1. Introduction
Atmospheric blocking represents a disruption of eastward migration of synoptic eddies and often causes persistent weather anomalies in the midlatitudes. Distribution and frequency of blocking episodes, together with their trend in a changing climate, are under active investigation (Barriopedro et al. 2006; Croci-Maspoli et al. 2007; Tyrlis and Hoskins 2008; Masato et al. 2013; Barnes et al. 2014; Mokhov et al. 2014). Blocking itself is a low-frequency phenomenon that can persist for a week or longer, but the role of high-frequency (synoptic) eddies and their feedback on the low-frequency dynamics in the block formation has been noted for some time (Berggren et al. 1949; Green 1977; Shutts 1983 and references therein; Colucci 1985, 2001; Mullen 1987; Altenhoff et al. 2008). Nakamura et al. (1997) show that the feedback from synoptic eddies accounts for the majority of observed block formation over the North Pacific but less than half of blocking events over Europe. A number of theoretical studies based on barotropic dynamics demonstrate that local interaction between a preexisting stationary diffluent flow and incident transient eddies can amplify a blocklike quasi-stationary feature (Shutts 1983; Haines and Marshall 1987; Anderson 1995; Luo 2005). However, identifying a precise condition that initiates blocking formation remains elusive.
In this article we revisit equivalent barotropic dynamics of a potential vorticity (PV) front, in which a train of short edge waves enters the domain from upstream and interacts with a diffluent jet downstream. It will be shown that this idealized model predicts a clear transition from steady downstream propagation of waves to highly deformed, blocklike wave tracks, depending on the upstream wave activity flux and the geometry of the incipient flow. The model dynamics is similar to the one considered by Swanson et al. (1997) except for two aspects. First, transient waves are fed continuously from upstream rather than as an isolated wave packet, and second, the layer depth is finite and allowed to vary meridionally to mimic the isentropic structure of PV near the extratropical tropopause. The finite Rossby radius only affects the basic state and the low-frequency response of the flow, since the transient wave forcing is in the shortwave (barotropic) limit and insensitive to the Rossby radius.
Another departure from Swanson et al. (1997) is that, in addition to the conservation of wave action (Bretherton and Garrett 1968), we will exploit the conservation of local wave activity (LWA) and associated local nonacceleration relation (Huang and Nakamura 2016, hereafter HN16) to predict the threshold behavior of wave envelope in a 1D quasilinear theory. As we will see below, given a longitudinally varying initial zonal flow and an upstream influx of wave activity, there is a threshold LWA beyond which no steady wave envelope exists. This allows one to predict the location of the onset of wave breaking. The line of argument is akin to Fyfe and Held (1990) for meridionally propagating Rossby waves and Wang and Fyfe (2000) for vertically propagating edge waves. We will then demonstrate numerically, both in 1D and fully nonlinear 2D calculations, that the wave amplitude undergoes a significant transformation once this threshold is reached. When the PV jump has significant contribution from the variation of layer thickness, the resulting 2D flow captures salient features of an atmospheric blocking in a manner similar to Shutts (1983).
The next section outlines the 1D theory. Section 3 describes the 2D model and the results of numerical experiments. Relevance to the observed atmospheric blocking will be discussed in the concluding section.
2. One-dimensional quasilinear theory for a slowly varying flow
a. Local wave activity and wave–zonal flow interaction
b. Transient behavior
The domain is discretized into 4097 grids and a standard finite difference method is used. We also add a small second-order diffusion term to (30) for numerical stability. [The diffusion coefficient is
Figure 2 depicts the numerical solutions for
Figure 3 shows an analogous result for
3. Two-dimensional nonlinear calculations
The preceding analysis shows that 1D quasi-linear edge waves develop a migratory shock in the envelope once the zonal flow is decelerated to one-half of the initial value. Since the discontinuity violates the premise of slow variation, the WKB approximation formally breaks down at this point. Similar loss of steadiness in wave activity was previously associated with the onset of wave breaking (Fyfe and Held 1990; Wang and Fyfe 2000). Since the shock formation may be viewed as a botched attempt by the wave envelope to overturn (i.e., to become multivalued), it indeed suggests an onset of wave breaking, or at least significant deformation of the wave envelope, in the full 2D problem. For a more accurate description of the flow evolution past the threshold, we shall perform direct numerical simulations in 2D. While the contour advection algorithm (e.g., Dritschel and Ambaum 1997) would be a natural extension to the above linear analysis, we adopt a finite-difference model on a rectangular domain instead, in which PV front is allowed to have a narrow but finite width similar to Harvey et al. (2016). The use of the finite-difference model is partly motivated by the ease with which to implement the open boundary condition. To describe the eddy–zonal flow interaction in as clean a setup as possible, transient waves are forced upstream and allowed to migrate downstream and exit the domain without reentering, just as in the 1D problem discussed earlier.
