1. Introduction
The equator to pole circulation in the winter stratosphere is primarily driven by the upward propagating waves from the troposphere. This large-scale dynamical process is called the Brewer–Dobson (B-D) circulation and can be studied using the transformed Eulerian mean (TEM) equations (Edmon et al. 1980; Andrews et al. 1987; Holton et al. 1995; Shepherd 2007; Birner and Bonisch 2011). The Eliassen–Palm (E-P) flux divergence, which represents the wave forcing that acts on the mean flow to cause wind and temperature variations, is the most important quantity in the TEM equations. Numerous studies have used the TEM equations together with the E-P flux divergence to study the interannual variation and long-term trends of the B-D circulation (Edmon et al. 1980; Seviour et al. 2012), the behavior of planetary wave activity (Hu and Tung 2002; Karpetchko and Nikulin 2004; Hu et al. 2005), the variability of the polar vortex (Waugh et al. 1999; Newman et al. 2001), the momentum balance of the stratosphere (Dima and Wallace 2007; Monier and Weare 2011), and the annual cycle in tropical tropopause temperature (Kerr-Munslow and Norton 2006; Randel et al. 2008; Randel and Jensen 2013), among other topics. The fidelity of these studies relies crucially on the accuracy and homogeneity of the datasets that are used to derive the E-P flux divergence and the residual circulation that approximates the B-D circulation.
The E-P flux divergence is a highly derived quantity. Its calculation requires nonlocal information such as the spatial and temporal departures of the primary variables (i.e., winds and temperatures) from their mean fields. It is therefore extremely difficult to estimate the E-P flux divergence directly using station-based measurements. In addition, the calculation involves not only estimates of high-frequency fluctuation of the wave fluxes at different altitudes and latitudes but also differential operators that are applied to the slowly varying background temperature gradient. All these complications can potentially cause biases in the climatology, interannual variability, and/or long-term trends of E-P flux divergence. The nonlocal and nonlinear operators may also amplify small errors that are associated with the primary variables to much larger errors in the E-P flux divergence. It is therefore important to gauge the uncertainties in estimating the E-P flux divergence.
The most commonly used tools to derive the E-P flux divergence are reanalysis datasets, which are normally constructed by a variety of observations that are assimilated by using numerical weather prediction models to give a coherent representation of the global atmosphere with uniform spatial and temporal coverage (Uppala et al. 2005; Dee and Uppala 2009). A major concern with the use of reanalyses is their accuracy and homogeneity in representing both the underlying dynamics and long-term trends (e.g., Sterl 2004; Bengtsson et al. 2007; Thorne and Vose 2010). In particular, regions with relatively large analysis increments (defined as the reanalysis minus the model first guess that is based on the 6-hourly model forecast) can induce errors in estimating radiative balance and temperature (Uppala et al. 2005; Dee and Uppala 2008, 2009). In addition to model errors and drifts, studies have also shown that reanalysis datasets tend to differ from each other, especially in regard to long-term trends (e.g., Bengtsson et al. 2007; Kobayashi et al. 2009). This is because low-frequency and trend uncertainties may be induced by observational errors, including instrument biases and changes in geographical coverage. Sudden changes induced by incorporating newly available radiance measurements are of a particular concern in causing biases in low-frequency variation (Simmons et al. 2014).
ERA-40 and ERA-Interim, the two major consecutive reanalysis datasets produced by the European Centre for Medium-Range Weather Forecasts (ECMWF), have been widely used for the study of atmospheric circulation and processes (Dee and Uppala 2009; Uppala et al. 2005). ERA-Interim, the newest reanalysis product of ECMWF, is known to have many improvements over ERA-40 (Dee and Uppala 2008, 2009; Dee et al. 2011a). It has much smaller analysis increments during winter at high latitudes, more realistic temperature trends and radiative budget, and more reliable low-frequency variability (Dee and Uppala 2009; Dee et al. 2011a; Screen and Simmonds 2011; Bracegirdle and Marshall 2012; Cornes and Jones 2013; Simmons et al. 2014). It also has better representations of the hydrological cycle in the tropics and subtropics and a more realistic B-D circulation in the stratosphere (Schoeberl et al. 2003; van Noije et al. 2004; Monge-Sanz et al. 2007, 2013; Dee et al. 2011b). Studies have yet to be undertaken to evaluate how the improvement may have affected the wave forcing estimates. Because it is extremely difficult to compare the wave forcing estimates directly against the observations, a comparative study may provide some insights into the uncertainties of estimating wave forcing based on reanalysis datasets.
