Potential Momentum and the equilibration of short Charney waves

by

Pablo Zurita-Gotor∗ GFDL/UCAR Visiting Scientist Program. Princeton, NJ 08542 and Richard S. Lindzen Program in Atmospheres, Oceans and Climate, MIT, Cambridge, MA 02139

Submitted to Meteorologische Zeitschrift Submitted: September 2005 Revised: March 2006



Corresponding author address: Dr. Pablo Zurita-Gotor GFDL, Rm 237

US Route 1. Princeton, NJ 08542 Email: [email protected]

Abstract

This paper proposes a new formalism for studying the extratropical circulation, based on the concept of potential momentum. Mathematically, the potential momentum is defined as the zonal momentum profile that produces the same potential vorticity distribution as the baroclinic term. Physically, easterly (westerly) potential momentum is associated with isentropic layers that open up (close down) with latitude. The former is the case in the extratropical troposphere, and gives a negative contribution to the interior potential vorticity gradient. Likewise, the surface temperature variance is interpreted in this framework as a reservoir of easterly potential momentum at the surface. With this redefinition of the mean flow, the wave-mean flow interaction problem can be recast in a form that is reminiscent of the barotropic framework. In particular, the eddy PV flux is the only wave-mean flow interaction term, and exchanges eddy pseudomomentum and mean flow momentum (which includes both the standard momentum and potential momentum) locally. On the other hand, the partition of the mean flow momentum between its physical and potential components is determined remotely, subject to the thermal wind constraint. This constraint is enforced by a residual circulation that exchanges both terms without affecting the total momentum or the full PV gradient. The potential momentum framework relates thermal homogenization at the surface to the net interior eddy absorption, regardless of whether this results from a meridional or a vertical EP convergence. This helps explain previous results by the same authors that a limited interior PV gradient does not constrain thermal homogenization at the surface, since in the 3D problem the waves can also refract meridionally to be absorbed at other latitudes.

1

1

1.

Introduction

2

In a recent two-paper series (Zurita-Gotor and Lindzen, 2004a, b), we have discussed the equi-

3

libration of short Charney waves, with the goal of understanding what prevents the eddies from

4

eliminating the extratropical surface temperature gradient. Since it is this temperature gradient

5

that provides the required sign change for instability (Charney and Stern, 1962), this is a central

6

question for theories of baroclinic adjustment, which assume a neutral mean state (see Zurita-

7

Gotor and Lindzen (2006b) for a review). A possible explanation was given by Lindzen (1993),

8

who proposed the Eady problem at the shortwave cutoff as the paradigm for the neutralized state.

9

However, it was noted by Zurita and Lindzen (2001) that the degree of PV homogenization in the

10

extratropical troposphere is comparable in the interior and at the surface, and that the interior PV

11

gradient only vanishes across a shallow region surrounding the steering level. These authors also

12

showed that short Charney waves (i.e., waves that see a small interior PV gradient) could also be

13

neutral when the PV gradient vanishes at the steering level alone.

14

Motivated by that analysis, Zurita-Gotor and Lindzen (2004a) investigated the equilibration of

15

short Charney waves in the 2D case, using the barotropic analog to the Charney problem proposed

16

by Lindzen et al. (1983). Their results supported the hypothesis that the baroclinic problem could

17

equilibrate through partial PV homogenization at the steering level. However, Zurita-Gotor and

18

Lindzen (2004b, ZL04 hereafter) found that in the 3D problem, this was only the case in the presence

19

of sufficiently strong surface friction. Otherwise, the interior PV gradient is greatly reinforced by

20

the meridional curvature of the enhanced barotropic jet and is no longer limited as for a short

21

Charney mode. ZL04 interpreted these results in terms of the full two-dimensional (in y and z)

22

redistribution of the zonal-mean momentum. The authors argued, based on the association of the

23

Eliassen-Palm (EP) flux F with a flux of easterly pseudomomentum (Edmon et al., 1980), that

24

the baroclinic equilibration may be interpreted in terms of the export of easterly momentum away 2

25

from the surface baroclinic zone. While in the purely baroclinic case this momentum can only be

26

transferred vertically (and might thus be limited by the depth of the modes), in the full 3D case

27

this constraint no longer applies because the easterly momentum can also be transferred laterally.

28

ZL04 implicitly assumed that the EP flux convergence ∇ · F = v 0 q 0 was associated with a local

29

modification of the mean flow. However, this association is not necessarily justified because, al-

30

though v 0 q 0 encapsulates the net wave-mean flow interaction, the mean flow responds in a non-local

31

manner to this forcing. To fix ideas, consider the 2D zonal momentum balance first: ∂U + αM U = v 0 ξ 0 ∂t

(1)

32

where αM is a frictional time scale. As discussed by Zurita-Gotor and Lindzen (2004a), the local

33

balance in equilibrium between the eddy vorticity flux and the mean flow forcing implies that the

34

equilibrated flow must maintain a negative (positive) PV gradient over those regions where the mean

35

flow acceleration is westerly (easterly), so that the eddy vorticity flux (assumed downgradient) can

36

support the mean flow imbalance against friction. In contrast, in the 3D case

37

∂U + αM U = f 0 v ∗ + v 0 q 0 ∂t 38

there is not such a local balance because the mean flow response involves a vertical redistribution

39

of the eddy forcing by the residual circulation. Hence, it is not possible in principle to associate

40

this eddy forcing with a local modification of the mean flow.

41

However, it will be shown in this paper that a generalization of the barotropic relation Eq. 1 can

42

still be obtained in the 3D case using the concept of potential momentum recently introduced by

43

Zurita-Gotor and Lindzen (2006a, ZL06 hereafter). This generalization provides the theoretical

44

justification for the picture proposed above. In particular, the surface temperature variance is

45

reinterpreted as an easterly momentum source, depleted by the surface eddy heat flux, while the 3

46

interior PV flux appears as a local easterly forcing for the mean flow. The theoretical framework is

47

presented in section 2, which essentially follows a parallel derivation in ZL06. Section 3 illustrates

48

these concepts by presenting some new diagnostics for the problem of equilibrating short Charney

49

waves. Section 4 concludes with a brief summary.

50

2.

51

Formulation

52

We start with the quasigeostrophic Boussinesq zonal-mean equations on the β plane:

The concept of potential momentum

∂U ∂ 0 0 + u v − f0 v a = −αM U ∂t ∂y ∂θ ∂ 0 0 v θ + w a Θz = −αT (θ − θ R ) + ∂t ∂y ∂v a ∂wa + =0 ∂y ∂z g ∂θ ∂U =− ∂z f0 Θ0 ∂y

(2) (3) (4) (5)

53

where the subscript a stands for ageostrophic variables. α M , αT are the linear damping coefficients

54

for Rayleigh friction and Newtonian cooling, and θ R is the radiative equilibrium profile to which

55

temperature is relaxed. We will assume α T to be constant, while αM is allowed to be height-

56

dependent (for instance, it may vanish above the boundary layer). Finally, Θ z is the reference

57

stratification of qg theory, which may also be height-dependent, and θ represents differences from

58

that reference state.

59

We will rewrite the thermodynamic equation in the form of a momentum equation so as to put

60

it into a form similar to that of Eq. 2, the momentum equation. We multiply Eq. 3 by −f 0 /Θz ,

61

integrate meridionally between y0 and y and differentiate with respect to z. Assuming that a 4

62

latitude y0 = 0 exists, such that v a = v 0 = 0 at y0 , the thermodynamic equation can then be

63

written: ∂ ∂M − ∂t ∂z

64



f0 0 0 v θ + f0 v a = −αT (M − M R ) Θz 

(6)

where M =−

Z

y 0

∂ ∂z

f0 θ dy 0 Θz 



(7)

65

is defined to be the potential momentum.

66

Note that M has momentum units, and that Eq.6 looks like a momentum equation. In particular,

67

the second term in that equation is interpreted in the Eliassen-Palm formalism as a vertical eddy

68

momentum flux (Edmon et al., 1980). This term can also be rewritten:

69

∂ − ∂z



f0 0 0 ∂ v θ = −v 0 Θz ∂z 



f0 0 ∂m0 θ = v0 Θz ∂y 

(8)

70

where m0 is the eddy component of the potential momentum M , and we took into account the

71

thermal wind relation for v 0 and the definition,

72

written in Transformed Eulerian Mean (TEM) form:

∂M ∂y

∂ = − ∂z



f0 Θz θ



. Eqs. 2, 6 can alternatively be

73

∂ ∂z



v0 θ0 Θz



∂U − f0 v ∗ = v 0 q 0 − αM U ∂t

(9)

  ∂M + f0 v ∗ = −αT M − M R ∂t

(10)

74

where v ∗ = v a −

75

Comparing Eqs. 9 and 10 (or Eqs 2 and 6), we can see that the residual/mean meridional circulation

76

simply converts potential momentum to physical momentum. We can eliminate this conversion term

77

by adding either set of equations together to obtain:

is the residual meridional velocity.

     ∂ ∂ 0 0 ∂  f0 0 0 U + M + αT M − M R + αM U = vθ − u v = v0 q0 ∂t ∂z Θz ∂y

5

(11)

78

As can be seen, in this formulation the forcing by the mean meridional circulation disappears, so

79

that the only dynamical forcing is the eddy PV flux v 0 q 0 . Hence, this equation can be regarded as

80

a generalization of the barotropic momentum equation Eq. 1. The main difference is that, in this

81

case, the eddy PV flux forces what we call the total momentum, which also includes the potential

82

momentum.

83

Eq. 11 can also be combined with the conservation of pseudomomentum: ∂ (M + U ) = v 0 q 0 + Dmean ∂t ∂ A = −v 0 q 0 + Deddy ∂t

(12) (13)

84

where A ≈ q 02 /2q y (see e.g., Andrews et al. (1987)), D mean = −αT (M − M R ) − αM U is the

85

mean flow forcing and the net eddy dissipation D eddy includes frictional, thermal and small scale

86

contributions (see ZL06 for details). These equations emphasize that there can only be a wave-

87

mean flow interaction when the eddy PV flux is non-zero (Charney and Drazin, 1961). For the

88

(time-mean) equilibrated state: −Dmean = v 0 q 0 = Deddy

(14)

89

which reflects a local balance everywhere between the forcing of the mean flow, the eddy-mean

90

flow interaction and the eddy dissipation. Hence, the mean flow is only adjusted (in the sense that

91

Dmean 6= 0) over those regions with non-zero PV flux. Based on the wave-geometry interpretation of

92

baroclinic instability (Lindzen et al. (1980), Lindzen (1988)), we expect the bulk of this adjustment

93

to occur at the surface and steering level (see also section 3).

6

94

Non-locality and relation to wave propagation

95

A conceptual advantage of this framework is that the total momentum M + U is forced locally by

96

the eddy PV flux (c.f., Eq. 12). In contrast, when only U is considered the mean flow response

97

to the forcing is non-local because the residual circulation spreads the forcing vertically. However,

98

note that M , as defined in 7, is a non-local function of the basic state, which depends on the

99

full isentropic structure equatorward of the given location. Hence, the non-local character of the

100

original equations has in a sense only been hidden in the definition of the basic state. Additionally,

101

although M + U is forced locally, its partition is determined remotely. At any time, this partition

102

must be such that thermal wind balance is satisfied: ∂ ∂2M = 2 ∂y ∂z

f02 ∂U N 2 ∂z

!

(15)

103

This constraint is enforced by the residual circulation, which exchanges both forms of momentum

104

according to Eqs. 9-10.

105

Non-locality should not be surprising since baroclinic instability is fundamentally associated with

106

the interaction of propagating waves with the mean flow (Lindzen, 1988). The fact that the eddy

107

PV flux may also be written as the divergence of a wave flux (Edmon et al., 1980) implies that

108

it cannot be determined locally: indeed, it has to integrate globally to zero (Bretherton, 1966).

109

Though Eq. 14 suggests a three-way local balance between the mean flow forcing, the wave-mean

110

flow interaction and the eddy dissipation, the locality of this balance only holds for the time-mean.

111

The adjustment of the basic state that defines D mean is still non-local.

7

112

Isentropic interpretation

113

The concept of potential momentum is somewhat similar to that of available potential energy.

114

While the concept of available potential energy exploits the sloping of the isentropes, the concept

115

of potential momentum is based on the non-uniformity of the isentropic thickness. As implied by

116

its name, M represents the zonal momentum that would be realized if the basic state stratification

117

were brought to its reference value Θ z while conserving the same potential vorticity distribution.

118

This is evident from its definition: ∂ ∂M =− ∂y ∂z

119



f0 θ Θz



which also allows us to rewrite the quasigeostrophic PV gradient: qy = β − ∂yy U + ∂yz (

f0 θ) = β − ∂yy (U + M ) Θz

(16)

120

For a stratified rotating fluid, an important component of the basic state potential vorticity is built

121

into the stretching term

122

would produce a meridional wind shear

123

which would result from bringing the isentropic thickness down to its reference value (which may

124

be height-dependent) at all latitudes poleward of y 0 .

125

This is illustrated in Fig. 1, which shows two idealized distributions of the isentropic thickness for

126

a generic isentropic layer θi , together with the corresponding distribution of M . Also shown is the

127

reference thickness (dashed). When the isentropic thickness is larger than in the reference state

128

(i.e.,

129

its reference value, the flow would generate an enhanced meridional westerly shear so as to decrease

130

its absolute vorticity and maintain constant potential vorticity. The reverse is also true, so that M

131

decreases with latitude when the isentropic thickness is smaller than in the reference state.

∂θ ∂z

∂ ∂z



f0 Θz θ



. If this term were entirely converted into relative vorticity, it ∂M ∂y .

Hence, M gives the net change in zonal momentum

< 0), M must increase with latitude. In other words, if the thickness were brought down to

8

132

For the sketch shown in the left panel, the thickness increases with latitude monotonically. If the

133

reference stratification is properly defined, this implies that the thickness must be smaller than

134

the reference thickness for the first half of the profile, and larger for the second. Consequently, M

135

initially decreases with latitude, and increases near the end. Taking into account that M (y 0 ) = 0,

136

this produces an M profile as shown. Moreover, when the mean stratification agrees with the

137

reference stratification

138

The reverse is true for the case shown in the right panel, for which the thickness decreases mono-

139

tonically with latitude. Hence, we can see that an easterly (westerly) potential momentum M is

140

associated with isentropic layers that open up (close down) with latitude, as represented schemat-

141

ically in Fig. 1. When the dependence of

142

Because of the boundary conditions M (y 0 ) = M (yL ) = 0, negative (positive) midlatitude values of

143

M are typically associated with positive (negative) values of M yy . From equation 16, this implies

144

that the M contribution to the PV gradient is negative (positive) when M is easterly (westerly). In

145

the troposphere, the slope of the isentropes increases with height (Stone and Nemet, 1996), which

146

implies that the isentropic layers open up with latitude and M is easterly. This gives a negative

147

contribution to the interior PV gradient, as discussed by Zurita and Lindzen (2001).

148

We can also interpret the generation of M in terms of the isentropic view put forward above. Eq.

149

10 shows that potential momentum can only be generated diabatically or converted adiabatically

150

from/into U . In the interior, this adiabatic conversion occurs through a poleward residual circula-

151

tion, that thins (thickens) the isentropic layers at low (high) latitudes. According to Fig. 1, this

152

generates easterly potential momentum, at the expense of an equal westerly U acceleration via the

153

Coriolis force. The opposite conversion occurs at the surface, where the return flow is equatorward.

154

The diabatic generation can be interpreted in a similar manner. Differential heating adds mass

155

to the warmer isentropic layers at low latitudes, and to the colder layers at high latitudes. For a

∂θ ∂z

has zero mean and M (yL ) = 0 as well at the right endpoint yL .

∂θ ∂z

on latitude is linear, M is quadratic in y.

9

156

stably stratified fluid, this implies that the lower troposphere isentropic layers are thinned at low

157

latitudes and thickened at high latitudes, whereas the opposite is true for the upper tropospheric

158

layers. Based on Fig. 1, this can be interpreted as a generation of easterly (westerly) potential

159

momentum at lower (upper) levels.

160

Surface potential momentum

161

Following Bretherton (1966), we can substitute the surface boundary condition for an isothermal

162

surface underneath a temperature jump. This temperature jump produces a delta-function poten-

163

tial momentum jet, as can be easily derived from Eq. 7: M (y, z) = −

f0 Θz

Z

y

θ δ(z) dy 0

at z = 0

(17)

0

164

where the temperature of the isothermal surface has again been subtracted. Since this temperature

165

can be arbitrarily chosen, we choose it to be the mean surface temperature between y 0 and yL ,

166

which makes the surface M vanish at both endpoints. Eq. 17 implies that a negative temperature

167

gradient as observed is equivalent to an easterly potential momentum reservoir. This also gives a

168

negative delta-function contribution to the PV gradient (c.f., Eq. 16).

169

An important consequence of the non-local definition of M is that its value may be sensitive to

170

the choice of endpoints y0 , yL . In the interior, this dependence disappears as long as y 0 is chosen

171

outside the adjusted region (assuming that the unperturbed stratification equals the reference

172

stratification, which is reasonable in qg theory). However, this is not the case at the surface because

173

the surface temperature at large positive and negative y is generally different from the mean surface

174

temperature, even if the far-range temperature gradient vanishes. As a result, the surface value of M

175

is ambiguous. This is discussed extensively by ZL06, who show that the ambiguity disappears when

176

the potential momentum correction M − M R is considered and y0 is chosen outside the adjusted 10

177

region. In other words, the choice of y 0 carries an implicit assumption on the meridional extent of

178

the domain over which the thermal field can be rearranged. ZL06 show that, in many instances,

179

differences in the surface potential momentum correction are due to the meridional expansion of

180

the adjusted domain. For this reason, they argue that potential momentum gives a more complete

181

measure of the eddy source than local instability measures, such as baroclinicity.

182

Another important result of ZL06 is that the vertically integrated potential momentum is conserved: Z



M dz = 0

Z

∞ 0

M R dz

(18)

183

where MR is the radiative equilibrium potential momentum, and the integrals include generalized

184

delta-function contributions at both boundaries. Moreover, each of the terms in Eq. 10 integrate

185

independently to zero. The easterly generation of interior M by the residual circulation is accompa-

186

nied by an equal generation of westerly M at the surface, associated with the return flow. Likewise,

187

the diabatic generation of easterly potential momentum at lower levels is compensated by an equal

188

generation of westerly M aloft.

189

General circulation

190

The general circulation can be described in terms of potential momentum as follows.

191

1. Diabatic processes generate easterly potential momentum at lower levels, and compensating

192

westerly potential momentum aloft. In particular, they force the surface temperature gradi-

193

ent, which can be interpreted as a reservoir of easterly potential momentum at the ground.

194

2. The surface easterly potential momentum is equivalent to a negative delta-function PV gradi-

195

ent which is needed for the existence of unstable baroclinic modes (Charney and Stern (1962), 11

196

Lindzen (1988)). The equilibration of the baroclinic waves produces downgradient PV fluxes,

197

thus positive (negative) PV fluxes at the surface (in the interior).

198

3. These eddy PV fluxes are balanced by a direct circulation v ∗ . The poleward interior branch

199

transforms (westerly) potential into physical momentum. The delta-function return flow de-

200

pletes the easterly reservoir of potential momentum/reduces the surface temperature variance.

201

4. The system reaches an equilibrium when the diabatic generation of westerly potential mo-

202

mentum in the interior is balanced by the easterly drag, and both are equal to the rate of

203

conversion from M into U by the poleward residual circulation.

204

In the time-mean, Eq. 14 represents a local balance everywhere between the restoration of the

205

mean flow, the eddy dissipation, and the wave-mean flow interaction. Although this is a powerful

206

constraint, in many instances the global integral of these quantities might be of more relevance.

207

For instance, a global value of Dmean provides an integral measure of the mean flow adjustment,

208

which would seem more relevant than its value at any given point. However, the main difficulty

209

for constructing global balances is the fact that the quantities in Eq. 14 are not sign-definite. In

210

particular, it is well known (Bretherton, 1966) that v 0 q 0 integrates globally to zero; Eq. 14 implies

211

that so must Dmean and Deddy in equilibrium. As a result, the global integral of Eq. 14 trivially

212

vanishes.

213

For this reason, it is more meaningful to consider the integral of these terms over the regions where

214

they are positive (which also equals minus the integral over the region with negative values, or

215

half the global integral of the absolute values). This is not merely a mathematical convenience,

216

but also physically meaningful. The association of the eddy PV flux with the divergence of a wave

217

flux v 0 q 0 = ∇ · F (Edmon et al., 1980) allows one to associate the regions with positive/negative

218

PV fluxes with a wave source/sink. Thus, the positive integral may be regarded as the net wave 12

219

source, which must also equal the net wave sink. Assuming that the PV fluxes are downgradient in

220

the time-mean, the separation between the regions with positive/negative PV fluxes is in practice

221

equivalent to the separation between the regions with negative/positive PV gradients, or the surface

222

and the interior troposphere (Held, 1999).

223

Thus, we define the intensity of the global circulation (following ZL06): C=−

Z Z

sur

Dmean =

Z Z

sur

αT (M − M R ) =

Z Z

int

Dmean = −

Z Z

v0 q0 = − int

Z Z

int

Deddy = ... (19)

224

where ’sur’ and ’int’ stand for the surface 1 and interior regions respectively. These integrals are

225

also equal to the net mass flux

226

As discussed above, we can associate the surface integral of D mean with the net wave source. On

227

the other hand, the potential momentum depletion:

RR

f0 v ∗ over each region (c.f., Eq. 10).

Z=−

Z Z

sur

(M − M R ) = C/αT

228

can be regarded as a measure of the thermal homogenization at the surface, i.e., how much the

229

surface temperature variance is reduced from radiative equilibrium. Eq. 19 shows that this term

230

is also related to the net eddy dissipation in the interior, which suggests that enhanced interior

231

dissipation might lead in some cases to enhanced thermal homogenization at the surface. Section

232

3 confirms this for a simple case. 1

For the argument discussed here, ’surface’ refers to the massless layer of qg theory (the delta-function). In an

actual multilayer/numerical model the surface would have a finite thickness, so that frictional contributions would also be lumped into the surface Dmean . For example, see ZL06 for the two-layer case.

13

233

3.

Some simple examples

234

Short Charney waves: a 3D view

235

In this section we illustrate the concepts introduced above for a very simple problem: the equili-

236

bration of short Charney waves. These are modes for which the net PV gradient in the interior is

237

smaller than the integrated delta function at the surface (Held (1978), Zurita and Lindzen (2001)).

238

Because these modes do not grow as fast as the most unstable mode, the latter would always

239

dominate in a seminfinite fluid. However, short Charney modes may still be relevant when longer

240

modes are prevented by the geometry, for instance if the meridional scale imposed by the jet con-

241

strains the depth of the modes (Lindzen, 1993). Zurita and Lindzen (2001) express the scale of

242

the modes in terms of the dimensionless parameter H/h, where H is the half Rossby depth and

243

h=

244

the ratio between the interior and boundary PV gradient. Short Charney modes are characterized

245

by H/h < 3.9, which is the value corresponding to the most unstable mode.

246

When the interior PV gradient is small, the condition that the PV fluxes integrate vertically to

247

zero requires a large interior diffusivity. Zurita and Lindzen (2001) show that this implies that

248

the modal PV flux must be strongly-peaked at the steering level because the linear diffusivity is

249

bounded away from it (Bretherton, 1966). This is also consistent with the generalized instability

250

mechanism of Lindzen (1988) in which eddy growth is due to stimulated wave emission by means of

251

the Kelvin-Orr mechanism in the neighborhood of the steering level (or by wave over-reflection as it

252

is referred to in Lindzen et al. (1980)). For weakly growing modes, this implies that the wave-mean

253

flow interaction is limited to a narrow region surrounding the steering level. However, the region

254

with non-zero PV flux broadens as ci increases (Lindzen et al., 1980), which may be interpreted in

255

terms of the time-dependent forcing at the steering level (Charney and Pedlosky, 1963).

2 1 f0 dU β N 2 dz



z=0+

is the Held scale (Held, 1978). This parameter may be regarded as a measure of

14

256

Consistent with these ideas, Zurita-Gotor and Lindzen (2004a) show that, when only short Char-

257

ney modes are allowed, the inviscid 2D problem equilibrates by eliminating the PV gradient at

258

the steering level alone, while still keeping a negative PV gradient at equilibration. The authors

259

speculate that this might explain why the surface temperature gradient is not eliminated, and

260

the tropospheric PV gradient vanishes at the steering level alone, provided that some external

261

mechanism like the jet width constrains the scale of the modes. However, ZL04 find that these

262

constraints do not hold in the 3D case. The reason is that in the 3D problem momentum is not

263

only redistributed vertically, but also meridionally. In other words, the net interior PV flux is not

264

constrained by a small interior PV gradient aloft as in the 2D case because the waves can also

265

propagate meridionally to be absorbed at other latitudes.

266

In practice, ZL04 find that meridional propagation is most important during an initial adjustment

267

stage. As the waves propagate meridionally and the convergent momentum fluxes accelerate the

268

jet, the flow develops a large horizontal curvature. This curvature contributes to the positive PV

269

gradient above the surface (or more generally above the region with negative potential vorticity

270

gradient), so that the mode no longer behaves as a short Charney wave. As a result, the PV flux

271

broadens around the steering level and the surface vertical shear is eliminated. ZL04 find that this

272

process is largely controlled by friction. With strong surface friction, the flow cannot develop the

273

required meridional curvature and the wave remains ’short’. Thus, PV is only homogenized locally

274

at the steering level, as in the 2D case. ZL04 explain these results noting that, according to the

275

TEM framework, the equilibration can be interpreted in terms of the export of easterly momentum

276

away from the surface. In a purely baroclinic adjustment all this easterly momentum is deposited

277

aloft along the column, but in the 3D problem the lateral export is also important.

278

This is most transparent in the potential momentum framework since Eq. 19 shows that thermal

279

homogenization at the surface, as measured by the potential momentum, is related to the net

15

280

interior PV flux, not just the along-column baroclinic component. We next apply these diagnostics

281

to the same sets of runs discussed by ZL04. The model used is a qg model forced by linear

282

relaxation to a ‘radiative equilibrium’ Charney-like basic state that has constant vertical shear

283

with height. The radiative equilibrium zonal wind is meridionally modulated by a Gaussian jet

284

with halfwidth σ = 2000 km and has a maximum vertical shear Λ(0) = 33 ms −1 /10 km at the

285

jet center. The model has rigid lids at the top and meridionally, but at enough distance to be

286

considered in practice unbounded. A small channel length (L = 4000 km) is chosen to prevent the

287

dominance of the most unstable mode and to allow only short Charney waves. The dimensionless

288

scale of the first harmonic, measured by H/h, is changed through β as described in ZL04. All

289

simulations are run for 800 days starting from radiative equilibrium, with averages calculated over

290

the last 400 days. Additional details about the model are given in ZL04.

291

Fig. 2 shows some diagnostics for a run with H/h = 1.6 and frictional timescale 1 day. Since the

292

vertical shear is constant with height (i.e., θ y 6= f (z), or ∂y θz = 0), the radiative equilibrium profile

293

has zero potential momentum in the interior. In contrast, panel A shows that the equilibrated

294

flow has developed, at equilibration, an easterly jet of potential momentum. Consistent with the

295

isentropic view put forward in section 2, panel B shows that the easterly potential momentum is

296

associated with isentropes that open up with latitude. However, note that the (scaled) isentropic

297

slope is large and that the changes in thickness at constant height are larger than following the

298

actual isentropes. It is noteworthy that the potential momentum jet of panel A is also comparable

299

to the actual jet of zonal momentum, shown in panel C. This is as required by the thermal wind

300

constraint (Eq. 15). Note that this also implies (panel D) that the PV gradients associated to the

301

horizontal curvature of the jet and stretching term (i.e., the potential momentum curvature, Eq.

302

16) are comparable.

303

Fig. 3A shows the (delta-integrated) surface potential momentum at equilibration (thin, solid) and

16

304

in radiative equilibrium (thin, dashed). Also shown is the vertically-integrated potential momen-

305

tum in the interior from Fig. 2A (thick, solid), which agrees exactly with the difference between

306

the two previous curves (thick, dashed). This is what should be expected, as the interior potential

307

momentum is initially zero and Eq. 18 implies that the vertically-integrated potential momen-

308

tum is conserved. The baroclinic equilibration can thus be interpreted in terms of the vertical

309

redistribution of the easterly M originally locked at the surface.

310

As discussed in section 2, the actual values of potential momentum at the surface depend on the

311

choice of endpoints y0 , yL , and are therefore ambiguous. This is illustrated in panel B of Fig. 3,

312

which shows the same curves of panel A for a different choice of the refefence latitude y 0 . As can be

313

seen, this gives different values of the surface M , both at equilibration and in radiative equilibrium.

314

Nevertheless, the difference between both curves should be the same, and equal to the integrated

315

M in the interior, as long as y0 is chosen within the unperturbed region (the small differences

316

between both panels are due to the violation of this condition for panel B). The main difference

317

between panels A and B is merely one of interpretation. When choosing the reference latitude,

318

one is implicitly assuming the size of the domain over which temperature can be rearranged, which

319

defines the maximum intensity of the circulation. Though the actual circulation C = α T (M − M R )

320

is the same with both choices of y0 , the theoretical maximum αT M R corresponding to full potential

321

momentum depletion is larger with the broader domain in panel A. The waves ’appear less efficient’

322

in reducing the surface potential momentum in that case because they homogenize temperature

323

over a narrower domain than anticipated. ZL06 show some examples in which differences in the

324

degree of thermal homogenization are due to differences in the meridional extent of the adjusted

325

domain, rather than to differences in the maximum baroclinicity.

326

Note that the interior potential momentum jet in Fig. 2 has a well-defined peak at the steering

327

level. This is not surprising since Eq. 14 implies that the potential momentum restoration must

17

328

equal the eddy PV flux locally. Because friction is sufficiently large, the mode still behaves like

329

a short 2D Charney wave and the wave-mean flow interaction/mean flow adjustment has a lot of

330

structure at the steering level. In contrast, Fig. 4 shows the potential momentum diagnostics for a

331

case with reduced surface friction (5 days). As can be seen, in this case the potential momentum

332

jet (and also the PV flux, not shown) have a much broader structure. This occurs as the flow

333

develops a strong barotropic jet (panel B) and a large horizontal curvature: Panel C shows that

334

this curvature produces a positive PV gradient (dashed line) that is nearly 3 times as large as

335

before. As a result, the mode no longer behaves as a short Charney mode: the PV flux is broader

336

and the interior jet of potential momentum jet is also stronger. Since the vertically integrated M

337

is conserved, this also implies stronger potential momentum depletion/thermal homogenization at

338

the surface (panel D).

339

Interior eddy dissipation and surface thermal homogenization

340

We showed above that for short Charney waves the degree of thermal homogenization at the surface

341

is controlled by the interior eddy absorption. When friction is weak, thermal homogenization

342

increases because the reinforced meridional curvature supports enhanced interior absorption. To

343

further stress this point, we next describe the sentivity of the potential momentum depletion M −

344

MR on the interior absorption. With this purpose, we have designed a series of idealized experiments

345

in which artificial wave dissipation is added in the interior. We simply add a term of the form ∂q 0 q0 = − + ... ∂t τ

346

to the non-zonal part of the flow. Thus, the waves are damped but the forcing of the basic state is

347

unchanged. All the remaining parameters are chosen as in Fig. 2.

348

From a linear point of view, this damping term should reduce the eddy amplitudes and fluxes, due 18

349

to its stabilizing effect. Indeed, in the limit of large damping, the radiative equilibrium solution

350

would be stable. Fig. 5A shows the linear growth rate as a function of τ . The linear growth rate

351

displays sensitivity to τ only when this damping is faster than the diabatic timescale; however,

352

even then the effect is moderate unless τ goes below 5 days or so.

353

On the other hand, the arguments of previous sections suggest that adding a wave sink in the

354

interior could actually lead to a stronger circulation. Indeed, in the strict non-dissipative limit

355

the non-acceleration theorem tells us that there cannot be a mean circulation. Fig. 5B shows the

356

dependence of the circulation (c.f., Eq. 19) on the wave damping timescale. As can be seen, the

357

circulation increases for nearly all timescales and reaches a maximum for τ ≈ 5 days, beyond which

358

value the direct damping effect presumably becomes important. The wave sink effect is clearly

359

significant: the circulation increases by as much as a 55% for τ = 5 days, and even for τ = 1 day it

360

is still a 45% stronger than in the control run. We have also performed experiments in which the

361

linear damping is added at the lower boundary alone. In that case the circulation always decreases

362

with the damping (not shown).

363

The increase in the circulation with increased damping may be surprising at first. Does this imply

364

that the eddy amplitude also increases with damping? As shown in panel C of Fig. 5, the answer

365

depends on the norm. The net interior eddy PV flux must of course increase to balance the

366

enhanced Deddy . However, this increase is mainly due to enhanced correlation between velocity

367

and PV perturbations. Although there is a weak eddy enstrophy maximum for τ = 5 days, the

368

meridional velocity variance always decreases with the damping.

369

Finally, Fig. 6 shows the structure of the interior eddy PV flux for the cases with no damping,

370

with interior damping with timescale τ = 5, 15 days and with boundary damping τ = 15 days.

371

Interior damping makes the PV flux less structured, as if the wave were deeper, whereas boundary

372

damping has the opposite effect and produces a shallower PV flux at the steering level. These 19

373

results suggest that it is the ratio of boundary to interior absorption, rather than between their

374

respective PV gradients, that defines whether a wave is short or long. Both would be equivalent

375

with uniform timescales, as in the standard linear problem.

376

The strong enhancement of the circulation as the interior absorption is enhanced should not be

377

surprising, as the defining characteristic of a short Charney wave is precisely the relative weakness

378

of the interior PV fluxes. However, this is not a general result: for instance, in the two-layer runs

379

of ZL06 the circulation always decreases with interior damping 2 . Although Eq. 19 implies that

380

increasing the net interior eddy dissipation |D eddy | should lead to enhanced thermal homogeniza-

381

tion, it is not obvious that artificially reducing the dissipative timescale somewhere necessarily

382

results in enhanced net absorption. In particular, the possibility cannot be discarded that the

383

local enhancement in the wave sink occurs at the expense of weakening other sinks, rather than

384

through an enhancement of the wave source. A specific example of how damping contributes to

385

the destabilization of baroclinic modes is given in Snyder and Lindzen (1988).

386

4.

387

In this paper we have discussed the equilibration of a forced-dissipative baroclinic system using

388

the generalized momentum framework recently derived by Zurita-Gotor and Lindzen (2006a). In

389

this framework, the thermal structure is recast in a momentum form using the concept of potential

390

momentum. The potential momentum is simply the zonal momentum profile that produces the

391

same PV distribution as the stretching term. When the total momentum is considered, the eddy

392

PV flux is the only forcing of the mean flow, which may be regarded as the mean flow counterpart to

Conclusion

2

Note that in the two-layer model the positive PV gradient is always larger than the negative PV gradient, due

to the presence of an upper rigid lid.

20

393

pseudomomentum conservation. On the other hand, the partition between potential and physical

394

momentum is subject to the thermal wind constraint. This is enforced by the residual circulation,

395

which exchanges potential and physical momentum without changing the total momentum.

396

We have also interpreted the equilibration of short Charney waves, previously discussed by Zurita-

397

Gotor and Lindzen (2004 a, b), using these concepts. In the new framework, the thermal ho-

398

mogenization at the surface is measured in terms of the potential momentum depletion. Unlike

399

baroclinicity, this measure also accounts for changes in the meridional domain over which surface

400

temperature is rearranged. Since the potential momentum depletion increases with eddy absorp-

401

tion everywhere, not just aloft the baroclinic zone, the along-column PV gradient does not limit

402

thermal homogenization as in the 2D problem. Likewise, we found that interior eddy dissipation

403

leads to enhanced thermal homogenization at the surface for a short Charney mode, due to the

404

increase in the interior PV fluxes.

405

The framework considered here is very idealized. Besides the QG limitation, the assumption that a

406

quiescent latitude exists (or equivalently, that the domain is meridionally unbounded) is obviously

407

unrealistic for the Earth’s atmosphere. A more common setup in the QG context involves the use of

408

close meridional walls. As discussed by Zurita-Gotor (2006), this could have an important impact

409

on the equilibrated system by constraining the inverse cascade. Neither setup is very realistic, but

410

we believe that the cleaner unbounded setup discussed here might be a useful reference paradigm for

411

understanding the atmospheric case. One can easily overcome this limitation by including boundary

412

terms at the reference latitude, an approach that could be useful for studying the interaction

413

between the Hadley cell and the extratropical eddies within the qg framework. We intend to follow

414

this approach in a subsequent study.

415

To conclude, we note that the total momentum equation (Eq. 11) is simply the y-integral of the

416

zonal-mean quasigeostrophic potential vorticity equation. For symmetric 2D flow, momentum pro21

417

vides a full description and vorticity is redundant. Likewise, symmetric 3D flow can be encapsulated

418

in terms of momentum alone, provided that the baroclinic contribution to the potential vorticity

419

is also rewritten as a momentum term. While the present momentum formulation offers a certain

420

conceptual elegance, it remains to be seen whether it offers serious practical advantages. At the

421

very least, however, the explicit partition of PV between M and U allows one to better distinguish

422

and quantify the roles of vertical and meridional propagation.

423

ACKNOWLEDGMENTS: P.Z-G was supported by the Visiting Scientist Program at the NOAA

424

Geophysical Fluid Dynamics Laboratory, administered by the University Corporation for Atmo-

425

spheric research. R.S.L. acknowledges financial support by NSF Grant ATM-9421195 and DOE

426

Grant FG02-93ER61673.

22

427

References

428

Andrews, D. G., J. R. Holton, and C. B. Leovy: 1987, Middle Atmospheric Dynamics. Academic

429

430

431

432

433

434

435

436

437

438

439

440

441

Press, 489 pp, first edition. Bretherton, F. P.: 1966, Critical layer instability in baroclinic flows. Q.J.Meteor.R.Soc., 92, 325– 334. Charney, J. G. and P. G. Drazin: 1961, Propagation of planetary scale disturbances from the lower into the upper atmosphere. J.Geophys.Res., 66, 83–110. Charney, J. G. and J. Pedlosky: 1963, On the trapping of unstable planetary waves on the atmosphere. J.Geophys.Res., 68, 6441–6442. Charney, J. G. and M. E. Stern: 1962, On the stability of internal baroclinic jets in a rotating atmosphere. J.Atmos.Sci., 19, 159–172. Edmon, H. J., B. J. Hoskins, and M. E. McIntyre: 1980, Eliassen-Palm cross-sections for the troposphere. J.Atmos.Sci, 37, 2600–2616. Held, I. M.: 1978, The vertical scale of an unstable baroclinic wave and its importance for eddy heat flux parameterizations. J.Atmos.Sci, 35, 572–576.

442

— 1999, The macroturbulence of the troposphere. Tellus, 51AB, 59–70.

443

Lindzen, R. S.: 1988, Instability of plane parallel shear flow (toward a mechanistic picture of how

444

it works). PAGEOPH , 126, 103–121.

445

— 1993, Baroclinic neutrality and the tropopause. J.Atmos.Sci, 50, 1148–1151.

446

Lindzen, R. S., B. F. Farrell, and K. K. Tung: 1980, The concept of wave overreflection and its

447

application to baroclinic instability. J.Atmos.Sci, 37, 44–63. 23

448

449

Lindzen, R. S., A. J. Rosenthal, and B. Farrell: 1983, Charney’s problem for baroclinic instability applied to barotropic instability. J.Atmos.Sci, 40, 1029–1034.

450

Snyder, C. and R. S. Lindzen: 1988, Upper level baroclinic instability. J.Atmos.Sci, 45, 2446–2459.

451

Stone, P. H. and B. Nemet: 1996, Baroclinic adjustment: a comparison between theory, observa-

452

453

454

455

456

457

458

459

460

461

462

tions, and models. J.Atmos.Sci., 53, 1663–1674. Zurita, P. and R. S. Lindzen: 2001, The equilibration of short Charney waves: implications for PV homogenization in the extratropical troposphere. J.Atmos.Sci, 58, 3443–3462. Zurita-Gotor, P.: 2006, The relation between baroclinic adjustment and turbulent diffusion in the two-layer model. Submitted, J.Atmos.Sci. Zurita-Gotor, P. and R. S. Lindzen: 2004a, Baroclinic equilibration and the maintenance of the momentum balance. Part I: a barotropic analog. J.Atmos.Sci, 61, 1469–1482. — 2004b, Baroclinic equilibration and the maintenance of the momentum balance. Part II: 3D results. J.Atmos.Sci, 61, 1483–1499. — 2006a, A generalized momentum framework for looking at baroclinic circulations. Accepted, J.Atmos.Sci.

463

— 2006b, Theories of baroclinic adjustment and eddy equilibration, The global circulation of the

464

atmosphere: phenomena, theory, challenge. T. Schneider and A. H. Sobel eds. Princeton UP.

24

465

466

List of Figures 1

whose thickness: (left) increases with latitude, (right) decreases with latitude. . . . . . . . .

467

468

Sketch illustrating the potential momentum distribution associated to an isentropic layer

2

Potential Momentum Diagnostics for the run with H/h = 1.6 and frictional timescale 1

469

day. (A) Interior potential momentum in m/s; (B) Percentage change in the equilibrium

470

stratification with respect to the reference stratification Θz ; (C) Zonal wind in m/s; and

471

(D) Vertically-averaged interior PV gradient contributions by the M (solid) and U (dashed)

472

curvatures (normalized by β). The isentropes are shown dashed for reference in panels A-C.

473

3

26

27

(A) Surface potential momentum at equilibration (thin, solid) and in radiative equilibrium

474

(thin, dashed) for the run with H/h = 1.6 and frictional timescale 1 day. Also shown are

475

their difference (thick, dashed) and the integrated potential momentum in the interior (thick,

476

solid). (B) As A, but for a different choice of the reference latitude y0 . Units are in m2 /s. .

28 29

477

4

As panels A, C, D in Fig. 2 and panel A in Fig. 3, but with frictional timescale 5 days. . . .

478

5

As a function of the interior eddy damping timescale τ (A) Growth rate, (B) Net Circulation

479

(c.f., Eq. 19) and (C) Eddy amplitude in the indicated norms. The solutions in panels B

480

and C are normalized with the τ = ∞ solution. . . . . . . . . . . . . . . . . . . . . . . .

481

482

6

30

Interior eddy PV flux for the interior and boundary eddy damping timescales indicated. Units are in 10−5 s−1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

25

31

θi

θi

Θz

yL

y0

yL

y0

M<0

Θz

M>0

Figure 1: Sketch illustrating the potential momentum distribution associated to an isentropic layer whose thickness: (left) increases with latitude, (right) decreases with latitude.

26

Potential Momentum

Stratification Adjustment

−8

z

z

−32

−16

24 32

−40

−32 −24

16

−24 8 −8

−16

A

B

y

y

Zonal Wind

PV Gradient

30

0.8 0.6

25

0.4 20

Myy, Uyy

0.2

z

15 10 5

0 −0.2 −0.4 −0.6 −0.8

5

−1

C

−1.2

y

M U

D y

Figure 2: Potential Momentum Diagnostics for the run with H/h = 1.6 and frictional timescale 1 day. (A) Interior potential momentum in m/s; (B) Percentage change in the equilibrium stratification with respect to the reference stratification Θz ; (C) Zonal wind in m/s; and (D) Vertically-averaged interior PV gradient contributions by the M (solid) and U (dashed) curvatures (normalized by β). The isentropes are shown dashed for reference in panels A-C.

27

5

2

5

x 10

2

x 10

1 1 0 0

−2

∫ M dz

∫ M dz

−1

−3 −4

−1

−2

−5 −3 −6

A −7

B −4

y

y

Figure 3: (A) Surface potential momentum at equilibration (thin, solid) and in radiative equilibrium (thin, dashed) for the run with H/h = 1.6 and frictional timescale 1 day. Also shown are their difference (thick, dashed) and the integrated potential momentum in the interior (thick, solid). (B) As A, but for a different choice of the reference latitude y0 . Units are in m2 /s.

28

Potential Momentum

Zonal Wind

A

10

5B

40

5

35

30

−16 −24

z

z

−32

25 20

−40

10 15

1520

−48 −8 −56 −8 −56 −64

−5

−5

y

y

PV Gradient

5

2

1 M U

C

x 10

Surface M

D 0

1.5

−1 1 −2 0.5

∫ M dz

yy

M , U

yy

−3

0

−4 −5

−0.5

−6 −1 −7 −1.5 −2

−8 −9

y

y

Figure 4: As panels A, C, D in Fig. 2 and panel A in Fig. 3, but with frictional timescale 5 days.

29

0.44

1.6

1.7

B

A τ=∞

0.435

1.6

1.4

1.5

1.2

Eddy amplitude

Circulation

Growth rate

0.43 1.4

0.425

C

v’q’

1.3

q’2

τ=∞

1

0.8

v’2

0.42

0.415

0.41

0

50

τ (days)

100

1.2

0.6

1.1

0.4

1

0

50

τ (days)

100

0.2

0

50

τ (days)

100

Figure 5: As a function of the interior eddy damping timescale τ (A) Growth rate, (B) Net Circulation (c.f., Eq. 19) and (C) Eddy amplitude in the indicated norms. The solutions in panels B and C are normalized with the τ = ∞ solution.

30

No damping

15−day interior damping A

B

−0.5

−1

−1

z

z

−0.5

−1.5 −3

−1.5 −3 −3.5

−2.5

−2

−2.5 −2

y

y

5−day interior damping

15−day boundary damping

C

D

−0.5 −1

z

z

−0.5 −2 −1 −3.5

−2.5

−1.5

−1.5

−2

−2.5 −3

y

y

Figure 6: Interior eddy PV flux for the interior and boundary eddy damping timescales indicated. Units are in 10−5 s−1

31

Potential Momentum and the equilibration of short ...

where αM is a frictional time scale. .... and the net eddy dissipation Deddy includes frictional, thermal and small scale ...... In the new framework, the thermal ho-.

235KB Sizes 0 Downloads 321 Views

Recommend Documents

The Equilibration of Short Charney Waves: Implications ...
Nov 15, 2001 - (Manuscript received 19 December 2000, in final form 30 April 2001) ..... zen 1990). Now, if we use Eqs. (4), (5), and the definition of h: H. 2. 2 hc |(0 .... Note that this definition is meaningful only .... tral solutions, as proved

Baroclinic Equilibration and the Maintenance of the ...
Jul 1, 2004 - Sketch illustrating the meaning of the different vertical scales for a short ..... by symmetry, this occurs at the expense of negative PV gradients in.

Short-Term Momentum and Long-Term Reversal in ...
finite unions of the sets C(st). The σ-algebras Ft define a filtration F0 ⊂ ... ⊂ Ft ⊂ . ..... good and a price system is given by q ≡ {q1 t , ..., qK t. }∞ t=0 . Agent i faces a state contingent solvency constraint, B ξ i,t(s), that limi

On the equilibration of a symmetrically unstable front ...
Cambridge, MA, 02139, USA. Email: [email protected] ... vertical fluxes induced by the. K-H instability were unable to change the bulk properties of the system.

Momentum, Impulse and Momentum Change.pdf
Determine the impulse (I), momentum change (Δp), momentum (p) and other values. A 7-ball collides with the 8-ball. A moving medicine ball is caught by a girl on ice skates. A car is at rest when it experiences a forward propulsion force to set it in

DENDROCLIMATIC POTENTIAL OF EARLYWOOD AND ...
DENDROCLIMATIC POTENTIAL OF EARLYWOOD AND LATEWOOD.pdf. DENDROCLIMATIC POTENTIAL OF EARLYWOOD AND LATEWOOD.pdf. Open.

On the equilibration of a symmetrically unstable front ...
density gradient in the x-direction is in 'thermal wind' balance with a velocity ..... The transfer of energy from the unstable symmetric modes to smaller scales via.

Mobilizing the Potential of Rural and Agricultural Extension - Food and ...
Extension includes technical knowledge and involves facilitation, brokering and ..... agricultural high schools) is an important component in efforts to enhance their ...... Increasing rural employment and incomes to make food more affordable.

Mobilizing the Potential of Rural and Agricultural Extension - Food and ...
extension in the coming years, there are multiple reasons to why hundreds of millions ... 1 World Bank and FAO, 2009, Awakening Africa's Sleeping Giant: Prospects for ... Extension includes technical knowledge and involves facilitation, brokering ...

Universality in the equilibration of quantum systems ...
Mar 11, 2010 - 2Department of Physics and Astronomy and Center for Quantum Information Science & Technology, University of Southern California,.

Universality in the equilibration of quantum ... - APS Link Manager
Mar 11, 2010 - 2Department of Physics and Astronomy and Center for Quantum Information Science & Technology, University of Southern California,.

The Momentum of Colliding Objects
Lab Preview. Directions: Answer these questions before you begin the Activity. 1. ... could an object that size change Earth's climate enough to cause the extinc-.

Derivation of the energy–momentum and Klein–Gordon ...
In a previous note, we have provided a formal derivation of the transverse Doppler shift of special relativity from the generalization of El Naschie's complex time. Here, we show that the relativistic energy–momentum equation, and hence the Kleinâ€

The potential and limitations of using carbon dioxide - Royal Society
A number of companies are already exploring these areas. It is likely that research currently underway will ... BEIS (Department for Business Energy & Industrial Strategy). 2015 Data ... products; and from approximately 1800Mt9 to 2000Mt10 of ...

Market Potential and the Rise of US Productivity ...
Apr 4, 2013 - data, gave us the green light to use some of it, and warned us against potential mis-use of various data series. Special ... government…it is free unrestricted trade in everything under the same ...... Netherlands (very open, not so r

pdf-1832\geology-and-hydrocarbon-potential-of-the-south-china ...
... the apps below to open or edit this item. pdf-1832\geology-and-hydrocarbon-potential-of-the-so ... f-joint-research-and-development-workshop-report.pdf.

Care of the Potential Organ Donor - dunkanesthesia
Dec 23, 2004 - quire hospitals to notify their local organ-procurement organization in a timely manner .... In all potential donors, hemodynamic management be- gins with an evaluation of ... guide the administration of vasoactive medications,.

pdf-08107\contesting-community-the-limits-and-potential-of-local ...
... apps below to open or edit this item. pdf-08107\contesting-community-the-limits-and-potential- ... es-defilippis-professor-robert-fisher-professor-eric.pdf.

THE POTENTIAL LIABILITY OF FEDERAL LAW-ENFORCEMENT ...
Under the doctrine formulated in Bivens v. ... 1 United States v. Archer, 486 ... to describe such people); see also Ashcroft v. Free Speech Coalition, 122 S.Ct. 1389, 1399 ... CHILD PORNOGRAPHY INVESTIGATIONS, HOWARD ANGLIN.pdf.

Care of the Potential Organ Donor - dunkanesthesia
Dec 23, 2004 - quire hospitals to notify their local organ-procurement organization in a timely manner .... In all potential donors, hemodynamic management be- gins with an evaluation of ... guide the administration of vasoactive medications,.

The elusive chemical potential
helium atoms are in thermal equilibrium at temperature T; we treat them as forming an ideal gas. What value should we anticipate for the number Nu of atoms in the upper volume, especially in comparison with the number Nl in the lower volume? We need