1. Introduction
Nakamura and Zhu (2010, hereafter NZ10) extended (1) for finite-amplitude Rossby waves and balanced eddies by introducing finite-amplitude wave activity (FAWA) based on the meridional displacement of quasigeostrophic potential vorticity (PV) from zonal symmetry. The formalism eliminates the cubic term from the right-hand side of (1) and extends the nonacceleration theorem (Charney and Drazin 1961) for an arbitrary eddy amplitude. This allows one to quantify the amount of the mean-flow modification by the eddy (Nakamura and Solomon 2010, 2011).
Despite its amenability to data, FAWA is a zonally averaged quantity and is incapable of distinguishing longitudinally isolated events, such as atmospheric blocking. In this article, we shall address this shortcoming by introducing local finite-amplitude wave activity (LWA). In essence, LWA quantifies longitude-by-longitude contributions to FAWA and, as such, recovers FAWA upon zonal averaging. As a first step into this topic, the present article concerns primarily the conservative dynamics of local eddy–mean flow interaction. Explicit representation of nonconservative dynamics (such as local diffusive flux of PV) will be deferred to a subsequent work. However, when observed data deviates from the theory, it may be readily interpreted as an indication of nonconservative effects. The material is organized as follows. Section 2 lays out the theory. Section 3 demonstrates the utility of LWA using idealized simulations with a barotropic vorticity equation on a sphere. We will compare LWA with one of the existing local metrics of finite-amplitude wave activity: impulse-Casimir wave activity (Killworth and McIntyre 1985; McIntyre and Shepherd 1987; Haynes 1988). As an application of LWA, a blocking episode that steered Superstorm Sandy to the U.S. East Coast in 2012 will be studied in section 4. Discussion and concluding remarks will follow in section 5.
2. Theory
a. Finite-amplitude wave activity
b. New formalism: The local finite-amplitude wave activity
1) Definitions
Although the FAWA formalism quantifies waviness in the PV contours and the associated mean-flow modification [see, for example, Solomon (2014) for stratospheric wave activity events], it is not suited to distinguish the longitudinal location of an isolated large-amplitude event such as blocking. To achieve this,
2) Local wave activity and PV gradient
Polvani and Plumb (1992) discuss two regimes of wave breaking in the context of vortex dynamics: major Rossby wave breaking that disrupts the vortex dynamics and microbreaking that only sheds filaments and does not affect the vortex significantly [see also Dritschel (1988)]. In terms of LWA, a major breaking would satisfy (19) as well as a large amplification in LWA
3) Local wave activity budget
c. Relationship to impulse-Casimir wave activity
Both wave activities obey similar equations [(20) and (24)], but while the ICWA equation is written entirely in terms of Eulerian quantities, the LWA equation involves line integrals and hence is Lagrangian in the meridional. A crucial difference arising from this is an extra meridional advection term
d. Local nonacceleration relation
The nonacceleration relation (4) shows conservation of the sum of zonal-mean zonal wind and wave activity in a frictionless barotropic flow, but it does not tell whether the deceleration of the zonal-mean wind is because of growth of a localized wave packet or simultaneous growth of multiple wave packets over longitudes. To understand the dynamics of a localized phenomenon such as blocking, it is desirable to characterize eddy–mean flow interaction over a regional scale.
3. Numerical experiment
a. Experimental setup
b. Comparison between and
The overall flow evolution is similar to that in HP87: the wave packet initially located on the northern side of the jet axis splits into poleward and equatorward-migrating tracks, and, as they approach critical lines at the flanks of the jet, they produce wave breaking. The initial vorticity pattern consists of six pairs of positive and negative anomalies (Fig. 3, top), but their strengths are not symmetric because of the addition of small-amplitude, secondary wave
The snapshots of absolute vorticity, LWA
c. Local negative correlation between and
For a zonal-mean state, the nonacceleration relation in (4) describes conservative eddy–mean flow interaction:
The opposite tendency of the two quantities is evident, particularly during the early stage of simulation. Also plotted in the top panel are the sum
Figure 6 extends the above analysis to the entire latitude circle by showing the longitude–time (Hovmöller) cross sections (Hovmöller 1949) of
We have repeated the analysis varying
4. Analysis of a blocking episode
Blocking is a phenomenon at midlatitudes in which a large-scale pressure anomaly remains stationary. The normal westerly winds in the mid- to upper troposphere are diverted meridionally along the blocking pattern, and the wind within the block is often replaced by easterlies. Lejenäs and Økland (1983) observed that blocking occurs at longitudes where the latitudinal average of the zonal wind at 500 hPa is easterly. Tibaldi and Molteni (1990) added an additional requirement that the average wind be westerly poleward of the block. Such description of blocking based on reversal of zonal wind is a kinematic statement. Given the potential of (36) to quantify the slowing down of the flow by finite-amplitude eddies, the formalism is well suited for identifying and investigating blocking events with meteorological data based on dynamics.
