Theoretical Foundation of the Linear Relationship between the Midlatitude Eddy Heat Flux and Fall-to-Spring Polar Ozone Buildup

Fumio Hasebe aFaculty of Environmental Earth Science, Hokkaido University, Sapporo, Japan

Search for other papers by Fumio Hasebe in
Current site
Google Scholar
PubMed
Close
,
Sayaka Kodera bNew Chitose Aviation Weather Station, Japan Meteorological Agency, Chitose, Japan

Search for other papers by Sayaka Kodera in
Current site
Google Scholar
PubMed
Close
, and
Hideharu Akiyoshi cNational Institute for Environmental Studies, Tsukuba, Japan

Search for other papers by Hideharu Akiyoshi in
Current site
Google Scholar
PubMed
Close
Open access

Abstract

It has been demonstrated that there is a globally unified linear relationship between the interannual variations of the fall-to-spring polar ozone accumulation and the winter-mean poleward eddy heat flux on the 100 hPa pressure surface. The foundation of this relationship is investigated using time-slice experiments on a chemistry–climate model with two levels of ozone-depleting substances (ODSs). The features of the transport field are interpreted by decomposing the horizontal ozone flux caused by the residual circulation into contributing processes including the eddy heat flux with the aid of the transformed Eulerian-mean momentum equation followed by rearrangement of terms. The linear relationship between the interannual variations of the fall-to-spring ozone buildup integrated poleward and above a reference point Pref on a meridional plane and the poleward eddy heat flux during the corresponding period at Pref is realized for each hemisphere implying that the interhemispheric unification should be treated with caution. This relationship is interpreted using the fact that the interannual variation of poleward ozone transport in the upper stratosphere is captured well by the vertical convergence of the constituent-based Eliassen–Palm (EP) flux (F˜), which is defined as the product of the constituent (ozone) mixing ratio and EP flux. The eddy momentum flux contributes to the meridional ozone transport in combination with the eddy heat flux in the form of the divergence of F˜, although it is not responsible for realizing the linear relationship. The dependence of the linearity on the location of Pref and ODS levels is discussed.

© 2023 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Fumio Hasebe, f-hasebe@ees.hokudai.ac.jp

Abstract

It has been demonstrated that there is a globally unified linear relationship between the interannual variations of the fall-to-spring polar ozone accumulation and the winter-mean poleward eddy heat flux on the 100 hPa pressure surface. The foundation of this relationship is investigated using time-slice experiments on a chemistry–climate model with two levels of ozone-depleting substances (ODSs). The features of the transport field are interpreted by decomposing the horizontal ozone flux caused by the residual circulation into contributing processes including the eddy heat flux with the aid of the transformed Eulerian-mean momentum equation followed by rearrangement of terms. The linear relationship between the interannual variations of the fall-to-spring ozone buildup integrated poleward and above a reference point Pref on a meridional plane and the poleward eddy heat flux during the corresponding period at Pref is realized for each hemisphere implying that the interhemispheric unification should be treated with caution. This relationship is interpreted using the fact that the interannual variation of poleward ozone transport in the upper stratosphere is captured well by the vertical convergence of the constituent-based Eliassen–Palm (EP) flux (F˜), which is defined as the product of the constituent (ozone) mixing ratio and EP flux. The eddy momentum flux contributes to the meridional ozone transport in combination with the eddy heat flux in the form of the divergence of F˜, although it is not responsible for realizing the linear relationship. The dependence of the linearity on the location of Pref and ODS levels is discussed.

© 2023 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Fumio Hasebe, f-hasebe@ees.hokudai.ac.jp

1. Introduction

Early attempts to describe the stratospheric general circulation by Brewer (1949) and Dobson (1956) were intended to explain the observed spatiotemporal distribution of trace gases. Although the speculated circulation, the Brewer–Dobson circulation (BDC), appeared plausible (Dütsch 1969), the mechanism of the spring maximum of total ozone, among others, remained unexplained, and the schematic model was criticized from the dynamical point of view by those who emphasized the contribution of stratospheric eddy processes (e.g., Newell 1961). The ozone transport features, investigated later by using general circulation models (GCMs) (e.g., Cunnold et al. 1975), showed that the stratospheric mean meridional circulation was characterized by ascending motion over the winter pole as well as in the summer hemisphere and descending flow in midlatitudes of the winter hemisphere. The upwelling in the high-latitude winter hemisphere was exactly the opposite of the downward circulation illustrated by Dütsch (1969) passing through the ozone maximum that was supposed to have resulted from stratospheric subsidence. Such an apparent conflict was explained by the near cancellation of the transport owing to mean flow and that from the eddy fluxes, and the residual of these two was responsible for creating the lower stratospheric ozone maximum. The conflict was soon resolved by Kida (1977a,b), who demonstrated that the Eulerian-mean meridional circulation shown by Cunnold et al. (1975) and the Lagrangian-mean circulation as in Brewer (1949) emerged from different averaging procedures of the same transport field.

A new theoretical framework for describing the circulation field was proposed by Andrews and McIntyre (1976). It relied on the Eliassen–Palm flux (EP flux) to define the transformed Eulerian-mean (TEM) circulation. The specific features of the TEM framework are that the eddy heat and momentum fluxes do not act separately but function in combination as a divergence of EP flux, and that the time-averaged residual circulation approximates the diabatic circulation and, thus, Lagrangian-mean circulation (Dunkerton 1978). The introduction of the TEM framework led to a paradigm shift. Haynes et al. (1991) proposed the “downward control mechanism” in which the BDC is driven by the convergence of EP flux owing to planetary Rossby waves and gravity waves that are excited in the troposphere and propagate into the stratosphere. The tropical ascending motion and the high-latitude descending flow result from mass conservation compensating for the stratospheric poleward flow driven in midlatitudes (Holton et al. 1995). The downward control principle also applies to the global-scale ozone transport. The spring maximum of total ozone is brought about by wintertime enhancement of the poleward ozone transport driven by breaking planetary waves; interhemispheric asymmetry in the ozone distribution is caused by more enhanced orographic excitation and subsequent propagation and breaking of the planetary waves in the Northern Hemisphere (NH) than in the Southern Hemisphere (SH).

Randel (1993) took the eddy heat flux at 60°N on 50 hPa during winter as a measure of the high-latitude winter wave activity and showed that the ozone increase in the high-latitude lower stratosphere corresponds well to the eddy heat flux. Fusco and Salby (1999) discussed the interannual variation of total ozone in terms of the stratospheric activity of breaking planetary waves. They showed that the vertical component of EP flux at 100 hPa averaged over the winter hemisphere in January corresponds well to the extratropical ozone increase. Newman et al. (2001) showed that the residual vertical velocity, averaged from a midlatitude reference latitude ϕr to the pole, is expressed by the eddy heat flux at ϕr on the corresponding pressure level. Based on this formulation, they showed high correlation between the eddy heat flux averaged from 45° to 75°N on 100 hPa and average temperature northward of 60°N on 50 hPa during the period from January to February. Randel et al. (2002) found high correlations between monthly mean eddy heat flux on 100 hPa and the total ozone tendency on several occasions such as the latitude ranges from 40° to 50°N in January and from 40° to 50°S in June.

The correspondence between the midlatitude eddy heat flux and the polar ozone buildup was generalized to a unified view without distinguishing between the NH and the SH. Weber et al. (2003, 2011) found a compact linear relationship between the wintertime midlatitude 100-hPa eddy heat flux and spring-to-fall ratio of polar (50°–90° latitudes) total ozone [see WMO (2018) for the update up to 2017]. Hereinafter, we refer to this as the Weber plot. Such a unified relationship is surprising in view of the interhemispheric contrast in the stratospheric transport field and the well-developed ozone hole in the Antarctic initiated by heterogeneous reactions on the surface of polar stratospheric clouds (PSCs). As extremely cold temperature below a threshold is required for PSC formation, it is plausible that the interannual variations of planetary wave activity have a nonlinear effect on the polar ozone amount in spring.

The purpose of the present study is to answer the following specific questions by using the meteorological fields simulated by a chemistry–climate model (CCM): Why is the eddy heat flux on a single pressure level (such as 100 hPa) highly correlated with the ozone amount in the polar stratosphere, and what is the theoretical foundation of the interhemispheric linear relationship revealed by Weber et al. (2003, 2011)? This paper is organized as follows. The reproducibility of the Weber plot is described together with the simulated meteorological fields in section 2. Section 3 explores the dynamical foundation of the Weber plot. The assumptions introduced to derive the governing equations are examined using the CCM ensemble simulations in section 4. The features of the transport field obtained from the ensemble runs are applied to interpret the linearity of the Weber plot in section 5. Section 6 discusses the transport problem and the effect of ozone chemistry, focusing on the linear relationship. Finally, the findings are summarized in section 7.

2. Time-slice ensemble experiments with a chemistry–climate model

Numerical simulations with the aid of CCMs have been conducted for studies of the stratosphere in which the interaction between dynamics and chemistry is crucially important. Weber et al. (2011) employed CCMs in the studies of the spring-to-fall ratio of polar total ozone dividing the period into the ages of low (from 1960 to 1985), maximum (from 1985 to 2010), and decreasing halogen loading (from 2010 to 2050). However, scenario experiments of this type are not ideal for studying the foundation of the linear relationship because the gradual shift of external forcing may obscure essential features of the phenomenon. In the present study, we use the CCM developed from the CCSR/NIES Atmospheric General Circulation Model (AGCM) to conduct time-slice experiments under fixed concentrations of greenhouse gas (GHG) and ozone-depleting substances (ODS). These are a part of the MIROC3.2 chemistry–climate model experiments (Akiyoshi et al. 2016) conducted under the Chemistry–Climate Model Initiative (CCMI). The AGCM is configured with T42 spectral truncation and 34 vertical layers with the top around 80 km. The present study deals with two sets of 500-yr experiments with the ODS fixed at the 1960 level (ODS1960) and 2000 level (ODS2000); the GHG concentration is fixed at the 2000 level (GHG2000) in all experiments. The sea surface temperature and sea ice extent, both given as boundary conditions, are taken from the products of air–sea interactive global warming experiments from CMIP5 MIROC-ESM under the GHG2000 condition. Simulations initialized on 1 January 2000 are continued for 510 years. The first 10 years are taken as a spinup period, and each of the 500 years from the 11th to 510th is regarded as a sample from 500 ensemble members.

