AMERICAN METEOROLOGICAL SOCIETY Journal of the Atmospheric Sciences EARLY ONLINE RELEASE This is a preliminary PDF of the author-produced manuscript that has been peer-reviewed and accepted for publication. Since it is being posted so soon after acceptance, it has not yet been copyedited, formatted, or processed by AMS Publications. This preliminary version of the manuscript may be downloaded, distributed, and cited, but please be aware that there will be visual differences and possibly some content differences between this version and the final published version. The DOI for this manuscript is doi: 10.1175/JAS-D-11-0115.1 The final published version of this manuscript will replace the preliminary version at the above DOI once it is available. If you would like to cite this EOR in a separate work, please use the following full citation: Røsting, B., and J. Kristjansson, 2011: The usefulness of piecewise potential vorticity inversion. J. Atmos. Sci. doi:10.1175/JAS-D-11-0115.1, in press. © 2011 American Meteorological Society LaTeX File (.tex, .sty, .cls, .bst, .bib) Common.Links.ClickHereToDownload 1 2 The usefulness of piecewise potential vorticity inversion by 3 Bjørn Røsting 4 1 2 1 and Jón Egill Kristjánsson The Norwegian Meteorological Institute, Oslo, Norway Department of Geosciences, University of Oslo, Norway 1 2 5 Abstract 6 It is today widely accepted that potential vorticity (PV) thinking is a highly 7 useful approach for understanding important aspects of dynamic meteorol- 8 ogy and for validation of output from state-of-the-art numerical weather pre- 9 diction (NWP) models. Egger (2008) recently presented a critical view on 10 piecewise potential vorticity inversion (PPVI). This was done by defining a 11 PV anomaly by retaining the observed PV field in a specific region, while 12 changing the observed PV fields to zero elsewhere. Inversion of such a modi- 13 fied PV field yields a flow vastly different from the observed. On the basis of 14 this result it was argued that piecewise potential vorticity inversion (PPVI) 15 is useless for understanding the dynamics of the flow. 16 In the present paper we argue that the results presented by Egger are 17 incomplete in the context of PPVI, since the complementary cases were not 18 considered and that the results also depend on the idealized model formula- 19 tions. The complementary case is defined by changing the observed PV to 20 zero in the specific region, while retaining the observed PV field elsewhere. 21 By including the complementary cases, we demonstrate that the stream- 22 function fields associated with the PV and boundary temperature anomalies 23 presented by Egger (2008), add up to yield the observed streamfunction field, 24 as expected if PPVI is to be valid. It follows that PPVI is indeed valid and 25 useful in these cases. 2 26 1 Introduction 27 Hoskins et al. (1985) provided a theoretical survey of the concept of potential 28 vorticity (PV), along with applications of PV and low level boundary temper- 29 ature anomalies to the interpretation of atmospheric dynamics. Though the 30 PV aspect of atmospheric dynamics had been well known for several decades, 31 being described by Rossby (1940), Ertel (1942) and Charney (1947), the pa- 32 per by Hoskins et al. introduced, most likely for the first time, the concept 33 of PV thinking. PV thinking means adopting PV in order to describe and 34 understand the dynamics of the atmosphere, such as e.g. baroclinic and 35 barotropic instability. This is achieved e.g. through PV inversion, by which 36 geopotential heights, winds and temperature fields are obtained, provided 37 that suitable balance and boundary conditions are specified. 38 The classical approach to baroclinic instability, described by Eady (1949), 39 is essentially based on the PV-concept, where the (quasi-geostrophic) PV 40 field confined between an upper lid, i.e. the tropopause, and the surface 41 are specified, along with the temperature distribution at lower and upper 42 boundaries. Analytical solutions of this problem yield growing waves with 43 the largest growth rate for wavelengths on the order of 4000 - 6000 km. 44 PV can also be applied in identifying errors in NWP analyses and simu- 45 lations (e.g. Mansfield 1996). This is achieved by comparing the PV fields 46 retrieved from numerical weather prediction (NWP) models with features 47 observed in water vapor (WV) images. This method is based on the strong 3 48 absorption of radiation in the 6.3-6.7 µm water vapor bands, hence PV- 49 rich stratospheric air is easily identified in WV images (e.g. Santurette and 50 Georgiev 2005). When errors in the PV fields are identified by an observed 51 mismatch between PV fields and WV features, it is possible to manually 52 correct the PV fields. The corrected PV fields are then inverted, yielding a 53 modified numerical analysis that is balanced and suitable for initializing nu- 54 merical reruns (e.g. Demirtas and Thorpe 1999, Manders et al. 2007, Røsting 55 et al 2003, Røsting and Kristjánsson 2006, 2008, Verkley et al. 2005). These 56 procedures represent an efficient use of PV thinking. 57 An important aspect of PV thinking is piecewise PV inversion (PPVI). 58 In PPVI, the PV field is partitioned, and the contributions of the different 59 parts of the PV field to the flow are quantified by piecewise inversion (Davis 60 1992, Kristjánsson et al 1999). The feasibility of such an approach for Ertel’s 61 nonlinear PV was first demonstrated by Davis and Emanuel (1991). 62 Criticism of PPVI pertaining to its usefulness for understanding atmo- 63 spheric dynamics was recently put forward by Egger (2008), referred to as 64 E2008 below. Though some of the arguments on PPVI in E2008 may be 65 pertinent, particularly the discussion of inverted tendency fields, others call 66 for comments. The criticism advanced in E2008 is based on experiments to 67 be discussed in section 3 below, in which the observed PV is replaced by 68 P V = 0 in a selected region. Inversion of such a modified PV field yields a 69 flow different from the observed. In E2008 it was claimed that such a result 70 implies that PPVI in general is useless for understanding the dynamics of 4 71 the flow. Below, we present evidence that this drastic claim is unjustified. 72 Methven and de Vries (2008) addressed the limitations of the mathematical 73 model describing the baroclinic cases of E2008 and pointed out the impor- 74 tance of dynamic evolution rather than pure diagnosis based on PPVI in 75 assessing the benefits of PV thinking. This paper is structured as follows: 76 77 In section 2 we give a brief presentation of the PV inversion technique used 78 in many case studies. In section 3 we address some of the cases presented in 79 E2008. A case study based on PPVI is presented in section 4, and discussion 80 and conclusions are provided in section 5. 81 2 82 83 84 Piecewise PV inversion We now recapture the PPVI technique, since knowledge of this technique is essential for the subsequent discussion. In general, studies requiring PV inversion adopt the Ertel PV : 1 q = η · ∇θ ρ (1) 85 where ρ is density, η is the absolute vorticity vector and ∇θ is the three 86 dimensional gradient of potential temperature. PV is expressed in PV-units, 87 defined as 1PVU=10−6 m2 s−1 Kkg −1 . 88 89 PV is changed due to diabatic effects and friction as shown in the following equation: 5 Dq ∂q 1 1 ≡ + u · ∇q = η · ∇θ̇ + ∇ × Fr · ∇θ Dt ∂t ρ ρ (2) 90 where u is the wind vector, Fr is the friction force and θ̇ denotes diabatic 91 heating. 92 93 The mathematical expression for the Ertel PV can be simplified by using isentropic coordinates (x,y,θ): q = (ζθ + f )(−g ∂θ ) ∂p (3) 94 where ζθ is relative vorticity on an isentropic surface and f is the coriolis 95 ∂θ is parameter (planetary vorticity), g is the acceleration of gravity and − ∂p 96 an expression of static stability. 97 98 Relations (2) and (3) show that for adiabatic and inviscid flow q is conserved for an air parcel moving on an isentropic surface. 99 Potential temperature fields (θb ) are specified at the lower and upper 100 boundaries, e.g. at 950 and 150 hPa and the θb anomalies can be regarded 101 as PV anomalies, i.e. a warm (cold) anomaly at the lower boundary can 102 be regarded as a positive (negative) PV anomaly. Likewise, a warm (cold) 103 anomaly at the upper boundary can be treated as a negative (positive) PV 104 anomaly (Hoskins et al. 1985). Hence such anomalies can be inverted sepa- 105 rately. 