You are on page 1of 22

Environmental control of tropical, subtropical, and extratropical

cyclone development over the North Atlantic Ocean: Idealized


numerical experiments

Wataru Yanase∗ and Hiroshi Niino


Atmosphere and Ocean Research Institute, The University of Tokyo, Kashiwa, Japan
Accepted Article


Correspondence to: Atmosphere and Ocean Research Institute, The University of Tokyo, Kashiwa, Japan, 5-1-5, Kashiwanoha,

Kashiwa, Chiba 277-8564, Japan. E-mail:wtryanase@gmail.com

The North Atlantic Ocean in autumn is characterized by a variety of cyclone activity:


tropical, subtropical, and extratropical cyclones are meridionally distributed from low
to high latitudes, and cyclones are more active in the western part than in the eastern
part. To examine to what degree the environmental fields explain the geographical
distribution of the various cyclones, idealized numerical experiments are conducted
using climatological environmental fields at different longitudes and latitudes. The
experiments qualitatively reproduce the geographical distribution of development and
types of cyclones, which enables an intercomparison of different types of cyclones on
the basis of their structure, energy budget analysis, and sensitivity experiments. It
is notable that subtropical cyclones can develop spontaneously in the climatological
environment which does not have an upper tropospheric disturbance in the initial fields.
While the simulated subtropical cyclones are sensitive to initial conditions, some of
them have structures that are in good agreement with those of observed subtropical
cyclones: a shallow warm-core structure, concentrated convection near the cyclone
centre embedded in synoptic-scale cloud on the north and east sides, and a narrow mid-
tropospheric potential vorticity anomaly near the cyclone centre under a wide upper-
tropospheric anomaly on the west side. The subtropical cyclones exhibit characteristics
intermediate between those of tropical cyclones and extratropical cyclones, notably
multi-scale structures with tropical characteristics on a small scale embedded in
extratropical characteristics on a large scale.

This article has been accepted for publication and undergone full peer review but has not
been through the copyediting, typesetting, pagination and proofreading process, which
may lead to differences between this version and the Version of Record. Please cite this
article as doi: 10.1002/qj.3227

This article is protected by copyright. All rights reserved.


Key Words: subtropical cyclone; hybrid cyclone; tropical cyclone; extratropical cyclone; idealized experiment; non-

hydrostatic model; North Atlantic Ocean

1. Introduction

Satellite images and weather charts show diversity in synoptic-scale cyclones developing across the global oceans. Two well-known
Accepted Article
cyclones are tropical cyclones (TCs) with organized cumulus convection at the low latitudes and extratropical cyclones (ECs) with

warm- and cold-frontal systems at the mid-latitudes. A TC develops through a positive feedback between vortex circulation and

diabatic processes including condensational heating over the warm oceans (Charney and Eliassen 1964; Ooyama 1969; Emanuel

1986; Montgomery and Smith 2014), whereas an EC develops through baroclinic processes within the region of large baroclinicity

(horizontal temperature gradient) in the mid-latitudes (Eady 1949; Charney 1949; Hoskins 1976; Catto 2016).

In addition to typical TCs and ECs, there are hybrid or intermediate types of cyclones between TCs and ECs. According to the

Glossary of Meteorology (American Meteorological Society 2017), subtropical cyclones (SCs) extract available potential energy

through baroclinic process, as ECs do, but they also receive some or most of their energy from condensational heating, as TCs do.

Guishard et al. (2009) showed that the increase in SC activity over the North Atlantic Ocean in September and October are consistent

with the environmental field in which the region of high sea surface temperature (SST) is overlapped with that of large baroclinicity. In

addition to the environmental fields, SCs often develop under the transient influence of upper tropospheric troughs (Evans and Guishard

2009; Bentley et al. 2016, 2017). SCs sometimes become TCs (Davis and Bosart 2003; Mauk and Hobgood 2012; Bentley et al. 2016)

through the tropical transition process (Davis and Bosart 2004; McTaggart-Cowan et al. 2013).

To understand the various types of synoptic-scale cyclones in a unified framework, Hart (2003) proposed a cyclone classification

algorithm called “cyclone phase space (CPS).” The CPS assesses cyclone types objectively based on their structure: a TC is a deep

warm-core cyclone, whereas an EC is a deep cold-core cyclone. An SC is intermediate between a TC and an EC, which has a shallow

warm-core (a warm-core in the lower troposphere but a cold-core in the upper troposphere) structure (Evans and Guishard 2009).

Yanase et al. (2014) applied the CPS classification to a Japanese reanalysis, JRA-25 (Onogi et al. 2007), and obtained a comprehensive

climatology of developing TCs, ECs and hybrid cyclones including SCs. They showed diverse cyclone activity over the North Atlantic

Ocean in autumn (Fig. 1). Firstly, there is a clear zonal contrast in cyclone activity around the subtropics: cyclones are more active

(develop more frequently) in the western part than in the eastern part. Secondly, hybrid cyclones are active in the subtropics, which is

in agreement with a number of previous studies on SCs and tropical transitions (Guishard et al. 2009; McTaggart-Cowan et al. 2013;

Bentley et al. 2016).

The geographical distribution of cyclone development is considered to be controlled by various processes. Cyclones can extract

energy from favourable environmental fields such as high SST and large baroclinicity. They are sometimes affected by transient external

forcing such as upper tropospheric troughs (Deveson and Browning 2002; Jones et al. 2003; Fischer et al. 2017), and tropical waves

(Ritchie and Holland 1999; Frank and Roundy 2006). The geographical distribution of cyclone development may also depend on the

life cycles of cyclones including extratropical transitions (Jones et al. 2003; Evans et al. 2017) and tropical transitions. Since these

processes work simultaneously and intricately in the real atmosphere, it is necessary to assess the individual processes independently

for understanding the cyclone distribution.

The present study focuses on the influence of environmental fields on spontaneous cyclone development that assumes no external

forcing such as transient upper tropospheric troughs. This basic subject has been developed by a number of theoretical studies

with different levels of simplification of cyclone dynamics: whereas the simplest models including linear models for ECs (Eady
This article is protected by copyright. All rights reserved.
Accepted Article

Figure 1. Climatological distribution of developing cyclones (crosses) obtained from the JRA-25 reanalysis using the tracking algorithm and CPS classification (Yanase
et al. 2014). Red, blue, green and grey crosses indicate the development of TCs, ECs, hybrid cyclones including SCs, and the others which does not fit the definitions,
respectively. Black circles show longitudes and latitudes for which the idealized experiments examine the influence of environmental fields on cyclone development; the
characters T, S, E and I are the locations of representative experiments for TCs, SCs, ECs, and inactive cyclones, respectively.

1949; Charney 1949) and axisymmetric models for TCs (Charney and Eliassen 1964; Ooyama 1969) are easy to understand, three-

dimensional high-resolution models with full physics can simulate multi-scale dynamics of cyclones (Hendricks et al. 2004; Coronel

et al. 2016). The latter are also useful for comparing different types of cyclones, because they assume no limitations on cyclone

structures such as a single wave-length or axisymmetry. Taking advantage of three-dimensional idealized experiments, Yanase and

Niino (2015) demonstrated that the climatological environments can reproduce the bimodal distribution of cyclones over the South

Indian Ocean in summer: TCs and ECs are active in the tropics and extratropics, respectively, whereas cyclones are inactive in the

subtropics. They assessed the influence of individual environmental factors including temperature, humidity, and vertical shear on

cyclone development quantitatively by sensitivity experiments. However, as the cyclone activity in the subtropics is different among

the oceans, a further study is necessary for understanding the cyclone dynamics in the intermediate environment between the tropics

and the extratropics.

In the present study, we apply the idealized experiments similar to Yanase and Niino (2015) to the North Atlantic Ocean in autumn,

which has more complex distribution of cyclones than the South Indian Ocean in summer, to address the following two questions. The

first question is “to what extent do the climatological environmental fields explain the geographical distribution of the development

of various types of cyclones?” It is a challenge to see if various cyclones, particularly SCs, are reproduced in a highly simplified

atmosphere without any external forcing. The set of experiments will enable an intercomparison between TCs, SCs and ECs in the

same framework of experimental designs using the same analysis methods. The second question is “what are the structures and

dynamics of the SCs in the simplified atmosphere with minimal ingredients (condensational heating and uniform baroclinicity)?”