The ensemble-mean time–latitude sections of the zonal-mean total ozone for ODS1960 and ODS2000 conditions are shown in the top two panels of Fig. 1. The northern midlatitude maximum attains its largest value in late February and March, and the phase of the maximum propagates to the north and reaches the Arctic in April to May in ODS1960 and a little later in May to June in ODS2000. The maximum in the southern midlatitudes is almost stationary at around 50°S. The springtime polar minimum under ODS1960 as well as ODS2000 is the most significant difference from the NH. The boundary is located at 60°S, where a steep latitudinal gradient develops in September and October associated with the formation of the polar vortex in winter to spring. The bottom panel of Fig. 1 shows the corresponding section derived from the European Centre for Medium-Range Weather Forecasts interim reanalysis (ERA-Interim; Dee et al. 2011). The averaging for 10 years (1995–2004), covering the sustained ODS maximum around the year 2000 (WMO 2018), is applied to compare with the ODS2000 simulation. There are noticeable differences in that the polar maxima during April (August) over the Arctic (Antarctic) in the ERA-Interim section are not reproduced in our CCM. This, together with the larger-than-observed ozone amount in the tropics, suggests that the poleward ozone transport is weak in the CCM. Fortunately, this deficiency will not be so serious for our purpose investigating the mutual relationship between the poleward transport across a latitude circle such as 50° and the accumulated ozone amount averaged in the whole region poleward of that latitude during fall to spring.

Fig. 1.
Fig. 1.

Seasonal variation of the zonal-mean total ozone (DU) obtained by averaging 500 ensemble members simulated by CCM under the conditions of ozone-depleting substances (ODS) at (top) 1960 levels (ODS1960) and (middle) 2000 levels (ODS2000) compared with that from (bottom) the ERA-Interim 10-yr average over 1995–2004.

Citation: Journal of the Atmospheric Sciences 80, 3; 10.1175/JAS-D-22-0023.1

Figure 2 shows the latitude–height sections of ensemble-mean ozone number density with the residual streamfunctions superposed for January, March, July, and September of ODS1960 (left column) and the difference between those for two ODS levels (ODS2000 minus ODS1960; right column). The seasonal march of the streamfunction is quite similar in ODS1960 and ODS2000. The similarity is also confirmed in the distributions of zonal-mean temperature, zonal wind, EP flux and its divergence (not shown). Some differences are noticeable, however, in the residual streamfunction indicating a strengthening of the BDC in the SH in January (top right). This is consistent with the increasing trend from 1960 to 2000 of the tropical upward mass flux at 70 hPa in December–February demonstrated in the studies of the dynamical effects of ozone depletion by McLandress et al. (2010). The ozone distribution, on the other hand, exhibits appreciable negative anomalies under ODS2000 especially in the springtime polar lower stratosphere.

Fig. 2.
Fig. 2.

Latitude–height sections of zonal-mean ozone number density (shade in green, ×1012 cm−3) superposed by the residual streamfunction with red (blue) contours for positive (negative) values at ±(0.1, 0.2, 0.3, 0.4, 0.5, 0.7, 1, 2, 4, 8, 15, 30, 50) (×102 kg m−1 s−1). The 500-member ensemble means for (left) ODS1960 and (right) difference between those for ODS1960 and ODS2000 simulations (ODS2000 minus ODS1960) for (top to bottom) January, March, July, and September. The difference field of the residual streamfunction is supplemented with dashed contours at ±(0.01, 0.02, 0.05) (×102 kg m−1 s−1).

Citation: Journal of the Atmospheric Sciences 80, 3; 10.1175/JAS-D-22-0023.1

The ODS dependence of the polar stratospheric ozone and temperature are shown as the time–height sections of the difference (ODS2000 minus ODS1960) in the area-weighted ensemble-mean values in Fig. 3. The polar region is defined as the area poleward of 50° latitude following Weber et al. (2011). Negative values of ozone perturbations seen in the upper stratosphere are brought about by enhanced ozone depletion due to gas-phase chemistry, whereas those in the lower stratosphere in spring are caused by halogen activation due to heterogeneous reactions on PSC surfaces under ODS2000. The temperature perturbations in the upper stratosphere appear as negative values that are more pronounced in summer. Those associated with heterogeneous chemistry appear a little later and at slightly lower altitudes than the enhanced ozone depletion in both Arctic and Antarctic regions with much greater amplitude in the Antarctic.

Fig. 3.
Fig. 3.

(top) Time–height sections of the difference between ODS1960 and ODS2000 conditions (ODS2000 minus ODS1960) in polar ozone mixing ratio (ppmv) area averaged in the region poleward of 50° latitude for the (left) Northern and (right) Southern Hemispheres. (bottom) As in the top row, but for the polar temperature (K).

Citation: Journal of the Atmospheric Sciences 80, 3; 10.1175/JAS-D-22-0023.1

Figure 4 compares the Weber plots obtained by two time-slice experiments with that produced using ERA-Interim from 1995 to 2010 as in Weber et al. (2011). The regression lines are drawn separately for the NH (black) and SH (red) in addition to that combining both hemispheres (blue). The statistics such as the slope of the linear fit are summarized in Table 1. We can see that the slope is more inclined, the magnitude of midlatitude eddy heat flux is larger, and the spring-to-fall ratio of polar total ozone is greater in the NH than in the SH in all three cases. The regression line obtained by combining data from both hemispheres shows a larger slope than those separately drawn for the NH and SH, and appears more like a line connecting the center of population of each hemisphere. The high correlation coefficient (>0.97) obtained by combining two hemispheres should be interpreted with caution as it results largely from the elongated distribution of data points making the covariance between different hemispheres (interclass variation) larger than that within each hemisphere (intraclass variation). The slope of the regression lines steepens, the value of the y intercept decreases, and the magnitude of the spring-to-fall ratio of polar total ozone reduces in ODS2000 especially in the SH. Interhemispheric differences such as the slope and y intercept of regression lines, the correlation coefficient, magnitude of eddy heat flux, and the spring-to-fall ratio in the ERA-Interim analysis correspond qualitatively well to those in CCM experiments with the values closer to those from ODS2000, which appears reasonable from the viewpoint of ODS dependence. All these features indicate that our CCM simulations are applicable to the investigation of the linear relationship of the Weber plot.

Fig. 4.
Fig. 4.

Scatterplots between the midlatitude winter-mean eddy heat flux at 100 hPa (K m s−1) and spring-to-fall ratio of polar total ozone derived from (top) ERA-Interim, and CCM under (middle) ODS1960, and (bottom) ODS2000. The analysis period of ERA-Interim is the same as that of Weber et al. (2011): from September 1995 (fall) to March 2010 (spring) for the Northern Hemisphere and from March 1996 (fall) to September 2010 (spring) for the Southern Hemisphere.

Citation: Journal of the Atmospheric Sciences 80, 3; 10.1175/JAS-D-22-0023.1

Table 1

Statistics for the scatterplots between midlatitude winter-mean eddy heat flux on 100 hPa and spring-to-fall ratio of polar total ozone. See Fig. 4 for the analysis and simulation conditions.

Table 1

3. Dynamical foundation of the linear relationship

This section explores the dynamical foundation of the linear relationship of the Weber plot. Examination of the validity of assumptions made during the derivation is postponed to section 4.

a. Governing equations

Consider the TEM continuity equation for a tracer with mixing ratio χ in log-pressure coordinates (Andrews et al. 1987) :
χ¯t+υ¯*aχ¯ϕ+w¯*χ¯z=S¯+1ρ0M.
All symbols and notations follow Andrews et al. (1987) except when stated otherwise. Here, the residual velocities (0,υ¯*,w¯*),
υ¯*υ¯1ρ0z(ρ0υθ¯θ¯z),
w¯*w¯+1acosϕϕ(cosϕυθ¯θ¯z),
satisfy the continuity equation
1acosϕϕ(υ¯*cosϕ)+1ρ0z(ρ0w¯*)=0,
and S and M in Eq. (1) represent chemical production/destruction and the eddy fluxes of tracer, respectively (see, e.g., Andrews et al. 1987; Brasseur et al. 1999).
Equation (1) is converted to flux form with the aid of Eq. (4):
χ¯t+1acosϕϕ(χ¯υ¯*cosϕ)+1ρ0z(ρ0χ¯w¯*)=S¯+1ρ0M.
If we denote total ozone expressed by the number of ozone molecules per unit area of Earth’s surface as X, and divide it into the sum of the amount integrated in the troposphere Xtr and that in the stratosphere and above Xst,
X¯stt=tzpn3¯dz=tzpnχ¯dz,
where n3 and n are number densities of ozone and air molecules, respectively, and zp is the tropopause height. Later in this paper, zp is taken as a time- and latitude-independent parameter that represents the altitude of the lower boundary for defining polar stratospheric ozone. By denoting the mean mass of air molecules as md, n = ρ/mdρ0/md leading to
mdX¯sttzpρ0χ¯tdz.
Because ρ0 → 0 as z → ∞, vertical integration of Eq. (5) from zp to the top of the atmosphere after multiplying both sides of the equation by ρ0 leads to
mdX¯stt+1acosϕzpϕ(ρ0χ¯υ¯*cosϕ)dz(ρ0χ¯w¯*)z=zp=zpρ0S¯dz+zpMdz.
Denoting the area integral from latitude ϕm to the pole (ϕp) of a quantity A in angle brackets,
A=2πa2ϕmϕpA¯cosϕdϕ,
the area integration of Eq. (8) from ϕm to ϕp after multiplying both sides of the equation by 2πa2 cosϕ reduces to
mdXstt=2πacosϕmzpρ0(χ¯υ¯*)ϕ=ϕmdz+2πa2ϕmϕp(ρ0χ¯w¯*)z=zpcosϕdϕ+zpρ0Sdz+2πa2ϕmϕp(zpMdz)cosϕdϕ.
This equation states that the ozone tendency integrated poleward and above a reference point Pref(ϕm, zp) [left-hand side (lhs)] is driven by the poleward ozone flux crossing latitude ϕm due to υ¯* (first term), the downward ozone flux crossing the tropopause poleward of ϕm due to w¯* (second term), the chemical source and sink (third term), and the divergence of eddy ozone flux (fourth term) integrated in the polar stratosphere. The appropriate position of Pref for realizing the linear relationship of the Weber plot will be sought in section 4.

