Anda di halaman 1dari 17

1894

JOURNAL OF CLIMATE

VOLUME 27

Objective Identification of the Intertropical Convergence Zone: Climatology


and Trends from the ERA-Interim
GARETH BERRY
Monash Weather and Climate, Monash University, Clayton, Victoria, Australia

MICHAEL J. REEDER
Monash Weather and Climate, and Centre of Excellence for Climate System Science, School of Mathematical Sciences,
Monash University, Clayton, Victoria, Australia
(Manuscript received 12 June 2013, in final form 7 October 2013)
ABSTRACT
An objective method for the identification of the intertropical convergence zone (ITCZ) in gridded numerical weather prediction datasets is presented. This technique uses layer- and time-averaged winds in the
lower troposphere to automatically detect the location of the ITCZ and is designed for use with datasets
including operational forecasts and climate model output. The method is used to create a climatology of ITCZ
properties from the Interim ECMWF Re-Analysis (ERA-Interim) dataset for the period 19792009 to serve
as an indicator of the techniques ability and a benchmark for future comparisons. The automatically generated objective climatology closely matches the results from subjective studies, showing a seasonal cycle in
which the oceanic ITCZ migrates meridionally and the land-based ITCZ features are predominantly summertime phenomena. Composites based on the phase of the El Ni~
noSouthern Oscillation index show a major
shift in the mean position and changes in intensity of the ITCZ in all ocean basins as the index varies. Under
La Ni~
na conditions, the ITCZ intensifies over the Maritime Continent and eastern Pacific, where the ITCZ
weakens over the central and equatorial eastern Pacific. An analysis of changes in the ITCZ and its divergence
during the period 19792009 indicates that the mean position of the ITCZ shifts southward in the western
Pacific and a broad global intensification of the convergence into ITCZ regions. The relationship between
tropical cyclogenesis and the ITCZ is also examined, finding that more than 50% of all tropical cyclones form
within 600 km of the ITCZ.

1. Introduction
The intertropical convergence zone (ITCZ) is one of the
most important components of the climate system. It is
associated with the seasonal migration of rainfall in the
tropics and the subtropical anticyclones. The ITCZ is
characterized by the lower-tropospheric convergence of the
northeasterly and southeasterly tropical trade winds and
a zonally elongated band of deep convection. On average,
air ascending in the deep convection moves poleward in the
upper troposphere and descends in the subtropics; the
ITCZ is essentially the ascending branch of the Hadley
circulation. The precise location of the ITCZ affects the

Corresponding author address: Gareth Berry, Monash Weather


and Climate, Monash University, Wellington Road, Clayton, VIC
3800, Australia.
E-mail: gareth.berry@monash.edu
DOI: 10.1175/JCLI-D-13-00339.1
2014 American Meteorological Society

weather conditions over a broad region as it changes the


position of convective rainfall at low latitudes and also the
suppression of rainfall within the subtropical anticyclones.
The seasonal-mean rainfall from the National Oceanic
and Atmospheric Administration (NOAA)/Climate Prediction Center (CPC) morphing technique (CMORPH;
Joyce et al. 2004) for the period 19982012 is shown in
Fig. 1. The time-mean ITCZ locations can be easily
inferred by the zonally elongated regions of high rainfall within approximately 158 of the equator in both
hemispheres. The day-to-day nature of the ITCZ can
be highly variable (Wang and Magnusdottir 2006) because of interactions with monsoon systems and synoptic
disturbances (including tropical cyclones). The aim of
the current study is to present an objective, automated
method for detecting the position and intensity of the
ITCZ and to use this method to describe the climatology
of these features and their recent trends in a global

1 MARCH 2014

BERRY AND REEDER

1895

FIG. 1. Seasonal averages of CMORPH precipitation (mm day21) for the period 19982012. (SON denotes the season
SeptemberNovember.)

reanalysis dataset. This work is intended to provide


a diagnostic tool for the analysis of climate models and
a reanalysis-based benchmark for comparison.
There is no agreed way to identify the location of the
ITCZ (Sadler 1975). Some previous work has used regions of low cloud-top temperature (e.g., Waliser and
Gautier 1993) or bands of high precipitation (e.g., Gu
et al. 2005), whereas other studies have used dynamical
fields, such as relative vorticity (e.g., Chan and Evans

2002), mean sea level pressure, or confluence. Each of


these methods has their advantages and disadvantages,
but all are able to locate the ITCZ. To the authors
knowledge, there is only one existing global climatology
of the ITCZ: that is the climatology by Waliser and
Gautier (1993), who used infrared satellite imagery over
a 17-yr period to describe the location of the ITCZ. This
study confirmed much of the common knowledge of how
the ITCZ location changes through the course of a year.

