Next Article in Journal
The Positive Effects of Linked Control Policy for Vessels Passing Through Locks on Air Quality—A Case Study of Yichang, China
Previous Article in Journal
Backward Integration of Nonlinear Shallow Water Model: Part 2: Vortex Merger
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Consistency of Changes in the Ascending and Descending Positions of the Hadley Circulation Using Different Methods

1
College of Ocean and Meteorology, Guangdong Ocean University, Zhanjiang 524088, China
2
South China Sea Institute of Marine Meteorology, Guangdong Ocean University, Zhanjiang 524088, China
3
CMA-GDOU Joint Laboratory for Marine Meteorology, Guangdong Ocean University, Zhanjiang 524088, China
4
National Meteorological Information Center, Beijing 100081, China
5
State Key Laboratory of Earth System Numerical Modeling and Application, Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100029, China
6
Wenling Meteorological Bureau, Wenling 317599, China
*
Authors to whom correspondence should be addressed.
Atmosphere 2025, 16(4), 367; https://doi.org/10.3390/atmos16040367
Submission received: 7 February 2025 / Revised: 16 March 2025 / Accepted: 19 March 2025 / Published: 24 March 2025
(This article belongs to the Section Meteorology)

Abstract

:
The shift in the intertropical convergence zone (ITCZ) and the poleward expansion of the Hadley circulation termini have attracted many investigations, since they affect the hydrological cycle and hence the societies and ecosystems in the tropical and subtropical areas. Using the observed precipitation and three atmospheric reanalysis data sets, different methods have been employed to quantify the changes in the ITCZ position, the Hadley circulation width, terminus position, and center intensity in both hemispheres over the global and seven longitudinal sections. It is found that the ITCZ position from the centroid method is closer to the equator over the global and ocean sections than that from the maximum precipitation method and the mass streamfunction, but the variability between different methods and data sets has significant correlations. The large spread of the ITCZ latitude is mainly from the different methods used. The ITCZ position has shifted away from the equator over 1983–2023, which is consistent across data sets, and the multi-method mean trend from five significant trends is 0.22 ± 0.12°/decade over this period. The south HC branch terminus is expanding poleward; this shift, computed using different methods and data sets, is consistent, and five out of seven are significant. The terminus position shift in the north branch is mixed, and most trends are insignificant except that from P-E. The global mean south branch circulation width has a significant increasing trend, contributed mainly by the northward shift in the ITCZ position; meanwhile, the north circulation width is shrinking insignificantly over 1983–2023. The cross-equatorial atmospheric energy transport AHT and the ITCZ position θITCZ from ERA5 are generally anti-correlated, and the correlation coefficients between AHT and θITCZ from different methods are all significant. The multi-method mean northward shift of θITCZ is 3.48 °PW−1.

1. Introduction

The Hadley circulation (HC) plays an important role in modulating the meridional distribution of temperature, humidity, and precipitation at low latitudes. Therefore, changes in the strength and position of the HC have profound climatic impacts in tropical and subtropical regions [1]. Recent studies have shown the tightening of the ascending branch and the poleward expansion of the descending branch of the Hadley circulation in the warming climate, based on observations, atmospheric reanalyses, and climate model simulations [2,3,4,5,6,7,8,9,10,11]. These changes will alter the intertropical convergence zone (ITCZ) position and precipitation spatial distribution over the tropical and subtropical areas, impacting society and ecosystems [12].
The ITCZ is the ascending branch of the Hadley circulation [10,13] and accounts for 32% of the global precipitation [14,15], making it the largest source of latent heating in the atmosphere [11,16]. Precipitation, especially the extreme precipitation, over the tropic area is increasing [17,18,19,20,21,22,23] with the rising temperature, due to more water vapor available in the atmosphere. The total column water vapor content has a trend of about 7%/K based on the Clausius–Clapeyron equation, and it has been confirmed by observations and model simulations [18,20]. The increasing trend of precipitation is particularly strong over the ITCZ location [14,24], consistent with the “wet getting wetter and dry getting drier” mechanism. The increasing trend of the ITCZ intensity is happening over the Central and Eastern Pacific [11,19,25,26] and other longitudinal sections [24,27].
The change in the ITCZ can not only influence the tropical area but can also affect global temperature and precipitation through an influence on the global energy distribution, because of the interaction between the cloud and radiation fields [9,10,28] and the energy transport by atmospheric circulation. Previous studies have shown that the ITCZ position will shift towards a warmer hemisphere [29], since the hemispheric energy imbalance drives the energy transport from the warmer hemisphere to the colder hemisphere, and the upper diverging branch of the Hadley circulation transports more energy than the lower branch due to the greater geopotential energy of the air in the upper branch [3]. The asymmetrical hemispheric heating due to the anthropogenic aerosol forcing and the greenhouse gas forcing may change the hemispheric energy imbalance and alter the ITCZ latitudinal position.
However, some existing studies showed that the position of the global mean ITCZ has not changed significantly under the warming climate [27,30], but the ITCZ width and intensity have changed significantly over recent decades [11,14,31] due to the combined effect of anthropogenic aerosols and greenhouse gases. The understanding of the physical processes of the ITCZ’s change is limited, and further research is needed [11,12,32,33,34].
The Hadley circulation has been reported to be widening due to the poleward expansion of the HC termini under the warming climate [34,35,36,37,38,39,40,41,42,43,44]. Some studies [35,45] have argued that the Hadley circulation widening is due to global warming and surface temperature gradient changes between the Southern and Northern Hemispheres. Using reanalysis and observational precipitation, Hu et al. [46] found that the Hadley cell half width coincided with changes in the ITCZ location. After analysis using model simulations, Lu et al. [4] suggested that the HC expansion is unlikely to originate from tropical processes, despite the fact that tropical heating is effective in driving the variation in the HC at interannual time scales. They found that the increase in the gross stability near the subtropics acts to suppress baroclinic instability, which is a critical factor controlling the limits of the outer boundaries of the HC. This is because the extratropical stabilization inhibits the breakdown of the thermally driven cell, allowing it to reach higher latitudes, resulting in the poleward expansion of the HC edges.
The position change of the ITCZ has been investigated by model simulations [30] and observations [27], but the observation period is too short to verify the consistency and robustness of the model results in previous studies; thus, longer time period observations are needed. The same is true for the verification of the poleward expansion of the Hadley circulation [3,4,40,46]. In this study, we will use the observed precipitation and atmospheric reanalysis data to re-visit the changes in the ITCZ position and the poleward expansion of the HC. Although the atmospheric reanalysis data are not real observations, a large amount of observations has been assimilated into the numerical weather forecasting model to obtain the consistent data set constrained by the laws of physics. The state-of-the-art reanalysis data are the optimized data from the combination of the model forecast and observations and can be regarded as the real state of the atmosphere. The objectives of this study are to use multiple methods, including the maximum precipitation and centroid methods, the mass streamfunction, and the extended Kuo–Eliassen (KE) equation, to check the consistency of the changes in the ITCZ position, Hadley circulation width, and terminus locations in two hemispheres. The KE equation is employed to identify the contributions of the friction and diabatic heating terms to the changes in these locations. The relationship between the hemispheric energy imbalance and the ITCZ position is further investigated.

2. Data and Methods

2.1. Data

In this study, the monthly means from the observed GPCP (Global Precipitation Climatology Project) precipitation are used. This is a merged data set containing data from land-based rain gauges, sounding observations, microwave radiometers, and infrared radiances from the Global Precipitation Climatology Centre [47], with a resolution of 2.5° × 2.5°; the GPCP has been widely used for long-term variation studies [20,21].
The data from the ECMWF (European Centre for Medium-Range Weather Forecast) fifth generation atmospheric reanalysis (ERA5) [48] are employed, including the total precipitation and the northward component of the wind velocity at each pressure level. The coverage period is from 1950–2023, and the horizontal resolution is 0.25° × 0.25°. There are 37 levels in the vertical direction. The ERA5 is the latest generation of the atmospheric reanalysis providing the atmospheric state close to reality due to the large amount of observational data being assimilated.
The MERRA2 (Modern-Era Retrospective Analysis for Research and Applications, version 2) is the latest atmospheric reanalysis of the modern satellite era produced by NASA’s Global Modeling and Assimilation Office (GMAO) [49]. MERRA-2 assimilates observations not available to its predecessor, MERRA, and includes updates to the Goddard Earth Observing System (GEOS) model and analysis scheme so as to provide a viable ongoing climate analysis beyond MERRA’s terminus. MERRA-2 includes the finite-volume dynamical core of a cube-sphere horizontal discretization at an approximate resolution of 0.5° × 0.625° and 72 hybrid-eta levels from the surface to 0.01 hPa.
The NCEP (National Center for Environmental Forecasting) CFSR (climate prediction system reanalysis) data set is also used here for comparison purposes [50]. It covers the time period from 1979 to the present. It is the first high-resolution global reanalysis data set coupling the atmosphere, oceans, land, and sea ice. Its horizontal resolution is T382 (approximately 38 km), with 64 layers of mixed coordinates used in the vertical direction. It has introduced the time-varying concentration of CO2 to more accurately reflect the climate change trend.
The daily data, including the air temperature, potential temperature, and wind velocity at each pressure level from ERA5, are also used to compute the friction and diabatic heating terms from the Kuo–Eliassen (KE) equation [51].
Since the ITCZ position differs depending on the geographic meridian, the tropical areas are divided into seven longitudinal sections following Liu et al. [46], including America (80–49° W), Africa (9–43.5° E), Atlantic (49° W–9° E), Indian Ocean (43.5–104° E), Western Pacific (104–162.5° E), Central Pacific (162.5° E–139.5° W), and Eastern Pacific (139.5–80° W).

2.2. Centroid Method

While the maximum zonal-mean precipitation can be used to determine the ITCZ position and intensity [27,52,53], the centroid method is also employed following Frierson and Hwang [3], Donohoe et al. [54], Burnett et al. [55], and Liao et al. [56] to calculate the ITCZ latitudinal position, in order to mitigate the effect of the low resolution in precipitation data on the ITCZ identification. The ITCZ latitudinal position ( θ I T C Z ) can be defined as
θ I T C Z = 25 25 ϕ × P × c o s ϕ d ϕ 25 25 P × c o s ϕ d ϕ
where ϕ is the latitude (°) and P the precipitation (mm/day). The integral is from 25° S to 25° N. The corresponding ITCZ intensity is defined as the mean precipitation of the two latitudes that are closest to the computed θ I T C Z latitude.

