid-winter. For westward flow, or for sufficiently strong eastward flow, 푚 is imag- inary and the waves decay exponentially in the vertical as exp(−훼푧) where

푁 훽 1/2 훼 = (푘2 + 푙2 − ) , (6.42) 푓0 푈 and the decay scale is smaller for shorter waves.

An interpretation One physical way to interpret the propagation criterion is to write (6.41) as |훽푘| 0 < |푈푘| < , (6.43) 푘2 + 푙2 allowing for the fact that 푘 may be negative. Now, in a resting medium (푈 = 0) the Rossby wave frequency has a maximum value when 푚 = 0 given by |훽푘| 휔 = , (6.44) 푘2 + 푙2 and the minimum frequency is zero. Let us suppose that the stationary Rossby waves are excited by flow of speed 푈 moving over bottom topog- raphy. If we move into the frame of reference of the flow then the waves are forced by a moving topography, and the frequency of the forcing is just |푈푘|. Thus, (6.43) is equivalent to saying that for oscillatory waves to exist the forcing frequency, |푈푘|, must lie within the frequency range of verti- cally propagating Rossby waves.

6.4.3 Heat Transport and Vertical Propagation If the group velocity in the 푧-direction, given by (6.38) is to be positive, then we require the product 푘푚 > 0. This has an important consequence for the heat transport. Remember that the buoyancy 푏, which is a proxy for , is given by 푓0∂휓/∂푧, and the northward velocity is 푣 = ∂휓/∂푥. Thus, the northward flux of heat, 퐻 say, is given by

∂휓 ∂휓 퐻 = 푣푏 = 푓 , (6.45) 0 ∂푧 ∂푥 where an overbar denotes a zonal average. To evaluate this expression it is simplest to suppose that 휓 = 휓0 cos(푘푥 + 푙푦 + 푚푧), where 휓0 is a real constant, in which case we obtain 1 퐻 = 푓 푘푚휓2 sin2(푘푥 + 푙푦 + 푚푧) = 푓 휓2푘푚, (6.46) 0 0 2 0 0 Chapter 6. Rossby Waves 118

Fig. 6.6: Idealised trajec- Wave packet collision. tory of two wavepackets, Ray theory fails. each with a different wave- length and moving with a different group velocity, as Trajectory 1 might be calculated using ray Time Trajectory 2 theory. If the wave packets collide ray theory must fail. Ray theory gives only the tra- jectory of the wave packet, not the detailed structure of the waves within a packet.

Distance, x

using the standard result that the average of sin2푥 over a wavelength is equal to 1/2. Thus, the heat flux, like the vertical component of the group velocity, is equal to 푘푚 multiplied by a positive quantity, and we may con- clude that an upward propagation of Rossby waves is associated with a polewards heat flux. A moment’s thought will reveal that the same conclu- sion holds in the Southern Hemisphere, even though 푓0 is negative there.

6.5 Ray Theory and Rossby Rays

In the real world most waves propagate in a medium that is inhomoge- neous; for example, in the Earth’s atmosphere and ocean the Coriolis pa- rameter varies with latitude and the stratification with height. In these cases it can be hard to obtain the solution of a wave problem by Fourier methods, even approximately. Nonetheless, the idea of signals propagat- ing at the group velocity is a robust one, and we can often obtain some of the information we want — and in particular the trajectory of a wave — using a recipe known as ray theory. If the background properties of the medium vary only slowly compared to the wavelength, we assume that the dispersion relation is satisfied locally. We then calculate the group velocity, and assume that the trajectory of the wave is along that group velocity. To implement this recipe, we first obtain a dispersion relation from the governing equation. In the homogeneous case we obtained a disper- sion relation 휔 = 푓(풌); that is the frequency is some function only of the wavenumber. In the inhomogeneous case we proceed the same way, and obtain a dispersion relation as if the parameters were fixed, but we then allow the parameters to vary in the dispersion relation; that is, we obtain a dispersion relation of the form 휔 = 푓(풌; 풙, 푡). For example, for barotropic 119 Appendix A: The WKB Approximation for Linear Waves

Rossby waves, the dispersion relation might now be 훽(푦) 휔 = 푈(푦) − . (6.47) 푘2 + 푙2 This is the same as the usual dispersion relation except that 푈 and 훽 are now taken to be slowly varying functions of position. This procedure is valid provided that 푈 and 훽 vary more slowly than the wavelengths of the Rossby waves, which is essentially the same condition used in the WKB approximation, discussed in the appendix below. The utility of ray theory is that it allows us to calculate the trajectory of a wave packet without fully solving the problem. Thus, suppose that a disturbance somewhere generates Rossby waves — perhaps a -surface temperature anomaly in the tropics. The waves propagate away from the disturbance following the group velocity, which we calculate using the dispersion relation. Thus, if the dispersion relation is given by (6.47) then the group velocity is given by