Seasonal climatology of PV variation across the tropopause on the 340-K isentropic surface. The second column describes the range of latitudes in which PV varies most. The variation of PV in each latitude band relative to the global range of 340-K PV is shown in the third column. The last column shows the percentage contribution of the thickness variation to the PV variation within each latitude band; 50% corresponds to β* = 1 in (50). Based on the 1979–2014 climatology computed with ERA-Interim Dee et al. (2011).
a. Weak forcing without thickness variation
First, we assume that there is no thickness variation
Figure 7 shows the corresponding changes in the zonal flow (left) and the LWA flux (right) at the axis of the jet
b. Strong forcing without thickness variation
The above case produces a steady wave envelope reminiscent of the quasilinear WKB solution. Now we increase the forcing to
The separatrices of
c. Strong forcing with thickness variation
In the preceding case the transient waves decelerate the zonal flow to the point that a significant transformation occurs in the wave envelope, eventually leading to a new 2D steady state. The longitudinal location at which this transformation occurs is largely consistent with the prediction by the 1D theory. However, despite the enormous deformation of the PV contours, the corresponding change in streamfunction is modest. The zonal flow does decrease below the threshold but it remains positive at
Now we introduce a cross-stream thickness variation by letting
By t* = 16 the incident transient waves are almost completely blocked. There is a precipitous drop (and reversal) of the zonal LWA flux from x* ≈ 12 to 2.7, yet virtually no flux leaks out downstream of
The role of the layer thickness variation in the block formation is qualitatively understood as follows. The meridional displacement of the PV contour induces a net (i.e., local phase average) PV anomaly whose sign is opposite across y = 0. Close to the jet axis, the net PV anomaly is about
d. Parameter sweep for and
We have repeated the experiments for a range of forcing amplitude
Model behavior as a function of
4. Summary and discussion
We have explored the role of interaction between a high-frequency transient wave train and a diffluent zonal flow in the low-frequency dynamics of an equivalent barotropic PV front. Central to the dynamics are (i) a positive feedback between the accumulation of LWA and the local deceleration of the zonal flow and (ii) the cross-stream variation of the layer thickness. The conservation of LWA for slowly modulated 1D quasilinear edge waves predicts that a steady WKB solution ceases to exist once the local zonal flow is decelerated below one-half of its original value [i.e., the violation of (29); see related discussions by Fyfe and Held (1990) and Wang and Fyfe (2000)]. The location at which this occurs may be predicted from the upstream LWA flux and the profile of the initial zonal flow downstream. In the 1D transient problem, the wave envelope develops a shock once this threshold is reached. An analogy to the traffic flow problem (Lighthill and Whitham 1955; Richards 1956) was noted.
We have tested the theoretical prediction with 2D nonlinear experiments, in which the zonal variation of the background flow is assumed to occur on the scale of the Rossby radius. The experiments agree with the theory in that once the predicted threshold is exceeded the wave envelope undergoes a significant transformation. When the PV discontinuity is associated entirely with the vorticity profile, the result is a partial deflection of LWA analogous to Swanson et al. (1997). In this case the wave envelope reorganizes into a 2D steady state, in which the waves are partially deflected sideways and partially transmitted downstream. When the PV jump is augmented by a sufficiently large thickness variation, the meridional displacement of PV leads to a self-organization of an anticyclone–cyclone pair across the jet axis, in a manner similar to atmospheric blocking. The zonal flow is reversed and the incident wave train is split in two tracks at the stagnation point. The incoming flux of LWA is entrained into the block and partially dissipated by mixing, but very little flux escapes to the downstream side of the block. From the last column of Table 1, the required threshold of thickness variation for block formation (~50%) is generally met across the extratropical tropopause.
The present work provides a paradigm in which a relatively clean threshold exists for a transition from a westerly jet to a blocked state in response to high-frequency wave forcing—a counterpart to the block formation through instability/nonlinearity in low-frequency dynamics (e.g., Swanson 2000; Cash and Lee 2000). In our last experiment (Fig. 10) the block evolution is very slow compared to the frequency of the wave forcing, yet the mechanism of feedback from high-frequency to low-frequency dynamics is not trivial. The phase structure of PV anomalies cascades to small scales in a diffluent shear flow, so each individual PV filament does not carry much dynamical significance. However, the envelope of the PV contours grows by absorbing LWA from upstream and this affects the large-scale circulation. Of the processes that control the wave envelope in our model, the zonal advective flux of LWA is most influential but PV mixing also plays an appreciable role.