2.3. Streamfunction Method

Based on the method deployed by Oort and Yienger [57] and Byrne et al. [14], the mass streamfunction ( ψ ) can be defined as
ψ ϕ ,   p = 2 π R c o s ϕ g p s p v ϕ ,   p ¯ d p
where g is the acceleration of gravity, R is the Earth’s radius, p is the pressure, ps is the surface pressure, and v ϕ ,   p ¯ is the zonal-mean meridional component of the wind velocity. The mass streamfunction can better reveal the properties of the Hadley circulation and its vertical structure over the ITCZ location. When applied to longitudinal sections, the v ϕ ,   p ¯ from the divergent wind must be employed [58].

2.4. Kuo–Eliassen Equation

The extended Kuo–Eliassen (KE) equation [51], which is derived from the zonally averaged momentum, thermodynamic, continuity, and thermal wind equations, is also employed here [51]. The extended KE equation allows us to evaluate the contributions of eddy-driven vertical momentum, heat fluxes, and other forcing terms to the HC change. The equation is presented as follows:
f 2 g 2 π R cos ϕ 2 ψ p 2 L f + g 2 π R ϕ 1 R cos ϕ ψ ϕ S 2 L s 2 + g 2 π R R d p ϕ [ T ] ϕ ψ p L T +   f g 2 π R cos ϕ [ u ] p 2 ψ ϕ p + 2 [ u ] p 2 ψ ϕ 2 ( [ u ] cos ϕ ) p 2 ψ p ( [ u ] cos ϕ ) p 2 2 ψ ϕ p L u = R d p [ Q ] ϕ D Q R d p ϕ ( [ v T ] cos ϕ ) ϕ D v T f [ F λ ] p D X + f 2 ( [ u v ] cos 2 ϕ ) ϕ p D u v +   f 2 [ u ω ] p 2 D u ω R d p p p 0 R d / c p ϕ [ ω θ ] p D ω θ
The KE equation is an elliptic second-order partial differential equation, where R d is the gas constant of dry air, ρ is the air density, f is the Coriolis parameter, and C p is the specific heat capacity. u, v, and ω represent the zonal, meridional, and vertical components of wind, respectively. λ is the longitude, T is the air temperature, and θ is the potential temperature. The prime denotes deviations from the zonal-mean and the monthly mean, and [∙] represents the regional mean and the annual (or monthly) mean. The left-hand side of the equation is an elliptic second-order linear differential operator. According to Chemke et al. [16] and Hess et al. [59], the main terms contributing to the streamfunction ψ are the diabatic heating Q = T t + u T R cos ϕ λ + v T R ϕ + ω T p ω T p R d C p and the vertical gradient of zonally averaged friction [ F λ ] computed from the zonally averaged momentum equation. [ F λ ] = u ] [ v c o s 2 ϕ R cos ϕ ϕ + u v c o s 2 ϕ R cos ϕ ϕ + ( [ u ] [ ω ] ) p + ( [ u ω ] ) p f [ v ] . u v and v T are the eddy momentum and heat fluxes. u ω and ω θ are the vertical eddy and heat fluxes.
On the right-hand side of the equation, D Q represents the meridional gradient of the zonal-mean diabatic heating, D v T is the zonal-mean meridional eddy heat flux, D X is the vertical gradient of the zonal-mean zonal friction, D u v represents the zonal-mean meridional eddy momentum flux, D u ω is the zonal-mean vertical eddy momentum flux, and D ω θ is the zonal-mean vertical eddy heat flux. On the left-hand side of the equation, the second-order linear differential operators L s 2 , L T , and L u describe the structure of the zonal-mean state of the atmosphere, including the static stability and spatial structure of the zonal-mean zonal wind. L f represents a constant term associated with the Coriolis force. The daily data from ERA5 are employed when the KE equation is used in the calculation.
The Successive Over-Relaxation Method (SOR) has been employed to iteratively solve Equation (3) to analytically obtain the streamfunction corresponding to each forcing term [16]. For example, by applying the operator L ( L f + L s 2 + L T + L U ) to the forcing term D X , we obtain the zonal friction contribution ( ψ X ) to the streamfunction. This method is applied analogously to the other forcing terms to fully resolve the streamfunction (please see Zaplotnik et al. [51] for details).

2.5. Useful Variables Related to the HC Description

Several useful variables have been defined based on the mass streamfunction. Figure 1a is the multiannual mean (1950–2023) mass streamfunction of DJF (December, January, and February) from the ERA5 data. Since the mass streamfunction integrated from the surface upward and the northward flow of the wind component is defined as positive, the north branch of the Hadley circulation has a positive mass transport and the south branch shows a negative mass transport. The ascending, or ITCZ, latitude (θITCZ) is defined as the average latitude of the zero streamfunction at each pressure level from 700–300 hPa near the equator, which represents the strongest ascending position [60]. The other two mean zero line positions averaged over 700–300 hPa in the Southern and Northern Hemispheres are defined as θS and θN, respectively, representing the subtropical termini of the Hadley circulation (Figure 1b). The maximum and minimum streamfunction values are defined as intensities of the two circulations, and the corresponding latitude and height are defined as the center position and height. The circulation width is defined as θITCZθS for the south branch and θNθITCZ for the north branch. The zero location of the difference between precipitation and evaporation (P-E) can also be utilized as an indicator of the subtropical dry zones where the HC termini are located [61], so it is also used in this study for comparison.

2.6. Atmospheric Heat Transport

Following Loeb et al. [62], the cross-equatorial atmospheric heat or energy transport (AHT) can be calculated by
A H T = 1 2 ( R T F S A E t )
where ∆ denotes the Southern Hemisphere minus the Northern Hemisphere. R T is the net radiative flux at the top of atmosphere, F S is the net surface energy flux, and A E is the total atmospheric energy. AHT is converted from Wm−2 to PW using 1 Wm−2 = 0.255 PW (assuming a hemispheric surface area of 2.55 × 1014 m2). The contribution from the atmospheric energy tendency ( A E t ) is small and can be neglected [62].
In this study, the two-tailed t-test was used for the significance test of the Pearson correlation coefficient [63], and the Mann–Kendall test is used to check the significance of the trend. Unless specified, the significance level is defined as 95%. The flowchart of the data analysis in this study is shown in Figure 2.

3. Results

3.1. Changes in Precipitation Latitudinal Distribution

The precipitation latitudinal distribution will be examined first using the GPCP, ERA5, MERRA2, and CSFR data (Figure 3), in order to check the changes in the ITCZ and HC terminus strength. The top row of Figure 3 shows the latitudinal variations in the multiannual mean precipitation over four time periods, i.e., 1980–1989, 1990–1999, 2000–2009, and 2010–2019, and the period mean precipitation is noted as P1, P2, P3, and P4, respectively. The heaviest mean precipitation location north of the equator corresponds to the ITCZ position (indicated by the dashed vertical black line in Figure 3a–d), and the next peak south of the ITCZ corresponds to the south Pacific convergence zone (SPCZ, also indicated by the dashed vertical black line). The minimum locations (indicated by the dashed vertical red line) near 30° S and 30° N are the termini of the HC in both hemispheres.
The bottom row (Figure 3e–h) shows the differences between the mean precipitation (∆P2 = P2 − P1, ∆P3 = P3 − P1, ∆P4 = P4 − P1). The peaks of the ITCZ strength from the later three time periods become stronger and stronger compared with that from the first time period (1980–1989) in GPCP (Figure 3e), implying the upward trend of the ITCZ strength under the warming climate. However the P4 peak is stronger than P1 and P2 but weaker than P3 from the ERA5, MERRA2, and CSFR data. In Figure 3e, the zero location of ∆P on the left of the ITCZ peak is very close, but the one on the right of the ITCZ peak is quite spread. The peak of ∆P2 is lower than ∆P3, but the width between two zero locations (on both sides of the ITCZ peak) is wider than that of ∆P3, indicating the narrower and stronger ITCZ from P2 to P3. However, though the peak of ∆P4 is stronger than ∆P3, its width between two zero locations is wider than that of ∆P3, and therefore, it is hard to conclude that the ITCZ is becoming narrower under the warming climate. For the GPCP data, the precipitation at the HC terminus in the Southern Hemisphere (Figure 3e) decreases from the first period (1980–1989) to the second period (1990–1999), and then it increases over the third period (2000–2009) and decreases over the last period (2010–2019). In the Northern Hemisphere, the precipitation at the HC terminus increases from the first period to the second period, and then it decreases over the third period (2000–2009) but increases again over the last period (2010–2019). In general, the precipitation changes at the HC termini in both hemispheres have complicated decadal changes, meriting further investigation. For the ERA5 data, similar results are also shown in Figure 3f.
When comparing Figure 3e–g, the ∆Ps around the SPCZ locations are completely different between GPCP, ERA5 and MERRA2. The ∆Ps are all negative in Figure 3e (GPCP), they are close to zero in Figure 3f (ERA5), they are much larger in Figure 3g (MERRA2), and the ∆Ps are even larger around the SPCZ than that around the ITCZ location. The ∆Ps in Figure 3h (CFSR) are also different from the others; the largest ∆P is located south of the ITCZ, and the ∆Ps around the SPCZ have both negative (∆P2) and positive (∆P3 and ∆P4) values. All these results indicate there are systematic differences between the data sets in precipitation properties around the ITCZ and SPCZ locations.