b. Poleward ozone transport expressed by eddy heat flux

To introduce eddy heat flux into the right-hand side (rhs) of Eq. (10), we use the zonal momentum equation (Andrews et al. 1987):
u¯t+υ¯*[1acosϕϕ(u¯cosϕ)f]+w¯*u¯z=1ρ0acosϕF+Y.
Here, F is the divergence of EP flux F = (0, F(ϕ), F(z)):
F(ϕ)=ρ0acosϕ(u¯zυθ¯θ¯zuυ¯),
F(z)=ρ0acosϕ{[f1acosϕϕ(u¯cosϕ)]υθ¯θ¯zuw¯},
F=1acosϕϕ(F(ϕ)cosϕ)+F(z)z,
and Y is the force due to unresolved small-scale eddies. Because the terms containing ϕ and z derivatives of u¯ are relatively small in the lhs of Eq. (11), the poleward ozone flux due to υ¯* is approximately given by
ρ0χ¯υ¯*=1facosϕ[χ¯acosϕϕ(F(ϕ)cosϕ)+χ¯F(z)z]+ρ0χ¯fu¯tρ0χ¯fY.
This equation describes the decomposition of the poleward ozone flux due to υ¯* (lhs) into terms representing the latitudinal and vertical convergences of EP flux multiplied by χ¯ (first and second terms of the rhs, respectively), a term associated with deceleration of the zonal wind (third term), and a term attributed to unresolved eddies in the zonal momentum equation [Eq. (11)] multiplied by χ¯ (fourth term).
As the eddy heat flux term containing υθ¯ is the major component of F(z), the second term of the rhs of Eq. (15) will be expressed by eddy heat flux if it is integrated by parts. Then vertical integration of both sides of this equation from zp to infinity will lead to the following if F(ϕ) is also expressed by its main component uυ¯:
zpρ0χ¯υ¯*dz=1fcosϕzpρ0χ¯acosϕϕ(uυ¯cos2ϕ)dz+(ρ0χ¯υθ¯θ¯z)z=zp+zpρ0χ¯zυθ¯θ¯zdz+1fzpρ0χ¯u¯tdz1fzpρ0χ¯Ydz.
We can approximate the poleward ozone flux by υ¯* by the eddy heat flux at z = zp if the second term is dominant in the rhs of this equation.
Alternatively, the eddy heat flux term is introduced, upon vertical integration such as that in Eq. (16), by multiplying both sides of Eq. (11) by χ¯ and rewriting χ¯F in terms of (χ¯F). In this article, we define F˜=(0,F˜(ϕ),F˜(z)) as
F˜χ¯F=(0,χ¯F(ϕ),χ¯F(z)),
and call it the constituent-based EP flux (C-EP flux) for the sake of simplicity, rather than using “EP flux multiplied by χ¯” each time it is referred to. Then, Eq. (15) is rewritten as
ρ0χ¯υ¯*=1facosϕ[1acosϕϕ(F˜(ϕ)cosϕ)+F˜(z)z]+ρ0χ¯fu¯tρ0fY˜,
where Y˜ includes all extra terms, in addition to χ¯Y, that arise from the above rewrite. Vertical integration of Eq. (18) above zp gives
zpρ0χ¯υ¯*dz=1facosϕzp[1acosϕϕ(F˜(ϕ)cosϕ)]dz+1facosϕ(F˜(z))z=zp+1fzpρ0χ¯u¯tdz1fzpρ0Y˜dz.
By retaining the leading terms of F˜(ϕ) and F˜(z), we have
zpρ0χ¯υ¯*dz=1fcosϕzp[ρ0acosϕϕ(χ¯uυ¯cos2ϕ)]dz+(ρ0χ¯υθ¯θ¯z)z=zp+1fzpρ0χ¯u¯tdz1fzpρ0Y˜dz.
This is merely a slight modification of Eq. (16), but the eddy heat flux on a single pressure level is introduced (instead of integration by parts) by the vertical integral of a quantity over the altitudes covering the whole stratosphere.
Now the path to the Weber plot is straightforward. If we assume that the horizontal flux term is dominant on the rhs of Eq. (10),
mdXstt=2πacosϕmzpρ0(χ¯υ¯*)ϕ=ϕmdz,
and that the second term (eddy heat flux term) is prevailing among others on the rhs of Eqs. (16) and (20), the poleward ozone flux due to υ¯* may be approximated by
zpρ0χ¯υ¯*dz=(ρ0χ¯υθ¯θ¯z)z=zp.
Then, the tendency of the integrated polar stratospheric ozone is approximated by the eddy heat flux at Pref:
mdXstt=2πacosϕm(ρ0χ¯υθ¯θ¯z)ϕ=ϕm,z=zp.
If the tropospheric ozone variation is so small that the tendency of stratospheric ozone 〈Xst〉 can be replaced by that of total ozone 〈X〉, the approximate form of the spring-to-fall ratio of polar total ozone is obtained by integrating Eq. (23) from fall to spring:
XspringXfall=1+2πacosϕmmdXfallfallspring(ρ0χ¯υθ¯θ¯z)ϕ=ϕm,z=zpdt.
This states that the spring-to-fall ratio of total ozone integrated poleward of Pref and the eddy heat flux integrated from fall to spring at Pref are linearly related as long as the time change of χ¯ during the period is small. This will be the theoretical foundation for the linearity of the Weber plot. Note that ϕm was taken as 50° for the estimation of total ozone, whereas the areal average between 45° and 75° was used for the calculation of eddy heat flux in Weber et al. (2011).

4. Examination of the governing equations

The assumptions having been made in the previous section need to be validated. Those applied to derive Eq. (21) are examined in section 4a. Those related to Eq. (23) are investigated, together with the vertical profile of transport terms constituting the ozone flux due to υ¯*, in section 4b. The results are applied to interpret the linearity of the Weber plot in section 5.

a. Approximation by horizontal flux

Here, we compare the time series of the four terms on the rhs of Eq. (10) to try to justify the approximation to Eq. (21). The latitude ϕm and altitude zp are temporarily set to 50° and 100 hPa, respectively, referring to Weber et al. (2011). The results from other choices are examined when necessary.

Figure 5 shows the ensemble-mean time series of each term on the rhs of Eq. (10). The time series of the first (black, denoted H.F), second (blue, V.F), and third (green, S) terms are calculated from model outputs, whereas in the present study we do not calculate M explicitly as was done by Randel et al. (1994). Instead, the fourth term (red, Res) is estimated as the residual obtained by subtracting the above three terms from the lhs of the equation, assuming negligible errors in the estimated terms. We readily confirm that the vertical flux crossing the tropopause zp (=100 hPa) is small in all cases. In the NH (top panels), the poleward ozone flux crossing the latitude ϕm (black) is the dominant term year-round, increasing from a minimum in summer to a maximum in January. This annual cycle is explained as the seasonal dependence of the planetary wave breaking and/or wave transience that take a maximum in winter. There is little difference between ODS1960 (left) and ODS2000 (right), except for an enhancement of chemical ozone loss (green) during winter to spring under ODS2000.

Fig. 5.
Fig. 5.

Tendency of polar stratospheric ozone above zp = 100 hPa integrated poleward of latitude ϕm = 50° (×108 DU km2 day−1) as shown by time series of 500-member ensemble-mean values on the rhs of Eq. (10). Black, blue, green, and red lines are the first, second, third, and fourth terms, respectively. Vertical bars are the spread (standard deviation) drawn by thinning out by 5 days for visual clarity.

Citation: Journal of the Atmospheric Sciences 80, 3; 10.1175/JAS-D-22-0023.1

In the SH (bottom), the poleward ozone flux owing to the residual circulation (black) is dominant in most of the year, as in the NH. However, the flux is suppressed in winter, creating dual maxima in fall (March) and spring (November). What is distinct from the features in the NH is the increased eddy transport (fourth term, red) surpassing the poleward flux by the residual circulation from October to December. However, this eddy transport is nearly canceled by simultaneously activated chemical ozone loss. Details of this phenomenon are analyzed in section 6b. Owing to this cancellation, the polar ozone tendency described by Eq. (10) is approximated by the horizontal flux (first term, H.F) also in the SH.

b. Approximation by eddy heat flux

For the validation of Eq. (23), we investigate if the poleward ozone flux crossing the latitude ϕm above the altitude zp is approximated by the eddy heat flux at Pref(ϕm, zp). Specifically, we examine whether or not Eq. (22) is a good approximation of Eqs. (16) and (20).

The top panels of Fig. 6 show the poleward ozone flux [lhs of Eq. (16); black, denoted H.F] decomposed into the latitudinal divergence of EP flux (blue, Fy), the first (red, Fz1) and the second (light blue, Fz2) terms originating from integration by parts of χ¯(F(z)/z), the component associated with zonal wind deceleration (orange, Ut), and the transport attributed to unresolved eddies (green, Res) under ODS1960. Those of ODS2000 are omitted as the dynamical fields are essentially similar to those of ODS1960 (section 2). We can see that the eddy heat flux term (red, Fz1) shows some enhancement in winter in the NH, although it does not constitute the major component of the poleward ozone flux (black, H.F) being smaller than the second term (light blue, Fz2) that arises from integration by parts.

Fig. 6.
Fig. 6.

(top) Time series of poleward ozone flux (×108 DU km2 day−1) expressed by Eq. (16). The 500-member ensemble-mean values under ODS1960 of the horizontal flux above Pref(ϕm, zp) (H.F; black), the latitudinal convergence of EP flux (Fy; blue), the eddy heat flux at Pref (Fz1; red), the additional term coming from integration by parts (Fz2; light blue), the zonal wind deceleration (Ut; orange), and the frictional term estimated as the residue (Res; green). (bottom) As in the top row, but for those expressed by Eq. (20): the horizontal flux (H.F; black), the latitudinal (F˜y; blue) and vertical (F˜z; red) convergences of F˜, the zonal wind deceleration (U˜t; orange), and the residual term (Res; green). ϕm and zp are taken to be 50° and 100 hPa, respectively. Vertical bars are the spread (standard deviation) drawn by thinning out by 5 days for visual clarity.

Citation: Journal of the Atmospheric Sciences 80, 3; 10.1175/JAS-D-22-0023.1

The bottom panels of Fig. 6 are the same as the top, but for those obtained from Eq. (20). There is scarcely any difference if Eq. (19) is used. Note that the first three terms on the rhs of Eq. (20) representing latitudinal and vertical convergences of F˜ (blue, F˜y, and red, F˜z) and zonal wind deceleration (orange, U˜t) are nearly the same as the first (blue, Fy), second (red, Fz1), and fourth (orange, Ut) terms of Eq. (16), respectively, meaning that the fourth term (vertical integral of ρ0Y˜) estimated as the residue in Eq. (20) roughly corresponds to the sum of the second term from integration by parts (light blue, Fz2) and the unresolved eddy term (green, Res) in Eq. (16). In the NH (bottom left), the poleward and equatorward ozone fluxes driven by the vertical (red, F˜z) and latitudinal (blue, F˜y) convergences, respectively, of F˜ tend to cancel each other, and the residual term (green, Res) is the major contributor to the poleward ozone flux due to υ¯*. In the SH (bottom right), the vertical convergence of F˜ (red, F˜z) is generally small throughout the year with a slight enhancement in winter, whereas the latitudinal convergence (blue, F˜y), contributing to the equatorward transport, is activated in winter to spring. The enhanced equatorward transport associated with the zonal wind deceleration (orange, U˜t) in spring reduces the poleward transport carried by the residual term (green, Res). Overall, we can see that the vertical integral of F˜(z) above zp, that is, the eddy heat flux at the altitude zp, is not a good approximation to the poleward ozone flux but that the term associated with Y˜, estimated as the residue, appears a better proxy for the meridional ozone flux on the seasonal time scale.