We have chosen our model setup so that the theory takes on a mathematically simplest form. Some of the assumptions invoked in the theory, including the shortwave (barotropic) and high-frequency limits, may seem restrictive. However our 2D numerical solutions show that as the wave enters a diffluent region its phase is compressed, suggesting that the shortwave assumption is indeed more appropriate where the wave interacts with the flow. While specific results like (29) do hinge on this assumption, we believe that the overall behavior of the model, such as the existence of wave breaking threshold, will remain unaffected in a more general setup insomuch as the dynamics is characterized by the along-stream propagation of transient waves. For example, the shock formation in the wave envelope only requires that the wave activity flux have a maximum with respect to wave activity. This is a consequence of nonlinearity in wave–mean flow interaction and is unlikely to be altered by the details of the parameter setting. As a test for the robustness of the result, we have run the 2D model with a much lower forcing frequency (ω* = 4) with other parameters unchanged from Fig. 10. The result is shown in Fig. 12. Compared with Fig. 10, the incident wave has a much greater wavelength, but its phase still collapses at the stagnation point, guiding finite-sized eddies along split paths. The eventual formation of blocking remains qualitatively similar to Fig. 10, although the streamfunction is appreciably noisier.
The present work concerns only the formation stage of blocking—our numerical experiments did not produce a well-defined life cycle of blocking. Even as a model of block formation, the extent to which this idealized study is applicable as a predictive tool remains to be seen, because in reality wave forcing and the response of the flow are not readily separable. For example, deep intrusion of tropical air with low PV into high latitudes occurs regularly from explosive cyclogenesis in the Northern Hemisphere winter storm tracks and it provides potent forcing for blocking (Colucci 1985). In Fig. 10 the emerging blocking high and cut-off low have comparable strengths owing to symmetry in the dynamics, but the symmetry may be broken by an asymmetric flow profile, Earth’s sphericity, or nonquasigeostrophic effects (Nakamura 1993). The behavior of our model is also distinct from those of the barotropic models on the β plane used in previous studies (e.g., Shutts 1983; Luo 2005), which produce blocks without the need for thickness variation. It appears that the piecewise-uniform PV model without thickness variation is prohibitive for block formation because the produced PV anomalies tend to be passive and incapable of self-organizing into large eddies.
Mutual reinforcement of the wave activity accumulation and the local deceleration of the zonal flow is corroborated by meteorological data. A recent work by Huang and Nakamura (2017) examines the regional budgets of column-averaged LWA over the Pacific and Atlantic storm tracks in the northern winter using the ERA-Interim products. Their analysis shows that covariance of the column-averaged LWA and the column-averaged zonal wind is largely negative and strongest in the jet exit (diffluent) regions (see Fig. 4 of their supporting information). Furthermore, they also show that on synoptic time scales the tendency of the column-averaged LWA is dominated by the convergence of the zonal LWA flux. This is consistent with (30). That the magnitude of the LWA–zonal flow covariation is greatest in the diffluent part of the zonal flow suggests that synoptic to intraseasonal variability of LWA occurs through the zonal flux variation in a manner described in this article. Future work will explore statistical connections between the upstream LWA flux and the frequency and magnitude of large LWA events in the diffluent regions.
Acknowledgments
This work has been supported by NSF Grant AGS-1563307. The ERA-Interim dataset used in this study (Dee et al. 2011) was obtained from the ECMWF data server (at http://apps.ecmwf.int/datasets/data/interim-AU2 full-daily/) with the horizontal resolution of 1.5°. Helpful discussions with Malte Jansen are gratefully acknowledged. Critiques by three anonymous reviewers also contributed to significant improvements in the presentation.
APPENDIX
Conservation of Quasilinear Local Wave Activity and Zonal Momentum
a. Local wave activity in a barotropic PV front
b. Local zonal flow in a barotropic PV front
REFERENCES
Altenhoff, A. M., O. Martius, M. Croci-Maspoli, C. Schwierz, and H. C. Davies, 2008: Linkage of atmospheric blocks and synoptic-scale Rossby waves: A climatological analysis. Tellus, 60A, 1053–1063, doi:10.1111/j.1600-0870.2008.00354.x.