This study undertakes a comparative study between ERA-40 and ERA-Interim to quantify the discrepancies in wave forcing, measured by the E-P flux divergence and the associated wave fluxes. We choose to compare these two ECMWF reanalyses mainly because of the well-documented improvements of ERA-Interim over ERA-40; these help in diagnosing the possible causes of the discrepancies. Our focus is on the height region from the upper troposphere to the upper stratosphere (500–1 hPa), where the zonal mean wave forcing is the main driver of the large-scale circulation, and the Northern Hemisphere (NH) winter mean of December–February (DJF), when both the wave amplitude and variability are largest. We first detect the regions with the largest E-P flux divergence discrepancies and identify the key wave fluxes that contribute the most to them. We then examine to what extent the E-P flux divergence discrepancies are linked to discrepancies in the residual circulation. Finally, we apply a changepoint detection method called the penalized maximal t test (PMT) to investigate the temporal consistency of the poleward eddy heat flux
2. Data and methods
a. Data
The 40-yr ECMWF Re-Analysis (ERA-40) was generated by using the ECMWF Integrated Forecast System (IFS) model and its 6-hourly three-dimensional variational data assimilation (3D-Var) system (Uppala et al. 2005). It covered the period from September 1957 to August 2002 and incorporated observations from in situ measurements, including balloons, radiosondes, dropsondes, aircraft, and ships, along with satellite observations, which only provided global coverage of radiance measurements from 1979 onward. The data ingestion involved approximately 7–9 × 106 observations at each time step. The assimilation model used had a spectral T159 grid, corresponding to a 1.125° grid spacing in latitude and longitude and 60 levels in the vertical between the surface and 0.1 hPa (~65 km). Analysis products on the 23 standard pressure surfaces from 1000 to 1 hPa are available for general use.
Covering the data-rich satellite era of 1979–present the Interim ECMWF Re-Analysis (ERA-Interim) is the ECMWF’s current comprehensive atmospheric reanalysis (Dee and Uppala 2009; Dee et al. 2011a). It makes use of the same observations as ERA-40 before September 2002, supplemented with ECMWF operational data afterward (Berrisford et al. 2011; Simmons et al. 2014) but with major improvements over ERA-40. Especially, the ECMWF’s operational four-dimensional variational data assimilation (4D-Var) system couples the dynamic variables more cohesively with the humidity and radiation than its previous 3D-Var analysis system. This ensures a realistic interaction of temperature, vertical velocity, and humidity both temporally and spatially. Improved correction of biases in satellite radiance data is also achieved through the use of an automated variational bias correction system that optimizes the consistency of multiple measurements (Dee 2005; Dee and Uppala 2009; Dee et al. 2011a). In addition, the ERA-Interim assimilation model has a spectral T255 grid, corresponding to a ~0.70° grid spacing in latitude and longitude. It represents a higher spatial resolution than ERA-40; hence smaller-scale waves are resolved explicitly. The increase in spatial resolution is one of the key factors contributing to the reduction of analysis increments of temperatures as well as to a more realistic representation of the B-D circulation, in addition to many other improvements, including better physical parameterization schemes for radiative transfer, data quality control, subgrid-scale orographic drag, humidity analysis, clouds, and surface/soil processes (Dee and Uppala 2009; Dee et al. 2011a). The ERA-Interim assimilation model uses the same vertical levels as ERA-40 but the data are made available at 37 levels between 1000 and 1 hPa, including the standard 23 levels used by ERA-40.
Our analysis is based on the overlapping 22 winters (i.e., the winters of 1979/80–2001/02) that are shared by both ERA-40 and ERA-Interim. For clarity and simplicity, the definition of a winter is based on January across this paper; for example, the DJF mean of the 1979/80 winter is numbered and stated as 1980 hereafter.
b. TEM equation and the E-P flux divergence


