1896

JOURNAL OF CLIMATE

However, it can be difficult to use a method like Waliser


and Gautiers to evaluate the performance of weather or
climate models, as cloud-top temperatures may not be
a standard output field and considerable manual interpretation may be required. Other studies have used
rainfall to define large convergence zones in climate
models [e.g., Brown et al. (2011) in examining the South
Pacific convergence zone (SPCZ)] with crude linear
fitting to join rainfall maxima in seasonal averages. This
type of analysis is problematic, since a region of seasonally high precipitation is not exclusively associated
with convergence zones and a linear fit of rainfall locations may not be applicable at shorter time scales and at
all points on the globe. Additionally, the precise rainfall
forecast depends on both the synoptic environment and
the model formulation. Here, a more sophisticated
objective method of deriving the ITCZ position is presented; this method uses lower-tropospheric convergence, which can be computed from standard dynamic
fields from numerical weather prediction (NWP) or
climate model output. The technique presented is fully
automated and objective allowing large amounts of data
(e.g., those produced for climate model intercomparison
projects) to be easily processed and quantitatively
compared.

2. Methodology
Previous studies (e.g., Sadler 1975) have shown that
ITCZs are zonally orientated linear features characterized by convergence in the lower troposphere, with largescale vigorous ascent and then divergence in the upper
troposphere. Over the oceans, ITCZ locations coincide
with a mean sea level pressure trough, but this relationship does not hold over landmasses, where the pressure
field is dominated by thermally driven features, such as
heat lows (e.g., Racz and Smith 1999). The time-mean
position of the ITCZ is well captured by a time averages
of many fields including rainfall (see Fig. 1), low-level
convergence, vorticity, midtropospheric vertical motion,
etc. In this study, the objective ITCZ position is defined in
NWP output from the layer mean divergence in the layer
between 1000 and 850 hPa. This layer, rather than a single
level, is chosen to allow the analysis to be carried out over
both land and ocean. Because it is well known that
weather systems (e.g., tropical cyclones, easterly waves,
and even thunderstorm clusters) have their own divergence pattern, the divergence field used for a particular day is a time mean, as this reduced the impact of these
transient features. An infrared satellite image of the
tropical Pacific from August 2004 is shown in Fig. 2a with
streamlines displaying the layer-mean wind between 1000
and 850 hPa, which are averaged for the proceeding and

VOLUME 27

following 72 h. Across the eastern portion of the ocean


basin (east of the dateline) there are scattered cold cloud
tops (indicative of deep convection) that are closely
aligned with confluent streamlines. West of the dateline,
the convection is more organized and lies along a shear
line (a streamline with strong cyclonic curvature) that lies
west-northwesteast-southeast, with its western end near
Taiwan. It is these two large-scale features in particular
that would be recognized as ITCZs by synoptic analysts
and what is desired to be detected automatically.
The core diagnostic for the objective detection of
ITCZs is the location at which the magnitude of horizontal gradient of divergence is equal to zero in these
time- and layer-averaged fields. This contour is overlaid
on the 1000850-hPa time-averaged divergence field
(derived from the wind field shown in Fig. 2a), which is
shaded in Fig. 2b. The lines where the gradient of divergence is equal to zero lines up precisely with the regions that correspond with the ITCZs determined
subjectively by inspection of the infrared imagery and
streamlines in Fig. 2a. However, this contour also highlights features that are clearly not ITCZs, as this diagnostic picks out both lines of maximum convergence
and divergence in both the tropics and extratropics.
Thus, following the techniques of Hewson (1998) and
Berry et al. (2011), other fields are used as masks to
isolate only the relevant features.
First the divergence field is used to find only those lines
in convergent regions; thus, only zero contours (which are
convergence lines) where the layer- and time-mean divergence is less than some threshold are retained. Figure
2c shows where the gradient of divergence is equal to zero
when those contours in divergent regions are erased. It
can be seen that some features that are not convergence
maxima are still retained, where there is an inflexion in
the horizontal gradient of divergence due to a relative
minimum within a convergent region (e.g., near 08 latitude and 1608E). Such features are eliminated using the
Laplacian of divergence, which is displayed by the underlying shading in Fig. 2c. The contours where the gradient of divergence is equal to zero are removed when
this quantity is less than a threshold value, giving the
contour field shown in Fig. 2d. This shows convergence
lines and is close to what an analyst might manually draw
given the fields in Fig. 2a. By definition, the ITCZ is exclusively found in the tropics, and consequently a thermodynamic mask is added to remove convergence lines
within nontropical air masses, where the convergence
lines could be associated with features such as midlatitude
cyclones. In the example shown, there is a long convergence line near the dateline that passes near Fiji (158S and
1808 longitude) and extends into the extratropics. At the
time shown, the southern end of this convergence line

1 MARCH 2014

BERRY AND REEDER

1897

FIG. 2. Example showing the steps involved in the objective ITCZ technique using ERA-Interim data centered on 1200 UTC 13 Aug
2004. (a) 1000850-hPa layer-mean streamlines, averaged for 72 h before and after valid time, overlaid on infrared satellite imagery
(shaded; K). (b) Divergence (scaled by 105 and colored according to legend) with solid contour showing where the gradient of divergence
is equal to zero. (c) Second derivative of divergence (scaled by 1013 and colored according to legend) with solid contour showing where the
gradient of divergence is equal to zero in convergent regions only. (d) 850-hPa uw (K; colored according to legend) with a black contour
showing where the horizontal gradient of divergence is equal to zero only in convergent regions and where the second gradient of
divergence exceeds zero. (e) Objective ITCZ locations (red contours) derived from the fields in (d) with 1000850-hPa layer-mean
streamlines, averaged for 72 h before and after valid time, overlaid on infrared satellite imagery (shaded; K).

intersects with a slow-moving midlatitude cyclone. Therefore, at least part of this convergence line may not conform
to the synoptic definition of an ITCZ. The tropics are
well defined by a warm moist air mass, so the 850-hPa
wet bulb potential temperature (uw) is used as a discriminator between the tropics and extratropics in this
study. The 850-hPa uw for the example time is shown by
the color shading in Fig. 2d, where it can be seen that the

tropical regions are characterized by high values, with


relatively little horizontal variation. Consequently, this
field is used as a final mask to retain only lines of convergence in the tropical regions.
When these two kinematic and one thermodynamic
masks are applied to the zero contour of the horizontal
gradient of divergence, a line joining algorithm used by
Berry et al. (2011) in the detection of fronts is employed.

1898

JOURNAL OF CLIMATE

The masking of the fields is conducted numerically, as


opposed to graphically for quantitative results and because the drawing of contours has a dependency on the
nature of the plotting program. This algorithm joins
points based on their proximity to their nearest neighbor; if the distance between the two points is less than
a chosen threshold, the points are joined to form an
ITCZ line segment. Only line segments that consist of
more than a specified number of individual points and
span more than 158 of longitude in total are retained as
the interest here is in large-scale features. The thresholds are customizable and must be modified according to
the resolution of the input data to produce synoptically
sensible results. The objective, numerically defined
ITCZ locations for the example time are displayed in
Fig. 2e. In this study, when a qualifying ITCZ line segment is encountered, the individual points along it are
examined further. The layer-mean wind components at
the grid points surrounding the identified ITCZ point
are recorded for the computation of composite kinematic
fields.
It is recognized that in addition to the ITCZ these
diagnostics will also pick out monsoon troughs and other
persistent large-scale tropical convergence zones, such
as the North African intertropical front (ITF; e.g., L
el
e
and Lamb 2010). In this study, no attempt it made to
separate these phenomena and they are all referred to as
the ITCZ. The rationale is they lack precise definitions
suitable for an automated scheme; for example, although the ITCZ is commonly defined by confluent
easterlies, whereas the monsoon trough is commonly
defined by westerlies on its equatorward side, the area
over which these westerlies must extend is not clear or
necessarily consistent across all regions and seasons. In
some instances these phenomena may blend into one
another, and thus the distinction might become semantic
rather than physical: broad convergence zones in the
moist tropics are likely to be associated with deep convection and large-scale ascent.
The ITCZ locations are computed using the Interim
European Centre for Medium-Range Weather Forecasts
(ECMWF) Re-Analysis (ERA-Interim). This global
dataset spans the period 19792010 and is used at 1.58
horizontal resolution. The threshold values used for the
background divergence and its Laplacian are set through
a subjective comparison with satellite imagery and the
ITCZ locations drawn on the unified surface analysis from
the U.S. National Weather Service. For the 1.58 ERAInterim data a five-point smoother was applied to the input wind field and it was determined that a maximum
background divergence of 21 3 1026 s21 and minimum
Laplacian of divergence of 2 3 10213 s21 gave the closest
match to the manually analyzed fields. The line joining