Anderson, J., 1995: A simulation of atmospheric blocking with a forced barotropic model. J. Atmos. Sci., 52, 2593–2608, doi:10.1175/1520-0469(1995)052<2593:ASOABW>2.0.CO;2.
Barnes, E. A., E. Dunn-Sigouin, G. Masato, and T. Woollings, 2014: Exploring recent trends in Northern Hemisphere blocking. Geophys. Res. Lett., 41, 638–644, doi:10.1002/2013GL058745.
Barriopedro, D., R. Garcia-Herrera, A. Lupo, and E. Hernandez, 2006: A climatology of Northern Hemisphere blocking. J. Climate, 19, 1042–1063, doi:10.1175/JCLI3678.1.
Berggren, R., B. Bolin, and C.-G. Rossby, 1949: An aerological study of zonal motion, its perturbations and break-down. Tellus, 1, 14–37, doi:10.3402/tellusa.v1i2.8501.
Bretherton, F. P., and C. J. R. Garrett, 1968: Wavetrains in ingomogeneous moving media. Proc. Roy. Soc. London, 302A, 529–554, doi:10.1098/rspa.1968.0034.
Butchart, N., K. Haines, and J. Marshall, 1989: A theoretical and diagnostic study of solitary waves and atmospheric blocking. J. Atmos. Sci., 46, 2063–2078, doi:10.1175/1520-0469(1989)046<2063:ATADSO>2.0.CO;2.
Cash, B., and S. Lee, 2000: Dynamical processes of block evolution. J. Atmos. Sci., 57, 3202–3218, doi:10.1175/1520-0469(2000)057<3202:DPOBE>2.0.CO;2.
Charney, J., and P. Drazin, 1961: Propagation of planetary-scale disturbances from the lower into the upper atmosphere. J. Geophys. Res., 66, 83–109, doi:10.1029/JZ066i001p00083.
Colucci, S., 1985: Explosive cyclogenesis and large-scale circulation changes: Implications for atmospheric blocking. J. Atmos. Sci., 42, 2701–2717, doi:10.1175/1520-0469(1985)042<2701:ECALSC>2.0.CO;2.
Colucci, S., 2001: Planetary-scale preconditioning for the onset of blocking. J. Atmos. Sci., 58, 933–942, doi:10.1175/1520-0469(2001)058<0933:PSPFTO>2.0.CO;2.
Croci-Maspoli, M., C. Schwierz, and H. Davies, 2007: A multifaceted climatology of atmospheric blocking and its recent linear trend. J. Climate, 20, 633–649, doi:10.1175/JCLI4029.1.
Dee, D., 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., and M. Ambaum, 1997: A contour-advective semi-Lagrangian numerical algorithm. Quart. J. Roy. Meteor. Soc., 123, 1097–1130, doi:10.1002/qj.49712354015.
Durran, D. R., 1991: The third-order Adams–Bathforth method: An attractive alternative to leapfrog time differencing. Mon. Wea. Rev., 119, 702–720, doi:10.1175/1520-0493(1991)119<0702:TTOABM>2.0.CO;2.
Fyfe, J., and I. Held, 1990: The two-fifths and one-fifth rules for Rossby wave breaking in the WKB limit. J. Atmos. Sci., 47, 697–705, doi:10.1175/1520-0469(1990)047<0697:TTFAOF>2.0.CO;2.
Green, J., 1977: The weather during July 1976: Some dynamical considerations of the drought. Weather, 32, 120–128, doi:10.1002/j.1477-8696.1977.tb04532.x.
Haines, K., and J. Marshall, 1987: Eddy-forced coherent structures as a prototype of atmospheric blocking. Quart. J. Roy. Meteor. Soc., 113, 681–704, doi:10.1002/qj.49711347613.
Harvey, B., J. Methven, and M. Ambaum, 2016: Rossby wave propagation on potential vorticity fronts with finite width. J. Fluid Mech., 794, 775–797, doi:10.1017/jfm.2016.180.
Hoskins, B., M. McIntyre, and A. 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.
Huang, C. S. Y., and N. Nakamura, 2016: Local finite-amplitude wave activity as a diagnostic of anomalous weather event. J. Atmos. Sci., 73, 211–229, doi:10.1175/JAS-D-15-0194.1.
Huang, C. S. Y., and N. Nakamura, 2017: Local wave activity budgets of the wintertime Northern Hemisphere: Implication for the Pacific and Atlantic storm tracks. Geophys. Res. Lett., doi:10.1002/2017GL073760, in press.