3.2. ITCZ Position

Using climate model simulations, Byrne et al. [14] reported that there is no robust change in the global mean ITCZ location over the twenty-first century. Liu et al. [27] also showed little change in ITCZ locations over different longitudinal sections from 2000 to 2018. This is revisited in this study employing different methods, including the maximum zonal-mean precipitation location (Max P) and the centroid method for the GPCP, ERA5, MERRA2, and CFSR data sets. The mean 300–700 hPa zero location of the mass streamfunction (Ψ) is also used for ERA5 and MERRA2 data. The time series of the ITCZ latitudinal position over the global and seven longitudinal sections are shown in Figure 4, and the corresponding trends of the time series over 1983–2023 are listed in Table 1. The annual mean time series of the global mean ITCZ latitudinal position are shown in Figure 4a. Nine out of ten trend values are positive, but only five trends are significant at the confidence level of 0.05, and all trends from the ERA5 using three different methods are significant (Table 1).
The ITCZ location from the centroid method is closer to the equator than that from other methods in all data sets over the global and ocean basins, but it is different over the land sections (Figure 4b,c). The ITCZ positions from the streamfunction (Ψ) are farther away from the equator, but the ITCZ positions computed from other methods using different data sets oscillate around the equator over America (Figure 4b). Over Africa, the ITCZ positions from all data sets and methods oscillate around the equator (Figure 4c). The trends over land sections are mixed, varying with data sets and methods. Although there are consistent positive trends over the Central and Eastern Pacific, only a few of them are significant. The large spread of the time series is mainly from the different methods used.
For the ITCZ position variability, the correlation coefficients (r) between the global mean ITCZ locations from different methods and data sets over 1983–2023 are listed in Table 2. Generally, there is a significant correlation in the variability between time series from the same method, but the correlation between time series from the Max P method and the centroid method is small and insignificant. This is also true for other sections except for the Indian Ocean, Central Pacific and the Eastern Pacific where the correlation is significantly high. Most of the correlation coefficients between time series from the mass streamfunctions (Ψ) and other methods are significantly high in Table 2.
The correlation coefficients between ITCZ locations computed from the centroid method using the ERA5, MERRA2, and GPCP data over the global and seven sections are high and significant, and the scatter plot is shown in Figure 5. The highest correlation is r = 0.97 in the Western and Central Pacific.

3.3. Other Hadley Circulation Properties

The circulation center intensity, height, and latitude of the global HC in the Southern and Northern Hemispheres are computed by the mass streamfunction using ERA5 and MERRA2 data, and their time series are plotted in Figure 6. The center intensities of the south HC branch have an overall significant negative trend (−0.10 × 1010 kg/s/decade, which indicates strengthening) over 1950–2023 from the ERA5 data, but it increases from 1950 to 1982 and decreases from 1982 to 2004 and then increase again (Figure 6a). The minimum center intensity corresponds to the 1997/98 El Niño event. The center intensity in the north HC branch shows general significantly increasing trends (0.13 × 1010 kg/s/decade) from 1950 to 2023 using ERA5 data; it reaches the maximum during the 1997/98 El Niño event (Figure 6b) and then decreases quickly. During the common time period, there is good agreement between the two data sets, and the correlation coefficients are r = 0.90 and 0.60, respectively, for the Southern and Northern Hemispheres (Table 3).
For the ERA5 data, the center height in the southern circulation shows little trend over 1950–1982 (Figure 6c), but with large fluctuations. It then increases from 1983 onward, leading to an overall significant upward trend of m = −4.39 hPa/decade. Meanwhile, the center height in the northern circulation has large interannual fluctuation without a significant trend (Figure 6d). The center height from MERRA2 follows the increasing trend in the south branch before 2004 but becomes flat after that, and it has opposite changes to the ERA5 result after 1995 in the north branch. The correlation coefficients for the center height between the two data sets are all low and insignificant for the global mean HC, but it is noticed from Table 3 that the correlations are very high and significant over the seven longitudinal sections.
The center latitudes have no obvious trends in both hemispheres, but the center latitude in the Northern Hemisphere has decadal changes, with a decreasing trend over 1950–1982 and an increasing trend from 1983–2023. However, none of the trends are significant. The correlation coefficients between the center latitudes from ERA5 and MERRA2 are high and significant over the global and seven sections (Table 3).
The annual mean time series of the terminus latitude and the circulation width of two HC branches over the global and seven sections are also calculated using ERA5 and MERRA2 data with different methods, and the results for the global mean are plotted in Figure 7. The global trends are listed in Table 4, and those for the seven sections are listed in Table 5. The terminus positions and HC width are calculated based on the mass streamfunction (ψ) and P-E for all sections, and the Kuo–Eliassen equation (ψKE) is only employed for the global HC using ERA5 data. The ψ Q K E and ψ X K E represent the diabatic heating and friction term contributions in the Kuo–Eliassen equation. The trend displayed in each panel is from the mass streamfunction over 1983–2023 using ERA5 and MERRA2 data, and the unit is °/decade.
The zonal-mean terminus positions computed from ψ (ERA5), P-E (ERA5), KE equation (ψKE), and P-E (MERRA2) show significant negative trends in the Southern Hemisphere over 1983–2023, implying a significant poleward shift (Figure 7a and Table 4). For the seven sections, the P-E method shows significant negative trends in the America, Atlantic, and Eastern Pacific sections, whereas there is almost no significant trend from the mass streamfunction. For the north HC branch, only the P-E method shows significant trends of 0.29 ± 0.11°/decade and 0.33 ± 0.11°/decade for the ERA5 and MERRA2 data, respectively, while all trends from ψ and ψKE, as well as its friction and diabatic heating terms, are insignificant (Figure 7b and Table 4).
The correlation coefficients between ψ (ERA5) and ψ (MERRA2) are listed in Table 6 for the global and seven sections, and all correlations are high and significant, showing strong variability agreement between the results from both the ERA5 and MERRA2 data sets in both hemispheres. In the south HC branch, the circulation width defined as θITCZθS has a significant increasing trend from ψ (ERA5), ψKE, and its two terms ( ψ Q K E and ψ X K E ) over 1983–2023 (Figure 7c and Table 4). The trend of the circulation width from ψ (ERA5) is significantly negative in the Northern Hemisphere, but the rest of them are significant (Table 4). The trends of the terminus latitude and HC circulation width over the seven sections are also listed in Table 5 for reference.
To check how the two branches of the Hadley circulation are related, the seasonal variabilities in the center intensity, terminus latitude, and circulation width are calculated based on ERA5 and MERRA2 data and are plotted in Figure 8. The strong seasonal variability is overall opposite in the two branches. When the center intensity of the south branch is the strongest in July (Figure 8a), the center intensity in the north branch is the weakest (Figure 8b). The terminus location in the south branch is closely related to the center intensity. The stronger the intensity, the closer the terminus is to the equator (Figure 8c). However, the situation is not that simple in the north branch (Figure 8d). From January to June, the terminus location moves southward with the decrease in the circulation intensity but moves northward when the intensity starts to increase until September, and then it moves southward again even though the intensity keeps increasing. Therefore, the northward movement of the terminus location only occurs from June to September, and this seasonal variability is also confirmed by the P-E method (Figure 8d).
The seasonal variability in the circulation width in the south branch is also consistent with the center intensity (Figure 8e). The stronger the intensity, the wider is the circulation. The maximum width occurs in July when the terminus location is the closest to the equator, so the width increase is mainly from the northward shift of the ITCZ position. The terminus location has little effect on the circulation width in the north branch (Figure 8f), implying the dominance of the ITCZ position movement in the circulation width θNθITCZ. In general, there are good agreements between ERA5 and MERRA2 data for the seasonal variability in the center intensity, terminus latitude, and circulation width of the Hadley circulation over the global and seven sections.
The seasonal variation shows strong links between the circulation center intensity, terminus location, and circulation width. The correlation coefficients (r) between the center intensity, θITCZ, terminus location, and circulation width from ψ in each month are calculated and listed in Table 7 for the ERA5 and MERRA2 data sets. The significant r values at a significance level of 0.05 are in bold. In the south branch and for ERA5 data, 10 out of 12 r values between the center intensity and θITCZ are negative, indicating that as the southern circulation intensifies, the θITCZ shifts northward. However, the r values are significant mainly in the winter (November-February). In the north branch, the r values are significantly negative in April–June, implying that the stronger the intensity, the more southward the θITCZ is. The maximum and minimum latitudes of θITCZ, θITCZ, and θITCZ in each month are listed in Table 8 for reference.
Both the annual mean center intensity and terminus latitude of the south branch have negative trends (Figure 6a and Figure 7a), but only the r value of −0.41 in November is significant (see Table 5). In the north branch, ten out of twelve r values between the center intensity and θN are negative, including two significant ones. Nine out of twelve r values between the intensity and circulation width in the south branch are negative (see Table 7), including the two significant ones, implying that the stronger the intensity, the wider is the circulation. This widening is mainly from the northward shift of the θITCZ. The significant r value of −0.36 occurs in December, while the r between the intensity and θITCZ also has a significant correlation of −0.59 in December. In the north branch, the r values of −0.34 and −0.45 between the intensity and circulation width are also significant in January and December, indicating the shrinking of the circulation width with the increasing intensity. However, a significant positive correlation of r = 0.40 occurs in June when the circulation intensity is the weakest and the width is the narrowest during the year (Figure 8b,f), so the circulation width increases with the increasing intensity. Therefore, the relationship is different in different seasons.

3.4. Relationship Between AHT and ITCZ Location

It has been identified that the ITCZ position shift is closely related to the cross-equatorial atmospheric energy transport AHT [54,62,64]. The monthly mean time series of the AHT and ITCZ position θITCZ from the centroid method using ERA5 precipitation are plotted in Figure 9a. It can be seen that the AHT and θITCZ are roughly in the opposite phase, and this is confirmed by the negative lead correlation (AHT leads θITCZ) in Figure 9b. The strongest correlation occurs when the AHT leads θITCZ by zero or one month, depending on the method used in the calculation of θITCZ. The scatter plot of the annual mean AHT and θITCZ is shown in Figure 9c, the results from ψ Q K E and ψ X K E are also plotted for reference. The correlation coefficients are all significant at the significance level of 0.05, and they vary from r = −0.31 to −0.58 (without results from ψ Q K E and ψ X K E ) with the method used. The regression slope also varies with the method used for the θITCZ calculation; it shows that the annual mean ITCZ position will move northward at a rate of 1.46 to 6.65 °PW−1, and the mean shift is 3.48 °PW−1, which is larger than the 2.7 °PW−1 from CMIP3 coupled model simulations [54].

