1. Introduction
Tropical cyclones (TCs) have the potential to lead to a loss of life and cause massive damage when making landfall. The damage is likely to increase with a larger and wealthier population concentrated in coastal areas (Pielke et al. 2008). TCs are projected to become more destructive (Grinsted et al. 2019) and intense TCs may become more frequent [e.g., Bender et al. (2010)] in a warmer climate.
Vertical wind shear, i.e., the vertical shear of horizontal wind, has long been linked to TC intensity. Increased vertical wind shear has been found to impede intensification by tilting the TC vortex and causing an outward flux of potential vorticity (PV) and equivalent potential temperature, thereby weakening the cyclone (Frank and Ritchie 2001). Further, vertical wind shear aids in ventilating the TC core by entraining relatively dry environmental air into convective updrafts (Emanuel et al. 2004; Tang and Emanuel 2010, 2012) and into the boundary layer (Riemer et al. 2010; Riemer and Laliberté 2015). Entrainment thereby reduces latent heating and buoyancy, and thus hinders convection. Furthermore, entrainment reduces the equivalent potential temperature in the convective updrafts, which leads to outflow at lower altitudes and thereby warmer temperatures. This reduces the intensity of the TC following the heat engine model of Emanuel (1986). Relative humidity at various pressure levels throughout the lower and midtroposphere has been linked to TC intensity and intensification rates, and reduced relative humidity has been found to reduce TC activity (e.g., Kaplan and DeMaria 2003; Emanuel et al. 2004; Hendricks et al. 2010; Wu et al. 2012). The ventilation index of Tang and Emanuel (2012) includes the combined effect of humidity and vertical wind shear. It has been found useful in the reconstruction of the annual cycle of North Atlantic TC activity, in particular the typically sharp increase in activity from July to August (Yang et al. 2021). Further, the ventilation index shows significant correlation with TC frequency and rapid intensification [i.e., an increase in maximum wind speed of at least 15.4 m s−1 in 24 h, see Kaplan and DeMaria (2003)] frequency within numerical modeling frameworks (Tang and Camargo 2014).
The large potential for damage that TCs can cause gives rise to the need for seasonal forecasting of TCs, which is an endeavor that has been undertaken for several decades. In the North Atlantic, the earliest attempts by Gray (1984a,b) used El Niño events, quasi-biennial oscillation phases and sea level pressure anomalies in the Caribbean during spring and early summer to forecast hurricane activity with useful skill. Since then, this approach of statistical forecasting has been continuously developed (e.g., Klotzbach and Gray 2004; Saunders and Lea 2005; Klotzbach 2007). Statistical forecasting generally makes use of the relationship between observable phenomena and environmental factors relevant to TC activity. For example, positive El Niño phases correlate with an increase in vertical wind shear and a decrease in TC activity in the tropical North Atlantic (Aiyyer and Thorncroft 2006).
Dynamical seasonal forecasts, where a numerical model is used to predict the future state of the atmosphere, can also be used to predict TC activity. In a purely dynamical approach, TCs within the forecast are tracked and evaluated directly. Even at low resolutions, climate models can produce TC-like structures, though the simulated TCs are larger and weaker than observed TCs (Manabe et al. 1970; Bengtsson et al. 1982). Still, while individual TCs are unrealistic, they provide useful information for forecasting the geographical and seasonal distribution of TCs (Bengtsson et al. 1995). Dynamical forecasting has since improved greatly, and provides skillful results (e.g., Thorncroft and Pytharoulis 2001; Vitart et al. 2007; Vecchi et al. 2014; Zhang et al. 2019). Dynamical forecasts can also be used in combination with statistical modeling in a hybrid approach, where the large-scale environment produced by the dynamical forecast is used to make statistical forecasts. Operational TC forecasts have generally been found to have good skill in predicting the number of TCs, and some provide useful information on landfall locations and regional activity (Klotzbach et al. 2019).
Forecasts for the 2013 North Atlantic hurricane season were consistently indicating an above average activity in terms of number of named storms, hurricanes and major hurricanes. In reality, there were 14 named storms, which is slightly above average, but only two category-1 hurricanes and no major hurricanes, which is far below average. Zhang et al. (2016) proposed that an increased number of anticyclonic Rossby wave breaking events reduced the humidity in the tropics through mixing with extratropical air and increased vertical wind shear. This prevented cyclones from intensifying despite other environmental factors being conducive to intense cyclones. Zhang et al. (2016) investigated the development of Rossby waves using PV, and drew special attention to the peculiar extreme equatorward position of the 2 potential vorticity unit (PVU; 1 PVU = 10−6 K kg−1 m2 s−1) contour on the 350-K isentropic surface during August. The 2-PVU contour can also be used to identify stratospheric PV streamers, which are elongated filaments that protrude from the high-PV stratosphere into the low-PV troposphere. These, in turn, can be used as a proxy for anticyclonic Rossby wave breaking (Wernli and Sprenger 2007; Béguin et al. 2013; Sprenger et al. 2017).
PV streamers and associated Rossby wave breaking events can be reliably identified (e.g., Bowley et al. 2019; Papin et al. 2020), and Rossby wave breaking has been studied in reanalysis data (e.g., Postel and Hitchman 1999; Scott and Cammas 2002; Abatzoglou and Magnusdottir 2006; Wernli and Sprenger 2007). However, predicting Rossby wave breaking based on numerical models may show biases in frequency and location (Barnes and Hartmann 2012). The location biases are not simply a direct result of biases in the general tropopause location, but of where along the tropopause models produce these events. The bias in Rossby wave breaking frequency is largest in the North Atlantic region, and the bias of the 2-PVU contour (i.e., the tropopause) latitude has been found to be lower than that of the Rossby wave breaking latitude during summer in climate models (Béguin et al. 2013).