106 The total PV (q) and boundary potential temperature (θb ) fields are de- 107 composed into mean and perturbation terms. Hence the total PV is expressed 6 108 as: q = q + q′ 109 and the expression for the boundary potential temperature field is analogous: θb = θb + θ′ b (5) The mean fields, q and θb are temporal or spatial means, with temporal 110 111 (4) means taken as an average over e.g. 48 hours. 112 The perturbation terms in equations (4) and (5) may be decomposed into 113 several anomalies. Usually one is interested in specific PV and θb anomalies 114 that appear to have a large influence on the phenomena of interest. Such 115 anomalies are usually clearly defined as temporally and spatially coherent 116 features. 117 PV anomalies that are far away from the region of interest, or difficult to 118 define temporally are generally defined as parts of a residual term, denoted 119 by qres . 120 Following Kristjánsson et al. (1999), the perturbation field in (4) is expressed 121 as: q′ = n X i=1 122 where qi denotes the anomalies. 7 qi (6) 123 For N selected PV anomalies the above relation becomes: ′ q = N X qi + qres (7) i=1 124 where the residual term is: n X qres = qi (8) i=N +1 125 The expressions for the θb fields are similar. 126 PPVI deals with obtaining uniquely the familiar fields of geopotential 127 height or streamfunction, winds and temperature distribution associated with 128 specific PV and θb anomalies. 129 To close the set of equations required for PV inversion a balance condition 130 is required, as well as lateral and horizontal boundary conditions. We adopt 131 Charney’s balance condition (Charney 1955) given by: ∇2 Φ = ∇ · (f ∇Ψ) + 2 132 ∂(∂Ψ/∂x, ∂Ψ/∂y) ∂(x, y) (9) where Φ is the geopotential and Ψ is the streamfunction. 133 The reason why Ertel’s PV (1) and the Charney balance condition (9) are 134 often preferred in PV inversion studies is that the flow is properly described 135 by (2) and (9), including flow with fairly large Rossby numbers, i.e. Ro ∼ 1, 136 e.g. allowing for strongly curved flow. The disadvantage of adopting these 137 equations directly is their non-linearity. This means that the streamfunction 8 138 fields associated with their respective PV and θb -anomalies, as well as con- 139 tributions from the mean and residual fields of PV and θb do not add up 140 to yield the total streamfunction field, as they do for quasi-geostrophic PV 141 (QPV). 142 Davis and Emanuel (1991) and Davis (1992) demonstrated how the non- 143 linearity problem can be resolved. For completeness this will now be briefly 144 recaptured. 145 Adopting the hydrostatic relation on the right hand side of (1) yields a 146 relation for q containing products of second derivatives of Ψ and Φ. See 147 e.g. relation (1.4) in Davis (1992) for details. Adopting this relation, Davis 148 and Emanuel (1991) demonstrated how Ertel’s PV can be inverted yielding 149 unambiguous results, as for inversion of QPV. The method is referred to as 150 the full linear (FL) method, because all terms are retained through collecting 151 the non-linear terms in the coefficient of the linear operator. 152 Davis (1992) presented two different PV inversion methods, i.e. subtrac- 153 tion from the total (ST) and addition to the mean (AM). He demonstrated 154 that the average of these two methods yields results close to those of the FL 155 method except for cases when the PV perturbations are much larger than 156 the mean PV-field. 157 158 The above description of the FL, ST and AM methods applies to boundary temperature (θb ) anomalies as well. 159 Rather than adopting the FL method we use the combination of the ST 160 and AM methods in the case study presented in section 4 below. The method 9 161 has also been succsessfully adopted in several case studies (e.g. Kristjánsson 162 et. al 1999, Thorsteinsson et. al 1999, Røsting and Kristjánsson 2006). 163 Though quasi-geostrophic theory is a weak anomaly theory, the inversion 164 of QPV is useful for demonstrating the principle of PV inversion and in 165 theoretical studies of atmospheric flows (e.g. the Eady model). The results 166 from such studies in E2008 led to the conclusion that PPVI in many cases fails 167 to contribute to our understanding of the dynamics of the flow. Since PPVI 168 has been adopted in many recent case studies (e.g. Bracegirdle and Gray 169 2009, McInnes et al. 2009) and constitutes a central part of PV thinking, 170 the criticism in E2008 needs to be addressed. 