In this section, we explore the extent to which the dynamics of a real blocking episode may be characterized based on the conservation relation in (36). In particular, we will study the blocking episode that steered Superstorm Sandy to the U.S. East Coast during October 2012 with the LWA formalism. The interior and surface LWA as well as the zonal wind are evaluated from the European Centre for Medium-Range Weather Forecasts (ECMWF) interim reanalysis product (ERA-Interim; Dee et al. 2011; http://apps.ecmwf.int/datasets/data/interim-full-daily/) at a horizontal resolution of 1.5° × 1.5°.
First, we evaluate PV from (3) on 49 equally spaced pressure pseudoheights, as described in Nakamura and Solomon (2010) (we assume that H = 7 km). Then, we compute
a. Overview of zonal wind and LWA in northern autumn 2012
Longitude–time (Hovmöller) diagrams for the barotropic components of zonal wind
b. Blocking episode around North American east coast during 27 October–2 November 2012
Now we focus on a single blocking episode that occurred during 27 October–2 November over the North Atlantic. (The longitudinal range of concern is marked by the black lines in Fig. 7.) This episode was characterized by a persistent blocking pattern in the mid- to upper troposphere and contributed to the steering of Superstorm Sandy at a right angle to the U.S. East Coast (Blake et al. 2013). Figures 8 and 9, respectively, show PV and the corresponding LWA
One might ask how the barotropic component (density-weighted vertical average) samples the vertical distribution of LWA associated with blocking. Figure 10 shows the vertical structure of
To examine the extent to which the local nonacceleration relation accounts for the simultaneous accumulation of LWA and deceleration of zonal flow, in Fig. 11 we show
There are several remarkable features from the plots. First, there is a strong negative correlation in the time series of
5. Summary and discussion
We have generalized the notion of FAWA introduced by NZ10 to LWA, a diagnostic for longitudinally localized wave events, and tested its utility in both a barotropic model and meteorological data. A significant advantage of LWA over the existing wave activity measures is that it carries over the nonacceleration relation of FAWA to regional scales, albeit within the WKB approximation. This explicitly attributes local deceleration of the zonal flow to accumulation of wave activity.
A robust negative correlation is found between
LWA dynamically connects the two criteria of blocking indices: 1) deceleration or even reversal of westerlies (Lejenäs and Økland 1983; Tibaldi and Molteni 1990) and 2) large amplitude of anomalies or gradient reversal in either geopotential height (at 500 hPa) (Barnes et al. 2012; Dunn-Sigouin and Son 2013) or potential temperature on constant potential vorticity surface (2 potential vorticity units) (Pelly and Hoskins 2003). Hoskins et al. (1985) suggests that meridional gradient reversal of potential temperature on a constant PV surface could imply a reversal of westerlies via the invertibility principle, but such a relation is not explicit. LWA can potentially serve as a blocking index because a large LWA will automatically lead to a significant deceleration of local zonal wind, to the extent that nonacceleration relation holds.
Acknowledgments
Constructive critiques of the original manuscript by three anonymous reviewers are gratefully acknowledged. This research has been supported by NSF Grant AGS-1151790. The ERA-Interim dataset used in this study was obtained from the ECMWF data server (http://apps.ecmwf.int/datasets/data/interim-full-daily/).