While there have been a number of idealized experiments on the development of TCs and ECs within simplified environmental fields

from various perspectives, idealized experiments on SCs have been conducted only by Davis (2010) to our knowledge; he simulated the

development of SCs on the south side of the mid-latitude jet stream, which incorporates cross-latitude influence between the subtropics

and the extratropics. Given increasing attention to SCs and their complex mechanism, more idealized experiments will promote an

understanding of SCs. Whereas Davis (2010) considered the environmental fields that are formed by convective heating during the

early time of the simulation, we examine the environmental fields based on the climatology of the real atmosphere.

The remainder of the present paper is organized as follows. Section 2 describes experimental design including the specification of

a nonhydrostatic model and environmental fields used for the numerical experiments. Section 3 presents the result of the idealized

experiments. Section 4 compares the cyclone characteristics between the idealized experiments and the real atmosphere, and discusses

the relationship between cyclone dynamics and environmental fields. Finally, section 5 summarizes the diversity in cyclone dynamics

over the North Atlantic Ocean in autumn.


This article is protected by copyright. All rights reserved.
Table 1. The Coriolis parameter and SST for the representative tropical (Ta), subtropical (Sa), extratropical (Ea) and inactive (Ia) experiments.

Environmental factor (unit) Ta Sa Ea Ia


The Coriolis parameter (×10−5 s−1 ) 3.8 7.3 10.3 5.0
Sea surface temperature (◦ C) 27.8 25.5 15.8 26.0

2. Methodology

2.1. Numerical model


Accepted Article

As in our previous study (Yanase and Niino 2015), three-dimensional numerical experiments are conducted using the Japan

Meteorology Agency Nonhydrostatic Model (JMA-NHM; Saito et al. 2006). The cloud processes are calculated by a bulk-type

microphysics scheme that considers water vapour, cloud water, cloud ice, rain, snow and graupel (Lin et al. 1983; Murakami 1990)

together with a convective parameterization scheme (Kain and Fritsch 1990). The planetary boundary layer is represented by the level-3

closure of Mellor-Yamada-Nakanishi-Niino turbulence scheme (Nakanishi and Niino 2004) with the surface-layer scheme developed

by Beljaars (1995). Radiative process and ocean feedback are excluded for simplicity.

The model domain is 8000 km, 3000 km and 24.52 km in the zonal, meridional and vertical directions, respectively. The zonal

boundary conditions are cyclic, whereas the meridional boundaries are free-slip and adiabatic walls. The horizontal grid spacing is 10

km throughout the model domain, whereas the vertical grid spacing increases linearly from 40 m at the lowest level to 1360 m at the

highest level of the domain.

2.2. Environmental factors

To examine cyclone development in different regions over the North Atlantic Ocean in autumn, we conduct 45 experiments using

the environmental fields at longitudes with a 10◦ interval between 20◦ W and 70◦ W and at latitudes with a 5◦ interval between 10◦ N

and 50◦ N as shown in Fig. 1 (black circles and characters E, S, T and I). We will mainly compare the results of four representative

experiments: tropical (Ta; 50◦ W, 15◦ N), subtropical (Sa; 50◦ W, 30◦ N), extratropical (Ea; 50◦ W, 45◦ N), and inactive experiments

(Ia; 30◦ W, 20◦ N). The climatological distribution of developing cyclones (Yanase et al. 2014) identifies apparently different types of

cyclones as defined by the CPS among these regions (see the crosses with different colours in Fig. 1).

The environmental fields prescribed in the idealized experiments are climatological fields for 30 years (1982 - 2011) which are

obtained from the JRA-25 atmospheric reanalysis (Onogi et al. 2007) and the NOAA’s optimally interpolated SST product (Reynolds

et al. 2002). The autumn environment is the average between September and November.

We consider five environmental factors: the Coriolis parameter, SST, and vertical profiles of potential temperature, relative humidity,

and zonal wind. The SST and three vertical profiles are averaged within the area of 20◦ in the zonal and meridional directions with its

centre at the longitude and latitude of interest, because the environmental field for a cyclone is considered to be larger than the cyclone

itself. To remove fluctuation of the vertical shear of zonal wind within the boundary layer, the zonal wind at the altitudes below 1 km

is extrapolated using the average gradient between the altitudes of 1 km and 1.5 km. The zonal wind is further shifted by a constant

value throughout the vertical domain so that the zonal wind vanishes at the sea level without any change in the vertical shear. The shift

in the zonal wind is necessary for minimizing the influence of surface friction on the environmental zonal flow. The environmental

factors for the representative experiments are given in Table 1 for the Coriolis parameter and SST, and in Fig. 2 for the vertical profiles

of potential temperature, relative humidity and zonal wind. Note that the vertical shear in the troposphere is westerly not only for the

representative experiments (Fig. 2c) but also for all the experiments.
This article is protected by copyright. All rights reserved.
(a) (b) (c)
18 18 18
T
16 16 S 16
E
14 14 I 14

12 12 12
Altitude (km)

Altitude (km)

Altitude (km)
10 10 10

8 8 8

6 6 6

4 T 4 4 T
S S
Accepted Article
2 E 2 2 E
I I
0 0 0
280 300 320 340 360 380 400 420 440 0 10 20 30 40 50 60 70 80 90 100 -10 0 10 20 30 40 50
Potential Temperature (K) Relative Humidity (/%) Zonal Wind (m/s)

Figure 2. Vertical profiles of environmental fields in the representative experiments. (a) Potential temperature (K), (b) relative humidity (percent), (c) zonal wind (ms−1 ).
The red, green, blue and grey curves are profiles for Ta, Sa, Ea and Ia.

Figure 3. Structure of the environmental fields prescribed in the experiments (a) Meridional distribution of the weighting function for zonal wind (Wu ). (b) Meridional
and vertical distribution of potential temperature (shading; the unit is K) and zonal wind (contour: the interval is 2 ms−1 ) for Ea. The white hatching indicates the southern
and northern boundary zones.

2.3. Design of three-dimensional environmental fields

To construct three dimensional environmental fields for the idealized experiments, horizontal distributions for the five environmental

factors are determined. We design a uniform horizontal distribution as much as possible, which is convenient for interpretation of the

results. The zonal distribution is uniform for all of the five factors. The meridional distribution of the Coriolis parameter is also uniform

(the f -plane approximation). However, the meridional distributions of the other factors are non-uniform because of two reasons: First,

meridional gradient of potential temperature is required to satisfy the thermal wind balance with the vertical shear of the zonal wind.

Second, the meridional temperature gradient results in higher temperature near the meridional boundary (e.g. the southern boundary for

the westerly shear) than at the centre of the model domain, which causes active cumulus convection near the boundary. Such convection

is sometimes organized into convective disturbances, and makes the interpretation of the results difficult.

After a number of preliminary experiments, we have found an experimental design that can suppress cumulus convection near the

meridional boundaries. We consider “boundary zones” within 500 km (Lb ) from the northern and southern boundaries. The rest of

the meridional domain of 3000 km (Ld ) is referred to as “a central zone” (2000 km; Lc ) which gives nearly uniform environment for

cyclone development.

To reduce the meridional temperature gradient in the boundary zones, the zonal wind is reduced by multiplying a weighting

function Wu . Figure 3a shows the meridional distribution of Wu , which is given by the combination of hyperbolic tangent functions of
This article is protected by copyright. All rights reserved.
meridional coordinate y :

Wu (y) = 0.5 + 0.25 × {tanh(y − 0.5Lb )/(0.5Lb ) + tanh(Ld − 0.5Lb − y)/(0.5Lb )}

Figure 3b shows the meridional-vertical distribution of zonal wind and potential temperature for Ea as an example.

The relative humidity is uniform in most of the domain. Therefore, warmer air contains more water vapour given the Clausius-

Clapeyron equation, which can cause active cumulus convection near the warmer boundary. To suppress the convection near the
Accepted Article
warmer boundary (e.g. the southern boundary for the westerly shear), an upper limit of the water vapour in the boundary zones is

set to the maximum value in the central zone. The meridional distribution of SST is prescribed so that the difference between SST

and atmospheric surface temperature is constant. Accordingly, warmer surface air is associated with higher SST, which supplies more

water vapour to cumulus convection. To suppress the convection near the warmer boundary, the upper limit of SST in the boundary

zones is set to the maximum value in the central zone. Note that these modifications to relative humidity and SST do not result in any

discontinuities in their meridional distribution.