4. Discussion and Conclusions

Using the ERA5 and MERRA2 atmospheric reanalysis data and GPCP precipitation, the variabilities in the ascending (ITCZ) and terminus locations; the center intensity, height, and latitude; the width of the Hadley circulation in both hemispheres over different longitudinal sections; and the relationship between the hemispheric energy imbalance and the ITCZ location are quantitatively investigated, in order to check the consistency between different methods and confirm the recent findings.
The ITCZ position from the centroid method is closer to the equator than that from the maximum precipitation (Max P) method and the mass streamfunction over the global and ocean sections, but it is different over the continent sections. There is a significant variability correlation between them. There is a positive trend in the ITCZ position over the main study period (1983–2023) from different methods and data sets over the globe, implying the shift of the ITCZ position away from the equator, which is consistent across data sets and methods; furthermore, five out of nine increasing trends are significant (Table 1), and the mean of the five significant trends is 0.22 ± 0.12°/decade. This result is different from previous findings [27,30] and is sensitive to the time period selected, and the reason behind it merits further investigation.
There are asymmetrical changes in the Hadley circulation in the Southern and Northern Hemispheres due to different ocean coupling processes [1]. In the southern circulation, there is a significant increasing trend of the center height (Figure 6c), while the increasing trend in the northern circulation is insignificant (Figure 6d). The southern circulation center latitude identified from the mass streamfunction has an insignificant trend towards the equator, while both methods show an insignificant northward shift of the circulation center in the northern circulation.
Similar to the results from previous investigations [3,4,40,46], the south branch terminus latitude did expand poleward, and five out of seven trends from the different methods and data sets are significantly negative (Table 4). The terminus position shift in the north branch is mixed, and most trends are insignificant except that from P-E, which shows a significant positive trend (Table 4). The annual mean global south branch circulation width has a significant increase from different methods and data sets, contributed mainly by the northward shift of the ITCZ position. The north circulation width is shrinking insignificantly, due to the more northward shift in the ITCZ position and the insignificant poleward expansion of the terminus location. The overall expansion found here is much smaller than the previous findings of Hu and Fu [46], who pointed out that the expansion is about 2 to 4.5 degrees latitude over 20–30 years since 1979.
There are strong seasonal variations in the circulation width, center intensity, and terminus position. In the south branch, the circulation is the strongest and the widest, and the terminus position is the closest to the equator in July (Figure 8c,e). Therefore, the overall feature is that the stronger the circulation is, the closer the terminus location to the equator and the wider the circulation, consistent with previous studies [43]. In the north branch, the circulation center intensity is the weakest and the circulation width is the narrowest in June. The terminus location shows complicated variability. It is the closest to the equator in June but reaches the most northern point in September. There is a rapid movement over June-September. The trends and correlations listed in Table 7 and Table 8 show distinct differences from month to month. These characteristics need further investigation.
The cross-equatorial atmospheric energy transport AHT and the ITCZ position θITCZ are generally anti-correlated as found in previous studies [54,62], and the correlation coefficients between AHT and θITCZ from different methods are all significant (Table 7). The multi-method mean northward shift of θITCZ is 3.48 °PW−1 in the current atmosphere state.
This study has focused on the verification and quantification of existing findings in Hadley circulation changes using GPCP precipitation and ERA5 and MERRA2 data, in order to check the consistency of the results from different methods. The changes in the center intensity of the Hadley circulation are not fully analyzed in this study, since there are still a lot of arguments about it. While the climate model simulations show a weakening of the HC intensity, the atmospheric reanalyses show a strengthening [16]. The tropical circulation should weaken with the warming climate, because the increasing rate in tropical precipitation cannot keep up with the rapid increase in the water vapor in the lower troposphere, implying that the exchange of mass between the boundary layer and the middle troposphere must decrease [65,66], therefore weakening the HC intensity. The discrepancy between atmospheric reanalysis and the climate model simulations may be due to the data assimilation processes [67], and further investigations are needed.

Author Contributions

C.L., Y.Z. and J.L. contributed to the idea, funding, supervision, paper writing and editing. Q.S. and J.Q. contributed to the data search and analysis, paper writing and editing. Y.X., N.C., X.L., K.Y., R.Z., Z.L., L.J., K.H., K.J. and N.Z. All contributed to the paper revision and discussion. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation of China (42275017, 42075036); the Guangdong Province University Student Innovation and Entrepreneurship Training Program (S202210566019); and the Postgraduate Education Innovation Project of Guangdong Ocean University (202144; 202253).

Data Availability Statement

The GPCP data can be downloaded from https://psl.noaa.gov/data/gridded/data.gpcp.html (accessed on 11 December 2024). ERA5 data can be available from https://cds.climate.copernicus.eu/datasets (accessed on 11 December 2024). MERRA2 data can be downloaded from https://disc.gsfc.nasa.gov/ (accessed on 11 December 2024) and CFSR data from https://rda.ucar.edu/ (accessed on 11 December 2024).