APPENDIX A
APPENDIX B
Derivation of (20)–(22)
Taking the time derivative and using the Leibniz rule and (9)
REFERENCES
Allen, D. R., and N. Nakamura, 2003: Tracer equivalent latitude: A diagnostic tool for isentropic transport studies. J. Atmos. Sci., 60, 287–304, doi:10.1175/1520-0469(2003)060<0287:TELADT>2.0.CO;2.
Andrews, D., and M. E. McIntyre, 1976: Planetary waves in horizontal and vertical shear: The generalized Eliassen–Palm relation and the mean zonal acceleration. J. Atmos. Sci., 33, 2031–2048, doi:10.1175/1520-0469(1976)033<2031:PWIHAV>2.0.CO;2.
Barnes, E. A., J. Slingo, and T. Woollings, 2012: A methodology for the comparison of blocking climatologies across indices, models and climate scenarios. Climate Dyn., 38, 2467–2481, doi:10.1007/s00382-011-1243-6.
Blake, E. S., T. B. Kimberlain, R. J. Berg, J. Cangialosi, and J. L. Beven II, 2013: Tropical cyclone report: Hurricane Sandy. National Hurricane Center Tropical Cyclone Rep. AL182012, 157 pp. [Available online at http://www.nhc.noaa.gov/data/tcr/AL182012_Sandy.pdf.]
Bühler, O., 2014: Waves and Mean Flows. Cambridge University Press, 374 pp.
Butchart, N., and E. Remsberg, 1986: The area of the stratospheric polar vortex as a diagnostic for tracer transport on an isentropic surface. J. Atmos. Sci., 43, 1319–1339, doi:10.1175/1520-0469(1986)043<1319:TAOTSP>2.0.CO;2.
Charney, J. G., and P. G. Drazin, 1961: On the transfer of energy in stationary mountain waves. J. Geophys. Res., 66, 83–110, doi:10.1029/JZ066i001p00083.
Dee, D. P., and Coauthors, 2011: The ERA-Interim reanalysis: Configuration and performance of the data assimilation system. Quart. J. Roy. Meteor. Soc., 137, 553–597, doi:10.1002/qj.828.
Dritschel, D. G., 1988: The repeated filamentation of two-dimensional vorticity interfaces. J. Fluid Mech., 194, 511–547, doi:10.1017/S0022112088003088.
Dunn-Sigouin, E., and S.-W. Son, 2013: Northern Hemisphere blocking frequency and duration in the CMIP5 models. J. Geophys. Res. Atmos., 118, 1179–1188, doi:10.1002/jgrd.50143.
Durran, D. R., 2013: Numerical Methods for Wave Equations in Geophysical Fluid Dynamics. Springer Science & Business Media, 516 pp.
Haynes, P. H., 1988: Forced, dissipative generalizations of finite-amplitude wave-activity conservation relations for zonal and nonzonal basic flows. J. Atmos. Sci., 45, 2352–2362, doi:10.1175/1520-0469(1988)045<2352:FDGOFA>2.0.CO;2.
Held, I. M., and P. J. Phillipps, 1987: Linear and nonlinear barotropic decay on the sphere. J. Atmos. Sci., 44, 200–207, doi:10.1175/1520-0469(1987)044<0200:LANBDO>2.0.CO;2.
Hoskins, B. J., M. McIntyre, and A. W. Robertson, 1985: On the use and significance of isentropic potential vorticity maps. Quart. J. Roy. Meteor. Soc., 111, 877–946, doi:10.1002/qj.49711147002.
Hovmöller, E., 1949: The trough-and-ridge diagram. Tellus, 1A, 62–66, doi:10.1111/j.2153-3490.1949.tb01260.x.
Killworth, P. D., and M. E. McIntyre, 1985: Do Rossby-wave critical layers absorb, reflect, or over-reflect? J. Fluid Mech., 161, 449–492, doi:10.1017/S0022112085003019.
Lejenäs, H., and H. Økland, 1983: Characteristics of Northern Hemisphere blocking as determined from a long time series of observational data. Tellus, 35A, 350–362, doi:10.1111/j.1600-0870.1983.tb00210.x.
Magnusdottir, G., and P. H. Haynes, 1996: Wave activity diagnostics applied to baroclinic wave life cycles. J. Atmos. Sci., 53, 2317–2353, doi:10.1175/1520-0469(1996)053<2317:WADATB>2.0.CO;2.