Equation (1) is assembled in this form so that the net effect of the wave forcing on the mean flow can be quantified. Its individual terms, however, may show contrasting or opposite behaviors (Edmon et al. 1980; Palmer 1981). Here, to identify the key flux terms that contribute most to the total wave forcing discrepancies, we not only analyze the total E-P flux divergence term
All the wave forcing quantities are calculated using data archived at 2.5° × 2.5° grid spacing and at the 23 pressure levels that are common to both reanalyses. As a result, the wave forcing estimated from this coarse resolution should primarily be dominated by the effect of large-scale Rossby waves. The derivatives involved in the E-P flux divergence and other quantities in Eq. (1) are all calculated using centered differences except for those at the top and bottom boundaries (i.e., 1000 and 1 hPa), where one-sided differences were employed. As such, the results at the boundaries are less reliable. In addition, all the calculations are performed on daily mean winds and temperatures and then averaged over the DJF season. We chose to use daily averages rather than the 6-hourly instantaneous records because the very high-frequency waves such as diurnal tides should make a negligible contribution toward the wave driving B-D circulation.
c. Diagnostic tools
ERA-40 and ERA-Interim describe the same circulation of Earth’s atmosphere. Ideally, there should be no difference between them in all the wave forcing quantities and in the residual circulation term
We apply the penalized maximal t test (Wang et al. 2007) to detect a significant sudden shift of mean in the wave forcing differences between the two reanalyses. A brief description of the method can be found in the appendix. To examine the principal contributor to the discontinuity, the PMT identification is separately applied to the total, stationary, and transient components of the wave forcing. This is because stationary waves are excited by the topography as well as land–sea heating contrast while transient waves are dominated by synoptic-scale weather patterns (Newman and Nash 2000). At each grid point, the total eddy heat flux
3. Results
a. Discrepancies in E-P flux divergence
Figures 1a and 1b show the climatology of DJF mean E-P fluxes (arrows) and E-P flux divergence term Ψ (contours) estimated from ERA-40 and ERA-Interim respectively. Both climatologies show that the wave forcing is marked by the upward and equatorward propagation of the E-P fluxes that are associated with the mainly negative E-P flux divergence term Ψ. There are two distinct peak regions of Ψ, one in the upper troposphere [~(200–300) hPa] and another in the upper stratosphere [~(1–3) hPa]. Another smaller peak can also be observed at high latitudes around 5–10 hPa.

