The severe El Niño events of 1997/1998 and 2015/2016 caused significant disruptions in Southeast Asia, particularly in Borneo, resulting in hazardous haze and acute water shortages. This study examines the influence of El Niño, the Indian Ocean Dipole (IOD), and the Madden–Julian oscillation (MJO) on regional climate, using time-series data from February 1993 to December 2020. Data from El Niño, IOD, and MJO indices were integrated with Landsat 5 and 8 land surface temperature records, allowing for a detailed analysis of their combined effects on regional temperature and precipitation patterns. Time-series trend decomposition and the generalized linear mixed model approach identified the Oceanic Niño Index (ONI) as a significant driver of temperature increases and dry spell occurrences during the peak El Niño years. On the other hand, ONI correlated strongly with mean monthly temperatures, underscoring its dominant influence. In addition, the IOD was found to significantly affect regional temperatures with a regression coefficient of 0.38867 (p = 0.0455), indicating its significant but less pronounced impact compared with ONI. These findings clarify the dynamics between key climate indices and their impact on regional climate extremes, offering critical insights for improving climate resilience and adaptation in tropical regions.

  • The Oceanic Niño Index significantly drove temperature increases and dry spells in Southeast Asia during peak El Niño years, while the Indian Ocean Dipole also impacted regional temperatures, providing essential insights for enhancing climate resilience and adaptation in tropical regions.

  • This study investigates the effects of El Niño, IOD, and MJO on Kuching City's climate using satellite data. It reveals that the IOD significantly affects regional temperatures, and the Oceanic Niño Index impacts heatwaves, offering insights into local climate resilience and adaptation strategies.

The years 1997/1998 and 2015/2016 marked the most intense El Niño events ever recorded in human history. These occurrences of the El Niño Southern Oscillation (ENSO) phenomenon are frequently associated with devastating droughts and floods (Kovats et al. 2003). ENSO consists of three distinct phases: El Niño, characterized by warm tropical Asia-Pacific sea surface temperatures; La Niña, characterized by cold tropical sea surface temperatures; and Neutral, which is neither El Niño nor La Niña. Each event possesses unique characteristics (Trenberth & Stepaniak 2001). Numerous studies have investigated the correlation between ENSO and temperature, including the works of Tan et al. (2020), Tavakol et al. (2020), and Eboy & Kemarau (2023). In Malaysia, the impact of the ENSO on climate variables is well documented and significant, contributing to notable climate variability (Tangang et al. 2012). Despite this, research remains sparse on how the Madden–Julian oscillation (MJO) and the Indian Ocean dipole (IOD), alongside ENSO, collaboratively influence local temperature fluctuations and broader climate patterns. This gap underscores the necessity of focused investigations into the combined effects of MJO, IOD, and ENSO on local land surface temperatures (LSTs), which could provide crucial insights for urban climate management and planning.

The decision to focus on a localized region for this pilot study is twofold. First, the unique geographical and climatic conditions of the Borneo region make it particularly susceptible to the effects of the IOD, MJO, and ENSO. This susceptibility provides a compelling case for studying the complex interplay and impacts of these climatic phenomena on local LST. Second, by conducting a focused study in a specific region, the aim is to establish a detailed and robust methodology that can be replicated or adapted for broader geographical analyses in future research. This approach not only fills a significant research gap by providing in-depth local data but also tests the effectiveness of integrating various climatic data sources on a smaller scale before their broader application.

In a comprehensive literature review, Islam et al. (2018) and Nurdiati et al. (2022) explored the effects of ENSO, IOD, and MJO on precipitation and temperature, considering the relevance of biomass burning. Lenka et al. (2022) focused on the impact of these phenomena on the Indian Summer Monsoon, and Huang et al. (2022) investigated extreme IOD events associated with ENSO and MJO at a regional scale. The complex relationship between El Niño, IOD, and MJO was highlighted due to their diverse spatial patterns (Huang et al. 2022). However, there is limited understanding of the influence of MJO, IOD, and ENSO on human comfort in urban areas, despite their significant contributions to climatological and meteorological anomalies such as heatwaves affecting urban populations (Garreaud et al. 2019; Tan et al. 2020; Eboy & Kemarau 2023). While some comprehensive studies have covered the entirety of Malaysia (Ibadullah et al. 2019; Tan et al. 2020), the relationship between the intensity of MJO, IOD, and El Niño events and local weather and climate remains limited, particularly in the context of satellite remote sensing (RS).