Further investigation is made on the vertical profiles of the quantities discussed above. Figure 7 shows the time–height section of each term of Eq. (15) at 50°N (left) and 50°S (right) under ODS1960. No appreciable difference is found for ODS2000, and similar features are confirmed from ERA-Interim analysis (not shown). Top panels show the poleward ozone flux due to υ¯* (lhs) indicating a stratospheric enhancement in winter and a persistent ozone transport at around 100 hPa. This spatiotemporal structure suggests the contribution by planetary wave driving for the former and synoptic waves and quasi-horizontal isentropic eddy mixing for the latter (e.g., Plumb 2002). Latitudinal convergence of EP flux (Fy), Figs. 7c and 7d, contributes to equatorward ozone transport in the lower stratosphere all year round and in the middle to upper stratosphere in winter. The poleward ozone transport is brought about by the vertical convergence of EP flux (Fz), Figs. 7e and 7f, exhibiting a similar structure to Figs. 7a and 7b, respectively, but with reduced magnitude. The zonal wind acceleration term (Ut), Figs. 7g and 7h, contributes to equatorward transport in spring. The dominant term is associated with unresolved eddies (Res), Figs. 7i and 7j. Because this term is obtained as the residue of all other terms, it also includes both physical and numerical diffusion and a variety of errors arising from transformation from spectral to gridpoint data and others. Model dependence especially on the parameterization of unresolved processes is a problem left for future studies.

Fig. 7.
Fig. 7.

Time–height sections of the poleward ozone flux (×10−9 kg m−2 s−1) at (left) 50°N and (right) 50°S described by Eq. (15). (a),(b) The ozone flux due to υ¯* is decomposed into the (c),(d) latitudinal and (e),(f) vertical convergences of EP flux, (g),(h) zonal wind acceleration, and (i),(j) the residual including those associated with small-scale eddies Y. Poleward (equatorward) flux is color-coded in red (blue). 500-member ensemble mean for ODS1960.

Citation: Journal of the Atmospheric Sciences 80, 3; 10.1175/JAS-D-22-0023.1

Figure 8 is the same as Fig. 7 except that those constituting Eq. (18) are shown with the terms contained in Y˜ expanded:
ρ0χ¯υ¯*=1facosϕ[1acosϕϕ(F˜(ϕ)cosϕ)+F˜(z)z]+1facosϕ(F(ϕ)aχ¯ϕ+F(z)χ¯z)+ρ0χ¯fu¯tρ0χ¯fY.
We can see that the latitudinal divergence of F˜ (F˜y), first term on the rhs of Eq. (25), Figs. 8c and 8d, is quite similar to that of the EP flux (Figs. 7c,d) and the contribution of the term proportional to χ¯/ϕ, third term denoted F × χy shown in Figs. 8k and 8l, is relatively small. The vertical convergence of F˜ (F˜z), second term shown in Figs. 8e and 8f, contributes to poleward ozone transport in the stratosphere above 20 hPa during fall to spring, whereas it is directed equatorward in the lower stratosphere except for the summer season, contributing to the transport opposite to the vertical convergence of the EP flux (Figs. 7e,f). This lower-stratospheric equatorward transport results from the positive vertical gradient of ozone mixing ratio and is compensated for by the poleward transport by the term proportional to χ¯/z (fourth term, F × χz) of Eq. (25), as can be seen from Figs. 8m and 8n.
Fig. 8.
Fig. 8.

As in Fig. 7, but those expressed by Eq. (25). Figures 7a and 7b are not replicated. (k),(l) The third term and (m),(n) the fourth term.

Citation: Journal of the Atmospheric Sciences 80, 3; 10.1175/JAS-D-22-0023.1

The fact that F˜ is divergent in the lower stratosphere while it is convergent in the upper stratosphere suggests that the eddy heat flux at 100 hPa is not a good proxy for the poleward ozone flux in the whole stratosphere and that the choice of 100 hPa for zp in constructing the Weber plot is too low to best express the fall-to-spring polar ozone buildup by the midlatitude eddy heat flux. On the other hand, higher values of zp will make the correspondence to total ozone weaker. Time series similar to those shown in the bottom panels of Fig. 6 but for zp taken as 50 hPa are shown in the top panels of Fig. 9. We can see some improvement from Fig. 6 (bottom) especially in the NH. That is, the contribution of the vertical convergence of F˜ is larger than that in the case of 100 hPa, and the seasonal enhancement during winter to spring is synchronized with the increase of poleward ozone transport due to υ¯*. If we want to make the contribution larger still, the choice of ϕm = 60° in addition to the change to zp = 50 hPa (bottom panels) is effective. In this case, the poleward ozone flux owing to vertical convergence of F˜ (red, F˜z) surpasses the total flux (black, H.F) numerically. This is made possible because the latitudinal convergence of F˜ (blue, F˜y) drives a strong equatorward ozone flux. As the latitudinal convergence of C-EP flux has nothing to do with the Weber plot, the choice of ϕm = 60° and zp = 50 hPa may improve the appearance of the linear relationship.

Fig. 9.
Fig. 9.

As in Fig. 6 (bottom), but (ϕm, zp) is (top) (50°, 50 hPa) and (bottom) (60°, 50 hPa).

Citation: Journal of the Atmospheric Sciences 80, 3; 10.1175/JAS-D-22-0023.1

5. Interpretation of the linear relationship

The results obtained above demonstrate that the fall-to-spring ozone accumulation poleward of Pref(ϕm, zp) is approximated by the horizontal ozone flux due to υ¯* vertically integrated above Pref (section 4a) and that the eddy heat flux at Pref may be used as a proxy for the poleward ozone flux if zp is taken to be 50 hPa [higher than the 100 hPa as in Weber et al. (2011)], although it is not the dominant contributor to the poleward ozone flux due to υ¯* integrated above Pref (section 4b). This reflects the fact that the transport driven by the vertical convergence of F˜ is poleward in the upper stratosphere, whereas it is equatorward in the lower stratosphere (Fig. 8). These features, characterizing the seasonally varying climatology of the contributing processes, are applied to interpret the linearity of the Weber plot, which illustrates the relationship between the interannual variations of the fall-to-spring integral of the lhs and rhs of Eq. (23) in the form of scatterplots. In doing so, the latitudinal averaging of the eddy heat flux between 45° and 75° (Weber et al. 2011) is not applied, and the accumulated raw ozone amount, instead of the spring-to-fall ratio, is analyzed. For the NH (SH), we take 1 September (1 March) as the start date in fall, denoted by t0, and 31 March (30 September) as the end date in spring (t1).

a. Description by horizontal flux

Integrating Eq. (21) from t0 to t1 for a given set of ϕm and zp, we have
md(Xstt1Xstt0)=2πacosϕmt0t1zpρ0(χ¯υ¯*)ϕ=ϕmdzdt.
The interannual variations of the variables on the lhs and rhs of the equation are shown as the scatterplots in Fig. 10. If we take ϕm = 50° and zp = 100 hPa (top row), the scatterplots obtained from CCM simulations show a well-aligned linear relationship between the two variables irrespective of the ODS levels. The fact that the slope of the regression line is less than 1 means that the amount of polar stratospheric ozone does not fully increase in response to the transported amount. As the downward flux across the bottom boundary is small (V.F shown in blue in Fig. 5), we expect that a part of the transported ozone is lost by chemical reactions and small-scale mixing. The importance of chemical ozone loss is also recognized by the smaller averages in ODS2000 than in ODS1960. As the standard deviation of ozone accumulation along the ordinate (y) is larger in ODS2000 than in ODS1960 [e.g., 0.5 for ODS1960 and 0.7 for ODS2000 (109 DU km2) in the NH], whereas that of the poleward ozone flux along the abscissa (x) does not change appreciably [e.g., 1.4 for ODS1960 and 1.3 for ODS2000 (109 DU km2) in the NH], the slope of the regression line is larger in ODS2000. The finding that the interhemispheric unified structure still holds under ODS2000 (Fig. 10b) suggests that the springtime polar ozone depletion, more pronounced in the Antarctic stratosphere (Fig. 3), has a limited effect, whereas the gas-phase ozone chemistry is important (section 6c). The results from ERA-Interim (Fig. 10c) do not show a unified structure and the regression line for the SH strongly depends on the value for 2002; the scatterplot shows no correlation if the year 2002 is excluded from the statistics.
Fig. 10.
Fig. 10.

Scatterplots of polar stratospheric ozone accumulation against the fall-to-spring integral of the poleward ozone flux across latitude ϕm above altitude zp [Eq. (26), (×109 DU km2)]. (top to bottom) (ϕm, zp) = (50°, 100 hPa), (50°, 50 hPa), and (60°, 50 hPa) under (left) ODS1960 and (center) ODS2000 CCM simulations, and (right) ERA-Interim from 1996 to 2010. Northern Hemisphere is in black and Southern Hemisphere is in red. Dashed red lines are derived by excluding the year 2002 that lies far to the right along the abscissa. Calculations are made referring to daily mean values on log-pressure coordinates taking the integration period [t0, t1] from 1 Sep to 31 Mar of the next year in the NH and from 1 Mar to 30 Sep for the SH.