Latitude–height cross section of the DJF mean E-P flux (arrows, 4 × 106 m3 s−2) and E-P flux divergence term Ψ (contours). Climatology from (a) ERA-40 and (b) ERA-Interim; and (c) composite difference (ERA40 − ERAInt). Solid (dashed) contours are positive (negative) divergence at intervals of ±0.6, ±1.8, ±3, ±4.2, … m s−1 day−1 in (a),(b) and ±0.3, ±0.9, ±1.5, ±2.1, … m s−1 day−1 in (c). The light (dark) shaded areas in (c) represent p ≤ 0.1 (0.05), estimated by two-sided Student’s t test. The plotted E-P flux vectors are shown as
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

Latitude–height cross section of the DJF mean E-P flux (arrows, 4 × 106 m3 s−2) and E-P flux divergence term Ψ (contours). Climatology from (a) ERA-40 and (b) ERA-Interim; and (c) composite difference (ERA40 − ERAInt). Solid (dashed) contours are positive (negative) divergence at intervals of ±0.6, ±1.8, ±3, ±4.2, … m s−1 day−1 in (a),(b) and ±0.3, ±0.9, ±1.5, ±2.1, … m s−1 day−1 in (c). The light (dark) shaded areas in (c) represent p ≤ 0.1 (0.05), estimated by two-sided Student’s t test. The plotted E-P flux vectors are shown as
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Latitude–height cross section of the DJF mean E-P flux (arrows, 4 × 106 m3 s−2) and E-P flux divergence term Ψ (contours). Climatology from (a) ERA-40 and (b) ERA-Interim; and (c) composite difference (ERA40 − ERAInt). Solid (dashed) contours are positive (negative) divergence at intervals of ±0.6, ±1.8, ±3, ±4.2, … m s−1 day−1 in (a),(b) and ±0.3, ±0.9, ±1.5, ±2.1, … m s−1 day−1 in (c). The light (dark) shaded areas in (c) represent p ≤ 0.1 (0.05), estimated by two-sided Student’s t test. The plotted E-P flux vectors are shown as
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Figure 1c shows the composite difference in the DJF mean E-P fluxes and E-P flux divergence term Ψ between the two reanalyses. The main feature of
Figure 2 shows the time series of DJF mean total E-P flux divergence term Ψ that are area-averaged over 45°–75°N at 3, 10, 50, and 100 hPa (top–bottom). At 3 hPa, noticeable discrepancies in both interannual variability and climatological mean can be observed with more negative Ψ values in ERA-40 than ERA-Interim. At 10 hPa, the difference is due mainly to the climatological mean with the ERA-Interim Ψ being more negative overall than that of ERA-40. At 50 hPa, a generally similar behavior to that at 3 hPa can be seen though the magnitude of the discrepancy is relatively smaller. At 100 hPa, the discrepancy is again dominated by a difference in the climatological mean with the ERA-40 Ψ being less negative than that of ERA-Interim. Over these four pressure levels, the climatological means of Ψ estimated from ERA-40 and ERA-Interim alternately exceed each other. The discrepancies are comparable to 15% of the interannual variability of Ψ at 10 hPa; this value increases to 45% at 100 hPa. There are also apparent trends in Ψ45–75N, especially at 100 hPa where upward trends are clearly noticeable in both ERA-40 and ERA-Interim estimates, and the trend of ERA-40 Ψ45–75N,100 hPa is distinctly steeper than that of ERA-Interim Ψ45–75N,100 hPa.

Time series of DJF mean E-P flux divergence term Ψ estimated from ERA-40 (blue dashed) and ERA-Interim (red dash-dotted) area-averaged over 45°–75°N and at (a) 3, (b) 10, (c) 50, and (d) 100 hPa.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

Time series of DJF mean E-P flux divergence term Ψ estimated from ERA-40 (blue dashed) and ERA-Interim (red dash-dotted) area-averaged over 45°–75°N and at (a) 3, (b) 10, (c) 50, and (d) 100 hPa.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Time series of DJF mean E-P flux divergence term Ψ estimated from ERA-40 (blue dashed) and ERA-Interim (red dash-dotted) area-averaged over 45°–75°N and at (a) 3, (b) 10, (c) 50, and (d) 100 hPa.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Figure 3 shows the composite differences of four of the individual terms that add up to the differences of the total wave forcing term Ψ. Because the climatology of

Composite differences (ERA40 − ERAInt) in the DJF E-P flux divergence terms (a) Ψ2, (b) Ψ3, (c) Ψ4, and (d) Ψ5. The contours and shadings are the same as in Fig. 1c with the exception that the statistical significance shading is omitted in regions of small differences and minor dynamical significance (<0.3 m s−1 day−1 in magnitude).
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