The link between TC activity, which is often described in terms of accumulated cyclone energy (ACE) (Bell et al. 2000), and PV streamers as a proxy for Rossby wave breaking has been investigated in several publications. Li et al. (2018) found that Rossby wave breaking events are generally uncorrelated with ACE on a short time scale of 8 days, using their entire data dataset from 1985 to 2013, but also found that they correlate during a number of individual years, which may point toward interannual variability of correlation. On a seasonal time scale, ACE correlates substantially and negatively with Rossby wave breaking frequency and the area of the involved PV streamers (Zhang et al. 2017), and a combined measure which takes into account frequency, size and the magnitude of the anomaly of PV streamers (Papin et al. 2020) on the 350-K isentropic surface. Rossby wave breaking can also enable tropical cyclogenesis (Takemura and Mukougawa 2021) via tropical transition, where a precursor extratropical cyclone develops tropical characteristics, such as a warm core and axisymmetry (Davis and Bosart 2004). McTaggart-Cowan et al. (2013) found that tropical cyclogenesis via tropical transition accounts for over a third of tropical cyclogenesis events in the North Atlantic basin. Despite the potential for cyclogenesis, it appears that the overall effect of an increased Rossby wave breaking frequency is detrimental to TC activity in terms of ACE (Zhang et al. 2017). TCs formed via tropical transition have been found to be less predictable than TCs formed via other genesis pathways (Wang et al. 2018), which suggests that changes in Rossby wave breaking frequency may affect predictability throughout a TC season.
The aim of this study is to assess if the latitude of the 2-PVU contour on multiple isentropic surfaces can potentially act as a predictor for seasonal ACE. The effects on environmental vertical wind shear and midtropospheric relative humidity are assessed and compared to the effects caused by Rossby wave breaking. Furthermore, it is investigated if changes in the position of the 2-PVU contour have an impact on other TC-related metrics concerning number and lifetime.
The following sections are organized as follows. Section 2 describes the methods and data used in this study. Section 3 describes the statistical link between the 2-PVU contour latitude and ACE. Section 4 describes the impact that the 2-PVU contour latitude exerts on environmental variables, and section 5 describes the exerted impact on TC count and lifetime metrics. Section 6 summarizes the results and states the key conclusions.
2. Data and methods
For the purpose of this study, the North Atlantic hurricane season is defined as beginning at 0000 UTC 1 June and ending at 0000 UTC 1 December. Only the North Atlantic basin is considered.
HURDAT2 data are further used to derive other metrics for seasonal activity. The first is the number of individual days per season where a cyclone of hurricane strength (categorized in HURDAT2 as HU) exists within the North Atlantic basin (hurricane days). The second is the number of individual days per season where a cyclone of at least tropical storm strength (categorized in HURDAT2 as TS or HU) exists within the North Atlantic basin (storm days), which thus includes hurricane days. The third is the number of named storms. The fourth is the number of named storms that make landfall, where named storms that make landfall multiple times are counted only once. The fifth is the lifetime of TCs, where the number of entries at synoptic hours and at which the TC is categorized as TD (tropical depression), TS or HU is used. All named terms that contain the word “storm” refer exclusively to TC systems, and not any other storm system, within this study.
Genesis location density is defined by first dividing the North Atlantic basin into squares of 5° side length. HURDAT2 data are then used to obtain the first entry for each cyclone in a set of years, as motivated in the corresponding section. The resulting number of cyclones per square is divided by the number of included cyclones, as this allows for a more meaningful comparison of two distinct sets of years with a different number of total cyclones.
ERA5 reanalysis data (Hersbach et al. 2020) are used to obtain PV, wind and relative humidity data. PV is interpolated to a 0.5° × 0.5° grid and to isentropic surfaces from 345 to 365 K in 5-K intervals. The algorithm used in Sprenger et al. (2017) is used to identify the 2-PVU contour on isentropic surfaces in terms of equidistant points along the contour. The latitude of the 2-PVU contour, here dubbed
The PV streamer detection algorithm of Sprenger et al. (2017), which is a refined version of the Wernli and Sprenger (2007) algorithm, is used to obtain PV streamer frequencies with a resolution of 0.5° × 0.5°. Other methods of obtaining PV streamer frequency, in particular the one used in Papin et al. (2020), identify the centroid of the streamer, and obtain from it the geographical distribution of the frequency. In comparison, the method of Sprenger et al. (2017) identifies all points on the specified grid where a PV streamer is present, and not only the centroid position. The frequency is defined as the ratio of time steps at which a PV streamer is present at a given location and the total number of time steps in the considered time interval. PV streamers are used as a proxy for anticyclonic Rossby wave breaking.
Various pressure levels have been used to describe midtropospheric humidity. Within this study, it is quantified by total precipitable water between 850 and 200 hPa in order to not be limited to a single or only few pressure levels. Here 850 and 200 hPa are chosen be consistent with the levels used in Zhang et al. (2016).
The main development region (MDR) is defined as the region from 10° to 20°N and from 20° to 80°W (Goldenberg and Shapiro 1996). The MDR is split into an eastern MDR (EMDR) and a western MDR (WMDR), with the border being at 50°W. Further, a high intensity region (HIR) is defined to represent the region where major hurricanes occur, based on Knapp et al. (2010). The corners of this region are at 30°N, 50°W; 30°N, 100°W; 21°N, 100°W; 10°N, 83°W; and 10°N,50°W, such that it encompasses the Gulf of Mexico and the western tropical North Atlantic basin. Both regions are shown in Fig. 1.