Citation: Journal of the Atmospheric Sciences 80, 3; 10.1175/JAS-D-22-0023.1

If 50 hPa is chosen for zp (Fig. 10, middle row), the sensitivity to ODS loading becomes higher in the SH (Fig. 10e); the ozone accumulation is negative and the regression line is no longer aligned with that of the NH. This is due to the increase of the slope associated with the decrease of the average and the increase of standard deviation (from 0.6 ± 0.2 to −2.2 ± 0.3) of the ozone accumulation; there is little change in the poleward ozone flux. In the ERA-Interim case (Fig. 10f), the polar stratospheric ozone accumulation becomes almost insensitive to the poleward ozone flux in the NH, and it is also true in the SH if the year 2002 is excluded. If ϕm is taken to be 60° (Fig. 10, bottom row), the unified structure between the two hemispheres is disrupted in all cases including ODS1960, ODS2000, and ERA-Interim without 2002 due to the appreciable decreases of the variation of the poleward ozone flux in the SH (Figs. 1 and 2).

b. Description by eddy heat flux

From Eq. (23), we have
md(Xstt1Xstt0)=2πacosϕmt0t1(ρ0χ¯υθ¯θ¯z)ϕ=ϕm,z=zpdt.
The interannual variations of the variables in the two sides of this equation are shown as the scatterplots in Fig. 11. In the case of ϕm = 50° and zp = 100 hPa (top panels), the regression lines shift appreciably to the left along the abscissa (note the change of scale from Fig. 10) and, in the case of CCM simulations (Figs. 11a,b), there are interhemispheric differences in the y intercept of the regression lines, although the parallel structure is maintained. The steepening of the regression lines in Fig. 11 suggests that the variations are smaller in the eddy heat flux than in the poleward ozone flux (Fig. 10) implying that the eddy heat flux at Pref is not the major driver of the poleward ozone flux on an interannual time scale. Data from ERA-Interim (Fig. 11c) do not exhibit systematic dependence of polar ozone accumulation on the eddy heat flux in the NH.
Fig. 11.
Fig. 11.

As in Fig. 10, but with polar stratospheric ozone accumulation approximated by eddy heat flux [Eq. (27)].

Citation: Journal of the Atmospheric Sciences 80, 3; 10.1175/JAS-D-22-0023.1

If we select 50 hPa for zp (middle panels), the variance of eddy heat flux becomes larger than that of 100 hPa whereas that of polar ozone accumulation becomes smaller (Figs. 11d,e). This improves the linearity of the scatterplot. The situation appears the same for ERA-Interim (Fig. 11f) if the 2002 data are excluded. The improvement is still greater if we take ϕm = 60° and zp = 50 hPa (bottom panels). In the case of ODS1960 (Fig. 11g), increased variance of the eddy heat flux and reduced interhemispheric difference in the polar ozone accumulation improve the alignment of the regression lines for NH and SH. However, the negative ozone accumulation under ODS2000 disrupts the alignment by shifting the plots downward in the SH (Fig. 11h) even though high linearity is maintained in each hemisphere. The results from ERA-Interim (Fig. 11i) also shows a structure similar to that of ODS2000 if the year 2002 is excluded.

In summary, the linear relationship between the interannual variations of the fall-to-spring buildup of ozone integrated above and poleward of Pref and the eddy heat flux during the corresponding period at Pref is realized for each hemisphere; latitudinal averaging of the eddy heat flux between 45° and 75° (Weber et al. 2011) is not essential. By choosing Pref at higher latitude and altitude such as Pref(60°, 50 hPa), the linearity between the two variables and the interhemispheric unified structure are improved. Under ODS2000, however, the unified structure is disrupted owing to the enhanced ozone depletion in the SH. These findings are interpreted using the fact that the eddy heat flux at Pref is derived from the vertical integral above Pref of the vertical convergence of F˜ and that the interannual variation of the wintertime poleward ozone flux is well-captured by F˜ in the upper stratosphere.

6. Discussion

a. Application of the constituent-based EP Flux to the transport problem

We have seen in Fig. 9 that the winter-to-spring enhancement of the vertical convergence of F˜ is accompanied by the simultaneous growth of the equatorward transport due to the latitudinal convergence of F˜. It is interesting to investigate such a mutual dependence in the interannual variations of the processes, not just the convergences of F˜, contributing to the poleward ozone transport. Such investigations will help us understand the physical processes underlying the linear relationship of the Weber plot. Figure 12 shows the scatterplots demonstrating the relationship between the time and altitude integrals of the poleward ozone flux due to υ¯* (ordinate) and those of each contributing process decomposed into the rhs terms of Eq. (25) (abscissa); integration is made from fall-to-spring above Pref(ϕm, zp), where (ϕm, zp) is taken as (50°, 100 hPa), (50°, 50 hPa), and (60°, 50 hPa). The terms related to the latitudinal ozone gradient multiplied by F(ϕ) and the zonal wind acceleration (third and fifth terms on the rhs, respectively) are omitted because they are small. The features for ODS2000 are similar (not shown). The latitudinal convergence of F˜ (first term, Figs. 12a1, 12a2, and 12a3, F˜y) shows appreciably negative values (equatorward transport) with an enhancement at Pref(60°, 50 hPa), although it shows little correlation with the ozone flux due to υ¯*. The vertical convergence of F˜ (second term, Figs. 12b1, 12b2, and 12b3, F˜z) is positive and contributes to the linear relationship showing a good correspondence to the poleward ozone flux. The magnitude of this term becomes larger as the altitude and latitude of Pref increase. The vertical ozone gradient multiplied by F(z) (fourth term, Figs. 12c1, 12c2, and 12c3, F × χz) is also positive and contributes to the linear relationship. Compared with the vertical convergence of F˜, the correlation coefficient with the poleward ozone flux is not as high; the magnitude of this term is larger at Pref(50°, 100 hPa) but becomes smaller as the altitude and latitude of Pref increase, especially at Pref(60°, 50 hPa). Unresolved eddies (sixth, Figs. 12d1, 12d2, and 12d3, Y), estimated by subtracting the values of all rhs terms [and those of vertical advection of zonal momentum omitted in Eq. (25)] except the sixth term from those of the lhs, do not show positive correlation with the poleward flux due to υ¯*. The magnitude of this term is comparable to the vertical convergence of F˜ at Pref(50°, 100 hPa) but becomes smaller at Pref(50°, 50 hPa) and Pref(60°, 50 hPa). The description in the C-EP flux framework thus indicates that no single process can explain the linearity of the poleward ozone flux. Because the interannual variation of the springtime polar ozone accumulation results mostly from that of planetary wave breaking in the upper stratosphere rather than that of synoptic waves and unresolved eddies, the interannual variation of the transport features is captured by the convergence of F˜. The linear relationship demonstrated by Weber et al. (2011) is an approximation that relies solely on the vertical convergence of F˜.

Fig. 12.
Fig. 12.

Decomposition, as in Eq. (25), of the fall-to-spring summed poleward ozone flux due to υ¯* integrated above Pref(ϕm, zp) (abscissa of Fig. 10 taken as ordinate) into terms representing (a) latitudinal and (b) vertical convergences of F˜, (c) the vertical gradient of ozone multiplied by F(z), and (d) the contribution of unresolved eddies. (top to bottom) (ϕm, zp) is (50°, 100 hPa), (50°, 50 hPa), and (60°, 50 hPa), respectively. NH (black) and SH (red) for ODS1960 condition.

Citation: Journal of the Atmospheric Sciences 80, 3; 10.1175/JAS-D-22-0023.1

This paper is aimed at understanding the Weber plot on the basis that describing the polar ozone buildup in terms of single-level eddy heat flux is physically meaningful. The C-EP flux system that was introduced for this purpose has proved useful as a mathematical framework suitable for such a description. It is worth giving a brief summary of the system here. With the introduction of (υ˜*,w˜*)(χ¯υ¯*,χ¯w¯*) in addition to F˜ by Eq. (17), the momentum equation [Eq. (11)] reduces to the following under the quasigeostrophic approximation:
χ¯u¯tfυ˜*=1ρ0acosϕF˜+Y˜.
Here, the vertical advection due to w˜* is omitted because it is relatively small in the stratosphere. Equation (28) may be written for every atmospheric component. The sum of all such equations becomes mathematically identical to the quasigeostrophic TEM momentum equation (χ = 1). Equation (28) and F˜ are, therefore, a component of the TEM momentum equation and a part of the EP flux for a specific constituent, respectively. From a constituent viewpoint, this equation can be interpreted differently, i.e., the meridional transport of constituents is driven by the convergence of the C-EP flux by analogy with the convergence of the EP flux driving poleward residual circulation (section 1). More specifically, the poleward ozone transport due to υ¯* is not driven by F˜(z) alone but by the combination with others such as F˜(ϕ) and Y˜. There is a clear distinction, however, in that the term Y˜ is often nonnegligible depending on the intrinsic spatiotemporal variations of the constituent field, which makes the argument over the general applicability of the C-EP flux system difficult.
The continuity equation for constituent-based residual circulation is obtained from that of residual circulation [Eq. (4)] with the aid of Eq. (1):
1acosϕϕ(υ˜*cosϕ)+1ρ0z(ρ0w˜*)=S¯χ¯t+1ρ0M.
The rhs of this equation will vanish and the continuity equation becomes two-dimensionally nondivergent under the conditions that the tracer is chemically inactive, the time scale is such that the steady state assumption safely holds, and the divergence of eddy ozone flux is negligible. Then the constituent-based residual streamfunction Ψ˜* may be defined in the form
υ˜*=1ρ0cosϕΨ˜*z,w˜*=1ρ0acosϕΨ˜*ϕ.
Although the above conditions are not fully met in the case of stratospheric ozone, the description using Ψ˜* may be useful for surface-released conservative species such as CO2 and SF6.