Composite differences (ERA40 − ERAInt) in the DJF E-P flux divergence terms (a) Ψ2, (b) Ψ3, (c) Ψ4, and (d) Ψ5. The contours and shadings are the same as in Fig. 1c with the exception that the statistical significance shading is omitted in regions of small differences and minor dynamical significance (<0.3 m s−1 day−1 in magnitude).
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Composite differences (ERA40 − ERAInt) in the DJF E-P flux divergence terms (a) Ψ2, (b) Ψ3, (c) Ψ4, and (d) Ψ5. The contours and shadings are the same as in Fig. 1c with the exception that the statistical significance shading is omitted in regions of small differences and minor dynamical significance (<0.3 m s−1 day−1 in magnitude).
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
In the middle-to-low latitude upper troposphere (0°–50°N, 200–500 hPa),
The poleward eddy potential heat flux

Latitude–height cross section of the DJF mean of the potential heat flux
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

Latitude–height cross section of the DJF mean of the potential heat flux
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Latitude–height cross section of the DJF mean of the potential heat flux
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Figure 5 elaborates this point further by showing the temporal variation of poleward eddy heat flux

Time series of DJF mean (top) total, (middle) stationary, and (bottom) transient components of the zonal mean eddy heat flux
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

Time series of DJF mean (top) total, (middle) stationary, and (bottom) transient components of the zonal mean eddy heat flux
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Time series of DJF mean (top) total, (middle) stationary, and (bottom) transient components of the zonal mean eddy heat flux
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
The eddy heat fluxes

Latitude–height cross section of composite differences of the DJF mean of the vertical gradient of potential temperature (a)
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

Latitude–height cross section of composite differences of the DJF mean of the vertical gradient of potential temperature (a)
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Latitude–height cross section of composite differences of the DJF mean of the vertical gradient of potential temperature (a)
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Figure 7 shows the vertical profile of DJF zonal mean temperature climatology

Vertical profile of (a) the DJF zonal mean temperature and (b) the differences between ERA-40 and ERA-Interim, showing area-weighted averages at 0°–35°N, 35°–75°N, and 75°–90°N.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

Vertical profile of (a) the DJF zonal mean temperature and (b) the differences between ERA-40 and ERA-Interim, showing area-weighted averages at 0°–35°N, 35°–75°N, and 75°–90°N.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Vertical profile of (a) the DJF zonal mean temperature and (b) the differences between ERA-40 and ERA-Interim, showing area-weighted averages at 0°–35°N, 35°–75°N, and 75°–90°N.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Figure 8 shows NH polar plots of DJF mean temperature differences

Polar stereographic plot of DJF mean temperature composite differences between ERA-40 and ERA-Interim at various pressure levels from 7 to 850 hPa. The hatched regions indicate statistical significance at p ≤ 0.05. Note that the value range of the color bars differs.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

Polar stereographic plot of DJF mean temperature composite differences between ERA-40 and ERA-Interim at various pressure levels from 7 to 850 hPa. The hatched regions indicate statistical significance at p ≤ 0.05. Note that the value range of the color bars differs.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Polar stereographic plot of DJF mean temperature composite differences between ERA-40 and ERA-Interim at various pressure levels from 7 to 850 hPa. The hatched regions indicate statistical significance at p ≤ 0.05. Note that the value range of the color bars differs.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
As well as showing significant discrepancies in the extratropical stratosphere, Fig. 3 also shows significant discrepancies in the upper troposphere. To illustrate the temporal variation of these tropospheric discrepancies, Fig. 9 shows the time series of DJF mean ERA-40 and ERA-Interim

Time series of DJF mean ERA-40 and ERA-Interim E-P flux divergence terms (left) Ψ3 and (right) Ψ5 that are area-weighted averages over 25°–50°N at 300 hPa.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

Time series of DJF mean ERA-40 and ERA-Interim E-P flux divergence terms (left) Ψ3 and (right) Ψ5 that are area-weighted averages over 25°–50°N at 300 hPa.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Time series of DJF mean ERA-40 and ERA-Interim E-P flux divergence terms (left) Ψ3 and (right) Ψ5 that are area-weighted averages over 25°–50°N at 300 hPa.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Figure 3 shows that

Time series of DJF ERA-40 and ERA-Interim E-P flux divergence terms Ψ5 that are area-weighted averages over 0°–10°N at (left) 300 and (right) 100 hPa.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

Time series of DJF ERA-40 and ERA-Interim E-P flux divergence terms Ψ5 that are area-weighted averages over 0°–10°N at (left) 300 and (right) 100 hPa.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Time series of DJF ERA-40 and ERA-Interim E-P flux divergence terms Ψ5 that are area-weighted averages over 0°–10°N at (left) 300 and (right) 100 hPa.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
b. Effect on the B-D circulation
This section investigates the extent to which the resolved wave forcing term Ψ is linked to the discrepancies in the B-D circulation by examining the momentum budget of the TEM equation. The first row of Fig. 11 shows the climatology of the DJF mean residual mean meridional circulation

Latitude–height cross section of the DJF mean residual mean meridional circulation (
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