Most RS studies have focused on developed countries, particularly in high-latitude regions, analyzing sea surface temperature (SST), analyzing the impact of land use change on LST and rainfall (Tangang et al. 2012; Houk et al. 2017; Sharma et al. 2020, 2023). As each El Niño event exhibits distinct intensity, formation processes, and duration, there is considerable interest in comparing the differences between the MJO, IOD, and El Niño events of 1997/1998 and 2015/2016 (Houk et al. 2017). Further research in this area can provide valuable insights into the intricate interactions between climate phenomena and their impact on local weather and climate conditions. Previous research has highlighted that the impacts of El Niño events can vary based on the geographical location of a region. Malaysia's location within the Maritime Continent (MC) makes it particularly susceptible to the ocean–atmospheric interactions influenced by ENSO, IOD, and MJO, as emphasized by Eboy & Kemarau (2023). These phenomena play crucial roles in shaping the climate patterns of the MC region. The significance of the MC lies in its position at the center of the tropical warm pool in Southeast Asia, making it a focal point of interest for researchers. The intricate interplay between the ocean and the atmosphere in this area significantly impacts monsoon activity and weather patterns. The eastern regions of Peninsular Malaysia, along with Sabah and Sarawak, are significantly impacted by the ENSO, which has been shown to increase the incidence of dryer-than-normal conditions by 20–30% (Mohmud 2018). These areas, prone to variable climate conditions influenced by ENSO, also experience interactions with the IOD and the MJO, though the collective impact of these systems on local temperatures and climate anomalies remains underexplored.

Addressing this research gap, this study focuses on these specific regions, which are critically affected by these climatic oscillations, to better understand their combined effects on LST. This knowledge is crucial for effective climate adaptation strategies, particularly in urban areas where temperature fluctuations have direct implications on human comfort and urban infrastructure. Researchers are particularly keen on understanding how ENSO, IOD, and MJO influence the climate of the MC, given their significant consequences for weather systems in Malaysia and the broader region. By exploring these influences, scientists aim to gain valuable insights into the region's climate variability and enhance the ability to predict and adapt to weather changes. In the context of the tropically situated study area, factors such as the monsoon, IOD, and MJO have a strong influence. Tan et al. (2020) and Kemarau & Eboy (2021a, 2021b, 2021c) have utilized temperature data from the Malaysia Meteorological Department (MMD) to demonstrate the impact of El Niño on temperature. However, these studies have not provided spatial information regarding the localized patterns of LST. In other words, while previous studies have acknowledged the impact of El Niño on temperature using MMD data, they have not specifically addressed the spatial distribution of LST at a local scale. This spatial information is crucial for understanding the localized effects of El Niño on LST patterns in the area of interest. The objective of this study is to examine the differences in oceanic and atmospheric conditions during extreme ENSO events and explore the synergy between these climate phenomena to provide a comprehensive picture of their climatic influence.

Location of the study area

The climate in Kuching is influenced by various factors, including the MJO, ENSO, IOD (Hua et al. 2013), and the Monsoon season (Yik et al. 2015). However, there is a specific research gap focusing on the impact of ENSO, MJO, and IOD during El Niño events. This study aims to provide new insights into understanding the climate in tropical cities, particularly in the MC. In recent years, the Borneo region, including Kuching City in Sarawak, has experienced ENSO events, climate changes, and heavy rainfall incidents (Mahmud 2018; Kemarau & Eboy (2021a, 2021b, 2021c). Severe floods occurred in Kuching in 2015, affecting numerous low-lying areas (Kemarau & Eboy (2021a, 2021b, 2021c). Similarly, in 2009, two major rainfall events resulted in widespread floods across Sarawak, including Kuching (Sa'adi et al. 2019). These occurrences have raised concerns and emphasized the need for further research to statistically analyze meteorological data and evaluate the significance of these changes (Kemarau & Eboy (2021a, 2021b, 2021c). Moreover, there is a lack of studies examining the collective specific effects of MJO, IOD, and temperature in the tropical region, particularly in the Kuching city area (Figure 1).
Figure 1

Location of the study area: Kuching City in Sarawak, Malaysian Borneo.

Figure 1

Location of the study area: Kuching City in Sarawak, Malaysian Borneo.

Close modal

Landsat satellite dataset

The dataset detailed in Table 1 comprises Landsat imagery, meticulously selected for clarity and absence of cloud coverage, essential for the accurate retrieval of LST and land cover data. The selection specifically targets years corresponding to significant El Niño events 1997/1998 and 2015/2016 due to their pronounced impact on regional climate conditions. This approach ensures the reliability and relevance of the RS data utilized in the analysis.

Table 1

Detailed data information

SatelliteThermal bandSpatial resolution (meters)Data acquisition
DateTime (hours)
Landsat 5 Band 6 30 18 October 1997 1,420 
Landsat 5 Band 6 30 19 November 1997 1,421 
Landsat 8 Band 6 30 29 March 1998 1,430 
Landsat 8 Band 10 30 23 July 2015 1,451 
Landsat 8 Band 10 30 23 June 2016 1,451 
Landsat 8 Band 10 30 11 September 2016 1,451 
SatelliteThermal bandSpatial resolution (meters)Data acquisition
DateTime (hours)
Landsat 5 Band 6 30 18 October 1997 1,420 
Landsat 5 Band 6 30 19 November 1997 1,421 
Landsat 8 Band 6 30 29 March 1998 1,430 
Landsat 8 Band 10 30 23 July 2015 1,451 
Landsat 8 Band 10 30 23 June 2016 1,451 
Landsat 8 Band 10 30 11 September 2016 1,451 

Oceanic Niño Index (ONI)

The Oceanic Niño Index (ONI), a reliable indicator for identifying El Niño and La Niña events, is widely recognized and accessible from the National Oceanic and Atmospheric Administration (NOAA) website at NOAA's ONI page (Huang et al. 2016). It tracks the development and intensity of El Niño or La Niña in the Pacific Ocean by calculating SST anomalies observed in the Niño region 3.4 (5°N–5°S, 120°W–170°W). El Niño is defined when the average SST anomaly over 3 months is at or above +0.5 °C, while La Niña is defined when the average SST anomaly is at or below −0.5 °C (National Oceanic & Atmospheric Administration Forecast Center 2019). ENSO events are categorized into five classes based on their intensity: weak (anomalies of 0.5–0.9 °C), moderate (1.0–1.4 °C), strong (1.5–1.9 °C), and very strong (≥2.0 °C) for El Niño events, with opposite classifications for La Niña events. Data accessed from the National Oceanic and Atmospheric Administration Climate Prediction Center.

Madden–Julian oscillation (MJO)

The MJO represents a prominent tropical weather pattern that fluctuates on a weekly to monthly basis. Data on the MJO can be accessed and downloaded from the Bureau of Meteorology's dedicated portal at http://www.bom.gov.au/climate/mjo/. It involves pulsating clouds and rainfall near the equator, moving eastward, and repeating every 30–60 days. Eight equatorial regions worldwide are depicted in the graphs, representing the MJO's intensity and progression. To evaluate the MJO's strength and location, the real-time multivariate MJO (RMM) 1 and RMM 2 indices combine upper and lower atmospheric cloud cover and wind patterns. The index within the central circle indicates a weak MJO, while positions outside the circle represent greater strength. Typically, the index moves counterclockwise as the MJO progresses from west to east. Figure 2 displays the MJO's intensity and progression when the ONI value is above 3, signifying an El Niño event.
Figure 2

Patterns and influences of ONI, MJO, and IOD on mean monthly temperature (MMD) of the study area during the selection years 1997/1998 and 2015/2016.

Figure 2

Patterns and influences of ONI, MJO, and IOD on mean monthly temperature (MMD) of the study area during the selection years 1997/1998 and 2015/2016.

Close modal

Indian Ocean dipole (IOD)

The IOD is a significant climatic pattern that influences weather conditions across the Indian Ocean region. Relevant datasets for the IOD can be accessed at NASA's Sea Level Portal, available at https://sealevel.jpl.nasa.gov/overlay-iod/, which provides comprehensive data essential for the analysis. In its positive phase, warm waters are pushed toward the western part of the Indian Ocean, while cold deep waters rise to the surface in the eastern Indian Ocean. The pattern is reversed during the negative phase. The intensity of the IOD is measured by the anomalous SST gradient between the western equatorial Indian Ocean (50°E–70°E and 10°S–10°N) and the southeastern equatorial Indian Ocean (90°E–110°E and 10°S–0°N). This gradient is known as the dipole mode index (DMI). A positive DMI indicates a positive IOD, while a negative DMI indicates a negative IOD. The DMI data can be obtained from the National Oceanic and Atmospheric Administration Working Group on Surface Pressure.

Methodology

The pre-processing stage involves several crucial steps, including geometric correction, atmospheric correction, and radiometric correction, to ensure their usefulness for temperature research (Sharma et al. 2023). Subsequently, the surface temperature of the thermal band is extracted for the specific datasets from 1997/1998 and 2015/2016. The process for inverting surface temperature from Landsat 5 and Landsat 8 imagery, available for download at https://earthexplorer.usgs.gov/, adheres to the methodologies established by Kemarau & Eboy (2019). This procedure ensures accurate temperature retrieval, which is essential for the analysis.

Table 2 presents waveband information for the Landsat 8 Operational Land Imager (OLI) thermal infrared sensor (TIRS) satellite. The Landsat Satellite 8 follows a sun-synchronous orbit at an altitude of 705 km near the Earth's poles. It is equipped with the Instrument OLI consisting of nine spectral bands and two thermal infrared sensors, as detailed in the table.

Table 2

Information of Landsat 8

BandWavelengthResolution (meters)
Band 1 Coastal aerosol (0.43–0.45 μm) 30 
Band 2 Blue (0.450–0.51 μm) 30 
Band 3 Green (0.53–0.59 μm) 30 
Band 4 Red (0.64–0.67 μm) 30 
Band 5 Near-infrared (0.85–0.88 μm) 30 
Band 6 Short-wave infrared (SWIR) 1(1.57–1.65 μm) 30 
Band 7 SWIR 2 (2.11–2.29 μm) 30 
Band 8 Panchromatic (PAN) (0.50–0.68 μm) 15 
Band 9 Cirrus (1.36–1.38 μm) 30 
Band 10 TIRS 1 (10.6–11.19 μm) 100 
Band 11 TIRS 2 (11.5–12.51 μm) 100 
BandWavelengthResolution (meters)
Band 1 Coastal aerosol (0.43–0.45 μm) 30 
Band 2 Blue (0.450–0.51 μm) 30 
Band 3 Green (0.53–0.59 μm) 30 
Band 4 Red (0.64–0.67 μm) 30 
Band 5 Near-infrared (0.85–0.88 μm) 30 
Band 6 Short-wave infrared (SWIR) 1(1.57–1.65 μm) 30 
Band 7 SWIR 2 (2.11–2.29 μm) 30 
Band 8 Panchromatic (PAN) (0.50–0.68 μm) 15 
Band 9 Cirrus (1.36–1.38 μm) 30 
Band 10 TIRS 1 (10.6–11.19 μm) 100 
Band 11 TIRS 2 (11.5–12.51 μm) 100 

Moreover, this study integrates data from the Landsat 8 OLI TIRS, Landsat 7 ETM + , and Landsat 4/5 TM RS systems. For detailed information regarding the paths associated with Landsat 5 TM, please refer to Table 3.

Table 3

Comprehensive band information for Landsat 4/5 TM

BandWavelengthResolution (meters)
Band 1 Blue (0.43–0.52 μm) 30 
Band 2 Green (0.52–0.60 μm) 30 
Band 3 Red (0.63–0.69 μm) 30 
Band 4 Near-infrared (0.76–0.90 μm) 30 
Band 5 Near-infrared (1.55–1.65 μm) 30 
Band 6 Thermal (10.40–12.50 μm) 30 
Band 7 Mid-Infrared (2.08–2.35 μm) 120 
BandWavelengthResolution (meters)
Band 1 Blue (0.43–0.52 μm) 30 
Band 2 Green (0.52–0.60 μm) 30 
Band 3 Red (0.63–0.69 μm) 30 
Band 4 Near-infrared (0.76–0.90 μm) 30 
Band 5 Near-infrared (1.55–1.65 μm) 30 
Band 6 Thermal (10.40–12.50 μm) 30 
Band 7 Mid-Infrared (2.08–2.35 μm) 120 

In this study, ENVI 5.3 software is employed for data pre-processing, which includes atmospheric correction and conversion of digital numbers (DNs) to surface temperature. Atmospheric correction plays a vital role in mitigating and removing atmospheric effects on thermal infrared data. The thermal bands in Landsat 4 and 5 TM are found in Band 6. Initially, the sensor captures radiance values in DNs, which are then processed using calculations based on previous studies (Chander et al. 2009), such as the following equation:
(1)

According to Equation (1) provided, the variables are defined as follows: represents the pixel value as radiance in W/(m2 sr μm), LMAXλ is the spectral radiance scale factor to Qcalmax in W/(m2 sr μm), and LMINλ is the spectral radiance scale factor to Qcalmin in W/(m2 sr μm). Qcalmax corresponds to the maximum calibrated pixel value (equivalent to LMAX) in DN, which is 255, while Qcalmin corresponds to the minimum calibrated pixel value (equivalent to LMIN) in DN (Table 4).

Table 4

LMAX and LMIN information for Landsat 5 TM

InformationValue
 15.303 
 1.238 
InformationValue
 15.303 
 1.238 

The information regarding LMAX and LMIN was obtained from the downloaded metadata accompanying the RS data. In the case of Landsat 8 OLI TIRS, surface radiance in the thermal spectrum is captured by the TIRS 10. To convert DNs to radiance spectra above the atmosphere, Equation (2) is applied:
(2)

In this study, the radiance spectrum is measured in units of W/(m2 sr μm), while ML represents the recalibrated gain for a specific band, also in units of W/(m2 sr μm). ML corresponds to the path radiance, measured in units of W/(m2 sr μm). The ENVI 5.3 software is employed, which provides thermal functions specifically designed for atmospheric correction purposes. Before conducting atmospheric correction, it is necessary to convert the thermal images into radiance format.

The primary objective of atmospheric correction for thermal bands is to minimize and remove the atmospheric effects present in thermal infrared data. This study utilizes the in-scene atmospheric correction (ISAC) algorithm proposed by Young et al. (2002). ISAC relies on radiance data acquired at the aperture to estimate the upwelling radiance and atmospheric transmission, as described in Equation (3). Furthermore, Equation (1) is employed to calculate the radiance at the aperture, assuming that objects with the highest radiance are black.
(3)

The term ‘radiance’ refers to the observed radiance at the aperture, while τ represents the atmospheric transmission. B(T) corresponds to the Planck function, and Lu represents the upwelling radiance due to the atmosphere. The ISAC algorithm evaluates the surface radiance emitted based on the brightness temperature of each pixel. The scatter plot of radiance for each band follows the Planck formula. A linear regression line is fitted to the upper portion of the scatter plot, which represents pixels approaching an emissivity of 1.0. This process estimates the transmission (slope) and upwelling (intercept) factors. These factors, referred to as ‘transmission’ and ‘upwelling’, approximate the surface radiance leaving the Earth for each pixel within the original cube. By plotting the transmission and upwelling (τ, Lu) and generating the cube, the ENVI 5.3 algorithm incorporates this information to perform thermal atmospheric correction using ISAC.