Mean 2-PVU contour in September on the 345–365-K isentropic surfaces using 1980–2017 climatology (solid) and 2013 (dashed) ERA5 data between 100° and 20°W. Gray regions denote the HIR and MDR, with the darker shading indicating the overlap of the two regions.
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1

Mean 2-PVU contour in September on the 345–365-K isentropic surfaces using 1980–2017 climatology (solid) and 2013 (dashed) ERA5 data between 100° and 20°W. Gray regions denote the HIR and MDR, with the darker shading indicating the overlap of the two regions.
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
Mean 2-PVU contour in September on the 345–365-K isentropic surfaces using 1980–2017 climatology (solid) and 2013 (dashed) ERA5 data between 100° and 20°W. Gray regions denote the HIR and MDR, with the darker shading indicating the overlap of the two regions.
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
3. as a predictor for ACE
This section explores
Figure 2 shows the correlation coefficient of monthly mean

Linear correlation coefficients of
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1

Linear correlation coefficients of
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
Linear correlation coefficients of
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
On the 350-K isentropic surface during September there is substantial correlation of

ERA5 1980–2017 climatological PV streamer frequency on the (top) 350- and (bottom) 360-K isentropic surface during (left) August and (right) September.
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1

ERA5 1980–2017 climatological PV streamer frequency on the (top) 350- and (bottom) 360-K isentropic surface during (left) August and (right) September.
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
ERA5 1980–2017 climatological PV streamer frequency on the (top) 350- and (bottom) 360-K isentropic surface during (left) August and (right) September.
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
On the 360-K isentropic surface, the correlation coefficient of September
On the 350-K isentropic surface, June
The
Autocorrelation coefficients of


On the 350-K isentropic surface, June
On the 360-K isentropic surface, the autocorrelation coefficient of June
In summary, September
4. The link between and environmental variables
The mechanisms by which
Zhang et al. (2016) argue that Rossby wave breaking leads to a reduction in humidity and an increase in vertical wind shear in the tropical North Atlantic. Multiple linear regression analysis based on 1980–2017 ERA5 data is thus used to assess whether there is a link between September
a. The HIR and the MDR
Table 2 summarizes the regression results for the HIR and MDR. Using only vertical wind shear as a predictor and using the 350-K isentropic surface, both regions show substantial correlation (R2 > 0.4, rows 1 and 2) of
Multiple linear regression results for September mean