McIntyre, M., and T. Shepherd, 1987: An exact local conservation theorem for finite-amplitude disturbances to non-parallel shear flows, with remarks on Hamiltonian structure and on Arnold’s stability theorems. J. Fluid Mech., 181, 527–565, doi:10.1017/S0022112087002209.
Methven, J., and P. Berrisford, 2015: The slowly evolving background state of the atmosphere. Quart. J. Roy. Meteor. Soc., 141, 2237–2258, doi:10.1002/qj.2518.
Nakamura, N., and A. Solomon, 2010: Finite-amplitude wave activity and mean flow adjustments in the atmospheric general circulation. Part I: Quasigeostrophic theory and analysis. J. Atmos. Sci., 67, 3967–3983, doi:10.1175/2010JAS3503.1.
Nakamura, N., and D. Zhu, 2010: Finite-amplitude wave activity and diffusive flux of potential vorticity in eddy–mean flow interaction. J. Atmos. Sci., 67, 2701–2716, doi:10.1175/2010JAS3432.1.
Nakamura, N., and A. Solomon, 2011: Finite-amplitude wave activity and mean flow adjustments in the atmospheric general circulation. Part II: Analysis in the isentropic coordinate. J. Atmos. Sci., 68, 2783–2799, doi:10.1175/2011JAS3685.1.
Pelly, J. L., and B. J. Hoskins, 2003: A new perspective on blocking. J. Atmos. Sci., 60, 743–755, doi:10.1175/1520-0469(2003)060<0743:ANPOB>2.0.CO;2.
Plumb, R. A., 1985: On the three-dimensional propagation of stationary waves. J. Atmos. Sci., 42, 217–229, doi:10.1175/1520-0469(1985)042<0217:OTTDPO>2.0.CO;2.
Polvani, L. M., and R. A. Plumb, 1992: Rossby wave breaking, microbreaking, filamentation, and secondary vortex formation: The dynamics of a perturbed vortex. J. Atmos. Sci., 49, 462–476, doi:10.1175/1520-0469(1992)049<0462:RWBMFA>2.0.CO;2.
Solomon, A., 2014: Wave activity events and the variability of the stratospheric polar vortex. J. Climate, 27, 7796–7806, doi:10.1175/JCLI-D-13-00756.1.
Solomon, A., and N. Nakamura, 2012: An exact Lagrangian-mean wave activity for finite-amplitude disturbances to barotropic flow on a sphere. J. Fluid Mech., 693, 69–92, doi:10.1017/jfm.2011.460.
Swanson, K., 2000: Stationary wave accumulation and the generation of low-frequency variability on zonally varying flows. J. Atmos. Sci., 57, 2262–2280, doi:10.1175/1520-0469(2000)057<2262:SWAATG>2.0.CO;2.
Takaya, K., and H. Nakamura, 2001: A formulation of a phase-independent wave-activity flux for stationary and migratory quasigeostrophic eddies on a zonally varying basic flow. J. Atmos. Sci., 58, 608–627, doi:10.1175/1520-0469(2001)058<0608:AFOAPI>2.0.CO;2.
Thuburn, J., and V. Lagneau, 1999: Eulerian mean, contour integral, and finite-amplitude wave activity diagnostics applied to a single-layer model of the winter stratosphere. J. Atmos. Sci., 56, 689–710, doi:10.1175/1520-0469(1999)056<0689:EMCIAF>2.0.CO;2.
Tibaldi, S., and F. Molteni, 1990: On the operational predictability of blocking. Tellus, 42A, 343–365, doi:10.1034/j.1600-0870.1990.t01-2-00003.x.
Wang, L., and N. Nakamura, 2015: Covariation of finite-amplitude wave activity and the zonal mean flow in the midlatitude troposphere: 1. Theory and application to the Southern Hemisphere summer. Geophys. Res. Lett., 42, 8192–8200, doi:10.1002/2015GL065830.
Williams, I. N., and S. J. Colucci, 2010: Characteristics of baroclinic wave packets during strong and weak stratospheric polar vortex events. J. Atmos. Sci., 67, 3190–3207, doi:10.1175/2010JAS3279.1.