Latitude–height cross section of the DJF mean residual mean meridional circulation (
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Latitude–height cross section of the DJF mean residual mean meridional circulation (
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
The key feature of the discrepancies in the residual circulation is the broadly positive
The second row of Fig. 11 shows the climatology of the nonconservative term
The magnitude of stratospheric
In the upper troposphere,
In the tropical upper troposphere,
c. Sudden change of mean in the eddy heat fluxes
Up to this point, the diagnostics have been based on the composite differences between the two datasets for their common period; they therefore do not address the discrepancies in long-term trends. Inhomogeneity in either temperature gradient
In this section, we use the PMT technique to detect any significant sudden departure of
Figure 12 shows the time series of DJF mean total, stationary, and transient eddy heat flux

Time series of DJF mean (top to bottom) total, stationary, and transient components of the zonal mean eddy heat fluxes that are area-weighted averages over 10°–30°N at 100 hPa. The right-hand y axis is for area-weighted averages of
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

Time series of DJF mean (top to bottom) total, stationary, and transient components of the zonal mean eddy heat fluxes that are area-weighted averages over 10°–30°N at 100 hPa. The right-hand y axis is for area-weighted averages of
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Time series of DJF mean (top to bottom) total, stationary, and transient components of the zonal mean eddy heat fluxes that are area-weighted averages over 10°–30°N at 100 hPa. The right-hand y axis is for area-weighted averages of
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
According to the three significance measures, a significant changepoint in the wave forcing difference
Figure 13a shows a NH polar plot of DJF mean eddy heat flux

Polar plots of DJF mean (a) ERA-Interim eddy heat flux at 100 hPa for the period of 1980–91 and (b) composite difference of the eddy heat flux difference
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

Polar plots of DJF mean (a) ERA-Interim eddy heat flux at 100 hPa for the period of 1980–91 and (b) composite difference of the eddy heat flux difference
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Polar plots of DJF mean (a) ERA-Interim eddy heat flux at 100 hPa for the period of 1980–91 and (b) composite difference of the eddy heat flux difference
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Figure 13b shows the difference plot of total
Figure 14 shows the time series of DJF mean total, stationary, and transient

As in Fig. 12, but for the eddy head flux over 45°–75°N at 10 hPa.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

As in Fig. 12, but for the eddy head flux over 45°–75°N at 10 hPa.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
As in Fig. 12, but for the eddy head flux over 45°–75°N at 10 hPa.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
Figure 15 shows the spatial characteristics of the 1998 sudden change of

As in Fig. 13, but at 10 hPa and with a changepoint in 1998.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1

As in Fig. 13, but at 10 hPa and with a changepoint in 1998.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
As in Fig. 13, but at 10 hPa and with a changepoint in 1998.
Citation: Journal of Climate 28, 6; 10.1175/JCLI-D-14-00356.1
4. Conclusions and discussion
We have here reported that significant discrepancies exist in the wave forcing estimated from ERA-40 and ERA-Interim during NH winter. When measured by the E-P flux divergence, three key regions are identified as having significant discrepancies. They are the entire high latitudes, the upper troposphere, and the extratropical upper stratosphere. The discrepancies in the high latitudes are marked by vertically alternating positive and negative anomalies of the E-P flux divergence. They are manifested as differences in the climatological mean between the two datasets and can account for up to 15%–45% of the interannual variability in the affected regions. Such discrepancies are due mainly to differences in the vertical gradient of potential temperature
Similar vertically alternating positive and negative anomalies were previously found in the analysis increments of temperature in many reanalysis datasets and are known to be caused by the presence of systematic bias between the data assimilation model and the satellite measurements (Uppala et al. 2005; Dee and Uppala 2009; Kobayashi et al. 2009). Such a bias has a larger magnitude and is more persistent in ERA-40 than ERA-Interim (Simmons et al. 2007; Dee and Uppala 2009). Recent studies indicate much closer agreement to observations by ERA-Interim compared to ERA-40, which is attributed to the advances in the ERA-Interim assimilation system, especially the various improvements of the ERA-Interim 4D-Var system over the previous 3D-Var system that was used by ERA-40 (e.g., Fueglistaler et al. 2009; Dee et al. 2011a; Simmons et al. 2010, 2014; Bracegirdle and Marshall 2012). For this reason, we suggest that the E-P flux divergence discrepancies at high latitudes are most likely due to the model drift induced by the data assimilation system, rather than observational errors.
In the middle-to-low latitude upper troposphere, the discrepancies in the E-P flux divergence are due largely to the bias in the vertical momentum flux
In the upper stratosphere, the E-P flux divergence discrepancies involve all the relevant flux terms and are associated with substantial differences in temperature as well as static stability. These discrepancies may be attributed to the relatively larger model bias in the region, where observations are sparse and model errors are large (Dee and Uppala 2009). Nevertheless, we find that the discrepancies between these two datasets become much reduced both in terms of interannual variability, climatological mean, and long-term trend if the wave forcing is measured by the poleward eddy heat flux
Based on the TEM momentum budget, we have shown that a stronger residual circulation is associated with ERA-40 than ERA-Interim, agreeing with previous studies (e.g., van Noije et al. 2004; Monge-Sanz et al. 2007; Dee and Uppala 2009; Monge-Sanz et al. 2013). However, the discrepancies in the residual circulation are only partially associated with the discrepancies in the resolved large-scale wave forcing. The majority of the discrepancies in the residual circulation are associated with the nonconservative term
The thermodynamic balance in the stratosphere is largely a balance between the radiative heating and the dynamical heating from the advection of the residual circulation (Andrews et al. 1987). Because the dynamical heating term in the thermodynamic budget of the TEM equations is the product of the residual velocity and the temperature gradients, an enhanced residual circulation should be associated with cooling in the low-latitude stratosphere and warming at high latitudes if the radiative heating is constant. However, the temperature difference
A sudden drop of the eddy heat flux difference
A subtler sudden drop in eddy heat flux difference
Several studies have found significant trends in stratospheric wave forcing (Newman and Nash 2000; Randel et al. 2002; Hu and Tung 2002) while others have found that the trends reverse in early and later winter with no significant trend in midwinter (Karpetchko and Nikulin 2004; Hu et al. 2005). Here, we have found that trends in the E-P flux divergences differ substantially between these two datasets. Sudden changes in either temperature gradient or eddy fluxes that are induced by inhomogeneity of observations are able to alter the respective trends and low-frequency variability in the wave forcing. Because of the highly derived nature of the E-P flux divergence, the trends estimated from such a quantity should be treated with extreme caution.
Nevertheless, we have found that the trends in the eddy heat flux
This comparative study of wave forcing estimated from ERA-40 and ERA-Interim provides an additional perspective for evaluating dynamic processes in the stratosphere and upper troposphere. It is noted that a comparative study like this cannot make a quantitative attribution in terms of which dataset is better and/or by how much. Our results nevertheless show that bias in static stability induced by temperature differences and/or radiative heating imbalance can potentially cause large uncertainty in the E-P flux divergence, endorsing the importance of reducing the analysis increments, especially the model drift, in assimilating temperatures. Our results also demonstrate the importance of the recently established Stratospheric Processes and their Role in Climate (SPARC) Reanalysis/Analysis Intercomparison Project (S-RIP) (Fujiwara et al. 2012; Fujiwara and Jackson 2013).
Acknowledgments
This study is part of the British Antarctic Survey Polar Science for Planet Earth Programme funded by the Natural Environment Research Council. We acknowledge use of ECMWF reanalysis datasets and documentation at http://www.ecmwf.int. We would also like to thank the three reviewers for their detailed and constructive comments.
APPENDIX
Penalized Maximal t Test






