The lack of correlation with precipitable water in the MDR differs strongly from the findings of Zhang et al. (2016), who state that humidity changes are linked to Rossby wave breaking events on the 350-K isentropic surface. However, this apparent contradiction can be reconciled. Zhang et al. (2016) use the month of August, and not September, in their analysis. Repeating the multiple linear regression analysis for August reveals that using precipitable water as a sole predictor yields a similar R2 value as when using vertical wind shear as a sole predictor (R2 = 0.23 and 0.27, respectively). Further, using precipitable water as the sole predictor for September results in a lower R2 value (R2 = 0.15), indicating that the link between
A plausible reason for the decline of precipitable water as a predictor for
Using the 360-K isentropic surface (rows 5–8), vertical wind shear remains the dominant predictor. However, using precipitable water as an additional predictor in the MDR increases the R2 value from 0.335 to 0.389, and the p value associated with precipitable as a predictor is quite low with p = 0.014. Precipitable water thus appears to be responsive to changes in
b. The WMDR and EMDR
Table 3 shows the multiple linear regression results of the WMDR and EMDR subregions. Using the 350-K isentropic surface and vertical wind shear as the sole predictor (rows 1 and 2), the correlation of
Using precipitable water as an additional predictor in the WMDR reduces the R2 value very slightly, and has an associated p value in excess of 0.4. Precipitable water changes due to changes in
Using the 360-K isentropic surface and vertical wind shear as the sole predictor, the R2 values decrease substantially in both subregions. However, using precipitable water as an additional predictor in the WMDR, the resulting R2 value of 0.48 is rather close to those predicting the 350-K isentropic surface in the WMDR (R2 = 0.46) and in the entire MDR (R2 = 0.49). Therefore, when using the 360-K isentropic surface, both vertical wind shear and precipitable water in the WMDR are important predictors for
c. PV streamer climatology on the 350- and 360-K isentropic surfaces
The PV streamer frequency in the North Atlantic region on the 360-K isentropic surface is shown in Fig. 3 compared to the frequency on the 350-K isentropic surface, as identified in ERA5 data during the 1980–2017 period. Note that the Sprenger et al. (2017) algorithm is used to identify PV streamers, which does not use the centroid position, but the entire area of the streamer feature. Therefore, the southward extent of streamers is represented more accurately than when the centroid position is used (see also section 2).
On the 350-K isentropic surface, the PV streamer frequency is reduced from August to September, and the maximum is shifted eastward, as also found by, e.g., Papin et al. (2020) tracking PV streamer centroid positions. PV streamer activity is thus shifted away from where TC activity is high (Knapp et al. 2010). On the 360-K isentropic surface, Abatzoglou and Magnusdottir (2006) show an increase in the number of Rossby wave breaking events from August to September in the North Atlantic region. However, this may be a result of their counting method. They counted events on the 360-K isentropic surface only if the event was not also present on the 350-K isentropic surface, such that an increase in the number of events on the 360-K isentropic surface may also be due to a reduction in vertical extent of events and not a true increase in the number of events. Figure 3 shows that there is a genuine increase in PV streamer frequency from August to September on the 360-K isentropic surface, as the region where PV streamers occur shifts eastward from the Pacific into the North Atlantic region. Notably, PV streamer frequency is maximal between 10° and 20°N, and is thus shifted into the WMDR. This strongly implies increased Rossby wave breaking in the WMDR during September, which enables mixing, and thus the comparatively strong response of precipitable water to changes in
d. quartiles on the 360-K isentropic surface
As 38 years are used within this study (1980–2017), the 9 years with the northernmost September
Figure 4 shows the PV streamer frequency for the southernmost and northernmost

Mean September PV streamer frequency of ERA5 1980–2017 (top) southernmost and (bottom) northernmost
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1

Mean September PV streamer frequency of ERA5 1980–2017 (top) southernmost and (bottom) northernmost
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
Mean September PV streamer frequency of ERA5 1980–2017 (top) southernmost and (bottom) northernmost
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1

(top) Mean September 500-hPa relative humidity of the ERA5 1980–2017 period and (bottom) the difference of the northernmost and southernmost
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1

(top) Mean September 500-hPa relative humidity of the ERA5 1980–2017 period and (bottom) the difference of the northernmost and southernmost
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
(top) Mean September 500-hPa relative humidity of the ERA5 1980–2017 period and (bottom) the difference of the northernmost and southernmost
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
Figure 6 shows a September mean 80°–50°W (i.e., the WMDR) zonal mean of zonal wind speed, the 2-PVU contour and the 350- and 360-K isentropic surfaces for the 1980–2017 period and the southernmost and northernmost

(left) ERA5 1980–2017 climatological September mean 80°–50°W zonal-mean zonal wind speed (black contours, from −8 to 8 m s−1, negative contours are dashed), 350- and 360-K potential temperature contours (red), and 2-PVU contour (blue line). (center) As in in the left panel, but for the nine southernmost September mean
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1

(left) ERA5 1980–2017 climatological September mean 80°–50°W zonal-mean zonal wind speed (black contours, from −8 to 8 m s−1, negative contours are dashed), 350- and 360-K potential temperature contours (red), and 2-PVU contour (blue line). (center) As in in the left panel, but for the nine southernmost September mean
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
(left) ERA5 1980–2017 climatological September mean 80°–50°W zonal-mean zonal wind speed (black contours, from −8 to 8 m s−1, negative contours are dashed), 350- and 360-K potential temperature contours (red), and 2-PVU contour (blue line). (center) As in in the left panel, but for the nine southernmost September mean
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
The reversal of upper-level wind direction between quartiles also has implications for vertical wind shear. The 850-hPa winds are easterly in the climatological mean as well as in the discussed quartiles. As the 200-hPa winds in the WMDR are mostly easterly in the northernmost

As in Fig. 5, but showing 200–850-hPa vertical wind shear.
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1

As in Fig. 5, but showing 200–850-hPa vertical wind shear.
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
As in Fig. 5, but showing 200–850-hPa vertical wind shear.
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
The combined effect of vertical wind shear and relative humidity, as quantified by the ventilation index, is shown in Fig. 8. Climatologically, the ventilation index has a local maximum off the northern coastline of South America, which is where there is also a climatological maximum in vertical wind shear. The climatological ventilation index is largest in the MDR at the northern boundary of the EMDR, where vertical wind shear is high and relative humidity is low. The quartile difference shows a sizeable decrease in ventilation index for the northernmost