171 We now discuss some of the results presented by E2008. The main purpose 172 of the following discussion is to investigate to what extent PPVI is valid and 173 useful in the cases adopted in E2008. 174 3 175 Results from elementary tests of piecewise PV inversion 176 In E2008, in connection with Figure 1, it was stated that the winds in region 177 D2 are induced by the positive PV anomaly Z1 in region D1. However, 178 basing our arguments on the PPVI technique (as described in the previous 179 section), we note that the winds induced by Z1 only contribute to the total 180 wind field in region D2 in the same way as one of the components qi in (6) 181 is only one of many contributions to the total perturbation field q ′ . To what 10 182 extent Z1 influences the winds in D2 depends on the distance between the 183 PV anomalies Z1 and A2 and their amplitudes. 184 The same argument is valid for the ”complementary case”, i.e. the winds 185 associated with the negative PV anomaly A2 contribute to the total wind 186 field in region D1 and may influence Z1 to a certain extent. The total stream- 187 function ψtotal in D2 is then obtained through adding the contributions from 188 A2 and Z1. The same procedure yields ψtotal in D1. Hence performing in- 189 versions for both regions D1 and D2 is a necessary requirement for assessing 190 the validity of PPVI. 191 3.1 192 The barotropic case - Rossby waves in zonal mean flow 193 In the equation for PV in a barotropic flow (e.g. Holton 2004) we may define 194 the planetary vorticity f0 + βy as the mean PV field. Introducing a height 195 increment η = g −1 f0 ψ, we obtain through linearization an expression for 196 perturbation PV: q ′ = ∇2 ψ − δ 2 ψ (10) 197 where δ −1 = (gH)1/2f0 −1 is the Rossby radius of deformation. H is the mean 198 height of the fluid. 199 200 We now consider a one-dimensional barotropic Rossby wave, hence equation (10) becomes 11 q′ = ∂2ψ − δ2ψ ∂x2 (11) 201 In this example from E2008, the observed streamfunction field is described 202 by ψob = P sin(kx) from which the observed QPV anomaly field is calculated 203 by inserting ψob in equation (11). 204 The case presented in E2008 now proceeds as follows: Region D1 is defined 205 for (-L/2 ≤ x ≤ 0) and D2 for (0≤ x ≤ L/2) as shown in Figure 2, reproduced 206 from E2008. Following E2008, the PV field is modified by letting q ′ = 0 in 207 the region D2. 208 209 Figure 2 shows, besides ψob , the streamfunction ψ ′ obtained through inversion of the modified PV (solid contours). 210 To test the validity of PPVI for the flow in this example we now define 211 the complementary case for the regions shown in Figure 2, where q ′ = 0 in 212 region D1, and with q ′ = qob in D2 : q′ = 0 in D1; (12) −(k 2 + δ 2 )P sin(kx) in D2. 213 Carrying out the inversion of PV in this case we obtain the streamfunction 214 ψ ′′ , assuming that, as for the case of ψ ′ , that ψ ′′ and ∂ψ ′′ /∂x are continuous 215 at the origin. The solution is: 12 ψ ′′ = −Bsinh[δ(x + L/4] in D1; (13) P sin(kx) + Bsinh[δ(x − L/4] in D2. 216 where B = −P k/[2δcosh(δL/4)], i.e. B is identical to the coefficient A in 217 the case presented in E2008. 218 219 The distribution of ψ ′′ in this complementary case (13) is antisymmetric to that displayed in Figure 2, with respect to the origin. 220 Addition of the two complementary fields yields for D1 and D2, ψ = 221 ψ ′ + ψ ′′ = P sin(kx) = ψob . Hence addition of the fields associated with 222 the PV in D1 and D2 yields the total, observed PV field, as expected for 223 piecewise PV inversion. 224 3.2 225 In the baroclinic case described in E2008 the flow is confined between a 226 surface plane and a rigid lid at the top, the height of the model atmosphere 227 being H. The dynamical development of the flow in this case is described by 228 the equations of the Eady model (e.g. Holton 2004, Egger 2008). Figure 3, 229 reproduced from E2008, shows the tropopause at Hs = 3H/4. 230 Baroclinic waves in zonal shear flow Following E2008, an observed field is given as a modal streamfunction : ψob = P sin(kx)sin[n(z − H)] = ψ̂(z)sin(kx) 13 (14) 231 232 233 234 235 The horizontal wavenumber is k = 2π/L, and n = 3π/H is the vertical wavenumber. Figure 3a, retrieved from E2008, shows the structure of the observed streamfunction ψob . Now using quasi-geostrophic PV, the perturbation is expressed by ′ qob = ∇2 ψob + (f02 /N02 )∂ 2 ψob /∂z 2 (15) 236 where N02 ≡ (g/θ0 )dθ0 /dz is the Brunt-Vaisala frequency squared, expressing 237 static stability, here assumed to be constant. 