Acknowledgments

We acknowledge the ECMWF for providing ERA5 data and the team making the GPCP data available. We thank two anonymous reviewers for reviewing this paper and providing constructive comments and suggestions.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chemke, R. Large Hemispheric Differences in the Hadley Cell Strength Variability Due to Ocean Coupling. Npj Clim. Atmos. Sci. 2022, 5, 1. [Google Scholar] [CrossRef]
  2. Byrne, M.P.; Pendergrass, A.G.; Rapp, A.D.; Wodzicki, K.R. Response of the Intertropical Convergence Zone to Climate Change: Location, Width, and Strength. Curr. Clim. Change Rep. 2018, 4, 355–370. [Google Scholar] [CrossRef] [PubMed]
  3. Frierson, D.M.W.; Hwang, Y.T. Extratropical influence on ITCZ shifts in slab ocean simulations of global warming. J. Clim. 2012, 25, 720–733. [Google Scholar]
  4. Lu, J.; Vecchi, G.A.; Reichler, T. Expansion of the Hadley cell under global warming. Geophys. Res. Lett. 2007, 34, L06805. [Google Scholar]
  5. Lu, J.; Deser, C.; Reichler, T. Cause of the widening of the tropical belt since 1958. Geophys. Res. Lett. 2009, 36, L03803. [Google Scholar] [CrossRef]
  6. Santer, B.D.; Wehner, M.F.; Wigley, T.M.L.; Sausen, R.; Meehl, G.A.; Taylor, K.E.; Ammann, C.; Arblaster, J.; Washington, W.M.; Boyle, J.S.; et al. Contributions of anthropogenic and natural forcing to recent tropopause height changes. Science 2003, 301, 479–483. [Google Scholar]
  7. Seidel, D.J.; Randel, W.J. Recent widening of the tropical belt: Evidence from tropopause observations. J. Geophys. Res. 2007, 112, D20113. [Google Scholar] [CrossRef]
  8. Seidel, D.J.; Fu, Q.; Randel, W.J.; Reichler, T.J. Widening of the tropical belt in a changing climate. Nat. Geosci. 2008, 1, 21–24. [Google Scholar]
  9. Su, H.; Jiang, J.H.; Neelin, J.D.; Shen, T.J.; Zhai, C.; Yue, Q.; Wang, Z.; Huang, L.; Choi, Y.S.; Stephens, G.L.; et al. Tightening of tropical ascent and high clouds key to precipitation change in a warmer climate. Nat. Commun. 2017, 8, 15771. [Google Scholar] [CrossRef]
  10. Tao, L.; Hu, Y.; Liu, J. Anthropogenic forcing on the Hadley Circulation in CMIP5 simulations. Clim. Dyn. 2016, 46, 3337–3350. [Google Scholar]
  11. Wodzicki, K.R.; Rapp, A.D. Long-term characterization of the Pacific ITCZ using TRMM, GPCP, and ERA-Interim. J. Geophys. Res. Atmos. 2016, 121, 3153. [Google Scholar]
  12. Mamalakis, A.; Randerson, J.T.; Yu, J.-Y.; Pritchard, M.S.; Magnusdottir, G.; Smyth, P.; Levine, P.A.; Yu, S.; Foufoula-Georgiou, E. Zonally contrasting shifts of the tropical rain belt in response to climate change. Nat. Clim. Change 2021, 11, 143–151. [Google Scholar] [CrossRef]
  13. Webster, P.J. The elementary Hadley circulation. In The Hadley Circulation: Present, Past, and Future; Diaz, H.F., Bradley, R.S., Eds.; Kluwer Academic: Dordrecht, The Netherlands, 2004; pp. 9–60. [Google Scholar]
  14. Byrne, M.P.; Schneider, T. Atmospheric dynamics feedback: Concept, simulations and climate implications. J. Clim. 2018, 31, 3249–3264. [Google Scholar] [CrossRef]
  15. Kang, S.M.; Shin, Y.; Xie, S.P. Extratropical forcing and tropical rainfall distribution: Energetics framework and ocean Ekman advection. Npj Clim. Atmos. Sci. 2018, 1, 20172. [Google Scholar] [CrossRef]
  16. Chemke, R.; Polvani, L.M. Opposite tropical circulation trends in climate models and in reanalyses. Nat. Geosci. 2019, 12, 528–532. [Google Scholar] [CrossRef]
  17. Adler, R.F.; Gu, G.; Sapiano, M.; Wang, J.J.; Huffman, G.J. Global Precipitation: Means, Variations and Trends During the Satellite Era (1979–2014). Surv. Geophys. 2017, 38, 679–699. [Google Scholar] [CrossRef]
  18. Allan, R.P.; Liu, C.; Zahn, M.; Lavers, D.A.; Koukouvagias, E.; Bodas-Salcedo, A. Physically consistent responses of the global atmospheric hydrological cycle in models and observations. Surv. Geophys. 2014, 35, 533–552. [Google Scholar] [CrossRef]
  19. Lau, K.M.; Wu, H.T. Detecting trends in tropical rainfall characteristics, 1979–2003. Int. J. Climatol. 2007, 27, 979. [Google Scholar]
  20. Liu, C.; Allan, R.P. Multisatellite observed responses of precipitation and its extremes to interannual climate variability. J. Geophys. Res. 2012, 117, D03101. [Google Scholar] [CrossRef]
  21. Liu, C.; Allan, R.P.; Huffman, G.J. Co-variation of temperature and precipitation in CMIP5 models and satellite observations. Geophys. Res. Lett. 2012, 39, L13803. [Google Scholar]
  22. Staten, P.W.; Lu, J.; Grise, K.M.; Davis, S.M.; Birner, T. Re-examining tropical expansion. Nat. Clim. Change 2018, 8, 768–775. [Google Scholar] [CrossRef]
  23. Staten, P.W.; Grise, K.M.; Davis, S.M.; Karnauskas, K.B.; Waugh, D.W.; Maycock, A.C.; Fu, Q.; Cook, K.; Adam, O.; Simpson, I.R.; et al. Tropical widening: From global variations to regional impacts. Bull. Am. Meteorol. Soc. 2020, 101, E897–E904. [Google Scholar] [CrossRef]
  24. Hari, V.; Villarini, G.; Karmakar, S.; Wilcox, L.J.; Collins, M. Northward propagation of the Intertropical Convergence Zone and strengthening of Indian summer monsoon rainfall. Geophys. Res. Lett. 2020, 47, e2020GL089823. [Google Scholar] [CrossRef]
  25. Gu, G.; Adler, R.F.; Huffman, G.J. Long-term changes/trends in surface temperature and precipitation during the satellite era (1979–2012). Clim. Dyn. 2016, 46, 1091. [Google Scholar] [CrossRef]
  26. Zhou, Y.; Xu, K.M.; Sud, Y.; Betts, A. Recent trends of the tropical hydrological cycle inferred from Global Precipitation Climatology Project and International Satellite Cloud Climatology Project data. J. Geophys. Res. 2011, 116, D09101. [Google Scholar] [CrossRef]
  27. Liu, C.; Liao, X.; Qiu, J.; Yang, Y.; Feng, X.; Allan, R.P.; Cao, N.; Long, J.; Xu, J. Observed variability of intertropical convergence zone over 1998–2018. Environ. Res. Lett. 2020, 15, 104011. [Google Scholar] [CrossRef]
  28. Bony, S.; Stevens, B.; Coppin, D.; Becker, T.; Reed, K.A.; Voigt, A.; Medeiros, B. Thermodynamic control of anvil cloud amount. Proc. Natl. Acad. Sci. USA 2016, 113, 8927–8932. [Google Scholar] [CrossRef]
  29. Schneider, T.; Bischoff, T.; Haug, G.H. Migrations and dynamics of the intertropical convergence zone. Nature 2014, 513, 45–53. [Google Scholar] [CrossRef]
  30. Donohoe, A.; Voigt, A. Why future shifts in tropical precipitation will likely be small: The location of the tropical rain belt and the hemispheric contrast of energy input to the atmosphere. In Climate Extremes: Patterns and Mechanisms; John Wiley and Sons: Hoboken, NJ, USA, 2017; pp. 115–226. [Google Scholar]
  31. Bellomo, K.; Clement, A.C. Evidence for weakening of the Walker circulation from cloud observations. Geophys. Res. Lett. 2015, 42, 7758–7766. [Google Scholar] [CrossRef]
  32. Byrne, M.P.; Schneider, T. Narrowing of the ITCZ in a warming climate: Physical mechanisms. Geophys. Res. Lett. 2016, 43, 350–375. [Google Scholar] [CrossRef]
  33. Hu, Y.; Zhou, C.; Liu, J. Observational evidence for poleward expansion of the Hadley circulation. Adv. Atmos. Sci. 2011, 28, 33–44. [Google Scholar] [CrossRef]
  34. Nguyen, H.; Evans, A.; Lucas, C.; Smith, I.; Timbal, B. The Hadley circulation in reanalyses: Climatology, variability, and change. J. Clim. 2013, 26, 3357–3376. [Google Scholar] [CrossRef]
  35. Adam, O.; Schneider, T.; Harnik, N. Role of changes in mean temperatures versus temperature gradients in the recent widening of the Hadley circulation. J. Clim. 2014, 27, 7450–7461. [Google Scholar]
  36. Ceppi, P.; Hartmann, D.L. Clouds and the atmospheric circulation response to warming. J. Clim. 2016, 29, 783–799. [Google Scholar] [CrossRef]
  37. Cubasc, U.; Stocker, T.F. Introduction. In Climate Change 2013: The Physical Science Basis; Cambridge University Press: Cambridge, UK, 2013; pp. 119–158. Available online: http://www.climatechange2013.org/images/report/WG1AR5_Chapter01_FINAL.pdf (accessed on 15 March 2021).
  38. Grise, K.M.; Davis, S.M.; Simpson, I.R.; Waugh, D.W.; Fu, Q.; Allen, R.J.; Staten, P.W. Recent tropical expansion: Natural variability or forced response? J. Clim. 2018, 32, 1551–1571. [Google Scholar] [CrossRef]
  39. Hu, Y.; Huang, H.; Zhou, C. Widening and weakening of the Hadley circulation under global warming. Sci. Bull. 2018, 63, 640–644. [Google Scholar] [CrossRef]
  40. Johanson, C.M.; Fu, Q. Hadley cell widening: Model simulations versus observations. J. Clim. 2009, 22, 2713–2725. [Google Scholar]
  41. Lionello, P.; D’Agostino, R.; Ferreira, D.; Nguyen, H.; Singh, M.S. The Hadley circulation in a changing climate. Ann. N. Y. Acad. Sci. 2024, 1534, 69–93. [Google Scholar] [CrossRef]
  42. Lucas, C.; Timbal, B.; Nguyen, H. The expanding tropics: A critical assessment of the observational and modeling studies. WIREs Clim. Change 2014, 5, 89–112. [Google Scholar]
  43. Seo, K.-H.; Yoon, S.-P.; Lu, J.; Hu, Y.; Staten, P.W.; Frierson, D.M.W. What controls the interannual variation of Hadley cell extent in the Northern Hemisphere: Physical mechanism and empirical model for edge variation. Npj Clim. Atmos. Sci. 2023, 6, 204. [Google Scholar] [CrossRef]