As in Fig. 5, but showing ventilation index in the MDR. The 99% confidence level is omitted to make the figure easier to read.
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1

As in Fig. 5, but showing ventilation index in the MDR. The 99% confidence level is omitted to make the figure easier to read.
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
As in Fig. 5, but showing ventilation index in the MDR. The 99% confidence level is omitted to make the figure easier to read.
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
The connection between September mean

September mean 2-PVU contour position for 1980–2017 (solid black), the years with the nine most equatorward September mean
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1

September mean 2-PVU contour position for 1980–2017 (solid black), the years with the nine most equatorward September mean
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
September mean 2-PVU contour position for 1980–2017 (solid black), the years with the nine most equatorward September mean
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
e. Redundancy of isentropic levels as predictors
The
Multiple linear regression analysis is again used to assess whether using both isentropic surfaces is superior to the use of only one. The square root of ACE is predicted instead of ACE to de-trend the residuals of the regression. This is necessary to ensure that the regression properly captures the relation between the dependent variable and the predictors. The latitudes where the correlation coefficient of
As in Table 2, but predicting the square root of ACE with September mean


Using the 360-K isentropic surfaces yields a far larger R2 value than using the 350-K isentropic surface (R2 = 0.50 versus R2 = 0.32). When both isentropic surfaces are used, the R2 value only increases very slightly. Further, the p value associated with the 350-K isentropic surface as a predictor is rather large (p > 0.2). There is thus no tangible benefit to using both isentropic surfaces, and the 360-K isentropic surface appears to be the superior predictor.
5. Impact on storm number and lifetime
From the definition of ACE, it follows that not only a reduced intensification via environmental variables can impact ACE, but also a reduction in named storm number or hurricane lifetime. This section assesses whether changes in September
Table 5 summarizes the correlation coefficients of September
Correlation coefficients of September mean


For all metrics, the difference from one isentropic surface to the other is generally rather small, with the ratio of TCs that make landfall showing the largest increase in correlation coefficient from 0.26 to 0.36. As all coefficients are positive, a more poleward
The change in correlation coefficient with the ratio of TCs that make landfall is of particular note, as the associated p value decreases from 0.11 on the 350-K isentropic surface to 0.02 on the 360-K isentropic surface. The change in the number of TCs that make landfall is therefore not only a result of there being more named storms when
The number of storm days and hurricane days can also be considered for September only, as these metrics do not cross from one month into the next. On both considered isentropic surfaces, the correlation coefficients are reduced slightly. On the 350-K isentropic surface, they are reduced to 0.58 and 0.52 for storm days and hurricane days, respectively, down from 0.67 to 0.60 for the entire North Atlantic hurricane season. On the 360-K isentropic surface, they are similarly reduced to 0.55 and 0.66 for storm days and hurricane days, respectively, from 0.66 to 0.69 for the entire North Atlantic hurricane season. The number of storm days and hurricane days are linked to the intensity of cyclones, which in turn is linked to the response of vertical wind shear and relative humidity patterns to
The lifetime of hurricanes can be affected by the location of their genesis. The northernmost and southernmost

Difference in genesis location density of the northernmost
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1

Difference in genesis location density of the northernmost
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
Difference in genesis location density of the northernmost
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
The bottom panel of Fig. 10 shows the difference in the equivalent genesis location density using only TCs that were formed in September. The favored cyclogenesis location to the north of the MDR for more equatorward
Table 6 shows the metrics used for correlations in Table 5, with the absolute numbers listed for the northernmost and southernmost

Tracks of TCs generated in September for (top) the nine southernmost September
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1

Tracks of TCs generated in September for (top) the nine southernmost September
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
Tracks of TCs generated in September for (top) the nine southernmost September
Citation: Journal of Climate 36, 8; 10.1175/JCLI-D-22-0479.1
Number of storm days (SD), hurricane days (HD), and named storms (NS), the lifetime of hurricanes (LTHU, in number of synoptic time steps), the number of TCs that make landfall (LF), and the ratio of TCs that make landfall to the number of named storms (qLF) for the northernmost