238 239 By inserting the expression for ψob into (15), an equation for the amplitude of the streamfunction in the x-z plane is obtained: ∂ 2 ψ̂ N02 q̂ 2 − γ ψ̂ = ∂z 2 f02 240 (16) where γ = kN0 /f0 and q̂ is the amplitude of the PV perturbation. 241 ′ With modifications of the observed PV, i.e. changing PV to qob = 0 in 242 ′ the troposphere, i.e. below Hs (D2), while retaining the original qob in the 243 stratosphere, i.e. above Hs (D1), Egger retrieved the solution of ψ̂ through 244 inversion of the modified PV field. 245 The coefficients are determined by the boundary conditions: 246 (i) ψ̂ = ψˆob at z = 0 and z = H 247 (ii) Continuity requirements for ψ̂ and ∂ ψ̂/∂z at the tropopause z = 3H/4. 248 The distribution of ψ̂ is shown in Figure 3b. 14 249 ′ In the complementary case, not considered by E2008, qob is confined to 250 ′ the troposphere (D2), while qob = 0 in the stratosphere (D1). This PV 251 distribution yields the following solution: 252 in D1, stratosphere 253 ψ̂ = Ee−γz + F eγz (17) ψ̂ = ψˆob + Ge−γz + Keγz (18) in D2, troposphere 254 The coefficients in (17) and (18) may be determined by using the boundary 255 conditions (i) and (ii). 256 Adding the streamfunctions in E2008 to the ones in (17) and (18) , in regions 257 D1 and D2 respectively, the observed streamfunction ψˆob is obtained. Hence, 258 as in the barotropic case, PPVI yields the desired results. 259 3.3 260 Baroclinic waves in zonal shear flow - impacts from the lower boundary temperature anomalies 261 The final example from E2008 to be discussed addresses the impact on the 262 flow from lower boundary temperature anomalies. 263 With suitable boundary conditions the atmospheric flow (in region D1) 264 associated with the temperature anomalies at the lower boundary (region 15 265 D2) becomes: ψ ∗ = −nP sin(kx)sinh[γ(z − H)]/γcosh(γH) 266 (19) The validity of PPVI in this case may be verified by again including the 267 ′ complementary case, where q ′ = qob in 0 < z ≤ H. 268 The atmospheric flow for the complementary case is: 269 in D1 (0 < z ≤ H) ψ̂ = ψˆob + Ae−γz + Beγz (20) 270 The coefficients in equation (20) are determined by the following boundary 271 conditions: 272 (i) At the top of the model atmosphere (z = H), ψ̂ = ψˆob = 0 273 (ii) At the lower boundary ∂ ψ̂/∂z = 0 as z → 0 274 275 Through application of the assumptions (i) - (ii) the inverted streamfunction for the complementary case is retrieved: ψ ∗∗ = P sin(kx)sin[n(z − H)] + nP sin(kx)sinh[γ(z − H)]/γcosh(γH) (21) 276 Addition of expressions (19) and (21), that describe the two complementary 277 cases, yields for region D1: 16 ψ = ψ ∗ + ψ ∗∗ = P sin(kx)sin[n(z − H)] ≡ ψob (22) 278 Thus, we find that the two fields obtained from piecewise QPV inversion for 279 the whole region, comprised by D1 and D2, again add to yield the observed 280 field. Hence, also for this case the criticism of PPVI by E2008 is refuted. 281 3.4 282 The cases presented in E2008 appear to depend strongly on the mathematical 283 formulation of the problems, including the necessity of introducing the condi- 284 tions ψ = 0, implying q ′ = 0 in a selected domain, as seen from e.g. equation 285 (12). We now compare the inversion of the PV anomalies (11) and (15) 286 with an example discussed by Thorpe and Bishop (1995) and Holton (2004). 287 Expressing the QPV anomaly (15) with a three dimensional Laplacian op- 288 erator through the transformation ẑ = (N0 /f0 )z and introducing spherical 289 coordinates, we get due to symmetry: The impact of an isolated QPV anomaly q0 1 ∂ 2 ∂ψ (r ) = r 2 ∂r ∂r 0 0 ≤ r ≤ r0 ; (23) r > r0 . 290 In this example the QPV anomaly is assumed to consist of a constant q0 291 within a ball-shaped region. 292 293 Relation (23) is solved by integration with the following continuity and boundary conditions: 17 294 (i) ψ is defined at the origin. 295 (ii) Continuity of ψ and ∂ψ/∂r at the boundary (r = r0 ) of the PV anomaly. 296 (iii) ψ → 0 when r → ∞. 297 The solution yields an induced cyclonic wind field that is strongest at the 298 surface of the PV anomaly, i.e. at r = r0 . Beyond the surface the winds 299 become evanescent with increasing distance (r), see Figure 6.10 in Holton, 300 2004 (originally from Thorpe and Bishop, 1995). 301 We now assume that D1 is the region 0 ≤ r ≤ r0 and D2 is the adjacent 302 region r > r0 . We note from (23) that this problem avoids the requirement 303 q ′ = 0 in D1. This means that PV inversion in this case yields the ψ-field 304 associated with a specific positive PV anomaly. 305 The solution of (23) is quite different from the ones presented in E2008, 306 e.g. the solution presented in Figure 2 for the barotropic case where the 307 associated winds appear to remain strong far away from the PV anomaly. In 308 fact the results from the case described by (23) have a strong resemblance to 309 NWP simulated flows associated with isolated PV anomalies. 310 4 311 We now present a case study describing a real weather situation, in which the 312 results from PPVI, using the technique described in section 2, i.e. combining 313 the ST and AM methods, explain properly important aspects of cyclogene- Case study 18 314 sis. This implies explaining the mutual intensification of the upper level PV 315 anomaly and the low level (boundary) temperature anomaly, as well as the 316 strong frontogenesis that took place. The fields in Figure 4 are based on 317 a successful NWP simulation of the severe North Atlantic winter storm of 318 10-11 January 2006. The position of the surface low is indicated by ”L”. 319 Two snapshots of the development are shown, the first one at 06UTC 10 320 January shown in Figures 4a,b and the second 6 hours later in Figures 4c,d. PPVI reveals the following sequence of events: 321 322 (i) An upper level positive PV (UPV) anomaly is shown in Figure 4a. The 323 winds obtained from PPVI are strong at the edge of the UPV anomaly and 324 weaken with distance away from it. There is pronounced low level temper- 325 ature advection by the wind associated with the UPV anomaly (Figure 4b). 326 Strong cold advection takes place west and south of the cyclone center, while 327 warm advection is pronounced in the warm frontal region north and north- 328 west of the surface low. This pattern of temperature advection enhances the 329 lower boundary warm anomaly observed at the cyclone center denoted by 330 ′′ 331 (ii) Six hours later (Figure 4c) the UPV anomaly has become stronger due 332 to the impacts from the winds induced by the strengthened lower boundary 333 warm anomaly. Figure 4d shows that the warm front has intensified and 334 the cold advection west and south of the cyclone center has become stronger 335 over the 6 hour period through the intensification of the winds related to 336 a stronger positive UPV anomaly. In turn the cold advection to the south L′′ . 19 337 of the cyclone center enhances the lower boundary warm anomaly observed 338 at the ′′ L′′ in Figure 4d. Hence the mutual interaction between the warm 339 anomaly at the lower boundary and the UPV anomaly continues and becomes 340 stronger. 341 Figure 4 only shows one selected UPV anomaly as well as low level bound- 342 ary temperature fields. Nevertheless, the UPV and boundary temperature 343 anomalies presented in Figure 4 have a crucial impact on the initial stage of 344 the cyclogenesis. This example illustrates that PPVI is very useful for analyzing instanta- 345 346 neous flow patterns, as well as the dynamics of cyclogenesis. 347 5 348 By presenting some examples of PV inversion, Egger (2008) claimed that 349 PPVI is generally not useful for understanding atmospheric dynamics. Here, 350 by reanalyzing some of the cases presented in E2008 we demonstrate that 351 there are no reasons to discard PPVI as a valid diagnostic tool. In E2008 the 352 observed PV field was changed to qob = 0 in a selected domain (D2), while 353 qob was retained in the remaining, adjacent domain (D1). Regarding the PV 354 field in the adjacent domain (D1) as the anomaly, inversion of this anomaly 355 for the whole integration region (D1 + D2) yields a streamfunction different 356 from the one that is observed. 357 Discussion and conclusions We have argued that this procedure is misleading, and that in order to 20 358 test the validity of PPVI it is necessary to invert the PV field also in the 359 complementary cases. The complementary cases are obtained by retaining 360 qob and letting qob = 0 in the selected (D2) and adjacent (D1) domains 361 respectively. 362 Addition of the field obtained through PV inversion in the complementary 363 case should then yield the total (observed) field if PPVI is valid. As we have 364 shown, by including the complementary cases, the formalism of piecewise PV 365 inversion indeed remains valid in the cases presented in E2008. 366 The examples presented in E2008 do not include interaction between PV 367 anomalies, a shortcoming also pointed out by Methven and de Vries (2008). 368 Such interactions constitute an important part of assessing the dynamics 369 through PPVI. The results from E2008 are also model dependent, this is 370 illustrated by the example given by equation (23). Inversion of the QPV 371 anomaly given by equation (23) yields a solution resembling that obtained 372 through PPVI for a real case presented in Figure 4. The solution of (23) 373 is very different from those of the cases presented in E2008, illustrating the 374 dependency of the results from PPVI on the mathematical and dynamical 375 formulation of the problem. 376 Based on the above examples and case study (Figure 4), it appears that 377 PPVI is a highly useful tool for diagnosing dynamical processes in the atmo- 378 sphere. In particular PPVI provides a method for assessing which regions of 379 the atmospheric flow that are important for specific dynamical developments. 380 Though the various contributions from individual PV anomalies, obtained 21 381 through PPVI, cannot be observed in the real atmosphere, addition of the 382 contributions from all the PV anomalies, including those from the resid- 383 ual and mean PV, yields the observed streamfunction field. This suggests 384 that the various contributions from different PV anomalies are conceptually 385 meaningful (e.g. Thorsteinsson et al. 1999). This is also confirmed by case 386 studies of real situations where frontogenesis and low level developments can 387 be understood through the impact from specific PV anomalies as shown in 388 section 4 (e.g. Davis and Emanuel 1991, Stoelinga 1996, Røsting et al. 2003, 389 Bracegirdle and Gray 2009). 390 An interesting example of application of PPVI was recently provided 391 by Hinssen et al (2011). They studied the impact of the stratospheric PV 392 anomaly associated with the polar winter vortex on the tropospheric flow. 393 Given the large horizontal dimension of the stratospheric polar vortex, an 394 associated large scale PV anomaly was defined, and its effect on the tro- 395 pospheric flow turned out to be noticeable. During sudden stratospheric 396 warmings (SSW) a breakup or substantial weakening of the stratospheric 397 PV anomaly and the associated polar vortex takes place. PPVI shows that 398 the westerly flow in the troposphere weakens and may even become easterly 399 in some regions after a SSW. 22 400 6 References 401 Bracegirdle,T.J. and Gray,S.L. 2009: The dynamics of a polar low assessed 402 using potential vorticity inversion. Q.J.R. Meteorol.Soc.,135,880-893 403 Charney,J.G. 1947: The dynamics of long waves in a baroclinic westerly 404 current. J.Meteor., 4, 135-163 405 Charney,J.G. 1955: The use of the primitive equations of motions in numer- 406 ical prediction. Tellus, 7, 22-26 407 Davis, C.A. 1992: Piecewise potential vorticity inversion. J.Atmos.Sci., 49, 408 1397-1411 409 Davis,C.A. and Emanuel,K.A. 1991: Potential vorticity diagnostics of cyclo- 410 genesis. Mon. Weather Rev., 119,1929-1953 411 Demirtas,M. and Thorpe,A.J. 1999: Sensitivity of short range weather fore- 412 casts to local potential vorticity modifications. Mon.Weather Rev., 127, 922- 413 939 414 Eady, E.T. 1949: Long waves and cyclone waves. Tellus 1(3), 33-52 415 Egger.J. 2008: Piecewise Potential Vorticity Inversion: Elementary Tests. J. 416 Atmos. Sci., 65, 2015-2024. 417 Ertel, H. 1942: Ein neuer hydrodynamischer Wirbelsatz. Meteorologische 418 Zeitschrift 59, 277-281 419 Hinssen, Y., van Delden, A. and Opsteegh,T. 2011: Influence of sudden 420 stratospheric warmings on tropospheric winds. Meteorologische Zeitschrift, 421 Vol. 20, No3, 259-266 23 422 Holton, J.R. 2004: An Introduction to Dynamic Meteorology. Fourth Edi- 423 tion. ISBN: 0-12-354015-1, Elsevier Academic Press 424 Hoskins,B.J., McIntyre,M.E. and Robertson,A.W. 1985: On the use and sig- 425 nificance of isentropic potential vorticity maps. Q.J.R. Meteorol.Soc., 111, 426 877-946 427 Kristjánsson, J.E., Thorsteinsson,S. and Ulfarsson,G.F. 1999: Potential vor- 428 ticity based interpretation of the ’Greenhouse Low’, 2-3 February 1991. Tel- 429 lus, 51A, 233-248 430 Manders, A.M.M., Verkley, W.T.M., Diepeveen, J.J. and Moene, A.R. 2007: 431 Application of a potential vorticity modification method to a case of rapid 432 cyclogenesis over the Atlantic Ocean. Q.J.R. Meteorol. Soc.,133, 1755-1770 433 Mansfield,D. 1996: The use of potential vorticity as an operational forecast- 434 ing tool. Meteorol.Appl.,3,195-210 435 McInnes, H., Kristjánsson, J.E., Schyberg, H. and Røsting, B. 2009: An 436 assessment of a Greenland lee cyclone during the Greenland Flow Distortion 437 experiment - an observational approach. Q.J.R. Meteorol.Soc.,135,1968-1985 438 439 Methven, J. and de Vries, H. 2008: Comments on “Piecewise Potential Vor- 440 ticity Inversion: Elementary Tests”. J. Atmos. Sci., 65, 3003-3008 441 Rossby, C. G. 1940: Planetary flow patterns in the atmosphere. Q.J.R.Meteorol.Soc., 442 66, 68-87 443 Røsting, B., Kristjánsson, J. E. and Sunde,J 2003: The sensitivity of numer24 444 ical simulations to initial modifications of potential vorticity – a case-study. 445 Q.J.R.Meteorol.Soc.,129, 2697-2718 446 Røsting, B. and Kristjánsson, J. E. 2006: Improving simulations of severe 447 winter storms by initial modification of potential vorticity in sensitive regions. 448 Q.J.R.Meteorol.Soc., 132, 2625-2652 449 Røsting, B. and Kristjánsson, J. E. 2008:A successful resimulation of the 7-8 450 January 2005 winter storm through initial potential vorticity modification in 451 sensitive regions. Tellus 60A, 604-619 452 Santurette,P. and Georgiev,C.G. 2005: Weather Analysis and Forecasting, 453 Applying Satellite Water Vapor Imagery and Potential Vorticity Analysis. 454 Elsevier Academic Press, ISBN: 0-12-619262-6 455 Stoelinga, M. T. 1996: A potential vorticity-based study of the role of di- 456 abatic heating and friction in a numerically simulated baroclinic cyclone. 457 Mon. Weather Rev.,124, 849-874 458 Thorpe,A.J. and C.H.Bishop, 1995: Potential vorticity and the electrostatics 459 analogy: Ertel-Rossby formulation. Q.J.R Meteorol.Soc., 121, 1477-1495 460 Thorsteinsson, S.T, Kristjánsson, J.E., Røsting, B., Erlingsson, V.,and Ul- 461 farsson, G.F, 1999: A diagnostic study of the Flateyri avalanche cyclone, 462 24-26 October 1995, using potential vorticity inversion, Mon. Weather Rev., 463 127,1072-1088 464 Verkley,W.T.M., Vosbeek,P.W.C.,and Moene,A.R., 2005: Manually adjust- 465 ing a numerical weather analysis in terms of potential vorticity using three 25 466 dimensional variational data assimilation, Q.J.R. Meteorol. Soc.,131, 1713- 467 1736 468 7 469 Fig.1. Schematic presentation of a domain D divided into subdomains D1 and 470 D2. Z1 and A2 represent positive and negative PV anomalies respectively. 471 Reproduced from Egger (2008). 472 Fig.2. Observed streamfunction (ψob ) in dotted lines, forming ridges R1 and 473 R2 and troughs T1 and T2 in regions D1 and D2 respectively. Solid lines 474 show ψ obtained from inversion of the PV field after replacing the observed 475 PV with PV=0 in region D2. Reproduced from Egger (2008). 476 Fig.3. Idealized presentation of a two-dimensional atmosphere with depth H 477 = 20km. The tropopause is at z = 3H/4 = Hs , whereas n = 3π/H, k = 478 2π/L are vertical and horizontal wavenumber respectively. In a) the observed 479 streamfunction ψob is displayed. Positive values are in solid contours, while 480 negative values are dashed. In b) the streamfunction is obtained through 481 inversion of the PV field after replacing the observed PV with PV = 0 in the 482 troposphere (D2). The contours show the streamfunction, contours as in a). 483 Reproduced from Egger (2008). 484 Fig.4. Fields based on a simulation by the Norwegian HIRLAM12 model. 485 The position of the surface low is indicated by ”L”. 486 a) A positive UPV anomaly (contours every 0.5 PVU) and the associated Figure captions 26 487 winds obtained through PPVI, presented at the 400 hPa level. Fields valid 488 at 06 UTC 10 January 2006. 489 b) Potential temperature (contours every 2K) at 950 hPa and the winds 490 (obtained through PPVI) at 900 hPa associated with the UPV anomaly 491 shown in a. The fields are valid at 06 UTC 10 January 2006. 492 c) Same as in a, but valid at 12 UTC 10 January 2006. 493 d) Same as in b, but valid at 12 UTC 10 January 2006. 27 28 29 30 31