  44. Xia, Y.; Hu, Y.; Liu, J. Comparison of trends in the Hadley circulation between CMIP6 and CMIP5. Sci. Bull. 2020, 65, 1667–1674. [Google Scholar] [CrossRef] [PubMed]
  45. Yang, H.; Li, Q.; Wang, K.; Sun, Y.; Sun, D. Decomposing the meridional heat transport in the climate system. Clim. Dyn. 2015, 44, 2751–2768. [Google Scholar] [CrossRef]
  46. Hu, Y.; Fu, Q. Observed poleward expansion of the Hadley circulation since 1979. Atmos. Chem. Phys. 2007, 7, 5229. [Google Scholar] [CrossRef]
  47. Adler, R.F.; Gu, G.; Wang, J.J.; Huffman, G.J.; Curtis, S.; Bolvin, D.T. Relationships between global precipitation and surface temperature on interannual and longer timescales (1979–2006). J. Geophys. Res. 2008, 113, D22104. [Google Scholar]
  48. Hersbach, H.; Bell, B.; Berrisford, P.; Hirahara, S.; Horányi, A.; Muñoz-Sabater, J.; Simmons, A. The ERA5 global reanalysis. Q. J. R. Meteorol. Soc. 2020, 146, 1–51. [Google Scholar] [CrossRef]
  49. Gelaro, R.; McCarty, W.; Suárez, M.J.; Todling, R.; Molod, A.; Takacs, L.; Randles, C.A.; Darmenov, A.; Bosilovich, M.G.; Reichle, R.; et al. The Modern-Era Retrospective Analysis for Research and Applications, Version 2 (MERRA-2). J. Clim. 2017, 30, 5419–5453. [Google Scholar]
  50. Saha, S.; Moorthi, S.; Wu, X.; Wang, J.; Nadiga, S.; Tripp, P.; Behringer, D.; Hou, Y.-T.; Chuang, H.-Y.; Iredell, M.; et al. The NCEP Climate Forecast System Version 2. J. Clim. 2014, 27, 2185–2208. [Google Scholar] [CrossRef]
  51. Zaplotnik, Ž.; Pikovnik, M.; Boljka, L. Recent Hadley circulation strengthening: A trend or multidecadal variability? J. Clim. 2022, 35, 4157–4176. [Google Scholar] [CrossRef]
  52. Philander, S.G.H.; Gu, D.; Halpern, D.; Lambert, G.; Lau, N.C.; Li, T.; Pacanowski, R.C. Why the ITCZ is mostly north of the equator. J. Clim. 1996, 9, 2958–2972. [Google Scholar] [CrossRef]
  53. Sultan, B.; Janicot, S. Abrupt shift of the ITCZ over West Africa and intra-seasonal variability. Geophys. Res. Lett. 2000, 27, 3353–3356. [Google Scholar] [CrossRef]
  54. Donohoe, A.; Marshall, J.; Ferreira, D.; Mcgee, D. The relationship between ITCZ location and cross-equatorial atmospheric heat transport: From the seasonal cycle to the last glacial maximum. J. Clim. 2013, 26, 3597–3618. [Google Scholar] [CrossRef]
  55. Burnett, A.C.; Sheshadri, A.; Silvers, L.G.; Robinson, T. Tropical cyclone frequency under varying SSTs in aquaplanet simulations. Geophys. Res. Lett. 2021, 48, e2020GL091980. [Google Scholar] [CrossRef]
  56. Liao, X.; Holloway, C.E.; Feng, X.; Liu, C.; Lyu, X.; Xue, Y.; Bao, R.; Li, J.; Qiao, F. Observed interannual relationship between ITCZ position and tropical cyclone frequency. J. Clim. 2023, 36, 5587–5603. [Google Scholar] [CrossRef]
  57. Oort, A.H.; Yienger, J.J. Observed interannual variability in the Hadley circulation and its connection to ENSO. J. Clim. 1996, 9, 2751–2767. [Google Scholar] [CrossRef]
  58. Schwendike, J.; Berry, G.J.; Reeder, M.J.; Jakob, C.; Govekar, P.; Wardle, R. Trends in the local Hadley and local Walker circulations. J. Geophys. Res. Atmos. 2015, 120, 7599–7618. [Google Scholar] [CrossRef]
  59. Hess, O.; Chemke, R. Anthropogenic forcings reverse a simulated multi-century naturally-forced Northern Hemisphere Hadley cell intensification. Nat. Commun. 2024, 15, 4001. [Google Scholar] [CrossRef]
  60. Stachnik, J.P.; Schumacher, C. A comparison of the Hadley circulation in modern reanalyses. J. Geophys. Res. 2011, 116, D22102. [Google Scholar] [CrossRef]
  61. Chemke, R.; Polvani, L.M. Exploiting the abrupt 4 × CO2 scenario to elucidate tropical expansion mechanisms. J. Clim. 2018, 32, 859–875. [Google Scholar] [CrossRef]
  62. Loeb, N.G.; Wang, H.; Cheng, A.; Kato, S.; Fasullo, J.; Xu, K.; Allan, R.P. Observational constraints on atmospheric and oceanic cross-equatorial heat transports: Revisiting the precipitation asymmetry problem in climate models. Clim. Dyn. 2016, 46, 9–10. [Google Scholar] [CrossRef]
  63. Hipel, K.W.; Mcleod, A.I. Time Series Modelling of Water Resources and Environmental Systems; Elsevier: Amsterdam, The Netherlands, 1994; p. 1012. [Google Scholar]
  64. Liu, C.; Allan, R.P.; Mayer, M.; Hyder, P.; Desbruyères, D.; Cheng, L.; Xu, J.; Xu, F.; Zhang, Y. Variability in the global energy budget and transports 1985–2017. Clim. Dyn. 2020, 55, 3381–3396. [Google Scholar] [CrossRef]
  65. Held, I.M.; Soden, B.J. Robust responses of the hydrological cycle to global warming. J. Clim. 2006, 19, 5686–5699. [Google Scholar]
  66. Vecchi, G.A.; Soden, B.J.; Wittenberg, A.T.; Held, I.M.; Leetmaa, A.; Harrison, M.J. Weakening of tropical Pacific atmospheric circulation due to anthropogenic forcing. Nature 2006, 441, 73–76. [Google Scholar] [CrossRef] [PubMed]
  67. D’Agostino, R.; Lionello, P. Evidence of Global Warming Impact on the Evolution of the Hadley Circulation in ECMWF Centennial Reanalyses. Clim. Dyn. 2016, 48, 3047–3060. [Google Scholar] [CrossRef]
Figure 1. (a) Multi−month (December-February over 1950–2023 from ERA5) mean streamfunction showing the Hadley circulations. (b) Diagram showing the ITCZ location θITCZ and the southern and northern termini (θS and θN) of the Hadley circulation. They are defined as the averaged latitude of the zero streamfunction at each pressure level from 700–300 hPa.
Figure 1. (a) Multi−month (December-February over 1950–2023 from ERA5) mean streamfunction showing the Hadley circulations. (b) Diagram showing the ITCZ location θITCZ and the southern and northern termini (θS and θN) of the Hadley circulation. They are defined as the averaged latitude of the zero streamfunction at each pressure level from 700–300 hPa.
Atmosphere 16 00367 g001
Figure 2. Flowchart of the data analysis in this study.
Figure 2. Flowchart of the data analysis in this study.
Atmosphere 16 00367 g002
Figure 3. The top row is the latitudinal variations of multiannual mean precipitation (P1, P2, P3, and P4) over four time periods (1980–1989, 1990–1999, 2000–2009, 2010–2019) from (a) GPCP, (b) ERA5, (c) MERRA2 and (d) CFSR. The bottom row is the difference between the mean precipitation (∆P2 = P2 − P1, ∆P3 = P3 − P1, ∆P4 = P4 − P1) for (e) GPCP, (f) ERA5, (g) MERRA2 and (h) CFSR data.
Figure 3. The top row is the latitudinal variations of multiannual mean precipitation (P1, P2, P3, and P4) over four time periods (1980–1989, 1990–1999, 2000–2009, 2010–2019) from (a) GPCP, (b) ERA5, (c) MERRA2 and (d) CFSR. The bottom row is the difference between the mean precipitation (∆P2 = P2 − P1, ∆P3 = P3 − P1, ∆P4 = P4 − P1) for (e) GPCP, (f) ERA5, (g) MERRA2 and (h) CFSR data.
Atmosphere 16 00367 g003
Figure 4. Time series of the (a) global annual mean ITCZ position calculated using the maximum precipitation (Max P) location and the centroid method from GPCP, ERA5, MERRA2, and CFSR. The mean 300–700 hPa zero location of the mass streamfunction (Ψ) is also computed for ERA5 and MERRA2 data. The time period is 1950–2023 for ERA5 and 1979–2023 for others. The same time series are also plotted for seven other longitudinal sections: (b) America, (c) Africa, (d) Atlantic, (e) Indian Ocean, (f) Western Pacific, (g) Central Pacific, and (h) Eastern Pacific. All lines are six-year running mean.
Figure 4. Time series of the (a) global annual mean ITCZ position calculated using the maximum precipitation (Max P) location and the centroid method from GPCP, ERA5, MERRA2, and CFSR. The mean 300–700 hPa zero location of the mass streamfunction (Ψ) is also computed for ERA5 and MERRA2 data. The time period is 1950–2023 for ERA5 and 1979–2023 for others. The same time series are also plotted for seven other longitudinal sections: (b) America, (c) Africa, (d) Atlantic, (e) Indian Ocean, (f) Western Pacific, (g) Central Pacific, and (h) Eastern Pacific. All lines are six-year running mean.
Atmosphere 16 00367 g004
Figure 5. Scatter plot between ITCZ locations calculated from the centroid method using the ERA5, MERRA2, and GPCP data over the (a) global, (b) America, (c) Africa, (d) Atlantic, (e) Indian Ocean, (f) Western Pacific, (g) Central Pacific, and (h) Eastern Pacific. (i,j) In the bottom row are the scatter plots between global mean ITCZ positions from the streamfunction (Ψ) and the centroid method. The x-axis in (i) is centroid GPCP, and it is the ITCZ location computed using the reanalysis data in (j). ERA5 data are in black, and the MERRA2 data are in red. The data period is from 1983 to 2023 using ERA5 data. The correlation coefficients between the global mean ITCZ locations are high using the streamfunction (Ψ) and the centroid method, but they are insignificant over seven sections.