VOLUME 27

algorithm was then set to scan 3.58 from a candidate ITCZ


point for nearby neighbors. To retain only the large-scale
features, ITCZ segments in the ERA-Interim must comprise of five or more individual points and the total zonal
extent of the segment must exceed 158 of longitude. The
850-hPa uw threshold used to define the tropics is determined by the annual mean of this field from ERA-Interim averaged between 308N and 308S, which is 300 K.
As mentioned earlier, it was decided to derive the
ITCZ locations based upon a 144-h running average of
the 1000850-hPa layer-mean winds in order to reduce
the effect of isolated synoptic events that have expected
time scales of approximately 25 days. The sensitivity of
the results to the choice of averaging period is summarized by Fig. 3, which shows the average ITCZ count for
August 2000 computed using different averaging periods. It can be seen that increasing the averaging period
has a qualitatively similar effect to increasing a plotting
threshold: the regions of highest feature counts become
more pronounced, but the overall pattern of relative
high and low counts is unchanged. Given that the overall
pattern is robust with varying averaging periods (consistent with the persistent nature of the ITCZ) the
overall results presented here are not changed substantially if this period is altered. Although the choice is
somewhat arbitrary, here a 6-day average is used on the
basis that many synoptic features of the tropical troposphere (tropical cyclones, African easterly waves, etc.)
have shorter periods.
The ITCZ locations are computed using fixed
thresholds for the entire ERA-Interim period and the
kinematic fields at the surrounding grid points are
retained for further calculations. Seasonal, monthly, and
annual averages are used to discuss the climatology of
these features and the linear changes during these averaging periods are used to discuss trends. The association
between the ITCZ and tropical cyclones is examined in
this study using the International Best Track Archive for
Climate Stewardship (IBTrACS) tropical cyclone dataset
(Knapp et al. 2010).

3. Results
a. Annual means
The ITCZ locations and composite kinematic quantities were computed from the ERA-Interim and the
annual averages for the period 19792010 are shown in
Fig. 4. The annual-mean ITCZ count (Fig. 4, top) shows
a pattern this is consistent with results from previous
studies (e.g., Waliser and Gautier 1993): well-defined
count maxima spanning the Northern Hemisphere Pacific
and Atlantic Oceans and a maximum in the southern
Indian Ocean and the south Pacific convergence zone

1 MARCH 2014

BERRY AND REEDER

FIG. 3. Sensitivity of ITCZ count for August 2000 produced using running averages of 8501000-hPa layer-mean
winds averaged from (top) 0 to (bottom) 610 days.

1899

1900

JOURNAL OF CLIMATE

FIG. 4. Annual means derived using the objective ITCZ locations computed using
ERA-Interim data for 19792010. (top) Count (month21; shaded), (middle) divergence (s21; shaded), and (bottom) relative vorticity (s21; shaded). Black contours in
the middle and bottom panels indicate where the count shown in the top panel exceeds 10 month21.

VOLUME 27

1 MARCH 2014

BERRY AND REEDER

(SPCZ). Smaller-scale regional maxima are observed in


association with the worlds monsoon systems in North
Africa, South America, and Australasia and near the
Indian subcontinent. These features are likely best described as monsoon troughs or intertropical fronts.
Over the oceans, the ITCZs are zonally orientated,
with the exception of the SPCZ, which has its characteristic north-northwestsouth-southeast tilt. The count
values over the oceans are diffused over a larger geographical region in the western parts of basins, suggesting larger variance in the seasonal or day-to-day
locations of the ITCZ. The ITCZ is most consistently
found in the eastern Pacific, near 908W, where there is
the highest global count constrained within a latitudinally thin band. In the eastern Pacific, there is also evidence of a second ITCZ in the Southern Hemisphere,
which has been observed frequently (e.g., Zhang 2001)
and is a persistent feature in both general circulation
models (e.g., Lin 2007) and aquaplanet-type model
experiments (e.g., Chao and Chen 2004; Fig. 1). This
Southern Hemisphere eastern Pacific ITCZ extends into
the central Pacific, where it merges with the SPCZ.
The annual-mean composite divergence along the
ITCZ is shown in the middle panel of Fig. 4. Note that at
each grid point this field is the average of only the times
when the ITCZ is identified and thus is a composite,
rather than a simple average. In an annual sense, the
strongest convergence occurs when ITCZs are located
over the landmasses, particularly West Africa and
northern Australia. Over the oceans, it is apparent that
when ITCZs exist in the east Pacific and Atlantic there is
significantly more convergence (approximately double)
than that found when ITCZs occur over other oceanic
basins. The corresponding map of composite relative
vorticity along the ITCZ is shown in the bottom panel of
Fig. 4, where the values in the Southern Hemisphere
have been multiplied by 21 for clarity. This demonstrates that the annual-mean composite relative vorticity
along the major ITCZ features is cyclonic but the cyclonic vorticity in Atlantic, SPCZ, eastern Pacific, and
central Pacific ITCZ maxima is relatively low. High
vorticity values are only found in these basins in the rare
occasions when the ITCZs are shifted poleward from their
climatological position (cf. Fig. 4. top). Compared to these
regions, the mean relative vorticity within the regions of
high ITCZ counts in the western Pacific and Indian Ocean
ITCZ are about 35 times greater, suggesting that these
may be somewhat different phenomena, such as monsoon
troughs (cf. Molinari and Vollaro 2013).

b. Variability
The seasonal cycle of ITCZ count is shown as maps in
Fig. 5 with zonal averages for some specific regions in

1901

Fig. 6. Over the continents, the ITCZ count and location


varies in phase with the climatology of the local monsoon;
for example, peak counts occur over North Africa in
JuneAugust (JJA) and over Australia during December
February (DJF), with very low counts during the local dry
seasons, as might be anticipated. Over the oceans, the
ITCZ are more persistent year-round features. In the
Atlantic, the ITCZ peak has the largest seasonal shift, as
it moves poleward by approximately 108 of latitude
between MarchMay (MAM) and JJA. During the
same period, ITCZ features become 21% more frequent. In the eastern Pacific, the Northern Hemisphere
ITCZ behaves like that in the Atlantic (albeit with
a smaller shift in latitude) and the Southern Hemisphere ITCZ is only truly evident for one season
(MAM). Further analysis indicates that this box only
exhibits simultaneous ITCZs in both hemispheres on
15% of the total analysis times. The western Pacific
northern ITCZ maximum also migrates poleward during summer, although the shift is small (approximately
38) and the feature count remains consistent throughout
the year. The SPCZ count is more interesting as from the
graph (Fig. 6b), it is evident that the Southern Hemisphere ITCZ count does not change during the year, but
the count spreads across a larger range of latitudes, extending as far poleward as 308S in DJF and/or MAM. In
the Indian Ocean this seasonal perspective reveals that
there is also a double ITCZ during most of the year. This
basin is dominated by a Southern Hemisphere ITCZ
that also migrates poleward during the local summer.
The second (Northern Hemisphere) ITCZ is present
just poleward of the equator during Austral summer and
disappears during boreal summer when convergence
zones associated with the Asian monsoon become evident
at higher latitudes.
The most significant large-scale interseasonal climate
mode is the El Ni~
noSouthern Oscillation (ENSO),
which is characterized by changes in the areal extent and
intensity of low SSTs in the equatorial east Pacific.
Given that convection over the oceans is expected to
have a relationship with the underlying SST, some variation of the ITCZ locations and/or intensity may be
anticipated with changes in the ENSO conditions. This is
determined here by examining the difference in ITCZ
count and divergence between warm and cold ENSO
phases, as defined by the Southern Oscillation index
(SOI) compiled by the Australian Bureau of Meteorology. Figure 7 shows these differences between an SOI
index of 110 (La Ni~
na) and 210 (El Ni~
no), such that
positive values in these plots indicate increased values
during La Ni~
na conditions. The most obvious change in
the nature of the ITCZ is seen in the Pacific Ocean and
is consistent with the changes in the underlying SST.

1902

JOURNAL OF CLIMATE

VOLUME 27

FIG. 5. Seasonal cycle of ITCZ (left) count (month21; shaded) and (right) divergence (s21; scaled by 105 and shaded).

Relative to El Ni~
no conditions, during La Ni~
na the
Northern Hemisphere cross-basin ITCZ is displaced
poleward by more than 58. The opposite appears true
with the Atlantic ITCZ, where in La Ni~
na conditions the
ITCZ is located farther equatorward. Around the Maritime Continent and in the eastern Indian Ocean, La Ni~
na
conditions increase the frequency of the ITCZ and the
convergence associated with it, as might be expected from
rainfall composites (e.g., Ropelewski and Halpert 1987).
During La Ni~
na conditions there is a dipole in both count
and mean convergence along the SPCZ, centered on its
approximate midpoint. This suggests that the western
portion of the SPCZ becomes more active, whereas the
eastern half becomes less active under La Ni~
na conditions.

c. Trends
The trends for annual-mean ITCZ count and divergence are shown in the top panel of Figs. 8 and 9,
respectively, as a percentage change relative to their 1979
2009 means. The lower panels in each of these figures
show zonal averages of the actual annual means as
a function of time across six specific boxes that have
interesting trends. Only regions where the trend is

significant at the 80% confidence level using a two-sided


Students t test are plotted in Figs. 8 and 9. As should be
expected, the statistically significant trends in both fields
are confined to the regions of high annual count. In the
Indian Ocean (region i), the ITCZ count has a negative
trend over much of the basin; in particular, the maximum in the Northern Hemisphere practically disappears
after the mid-1990s. In the southern Indian Ocean, the
ITCZ count has a slight negative trend, but the convergence in that region has increased near the count
maximum. In terms of total vertical mass flux, these factors will offset one another and may not change the total
significantly. At the longitude of the Malay Peninsula
(region ii) there are strong trends in both ITCZ count
and divergence that indicate a strengthening of the
ITCZ along the western coast of Sumatra. It appears
that prior to the mid-1980s the ITCZ was essentially
present only in the Southern Hemisphere and has become less frequent over the ERA-Interim period. After
approximately 1988 there is an abrupt increase in the
ITCZ count along the equator and in the Northern
Hemisphere, which could be linked to changes in the
SOI during this period.

1 MARCH 2014

BERRY AND REEDER

1903

FIG. 6. Seasonal-mean ITCZ feature counts, averaged in longitude for specific boxes shown in the map.

A large positive trend in ITCZ count exists between


Papua New Guinea and Borneo, just south of the equator. From the timelatitude plots (region iii), it is evident
that the majority of this trend is associated with the appearance and intensification of the ITCZ between the
equator and 58S from approximately 1990 onward. In the

western Pacific, the count in both the equatorial ITCZ


and SPCZ have negative trends on their northern flank
and positive trends on their southern flank, suggesting
that the whole pattern has shifted to the south during the
reanalysis period. The timelatitude plot of the annualmean count in this box (region iv) supports this, showing

1904

JOURNAL OF CLIMATE

VOLUME 27

FIG. 7. (top) Difference in annual-mean ITCZ feature count between positive SOI (.10 units) and negative SOI (,210 units).
(bottom) As in (top), but for the divergence field along objectively identified ITCZ features. The black contour in each shows where the
annual-mean count is equal to 10 month21.

a perceptible southward movement of the count maxima


in both hemispheres throughout this period. Globally, the
largest significant trends in both quantities occurs over
the Amazon (region v), where both the number and
strength (i.e., divergence) of ITCZ features have increased over the ERA-Interim period. The timelatitude
plots indicate that these trends are caused by an abrupt
shift in the time-mean location of the ITCZ. It is apparent
that around 2003 the ITCZ locations move from being
predominantly on or just north of the equator (and over
the ocean) to being mostly south of the equator. The divergence along ITCZs along both latitudes decreases
during this period but most strongly in the Southern
Hemisphere. The physical reason for this change is not
clear; thus, this should be investigated further in the future. However, given such an abrupt change, the possibility of changes in the observing system used in the
reanalysis should not be discounted. In the eastern tropical Atlantic (region vi) the ITCZ counts have increased
near the equator, whereas the composite divergence
along these ITCZ has not changed. The timelatitude
plot indicates that this positive trend in count arises because of a gradual southward expansion of high ITCZ
counts from the Northern Hemisphere.

d. The ITCZ and tropical cyclogenesis


It is well known that the background cyclonic relative
vorticity and high-tropospheric humidity of the ITCZ
provides a conducive environment for the formation of
tropical cyclones (e.g., McBride 1995). Molinari and
Vollaro (2013) noted that many previous studies had
stated that a large percentage of western North Pacific
tropical cyclones (TCs) formed within the monsoon
trough without a clear definition of the term. These
authors wanted to determine more precisely how many
TCs developed in the northern west Pacific monsoon
trough. Some objective definitions, based on the 850-hPa
relative vorticity field, were developed, and they found
that 50%100% of tropical cyclones formed within this
region, depending on their precise definition and the state
of ENSO. With the objective ITCZ dataset, here a similar
question can be posed globally: that is, how close is TC
genesis to the ITCZ?
Using the IBTrACS tropical cyclone dataset, the distance between TC genesis and objectively identified
ITCZ features is considered. Because the ITCZ identification methodology uses a 6-day running average (672 h)
of the wind field, and the TCs themselves are associated

1 MARCH 2014

BERRY AND REEDER

1905

FIG. 8. (top) Linear trend (total change; %) of annual-mean ITCZ count for the period 19792010 with regions of interest denoted by
rectangles labeled ivi. Trends shown are significant at the 80% confidence level using a two-sided Students t test. (bottom) Time
longitude averages of the annual ITCZ counts for the regions ivi in the top panel.

with strong low-level convergence, the distance of a TC


genesis location from ITCZ locations derived from wind
field centered 3 days previously is considered. This time
offset, along with the other tracking criteria, reduces the

possibility of the synoptic-scale convergence associated


with an isolated TC being wrongly aliased as an ITCZ.
The computed mean distance of TC genesis locations
from the closest ITCZ is displayed in Fig. 10. As should be

1906

JOURNAL OF CLIMATE

VOLUME 27

FIG. 9. As in Fig. 8, but showing the total trend (%) and annual means (1025 s21) for the 1000850-hPa layer-mean divergence.

expected, the map (Fig. 10a) shows that this distance is low
where the ITCZ count (Fig. 4, top) is high, as the ITCZ
counts and TC genesis locations are both peaked in the
same narrow geographical areas. Farther from the equator
the mean distance increases, presumably as a consequence
of both the lower ITCZ count and alternative pathways for cyclogenesis (McTaggart-Cowan et al. 2013).

The histogram indicates that globally 50% of TCs form


within 600 km of an objectively identified ITCZ.

4. Discussion and conclusions


In this study, an objective method has been developed for the purpose of identifying the ITCZ in gridded

1 MARCH 2014

BERRY AND REEDER

1907

FIG. 10. (a) Mean distance (km; shaded) between tropical cyclone genesis location in the IBTrACS dataset and closest objectively
identified ITCZ with 19792010 total tropical cyclone genesis count (contoured every 5 above 5). (b) Histograms showing the distribution
of distances between tropical cyclone genesis locations and nearest ITCZ using the IBTrACS dataset.

numerical datasets based upon the time- and layer-averaged horizontal wind field in the lowest part of the
troposphere. On day-to-day time scales (Fig. 2), the ITCZ
locations are consistent with synoptic experience and
expectations and thus have been used here to examine
ITCZ behavior and trends over a 30-yr period in the
ERA-Interim dataset. It is found that the overall distribution of the ITCZ count is consistent with subjective or
satellite analyses (Fig. 4), with elongated zones occurring
over the tropical oceans and monsoon regions. Most of
these zones move poleward during summer and equatorward in winter. These results are certainly not new, but
what is new is that an automated objective methodology
has been employed to detect the location of these features and give more detailed information about them. An
important result is that a qualitative and repeatable analysis of the impact of ENSO and the recent (30 yr) trends in
ITCZ properties has been given. The work presented here

provides one of the first automated, objective reference


studies of ITCZ features, which can be repeated and
quantitatively compared with other gridded datasets,
such as other reanalysis products or climate models.
The most persistent ITCZ is found over the eastern
Pacific Ocean, where the ITCZ is confined to a narrow
latitudinal band year-round, which likely reflects the
variability of the underlying sea surface temperatures
and the flow around the adjacent subtropical anticylones. The Northern Hemisphere east Pacific ITCZ
also possesses the strongest convergence of all the oceanic ITCZ regions during all seasons. The count statistics also pick out a second, weaker and less frequent
equatorial ITCZ that extends across the entire Pacific
during MAM (see Fig. 5, left), which intersects the
SPCZ near the dateline. There are also two ITCZ
maxima in the Indian Ocean present throughout the
year, although they are most pronounced during the

1908

JOURNAL OF CLIMATE

same season, which is reminiscent of double ITCZs that


appear in general circulation models when forced with
observed sea surface temperatures (e.g., Lin 2007).
The analysis here picks out the SPCZ rather differently from the other tropical convergence zones. Not
only does it have a characteristic tilt that means it extends across a range of latitudes, it is the only major
ITCZ-type feature that does not shift meridionally
during the course of the year. The SPCZ core remains at
the same latitude and activity spreads farther poleward
during Austral summer. This may highlight some different dynamics associated with the SPCZ, such as the
interaction with subtropical fronts (Berry et al. 2011).
The statistically significant annual trends of ITCZ
properties are confined to regions of climatologically high
ITCZ counts. Over most regions, the convergence within
the oceanic ITCZ has increased over the last three decades
consistent with the strengthening of the Hadley circulation
noted in previous studies (e.g., Stachnik and Schumacher
2011; Mitas and Clement 2005). In the western Pacific,
there is a pattern in both ITCZ count and divergence that
suggests that the equatorial ITCZ and SPCZ have shifted
southward, which is supported by the timelatitude plots in
Figs. 8 and 9. Using objective fronts to identify thermal
discontinuities across the SPCZ, a similar change in the
SPCZ was noted by Berry et al. (2011). The physical
mechanism responsible for this shift is unclear: the ITCZ
location is likely determined by a combination of the underlying surface (e.g., locations of high and low SSTs) and
the large-scale flow in both the tropics and extratropics.
Over land, one particularly interesting trend is found in
northern South America; here there is a strong positive
trend in the ITCZ count and the mean convergence,
whereas just offshore there is a negative trend in both,
associated with an abrupt shift in the ITCZ around 2003.
The physical mechanism driving this is not clear, but it may
be associated with observed changes in the sea surface
temperatures along the South American coast (see, e.g.,
Evans and Braun 2012; Lumpkin and Garzoli 2011).
By comparing the objective ITCZ locations with data
from the IBTrACS dataset, it was found that more than
50% of TCs formed within 600 km of the ITCZ. From
a synoptic perspective, this result is unsurprising as
the ITCZ can provide the background cyclonic vorticity required for cyclogenesis and it is found that TC
genesis and ITCZ are generally peaked in the same
geographical location, consistent with recent work by
Molinari and Vollaro (2013). The natural extension of
this type of analysis is in the analysis of output from climate models that are either too coarse to resolve TCs or
only store data averaged over a period. If these simulations suggest changes in the location or intensity of the
ITCZ, this may be used as a proxy to infer general

VOLUME 27

changes in TC genesis and help to test hypotheses of


changing TC activity.
The diagnostic methods shown here are simple to
apply in any gridded dataset with the necessary lowertropospheric input fields (i.e., zonal and meridional wind
components). Because the ITCZ is an important fixture
of the tropical climate, it is recommended that future
studies apply these calculations to output from climate
models. Doing so will allow the skill of climate models in
the tropics in the present day to be objectively determined. Consequently, different models or different versions of the same model can be easily and quantitatively
compared. The diagnostics will also permit an automated
and consistent assessment of the future state of the ITCZ
and forecasts of how conditions could change.
Acknowledgments. The authors wish to thank three
anonymous reviewers for their helpful comments. This
work is funded by Australian Research Council Grant
FS100100081.
REFERENCES
Berry, G., C. Jakob, and M. Reeder, 2011: Recent global trends in
atmospheric fronts. Geophys. Res. Lett., 38, L21812, doi:10.1029/
2011GL049481.
Brown, J. R., S. B. Power, F. P. Delage, R. A. Colman, A. F. Moise,
and B. F. Murphy, 2011: Evaluation of the South Pacific convergence zone in IPCC AR4 climate model simulations of the
twentieth century. J. Climate, 24, 15651582.
Chan, S. C., and J. L. Evans, 2002: Comparison of the structure of
the ITCZ in the west Pacific during the boreal summers of
198993 using AMIP simulations and ECMWF reanalysis.
J. Climate, 15, 35493568.
Chao, W. C., and B. Chen, 2004: Single and double ITCZ in an
aqua-planet model with constant sea surface temperature and
solar angle. Climate Dyn., 22, 447459.
Evans, J. L., and A. Braun, 2012: A climatology of subtropical
cyclones in the South Atlantic. J. Climate, 25, 73287340.
Gu, G., R. F. Adler, and A. H. Sobel, 2005: The eastern Pacific ITCZ
during the boreal spring. J. Atmos. Sci., 62, 11571174.
Hewson, T. D., 1998: Objective fronts. Meteor. Appl., 5, 3765.
Joyce, R. J., J. E. Janowiak, P. A. Arkin, and P. Xie, 2004: CMORPH:
A method that produces global precipitation estimates from
passive microwave and infrared data at high spatial and temporal resolution. J. Hydrometeor., 5, 487503.
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, 363376.
Lele, I., and P. J. Lamb, 2010: Variability of the intertropical front
(ITF) and rainfall over the West African SudanSahel zone.
J. Climate, 23, 39844004.
Lin, J.-L., 2007: The double-ITCZ problem in IPCC AR4 coupled
GCMs: Oceanatmosphere feedback analysis. J. Climate, 20,
44974525.
Lumpkin, R., and S. Garzoli, 2011: Interannual to decadal changes
in the western South Atlantics surface circulation. J. Geophys.
Res., 116, C01014, doi:10.1029/2010JC006285.

1 MARCH 2014

BERRY AND REEDER

McBride, J. L., 1995: Tropical cyclone formation. Global Perspectives


on Tropical Cyclones, R. L. Elsberry, Ed., World Meteorological Organization, 63105.
McTaggart-Cowan, R., T. J. Galarneau Jr., L. F. Bosart, R. W.
Moore, and O. Martius, 2013: A global climatology of
baroclinically influenced tropical cyclogenesis. Mon. Wea. Rev.,
141, 19631989.
Mitas, C. M., and A. Clement, 2005: Has the Hadley cell been
strengthening in recent decades? Geophys. Res. Lett., 32, L03809,
doi:10.1029/2004GL021765.
Molinari, J., and D. J. Vollaro, 2013: What percentage of western
North Pacific tropical cyclones form within the monsoon
trough? Mon. Wea. Rev., 141, 499505.
R
acz, Z., and R. K. Smith, 1999: The dynamics of heat lows. Quart.
J. Roy. Meteor. Soc., 125, 225252.

1909

Ropelewski, C. F., and M. S. Halpert, 1987: Global and regional


scale precipitation patterns associated with the El Ni~
no/
Southern Oscillation. Mon. Wea. Rev., 115, 16061626.
Sadler, J. C., 1975: The upper tropospheric circulation over the
global tropics. University of Hawaii Department of Meteorology Rep. UHMET-75-05, 35 pp.
Stachnik, J. P., and C. Schumacher, 2011: A comparison of the
Hadley circulation in modern reanalyses. J. Geophys. Res.,
116, D22102, doi:10.1029/2011JD016677.
Waliser, D. E., and C. Gautier, 1993: A satellite-derived climatology of the ITCZ. J. Climate, 6, 21622174.
Wang, C.-C., and G. Magnusdottir, 2006: The ITCZ in the central
and eastern Pacific on synoptic time scales. Mon. Wea. Rev.,
134, 14051421.
Zhang, C., 2001: Double ITCZs. J. Geophys. Res., 106 (D11),
11 78511 792.

Copyright of Journal of Climate is the property of American Meteorological Society and its
content may not be copied or emailed to multiple sites or posted to a listserv without the
copyright holder's express written permission. However, users may print, download, or email
articles for individual use.

Anda mungkin juga menyukai