One may argue against the use of the C-EP flux system because it is simply a rearrangement of the equations and does not deserve to be given a specific name. In fact, F˜ is merely the EP flux multiplied by χ¯ and the terms shown in Figs. 12b112b3 and 12c112c3 are equivalent to the second and third terms, respectively, on the rhs of Eq. (16) which is derived without using F˜. Still, the C-EP flux system is useful for understanding the Weber plot as more than an empirical relationship: it provides a mathematical basis as to why the eddy heat flux at a particular reference point Pref can approximate the poleward constituent (ozone) transport vertically integrated above Pref. The divergence of F˜ that implies a contribution to the equatorward transport in the lower stratosphere tells us how Pref should be chosen for better realization of the linearity. Despite the difficulty in generalizing the importance of the term Y˜, it is a framework with which to describe the poleward constituent transport in terms of the divergence of F˜ that sheds light on the role of eddy momentum flux as well as that of eddy heat flux. Ultimately, the usefulness of this system will depend on whether it provides an advantage over the system described by Eq. (16), recognizing it as a useful extension of the TEM momentum equation to the meridional constituent transport.

b. Enhanced ozone depletion in the Antarctic late spring

We found a remarkable enhancement of the eddy ozone flux (M) and the simultaneous activation of chemical ozone losses during October to December in the Antarctic stratosphere (Fig. 5). This takes place, accompanying the mixing of ozone-depleted air and ozone-rich air that had been separated by the PV mixing barrier along the vortex edge, associated with the transition to the summer circulation. The ozone losses are brought about by the mixing in of high-concentration NOx (mostly NO and NO2) from outside of the polar vortex (Akiyoshi et al. 2004). Some details on this process are shown in Fig. 13, taking the model year 2019 as an example. Note that there is no correspondence to the year 2019 of the real atmosphere. Figure 13a shows the time series similar to those of Fig. 5, but for one of the 500 ensemble members. Figure 13b is the time–height section of the chemical ozone production rate averaged in the area poleward of 50°S. We can see that the ozone depletion that appears at around the end of October on the 5 hPa level propagates downward until early December. Figures 13c–h show the horizontal distributions of the mixing ratios of ozone (left) and nitrogen oxides (center) and chemical production rate of ozone (right) superposed on the contours of potential vorticity on the 850 K isentropic surface (around 10 hPa) for 25 October (middle) and 5 November (bottom) 2019. Since daily values of NOx are not available in our model output, nitrogen oxides in this figure are defined by subtracting HNO3 from total reactive nitrogen. On 25 October when the polar vortex is still present, ozone is distributed in an almost axially symmetric structure and nitrogen oxides have extremely low values in the polar vortex, while the chemical ozone production rate has hardly any systematic features. When the vortex becomes unstable and the mixing associated with the vortex breaking is initiated (5 November), the area of low ozone concentration spreads in the elongated vortex and scatters around the small patches torn from the main vortex, while some organized structure emerges in the chemical ozone production rate. In general, the enhanced ozone depletion is seen in the region where the low-latitude air migrates closer to the pole along PV contours associated with the polar vortex deformation. What is interesting here is that the high depletion rates are distributed also inside the vortex. This will be due to the nitrogen oxides penetrating across the PV barrier marked by the contours around −200 PVU (labeled as −20, e.g., between 40° and 120°W). The longitude of penetration found in the Western Hemisphere on 5 November migrates day by day (not shown). This will be the source of chemical ozone depletion compensating for the ozone increase due to eddy ozone transport.

Fig. 13.
Fig. 13.

(a) As in Fig. 5, but extracted for the model year 2019, (b) time–height section of the chemical ozone production rate integrated poleward of 50°S (×1018 m−1 day−1). (c)–(e) Horizontal distributions of the mixing ratios of ozone (ppmv) and nitrogen oxides (ppbv), and ozone production rate (×106 cm−3 s−1), respectively, superposed on contours of potential vorticity (×10 PVU) on the 850 K isentrope for 25 Oct 2019. The meridian of 0° longitude is directed from the center to the left. (f)–(h) As in (c)–(e), respectively, but for 5 Nov 2019.

Citation: Journal of the Atmospheric Sciences 80, 3; 10.1175/JAS-D-22-0023.1

c. ODS-dependent variations

We have seen that the effect of ozone chemistry appears as the smaller values of y intercept and the larger slope in ODS2000 than in ODS1960 in the Weber plot (Figs. 4 and 11). The former is the direct consequence of enhanced gas-phase ozone chemistry under ODS2000 (section 5). The improved linearity associated with the change of zp from 100 to 50 hPa in Fig. 11, combined with the altitude dependence of the homogeneous and heterogeneous chemistry on stratospheric ozone (Fig. 3), supports this interpretation. This is also consistent with the notion that the long-term decrease of March total ozone relies mostly on halogen-driven ozone depletion occurring all year round rather than the winter-to-spring heterogeneous chemistry on PSCs (Chipperfield and Jones 1999). The latter comes from the growth of the variance of ozone accumulation even though the dynamical fields remain mostly unchanged. The investigation on this mechanism is left for future studies.

The ODS-dependent variations in polar-stratospheric ozone could be investigated further by using the linear relationship between the polar stratospheric ozone accumulation and the fall-to-spring poleward ozone flux caused by residual circulation [Eq. (26), Fig. 10]. The left panel of Fig. 14 is the same as Fig. 10c, except that all years from 1979 to 2018 available from ERA-Interim are shown. The dashed line superposed on the plots shows a slope of 0.28, which is the gradient of the regression line between the poleward ozone flux and associated ozone accumulation for ODS1960 (Fig. 10a). As the 1960 data are not available in ERA-Interim, the dashed line is configured to contain the 1979 data point of the NH, assuming that the effect of the 1960–79 increase in ODS on the slope is negligible. Almost all of the plotted data points fall below this line in both hemispheres, indicating that, for every year, the fall-to-spring seasonal accumulation of polar ozone was less than that of 1979 because of the enhanced ozone depletion under the ODS burden that increased from 1979 onward. The scatterplot of Fig. 14 (left panel) is converted to two time series, one for each hemisphere, in the top panel on the right by sorting the data points by time. The line plots connecting plus (+) symbols in the bottom panel are the same as those of the top panel, except that the deviations from the dashed line rather than the raw values in the left panel are presented. The time series of the plus symbols show the temporal evolution of the components attributable to the enhanced chemical ozone loss relative to 1979. However, the attribution is not exclusive because the fluctuations due to transport that appear to be aligned with the regression line are superposed. As the ODS levels changed over a decadal time scale and the dynamical fields remained almost the same between ODS1960 and ODS2000 despite the prevailing year-to-year variations, the ODS-dependent changes could be extracted with the aid of a low-pass filter. The heavy dashed lines are the 3-yr running means applied to the time series of the plus symbols, whereas the line plots connecting squares (□) are the residual obtained by subtracting the ODS-dependent changes (dashed line) from the raw time series (top-right panel). The residual time series approximately represent the variations in poleward ozone flux. Differences between the two hemispheres are attributable mainly to the ozone transport (□), and the small negative bias of the SH relative to the NH (dashed lines) is the result of enhanced ozone depletion in the SH. The slow decrease that persists until around 2000 and a possible recovery thereafter reflect the temporal development of ODS level. See, e.g., Figs. 1–18 of WMO (2018) as for its time evolution.

Fig. 14.
Fig. 14.

(left) As in Fig. 10c, but showing all available data from ERA-Interim. The dashed line passes through the NH data point for 1979 (i.e., the period from September 1979 to March 1980), with the slope of 0.28 taken from Fig. 10a. (right) Time series of Arctic (black) and Antarctic (red) stratospheric ozone accumulation from fall to spring obtained by sorting the data in the left panel by time. (top right) Raw values and (bottom right) lines connecting plus (+) symbols show deviations from the dashed line in the left panel. The heavy dashed lines are 3-yr running means. The lines connecting squares (□) show deviations of the raw values (top panel) from the running means.

Citation: Journal of the Atmospheric Sciences 80, 3; 10.1175/JAS-D-22-0023.1

7. Conclusions

The linear relationship between the wintertime midlatitude 100-hPa eddy heat flux and spring-to-fall ratio of polar total ozone (Weber et al. 2011) is investigated using MIROC3.2 CCM simulations (Akiyoshi et al. 2018). It is found that the fall-to-spring accumulation of polar stratospheric ozone, poleward and above Pref(ϕm, zp), is linearly correlated with the horizontal ozone flux owing to the residual circulation vertically integrated above Pref if appropriate values such as zp = 100 hPa and ϕm = 50° are set. This flux is decomposed into contributing processes including the eddy heat flux with the aid of the TEM momentum equation followed either by integration by parts or by introducing the constituent-based EP flux (C-EP flux; F˜). Defined as the conventional EP flux multiplied by the constituent (ozone) mixing ratio, F˜ is a quantity whose vertical convergence corresponds to the eddy heat flux at Pref in the transport equation vertically integrated above Pref. Because the interannual variation of the springtime polar ozone accumulation results mostly from that of planetary wave breaking in the upper stratosphere, the interannual variation of the transport features is captured by the convergence of F˜. While the horizontal convergence of F˜ describes the contribution of the eddy momentum flux to the meridional ozone transport on the seasonal time scale, the vertical convergence of F˜, thus the eddy heat flux at Pref, shows a clear linear relationship with the interannual variations of the fall-to-spring buildup of polar stratospheric ozone for each hemisphere even though it is not the dominant contributor to the poleward ozone flux due to υ¯* in the seasonally varying climatology. The results indicate that the linear relationship of Weber et al. (2011) can be justified on the basis of an approximation that considers the vertical convergence of F˜ alone and that the interhemispheric unification should be treated with caution.

Acknowledgments.

This work was supported by the Environment Research and Technology Development Fund (JPMEERF20172009) of the Ministry of the Environment, Japan, and by JSPS KAKENHI Grants JP18KK0289, JP19K03961, and JP20H01977. The CCM calculations were performed using a supercomputer system at CGER, NIES. We appreciate the helpful and constructive comments from T. Hirooka and the three anonymous reviewers that greatly improved this manuscript. We are also grateful to Y. Yamashita for carrying out the 500 ensemble simulations.

Data availability statement.

The model outputs are available from the corresponding author on request. Example code for reading (in Fortran) is also provided.

REFERENCES

  • Akiyoshi, H., T. Sugita, H. Kanzawa, and N. Kawamoto, 2004: Ozone perturbations in the Arctic summer lower stratosphere as a reflection of NOx chemistry and planetary scale wave activity. J. Geophys. Res., 109, D03304, https://doi.org/10.1029/2003JD003632.

    • Search Google Scholar
    • Export Citation
  • Akiyoshi, H., T. Nakamura, T. Miyasaka, M. Shiotani, and M. Suzuki, 2016: A nudged chemistry-climate model simulation of chemical constituent distribution at northern high-latitude stratosphere observed by SMILES and MLS during the 2009/2010 stratospheric sudden warming. J. Geophys. Res. Atmos., 121, 13611380, https://doi.org/10.1002/2015JD023334.

    • Search Google Scholar
    • Export Citation
  • Akiyoshi, H., M. Kadowaki, Y. Yamashita, and T. Nagatomo, 2018: Dependency of the areal extent of the northern midlatitude low-total ozone region on the greenhouse gas concentrations (in Japanese). Proc. 2018 Fall Meeting, Tokyo, Japan, Meteorological Society of Japan, P381.

  • Andrews, D. G., 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, 20312048, https://doi.org/10.1175/1520-0469(1976)033<2031:PWIHAV>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Andrews, D. G., J. R. Holton, and C. B. Leovy, 1987: Middle Atmosphere Dynamics. Academic Press, 489 pp.

  • Brasseur, G. P., J. J. Orlando, and G. S. Tyndall, Eds., 1999: Atmospheric Chemistry and Global Change. Oxford University Press, 688 pp.