Figure 5. Scatter plot between ITCZ locations calculated from the centroid method using the ERA5, MERRA2, and GPCP data over the (a) global, (b) America, (c) Africa, (d) Atlantic, (e) Indian Ocean, (f) Western Pacific, (g) Central Pacific, and (h) Eastern Pacific. (i,j) In the bottom row are the scatter plots between global mean ITCZ positions from the streamfunction (Ψ) and the centroid method. The x-axis in (i) is centroid GPCP, and it is the ITCZ location computed using the reanalysis data in (j). ERA5 data are in black, and the MERRA2 data are in red. The data period is from 1983 to 2023 using ERA5 data. The correlation coefficients between the global mean ITCZ locations are high using the streamfunction (Ψ) and the centroid method, but they are insignificant over seven sections.
Atmosphere 16 00367 g005
Figure 6. Annual mean time series of the global mean circulation center (a,b) intensity, (c,d) height, and (e,f) latitude for the southern and northern circulations. The black line is from ERA5, and the red one is from MERRA2.
Figure 6. Annual mean time series of the global mean circulation center (a,b) intensity, (c,d) height, and (e,f) latitude for the southern and northern circulations. The black line is from ERA5, and the red one is from MERRA2.
Atmosphere 16 00367 g006
Figure 7. Annual mean time series of the (a,b) terminus positions and (c,d) width of the Hadley circulation in Southern and Northern Hemispheres. They are calculated based on the mass streamfunction (ψ), Kuo–Eliassen equation (ψKE), and P-E. The ψ Q K E and ψ X K E represent the diabatic heating and friction term contributions in the Kuo−Eliassen equation. The trend displayed in each panel is for the mass streamfunction over 1983–2023, and the unit is °/decade.
Figure 7. Annual mean time series of the (a,b) terminus positions and (c,d) width of the Hadley circulation in Southern and Northern Hemispheres. They are calculated based on the mass streamfunction (ψ), Kuo–Eliassen equation (ψKE), and P-E. The ψ Q K E and ψ X K E represent the diabatic heating and friction term contributions in the Kuo−Eliassen equation. The trend displayed in each panel is for the mass streamfunction over 1983–2023, and the unit is °/decade.
Atmosphere 16 00367 g007
Figure 8. Multiannual mean (1983–2023) seasonal variability of the (a,b) center intensity, (c,d) terminus latitude, and (e,f) circulation width of the Hadley circulation in the Southern and Northern Hemispheres. The error bar is one standard deviation. They are computed using the mass streamfunction (ψ), KE equation, and P-E, and the data sets used are ERA5 and MERRA2.
Figure 8. Multiannual mean (1983–2023) seasonal variability of the (a,b) center intensity, (c,d) terminus latitude, and (e,f) circulation width of the Hadley circulation in the Southern and Northern Hemispheres. The error bar is one standard deviation. They are computed using the mass streamfunction (ψ), KE equation, and P-E, and the data sets used are ERA5 and MERRA2.
Atmosphere 16 00367 g008
Figure 9. (a) Monthly time series of cross-equatorial atmospheric energy transport AHT and the ITCZ position θITCZ. AHT is multiplied by 3, and the unit is PW (Petawatt = 1015 W). (b) Correlation coefficient between AHT and θITCZ. The month is the time that the AHT leads θITCZ. (c) Scatter plot of the annual mean AHT and θITCZ. The regression slope and correlation coefficient are also displayed in Table 9.
Figure 9. (a) Monthly time series of cross-equatorial atmospheric energy transport AHT and the ITCZ position θITCZ. AHT is multiplied by 3, and the unit is PW (Petawatt = 1015 W). (b) Correlation coefficient between AHT and θITCZ. The month is the time that the AHT leads θITCZ. (c) Scatter plot of the annual mean AHT and θITCZ. The regression slope and correlation coefficient are also displayed in Table 9.
Atmosphere 16 00367 g009
Table 1. Trends of the annual mean ITCZ latitudinal position from 1983 to 2023 over the global and seven longitudinal sections. They are computed from the maximum precipitation (Max P), the centroid method, and the mass streamfunction ( ψ ) methods. The unit is °/decade. m is the trend, ∆m is the error of m, and p in parentheses is the confidence level.
Table 1. Trends of the annual mean ITCZ latitudinal position from 1983 to 2023 over the global and seven longitudinal sections. They are computed from the maximum precipitation (Max P), the centroid method, and the mass streamfunction ( ψ ) methods. The unit is °/decade. m is the trend, ∆m is the error of m, and p in parentheses is the confidence level.
GlobalAmericaAfricaAtlanticIndian OceanWestern PacificCentral PacificEastern Pacific
m ± ∆m (p)m ± ∆m (p)m ± ∆m (p)m ± ∆m (p)m ± ∆m (p)m ± ∆m (p)m ± ∆m (p)m ± ∆m (p)
Max P (GPCP)0.28 ± 0.12 (0.06)0.41 ± 0.31 (0.21)0.03 ± 0.21 (0.37)0.17 ± 0.25 (0.25)0.21 ± 0.26 (0.28)0.47 ± 0.32 (0.14)0.31 ± 0.30 (0.37)0.02 ± 0.22 (0.77)
Max P (ERA5)0.37 ± 0.14 (0.01)0.44 ± 0.19 (0.11)−0.83 ± 0.21 (<0.01)0.03 ± 0.16 (0.67)−0.62 ± 0.27 (0.04)−0.32 ± 0.24 (0.33)0.42 ± 0.24 (0.11)0.12 ± 0.17 (0.39)
Max P (CFSR)0.17 ± 0.14 (0.23)−0.31 ± 0.23 (0.02)0.09 ± 0.26 (0.66)−0.07 ± 0.1 (0.12)0.76 ± 0.19 (<0.01)−0.56 ± 0.30 (0.13)0.42 ± 0.27 (0.45)0.36 ± 0.22 (0.16)
Max P (MERRA2)0.30 ± 0.13 (0.01)−0.05 ± 0.19 (0.77)−0.10 ± 0.12 (0.74)0.08 ± 0.16 (0.27)0.93 ± 0.29 (<0.01)−0.15 ± 0.23 (0.25)0.42 ± 0.29 (0.29)0.10 ± 0.15 (0.63)
Centroid
(GPCP)
0.13 ± 0.04 (<0.01)0.08 ± 0.08 (0.32)0.15 ± 0.04 (<0.01)0.14 ± 0.05 (0.02)0.19 ± 0.07 (<0.01)0.06 ± 0.10 (0.87)0.05 ± 0.13 (0.90)0.18 ± 0.10 (0.11)
Centroid
(ERA5)
0.0 ± 0.04 (0.39)−0.41 ± 0.12 (0.01)−0.08 ± 0.06 (0.21)−0.20 ± 0.06 (<0.01)0.15 ± 0.06 (0.10)−0.09 ± 0.09 (0.20)0.09 ± 0.11 (0.58)0.11 ± 0.07 (0.09)
Centroid
(CFSR)
0.08 ± 0.03 (0.04)0.31 ± 0.08 (<0.01)−0.18 ± 0.04 (<0.01)0.14 ± 0.05 (0.02)0.05 ± 0.06 (0.65)−0.01 ± 0.09 (0.65)0.08 ± 0.09 (0.55)0.18 ± 0.08 (0.02)
Centroid
(MERRA2)
0.05 ± 0.04 (0.11)−0.11 ± 0.08 (0.25)0.02 ± 0.06 (0.63)0.11 ± 0.06 (0.35)0.14 ± 0.08 (0.05)−0.03 ± 0.08 (0.66)0.09 ± 0.11 (0.55)0.06 ± 0.09 (0.45)
ψ   ( E R A 5 ) 0.24 ± 0.07 (<0.01)1.59 ± 0.68 (0.02)0.10 ± 0.37 (0.49)0.01 ± 0.34 (0.74)−0.14 ± 0.19 (0.65)0.97 ± 0.47 (0.10)0.07 ± 0.74 (0.69)0.31 ± 0.48 (0.44)
ψ   (MERRA2) 0.05 ± 0.07 (0.68)1.34 ± 0.61 (0.02)0.35 ± 0.35 (0.44)0.26 ± 0.27 (0.58)−0.16 ± 0.23 (0.20)−0.44 ± 0.58 (0.36)0.80 ± 0.59 (0.13)0.52 ± 0.45 (0.23)
Table 2. Correlation coefficients between the global mean ITCZ position time series from different methods and data sets in Figure 4. The significant values at the confidence level of 0.05 are in bold.
Table 2. Correlation coefficients between the global mean ITCZ position time series from different methods and data sets in Figure 4. The significant values at the confidence level of 0.05 are in bold.
Max P (ERA5)Max P (CFSR)Max P (MERRA2)Centroid (GPCP)Centroid (ERA5)Centroid (CFSR)Centroid (MERRA2)Ψ (ERA5)Ψ (MERRA2)
Max P (GPCP)0.770.750.470.230.240.030.260.380.33
Max P (ERA5) 0.770.400.110.13−0.160.160.270.20
Max P (CFSR) 0.240.010.01−0.090.100.180.17
Max P (MERRA2) 0.070.03−0.20−0.070.360.11
Centroid (GPCP) 0.960.770.930.770.68
Centroid (ERA5) 0.770.960.760.72
Centroid (CFSR) 0.850.520.65
Centroid (MERRA2) 0.690.74
Ψ (ERA5) 0.79
Table 3. Correlation coefficients between the results computed from the ERA5 and MERRA2 data sets using the mass streamfunction over 1983–2023. The computed values are the HC center intensity, height, and latitude over the global and seven longitudinal sections. The trend unit is 1010 kg/s/decade for intensity, hPa/decade for center height, and °/decade for center latitude. The significant values at the confidence level of 0.05 are in bold.
Table 3. Correlation coefficients between the results computed from the ERA5 and MERRA2 data sets using the mass streamfunction over 1983–2023. The computed values are the HC center intensity, height, and latitude over the global and seven longitudinal sections. The trend unit is 1010 kg/s/decade for intensity, hPa/decade for center height, and °/decade for center latitude. The significant values at the confidence level of 0.05 are in bold.
IntensityHeightLatitude
RegionSouthNorthSouthNorthSouthNorth
Global0.900.600.170.220.850.88
America0.940.840.490.730.600.72
Atlantic0.890.900.510.530.700.45
Africa0.690.510.470.500.560.48
Indian Ocean0.740.810.420.210.790.63
Western Pacific0.810.820.720.520.860.71
Central Pacific0.920.910.850.790.780.76
Eastern Pacific0.920.900.720.730.730.65
Table 4. Trends of the annual mean global mean terminus position and width of the southern and northern Hadley circulations over 1983–2023 from ERA5 and MERRA2 using different methods. The unit is °/decade. m is the trend, ∆m is the error of m, and p in parentheses is the confidence level. The significant values at the confidence level of 0.05 are in bold.
Table 4. Trends of the annual mean global mean terminus position and width of the southern and northern Hadley circulations over 1983–2023 from ERA5 and MERRA2 using different methods. The unit is °/decade. m is the trend, ∆m is the error of m, and p in parentheses is the confidence level. The significant values at the confidence level of 0.05 are in bold.
DataSouth Branch
Terminus Position
North Branch Terminus PositionSouth Circulation WidthNorth Circulation Width
Ψ (ERA5)−0.21 ± 0.06 (<0.01)−0.12 ± 0.10 (0.29)0.45 ± 0.08 (<0.01)−0.31 ± 0.11 (0.01)