LeVeque, R., 2002: Finite Volume Methods for Hyperbolic Problems. Cambridge University Press, 580 pp.
Lighthill, M., and G. Whitham, 1955: On kinematic waves. II. A theory of traffic flow on long crowded roads. Proc. Roy. Soc. London, 229A, 317–345, doi:10.1098/rspa.1955.0089.
Luo, D., 2005: Barotropic envelope Rossby soliton model for block-eddy interaction. Part I: Effect of topography. J. Atmos. Sci., 62, 5–21, doi:10.1175/1186.1.
Masato, G., B. Hoskins, and T. Woollings, 2013: Winter and summer Northern Hemisphere blocking in CMIP5 models. J. Climate, 26, 7044–7059, doi:10.1175/JCLI-D-12-00466.1.
McWilliams, J., 1980: An application of equivalent modons to atmospheric blocking. Dyn. Atmos. Oceans, 5, 43–66, doi:10.1016/0377-0265(80)90010-X.
Mokhov, I., A. Timazhev, and A. Lupo, 2014: Changes in atmospheric blocking characteristics within Euro-Atlantic region and Northern Hemisphere as a whole in the 21st century from model simulations using RCP anthropogenic scenarios. Global Planet. Change, 122, 265–270, doi:10.1016/j.gloplacha.2014.09.004.
Mullen, S., 1987: Transient eddy forcing of blocking flows. J. Atmos. Sci., 44, 3–22, doi:10.1175/1520-0469(1987)044<0003:TEFOBF>2.0.CO;2.
Nakamura, N., 1993: Momentum flux, flow symmetry, and the nonlinear barotropic governor. J. Atmos. Sci., 50, 2159–2179, doi:10.1175/1520-0469(1993)050<2159:MFFSAT>2.0.CO;2.
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.
Nakamura, H., M. Nakamura, and J. L. Anderson, 1997: The role of high- and low-frequency dynamics in blocking formation. Mon. Wea. Rev., 125, 2074–2093, doi:10.1175/1520-0493(1997)125<2074:TROHAL>2.0.CO;2.
Orlanski, I., 1976: A simple boundary condition for unbounded hyperbolic flows. J. Comput. Phys., 21, 251–269, doi:10.1016/0021-9991(76)90023-1.
Rex, D., 1950: Blocking action in the middle troposphere and its effect upon regional climate. Tellus, 2, 196–211, doi:10.3402/tellusa.v2i3.8546.
Richards, P. I., 1956: Shock waves on highways. Oper. Res., 4, 42–51.
Shutts, G. J., 1983: Propagation of eddies in diffluent jetstreams: Eddy forcing of ‘blocking’ flow fields. Quart. J. Roy. Meteor. Soc., 109, 737–761, doi:10.1002/qj.49710946204.
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.
Swanson, K., P. Kushner, and I. Held, 1997: Dynamics of barotropic storm tracks. J. Atmos. Sci., 54, 791–810, doi:10.1175/1520-0469(1997)054<0791:DOBST>2.0.CO;2.
Treiber, M., and A. Kesting, 2013: Traffic Flow Dynamics: Data, Models and Simulation. Springer, 506 pp.
Tyrlis, E., and B. Hoskins, 2008: Aspects of a Northern Hemisphere blocking climatology. J. Atmos. Sci., 65, 1638–1652, doi:10.1175/2007JAS2337.1.
Wang, H., and J. Fyfe, 2000: Onset of edge wave breaking in an idealized model of the polar stratospheric vortex. J. Atmos. Sci., 57, 956–966, doi:10.1175/1520-0469(2000)057<0956:OOEWBI>2.0.CO;2.
Wang, L., and N. Nakamura, 2016: Covariation of finite-amplitude wave activity and the zonal-mean flow in the midlatitude troposphere. Part II: Eddy forcing spectra and the periodic behavior in the Southern Hemisphere summer. J. Atmos. Sci., 73, 4731–4752, doi:10.1175/JAS-D-16-0091.1.
Yamazaki, A., and H. Itoh, 2009: Selective absorption mechanism for the maintenance of blocking. Geophys. Res. Lett., 36, L05803, doi:10.1029/2008GL036770.
Zhu, D., and N. Nakamura, 2010: On the representation of Rossby waves on the β-plane by a piecewise uniform potential vorticity distribution. J. Fluid Mech., 664, 397–406, doi:10.1017/S002211201000457X.