  • Brewer, A. W., 1949: Evidence for a world circulation provided by the measurements of helium and water vapour distribution in the stratosphere. Quart. J. Roy. Meteor. Soc., 75, 351363, https://doi.org/10.1002/qj.49707532603.

    • Search Google Scholar
    • Export Citation
  • Chipperfield, M. P., and R. L. Jones, 1999: Relative influences of atmospheric chemistry and transport on Arctic ozone trends. Nature, 400, 551554, https://doi.org/10.1038/22999.

    • Search Google Scholar
    • Export Citation
  • Cunnold, D., F. Alyea, N. Phillips, and R. Prinn, 1975: A three-dimensional dynamical-chemical model of atmospheric ozone. J. Atmos. Sci., 32, 170194, https://doi.org/10.1175/1520-0469(1975)032<0170:ATDDCM>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Dee, D. P., and Coauthors, 2011: The ERA-Interim reanalysis: Configuration and performance of the data assimilation system. Quart. J. Roy. Meteor. Soc., 137, 553597, https://doi.org/10.1002/qj.828.

    • Search Google Scholar
    • Export Citation
  • Dobson, G. M. B., 1956: Origin and distribution of the polyatomic molecules in the atmosphere. Proc. Roy. Soc., 236A, 187193, https://doi.org/10.1098/rspa.1956.0127.

    • Search Google Scholar
    • Export Citation
  • Dunkerton, T., 1978: On the mean meridional mass motions of the stratosphere and mesosphere. J. Atmos. Sci., 35, 23252333, https://doi.org/10.1175/1520-0469(1978)035<2325:OTMMMM>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Dütsch, H. U., 1969: Atmospheric ozone and ultraviolet radiation. World Survey of Climatology, H. E. Landsberg, Ed., Vol. 4, Elsevier, 383–431.

  • Fusco, A. C., and M. L. Salby, 1999: Interannual variations of total ozone and their relationship to variations of planetary wave activity. J. Climate, 12, 16191629, https://doi.org/10.1175/1520-0442(1999)012<1619:IVOTOA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Haynes, P. H., C. J. Marks, M. E. McIntyre, T. G. Shepherd, and K. P. Shine, 1991: On the “downward control” of extratropical diabatic circulations by eddy-induced mean zonal forces. J. Atmos. Sci., 48, 651678, https://doi.org/10.1175/1520-0469(1991)048<0651:OTCOED>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Holton, J. R., P. H. Haynes, M. E. McIntyre, A. R. Douglass, R. B. Rood, and L. Pfister, 1995: Stratosphere-troposphere exchange. Rev. Geophys., 33, 403439, https://doi.org/10.1029/95RG02097.

    • Search Google Scholar
    • Export Citation
  • Kida, H., 1977a: A numerical investigation of the atmospheric general circulation and stratospheric-tropospheric mass exchange: I. Long-term integration of a simplified general circulation model. J. Meteor. Soc. Japan, 55, 5270, https://doi.org/10.2151/jmsj1965.55.1_52.

    • Search Google Scholar
    • Export Citation
  • Kida, H., 1977b: A numerical investigation of the atmospheric general circulation and stratospheric-tropospheric mass exchange: II. Lagrangian motion of the atmosphere. J. Meteor. Soc. Japan, 55, 7188, https://doi.org/10.2151/jmsj1965.55.1_71.

    • Search Google Scholar
    • Export Citation
  • McLandress, C., A. I. Jonsson, D. A. Plummer, M. C. Reader, J. F. Scinocca, and T. G. Shepherd, 2010: Separating the dynamical effects of climate change and ozone depletion. Part I: Southern Hemisphere stratosphere. J. Climate, 23, 50025020, https://doi.org/10.1175/2010JCLI3586.1.

    • Search Google Scholar
    • Export Citation
  • Newell, R. E., 1961: The transport of trace substances in the atmosphere and their implications for the general circulation of the stratosphere. Pure Appl. Geophys., 49, 137158, https://doi.org/10.1007/BF01992149.

    • Search Google Scholar
    • Export Citation
  • Newman, P. A., E. R. Nash, and J. E. Rosenfield, 2001: What controls the temperature of the Arctic stratosphere during the spring? J. Geophys. Res., 106, 19 99920 010, https://doi.org/10.1029/2000JD000061.

    • Search Google Scholar
    • Export Citation
  • Plumb, R. A., 2002: Stratospheric transport. J. Meteor. Soc. Japan, 80, 793809, https://doi.org/10.2151/jmsj.80.793.

  • Randel, W. J., 1993: Global variations of zonal mean ozone during stratospheric warming events. J. Atmos. Sci., 50, 33083321, https://doi.org/10.1175/1520-0469(1993)050<3308:GVOZMO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Randel, W. J., B. A. Boville, J. C. Gille, P. L. Bailey, S. T. Massie, J. B. Kumer, J. L. Mergenthaler, and A. E. Roche, 1994: Simulation of stratospheric N2O in the NCAR CCM2: Comparison with CLAES data and global budget analyses. J. Atmos. Sci., 51, 28342845, https://doi.org/10.1175/1520-0469(1994)051<2834:SOSNIT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Randel, W. J., F. Wu, and R. Stolarski, 2002: Changes in column ozone correlated with the stratospheric EP flux. J. Meteor. Soc. Japan, 80, 849862, https://doi.org/10.2151/jmsj.80.849.

    • Search Google Scholar
    • Export Citation
  • Weber, M., S. Dhomse, F. Wittrock, A. Richter, B.-M. Sinnhuber, and J. P. Burrows, 2003: Dynamical control of NH and SH winter/spring total ozone from GOME observations in 1995–2002. Geophys. Res. Lett., 30, 1583, https://doi.org/10.1029/2002GL016799.

    • Search Google Scholar
    • Export Citation
  • Weber, M., S. Dikty, J. P. Burrows, H. Garny, M. Dameris, A. Kubin, J. Abalichin, and U. Langematz, 2011: The Brewer-Dobson circulation and total ozone from seasonal to decadal time scales. Atmos. Chem. Phys., 11, 11 22111 235, https://doi.org/10.5194/acp-11-11221-2011.

    • Search Google Scholar
    • Export Citation
  • WMO, 2018: Scientific assessment of ozone depletion: 2018. Global Ozone Research and Monitoring Project Rep. 58, 588 pp.

Save
  • Akiyoshi, H., T. Sugita, H. Kanzawa, and N. Kawamoto, 2004: Ozone perturbations in the Arctic summer lower stratosphere as a reflection of NOx chemistry and planetary scale wave activity. J. Geophys. Res., 109, D03304, https://doi.org/10.1029/2003JD003632.

    • Search Google Scholar
    • Export Citation
  • Akiyoshi, H., T. Nakamura, T. Miyasaka, M. Shiotani, and M. Suzuki, 2016: A nudged chemistry-climate model simulation of chemical constituent distribution at northern high-latitude stratosphere observed by SMILES and MLS during the 2009/2010 stratospheric sudden warming. J. Geophys. Res. Atmos., 121, 13611380, https://doi.org/10.1002/2015JD023334.

    • Search Google Scholar
    • Export Citation
  • Akiyoshi, H., M. Kadowaki, Y. Yamashita, and T. Nagatomo, 2018: Dependency of the areal extent of the northern midlatitude low-total ozone region on the greenhouse gas concentrations (in Japanese). Proc. 2018 Fall Meeting, Tokyo, Japan, Meteorological Society of Japan, P381.

  • Andrews, D. G., 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, 20312048, https://doi.org/10.1175/1520-0469(1976)033<2031:PWIHAV>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Andrews, D. G., J. R. Holton, and C. B. Leovy, 1987: Middle Atmosphere Dynamics. Academic Press, 489 pp.

  • Brasseur, G. P., J. J. Orlando, and G. S. Tyndall, Eds., 1999: Atmospheric Chemistry and Global Change. Oxford University Press, 688 pp.

  • Brewer, A. W., 1949: Evidence for a world circulation provided by the measurements of helium and water vapour distribution in the stratosphere. Quart. J. Roy. Meteor. Soc., 75, 351363, https://doi.org/10.1002/qj.49707532603.

    • Search Google Scholar
    • Export Citation
  • Chipperfield, M. P., and R. L. Jones, 1999: Relative influences of atmospheric chemistry and transport on Arctic ozone trends. Nature, 400, 551554, https://doi.org/10.1038/22999.

    • Search Google Scholar
    • Export Citation
  • Cunnold, D., F. Alyea, N. Phillips, and R. Prinn, 1975: A three-dimensional dynamical-chemical model of atmospheric ozone. J. Atmos. Sci., 32, 170194, https://doi.org/10.1175/1520-0469(1975)032<0170:ATDDCM>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Dee, D. P., and Coauthors, 2011: The ERA-Interim reanalysis: Configuration and performance of the data assimilation system. Quart. J. Roy. Meteor. Soc., 137, 553597, https://doi.org/10.1002/qj.828.

    • Search Google Scholar
    • Export Citation
  • Dobson, G. M. B., 1956: Origin and distribution of the polyatomic molecules in the atmosphere. Proc. Roy. Soc., 236A, 187193, https://doi.org/10.1098/rspa.1956.0127.

    • Search Google Scholar
    • Export Citation
  • Dunkerton, T., 1978: On the mean meridional mass motions of the stratosphere and mesosphere. J. Atmos. Sci., 35, 23252333, https://doi.org/10.1175/1520-0469(1978)035<2325:OTMMMM>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Dütsch, H. U., 1969: Atmospheric ozone and ultraviolet radiation. World Survey of Climatology, H. E. Landsberg, Ed., Vol. 4, Elsevier, 383–431.

  • Fusco, A. C., and M. L. Salby, 1999: Interannual variations of total ozone and their relationship to variations of planetary wave activity. J. Climate, 12, 16191629, https://doi.org/10.1175/1520-0442(1999)012<1619:IVOTOA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Haynes, P. H., C. J. Marks, M. E. McIntyre, T. G. Shepherd, and K. P. Shine, 1991: On the “downward control” of extratropical diabatic circulations by eddy-induced mean zonal forces. J. Atmos. Sci., 48, 651678, https://doi.org/10.1175/1520-0469(1991)048<0651:OTCOED>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Holton, J. R., P. H. Haynes, M. E. McIntyre, A. R. Douglass, R. B. Rood, and L. Pfister, 1995: Stratosphere-troposphere exchange. Rev. Geophys., 33, 403439, https://doi.org/10.1029/95RG02097.