Ψ (MERRA2)0.01 ± 0.07 (0.97)0.05 ± 0.10 (0.78)0.07 ± 0.08 (0.51)0.01 ± 0.12 (0.96)
ψ X K E −0.22 ± 0.08 (0.02)−0.01 ± 0.17 (0.97)0.28 ± 0.11 (0.04)−0.15 ± 0.18 (0.47)
ψ Q K E −0.12 ± 0.13 (0.33)0.17 ± 0.15 (0.11)0.65 ± 0.18 (<0.01)−0.33 ± 0.18 (0.07)
ψKE−0.16 ± 0.06 (0.01)0.05 ± 0.09 (0.51)0.34 ± 0.11 (0.01)−0.07 ± 0.16 (0.61)
P-E (ERA5)−0.27 ± 0.07 (<0.01)0.29 ± 0.11 (0.04)
P-E (MERRA2)−0.30 ± 0.06 (<0.01)0.33 ± 0.11 (<0.01)
Table 5. Trends of the annual mean terminus position and width of the southern and northern Hadley circulations over seven sections. They are calculated using ERA5 and MERRA2 data from 1983 to 2023 with different methods. The unit is °/decade. m is the trend, ∆m is the error of m, and p in parentheses is the confidence level.
Table 5. Trends of the annual mean terminus position and width of the southern and northern Hadley circulations over seven sections. They are calculated using ERA5 and MERRA2 data from 1983 to 2023 with different methods. The unit is °/decade. m is the trend, ∆m is the error of m, and p in parentheses is the confidence level.
AmericaAfricaAtlanticIndian OceanWestern PacificCentral PacificEastern Pacific
Data Sourcem ± ∆m (p)m ± ∆m (p)m ± ∆m (p)m ± ∆m (p)m ± ∆m (p)m ± ∆m (p)m ± ∆m (p)
South branchTerminus
position
ψ   (ERA5)0.05 ± 0.19 (0.92)−0.08 ± 0.21 (0.94)−0.27 ± 0.23 (0.29)−0.06 ± 0.13 (0.57)−0.34 ± 0.26 (0.51)1.43 ± 0.87 (0.08)0.27 ± 0.29 (0.63)
ψ   (MERRA2)0.13 ± 0.24 (0.81)0.06 ± 0.18 (0.35)−0.06 ± 0.19 (0.92)−0.13 ± 0.13 (0.47)−0.24 ± 0.26 (0.44)0.19 ± 0.96 (0.68)−0.37 ± 0.35 (0.05)
P-E (ERA5)−0.85 ± 0.24 (<0.01)−0.15 ± 0.05 (<0.01)−0.54 ± 0.16 (<0.01)0.15 ± 0.11 (0.15)−0.25 ± 0.12 (0.15)1.27 ± 0.65 (0.07)−0.62 ± 0.21 (<0.01)
P-E (MERRA2)−1.27 ± 0.30 (<0.01)−0.09 ± 0.05 (0.15)−0.73 ± 0.15 (<0.01)0.29 ± 0.11 (0.03)0.07 ± 0.18 (0.97)−0.09 ± 0.69 (0.83)−0.73 ± 0.18 (<0.01)
WidthERA51.54 ± 0.78 (0.05)0.19 ± 0.36 (0.65)0.29 ± 0.35 (0.48)−0.08 ± 0.22 (1.00)1.32 ± 0.47 (0.01)−1.36 ± 1.44 (0.51)0.04 ± 0.56 (0.80)
MERRA21.22 ± 0.67 (0.07)0.29 ± 0.32 (0.26)0.32 ± 0.31 (0.36)−0.03 ± 0.27 (0.49)−0.20 ± 0.58 (0.61)0.62 ± 1.25 (0.69)0.90 ± 0.59 (0.19)
North branchTerminus
position
ψ   (ERA5)0.47 ± 0.31 (0.07)−0.19 ± 0.10 (0.06)0.15 ± 0.17 (0.54)0.26 ± 0.10 (0.03)−1.31 ± 0.95 (0.31)−0.18 ± 0.35 (0.48)−0.28 ± 0.25 (0.09)
ψ   (MERRA2)0.07 ± 0.30 (0.99)−0.30 ± 0.14 (0.04)−0.06 ± 0.17 (0.60)0.16 ± 0.11 (0.15)−0.76 ± 0.62 (0.21)0.18 ± 0.35 (0.71)−0.38 ± 0.27 (0.15)
P-E (ERA5)0.15 ± 0.18 (0.52)0.19 ± 0.14 (0.21)0.02 ± 0.18 (0.60)−0.12 ± 0.39 (0.80)0.64 ± 0.37 (0.02)0.17 ± 0.11 (0.17)0.27 ± 0.19 (0.14)
P-E (MERRA2)0.69 ± 0.19 (<0.01)0.22 ± 0.18 (0.31)0.43 ± 0.15 (<0.01)0.03 ± 0.42 (0.99)0.19 ± 0.43 (0.24)0.34 ± 0.10 (<0.01)−0.10 ± 0.22 (0.61)
WidthERA5−1.12 ± 0.64 (0.09)−0.30 ± 0.41 (0.40)0.14 ± 0.45 (0.92)0.40 ± 0.19 (0.23)−2.28 ± 0.99 (0.06)−0.25 ± 0.59 (0.58)−0.59 ± 0.67 (0.24)
MERRA2−1.27 ± 0.59 (0.03)−0.65 ± 0.44 (0.21)−0.32 ± 0.37 (0.58)0.33 ± 0.24 (0.17)−0.32 ± 0.83 (0.69)−0.62 ± 0.49 (0.21)−0.90 ± 0.65 (0.09)
Table 6. Correlation coefficients between the results from ψ (ERA5) and ψ (MERRA2) over 1983–2023 for the terminus latitude and circulation width of HC. p is the confidence level.
Table 6. Correlation coefficients between the results from ψ (ERA5) and ψ (MERRA2) over 1983–2023 for the terminus latitude and circulation width of HC. p is the confidence level.
South BranchNorth Branch
Terminus LatitudepCirculation WidthpTerminus Latitude Circulation Widthp
Global0.61<0.010.57<0.010.88<0.010.83<0.01
America0.71<0.010.48<0.010.49<0.010.370.02
Africa0.71<0.010.66<0.010.90<0.010.75<0.01
Atlantic0.77<0.010.52<0.010.76<0.010.75<0.01
Indian Ocean0.76<0.010.57<0.010.68<0.010.58<0.01
Western Pacific0.86<0.010.55<0.010.60<0.010.350.02
Central Pacific0.43<0.010.58<0.010.84<0.010.63<0.01
Eastern Pacific0.84<0.010.76<0.010.87<0.010.87<0.01
Table 7. Correlation coefficient (r) between monthly mean center intensity, θITCZ, terminus location, and circulation width. The significant r values are in bold after applying the two-tailed test using Pearson critical values at the level of 5%. The significant values at the confidence level of 0.05 are in bold.
Table 7. Correlation coefficient (r) between monthly mean center intensity, θITCZ, terminus location, and circulation width. The significant r values are in bold after applying the two-tailed test using Pearson critical values at the level of 5%. The significant values at the confidence level of 0.05 are in bold.
JANFEBMARAPRMAYJUNJULAUGSEPOCTNOVDEC
ERA5Southern
circulation
r (Intensity and θITCZ)−0.21−0.340.01−0.17−0.240.16−0.07−0.01−0.22−0.06−0.51−0.59
r (Intensity and θS)0.18−0.15−0.030.080.03−0.080.08−0.09−0.05−0.11−0.41−0.26
r (Intensity and width)−0.35−0.150.03−0.19−0.180.16−0.090.03−0.160.07−0.23−0.36
Northern
circulation
r (Intensity and θITCZ)0.230.250.05−0.47−0.57−0.47−0.20−0.28−0.18−0.55−0.060.35
r (Intensity and θN)−0.33−0.15−0.22−0.30−0.060.09−0.210.08−0.11−0.02−0.08−0.31
r (Intensity and width)−0.34−0.27−0.180.330.300.40−0.120.20−0.000.170.02−0.45
MERRA2Southern
circulation
r (Intensity and θITCZ)−0.33−0.29−0.10−0.27−0.21−0.10−0.14−0.08−0.34−0.16−0.58−0.65
r (Intensity and θS)0.130.14−0.160.02−0.29−0.210.09−0.110.01−0.23−0.57−0.12
r (Intensity and width)−0.45−0.320.04−0.230.04−0.00−0.15−0.00−0.230.12−0.13−0.52
Northern
circulation
r (Intensity and θITCZ)0.150.19−0.05−0.25−0.56−0.15−0.28−0.23−0.06−0.40−0.190.43
r (Intensity and θN)−0.140.03−0.09−0.42−0.260.14−0.340.07−0.18−0.19−0.03−0.24
r (Intensity and width)−0.19−0.10−0.01−0.020.230.25−0.260.17−0.16−0.050.15−0.50
Table 8. Maximum and minimum latitudes of θITCZ, θITCZ, and θITCZ. Positive values represent the Northern Hemisphere, and negative values represent the Southern Hemisphere.
Table 8. Maximum and minimum latitudes of θITCZ, θITCZ, and θITCZ. Positive values represent the Northern Hemisphere, and negative values represent the Southern Hemisphere.
JANFEBMARAPRMAYJUNJULAUGSEPOCTNOVDEC
ERA5ITCZ
position
Maximum θITCZJANFEBMARAPRMAYJUNJULAUGSEPOCTNOVDEC
Minimum θITCZ−11.1°−10.6°−6.2°4.1°9.3°20.0°22.2°20.4°20.0°10.0°4.9°−5.0°
Southern
circulation
Maximum θS−16.4°−15.4°−11.2°−4.9°4.2°10.8°14.5°14.7°12.2°6.7°−4.3°−14.7°
Minimum θS−32.6°−33.1°−32.3°−29.7°−26.3°−25.3°−26.5°−27.0°−26.8°−27.1°−24.7°−29.3°
Northern
circulation
Maximum θN−37.3°−38.5°−37.8°−35.4°−31.5°−30.6°−29.9°−30.6°−31.0°−32.2°−33.8°−36.6°
Minimum θN30.4°29.9°30.0°27.9°27.9°30.0°40.2°41.6°40.8°38.8°34.6°31.9°
MERRA2ITCZ
position
Maximum θITCZ−11.4°−11.1°−6.8°2.9°9.1°20.0°20.8°19.6°16.7°10.9°5.3°−4.3°
Minimum θITCZ−17.0°−15.8°−11.5°−5.0°3.9°10.7°14.0°14.5°12.6°7.5°−4.6°−15.2°
Southern
circulation
Maximum θS−27.8°−30.6°−30.5°−27.3°−25.2°−24.9°−25.6°−25.6°−25.1°−24.5°−23.0°−25.4°
Minimum θS−35.7°−37.7°−36.5°−34.3°−29.7°−28.8°−28.5°−28.2°−29.7°−30.4°−31.9°−34.4°
Northern
circulation
Maximum θN30.4°29.5°28.1°27.6°26.6°32.1°40.8°42.6°41.7°39.1°34.8°32.2°
Minimum θN26.7°24.7°23.7°23.0°22.0°21.3°23.9°31.1°32.9°29.8°28.6°26.8°
Table 9. Correlation coefficients between the AHT and different θITCZ, as well as the regression slopes between them (in units of °/PW). The significant r is in bold after applying the two-tailed test using Pearson critical values at the level of 5%.
Table 9. Correlation coefficients between the AHT and different θITCZ, as well as the regression slopes between them (in units of °/PW). The significant r is in bold after applying the two-tailed test using Pearson critical values at the level of 5%.
Max PCentroidψΨKE ψ Q K E ψ X K E
Correlation
coefficient
−0.58−0.53−0.51−0.31−0.62−0.35
Slope
(°/PW)
−6.65 ± 1.43−1.46 ± 0.37−3.04 ± 0.79−2.78 ± 1.30−6.82 ± 1.32−2.15 ± 0.90
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Su, Q.; Liu, C.; Zhang, Y.; Qiu, J.; Li, J.; Xue, Y.; Cao, N.; Liao, X.; Yang, K.; Zheng, R.; et al. Consistency of Changes in the Ascending and Descending Positions of the Hadley Circulation Using Different Methods. Atmosphere 2025, 16, 367. https://doi.org/10.3390/atmos16040367

AMA Style

Su Q, Liu C, Zhang Y, Qiu J, Li J, Xue Y, Cao N, Liao X, Yang K, Zheng R, et al. Consistency of Changes in the Ascending and Descending Positions of the Hadley Circulation Using Different Methods. Atmosphere. 2025; 16(4):367. https://doi.org/10.3390/atmos16040367

Chicago/Turabian Style

Su, Qianye, Chunlei Liu, Yu Zhang, Juliao Qiu, Jiandong Li, Yufeng Xue, Ning Cao, Xiaoqing Liao, Ke Yang, Rong Zheng, and et al. 2025. "Consistency of Changes in the Ascending and Descending Positions of the Hadley Circulation Using Different Methods" Atmosphere 16, no. 4: 367. https://doi.org/10.3390/atmos16040367

APA Style

Su, Q., Liu, C., Zhang, Y., Qiu, J., Li, J., Xue, Y., Cao, N., Liao, X., Yang, K., Zheng, R., Liang, Z., Jin, L., Huang, K., Jin, K., & Zhou, N. (2025). Consistency of Changes in the Ascending and Descending Positions of the Hadley Circulation Using Different Methods. Atmosphere, 16(4), 367. https://doi.org/10.3390/atmos16040367

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop