You are on page 1of 15

Atmospheric Research 299 (2024) 107214

Contents lists available at ScienceDirect

Atmospheric Research
journal homepage: www.elsevier.com/locate/atmosres

Combined influence of El Niño, IOD and MJO on the Indian Summer


Monsoon Rainfall: Case Study for the years 1997 and 2015
Satyaban B. Ratna a, *, C.T. Sabeerali a, Tanu Sharma a, b, D.S. Pai c, M. Mohapatra c
a
Climate Research and Services, India Meteorological Department, Pune, India
b
Savitribai Phule Pune University, Maharashtra, India
c
India Meteorological Department, New Delhi, India

A R T I C L E I N F O A B S T R A C T

Keywords: To find out how El Niño, the positive Indian Ocean Dipole (IOD), and the Madden-Julian Oscillation (MJO) affect
El Niño the Indian Summer Monsoon Rainfall (ISMR), this study looks at two different monsoon years: 1997 (normal
Indian Ocean Dipole ISMR) and 2015 (below-normal ISMR). These two years were ideal for examining their combined effects on the
Indian Summer Monsoon Rainfall
performance of ISMR because they both experienced super El Niño conditions over the Pacific Ocean and positive
Madden-Julian Oscillation
Sea surface temperature
IOD over the Indian Ocean. In these two years, spatial diversity associated with sea surface temperature (SST)
Large-scale circulation was found over the Indian Ocean as well as the Pacific Ocean. The analysis of the large-scale circulation showed a
distinct difference: the below-normal monsoon in 2015 was due to weak southwesterly flow toward the Indian
landmass as well as anomalous subsidence compared to 1997. Analysis of rainy and dry days shows that 2015
had a greater (less) number of dry (rainy) days compared to 1997, contributing to a below-normal ISMR in 2015.
The study also looked at the role of MJO in the observed rainfall anomalies over India during these two years,
when there was both El Niño and positive IOD. This was done by looking at the composite of rainfall anomalies
for different MJO phases. In 1997, during the monsoon season, MJO showed prolonged activity over the Indian
Ocean (favorable MJO phases), which contributed to normal ISMR. However, in 2015, MJO was found active for
many days over the western Pacific (unfavorable MJO phases), which contributed to below-normal ISMR.
Overall, this case study highlights the complex interactions between El Niño, IOD, MJO, and monsoon dynamics,
emphasizing the need for a comprehensive study of these factors to understand the behaviour of ISMR.

1. Introduction socio-economics of the region (Gadgil and Gadgil, 2006; Revadekar and
Preethi, 2012). Given its importance, the accurate and improved pre­
The Indian Summer Monsoon is one of the most energetic monsoon diction of ISMR is still the subject of extensive and active research
systems in the world, experienced during the period of June through among scientists the world over.
September (JJAS) each year (Sontakke et al., 1993). Approximately The variability of the ISMR is primarily influenced by large-scale
75–90% of the annual rainfall in most parts of the country is received climate phenomena such as El Niño Southern Oscillations (ENSO), the
during this four-month period (Parthasarathy et al., 1994; Shukla and Indian Ocean Dipole (IOD), and the Madden-Julian Oscillation (MJO).
Huang, 2016). The lives of millions of people across the Indian sub­ The first two are well-known climate drivers of interannual variability in
continent depend on the annual cycle of monsoon rainfall to satisfy the the global climate and the third is the climate driver of global intra­
water requirement for irrigation, drinking, power generation, etc. seasonal variability. Previous studies have extensively studied the
(Rajeevan and Pai, 2007; Saha and Ghosh, 2019). The Indian Summer impact of each of these climate drivers on ISMR (Ju and Slingo, 1995;
Monsoon Rainfall (ISMR) displays strong sub-seasonal and interannual Rasmusson and Carpenter, 1983; Webster and Yang, 1992; Saji et al.,
variability (Krishnamurti and Bhalme, 1976; Shukla, 1987; Goswami 1999; Ashok et al., 2004; Ajayamohan et al., 2008; Sikka and Ratna,
et al., 1999; Chatterjee and Goswami, 2004; Wang et al., 2006; Anna­ 2011; Pai et al., 2011). However, understanding the relative impact of
malai and Slingo, 2001; Sabeerali et al., 2014, 2017). As a result, the these large-scale climate forcings on the ISMR each year is crucial for the
variability of ISMR has a profound impact on the agricultural sector and improved predictions of the ISMR. In addition to the aforementioned

* Corresponding author.
E-mail address: satyaban.ratna@imd.gov.in (S.B. Ratna).

https://doi.org/10.1016/j.atmosres.2023.107214
Received 18 August 2023; Received in revised form 2 December 2023; Accepted 27 December 2023
Available online 2 January 2024
0169-8095/© 2024 Elsevier B.V. All rights reserved.
S.B. Ratna et al. Atmospheric Research 299 (2024) 107214

large-scale forcings, the interannual variability of the ISMR is also The composite rainfall anomaly associated with pure El Niño and
influenced by other climate forcing such as the interannual variability of pure La Niña events (the years when IOD is in neutral phase) reveals a
the tropical Atlantic Ocean, characterized by phenomena such as the nearly stationary anomalous pattern, i.e. most parts of India receive
Atlantic Zonal Mode (AZM) or Atlantic Niño (Kucharski et al., 2007, above (below) normal rainfall during La Niña (El Niño) events (Fig. 1a
2009; Pottapinjara et al., 2016; Sabeerali et al., 2019, 2022) and the and b). A study by Kumar et al. (1999) shows that the ENSO-ISMR
variability of sea surface temperature (SST) in the extratropical Pacific relationship weakened post the 1980s, attributing this change to
and Atlantic Oceans (Chattopadhyay et al., 2015). However, the impact global warming. However, contrary to this hypothesis, the ENSO-ISMR
of extra-tropical SST is mainly experienced during the non-ENSO years. relationship regained statistical significance after 1999 (Yang and
The ENSO is characterized by irregular periodic variations in winds Huang, 2021; Yu et al., 2021). Adding to this discussion, a recent study
and SSTs in the tropical central or eastern Pacific Ocean. It is widely by Mahendra et al. (2023) suggested that changes in the ENSO-ISMR
recognized as a significant driver of global climate variations. During the relationship are influenced by the regional response of ISMR. The
warm phase of ENSO, known as El Niño, the rainfall over the Indian major drivers responsible for the temporal variability of the ENSO-ISMR
landmass tends to be suppressed. Conversely, the cold phase of ENSO, relationship across different regions in India are not completely under­
known as La Niña, often enhances rainfall over the Indian subcontinent. stood till now (Athira et al., 2023).
The anomalous changes in the tropical Walker circulation are respon­ The IOD is another ocean-atmosphere coupled mode of variability
sible for the variations in the ISMR during the contrasting phases of which is characterized by an anomalous SST difference between the
ENSO (Kumar et al., 1999; Ashok et al., 2004). The El Niño events before tropical western and eastern parts of the Indian Ocean (Saji et al., 1999).
1997 (1982 and 1987) badly affected ISMR (14 and 19% deficient The IOD has an impact on ISMR and it also influences the ENSO-ISMR
respectively, based on India Meteorological Department record). How­ relationship (Ashok et al., 2001; Cherchi et al., 2021). The positive
ever, during 1997 El Niño, ISMR was normal, which indicated a and negative phases of the IOD exhibit asymmetric rainfall variability
decreased influence of El Niño on ISMR. This led the research commu­ over India (Behera and Ratnam, 2018). Specifically, a positive IOD leads
nity to re-examine the ENSO-ISMR relationship (Kumar et al., 1999; to enhanced ISMR (Ratna et al., 2021). The composite rainfall anomaly
Krishnamurthy and Goswami, 2000; Sarkar et al., 2004). associated with the pure positive IOD phase and pure negative IOD

Fig. 1. Composite rainfall anomaly (mm/month) for the monsoon season (JJAS) during the period 1951–2022 for a) pure El Niño, b) pure La Niña, c) pure positive
IOD, and d) pure negative IOD events. The dotted marks indicate statistical significance at the 90% level using a two-tail Sudent’s-t-test.

2
S.B. Ratna et al. Atmospheric Research 299 (2024) 107214

phase (the years when ENSO is in neutral phase) shows both positive and results in Section 3. This will be followed by a summary and conclusion
negative rainfall anomalies during the IOD phases in the 1951–2022 in Section 4.
period (Fig. 1c, d).
The Madden-Julian Oscillation (MJO) is a major tropical phenome­ 2. Methodology
non with a typical period of 30–60 days, first discovered by Madden and
Julian in 1971, is one of the leading modes of intraseasonal variability of To identify El Niño, La Niña, and Neutral ENSO years, we used the
ISMR. The MJO plays a significant role in modulating weather and Extended Reconstructed monthly Sea Surface Temperature version 5
climate patterns across the globe, especially in the tropics (Zhang, 2005, (ERSSTv5) data during the June–September (JJAS) season. The ERSST
2013). It is characterized by the eastward propagation of large areas of values are averaged over the Niño3.4 region in the equatorial Pacific
enhanced and suppressed convection along the equator, which exerts a Ocean (5◦ N − 5◦ S and 120◦ W − 170◦ W). The analysis covers the period
significant influence on ISMR (Madden and Julian, 1971; Matthews, from 1981 to 2022. All the seasonal, monthly, and daily anomalies used
2000; Zhang, 2005). Depending on the phase and location of active in this analysis are calculated with respect to 1981–2022 climatology
MJO, it can have both positive and negative impacts on the ISMR on the except for Fig. 1; This figure incorporates data from the period
intra-seasonal scale (Pai et al., 2011; Saith and Slingo, 2006; Bhatla 1951–2022, in order to encompass a greater number of pure ENSO and
et al., 2017; Dey et al., 2022). The MJO influences ISMR by altering the pure IOD events in the composite. We identified pure El Niño (La Niña)
distribution of the large-scale convection, circulation, and moisture years as those years in which ENSO is in a positive (negative) phase but
transport patterns over the Indo-Pacific region throughout the season IOD is neutral. Similarly, we identified pure positive (negative) IOD
(Singh et al., 1992). years as those years in which the IOD phase is positive (negative) but the
As every El Niño event is different due to its temporal and spatial ENSO is in neutral phase over the Pacific Ocean. The linear trend from
diversity (Capotondi et al., 2015) its influence on the monsoon depends the SST anomaly time series is removed before employing the analysis.
on regional details of the atmospheric response to the SST forcing. Slingo The El Niño (La Niña) years are defined as those in which the detrended
and Annamalai, (1999) discussed that in the year 1997, despite the JJAS SST anomalies are greater (smaller) than 0.5 ◦ C (− 0.5 ◦ C) over the
strong El Niño condition, ISMR was normal because of active local Niño3.4 region. All remaining years, where the detrended JJAS SST
north-south Hadley circulation during the established phase (July and anomalies fall between - 0.5 ◦ C to 0.5 ◦ C, are considered Neutral ENSO
August) of monsoon, the absence of prolonged break spells, and high years. To find the phase of the Indian Ocean Dipole (IOD), we calculate
intensity of monsoon depressions during August. In contrast, during the Dipole Mode Index (DMI). The DMI is derived from the difference in
2015 El Niño the ISMR was below-normal. The meridional SST gradient detrended SST anomalies during the JJAS season between two regions:
across the central-eastern Pacific modulated the ISMR response to El western equatorial Indian Ocean (WEIO: 50◦ E− 70◦ E and 100S − 100N)
Niño in 2015, through sub-tropical convective activities and the changes and eastern equatorial Indian Ocean (EEIO: 90◦ E− 110◦ E and 100S −
in the east–west Walker circulation over Indo-Pacific region (Mujumdar 00N). Positive (Negative) IOD years are identified when the DMI, ex­
et al., 2017). Previous studies show that the modes of interannual ceeds 0.245 ◦ C (− 0.245 ◦ C), which is the value for one standard devi­
variability, including ENSO (DeMott et al., 2018), IOD (Wilson et al., ation of DMI. All remaining years, where the DMI values fall within the
2013), and the modes of intra-seasonal variability, including Quasi- range of one standard deviation, are considered Neutral IOD years. We
Biennial Oscillation (QBO, Nishimoto and Yoden, 2017), influence the have calculated the percentage rainfall anomaly for all India and iden­
intensity and propagation of MJO. On the other hand, some studies show tified the year as above-normal (below-normal) if the percentage rainfall
that the MJO influences dominant modes of climate variability, anomaly is greater (less) than 10% (− 10%). Rest of the years are iden­
including the ENSO (Hendon et al., 2007; Kessler and Kleeman, 2000; tified as normal rainfall years. To check the MJO variability we used the
McPhaden, 1999), and IOD (Rao and Yamagata, 2004). It is important to MJO phase space diagram (Wheeler and Hendon, 2004) which was
understand the indirect ocean feedback on temporal scales other than constructed using the Real-time Multivariate MJO series 1 (RMM1) and
intraseasonal, which interacts with MJO through the intermediate pro­ 2 (RMM2). These series were generated from the pair of Principle
cess (Jiang et al., 2020). We did not find any study analyzing the role of Component (PC) time series using the combined fields of near-
MJO on ISMR in the presence of El Niño and positive IOD. The MJO can equatorially averaged 850 hPa, 200 hPa zonal wind, and outgoing
modulate the impact of El Niño and IOD on ISMR during the season, longwave radiation (OLR). During the JJAS seasons of 1997 and 2015,
depending upon its location and intensity. Hence, the performance of the number of dry, rainy, and heavy rainfall days are calculated on each
ISMR can, to some extent, be influenced by whether the MJO predom­ grid point over India. A dry day is defined when the received daily
inantly resided in its favorable/unfavorable phase for most parts of the rainfall is 0 mm, indicating no rainfall occurred on that day. A rainy day
season, based on its year-to-year variability. is determined when the daily rainfall value exceeds 2.5 mm. Similarly, a
This study attempts to understand the combined impact of super El heavy rainfall day is defined when the daily rainfall value exceeds 64.5
Niño, positive IOD and MJO on ISMR, with a specific focus on the case mm as per the criteria followed by the India Meteorological Department
studies of 1997 and 2015. ‘Super El Niño’ (also called extreme El Niño) (IMD).
events are selected based on the intensity of the SST anomaly in recent
decades near the central-eastern equatorial Pacific Ocean (Marjani et al., 3. Results
2019; Wang and Wang, 2021). The influence of super El Niño on global
weather and climate is significantly different compared to the ordinary 3.1. SST pattern over the Pacific and Indian Ocean
El Niño (Bi et al., 2017; Qian and Guan, 2018). These two years expe­
rienced super El Niño as well as positive IOD conditions. Therefore, this The time series of Niño3.4 and IOD indices (DMI) for the JJAS season
study gives us the opportunity to understand the ISMR response to the from 1981 to 2022 are given in Fig. 2a. It is clear that the years 1997 and
combined impact of ENSO and IOD, along with the MJO. By utilizing 2015 were characterized by the super El Niño events. The El Niño event
historical observational data, we seek to elucidate the complex in­ for which the SST anomaly over the Niño3.4 region exceeds 1.5 ◦ C in the
teractions and feedback mechanisms between these three climate JJAS season, is characterized as ‘Super El Niño’ (Marjani et al., 2019;
drivers and the ISMR response. The findings of this case study of two Wang and Wang, 2021). During the JJAS season of 1997 and 2015, the
years: 1997 and 2015 will contribute to a better understanding of the Niño3.4 region was the warmest on record placing these two years as
complex feedback mechanisms among different large-scale climate extreme El Niño events since 1950 (http://www.wmo.int; http://www.
forcings (ENSO, IOD, and MJO) and their combined influence on ISMR. bom.gov.au/climate/enso; http://www.cpc.ncep.noaa.gov). Interest­
In the subsequent sections, we will provide a detailed discussion of ingly, positive IOD conditions were observed during these two years
the methodology in Section 2, followed by the presentation of the main over the Indian Ocean. We found that during the JJAS season, the DMI in

3
S.B. Ratna et al. Atmospheric Research 299 (2024) 107214

Fig. 2. Time series of a) Niño3.4 and DMI Index b) percentage rainfall anomaly for the June–September (JJAS) season from 1981 to 2022. Positive and negative
phases of ENSO and IOD are represented for each year.

1997 measured at 0.246 ◦ C, indicating a weaker value compared to 2015 the subtropical north-eastern Pacific to the equatorial CP. Fig. 3c, sug­
(0.354 ◦ C). It is important to note a previous study’s findings indicating gests the presence of an eastern Pacific (EP), also known as canonical
that in 2015, the DMI was weaker than the mean DMI in 1997 and 1982 type El Niño in 1997 and a combination of the EP and El Niño Modoki-II
as depicted in Fig. 1 of Liu et al., 2017. However, it is crucial to (EM-II) type El Niño in 2015. Our finding is similar to the previous study
emphasize that these results specifically pertain to the months of by Paek et al. (2017) which showed that during the 2015 El Niño, the
October and November, not the JJAS season. It is clearly seen from the center of positive SST anomalies was westward with mixed character­
time series that 1997 and 2015 are the only two years where super El istics of the CP El Niño and EP El Niño.
Niño and positive IOD was observed. Time series for the percentage There are noticeable variations in SST anomalies over the eastern
rainfall anomaly for all India during the JJAS season shows that the year and western equatorial Indian Oceans. The association of ISMR with the
1997 experienced normal ISMR while 2015 was below-normal monsoon equatorial Indian Ocean is as important as its association with the
year in the presence of super El Niño and positive IOD (Fig. 2b). equatorial Pacific Ocean (Gadgil et al., 2003). In 2015, the SST anomaly
To gain further insight, Fig. 3. provides the spatial pattern of the SST over the EEIO was warmer compared to 1997 (Fig. 3a and b). A warm
anomaly for the years 1997, 2015 and the difference (1997–2015). This EEIO drives easterly winds in the Indo–Pacific sector as a Gill response
figure shows the strength and spatial extent of El Niño in the Pacific and resultant circulation opposes the monsoon low-level circulation
Ocean. The comparison of SST anomaly between 1997 and 2015 in­ hence weakening the ISMR in 2015 (Goswami, 2023). It was also seen
dicates that, in 1997, the SST anomaly was warmer over the equatorial that during 1997, there were negative OLR anomalies over the WEIO
eastern Pacific and colder in the subtropical north-eastern Pacific and (Fig. S1b and S1f), which indicates enhanced convection over WEIO;
equatorial central Pacific Ocean compared to 2015 (Fig. 3c). Wang and favors ISMR (Gadgil et al., 2004). Also, tropical climate modes across the
Wang (2013), classified central Pacific (CP) El Niño in two different globe influence the pressure gradient over the Indian Ocean in zonal as
types. The El Nino Modoki (EM-I), which shows warm SST anomalies well as meridional direction, leading to the modulation in ISMR
over the equatorial CP, and El Niño Modoki II (EM-II), which display an (Chakraborty and Singhai, 2021). It was seen that there was a negative
asymmetric distribution with the warm SST anomalies extending from sea level pressure (SLP) anomaly near the north Indian region and a

4
S.B. Ratna et al. Atmospheric Research 299 (2024) 107214

Fig. 3. SST anomaly (color contour, in ◦ C) and wind vector anomaly (m s− 1) at 850 hPa during June-Sept Season for the years a) 1997, b) 2015 and c) difference
between 1997 and 2015.

positive SLP over the Indian Ocean, creating a stronger north-south/ equatorial flow and strengthened the moisture-containing low-level
meridional gradient in 1997 (Fig. S1a) compared to 2015 (Fig. S1c). south-westerly winds which contributed to higher rainfall in 1997
The stronger north-south pressure gradient enhanced the cross- compared to 2015.

Fig. 4. Monthly SST anomaly over Niño3.4, Western Equatorial Indian Ocean (WEIO), Eastern Equatorial Indian Ocean (EEIO), and Dipole Mode Index (DMI: SST
anomaly difference WEIO-EEIO) for the years a) 1997 and b) 2015.

5
S.B. Ratna et al. Atmospheric Research 299 (2024) 107214

The monthly evolution of the Niño3.4 index indicates a similar We also verified the daily rainfall anomaly series for all India and the
evolution of El Niño in both 1997 and 2015 (Fig. 4). In both years, El core monsoon zone (18◦ N to 28◦ N and 65◦ E to 88◦ E; Rajeevan et al.,
Niño started around April–May and gradually strengthened as the 2010) for the years 1997 and 2015 and compared them with the
monsoon season progressed. Similarly, from June to September the climatology (Fig. 6). It is evident that there were a greater number of
monthly evolution of IOD was quite similar in both 1997 and 2015, with break spells (normalized rainfall anomaly over core monsoon zone, less
a slight difference in the beginning of IOD between these two years. In than − 1 continuously for three or more days) during the 2015 monsoon
1997, the IOD evolved slowly around June and maintained its strength season compared to 1997. The increased occurrence of break spells in
during the monsoon season of that year (Fig. 4a). However, in 2015, the 2015 contributed to reduced rainfall in 2015 compared to 1997.
IOD peaked in July and more or less maintained its strength throughout As there are significant differences in the spatial distribution of
the monsoon season. The main difference was observed after the seasonal rainfall anomalies between the years 1997 and 2015 over India,
monsoon season (October, November and December). After monsoon we also explored the total number of rainy days, dry days, and heavy
season, the DMI signal peaked in 1997 whereas DMI weakened and rainfall days that contributed to seasonal mean rainfall (Fig. 7). In 2015,
dropped to almost neutral IOD condition in 2015 (Fig. 4b). The evolu­ the total number of dry days was higher and more widespread compared
tion of El Niño and positive IOD events during the JJAS season in the to 1997 (Fig. 7a, b, and c), especially in central and northern India,
years 1997 and 2015 and the similarity in their intensity during these representing the differences in the seasonal rainfall anomaly.
two years give a unique opportunity to study the combined impact of El Conversely, in 1997, the dry days were more frequent in the northern
Niño and IOD on the ISMR characteristics. By analyzing these specific parts of south peninsular India compared to 2015. However, the number
years in detail, we can gain more insights into how the MJO modulates of rainy days was higher in 1997 compared to 2015, particularly in
the combined and concurrent impact of El Niño and positive IOD on the northeast India, central India, and extreme north India (Fig. 7d, e, and f).
ISMR. Moreover, the spatial pattern of heavy rainfall days displays a minimal
difference between 1997 and 2015, except for a higher number of heavy
3.2. Observed ISMR features of 1997 and 2015 rainfall days in 1997 along the west coast of India (Fig. 7g, h, and i). The
distribution of the number of dry days and rainy days in these two years
The spatial pattern of rainfall during the JJAS season for these two contributed to the seasonal total rainfall. It was clearly observed that in
years exhibited significant variations across different parts of India 2015, a higher number of dry days and a lower number of rainy days
(Fig. 5). In 1997, despite the presence of the super El Niño, the country contributed to below-normal rainfall compared to 1997. Also, the large
received normal rainfall in the monsoon season (Fig. 2b). However, the number of heavy rainfall days over the Western Ghat regions in 1997
spatial distribution of rainfall in 1997 varied significantly, with dry and might also have contributed to the higher seasonal total rainfall in 1997
wet anomalies over different parts of the country. Most parts of penin­ compared to 2015 (Fig. 7g, h, and i).
sular India and the foothills of the Himalayas faced below-normal In short, despite the presence of super El Niño and positive IOD
rainfall in the 1997 monsoon season (Fig. 5a). Conversely, most parts conditions in 1997 and 2015, the rainfall over India was normal in 1997,
of east-central India, west-central India, the west coast of India, and while it was below-normal in 2015. This contrast in the spatial distri­
extreme north India received above-normal rainfall in the 1997 bution of rainfall patterns, which varied significantly between 1997 and
monsoon season (Fig. 5a). In contrast, in 2015, despite the presence of 2015, highlights the complex nature of monsoon dynamics and other
positive IOD conditions, the rainfall during the monsoon season was factors influencing the ISMR.
below-normal in most parts of the country (Fig. 5f). The rainfall was
above-normal in some areas in northwest India and in a few pockets of 3.3. Large-scale circulation
northeast India and adjoining east India (Fig. 5f) and the country as a
whole experienced below-normal rainfall in 2015 (Fig. 2b). The analysis The general hypothesis suggests that the strength of ISMR is influ­
above also indicates that the impact on northwest India is almost the enced by the modulation of the east-west oriented Walker circulation
same for both years (1997 and 2015), but the major difference seems to during ENSO years. During El Niño years, there is a subsidence over the
be over central India, particularly in the eastern part of central India. west Pacific and southeast Asia, including India, while an anomalous
The monthly distribution of rainfall in 1997 shows above-normal ascending motion occurs over the east/central Pacific due to warm SST
rainfall in the north and northwest during June and August months anomalies, and vice versa during La Niña years (Kumar et al., 1999; Pai
(Fig. 5b and Fig. 5d), whereas the same regions received below-normal et al., 2014; Cherchi et al., 2021). The changes in Walker circulation
rainfall during July and September (Fig. 5c and Fig. 5e). Also, the central during the El Niño years of 1997 and 2015 are clearly evident in Figs. 8
region received below-normal rainfall in June and September (Fig. 5b and 9. In both 1997 and 2015, there was a significant longitudinal shift
and e), but normal rainfall in July and August (Fig. 5c and d). The South in the Walker cell, which corresponded to the eastward movement of
Peninsula of India, which received below-normal rainfall from June to warm SST anomalies. It is characterized by an anomalous upward mo­
August (Fig.5b, c, and d), made a recovery and had above-normal tion over the eastern and central Pacific accompanied by a compensating
rainfall in September (Fig. 5e). The monthly rainfall distribution in anomalous downward motion over the eastern hemisphere, especially
June 2015 (Fig. 5g) shows that most parts of the country received above- over the west Pacific and east Indian Ocean regions (Fig. 8). However, in
normal rainfall, resulting in above average rainfall for India as a whole 2015, the upward motion over the eastern and central Pacific, as well as
during that particular month. However, the all-India average rainfall the downward motion over the western Pacific and east Indian Ocean
was below normal in July, August, and September 2015. In July 2015, regions, were considerably stronger compared to 1997. In 2015, an
the rainfall was deficient in peninsular India, the foothills of the anomalous downward motion was witnessed along the Indian longi­
Himalayas, and northeast India (Fig. 5h). However, central India and tudes, which was stronger than in 1997. This pattern is further
northwest India experienced deficient rainfall in August and September confirmed by the vertical velocity anomaly plot at 500 hPa (Fig. 9). In
2015 (Fig. 5i and j). This pattern of deficient rainfall deviated from the 1997, a weak subsidence was noted over central and northwest India at
expected pattern of rainfall during El Niño years. Overall, the rainfall 500 hPa, while the rest of the region experienced a weak upward mo­
was better during the active monsoon months (July and August) in 1997 tion, as evident in 500 hPa omega anomaly (Fig. 9a). In contrast, in
whereas the rainfall was better during the onset and withdrawal months 2015, the omega at 500 hPa showed anomalously strong subsidence
(June and September) in 2015. Climatologically, India receives more over most of the Indian subcontinent, excluding northeast India, when
rainfall in July and August, compared to June and September. So, the compared to 1997 (Fig. 9). This higher subsidence in 2015 over the
higher rainfall received in India during the active monsoon months in Indian landmass coincided with lesser rainfall over India in 2015
1997 contributed to more seasonal rainfall compared to 2015. compared to 1997 (Fig. 5a and f).

6
S.B. Ratna et al. Atmospheric Research 299 (2024) 107214

Fig. 5. Rainfall anomaly (mm/month) for the 1997 (a) JJAS season, (b) June, (c) July, (d) August, and (e) September. Equivalent anomalies for the year 2015 are
represented in (f) to (j), respectively. Panels (k) to (o) depict the difference between 1997 and 2015 in terms of rainfall anomalies, corresponding to the same months
as in panels (a) to (e).

7
S.B. Ratna et al. Atmospheric Research 299 (2024) 107214

Fig. 6. Daily time series for all India (a and c) and core monsoon zone (b and d) for the monsoon season (1 June to 30 September rainfall, in mm day − 1) for the years
1997 and 2015.

Fig. 10 shows the 850 hPa velocity potential anomalies during 1997 positive IOD events. So, to further investigate the factors contributing to
and 2015. Consistent with the modulation of Walker cell anomalies the rainfall variability during these two years, the role of the MJO was
during these years, the 850 hPa velocity potential shows anomalous explored.
convergence over the east and central Pacific, coupled with anomalous The daily anomalies of zonal wind at 850 hPa show the active
divergent motion over the west Pacific and southeast Asian region. It is presence of westerly anomalies over the Pacific Ocean during August
seen that the extent of the divergent region over the Indian landmass and September 1997, in contrast, a westerly anomaly was observed in
was less pronounced in 1997 compared to 2015. Interestingly, in 2015, a July 2015 but gradually weakened thereafter (Fig. S2). The power
significant portion of India showed an anomalous divergence at lower spectrum analysis done by Li et al. (2023) for the years 1997 and 2015
levels compared to 1997 (Fig. 10). Although, there was an El Niño revealed considerable difference in the MJO periodicity. Their analysis
impact for both years, in 1997, anomalous westerlies (Fig. 9a and highlighted strong peaks of MJO at 45 days and 90 days periods from
Fig. S1a) entering the Indian landmass, resulting in more moisture February to September 1997, while in 2015, peaks were observed at 40
convergence into the region, and contributed to a higher number of days and 60 days periods.
heavy rainfall days over the western Ghats (Fig. 7g-i) and also to the The MJO phase diagram (Wheeler and Hendon, 2004) for the period
total seasonal rainfall (Fig. 5a, f and k). In comparison, in 2015, in the of 1 June to 30 September in 1997 and 2015 is analyzed (Fig. 11). In
absence of such low-level anomalous westerlies (Fig. 9b and Fig. S1b), 1997, the MJO was in Phases 7 and 8 during June and then transitioned
there were fewer heavy rainfall days over the western Ghats as well as to Phases 1 and 2. In July, it progressed from phase 3 to phase 6. Toward
less seasonal total rainfall. The role of low-level westerlies in these two the end of July, the MJO weakened and maintained a weak intensity
seasons may be due to the differences in the strength and patterns of the until the end of the monsoon season (i.e., for the full months of August
SST anomaly over the Arabian Sea associated with the positive IOD and September). It is worth noting that MJO phases 3–6 are known to be
(Fig. 3a and b). This is because, although both the years are associated favorable phases for ISMR, as these phases typically bring a good
with positive IOD, the SST anomaly over the EEIO is negative (positive) amount of rainfall to India during the monsoon period (Pai et al., 2011).
in 1997(2015) as shown in Fig. 3a and b. A study by Yadav and Roxy Conversely, phases 7, 8, 1, and 2 are known to be unfavorable phases for
(2019) discusses that warm SST over EEIO has a negative correlation the ISMR. The strong MJO signal is defined, when the RMM index is
with the rainfall over the northern parts of India. equal or >1. As per the RMM phase space diagram, the MJO was in
These observations bring out the distinct differences in the Walker favorable phases (Phases 3 to 6) for 27 days during the monsoon season
circulation and atmospheric dynamics between the El Niño events of of 1997 (Table 1), which coincided with good rainfall over India
1997 and 2015. The stronger upward and downward motions over the (although there were regions that also received negative rainfall
western hemisphere and eastern hemisphere in 2015 suggest a stronger anomalies), as shown in Figs. 11 and 12. In 1997, it was seen that the
and more localized impact on weather patterns, which may have MJO was in the unfavorable phases (Phases 7, 8,1 and 2) for 30 days
contributed to the strong drought-like conditions over India in 2015. (Table 1), coincident with below-normal rainfall over most parts of India
(Fig. 12).
In June 2015, the MJO travelled from Phase 1 to 6, and then in July,
3.4. MJO characteristics it was in Phases 7, 8, 1, and 2. However, during August and September,
the MJO mostly remained in a weak phase. Analysis of the MJO phase
Despite the different temporal scales of ENSO, IOD and MJO, they diagram shows that the MJO was in favorable phases for 27 days during
exhibit close interrelationship and influence weather patterns over the monsoon season of 2015, while there were 52 days where the MJO
different parts of the globe. Previous studies show that the MJO in­ was in unfavorable phases (Table 1). A previous study by Chen et al.
fluences dominant modes of climate variability, including the ENSO (2016) shows that during the CP El Niño, the horizontal as well as
(Hendon et al., 2007; Kessler and Kleeman, 2000; McPhaden, 1999), and vertical moisture advection over the CP is stronger than during the EP El
IOD (Rao and Yamagata, 2004). On the other hand, the year-to-year Niño, resulting in the further eastward propagation of the MJO. During
variability of MJO is found to be influenced by ENSO (Hendon et al., 2015, the extratropical circulations originated from northeastern Pa­
2007; Marshall et al., 2016) through the Indonesian Throughflow (ITF) cific, reached the equator and enhanced the activities of MJO over the
and air-sea interactions, connecting the tropical Pacific and Indian Pacific (Li et al., 2023), which are unfavorable for ISMR. India received
Oceans (Gordon, 2005). Recently, many researchers have focused on the good rainfall activity during the favorable phases (Phase 3, 4, and 5), as
relationship between MJO and two types of El Niño: central Pacific (CP) shown in Fig. 12. However, the large number of unfavorable days during
and eastern Pacific (EP) El Niño (Feng et al., 2015; Yuan et al., 2015). A phases 7, 8, 2 and 1 (Fig. 11) contributed to a large deficiency in sea­
significant variation in the spatial distribution of rainfall was observed sonal rainfall (Fig. 5f).
between the years 1997 and 2015 against the backdrop of El Niño and

8
S.B. Ratna et al. Atmospheric Research 299 (2024) 107214

Fig. 7. The number of dry days (daily rainfall = 0 mm) during the JJAS season of a) 1997 b) 2015 and c) the difference between 1997 and 2015. d), e), f) and g), h),
i) are the same as a), b), and c) but for the number of rainy days (daily rainfall >2.4 mm) and the number of heavy rainfall days (daily rainfall >64.5 mm)
respectively.

By comparing the MJO-associated rainfall in 1997 and 2015, it was 4. Summary and conclusions
observed that India received more rainfall in 2015 during the favorable
phases compared to 1997. However, the negative rainfall anomaly was It is well known that ENSO has a negative relationship with ISMR,
stronger in 2015 during unfavorable phases compared to 1997. This whereas there is an asymmetric relationship between IOD and rainfall
indicates that although there was the same number of MJO favorable conditions in different parts of India. However, it is complex to under­
days in 1997 and 2015, the large number of unfavorable days in 2015 stand the rainfall variability over India when multiple climate modes
led to a significant deficiency in rainfall over India, contributing to such as ENSO and IOD co-occurred during the same season.
below-normal seasonal rainfall over India. Many studies discuss the individual impacts of the ENSO, IOD, and
MJO on ISMR but the complex interplay among them is not fully un­
derstood. While some studies show little relationship between the MJO

9
S.B. Ratna et al. Atmospheric Research 299 (2024) 107214

and ENSO (Slingo and Annamalai, 1999; Son et al., 2016), others show
that MJO amplitude and propagation can modulate ENSO-like large-
scale mode (DeMott et al., 2018; Gonzalez and Jiang, 2019). The MJO-
ENSO relationship is further complicated by the diversity of ENSO
events (Gushchina and Dewitte, 2012; Feng et al., 2015). Moreover,
there are not many studies that has discussed the role of MJO on the
ISMR in the presence of super El Niño and positive IOD in the same
season. In this study, we investigate the combined influence of super El
Niño and positive IOD that drive the interannual variability of ISMR
along with the eastward moving MJO that drives the intra-seasonal
variability of the ISMR. Our goal in this study is to examine the
distinct performance of ISMR during 1997 and 2015 which coincides
with super El Niño and positive IOD. Using the observational data of the
recent 40 years, we found both 1997 and 2015, to be the strongest El
Fig. 8. Walker circulation anomalies over the Indo-Pacific region during a) Niño events in history, accompanied by positive IOD conditions, making
1997 and b) 2015 JJAS monsoon season. Color contour shows vertical wind them ideal for studying the combined effect of the three climate drivers
velocity (w ✕ 103 m s− 1, positive (negative) values represent subsidence (rising (ENSO, IOD, and MJO) on ISMR.
motion)). Vectors are plotted using zonal (u) and vertical (w) wind, from 1000 The observed analysis revealed that despite the presence of the same
hPa to 100 hPa. Values averaged over the latitude range of 30 0N-5 0S. climate drivers (El Niño and positive IOD) in 1997 and 2015, 1997 was a
normal monsoon year, whereas 2015 was a below-normal monsoon year

Fig. 9. Vertical wind velocity anomaly (102 ✕ Pa s− 1, positive (negative) values represent subsidence (rising motion)) at 500 hPa and horizontal wind vector
anomaly (m s− 1) at 850 hPa for the years a) 1997 and b) 2015 monsoon season.

Fig. 10. Velocity Potential Anomaly Composite (shaded, 106 ✕ m2 s− 1) and divergent wind vector (m s− 1) at 850 hPa for the years a) 1997 and b) 2015 during the
JJAS monsoon season.

10
S.B. Ratna et al. Atmospheric Research 299 (2024) 107214

Fig. 11. MJO phase space plot for the years a) 1997 and b) 2015.

Table 1
The number of days with favorable and unfavorable phases during the JJAS seasons of 1997 and 2015.
Number of Days 1997 2015

Jun Jul Aug Sep JJAS Jun Jul Aug Sep JJAS

Favorable MJO Phase 2 23 – 2 27 20 – – 7 27


(3, 4, 5, 6)
Unfavorable MJO Phase (1, 2, 7, 8) 26 – 1 3 30 10 30 7 5 52

for the Indian region. In 1997, despite the strongest El Niño the Indian Apart from this, the negative OLR anomaly over the WEIO and the
subcontinent received normal rainfall during the monsoon season, but stronger north-south pressure gradient enhances the strength of the
with significant spatial variations. In contrast, 2015 witnessed below- moisture-containing low-level south-westerly winds which play a
normal rainfall across most parts of the country, with only a few areas crucial role in ISMR, contributing to a normal monsoon year in 1997.
in the northwest and pockets of northeast and east India receiving However, during 2015, the north-south pressure gradient was weak and
above-normal rainfall. The results showed that 2015 had a higher the EEIO was relatively warmer, which led to weaker divergence and
number of dry days, fewer rainy days, and more days with unfavorable weaker low-level westerly flow (e.g. Mujumdar et al., 2017) toward the
MJO phases compared to 1997, which led to below-normal monsoon Indian landmass, and hence resulted in below-normal monsoon rainfall.
rainfall in India. Such results encourage future studies to analyze mul­ Although, there was El Niño and positive IOD present, apart from the
tiscale (from the diurnal to ENSO cycles) interactions (e.g. Yoneyama SST variations, the MJO played a crucial role in both years. In 1997 MJO
and Zhang, 2020). activity was prominent near the Indian Ocean during the monsoon
A schematic highlighting the features and mechanism regulating season, which is favorable for ISMR. In contrast to this, during 2015
ISMR during 1997 and 2015, is represented in Fig. 13. The distribution MJO showed prolonged activity near the western Pacific Ocean which is
of SST anomalies during these years revealed the spatial variation of El considered an unfavorable phase for ISMR. Although 2015 was not pure
Niño. The year 1997 was EP El Niño whereas 2015 was a combination of CP El Niño but a combination of EP and CP El Niño, our findings for the
the EP type (canonical) El Niño and CP type or El Nino Modoki. The case study of these two years are consistent with a prior study conducted
occurrence of the CP type El Niño in 2015 was also highlighted in pre­ by Wang et al. (2018) which shows that during EP (CP) El Niño years the
vious studies (Mujumdar et al., 2017; Marshall et al., 2016; Paek et al., MJO shows weaker (stronger) activity over the western Pacific.
2017). We also observed significant variations in the SST anomalies over Despite the attempt to understand the combined influence of the
the EEIO. In 1997, EEIO experienced a negative SST anomaly. The study various climate drivers, it is important to acknowledge some limitations
by Behera and Ratnam (2018) and Ratna et al. (2021) discuss that the of our study. Firstly, this study focuses for the year 1997 and 2015 as
low-level divergence associated with the negative SST anomaly over the both are the two only super El Niño years with a Niño3.4 index of
EEIO could enhance the low-level westerlies and southerlies from the >1.5 ◦ C and all the results are based on the comparative analysis of these
EEIO toward India. Despite the positive IOD and super El Niño condi­ two years. Secondly, 1982 was also an El Niño and positive IOD year,
tions, the spatial pattern of rainfall during the summer monsoon seasons which we did not include in our analysis because El Niño has a weaker
(JJAS) of 1997 and 2015 showed notable variations between these two intensity (with a Niño3.4 index of <1 ◦ C). Also, the MJO phase diagram
years. Due to the zonal pressure gradient, the east-west-oriented Walker data is available only from 1974 which gives limited opportunities to
cell associated with El Niño and IOD conditions in the Indo-Pacific include a larger number of case studies together with El Niño and pos­
oceans contributed to the strong anomalous subsidence over most itive IOD. Lastly, our results are based on the observational and rean­
parts of the Indian subcontinent in 2015. alysis datasets. So, for a deeper understanding of the phenomenon and

11
S.B. Ratna et al. Atmospheric Research 299 (2024) 107214

Fig. 12. Daily rainfall anomaly composites (mm) for 8 phases of MJO during the JJAS seasons of a) 1997 and b) 2015.

for identifying the relative contributions of individual factor, some and their complex interplay to better understand the monsoon rainfall
modeling studies are required, which is beyond the scope of this study. variability over India.
However, Marshall et al. (2016) conducted a model study revealing that
the anomalous SST in the central Pacific during the 2015 El Niño Funding statement
stimulated intensified MJO activity in the western Pacific. This coincides
with our observed analysis that MJO was active over western Pacific for This research did not receive any specific grant from funding
large number of days during monsoon season of 2015, which was un­ agencies in the public, commercial, or not-for-profit sectors.
favorable for ISMR.
In summary, the years 1997 and 2015, with super El Niño and pos­ CRediT authorship contribution statement
itive IOD conditions, displayed significant differences in the spatial
distribution of rainfall over India. It was observed that the phases of Satyaban B. Ratna: Conceptualization, Formal analysis, Investiga­
MJO have a crucial role in shaping the rainfall pattern over India during tion, Methodology, Resources, Writing – original draft, Writing – review
such El Niño and positive IOD years. These findings highlight the com­ & editing. C.T. Sabeerali: Writing – original draft, Writing – review &
plex nature of monsoon dynamics and the influence of various factors, editing. Tanu Sharma: Software, Visualization, Writing – original draft,
including El Niño, positive IOD, MJO, and atmospheric circulation Writing – review & editing. D.S. Pai: Writing – review & editing. M.
patterns. The contrasting rainfall patterns between 1997 and 2015 Mohapatra: Writing – review & editing.
emphasize the need for a comprehensive understanding of these factors

12
S.B. Ratna et al. Atmospheric Research 299 (2024) 107214

Fig. 13. Schematic diagram representing the factors contributing to normal ISMR during 1997 and below-normal ISMR during 2015. The oceanic and atmospheric
conditions over the Indo-Pacific region are shown. Over the ocean, the red (blue) shading represents a positive (negative) SST anomaly. Thickness of red (upward
motion) and blue (subsidence) arrows represent the strength of Walker circulation. The locations of prolonged MJO activity and strength of southwesterly flow are
just illustrations and not meant to be precise. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of
this article.)

Declaration of competing interest Ashok, K., Guan, Z., Yamagata, T., 2001. Impact of the indian ocean dipole on the
relationship between the Indian Monsoon Rainfall and ENSO. Geophys. Res. Lett. 28
(23), 4499–4502. https://doi.org/10.1029/2001GL013294.
The authors declare no competing interest. Ashok, K., Guan, Z., Saji, N.H., Yamagata, T., 2004. Individual and combined influences
of ENSO and the Indian Ocean dipole on the Indian summer monsoon. J. Clim. 17
Data availability (16), 3141–3155.
Athira, K.S., Roxy, M.K., Dasgupta, P., Saranya, J.S., Singh, V.K., Attada, R., 2023.
Regional and temporal variability of Indian summer monsoon rainfall in relation to
We used the National Oceanic and Atmospheric Administration El Niño southern oscillation. Sci. Rep. 13, 12643. https://doi.org/10.1038/s41598-
(NOAA) Extended Reconstructed monthly Sea Surface Temperature 023-38730-5.
Behera, S.K., Ratnam, J.V., 2018. Quasi-asymmetric response of the Indian summer
version 5 (ERSSTv5) dataset (Huang et al., 2017) for the period monsoon rainfall to opposite phases of the IOD. Sci. Rep. 8, 123. https://doi.org/
1981–2022, which is accessible on their website at https://psl.noaa.gov. 10.1038/s41598-017-18396-6.
We also used IMD high spatial resolution (0.25◦ x 0.25◦ ) gridded rainfall Bhatla, R., Singh, M., Pattanaik, D.R., 2017. Impact of Madden-Julian oscillation on
onset of summer monsoon over India. Theor. Appl. Climatol. 128, 381–391.
data for the period 1981–2022 (Pai et al., 2014) available at Bi, B., Zhang, X., Dai, K., 2017. Characteristics of 2016 severe convective weather and
https://www.imdpune.gov.in/cmpg/Griddata/Rainfall_25_NetCDF. extreme rainfalls under the background of super El Niño. Chin. Sci. Bull. 62,
html. For zonal (u), meridional (v), and vertical wind velocity (omega), 928–937. https://doi.org/10.1360/n972016-01136.
Capotondi, A., Wittenberg, A.T., Newman, M., Lorenzo, E.D., Yu, J.-Y., Braconnot, P.,
Sea Level Pressure (SLP) and Outgoing Long Radiation (OLR), we used et al., 2015. Understanding ENSO diversity. Bull. Am. Meteorol. Soc. 96 (6),
monthly mean Reanalysis 1 data (Kalnay et al., 1996) for the period 921–938. https://doi.org/10.1175/BAMS-D-13-00117.1.
1981–2022 provided by the National Centers for Environmental Chakraborty, A., Singhai, P., 2021. Asymmetric response of the Indian summer monsoon
to positive and negative phases of major tropical climate patterns. Sci. Rep. 11,
Prediction-National Center for Atmospheric Research (NCEP-NCAR) and
22561. https://doi.org/10.1038/s41598-021-01758-6.
is available at https://psl.noaa.gov/data/gridded/data.ncep.reanalysis. Chatterjee, P., Goswami, B.N., 2004. Structure, genesis and scale selection of the tropical
html. The Madden-Julian Oscillation (MJO) index was obtained from quasi-biweekly mode. Quart. J. Roy. Meteor. Soc. 130 (599), 1171–1194. https://
the Bureau of Meteorology (BoM) (http://www.bom.gov. doi.org/10.1256/qj.03.133 w.
Chattopadhyay, R., Phani, R., Sabeerali, C.T., Dhakate, A.R., Salunke, K.D.,
au/climate/mjo/graphics/rmm.74toRealtime.txt). Mahapatra, S., Rao, A.S., Goswami, B.N., 2015. Influence of extratropical sea-surface
temperature on the Indian summer monsoon: an unexplored source of seasonal
Acknowledgments predictability. Quart. J. Roy. Meteor. Soc. 141 (692), 2760–2775.
Chen, X., Ling, J., Li, C., 2016. Evolution of the madden–julian oscillation in two types of
El Niño. J. Clim. 29, 1919–1934. https://doi.org/10.1175/jcli-d-15-0486.1.
The authors thank three anonymous reviewers for the valuable Cherchi, A., Terray, P., Ratna, S.B., Sankar, S., Sooraj, K.P., Behera, S., 2021. Chapter 8 -
comments and suggestions which helped in improving the quality of the Indian Ocean Dipole Influence on Indian Summer Monsoon and ENSO: a review,
Indian Summer Monsoon Variability. Elsevier, pp. 157–182. https://doi.org/
manuscript. Tanu Sharma acknowledges the research fellowship support 10.1016/B978-0-12-822402-1.00011-9. ISBN 9780128224021.
from the MRFP (Ministry of Earth Sciences Research Fellowship Pro­ DeMott, C.A., Wolding, B.O., Maloney, E.D., Randall, D.A., 2018. Atmospheric
gram) Project, MoES, Government of India. mechanisms for MJO decay over the Maritime Continent. J. Geophys. Res. Atmos.
123, 5188–5204. https://doi.org/10.1029/2017JD026979.
Dey, A., Chattopadhyay, R., Joseph, S., Kaur, M., Mandal, R., Phani, R., Sahai, A.K.,
Appendix A. Supplementary data Pattanaik, D.R., 2022. The intraseasonal fluctuation of Indian summer monsoon
rainfall and its relation with monsoon intraseasonal oscillation (MISO) and Madden
Julian oscillation (MJO). Theor. Appl. Climatol. 148 (1–2), 819–831.
Supplementary data to this article can be found online at https://doi.
Feng, J., Liu, P., Chen, W., Wang, X., 2015. Contrasting Madden-Julian Oscillation
org/10.1016/j.atmosres.2023.107214. activity during various stages of EP and CP El Niños. Atmos. Sci. Lett. 16, 32–37.
https://doi.org/10.1002/asl2.516.
References Gadgil, S., Gadgil, S., 2006. The Indian monsoon, GDP and agriculture. Econ. Polit. Wkly.
41, 4887–4895.
Gadgil, S., Vinayachandran, P.N., Francis, P.A., 2003. Droughts of the Indian summer
Ajayamohan, R.S., Rao, S.A., Yamagata, T., 2008. Influence of Indian Ocean dipole on monsoon: Role of clouds over the Indian Ocean. Curr. Sci. 85 (12), 1713–1719.
poleward propagation of boreal summer intraseasonal oscillations. J. Clim. 21 (21), http://www.jstor.org/stable/24109976.
5437–5454. Gadgil, S., Vinayachandran, P.N., Francis, P.A., Gadgil, S., 2004. Extremes of the Indian
Annamalai, H., Slingo, J.M., 2001. Active/break cycles: diagnosis of the intraseasonal summer monsoon rainfall, ENSO and equatorial Indian Ocean oscillation. Geophys.
variability of the Asian summer monsoon. Clim. Dyn. 18, 85–102. Res. Lett. 31, L12213. https://doi.org/10.1029/2004GL019733.

13
S.B. Ratna et al. Atmospheric Research 299 (2024) 107214

Gonzalez, A.O., Jiang, X., 2019. Distinct propagation characteristics of intraseasonal Pottapinjara, V., Girishkumar, M.S., Sivareddy, S., Ravichandran, M., Murtugudde, R.,
variability over the tropical West Pacific. J. Geophys. Res. Atmos. 124, 5332–5351. 2016. Relation between the upper ocean heat content in the equatorial Atlantic
https://doi.org/10.1029/2018JD029884. during boreal spring and the Indian monsoon rainfall during June–September. Int. J.
Gordon, A.L., 2005. Oceanography of the Indonesian seas and their throughflow. Climatol. 36 (6), 2469–2480.
Oceanog. 18 (4), 14–27. https://doi.org/10.5670/oceanog.2005.01. Qian, D., Guan, Z., 2018. Different features of super and regular El Niño events and their
Goswami, B.B., 2023. Role of the eastern equatorial Indian Ocean warming in the Indian impacts on the variation of the West Pacific subtropical high. Acta. Meteor. Sin.
summer monsoon rainfall trend. Clim. Dyn. 60, 427–442. https://doi.org/10.1007/ 394–407 https://doi.org/10.11676/qxxb2018.011.
s00382-022-06337-7. Rajeevan, M., Pai, D.S., 2007. On the El Nino-Indian Monsoon predictive relationships.
Goswami, B.N., Krishnamurthy, V., Annmalai, H., 1999. A broad-scale circulation index Geophys. Res. Lett. 34 https://doi.org/10.1029/2006GL028916.
for the interannual variability of the Indian summer monsoon. Quart. J. Roy. Meteor. Rajeevan, M., Gadgil, S., Bhate, J., 2010. Active and break spells of the Indian summer
Soc. 125 (554), 611–633. monsoon. J. Earth. Syst. Sci. 119, 229–247. https://doi.org/10.1007/s12040-010-
Gushchina, D., Dewitte, B., 2012. Intraseasonal tropical atmospheric variability 0019-4.
associated with the two flavors of El Niño. Mon. Weather Rev. 140, 3669–3681. Rao, S.A., Yamagata, T., 2004. Abrupt termination of Indian Ocean dipole events in
https://doi.org/10.1175/mwr-d-11-00267.1. response to intraseasonal disturbances. Geophys. Res. Lett. 31, L19306. https://doi.
Hendon, H.H., Wheeler, M.C., Zhang, C., 2007. Seasonal dependence of the MJO-ENSO org/10.1029/2004GL020842.
relationship. J. Clim. 20, 531–543. https://doi.org/10.1175/jcli4003.1. Rasmusson, E.M., Carpenter, T.H., 1983. The relationship between eastern equatorial
Huang, B., Peter, W.T., et al., 2017. Extended reconstructed sea surface temperature Pacific Sea surface temperatures and rainfall over India and Sri Lanka. Mon. Weather
version 5 (ERSSTv5), upgrades, validations, and intercomparisons. J. Clim. https:// Rev. 111 (3), 517–528.
doi.org/10.1175/JCLI-D-16-0836.1. Ratna, S.B., Cherchi, A., Osborn, T.J., Joshi, M., Uppara, U., 2021. The Extreme positive
Jiang, X., Adames, Á.F., Kim, D., Maloney, E.D., Lin, H., Kim, H., et al., 2020. Fifty years Indian Ocean Dipole of 2019 and Associated Indian Summer Monsoon Rainfall
of research on the Madden-Julian Oscillation: recent progress, challenges, and Response. Geophys. Res. Lett. https://doi.org/10.1029/2020GL091497.
perspectives. J. Geophys. Res. Atmos. 125, e2019JD030911 https://doi.org/ Revadekar, J., Preethi, B., 2012. Statistical analysis of the relationship between summer
10.1029/2019JD030911. monsoon precipitation extremes and foodgrain yield over India. Int. J. Climatol. 32,
Ju, J., Slingo, J., 1995. The Asian summer monsoon and ENSO. Quart. J. Roy. Meteorol. 419–429.
Soc. 121 (525), 1133–1168. Sabeerali, C.T., Rao, S.A., George, G., Rao, D.N., Mahapatra, S., Kulkarni, A.,
Kalnay, E., Kanamitsu, M., Kistler, R., Collins, W., Deaven, D., Gandin, L., et al., 1996. Murtugudde, R., 2014. Modulation of monsoon intraseasonal oscillations in the
The NCEP/NCAR 40-year reanalysis project. Bull. Am. Meteorol. Soc. 77 (3), recent warming period. J. Geophys. Res. Atmos. 119 (9), 5185–5203.
437–472. https://doi.org/10.1175/1520-0477(1996)077<0437:TNYRP>2.0.CO;2. Sabeerali, C.T., Ajayamohan, R.S., Giannakis, D., Majda, A.J., 2017. Extraction and
Kessler, W.S., Kleeman, R., 2000. Rectification of the Madden-Julian Oscillation into the prediction of indices for monsoon intraseasonal oscillations: an approach based on
ENSO cycle. J. Clim. 13, 3560–3575. nonlinear Laplacian spectral analysis. Clim. Dyn. 49 (9–10), 3031–3050.
Krishnamurthy, V., Goswami, B.N., 2000. Indian monsoon-ENSO relationship on Sabeerali, C.T., Ajayamohan, R.S., Bangalath, H.K., Chen, N., 2019. Atlantic zonal mode:
interdecadal timescale. J. Clim. 13, 579–595. an emerging source of Indian summer monsoon variability in a warming world.
Krishnamurti, T.N., Bhalme, H.N., 1976. Oscillations of a monsoon system. Part I. Geophys. Res. Lett. 46 (8), 4460–4467.
Observational aspects. J. Atmos. Sci. 33 (10), 1937–1954. Sabeerali, C.T., Ajayamohan, R.S., Praveen, V., 2022. Atlantic zonal mode-monsoon
Kucharski, F., Bracco, A., Yoo, J.H., Molteni, F., 2007. Low-frequency variability of the teleconnection in a warming scenario. Clim. Dyn. 1–15.
Indian monsoon–ENSO relationship and the tropical Atlantic: the “weakening” of the Saha, A., Ghosh, S., 2019. Can the weakening of Indian monsoon be attributed to
1980s and 1990s. J. Clim. 20 (16), 4255–4266. anthropogenic aerosols? Environ. Res. Commun. 1, 061006.
Kucharski, F., Bracco, A., Yoo, J.H., Tompkins, A.M., Feudale, L., Ruti, P., Dell’Aquila, A., Saith, N., Slingo, J., 2006. The role of the Madden–Julian Oscillation in the El Nino and
2009. A Gill–Matsuno-type mechanism explains the tropical Atlantic influence on Indian drought of 2002. Int. J. Climatol. 26 (10), 1361–1378.
African and Indian monsoon rainfall. Quart. J. Roy. Meteor. Soc. 135 (640), Saji, N.H., Goswami, B.N., Vinayachandran, P.N., Yamagata, T., 1999. A dipole mode in
569–579. the tropical Indian Ocean. Nat. 401 (6751), 360–363.
Kumar, K.K., Rajagopalan, B., Cane, M.A., 1999. On the weakening relationship between Sarkar, S., Singh, R.P., Kafatos, M., 2004. Further evidences for the weakening
the Indian monsoon and ENSO. Sci. 284 (5423), 2156–2159. https://doi.org/ relationship of Indian rainfall and ENSO over India. Geophys. Res. Lett. 31, L13209.
10.1126/science.284.5423.2156. https://doi.org/10.1029/2004GL020259.
Li, L., Chen, X., Li, C., Li, X., Yang, M., 2023. Comparison of Madden-Julian oscillation in Shukla, J., 1987. Interannual variability of monsoons. Monsoons 14, 399–464.
three super El Niño events. Front. Earth Sci. 10, 1021953. https://doi.org/10.3389/ Shukla, R.P., Huang, B., 2016. Interannual variability of the Indian summer monsoon
feart.2022.1021953. associated with the air–sea feedback in the northern Indian Ocean. Clim. Dyn. 46,
Liu, L., Yang, G., Zhao, X., Feng, L., Han, G., Wu, Y., Yu, W., 2017. Why was the Indian 1977–1990. https://doi.org/10.1007/s00382-015-2687-x.
Ocean Dipole Weak in the Context of the Extreme El Nino in 2015? J. Clim. 30, Sikka, D.R., Ratna, S.B., 2011. On improving the ability of a high resolution atmospheric
4755–4761. https://doi.org/10.1175/JCLI-D-16-0281.1. general circulation model for dynamical seasonal prediction of the extreme seasons
Madden, R.A., Julian, P.R., 1971. Detection of a 40–50 day oscillation in the zonal wind of the Indian summer monsoon. Mausam 62 (3), 339–360.
in the tropical Pacific. J. Atmos. Sci. 28 (5), 702–708. Singh, S.V., Kripalani, R.H., Sikka, D.R., 1992. Interannual variability of the
Mahendra, N., Chowdary, J.S., Darshana, P., Sunitha, P., Parekh, A., Gnanaseelan, C., Madden–Julian oscillations in Indian summer monsoon rainfall. J. Clim. 973–978.
2023. Interdecadal modulation of interannual ENSO-Indian summer monsoon Slingo, J.M., Annamalai, H., 1999. 1997: the El nino of the century and the response of
rainfall teleconnections in observations and CMIP6 models: Regional patterns. Int. J. the indian summer monsoon. Mon. Weather Rev. 128, 1778–1797.
Climatol. 41, 2528–2552. https://doi.org/10.1002/joc.6973. Son, S.-W., Lim, Y., Yoo, C., Hendon, H.H., Kim, J., 2016. Stratospheric control of the
Marjani, S., Alizadeh-Choobari, O., Irannejad, P., 2019. Frequency of extreme El Niño Madden-Julian Oscillation. J. Clim. 30, 1909–1922. https://doi.org/10.1175/JCLI-
and La Niña events under global warming. Clim. Dyn. 53, 5799–5813. https://doi. D-16-0620.1.
org/10.1007/s00382-019-04902-1. Sontakke, N.A., Pant, G.B., Singh, N., 1993. Construction of all-India summer monsoon
Marshall, A.G., Hendon, H.H., Wang, G., 2016. On the role of anomalous ocean surface rainfall series for the period 1844-1991. J. Clim. 6, 1807–1811.
temperatures for promoting the record Madden Julian oscillation in March 2015. Wang, C., Wang, X., 2013. Classifying El Niño Modoki I and II by Different Impacts on
Geophys. Res. Lett. 43, 472–481. https://doi.org/10.1002/2015GL066984. Rainfall in Southern China and Typhoon Tracks. J. Clim. 26 (4), 1322–1338.
Matthews, A.J., 2000. Propagation mechanisms for the Madden-Julian oscillation. Quart. Wang, J.W., Wang, C., 2021. Joint boost to super El Nino from the Indian and Atlantic
J. Roy. Meteor. Soc. 126 (569), 2637–2651. Oceans. J. Clim. 34, 4937–4954. https://doi.org/10.1175/JCLI-D-20-0710.1.
McPhaden, M.J., 1999. Genesis and evolution of the 1997–98 El Niño. Sci 283, 950–954. Wang, B., Yang, S., Lau, W.K.M., 2006. Interannual variability of the Asian monsoon.
Mujumdar, M., Sooraj, K.P., Krishnan, R., et al., 2017. Anomalous convective activity Asian Monsoon 259–293.
over sub-tropical East Pacific during 2015 and associated boreal summer monsoon Wang, L., Li, T., Chen, L., Behera, S.K., Nasuno, T., 2018. Modulation of the MJO
teleconnections. Clim. Dyn. 48, 4081–4091. https://doi.org/10.1007/s00382-016- intensity over the equatorial western Pacific by two types of El Niño. Clim. Dyn. 51,
3321-2. 687–700. https://doi.org/10.1007/s00382-017-3949-6.
Nishimoto, E., Yoden, S., 2017. Influence of the stratospheric quasi-biennial oscillation Webster, P.J., Yang, S., 1992. Monsoon and ENSO: selectively interactive systems. Quart.
on the Madden-Julian Oscillation during austral summer. J. Atmos. Sci. 74, J. Roy. Meteor. Soc. 118 (507), 877–926.
1105–1125. https://doi.org/10.1175/JAS-D-16-0205.1. Wheeler, M., Hendon, H.H., 2004. An all-season real-time multivariate MJO index:
Paek, H., Yu, J.Y., Qian, C., 2017. Why were the 2015/2016 and1997/1998 extreme El Development of an index for monitoring and prediction. Mon. Weather Rev. 132,
Niños different? Geophys. Res. Lett. 44, 1848–1856. https://doi.org/10.1002/ 1917–1932.
2016GL071515. Wilson, E.A., Gordon, A.L., Kim, D., 2013. Observations of the Madden Julian Oscillation
Pai, D.S., Bhate, J., Sreejith, O.P., et al., 2011. Impact of MJO on the intraseasonal during Indian Ocean Dipole events. J. Geophys. Res. Atmos. 118, 2588–2599.
variation of summer monsoon rainfall over India. Clim. Dyn. 36, 41–55. https://doi. https://doi.org/10.1002/jgrd.50241.
org/10.1007/s00382-009-0634-4. Yadav, R.K., Roxy, M.K., 2019. On the relationship between North Indian Summer
Pai, D.S., Sridhar, Latha, Rajeevan, M., Sreejith, O.P., Satbhai, N.S., Mukhopadhyay, B., Monsoon and East Equatorial Indian Ocean Warming. Glob. Planet. Chang. 179,
2014. Development of a new high spatial resolution (0.25◦ X 0.25◦ ) Long period 23–32. https://doi.org/10.1016/j.gloplacha.2019.05.001.
(1901-2010) daily gridded rainfall data set over India and its comparison with Yang, X., Huang, P., 2021. Restored relationship between ENSO and Indian summer
existing data sets over the region. Mausam 65, 1, 1–18. monsoon rainfall around 1999/2000. Innovat. 2 (2) https://doi.org/10.1016/j.
Parthasarathy, B., Munto, A.A., Kothawale, D.R., 1994. All-India monthly and seasonal xinn.2021.100102. ISSN 2666-6758.
rainfall series: 1871–1993. Theor. Appl. Climatol. 49, 217–224. Dio:10.1007/ Yoneyama, K., Zhang, C., 2020. Years of the Maritime Continent. Geophys. Res. Lett. 47,
BF00867461. e2020GL087182 https://doi.org/10.1029/2020GL087182.

14
S.B. Ratna et al. Atmospheric Research 299 (2024) 107214

Yu, S.-Y., Fan, L., Zhang, Y., Zheng, X.-T., Li, Z., 2021. Reexamining the Indian Summer Zhang, C., 2005. Madden-julian oscillation. Rev. Geophys. 43 (2) https://doi.org/
Monsoon Rainfall–ENSO relationship from its recovery in the 21st century: Role of 10.1029/2004RG000158.
the Indian Ocean SST anomaly associated with types of ENSO evolution. Geophys. Zhang, C., 2013. Madden–Julian oscillation: Bridging weather and climate. Bull. Am.
Res. Lett. 48, e2021GL092873 https://doi.org/10.1029/2021GL092873. Meteorol. Soc. 94 (12), 1849–1870.
Yuan, Y., Li, C., Ling, J., 2015. Different MJO activities between EP El Niño and CP El
Niño. Sci. China Earth Sci. 45, 318–334. https://doi.org/10.1360/zd-2015-45-3-
318.

15

You might also like