Furthermore, cumulus convection in the boundary zones is also suppressed by Rayleigh damping. The damping is strongest at the

boundaries where its e-folding time is 10 hours, decreases linearly with increasing the distance from the boundaries, and vanishes at

the distance Lb from the boundaries. The Rayleigh damping together with the modified meridional distribution of the environmental

factors described above effectively suppress convection in the boundary zones.

2.4. Time integration

To trigger initial development of a cyclone, an axisymmetric vortex is given at the centre of the model domain, which is superimposed

on the environmental field. The horizontal and vertical structures of the initial vortex are designed as simply as possible, because the

structure of a developing cyclone is unknown in advance. The tangential wind of the vortex is given by

 
2r/rmax z
V = Vmax 1−
1 + (r/rmax )2 zt

for the altitudes of z ≤ zt , where Vmax = 10 ms−1 , rmax = 250 km, and zt = 10 km. The radial distribution of the initial vortex is

identical to that in Emanuel and Rotunno (1989), whereas the vertical distribution is limited in the lower troposphere. The vortex is

accompanied by a potential temperature anomaly that satisfies a thermal wind balance with the tangential wind. Yanase and Niino

(2015) confirmed that maximum intensity and characteristics of TCs and ECs have little sensitivity to the structure of initial vortices.

The integration time is as long as 1200 hours, because some cyclones start to develop after several hundred hours. Section 4 discusses

the reason of slow development in the idealized experiments. The physical variables are output at every 10 hours.

During the time integration, the environmental field tends to be modified by diffusion, meso-scale convection, and synoptic-scale

cyclones. Because the modification of the environmental fields makes the interpretation of the results difficult, the environmental fields

are maintained using a spectral nudging. The spectral nudging only works on the zonally averaged fields (i.e. the wave number of zero)

so that the zonally uniform environmental fields are effectively maintained with minimal influence on cyclone dynamics. The nudging

is applied to zonal wind, potential temperature, and humidity with the e-folding time of 10 hours.

2.5. Analysis of cyclones

The horizontal grid spacing of 10 km resolves not only a synoptic-scale cyclone, but also convective systems of 10 ∼ 100 km, which

modify wind and pressure fields of the cyclone locally and temporarily, and may obscure the structure of the cyclone. Therefore, in
This article is protected by copyright. All rights reserved.
addition to the model outputs on the original grid of 10 km (referred to as “raw fields”), we analyse the “filtered fields” that do not

contain wavelength less than 500 km.

Because of the large model domain and the long integration time, the experiment may produce more than one cyclone, which

develops from the initial vortex or form spontaneously. We detect all the cyclones throughout the integration time, and focus on the most

intense cyclone. The centres of cyclones are defined by minima of sea level pressure (SLP). Then, the life-cycles of individual cyclones

are tracked using a nearest-neighbour algorithm: the algorithm simply connect the locations of cyclone centres between consecutive

time steps, if the centre at a given time step is found near the centre at the previous time step. Because small-scale convective systems
Accepted Article
may modify the location of the SLP minimum temporarily, the filtered fields are used for the cyclone tracking to reduce fluctuation

and miscalculation of tracks. On the other hand, the intensity of a cyclone is assessed using the raw fields of SLP, because the filtered

fields are too smooth. The intensity of a cyclone is defined by a minimum SLP anomaly (Pmin ), where the anomaly is defined by the

deviation from the zonal mean.

The development stage and maximum intensity of a cyclone are determined based on the 100-hour running mean of Pmin , which

effectively removes fluctuations in the raw value of the output at every 10 hours, as shown later. The “maximum intensity” of a cyclone

is defined as the lowest Pmin during its life cycle. The end of the “development stage” is the time when the product of Pmin (t) and

(Pmin (t) − Pmin (t − ∆t))/∆t is largest, where ∆t is 100 hours. The beginning of the “development stage” is the time when the

magnitude of Pmin exceeds 25 percent of that at the end of the development stage.

The cyclones are classified using the thermal wind index −VT of the CPS classification (Hart 2003), which is determined from the

vertical gradient of the difference between maximum and minimum geopotential height within the radius of 500 km from the cyclone

centre. Positive and negative values of −VT indicate a warm-core and cold-core structure, respectively. The index −VT is evaluated

over the layers in the lower (600-900 hPa; −VTL ) and upper (300-600 hPa; −VTU ) troposphere. The present study classifies cyclones

into three types: a TC for a deep warm-core structure (−VTL > 0 and −VTU > 0), an EC for a deep cold-core structure (−VTL < 0

and −VTU < 0), and an SC for a shallow warm-core structure (−VTL > 0 and −VTU < 0). Note that Yanase et al. (2014) call the third

type a hybrid cyclone, because it may include not only SCs but also tropical or extratropical transitioning cyclones within changing

environment in the real atmosphere. While Yanase et al. (2014) considered slightly different thresholds including ill-defined cyclones,

the present study classifies all the cyclones into one of the three types.

2.6. Energy budget analysis

To understand development mechanism of cyclones, the budget of eddy available potential energy (EPE) is also examined. The energy

budget analysis in a Cartesian coordinate (x,y ,z ) system is calculated based on the anelastic approximation as in Yanase and Niino

(2015). The EPE is defined as


αθ02
Pe ≡ ,
2

where θ is the potential temperature and


 −1
ρ0 g ∂(θ0 + θ)
α≡
θ0 ∂z

(ρ0 and θ0 are the reference density and potential temperature, respectively, which are functions of altitude only). The overbar denotes

zonal average, and the prime denotes the deviation from the average.

The equation for the EPE budget becomes

∂P e ∂θ αθ0 0 0 ρ0 g 0 0
= −α v 0 θ0 + qθ − w θ + Res.
∂t ∂y cp T0 θ0

This article is protected by copyright. All rights reserved.


(a) Control
0

-10

-20

-30

∆Prs (hPa)
-40

-50

-60
Ta
-70 Sa
Accepted Article
Ea
-80 Ia
0 200 400 600 800 1000 1200
Time (hr)
(b) No-shear
0

-10

-20

-30
∆Prs (hPa)

-40

-50

-60
Tu0
-70 Su0
Eu0
-80 Iu0
0 200 400 600 800 1000 1200
Time (hr)

Figure 4. Time evolution of Pmin in the representative experiments: (a) the control experiments and (b) the no-shear experiments. The red, green, blue, and grey curves
are Ta, Sa, Ea and Ia, respectively. Thin and thick curves are raw values and 100-hour running means, respectively.

where v is the meridional velocity, w is the vertical velocity, q is the condensational heating, g is the gravitational acceleration, cp is

the specific heat capacity, T0 is the reference temperature, and Res is the residual terms including subgrid-scale dissipation. The first

term on the right hand side is the conversion from mean available potential energy to EPE (referred to as “baroclinic term”), the second

is the generation of EPE by condensational heating (referred to as “diabatic term”), and the third is the conversion from eddy kinetic

energy to EPE. The present paper compares the baroclinic and diabatic terms. The energy and its budget is calculated using the filtered

data.

3. Results

3.1. Development and types of cyclones

The four representative experiments (Ta, Sa, Ea and Ia) result in different development and structure of cyclones. Figure 4a shows the

time evolution of Pmin for the representative experiments. Cyclones develop for Ta, Sa, and Ea (Pmin ≤ −10 hPa), whereas a cyclone

hardly develop for Ia (Pmin > −10 hPa). The most intense cyclone develops for Ta. While the cyclone in Sa starts to develop after 600

h, it develops more intense than that for Ea. The development stages defined in subsection 2.5 for Ta, Sa, and Ea are 80-380 h, 400-890

h, and 160-590 h, respectively.

Figure 5 shows cloud patterns of cyclones at their development stages. The cyclone in Ta has an axisymmetric structure with a

cloud-free eye at its centre at 350 h (Fig. 5a), which resembles a TC observed in satellite imagery. The cyclone in Ea has a larger

structure with a cloudy area on the east and north sides at 530 h (Fig. 5d), which is similar to an EC. The cyclone in Sa has a cloudy

area on the east side at early development stage (650 h; Fig. 5b). At later development stage (830 h; Fig. 5c), the cyclone has intense

convection near the centre in addition to the cloudy area on the east and north sides. This characteristic resembles some SCs in satellite
This article is protected by copyright. All rights reserved.
Accepted Article (a) Ta (350hr) (b) Sa (650hr) (c) Sa (830hr)

(d) Ea (530hr) (e) Sa-RDM (600hr) (f) Sa-RDM (850hr)

Figure 5. Cloud pattern (shading) and SLP (contours; interval is 5 hPa) in the representative experiments (Ta, Sa, and Ea) and the sensitivity experiment (Sa-RDM). (a)
Ta at 350 h, (b) Sa at 650 h, (c) Sa at 830 h, (d) Ea at 530 h, (e) Sa-RDM at 600 h, and (f) Sa-RDM at 850 h The cloud pattern is represented by vertically integrated
condensed water (the sum of cloud water, cloud ice, rain, snow and graupel). Centres of cyclones are located at the zonal centres of the panels, which show 3000 km ×
3000 km domains.

imagery, though the cloud patterns are different depending on cases (e.g. Fig. 9 in Evans and Guishard (2009); Figs. 1, 3 in Guishard

et al. (2009)).

It is necessary to confirm whether the results are sensitive to initial disturbance or not. Because Yanase and Niino (2015) have

already confirmed that the maximum intensity of cyclones has little sensitivity to initial disturbances in the tropical and extratropical

experiments, we will examine the sensitivity in the subtropical experiment. The sensitivity experiments are conducted for five sets of

radius of maximum wind (rmax ): 100, 150, 200, 250 and 300 km (referred to as Sa-R100, Sa-R150, Sa-R200, Sa-R250 and Sa-R300,

respectively). Note that the control experiment Sa is equivalent to Sa-R250. In addition, we also conduct an experiment in which a

random noise of potential temperature with the amplitude of 0.5 K is given as an initial disturbance (referred to as Sa-RDM). Figure

6a shows the time evolution of Pmin for the sensitivity experiments. All the experiments show cyclone development in the subtropical

environment. However, the timing of the cyclone development varies among the experiments, and shows no systematic relationship

to the size of the initial disturbances. It is notable that a cyclone can develop even from a random noise of potential temperature;

a similar characteristic is also found in the corresponding experiment for Ea (not shown), whereas cyclones hardly develop in the

corresponding experiment for Ta. Figures 5e and 5f show cloud patterns of the cyclone in Sa-RDM. At an early development stage

(600 h), the cyclone has small organized convection near the centre, which is different from Sa (Fig. 5b). The convection is organized

within a positive anomaly of equivalent potential temperature that shows a wavy pattern with a wavelength of 1500 ∼ 2000 km in the

lower troposphere (not shown). At later stage (850 h), however, the cyclone becomes larger, and has cloudy area on the east side of

the organized convection near the centre, which is similar to Sa (Fig. 5c). Thus, the subtropical experiment is more sensitive to initial

condition than the tropical and extratropical environments. Hereafter, the present study focuses on the result of the control experiment

(Sa) for which rmax of the initial vortex is 250 km, because it shows characteristics that are comparable with previous observational

studies.
This article is protected by copyright. All rights reserved.
(a) Control
0

-10

-20

-30

∆Prs (hPa)
-40

-50
Sa-R100
-60 Sa-R150
Sa-R200
-70 Sa-R250
Accepted Article
Sa-R300
-80 Sa-RDM
0 200 400 600 800 1000 1200
Time (hr)
(b) No-shear
0

-10

-20

-30
∆Prs (hPa)

-40

-50
Su0-R100
-60 Su0-R150
Su0-R200
-70 Su0-R250
Su0-R300
-80 Su0-RDM
0 200 400 600 800 1000 1200
Time (hr)

Figure 6. Time evolution of Pmin in the experiments with different initial disturbances: (a) Sa and (b) Su0 .

200
Ta
Sa
Ea
100

-100

-200
0 200 400 600 800 1000 1200
Time (hr)

Figure 7. Time evolution of the lower and upper thermal wind indices in the CPS classification (solid and dashed curves, respectively). The red, green and blue curves are
Ta, Sa and Ea, respectively. Thin and thick curves are raw values and 100-hour running means, respectively.

Figure 7 shows the time evolution of the thermal wind indices of the CPS classification. The cyclone in Ta has a deep warm-core

structure (positive lower and upper indices) almost throughout the life-cycle, which is a typical characteristic of a TC. The cyclone

in Ea has a deep cold-core structure (negative lower and upper indices) during most of the development stage before ∼ 550 h, but

subsequently has a shallow warm-core structure (positive lower and negative upper indices) during the mature stage. This evolution of

the CPS indices resembles a life cycle of an EC that undergoes a warm seclusion (e.g. Fig. 8 in Hart (2003)). While the lower index in

Sa fluctuates between positive and negative values, its average is positive during the development stage. Together with a negative upper

index, the cyclone with a shallow warm-core structure in Sa can be classified as an SC, which is consistent with the definition in Evans

and Guishard (2009).

To confirm whether the results of the representative experiments are robust or not, Fig. 8 summarizes the maximum intensity

and cyclone types averaged in the development stage for all the experiments. The experiments well reproduce the climatological

zonal contrast of cyclone development around the subtropics: the strong development in the idealized experiments in the western

This article is protected by copyright. All rights reserved.


Accepted Article

Figure 8. Maximum intensity and cyclone types for all the experiments. The maximum intensity is classified into strong development (Pmin ≤ −20 hPa; double circles),
weak development (-20 hPa < Pmin ≤ −10hPa; open single circles), and no development (Pmin > −10hPa; closed circles). The cyclone types are classified into TCs
(red circles), ECs (blue circles) and SCs (green circles) based on the CPS indices averaged in the development stage. Crosses are the same as in Fig. 1. Contours are SST
of 25, 26 and 27◦ C, whereas grey shading indicate vertical shear between 200 and 850 hPa exceeding 15 ms−1 .

part corresponds to frequent cyclone development in the climatology; and weak or no development in the former in the eastern

part corresponds to inactive development in the latter. The experiments also show the diversity in cyclone types in the western part,

particularly SCs in the subtropics. Note that all the experiments are conducted using only an initial vortex with the radius of maximum

wind of 250 km, because the computational resource is limited. Here, the sensitivity of the results to initial condition is examined only

for the representative experiments; all the experiments using the initial vortex with rmax of 100, 150, 200, 250 and 300 km consistently

show deep warm-core, shallow warm-core, and deep cold-core structures, for Ta, Sa, and Ea, respectively. While the development and

types of the cyclones are somewhat sensitive to the initial disturbance (Fig. 6) and the definition of the development stage (Fig. 7), Fig.

8 confirms the robust contrast in cyclone activity and types between different regions.

3.2. Structures

The previous subsection has shown that the cloud pattern and the CPS classification for Ta, Sa, and Ea are consistent with those of a

TC, SC, and EC (Figs. 5 and 7). The present subsection analyses the structures of the cyclones at their development stages in more

detail.

The equivalent potential temperature and horizontal winds at the altitude of ∼1.5 km near the three types of representative cyclones

are shown in Fig. 9. The equivalent potential temperature for Ta is high around the cyclone centre (Fig. 9a), which is similar to the

structure of a typical TC. On the other hand, the equivalent potential temperature in Ea is high on the east side of the cyclone, which

is advected by the southerly wind and form the cloudy area as seen in Fig. 5d. These structures are qualitatively similar to ECs. The

structure in Sa looks intermediate between those for Ta and Ea: the high equivalent potential temperature occurs around the centre and

on the east side of the cyclone, which is consistent with the cloud pattern as shown in Fig. 5c.

Vertical structures are also different between the experiments. Figure 10 shows pressure anomaly and potential temperature anomaly

on zonal-vertical cross sections through the cyclone centres, where the anomaly field is defined as deviations from the zonal mean. The

cyclone for Ta has an upright trough axis around the cyclone centre, and warm anomaly particularly in the upper troposphere (Fig. 10a),

which are typical characteristics of observed TCs. In contrast, the cyclone for Ea shows westward tilt of the trough axis and eastward

tilt of warm anomaly with increasing altitude (Fig. 10c), which is consistent with the structure of observed ECs or theoretical baroclinic

waves. The cyclone for Sa shows intermediate characteristics between those for Ta and Ea (Fig. 10b): the trough axis is upright on a

small scale near the cyclone centre as in Ta, but is tilted westward with increasing altitude on a large scale as in Ea. Thus, the cyclone

for Sa has a multi-scale structure: a TC-like structure on the small scale and an EC-like structure on the larger scale.

This article is protected by copyright. All rights reserved.


Accepted Article (a) Ta (350hr)

(b) Sa (830hr)

(c) Ea (530hr)

Figure 9. Equivalent potential temperature (shading; K) and horizontal wind (arrows) at the altitude of ∼1.5 km, and negative SLP anomaly (contours; interval is 10 hPa).
(a) Ta at 350 h, (b) Sa at 830 h and (c) Ea at 530 h.

The multi-scale structure in Sa is also found in the potential vorticity (PV) field. Figure 11 shows PV on the zonal-vertical cross

section through the cyclone centre. A region with large PV exists on a small scale near the cyclone centre in the middle troposphere,

but on a large scale it exists on the west side of the cyclone in the upper troposphere. Such a characteristic is comparable with the

composite of SCs observed over the North Atlantic Ocean (Fig. 6 in Evans and Guishard (2009))

3.3. EPE budget

The source of EPE is a useful metric for understanding the development mechanism of cyclones. Figure 12 shows the time evolution of

baroclinic and diabatic terms in the EPE budget averaged over the entire horizontal domain and for the altitudes below 20 km for Ta, Sa

and Ea. The major energy source for Ta is diabatic term, which indicates that the cyclone develop through the condensational heating

as a typical TC does. On the other hand, the energy sources for Ea are both baroclinic term and diabatic term, for which the former is

an essential process for a baroclinic instability. The ratio of the diabatic term to the baroclinic term for Sa is intermediate between Ta

and Ea: the diabatic term is dominant, whereas the baroclinic term also works.

Figure 13 shows the meridional-vertical distribution of the baroclinic and diabatic terms at development stages of the cyclones. The

diabatic term for Ta is deep and concentrated around the cyclone centre. For Ea, the baroclinic term is largest in the lower troposphere
This article is protected by copyright. All rights reserved.
Accepted Article (a) Ta (350hr)

(b) Sa (830hr)

(c) Ea (530hr)

Figure 10. Zonal-vertical sections through the cyclone centre for potential temperature anomaly (shading; K) and negative pressure anomaly (contours; interval is 5 hPa).
(a) Ta at 350 h, (b) Sa at 830 h and (c) Ea at 530 h.

Figure 11. Zonal-vertical section through the cyclone centre for PV (shading; PVU) and potential temperature (contours; interval is 3K) for Sa.

and is distributed on a large scale in the meridional direction, whereas the diabatic term occurs on a smaller scale. The distribution of the

EPE terms for Sa is again intermediate between those for Ta and Ea: the baroclinic term is widely distributed in the lower troposphere

as for Ea, and deep diabatic term is concentrated near the cyclone centre as for Ta, although shallow diabatic term is widely distributed
This article is protected by copyright. All rights reserved.
Ta
20 Sa
Ea

Pe budget (J m-3 day-1)


15

10

5
Accepted Article
0

0 100 200 300 400 500 600 700 800 900 1000 1100 1200
Time (hr)

Figure 12. Time evolution of diabatic term (solid curves) and baroclinic term (dashed curves) in the EPE budget for Ta (red), Sa (green) and Ea (blue).

(a) Ta (350hr)

(b) Sa (830hr)

(c) Ea (530hr)

Figure 13. Meridional-vertical distribution of EPE (shading; Jm−3 ), diabatic term (orange contour) and baroclinic term (purple contour) in the EPE budget for (a) Ta at
350 h, (b) Sa at 830 h, and (c) Ea at 530 h (contour intervals are 20, 5 and 20 Jm−3 day−1 , respectively).

in the lower troposphere. Thus, Sa shows a multi-scale distribution in the EPE budget: TC-like characteristics on a small scale and

EC-like characteristics on a large scale.

To confirm the EPE budget for all the experiments, the ratio of the baroclinic term to the diabatic term is examined during the

development stage (Fig. 14). At the latitudes lower than 30◦ N, the diabatic term is the dominant source of the EPE. The ratio becomes

larger with increasing latitude, and exceed unity at the latitudes higher than 40◦ N. Thus, the ratio of the baroclinic term to the diabatic

term also discriminates the cyclone types.


This article is protected by copyright. All rights reserved.
Accepted Article

Figure 14. As in Fig. 8, but for the ratio of baroclinic term to diabatic term in the EPE budget (colours of circles) for all the experiments.

Figure 15. As in Fig. 8, but for the no-shear experiments.

3.4. Sensitivity to vertical shear

The vertical shear of the environmental fields (Fig. 2) can have positive or negative influence on the cyclone development depending

on the cyclone types. To examine the influence of the vertical shear, we conduct “no-shear” experiments, in which the vertical shear

is removed from the control experiments. Because the zonal wind at the sea level is fixed to zero, the “no-shear” experiments has no

zonal wind. In addition, the thermal wind balance gives no meridional gradient of potential temperature, i.e. no baroclinicity.

Figure 4b shows the time evolution of Pmin in the representative no-shear experiments (see also Fig. 4a for the control experiments).

The no-shear experiments for Ta, Sa, Ea, and Ia are referred to as Tu0 , Su0 , Eu0 , and Iu0 , respectively, where the subscript u0 indicates

that the zonal wind is set to zero. Not surprisingly, a cyclone cannot develop in Eu0 , because the vertical shear or baroclinicity is

indispensable for the development of ECs. As would also be expected, a cyclone in Tu0 develop faster than that in Ta, because the

vertical shear is unfavourable for the development of a TC. However, the maximum intensity in Tu0 (-63.9 hPa) is little different from

that in Ta (-65.0 hPa). More drastic influence of the vertical shear is found in Iu0 , where a cyclone develops in contrast to Ia. Because

the cyclone in Iu0 shows an axisymmetric cloud pattern with a cloud-free eye at the centre (not shown), the cyclone resembles a

TC. A cyclone in Su0 also develops faster than that in Sa, although the difference in the maximum intensity is little between the two

experiments (-34.6 hPa for Sa and -34.4 hPa for Su0 ). The cloud pattern of the cyclone in Su0 becomes similar to a TC rather than an

SC (not shown).

As the cyclone development in Sa is sensitive to the initial disturbance (Fig. 6a), the sensitivity to the initial disturbance is also

examined for Su0 . Figure 6b shows the result of the same set of sensitivity experiments as in Fig. 6a except for Su0 . Most of the
This article is protected by copyright. All rights reserved.
cyclones in the no-shear experiments develop faster than those in the control experiments, and becomes less sensitive to the initial

disturbance. Exceptionally, the random noise of potential temperature in Su0 -RDM trigger no cyclones (no curve for Su0 -RDM in

Fig. 6b). This result is similar to the tropical experiment (Ta-RDM; not shown), where a TC cannot develop from the random noise.

Therefore, the vertical shear in Sa supports initial organization of cumulus convection into the cyclones, while it delays the development

of the cyclones.

The results of all the no-shear experiments are summarized in Fig. 15 (see also Fig. 8 for the control experiments), where only TC

dynamics dominates. Apparently, ECs in the extratropics hardly develop. At the latitudes between 15◦ N ∼ 25◦ N in the eastern part,
Accepted Article
several cyclones become active and strong as TCs. This implies that the cyclones in the eastern North Atlantic are suppressed by the

strong vertical shear in the control experiments (grey shading in Fig. 8, 15). The cyclones are still more active in the western part than

in the eastern part, because of the SST contrast. Most of the SCs in the control experiment become TCs in the no-shear experiments. A

few exceptions are the SCs at 35◦ N, where the vertical shear seems to be more important for the development of SCs, although more

experiments are necessary to draw a robust conclusion.

4. Discussion

4.1. Advantages and disadvantages of the idealized experiments

The first purpose of the present idealized experiments is to examine to what extent the climatological environment explains the

geographical distribution of the development for various types of cyclones over the North Atlantic Ocean. The experiments have

demonstrated that the climatological environment can reproduce the geographical distribution of cyclones qualitatively as follows: (1)

the zonal contrast in the cyclone activity around the subtropics, and (2) the diverse cyclones including TCs, SCs, and ECs at different

latitudes in the western North Atlantic Ocean.

On the other hand, the cyclones in the experiments grow more slowly than those in the real atmosphere, presumably due to

the experimental design and several missing processes. The experiment is designed not only to deal with various environments

from the tropics to the extratropics, but also to suppress any disturbances near the meridional boundaries. Therefore, we employ a

smoothed environmental field (the regional average in 20◦ longitudes and latitudes) and the boundary zones with the damping and the

modified environment. While these designs successfully suppress disturbances in the boundary zones, they partly weaken the cyclone

development in the central zone. More sophisticated design may improve the growth of cyclones.

In addition, the slow growth of cyclones may also result from several missing processes in the idealized experiments. Firstly, the

seasonal mean fields of the 30-year climatology filters out some temporal variation in the environmental fields. Therefore, cyclones

in the real atmosphere may develop in a more favourable environment than in the climatological fields: for example, TCs in the real

atmosphere may grow faster within temporarily reduced vertical shear as in the no-shear experiment (Fig. 4b). There are a wide range

of temporal modulations in the real atmosphere. In the inter-annual time scale, ENSO modulates the distribution of TCs (Camargo et al.

2007). In the intra-seasonal time scale, the Madden Julian Oscillation and the boreal summer intra-seasonal oscillation can modulate

the activity of TCs (Camargo et al. 2009; Yanase et al. 2012). Furthermore, the experiments exclude the influence of transient external

forcing including upper tropospheric troughs (Petterssen and Smebye 1971; Deveson and Browning 2002; Bentley et al. 2017; Fischer

et al. 2017) and tropical waves (Ritchie and Holland 1999; Frank and Roundy 2006). Secondly, the present idealized experiments do not

consider the influence of several environmental factors. While the vertical motion may affect the development of cyclones, it is difficult

to maintain vertical motion in idealized experiments because of mass conservation and vertical advection of potential temperature and

water vapour. The horizontal shear is another environmental factor that may affect the dynamics of cyclones (Wernli et al. 1998; Davis

2010).
This article is protected by copyright. All rights reserved.
The present study has no answer to the question about how much the aforementioned missing processes affect the geographical

distribution of cyclones. The relative importance of these processes should be assessed quantitatively in a future study. Nevertheless,

it is noteworthy that the climatological fields with only five environmental factors have potential for reproducing the geographical

distribution of development and types of cyclones even if the other processes are not included. This result supports the previous studies

that attribute the climatology of cyclones to environmental fields (Camargo et al. 2007; Roberts et al. 2015).

4.2. Subtropical cyclones


Accepted Article
Our idealized experiments simulate SCs in the subtropics over the western North Atlantic Ocean, which are defined as cyclones with

shallow warm-core structures in the present study. A number of previous studies have reported the formation of SCs in this region,

where the environment of high SST is overlapped with that of large baroclinicity (Guishard et al. 2009). In the real atmosphere,

SCs often form under the influence of an equatorward extension of cold troughs or PV anomalies in the upper troposphere (Evans and

Guishard 2009; Bentley et al. 2016); Bentley et al. (2017) classified them into cutoff lows, meridional troughs, and zonal troughs, where

∼61 % of the analyzed SCs are associated with anticyclonic wave breaking. On the other hand, no upper tropospheric disturbances are

given in the initial field in our idealized experiments. Therefore, the second purpose of the present study is to compare the SCs in our

highly simplified experiments with those in previous observational and numerical studies.

In the Sa experiment, the SC has intense organized convection near the cyclone centre and cloudy area on the east side (Fig. 5c).

The former structure resembles a TC, whereas the latter an EC. Such characteristics resemble observed cloud patterns of some SCs and

tropical transitions (Hebert and Poteat 1975; Evans and Guishard 2009; Guishard et al. 2009; Mauk and Hobgood 2012; Pinto et al.

2013), although the patterns vary from case to case. The cloudy area on the east side of the cyclone occurs where the high equivalent

potential temperature is advected by the southerly wind in the lower troposphere (Fig. 9b). This is consistent with the composite

analysis of SCs over the South Atlantic Ocean in Gozzo et al. (2014), which showed the poleward advection of warm and moist air and

consequent upward motion on the east side of SCs.

The SC is accompanied by a narrow positive PV anomaly with a warm core in the middle troposphere at the cyclone centre and

by a wide positive PV anomaly with a cold core on the west side of the cyclone centre (Fig. 11). These characteristics resemble the

composite PV fields of SCs over the North Atlantic Ocean (Evans and Guishard 2009) and over the South Atlantic Ocean (Gozzo

et al. 2014). It should be noted that the upper tropospheric PV anomaly in the Sa experiment forms spontaneously through baroclinic

process, while that in observations often originates from the higher latitudes. Based on idealized experiments and quasi-geostrophic

omega-equation, Davis (2010) showed that a positive PV anomaly in the upper troposphere dynamically forces updrafts on its east

side, which results in deep convection and a diabatically-produced PV anomaly in the middle troposphere.

The EPE budget in the subtropical experiments (Fig. 14) shows that both diabatic and baroclinic processes produce EPE, where the

former is larger than the latter. In contrast, the EPE budget in the extratropical experiments shows baroclinic term larger than diabatic

term. Pinto et al. (2013) showed a similar energy cycle for an SC Anita in 2010 over the South Atlantic Ocean: both baroclinic and

diabatic processes produce EPE during a shallow warm-core stage, whereas the baroclinic process becomes larger than the diabatic

process during a deep cold-core stage.

In the idealized experiments in Davis (2010), SCs are associated with condensational heating, moderate and deep westerly shear, and

near absence of lower tropospheric baroclinicity. The SC in our Sa experiment is also characterized by large condensational heating

and moderate and deep westerly shear (13.0 ms−1 between 850 hPa and 200 hPa). Davis (2010) proposed the PV1/PV2 index, which

is the ratio of the near-surface baroclinicity to lower-to-middle-tropospheric PV. The SC in our experiment has smaller PV1/PV2 (∼

1; not shown) than the EC for Ea does (PV1/PV2 is 2 ∼ 8), which is in agreement with Davis (2010). However, the SCs in Davis

(2010) have PV1/PV2 much smaller than unity, which is attributed to the absence of near-surface baroclinicity in their environmental

This article is protected by copyright. All rights reserved.


field. This results from the difference in the experimental design: the environmental field in Davis (2010) is formed by convective

heating during the early time of the simulation, whereas that in our experiments is obtained from the climatological fields, which is

maintained by the spectral nudging. The lower-tropospheric baroclinicity in our experiment causes wavy modulation confined to the

lower troposphere (not shown), and occasionally organizes cumulus convection into a cyclone in some subtropical experiments as seen

in the early stage in Sa-RDM (Fig. 5e). Thus, the cyclones in the subtropical experiments seem to be organized not only by upper-

tropospheric baroclinic processes such as Rossby wave on the tropopause (Davis 2010), but also by lower-tropospheric processes such

as diabatic Rossby vortex (Moore and Montgomery 2005; Conzemius et al. 2007). Common and different characteristics between SCs
Accepted Article
should be discussed based on further idealized experiments, and should also be compared with disturbances observed in the subtropical

environments.

4.3. Relationship between cyclones and environment

The idealized experiments reproduce the different types of cyclones in different regions: TCs, SCs, ECs and inactive cyclones in

the tropical, subtropical, extratropical and eastern North Atlantic Ocean, respectively. This confirms that the environmental fields

control the development and types of cyclones. Figures 16a and 16b show the climatological distribution of potential intensity (PI) and

baroclinicity (BC) in autumn. The PI index assesses the environment for TCs based on an estimate for the maximum intensity of a TC

in a steady state (Bister and Emanuel 2002). While PI is largely dependent on SST, it also considers the vertical profiles of atmospheric

temperature and humidity to incorporate the stability of the troposphere. High PI occurs around the tropical ocean (Fig. 16a), which is

in agreement with the active TC region. The BC index assesses the environment for ECs based on the Eady growth rate of baroclinic

instability, which depends on the Coriolis parameter, vertical shear (or horizontal gradient of temperature) and static stability. (Lindzen

and Farrell 1980; Hoskins and Valdes 1990). High BC occurs in the mid-latitudes (Fig. 16b), which explains the active EC region.

To assess the direct relationship between cyclone development and their environmental fields, the environments for climatologically

analysed and numerically reproduced cyclones are plotted on the diagram of environmental parameter space (Yanase et al. 2014), in

which PI and BC are taken as the horizontal and vertical coordinates, respectively (Fig. 16c). In this diagram, the development of

cyclones in the idealized experiments (single open circles and double circles) are distributed from the lower right (high PI and low

BC) to the upper left (low PI and high BC). The lower right and upper left parts are characterized by TCs (red circles) and ECs (blue

circles), respectively, which is consistent with the climatological cyclone development obtained from the JRA-25 reanalysis (red and

blue crosses). In the intermediate part (moderate PI and moderate BC), SCs develop in several experiments (green circles), which

is also consistent with the climatological study (green crosses). Previous studies also suggested that SCs occur in the intermediate

environment between the tropics and extratropics (Guishard et al. 2009). However, cyclones fail to develop in several experiments in

the intermediate part, when PI and BC are relatively low, which is similar to the idealized experiments on cyclones over the South

Indian Ocean in summer (Yanase and Niino 2015).

While PI and BC are useful indices for a quick look at the environment for TCs and ECs, the role of BC should be carefully discussed.

As described above, BC depends on the Coriolis parameter and vertical shear. The sensitivity experiments have demonstrated that the

vertical shear has different influences on cyclones depending on regions. The vertical shear is indispensable for the development of ECs

in the extratropics. On the other hand, the vertical shear suppresses the development of TCs around the tropics and subtropics, which is

remarkable in the eastern North Atlantic at the latitudes between 15◦ N and 25◦ N (Figures 8 and 15). Note that TC development (Fig.

1) and genesis (Fig. 7a in McTaggart-Cowan et al. (2013)) along the longitude of ∼ 40◦ W are more active around 30◦ N than around

20◦ N. This characteristic cannot be explained by SST alone which monotonically decreases with increasing latitude. The sensitivity

experiments have demonstrated that the vertical shear is one of the factors that can explain this meridional distribution of TCs, although

we cannot discard the possible influences of other environmental factors (humidity and vertical motion), external forcing, and the tracks

This article is protected by copyright. All rights reserved.


Accepted Article (a) Potential Intensity

(b) Baroclinicity

(c) PI-BC space

Figure 16. Environmental fields of cyclones during the autumn season. (a) Geographical distribution of climatological potential intensity. (b) Geographical distribution
of climatological baroclinicity. (c) Cyclone development in the idealized experiments (circles) and climatological analysis (crosses) plotted on the parameter space of
potential intensity (horizontal axis) and baroclinicity (vertical axis). As in Fig. 8, the intensity of cyclones are classified into three types for the idealized experiments:
strong development (Pmin ≤ −20hPa; double circles), weak development (-20 hPa < Pmin ≤ −10hPa; open single circles), and no development (Pmin > −10hPa;
closed circles). The cyclones are classified into TCs (red circles), ECs (blue circles) and SCs (green circles).

of TCs. In the subtropics, the vertical shear has more complicated influence on SCs (Fig. 6). We are analysing the role of vertical shear

in the intermediate environment between the tropics and extratropics in more detail, and will report it in a future paper.

5. Summary

To understand the influence of environmental fields on a wide variety of cyclones over the North Atlantic Ocean, we have examined

the development of cyclones within the climatological environment by a series of idealized numerical experiments. The experiments

reproduce the contrast between the active cyclones in the western part and the inactive cyclones in the eastern part. The experiments

also reproduce the various cyclones in the western part: not only TCs (deep warm-core cyclones) in the tropics and ECs (deep cold-core

cyclones) in the extratropics, but also SCs (shallow warm-core cyclones) in the subtropics. Thus, the development and types of cyclones

are explained by the environmental fields qualitatively, whereas the growth rate of cyclones should be affected by additional processes

This article is protected by copyright. All rights reserved.


such as upper tropospheric troughs. While the vertical shear delays the development of the SCs, it supports the initial organization of

cumulus convection into the cyclones. The vertical shear suppresses the development of TCs particularly in the eastern North Atlantic

Ocean. Although the tropical and extratropical experiments themselves merely confirm the state of knowledge about TCs and ECs, the

series of experiments provide a useful intercomparison of the wide spectrum of cyclone types in the same framework.

The present study is the first attempt at idealized experiments on SCs using the climatological environmental fields. Several

characteristics of the simulated SCs resemble those in the real atmosphere. The cloud pattern shows small-scale organized convection

near the cyclone centre like a TC and a cloud-covered area on the north and east sides of the centre like an EC. The PV structure shows
Accepted Article
a small-scale positive anomaly in the middle troposphere near the cyclone centre together with a large-scale positive anomaly in the

upper troposphere on the west side of the centre. The EPE is produced from both diabatic and baroclinic processes. However, SCs in

the real atmosphere often form under the influence of an equatorward extension of cold troughs in the upper troposphere, which is not

prescribed in the initial fields of the present experiments. While the influence of the upper tropospheric troughs are different between

the real atmosphere and the idealized experiments, it is noteworthy that the subtropical environment has a potential for developing the

cyclones that have several characteristics similar to those of observed SCs as mentioned above. Although most of the characteristics of

the simulated SCs are intermediate between those of TCs and ECs, as might be expected, it is notable that the SC shows multi-scale

structures with small TC-like structures embedded in large EC-like structures: the cloud pattern (Fig. 5), the vertical structure of the

trough axis (Fig. 10), and the EPE budget (Fig. 13). As characteristics of SCs vary among the experiments in the present study (Figs.

5 and 6), and also among those in Davis (2010), additional idealized experiments from various perspectives will promote a further

understanding of basic dynamics of SCs.

In the real atmosphere, the dynamics of individual cyclones would be controlled by various processes such as the uniform

environmental fields considered in the present study, and other factors (e.g., environmental vertical motion, temporal and spatial

variations of the environmental fields, and the influence of external forcing). Since these processes work simultaneously and intricately,

it is generally difficult to elucidate their individual influence on cyclone dynamics in case studies. In this context, the present idealized

experiments have clarified to what extent the simplified climatological environmental fields can explain the geographical distribution

of cyclones. The present findings will be useful for understanding the environments for a spectrum of cyclones including TCs, SCs

(hybrid cyclones), and ECs comprehensively within various climate in the past, present, and future. Further idealized experiments are

also necessary for understanding more realistic processes such as the influence of upper tropospheric troughs, the environment with

horizontal shear and convergence, and time-dependent environment for transitioning cyclones.

Acknowledgement

We acknowledge helpful comments by Keita Iga, Eigo Tochimoto, Akiyoshi Wada, Masato Sugi, Masaki Satoh, Kensuke Nakajima,

Yoshiyuki Hayashi, and two anonymous reviewers. The atmosphere and ocean datasets made use of JRA-25/JCDAS and NOAA OI-

SST. Potential intensity was calculated using the FORTRAN program from the website of Dr. Kerry A. Emanuel. This research was

mainly conducted using the HITACHI SR16000 System (yayoi) in the Information Technology Center, The University of Tokyo. This

research was partly supported by JSPS KAKENHI Grant Numbers 23740349 and 24244074, and FLAGSHIP 2020, MEXT within

priority study 4 (Advancement of Meteorological and Global Environmental Predictions Utilizing Observational “Big Data”).

References

American Meteorological Society. 2017. Subtropical cyclone. Glossary of Meteorology. [Available online at http://glossary.ametsoc.org/wiki/
Subtropical_cyclone].

Beljaars A. 1995. The parametrization of surface fluxes in largescale models under free convection. Quart. J. Roy. Meteor. Soc. 121: 255–270.
This article is protected by copyright. All rights reserved.
Bentley AM, Bosart LF, Keyser D. 2017. Upper-tropospheric precursors to the formation of subtropical cyclones that undergo tropical transition in the North
Atlantic Basin. Mon. Wea. Rev. 145: 503–520.

Bentley AM, Keyser D, Bosart LF. 2016. A Dynamically Based Climatology of Subtropical Cyclones that Undergo Tropical Transition in the North Atlantic
Basin. Mon. Wea. Rev. 144: 2049–2068.

Bister M, Emanuel KA. 2002. Low frequency variability of tropical cyclone potential intensity. 1. interannual to interdecadal variability. J. Geophys. Res. 107,
doi:10.1029/2001JD000776.

Camargo SJ, Emanuel KA, Sobel AH. 2007. Use of a genesis potential index to diagnose ENSO effects on tropical cyclone genesis. J. Climate 20: 4819–4834.

Camargo SJ, Wheeler MC, Sobel AH. 2009. Diagnosis of the MJO modulation of tropical cyclogenesis using an empirical index. J. Atmos. Sci. 66: 3061–3074,
Accepted Article
doi:10.1175/2009JAS3101.1.

Catto JL. 2016. Extratropical cyclone classification and its use in climate studies. Rev. Geophys. 54: 486–520.

Charney JG. 1949. The dynamics of long waves in a baroclinic westerly current. J. Meteor. 4: 135–162.

Charney JG, Eliassen A. 1964. On the growth of the hurricane depression. J. Atmos. Sci. 21: 68–75.

Conzemius RJ, Moore RW, Montgomery MT. 2007. Mesoscale convective vortex formation in a weakly sheared moist neutral environment. J. Atmos. Sci. 64:
1443–1466.

Coronel B, Ricard D, Rivière G, Arbogast P. 2016. Coldconveyorbelt jet, sting jet and slantwise circulations in idealized simulations of extratropical cyclones.
Quart. J. Roy. Meteor. Soc. 142: 1781–1796.

Davis C, Bosart L. 2003. Baroclinically induced tropical cyclogenesis. Mon. Wea. Rev. 131: 2730–2747.

Davis CA. 2010. Simulations of subtropical cyclones in a baroclinic channel model. J. Atmos. Sci. 67: 2871–2892.

Davis CA, Bosart LF. 2004. The TT problem: Forecasting the tropical transition of cyclones. Bull. Amer. Meteor. Soc. 85: 1657–1662.

Deveson ACL, Browning KA. 2002. A classification of FASTEX cyclones using a heightattributable quasigeostrophic verticalmotion diagnostic. Quart. J. Roy.
Meteor. Soc. 128: 93–117.

Eady ET. 1949. Long waves and cyclone waves. Tellus 1: 33–52.

Emanuel KA. 1986. An air-sea interaction theory for tropical cyclones. Part I: steady-state maintenance. J. Atmos. Sci. 43: 585–604.

Emanuel KA, Rotunno R. 1989. Polar lows as arctic hurricanes. Tellus 41A: 1–17.

Evans C, Wood KM, Aberson SD, Archambault HM, Milrad SM, Bosart LF, Corbosiero KL, Davis CA, Pinto JRD, Doyle J, Fogarty C, Galarneau Jr TG, Grams
CM, Griffin KS, Gyakum J, Hart RE, Naoko K, Lentink HS, McTaggart-Cowan R, Perrie W, Quinting JFD, Reynolds CA, Riemer M, Ritchie EA, Sun Y,
Zhang F. 2017. The extratropical transition of tropical cyclones. Part I: Cyclone evolution and direct impacts. Mon. Wea. Rev. 145: 4317–4344.

Evans JL, Guishard MP. 2009. Atlantic subtropical storms. part I: diagnostic criteria and composite analysis. Mon. Wea. Rev. 137: 2065–2080.

Fischer MS, Tang BH, Corbosiero KL. 2017. Assessing the Influence of upper-tropospheric troughs on tropical cyclone intensification rates after genesis. Mon.
Wea. Rev. 145: 1295–1313.

Frank WM, Roundy PE. 2006. The role of tropical waves in tropical cyclogenesis. Mon. Wea. Rev. 134: 2397–2417.

Gozzo LF, da Rocha RP, Reboita MS, Sugahara S. 2014. Subtropical cyclones over the Southwestern South Atlantic: Climatological aspects and case study. J.
Climate 27: 8543–8562.

Guishard M, Evans J, Hart R. 2009. Atlantic subtropical storms. Part II: Climatology. J. Climate 22: 3574–3594.

Hart RE. 2003. A cyclone phase space derived from thermal wind and thermal asymmetry. Mon. Wea. Rev. 131: 585–616.

Hebert PH, Poteat KO. 1975. A satellite classification technique for subtropical cyclones. NOAA Tech. Memo. NWS SR-83 : 25 pp.

Hendricks EA, Montgomery MT, Davis CA. 2004. The role of “vortical” hot towers in the formation of tropical cyclone Diana (1984). J. Atmos. Sci. 61:
1209–1232.

Hoskins BJ. 1976. Baroclinic waves and frontogenesis Part I: Introduction and Eady waves. Quart. J. Roy. Meteor. Soc. 102: 103–122.

Hoskins BJ, Valdes PJ. 1990. On the existence of storm-tracks. J. Atmos. Sci. 47: 1854–1864.

Jones SC, Harr PA, Abraham J, Bosart LF, Bowyer BJ, Evans JL, Hanley DE, Hanstrum BN, Hart RE, Lalaurette F, Sinclair MR, Smith RK, Thorncroft C.
2003. The extratropical transition of tropical cyclones: Forecast challenges, current understanding, and future directions. Wea. Forecasting 18: 1052–1092.

Kain J, Fritsch J. 1990. A one-dimensional entraining/detraining plume model and its application in convective parameterization. J. Atmos. Sci. 47: 2784–2802.

Lin Y, Farley R, Orville H. 1983. Bulk parameterization of the snow field in a cloud model. J. Appl. Meteor. 22: 1065–1092.

Lindzen RS, Farrell B. 1980. A simple approximate result for the maximum growth rate of baroclinic instabilities. J. Atmos. Sci. 37: 1648–1654.

This article is protected by copyright. All rights reserved.


Mauk RG, Hobgood JS. 2012. Tropical cyclone formation in environments with cool SST and high wind shear over the northeastern Atlantic Ocean. Wea.
Forecasting 27: 1433–1448.

McTaggart-Cowan R, Galarneau Jr TJ, Bosart LF, Richard WM, Martius O. 2013. A global climatology of baroclinically influenced tropical cyclogenesis. Mon.
Wea. Rev. 141: 1963–1989.

Montgomery MT, Smith RK. 2014. Paradigms for tropical cyclone intensification. Aust. Meteor. Oceanogr. J. 64: 37–66.

Moore RW, Montgomery MT. 2005. Analysis of an idealized, three-dimensional diabatic Rossby vortex: A coherent structure of the moist baroclinic atmosphere.
J. Atmos. Sci. 62: 2703–2725.

Murakami M. 1990. Numerical modeling of dynamical and microphysical evolution of an isolated convective cloud - The 19 July 1981 CCOPE cloud. J. Meteor.
Accepted Article
Soc. Japan 68: 107–128.

Nakanishi M, Niino H. 2004. An improved MellorYamada level-3 model with condensation physics: Its design and verification. Boundary Layer Meteorol. 112:
1–31.

Onogi K, Tsutsui J, Koide H, Sakamoto M, Kobayashi S, Hatsushika H, Matsumoto T, Yamazaki N, Kamahori H, Takahashi K, Kadokura S, Wada K, Kato K,
Oyama R, Ose T, Mannoji N, Taira R. 2007. The JRA-25 reanalysis. J. Meteor. Soc. Japan 85: 369–432.

Ooyama K. 1969. Numerical simulation of the life cycle of tropical cyclones. J. Atmos. Sci. 26: 3–40.

Petterssen S, Smebye SJ. 1971. On the development of extratropical cyclones. Quart. J. Roy. Meteor. Soc. 97: 457–482.

Pinto JRD, Reboita MS, Rocha RP. 2013. Synoptic and dynamical analysis of subtropical cyclone Anita (2010) and its potential for tropical transition over the
South Atlantic Ocean. J. Geophys. Res. Atmos. 118: 10 870–10 883.

Reynolds RW, Rayner NA, Smith TM, Stokes DC, Wang W. 2002. An improved in situ and satellite SST analysis for climate. J. Climate 15: 1609–1625.

Ritchie EA, Holland GJ. 1999. Large-scale patterns associated with tropical cyclogenesis in the western Pacific. Mon. Wea. Rev. 127: 2027–2043.

Roberts MJ, Vidale PL, Mizielinski MS, Demory ME, Schiemann R, Strachan J, Hodges K, Bell R, Camp J. 2015. Tropical cyclones in the UPSCALE ensemble
of high-resolution global climate models. J. Climate 28: 574–596.

Saito K, Fujita T, Yamada Y. 2006. The operational JMA nonhydrostatic mesoscale model. Mon. Wea. Rev. 134: 1266–1298.

Wernli H, Fehlmann R, Lüthi D. 1998. The effect of barotropic shear on upper-level induced cyclogenesis: Semigeostrophic and primitive equation numerical
simulations. J. Atmos. Sci. 55: 2080–2094.

Yanase W, Niino H. 2015. Idealized Numerical Experiments on Cyclone Development in the Tropical, Subtropical, and Extratropical Environments. J. Atmos.
Sci. 72: 3699–3714.

Yanase W, Niino H, Hodges KI, Kitabatake N. 2014. Parameter spaces of environmental fields responsible for cyclone development from tropics to extratropics.
J. Climate 27: 652–671.

Yanase W, Satoh M, Taniguchi H, Fujinami H. 2012. Seasonal and intraseasonal modulation of tropical cyclogenesis environment over the Bay of Bengal during
the extended summer monsoon. J. Climate 25: 2914–2930.

This article is protected by copyright. All rights reserved.

You might also like