    • Search Google Scholar
    • Export Citation
  • Kida, H., 1977a: A numerical investigation of the atmospheric general circulation and stratospheric-tropospheric mass exchange: I. Long-term integration of a simplified general circulation model. J. Meteor. Soc. Japan, 55, 5270, https://doi.org/10.2151/jmsj1965.55.1_52.

    • Search Google Scholar
    • Export Citation
  • Kida, H., 1977b: A numerical investigation of the atmospheric general circulation and stratospheric-tropospheric mass exchange: II. Lagrangian motion of the atmosphere. J. Meteor. Soc. Japan, 55, 7188, https://doi.org/10.2151/jmsj1965.55.1_71.

    • Search Google Scholar
    • Export Citation
  • McLandress, C., A. I. Jonsson, D. A. Plummer, M. C. Reader, J. F. Scinocca, and T. G. Shepherd, 2010: Separating the dynamical effects of climate change and ozone depletion. Part I: Southern Hemisphere stratosphere. J. Climate, 23, 50025020, https://doi.org/10.1175/2010JCLI3586.1.

    • Search Google Scholar
    • Export Citation
  • Newell, R. E., 1961: The transport of trace substances in the atmosphere and their implications for the general circulation of the stratosphere. Pure Appl. Geophys., 49, 137158, https://doi.org/10.1007/BF01992149.

    • Search Google Scholar
    • Export Citation
  • Newman, P. A., E. R. Nash, and J. E. Rosenfield, 2001: What controls the temperature of the Arctic stratosphere during the spring? J. Geophys. Res., 106, 19 99920 010, https://doi.org/10.1029/2000JD000061.

    • Search Google Scholar
    • Export Citation
  • Plumb, R. A., 2002: Stratospheric transport. J. Meteor. Soc. Japan, 80, 793809, https://doi.org/10.2151/jmsj.80.793.

  • Randel, W. J., 1993: Global variations of zonal mean ozone during stratospheric warming events. J. Atmos. Sci., 50, 33083321, https://doi.org/10.1175/1520-0469(1993)050<3308:GVOZMO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Randel, W. J., B. A. Boville, J. C. Gille, P. L. Bailey, S. T. Massie, J. B. Kumer, J. L. Mergenthaler, and A. E. Roche, 1994: Simulation of stratospheric N2O in the NCAR CCM2: Comparison with CLAES data and global budget analyses. J. Atmos. Sci., 51, 28342845, https://doi.org/10.1175/1520-0469(1994)051<2834:SOSNIT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Randel, W. J., F. Wu, and R. Stolarski, 2002: Changes in column ozone correlated with the stratospheric EP flux. J. Meteor. Soc. Japan, 80, 849862, https://doi.org/10.2151/jmsj.80.849.

    • Search Google Scholar
    • Export Citation
  • Weber, M., S. Dhomse, F. Wittrock, A. Richter, B.-M. Sinnhuber, and J. P. Burrows, 2003: Dynamical control of NH and SH winter/spring total ozone from GOME observations in 1995–2002. Geophys. Res. Lett., 30, 1583, https://doi.org/10.1029/2002GL016799.

    • Search Google Scholar
    • Export Citation
  • Weber, M., S. Dikty, J. P. Burrows, H. Garny, M. Dameris, A. Kubin, J. Abalichin, and U. Langematz, 2011: The Brewer-Dobson circulation and total ozone from seasonal to decadal time scales. Atmos. Chem. Phys., 11, 11 22111 235, https://doi.org/10.5194/acp-11-11221-2011.

    • Search Google Scholar
    • Export Citation
  • WMO, 2018: Scientific assessment of ozone depletion: 2018. Global Ozone Research and Monitoring Project Rep. 58, 588 pp.

  • Fig. 1.

    Seasonal variation of the zonal-mean total ozone (DU) obtained by averaging 500 ensemble members simulated by CCM under the conditions of ozone-depleting substances (ODS) at (top) 1960 levels (ODS1960) and (middle) 2000 levels (ODS2000) compared with that from (bottom) the ERA-Interim 10-yr average over 1995–2004.

  • Fig. 2.

    Latitude–height sections of zonal-mean ozone number density (shade in green, ×1012 cm−3) superposed by the residual streamfunction with red (blue) contours for positive (negative) values at ±(0.1, 0.2, 0.3, 0.4, 0.5, 0.7, 1, 2, 4, 8, 15, 30, 50) (×102 kg m−1 s−1). The 500-member ensemble means for (left) ODS1960 and (right) difference between those for ODS1960 and ODS2000 simulations (ODS2000 minus ODS1960) for (top to bottom) January, March, July, and September. The difference field of the residual streamfunction is supplemented with dashed contours at ±(0.01, 0.02, 0.05) (×102 kg m−1 s−1).

  • Fig. 3.

    (top) Time–height sections of the difference between ODS1960 and ODS2000 conditions (ODS2000 minus ODS1960) in polar ozone mixing ratio (ppmv) area averaged in the region poleward of 50° latitude for the (left) Northern and (right) Southern Hemispheres. (bottom) As in the top row, but for the polar temperature (K).

  • Fig. 4.

    Scatterplots between the midlatitude winter-mean eddy heat flux at 100 hPa (K m s−1) and spring-to-fall ratio of polar total ozone derived from (top) ERA-Interim, and CCM under (middle) ODS1960, and (bottom) ODS2000. The analysis period of ERA-Interim is the same as that of Weber et al. (2011): from September 1995 (fall) to March 2010 (spring) for the Northern Hemisphere and from March 1996 (fall) to September 2010 (spring) for the Southern Hemisphere.

  • Fig. 5.

    Tendency of polar stratospheric ozone above zp = 100 hPa integrated poleward of latitude ϕm = 50° (×108 DU km2 day−1) as shown by time series of 500-member ensemble-mean values on the rhs of Eq. (10). Black, blue, green, and red lines are the first, second, third, and fourth terms, respectively. Vertical bars are the spread (standard deviation) drawn by thinning out by 5 days for visual clarity.

  • Fig. 6.

    (top) Time series of poleward ozone flux (×108 DU km2 day−1) expressed by Eq. (16). The 500-member ensemble-mean values under ODS1960 of the horizontal flux above Pref(ϕm, zp) (H.F; black), the latitudinal convergence of EP flux (Fy; blue), the eddy heat flux at Pref (Fz1; red), the additional term coming from integration by parts (Fz2; light blue), the zonal wind deceleration (Ut; orange), and the frictional term estimated as the residue (Res; green). (bottom) As in the top row, but for those expressed by Eq. (20): the horizontal flux (H.F; black), the latitudinal (F˜y; blue) and vertical (F˜z; red) convergences of F˜, the zonal wind deceleration (U˜t; orange), and the residual term (Res; green). ϕm and zp are taken to be 50° and 100 hPa, respectively. Vertical bars are the spread (standard deviation) drawn by thinning out by 5 days for visual clarity.

  • Fig. 7.

    Time–height sections of the poleward ozone flux (×10−9 kg m−2 s−1) at (left) 50°N and (right) 50°S described by Eq. (15). (a),(b) The ozone flux due to υ¯* is decomposed into the (c),(d) latitudinal and (e),(f) vertical convergences of EP flux, (g),(h) zonal wind acceleration, and (i),(j) the residual including those associated with small-scale eddies Y. Poleward (equatorward) flux is color-coded in red (blue). 500-member ensemble mean for ODS1960.

  • Fig. 8.

    As in Fig. 7, but those expressed by Eq. (25). Figures 7a and 7b are not replicated. (k),(l) The third term and (m),(n) the fourth term.

  • Fig. 9.

    As in Fig. 6 (bottom), but (ϕm, zp) is (top) (50°, 50 hPa) and (bottom) (60°, 50 hPa).

  • Fig. 10.

    Scatterplots of polar stratospheric ozone accumulation against the fall-to-spring integral of the poleward ozone flux across latitude ϕm above altitude zp [Eq. (26), (×109 DU km2)]. (top to bottom) (ϕm, zp) = (50°, 100 hPa), (50°, 50 hPa), and (60°, 50 hPa) under (left) ODS1960 and (center) ODS2000 CCM simulations, and (right) ERA-Interim from 1996 to 2010. Northern Hemisphere is in black and Southern Hemisphere is in red. Dashed red lines are derived by excluding the year 2002 that lies far to the right along the abscissa. Calculations are made referring to daily mean values on log-pressure coordinates taking the integration period [t0, t1] from 1 Sep to 31 Mar of the next year in the NH and from 1 Mar to 30 Sep for the SH.

  • Fig. 11.

    As in Fig. 10, but with polar stratospheric ozone accumulation approximated by eddy heat flux [Eq. (27)].

  • Fig. 12.

    Decomposition, as in Eq. (25), of the fall-to-spring summed poleward ozone flux due to υ¯* integrated above Pref(ϕm, zp) (abscissa of Fig. 10 taken as ordinate) into terms representing (a) latitudinal and (b) vertical convergences of F˜, (c) the vertical gradient of ozone multiplied by F(z), and (d) the contribution of unresolved eddies. (top to bottom) (ϕm, zp) is (50°, 100 hPa), (50°, 50 hPa), and (60°, 50 hPa), respectively. NH (black) and SH (red) for ODS1960 condition.

  • Fig. 13.

    (a) As in Fig. 5, but extracted for the model year 2019, (b) time–height section of the chemical ozone production rate integrated poleward of 50°S (×1018 m−1 day−1). (c)–(e) Horizontal distributions of the mixing ratios of ozone (ppmv) and nitrogen oxides (ppbv), and ozone production rate (×106 cm−3 s−1), respectively, superposed on contours of potential vorticity (×10 PVU) on the 850 K isentrope for 25 Oct 2019. The meridian of 0° longitude is directed from the center to the left. (f)–(h) As in (c)–(e), respectively, but for 5 Nov 2019.

  • Fig. 14.

    (left) As in Fig. 10c, but showing all available data from ERA-Interim. The dashed line passes through the NH data point for 1979 (i.e., the period from September 1979 to March 1980), with the slope of 0.28 taken from Fig. 10a. (right) Time series of Arctic (black) and Antarctic (red) stratospheric ozone accumulation from fall to spring obtained by sorting the data in the left panel by time. (top right) Raw values and (bottom right) lines connecting plus (+) symbols show deviations from the dashed line in the left panel. The heavy dashed lines are 3-yr running means. The lines connecting squares (□) show deviations of the raw values (top panel) from the running means.

All Time Past Year Past 30 Days
Abstract Views 113 0 0
Full Text Views 5210 4161 913
PDF Downloads 636 344 7