6. Summary and conclusions
The metric
Multiple linear regression analysis was used to assess the link between September
The northernmost and southernmost quartiles of
Finally, the correlation of September
We recommend conducting further research on how specifically the
Acknowledgments.
The authors thank the three anonymous reviewers for their feedback, which has been used to improve the quality of the manuscript. ERA5 data were downloaded from the Copernicus Climate Change Service (C3S) Climate Data Store. The results contain modified Copernicus Climate Change Service information 2020. Neither the European Commission nor ECMWF is responsible for any use that may be made of the Copernicus information or data it contains.
Data availability statement.
All used ERA5 data are available at Copernicus Climate Change Service (C3S) (2017): ERA5: Fifth generation of ECMWF atmospheric reanalyses of the global climate, Copernicus Climate Change Service Climate Data Store (CDS). Monthly averaged data on single levels: https://doi.org/10.24381/cds.6860a573; monthly averaged data on pressure levels: https://doi.org/10.24381/cds.6860a573; and hourly data on pressure levels: https://doi.org/10.24381/cds.bd0915c6. The HURDAT2 dataset is available at the NHC Data Archive under https://www.nhc.noaa.gov/data/.
REFERENCES
Abatzoglou, J. T., and G. Magnusdottir, 2006: Planetary wave breaking and nonlinear reflection: Seasonal cycle and interannual variability. J. Climate, 19, 6139–6152, https://doi.org/10.1175/JCLI3968.1.
Aiyyer, A. R., and C. Thorncroft, 2006: Climatology of vertical wind shear over the tropical Atlantic. J. Climate, 19, 2969–2983, https://doi.org/10.1175/JCLI3685.1.
Barnes, E. A., and D. L. Hartmann, 2012: Detection of Rossby wave breaking and its response to shifts of the midlatitude jet with climate change. J. Geophys. Res., 117, D09117, https://doi.org/10.1029/2012JD017469.
Béguin, A., O. Martius, M. Sprenger, P. Spichtinger, D. Folini, and H. Wernli, 2013: Tropopause level Rossby wave breaking in the Northern Hemisphere: A feature-based validation of the ECHAM5-HAM climate model. Int. J. Climatol., 33, 3073–3082, https://doi.org/10.1002/joc.3631.
Bell, G. D., and Coauthors, 2000: Climate assessment for 1999. Bull. Amer. Meteor. Soc., 81 (6), S1–S50, https://doi.org/10.1175/1520-0477(2000)81[s1:CAF]2.0.CO;2.
Bender, M. A., T. R. Knutson, R. E. Tuleya, J. J. Sirutis, G. A. Vecchi, S. T. Garner, and I. M. Held, 2010: Modeled impact of anthropogenic warming on the frequency of intense Atlantic hurricanes. Science, 327, 454–458, https://doi.org/10.1126/science.1180568.
Bengtsson, L., H. Böttger, and M. Kanamitsu, 1982: Simulation of hurricane-type vortices in a general circulation model. Tellus, 34A, 440–457, https://doi.org/10.3402/tellusa.v34i5.10830.
Bengtsson, L., M. Botzet, and M. Esch, 1995: Hurricane-type vortices in a general circulation model. Tellus, 47A, 175–196, https://doi.org/10.3402/tellusa.v47i2.11500.
Bister, M., and K. A. Emanuel, 2002 : Low frequency variability of tropical cyclone potential intensity 1. Interannual to interdecadal variability. J. Geophys. Res., 107, 4801, https://doi.org/10.1029/2001JD000776.
Bowley, K. A., J. R. Gyakum, and E. H. Atallah, 2019: A new perspective toward cataloging Northern Hemisphere Rossby wave breaking on the dynamic tropopause. Mon. Wea. Rev., 147, 409–431, https://doi.org/10.1175/MWR-D-18-0131.1.
Bryan, G. H., 2008: On the computation of pseudoadiabatic entropy and equivalent potential temperature. Mon. Wea. Rev., 136, 5239–5245, https://doi.org/10.1175/2008MWR2593.1.
Davis, C. A., and L. F. Bosart, 2004: The TT problem: Forecasting the tropical transition of cyclones. Bull. Amer. Meteor. Soc., 85, 1657–1662, https://doi.org/10.1175/BAMS-85-11-1657.
Emanuel, K. A., 1986: An air–sea interaction theory for tropical cyclones. Part I: Steady-state maintenance. J. Atmos. Sci., 43, 585–605, https://doi.org/10.1175/1520-0469(1986)043<0585:AASITF>2.0.CO;2.
Emanuel, K. A., C. DesAutels, C. Holloway, and R. Korty, 2004: Environmental control of tropical cyclone intensity. J. Atmos. Sci., 61, 843–858, https://doi.org/10.1175/1520-0469(2004)061<0843:ECOTCI>2.0.CO;2.
Fitzpatrick, P. J., J. A. Knaff, C. W. Landsea, and S. V. Finley, 1995: Documentation of a systematic bias in the aviation model’s forecast of the Atlantic tropical upper-tropospheric trough: Implications for tropical cyclone forecasting. Wea. Forecasting, 10, 433–446, https://doi.org/10.1175/1520-0434(1995)010<0433:DOASBI>2.0.CO;2.
Frank, W. M., and E. A. Ritchie, 2001: Effects of vertical wind shear on the intensity and structure of numerically simulated hurricanes. Mon. Wea. Rev., 129, 2249–2269, https://doi.org/10.1175/1520-0493(2001)129<2249:EOVWSO>2.0.CO;2.
Gettelman, A., W. D. Collins, E. J. Fetzer, A. Eldering, F. W. Irion, P. B. Duffy, and G. Bala, 2006: Climatology of upper-tropospheric relative humidity from the atmospheric infrared sounder and implications for climate. J. Climate, 19, 6104–6121, https://doi.org/10.1175/JCLI3956.1.
Goldenberg, S. B., and L. J. Shapiro, 1996: Physical mechanisms for the association of El Niño and West African rainfall with Atlantic major hurricane activity. J. Climate, 9, 1169–1187, https://doi.org/10.1175/1520-0442(1996)009<1169:PMFTAO>2.0.CO;2.
Gray, W. M., 1984a: Atlantic seasonal hurricane frequency. Part I: El Niño and 30 mb quasi-biennial oscillation influences. Mon. Wea. Rev., 115, 1649–1668, https://doi.org/10.1175/1520-0493(1984)112<1649:ASHFPI>2.0.CO;2.
Gray, W. M., 1984b: Atlantic seasonal hurricane frequency. Part II: Forecasting its variability. Mon. Wea. Rev., 112, 1669–1683, https://doi.org/10.1175/1520-0493(1984)112<1669:ASHFPI>2.0.CO;2.
Grinsted, A., P. Ditlevsen, and J. H. Christensen, 2019: Normalized U.S. hurricane damage estimates using area of total destruction, 1900–2018. Proc. Natl. Acad. Sci. USA, 116, 23 942–23 946, https://doi.org/10.1073/pnas.1912277116.
Hendricks, E. A., M. S. Peng, B. Fu, and T. Li, 2010: Quantifying environmental control on tropical cyclone intensity change. Mon. Wea. Rev., 138, 3243–3271, https://doi.org/10.1175/2010MWR3185.1.
Hersbach, H., and Coauthors, 2020: The ERA5 global reanalysis. Quart. J. Roy. Meteor. Soc., 146, 1999–2049, https://doi.org/10.1002/qj.3803.
Kaplan, J., and M. DeMaria, 2003: Large-scale characteristics of rapidly intensifying tropical cyclones in the North Atlantic basin. Wea. Forecasting, 18, 1093–1108, https://doi.org/10.1175/1520-0434(2003)018<1093:LCORIT>2.0.CO;2.
Klotzbach, P. J., 2007: Recent developments in statistical prediction of seasonal Atlantic basin tropical cyclone activity. Tellus, 59A, 511–518, https://doi.org/10.1111/j.1600-0870.2007.00239.x.
Klotzbach, P. J., and W. M. Gray, 2004: Updated 6–11-month prediction of Atlantic basin seasonal hurricane activity. Wea. Forecasting, 19, 917–934, https://doi.org/10.1175/1520-0434(2004)019<0917:UMPOAB>2.0.CO;2.
Klotzbach, P. J., and Coauthors, 2019: Seasonal tropical cyclone forecasting. Trop. Cyclone Res. Rev., 8, 134–149, https://doi.org/10.1016/j.tcrr.2019.10.003.
Knapp, K. R., M. C. Kruk, D. H. Levinson, H. J. Diamond, and C. J. Neumann, 2010: The International Best Track Archive for Climate Stewardship (IBTrACS). Bull. Amer. Meteor. Soc., 91, 363–376, https://doi.org/10.1175/2009BAMS2755.1.
Landsea, C. W., and J. L. Franklin, 2013: Atlantic hurricane database uncertainty and presentation of a new database format. Mon. Wea. Rev., 141, 3576–3592, https://doi.org/10.1175/MWR-D-12-00254.1.
Li, W., Z. Wang, G. Zhang, M. S. Peng, S. G. Benjamin, and M. Zhao, 2018: Subseasonal variability of Rossby wave breaking and impacts on tropical cyclones during the North Atlantic warm season. J. Climate, 31, 9679–9695, https://doi.org/10.1175/JCLI-D-17-0880.1.
Manabe, S., J. L. Holloway, and H. M. Stone, 1970: Tropical circulation in a time-integration of a global model of the atmosphere. J. Atmos. Sci., 27, 580–613, https://doi.org/10.1175/1520-0469(1970)027<0580:TCIATI>2.0.CO;2.
McTaggart-Cowan, R., T. J. Galarneau, L. F. Bosart, R. W. Moore, and O. Martius, 2013: A global climatology of baroclinically influenced tropical cyclogenesis. Mon. Wea. Rev., 141, 1963–1989, https://doi.org/10.1175/MWR-D-12-00186.1.
Papin, P. P., L. F. Bosart, and R. D. Torn, 2020: A feature-based approach to classifying summertime potential vorticity streamers linked to Rossby wave breaking in the North Atlantic basin. J. Climate, 33, 5953–5969, https://doi.org/10.1175/JCLI-D-19-0812.1.
Pielke, R. A., Jr., J. Gratz, C. W. Landsea, D. Collins, M. A. Saunders, and R. Musulin, 2008: Normalized hurricane damage in the United States: 1900–2005. Nat. Hazards Rev., 9, 29–42, https://doi.org/10.1061/(ASCE)1527-6988(2008)9:1(29).
Postel, G. A., and M. H. Hitchman, 1999: A climatology of Rossby wave breaking along the subtropical tropopause. J. Atmos. Sci., 56, 359–373, https://doi.org/10.1175/1520-0469(1999)056<0359:ACORWB>2.0.CO;2.
Riemer, M., and F. Laliberté, 2015: Secondary circulation of tropical cyclones in vertical wind shear: Lagrangian diagnostic and pathways of environmental interaction. J. Atmos. Sci., 72, 3517–3536, https://doi.org/10.1175/JAS-D-14-0350.1.
Riemer, M., M. T. Montgomery, and M. E. Nicholls, 2010: A new paradigm for intensity modification of tropical cyclones: Thermodynamic impact of vertical wind shear on the inflow layer. Atmos. Chem. Phys., 10, 3163–3188, https://doi.org/10.5194/acp-10-3163-2010.
Saunders, M. A., and A. S. Lea, 2005: Seasonal prediction of hurricane activity reaching the coast of the United States. Nature, 434, 1005–1008, https://doi.org/10.1038/nature03454.
Scott, R. K., and J.-P. Cammas, 2002: Wave breaking and mixing at the subtropical tropopause. J. Atmos. Sci., 59, 2347–2361, https://doi.org/10.1175/1520-0469(2002)059<2347:WBAMAT>2.0.CO;2.
Sprenger, M., and Coauthors, 2017: Global climatologies of Eulerian and Lagrangian flow features based on ERA-Interim. Bull. Amer. Meteor. Soc., 98, 1739–1748, https://doi.org/10.1175/BAMS-D-15-00299.1.
Takemura, K., and H. Mukougawa, 2021: Tropical cyclogenesis triggered by Rossby wave breaking over the western North Pacific. SOLA, 17, 164–169, https://doi.org/10.2151/sola.2021-029.
Tang, B., and K. A. Emanuel, 2010: Midlevel ventilation’s constraint on tropical cyclone intensity. J. Atmos. Sci., 67, 1817–1830, https://doi.org/10.1175/2010JAS3318.1.
Tang, B., and K. A. Emanuel, 2012: A ventilation index for tropical cyclones. Bull. Amer. Meteor. Soc., 93, 1901–1912, https://doi.org/10.1175/BAMS-D-11-00165.1.
Tang, B., and S. J. Camargo, 2014: Environmental control of tropical cyclones in CMIP5: A ventilation perspective. J. Adv. Model. Earth Syst., 6, 115–128, https://doi.org/10.1002/2013MS000294.
Thorncroft, C., and I. Pytharoulis, 2001: A dynamical approach to seasonal prediction of Atlantic tropical cyclone activity. Wea. Forecasting, 16, 725–734, https://doi.org/10.1175/1520-0434(2002)016<0725:ADATSP>2.0.CO;2.
Vecchi, G. A., and Coauthors, 2014: On the seasonal forecasting of regional tropical cyclone activity. J. Climate, 27, 7994–8016, https://doi.org/10.1175/JCLI-D-14-00158.1.
Vitart, F., and Coauthors, 2007: Dynamically-based seasonal forecasts of Atlantic tropical storm activity issued in June by EUROSIP. Geophys. Res. Lett., 34, L16815, https://doi.org/10.1029/2007GL030740.
Wang, Z., W. Li, M. S. Peng, X. Jiang, R. McTaggart-Cowan, and C. A. Davis, 2018: Predictive skill and predictability of North Atlantic tropical cyclogenesis in different synoptic flow regimes. J. Atmos. Sci., 75, 361–378, https://doi.org/10.1175/JAS-D-17-0094.1.
Wernli, H., and M. Sprenger, 2007: Identification and ERA-15 climatology of potential vorticity streamers and cutoffs near the extratropical tropopause. J. Atmos. Sci., 64, 1569–1586, https://doi.org/10.1175/JAS3912.1.
Wu, L., and Coauthors, 2012: Relationship of environmental relative humidity with North Atlantic tropical cyclone intensity and intensification rate. Geophys. Res. Lett., 39, L20809, https://doi.org/10.1029/2012GL053546.
Yang, W., T.-L. Hsieh, and G. A. Vecchi, 2021: Hurricane annual cycle controlled by both seeds and genesis probability. Proc. Natl. Acad. Sci. USA, 118, e2108397118, https://doi.org/10.1073/pnas.
Zhang, G., Z. Wang, T. J. Dunkerton, M. S. Peng, and G. Magnusdottir, 2016: Extratropical impacts on Atlantic tropical cyclone activity. J. Atmos. Sci., 73, 1401–1418, https://doi.org/10.1175/JAS-D-15-0154.1.
Zhang, G., Z. Wang, M. S. Peng, and G. Magnusdottir, 2017: Characteristics and impacts of extratropical Rossby wave breaking during the Atlantic hurricane season. J. Climate, 30, 2363–2379, https://doi.org/10.1175/JCLI-D-16-0425.1.
Zhang, G., H. Murakami, R. Gudgel, and X. Yang, 2019: Dynamical seasonal prediction of tropical cyclone activity: Robust assessment of prediction skill and predictability. Geophys. Res. Lett., 46, 5506–5515, https://doi.org/10.1029/2019GL082529.