Here, we use three measures to evaluate the significance of any detected changepoint. The first measure is the chance of a changepoint occurring at the detected position, the second measure is the significance of the mean-shift magnitude Δμ, and the third measure is the significance of the maximum value of the penalized t statistics
REFERENCES
Andrews, D. G., J. R. Holton, and C. B. Leovy, 1987: Middle Atmosphere Dynamics. Academic Press, 489 pp.
Bengtsson, L., and Coauthors, 2007: The need for a dynamical climate reanalysis. Bull. Amer. Meteor. Soc., 88, 495–501, doi:10.1175/BAMS-88-4-495.
Berrisford, P., and Coauthors, 2011: Atmospheric conservation properties in ERA-Interim. Quart. J. Roy. Meteor. Soc., 137, 1381–1399, doi:10.1002/qj.864.
Birner, T., and H. Bonisch, 2011: Residual circulation trajectories and transit times into the extratropical lowermost stratosphere. Atmos. Chem. Phys., 11, 817–827, doi:10.5194/acp-11-817-2011.
Birner, T., D. W. J. Thompson, and T. G. Shepherd, 2013: Up-gradient eddy fluxes of potential vorticity near the subtropical jet. Geophys. Res. Lett., 40, 5988–5993, doi:10.1002/2013GL057728.
Bracegirdle, T. J., and G. J. Marshall, 2012: The reliability of Antarctic tropospheric pressure and temperature in the latest global reanalyses. J. Climate, 25, 7138–7146, doi:10.1175/JCLI-D-11-00685.1.
Butchart, N., and Coauthors, 2010: Chemistry–climate model simulations of twenty-first-century stratospheric climate and circulation changes. J. Climate, 23, 5349–5374, doi:10.1175/2010JCLI3404.1.
Cornes, R., and P. Jones, 2013: How well does the ERA-Interim reanalysis replicate trends in extremes of surface temperature across Europe? J. Geophys. Res. Atmos., 118, 10 262–10 276, doi:10.1002/jgrd.50799.
Dee, D. P., 2005: Bias and data assimilation. Quart. J. Roy. Meteor. Soc., 131, 3323–3343, doi:10.1256/qj.05.137.
Dee, D. P., and S. Uppala, 2008: Variational bias correction in ERA-Interim. ECMWF Tech. Memo. 575, 26 pp.
Dee, D. P., and S. Uppala, 2009: Variational bias correction of satellite radiance data in the ERA-Interim reanalysis. Quart. J. Roy. Meteor. Soc., 135, 1830–1841, doi:10.1002/qj.493.
Dee, D. P., and Coauthors, 2011a: 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.
Dee, D. P., E. Källén, A. J. Simmons, and I. Haimberger, 2011b: Comments on “Reanalyses suitable for characterizing long-term trends.” Bull. Amer. Meteor. Soc., 92, 65–70, doi:10.1175/2010BAMS3070.1.
Dima, I., and J. Wallace, 2007: Structure of the annual-mean equatorial planetary waves in the ERA-40 reanalyses. J. Atmos. Sci., 64, 2862–2880, doi:10.1175/JAS3985.1.
Edmon, H., B. Hoskins, and M. McIntyre, 1980: Eliassen-Palm cross sections for the troposphere. J. Atmos. Sci., 37, 2600–2616, doi:10.1175/1520-0469(1980)037<2600:EPCSFT>2.0.CO;2.
Fueglistaler, S., B. Legras, A. Beljaars, J.-J. Morcrette, A. Simmons, A. M. Tompkins, and S. Uppala, 2009: The diabatic heat budget of the upper troposphere and lower/mid stratosphere in ECMWF reanalyses. Quart. J. Roy. Meteor. Soc., 135, 21–37, doi:10.1002/qj.361.
Fujiwara, M., and D. Jackson, 2013: SPARC Reanalysis Intercomparison Project (S-RIP) planning meeting. SPARC Newsletter, No. 41, SPARC Office, Toronto, Ontario, Canada, 52–55.
Fujiwara, M., S. Polavarapu, and D. Jackson, 2012: A proposal of the SPARC Reanalysis/Analysis Intercomparison Project. SPARC Newsletter, No. 38, SPARC Office, Toronto, Ontario, Canada, 14–17.
Holton, J. R., 1983: The influence of gravity wave breaking on the general circulation of the middle atmosphere. J. Atmos. Sci., 40, 2497–2507, doi:10.1175/1520-0469(1983)040<2497:TIOGWB>2.0.CO;2.
Holton, J. R., P. Haynes, M. McIntyre, A. Douglass, R. Rood, and L. Pfister, 1995: Stratosphere–troposphere exchange. Rev. Geophys., 33, 403–439, doi:10.1029/95RG02097.
Hu, Y., and K. Tung, 2002: Interannual and decadal variations of planetary wave activity, stratospheric cooling, and Northern Hemisphere annular mode. J. Climate, 15, 1659–1673, doi:10.1175/1520-0442(2002)015<1659:IADVOP>2.0.CO;2.
Hu, Y., K. Tung, and J. Liu, 2005: A closer comparison of early and late-winter atmospheric trends in the Northern Hemisphere. J. Climate, 18, 3204–3216, doi:10.1175/JCLI3468.1.
Iwasaki, T., H. Hamada, and K. Miyazaki, 2009: Comparisons of Brewer-Dobson circulations diagnosed from reanalyses. J. Meteor. Soc. Japan, 87, 997–1006, doi:10.2151/jmsj.87.997.
Karpetchko, A., and G. Nikulin, 2004: Influence of early winter upward wave activity flux on midwinter circulation in the stratosphere and troposphere. J. Climate, 17, 4443–4452, doi:10.1175/JCLI-3229.1.
Kerr-Munslow, A., and W. Norton, 2006: Tropical wave driving of the annual cycle in tropical tropopause temperatures. Part I: ECMWF analyses. J. Atmos. Sci., 63, 1410–1419, doi:10.1175/JAS3697.1.
Kobayashi, S., M. Matricardi, D. Dee, and S. Uppala, 2009: Toward a consistent reanalysis of the upper stratosphere based on radiance measurements from SSU and AMSU-A. Quart. J. Roy. Meteor. Soc., 135, 2086–2099, doi:10.1002/qj.514.
Lindzen, R. S., 1981: Turbulence and stress owing to gravity wave and tidal breakdown. J. Geophys. Res., 86, 9707–9714, doi:10.1029/JC086iC10p09707.
Lu, H., T. J. Bracegirdle, T. Phillips, A. Bushell, and L. Gray, 2014: Mechanisms for the Holton–Tan relationship and its decadal variation. J. Geophys. Res. Atmos., 119, 2811–2830, doi:10.1002/2013JD021352.
McFarlane, N. A., 1987: The effect of orographically excited gravity wave drag on the general circulation of the lower stratosphere and troposphere. J. Atmos. Sci., 44, 1775–1800, doi:10.1175/1520-0469(1987)044<1775:TEOOEG>2.0.CO;2.
McLandress, C., and T. Shepherd, 2009: Simulated anthropogenic changes in the Brewer–Dobson circulation, including its extension to high latitudes. J. Climate, 22, 1516–1540, doi:10.1175/2008JCLI2679.1.
Monge-Sanz, B., M. Chipperfield, A. Simmons, and S. Uppala, 2007: Mean age of air and transport in a CTM: Comparison of different ECMWF analyses. Geophys. Res. Lett., 34, L04801, doi:10.1029/2006GL028515.
Monge-Sanz, B., M. Chipperfield, D. Dee, A. Simmons, and S. Uppala, 2013: Improvements in the stratospheric transport achieved by a chemistry transport model with ECMWF (re)analyses: Identifying effects and remaining challenges. Quart. J. Roy. Meteor. Soc., 139, 654–673, doi:10.1002/qj.1996.
Monier, E., and B. Weare, 2011: Climatology and trends in the forcing of the stratospheric ozone transport. Atmos. Chem. Phys., 11, 6311–6323, doi:10.5194/acp-11-6311-2011.
Newman, P., and E. Nash, 2000: Quantifying the wave driving of the stratosphere. J. Geophys. Res., 105, 12 485–12 497, doi:10.1029/1999JD901191.
Newman, P., E. Nash, and J. Rosenfield, 2001: What controls the temperature of the Arctic stratosphere during the spring? J. Geophys. Res., 106, 19 999–20 010, doi:10.1029/2000JD000061.
Palmer, T. N., 1981: Diagnostic study of a wavenumber-2 stratospheric sudden warming in a transformed Eulerian-mean formalism. J. Atmos. Sci., 38, 844–855, doi:10.1175/1520-0469(1981)038<0844:DSOAWS>2.0.CO;2.
Plumb, R., 1985: An alternative form of Andrews’ conservation law for quasi-geostrophic waves on a stead, nonuniform flow. J. Atmos. Sci., 42, 298–300, doi:10.1175/1520-0469(1985)042<0298:AAFOAC>2.0.CO;2.
Randel, W., and E. Jensen, 2013: Physical processes in the tropical tropopause layer and their roles in a changing climate. Nat. Geosci., 6, 169–176, doi:10.1038/ngeo1733.
Randel, W., R. Garcia, and F. Wu, 2002: Time-dependent upwelling in the tropical lower stratosphere estimated from the zonal-mean momentum budget. J. Atmos. Sci., 59, 2141–2152, doi:10.1175/1520-0469(2002)059<2141:TDUITT>2.0.CO;2.
Randel, W., R. Garcia, and F. Wu, 2008: Dynamical balances and tropical stratospheric upwelling. J. Atmos. Sci., 65, 3584–3595, doi:10.1175/2008JAS2756.1.
Schoeberl, M., A. Douglass, Z. Zhu, and S. Pawson, 2003: A comparison of the lower stratospheric age spectra derived from a general circulation model and two data assimilation systems. J. Geophys. Res., 108, 4113, doi:10.1029/2002JD002652.
Screen, J., and I. Simmonds, 2011: Erroneous Arctic temperature trends in the ERA-40 reanalysis: A closer look. J. Climate, 24, 2620–2627, doi:10.1175/2010JCLI4054.1.
Seviour, W., N. Butchart, and S. Hardiman, 2012: The Brewer–Dobson circulation inferred from ERA-Interim. Quart. J. Roy. Meteor. Soc., 138, 878–888, doi:10.1002/qj.966.
Shepherd, T., 2007: Transport in the middle atmosphere. J. Meteor. Soc. Japan, 85B, 165–191, doi:10.2151/jmsj.85B.165.
Simmons, A. J., S. Uppala, D. Dee, and S. Kobayashi, 2007: ERA-Interim: New ECMWF reanalysis products from 1989 onwards. ECMWF Newsletter, No. 110, ECMWF, Reading, United Kingdom, 25–35.
Simmons, A. J., K. M. Willett, P. D. Jones, P. W. Thorne, and D. P. Dee, 2010: Low-frequency variations in surface atmospheric humidity, temperature, and precipitation: Inferences from reanalyses and monthly gridded observational data sets. J. Geophys. Res., 115, D01110, doi:10.1029/2009JD012442.
Simmons, A. J., P. Poli, D. P. Dee, P. Berrisford, H. Hersbach, S. Kobayashi, and C. Peubey, 2014: Estimating low-frequency variability and trends in atmospheric temperature using ERA-Interim. Quart. J. Roy. Meteor. Soc., 140, 329–353, doi:10.1002/qj.2317.
Sterl, A., 2004: On the (in)homogeneity of reanalysis products. J. Climate, 17, 3866–3873, doi:10.1175/1520-0442(2004)017<3866:OTIORP>2.0.CO;2.
Thorne, P., and R. Vose, 2010: Reanalyses suitable for characterizing long-term trends: Are they really achievable? Bull. Amer. Meteor. Soc., 91, 353–361, doi:10.1175/2009BAMS2858.1.
Uppala, S., and Coauthors, 2005: The ERA-40 Re-Analysis. Quart. J. Roy. Meteor. Soc., 131, 2961–3012, doi:10.1256/qj.04.176.
van Noije, T., H. Eskes, M. van Weele, and P. van Velthoven, 2004: Implications of the enhanced Brewer-Dobson circulation in European Centre for Medium-Range Weather Forecasts reanalysis ERA-40 for the stratosphere-troposphere exchange of ozone in global chemistry transport models. J. Geophys. Res., 109, D19308, doi:10.1029/2004JD004586.
Wang, X., Q. Wen, and Y. Wu, 2007: Penalized maximal t test for detecting undocumented mean change in climate data series. J. Appl. Meteor. Climatol., 46, 916–931, doi:10.1175/JAM2504.1.
Waugh, D., W. Randel, S. Pawson, P. Newman, and E. Nash, 1999: Persistence of the lower stratospheric polar vortices. J. Geophys. Res., 104, 27 191–27 201, doi:10.1029/1999JD900795.
Wright, J. S., and S. Fueglistaler, 2013: Large differences in reanalyses of diabatic heating in the tropical upper troposphere and lower stratosphere. Atmos. Chem. Phys., 13, 9565–9576, doi:10.5194/acp-13-9565-2013.