2 2 푥 훽(푘 − 푙 ) 푦 2푘푙훽 푐푔 = 푈 + 2 , 푐푔 = 2 , (6.48a,b) (푘2 + 푙2) (푘2 + 푙2) where 훽 and 푈 are functions of space. Thus, the position of the wave packet, 풙, is given by solving (often numerically) the ray equations

d푥 푥 d푦 푦 = 푐 , = 푐푔 , (6.49) d푡 푔 d푡 푥 푦 where 푐푔 and 푐푔 are evaluated using (6.48). Figure 6.6 shows schemati- cally how a wave packet might then propagate. In the Rossby-wave case here, the frequency is constant (determined by the source of the waves), as is the 푥-wavenumber, 푘, because 푈 and 훽 are only functions of 푦. As we move along the ray, we recalculate the 푦-wavenumber, 푙, at each point using the dispersion relation and then calculate the group velocity at each point, and thence the trajectory. When this is done on Earth, one some- times can see Rossby waves propagating from the tropics to midlatitude In 1926, G. Wentzel, H. A. that bring about long-range correlations in weather patterns known as Kramers and L. Brillouin each ‘teleconnections’. presented a method to find approximate solutions of the Schrödinger equation in ♦ Appendix A: The WKB Approximation quantum mechanics, and the technique became known as The WKB method is a way of finding approximate solutions to certain the WKB method. It turns linear differential equations in which the term with the highest derivative out that the technique had is multiplied by a small parameter. In particular, WKB theory can be used already been discovered by to find approximate solutions to wave equations in which the coefficients Harold Jeffreys, a mathemat- vary slowly in space or time. Consider an equation of the form ical geophysicist, a few years prior to that, and in fact the d2휉 origins of the method go back + 푚2(푧)휉 = 0. (6.50) d푧2 to the early nineteenth cen- tury. Evidently, methods are Such an equation commonly arises in wave problems. If 푚2 is positive the named after the last people to equation has wavelike solutions, and if 푚 is constant the solution has the discover them. Chapter 6. Rossby Waves 120

harmonic form i푚푧 휉 = Re 퐴0 e , (6.51)

where 퐴0 is a complex constant. If 푚 varies only slowly with 푧 (meaning that the variations in 푚 only occur on a scale much longer than 1/푚) one might reasonably expect that the harmonic solution above would provide a decent first approximation; that is, we expect the solution to locally look like a plane wave with local wavenumber 푚(푧). However, we might also expect that the solution would not be exactly of the form exp(i푚(푧)푧), be- cause the phase of 휉 is 휃(푧) = 푚푧, so that d휃/d푧 = 푚+푧d푚/d푧 ≠ 푚. Thus, in (6.51) 푚 is not the wavenumber unless 푚 is constant. The condition that variations in 푚, or in the wavelength 휆 ∼ 푚−1, occur only slowly may be variously expressed as

∂휆 ∂푚−1 ∂푚 휆 | | ≪ 휆 or | | ≪ 1 or | | ≪ 푚2. (6.52a,b,c) ∂푧 ∂푧 ∂푧 This condition will generally be satisfied if variations in the background state, or in the medium, occur on a scale much longer than the wavelength.

The Solution Let us seek solutions of (6.50) in the form

휉 = 퐴(푧)ei휃(푧), (6.53)

where 퐴(푧) and 휃(푧) are both real. Using (6.53) in (6.50) yields

d퐴 d휃 d2휃 d휃 2 d2퐴 i[2 + 퐴 ] + [퐴 ( ) − − 푚2퐴] = 0. (6.54) d푧 d푧 d푧2 d푧 d푧2 The terms in square brackets must each be zero. The WKB approxi- mation is to assume that the amplitude varies sufficiently slowly that |퐴−1d2퐴/d푧2| ≪ 푚2, and hence that the term involving d2퐴/d푧2 may be neglected. The real and imaginary parts of (6.54) become

d휃 2 d퐴 d휃 d2휃 ( ) = 푚2, 2 + 퐴 = 0. (6.55a,b) d푧 d푧 d푧 d푧2 The solution of the first equation above is

휃 = ± ∫ 푚 d푧, (6.56)

and substituting this into (6.55b) gives d퐴 d푚 2 푚 + 퐴 = 0, with solution 퐴 = 퐴 푚−1/2, (6.57a,b) d푧 d푧 0

where 퐴0 is a constant. Using (6.56) and (6.57b) in (6.53) gives us

−1/2 휉(푧) = 퐴0 푚 exp (±i ∫ 푚 d푧) , (6.58) 121 Problems and this is the WKB solution to (6.50). In terms of real quantities the solution may be written

−1/2 −1/2 휉(푧) = 퐵0 푚 cos (∫푚 d푧) + 퐶0 푚 sin (∫푚 d푧) , (6.59) where 퐵0 and 퐶0 are real constants. Using (6.55a) and the real part of (6.54) we see that the condition for the validity of the approximation is that

d2퐴 |퐴−1 | ≪ 푚2, (6.60a,b) d푧2 which using (6.57b) is

1 d2푚−1/2 | | ≪ 푚2. (6.61) 푚−1/2 d푧2

Equation (6.52) expresses a similar condition to (6.61).

Notes and References Many books discuss waves in fluids with the ones by LeBlond & Mysak (1980), Pedlosky (2003) and Gill (1982) having a geophysical emphasis. Accessible intro- ductions to WKB theory can be found in Simmonds & Mann (1998) and Holmes (2013). An example of the application of WKB theory to the atmosphere is given in Hoskins & Karoly (1981).

Problems 6.1 Rossby waves and Galilean invariance in shallow water. Consider the flat- bottomed shallow-water quasi-geostrophic equations in standard nota- tion, D 푓 휂 (휁 + 훽푦 − 0 ) = 0. (P6.1) D푡 퐻 (a) How is 휁 related to 휂? Express 푢, 푣, 휂 and 휁 in terms of a streamfunction. (b) Linearize (P6.1) about a state of rest, and show that the resulting sys- tem supports two-dimensional Rossby waves. Discuss the limits in which the wavelength is much shorter or much longer than the defor- mation radius. (c) Now linearize (P6.1) about a geostrophically balanced state that is trans- lating uniformly eastwards. This means that: 푢 = 푈 + 푢′ and 휂 = 휂(푦) + 휂′, where 휂(푦) is in geostrophic balance with 푈. Obtain an expression for the form of 휂(푦). Obtain the dispersion relation for Rossby waves in this system. Show that their speed is different from that obtained by adding a constant 푈 to the speed of Rossby waves in part (b). The problem is therefore not Galilean invariant — why? 6.2 Horizontal propagation of Rossby waves. Consider barotropic Rossby waves obeying the dispersion relation 휔 = 푈푘 − 훽푘/((푘2 + 푙2), where 푈 and/or 훽 vary slowly with latitude. Chapter 6. Rossby Waves 122

(i) By re-arranging this expression to obtain an expression for the meridional wavenumber, or otherwise, show that the Rossby waves can only propagate in the horizontal if 훽 0 < 푈 − 푐 < , (P6.2) 푘2 where 푐 = 휔/푘. (ii) If the waves approach a latitude where 푈 = 푐 (a ‘critical latitude’) show that the meridional wavenumber 푙 becomes large but the group velocity in the 푦-direction becomes small. Show that Rossby waves generated in the atmosphere in midlatitudes are unlikely to propa- gate into the tropics (where the mean flow is westward). (iii) Suppose that the Rossby waves approach a latitude where 푈 − 푐 = 훽/푘2. Calculate the 푥- and 푦-components of the group velocity in this limit and infer that a wave will turn away from such a latitude. 6.3 Consider barotropic Rossby waves with an infinite deformation radius. Obtain an expression for the group velocity in the presence of a uniform eastward mean flow. Show that for stationary waves the group velocity is eastward relative to Earth’s surface, and hence deduce that energy propa- gation is downstream of topographic sources. Is this still true if the defor- mation radius is finite? Is it true if the mean flow is westward? 6.4 Rossby waves in the two-level model. (a) Consider a two-layer (or two-level) quasi-geostrophic system on a 훽-plane, and you may consider the layers to be of equal depth. Lin- earize the equations about a state of rest, and show that they may be transformed into two uncoupled equations. Give a physical interpre- tation of what these equations represent. (b) Obtain the dispersion relations for this system (there are two). Show that one of the dispersion relations corresponds to the synchronous, depth-independent motion in the two layers. What does the other one correspond to? Obtain the phases and group velocities of these waves and determine the conditions under which they are eastward or westward. 6.5 In realistic calculations of the vertical propagation of Rossby waves one must take into account the vertical variation of density. Carry through the calculation leading to the Charney–Drazin condition, either using pres- sure (or log-pressure) coordinates or using the anelastic version of quasi-

geostrophy, using (5.66) with ̃휌= 휌0 exp(−푧/퐻). Show that the condition for propagation analogous to (6.41) becomes 훽 0 < 푈 < , (P6.3) 푘2 + 푙2 + 훾2

and obtain an expression for 훾.