In the case of Landsat 5 TM, information regarding the sensor is stored in the Landsat 5 TM metadata file. Assuming the Earth's surface behaves as a black body emitter with a spectral radiance, the radiance spectrum obtained from the sensor is then converted into an effective sensor brightness temperature using the following equation (Chander et al. 2009):
(4)
In Equation (4), represents the effective sensor brightness temperature in Kelvin (K), while k1 and k2 are calibration constants. For Landsat 5 TM, the values for k1 and k2 are set to 607.76 and 1,260.56 K, respectively. It is important to note that the sensor brightness temperature obtained from Equation (4) is related to a black body. To estimate the surface temperature of a gray body, the emissivity of the surface spectrum needs to be considered. This can be achieved by calculating the improved surface temperature emissivity using the following equation proposed by Artis & Carnahan (1982):
(5)

Surface temperature is typically measured in Kelvin (K). In the context of the mentioned equation, TB represents the effective sensor brightness temperature in K. The variable λ denotes the wavelength of the emitted light, expressed in meters. For the Landsat 5 TM thermal bands, the effective wavelengths are 11.457, 11.269, and 10.904 μm, based on Equation (6) with α = 1.438 × 10−2 mK. The symbol ε represents the surface emissivity.

To convert the surface temperature from Kelvin to degrees Celsius (°C), Equation (6) is applied:
(6)

The thermal surface radiance is captured by the TIRS band 10 on Landsat 8 OLI. Pre-processing is conducted to eliminate atmospheric effects from the thermal data, and the surface temperature is derived using the algorithm proposed by He et al. (2019).

To convert the DNs to spectral radiance L above the atmosphere, Equation (7) is applied:
(7)

In Equation (7), Lλ represents the spectral radiance in units of W/(m sr μm)2. ML corresponds to the rescaled gain specific to a particular band, measured in units of W/(m sr μm)2. AL accounts for the band-specific rescaling bias, also measured in units of W/(m sr μm)2. It is important to note that the rescaling bias is applied to a specific band.

To extract the sensor brightness temperature TB in Kelvin (K) from TIRS band 10, similar to the OLI sensor, the following equation is utilized:
(8)

In Equation (8), K_1 and K_2 are predefined constants set to 774.89 (in W/(m sr μm)2) and 1,321.08 (in W), respectively. The surface temperature is determined by correcting the emitted radiance of the ground (B) in Kelvin (K) using the mono-window algorithm proposed by Qin et al. (2001).

Normalized difference vegetation index (NDVI) is calculated using Equation (9), where ρNIR and ρRED represent the values of the near-infrared and red bands, respectively. Calculating NDVI is important as it closely relates to the vegetation cover fraction (Pv). Pv represents the percentage of vegetation in the pixel, NDVIs (normalized difference vegetation index for soil) represents the NDVI value characteristic of bare soil within the same area, NDVIv (normalized difference vegetation index for vegetation) represents the NDVI value typical of dense, healthy vegetation within the study area and are calculated based on Carlson & Ripley (1997) using the following equation:
(9)
Pv is also related to the emissivity (E), which is calculated using the following equation:
(10)

In Equation (10), ε represents the land emissivity, εv represents the emissivity of pure vegetation (set to 0.985), εg represents the emissivity of bare soil (set to 0.960), and (A parameter representing the adjustment to an average value of 0.01, used to refine emissivity calculations based on specific environmental conditions) is a parameter adjusted to an average of 0.01.

Finally, the surface temperature in Kelvin is converted to Celsius using the following equation:
(11)

In Equation (11), TB represents the sensor brightness temperature in Kelvin (K), λ is wavelength of the radiation (specific to the sensor's operational band), C2 is the second radiation constant in Planck's law equation, ε is the overall emissivity used in the conversion from brightness temperature to actual surface temperature, and ρ = hc/σ, where σ is the Boltzmann constant (1.38 × 10−23 J/K), h is the Planck constant (6.626 × 10−34J/s), c is the speed of light (2.998 × 108m/s), and ε is the emissivity.

After generating the LST maps, a correlation analysis were performed between the LST values and the mean temperature recorded between 1,400 and 1,500 h on the dates coinciding with satellite captures in the study area. Meteorological data from the MMD were used for this analysis. It is important to note that the pixel coordinates in the LST maps correspond to the locations of the meteorological stations to validate the LST from the Landsat satellite.

Next, the spatial patterns of the LST maps were compared between the El Niño years of 1997/1998 and 2015/2016. This comparison aimed to identify any temperature distribution differences during these two El Niño events.

Moreover, this study performed statistical comparisons on the mean, maximum, and minimum LST values extracted from the LST maps of the years 1997/1998 and 2015/2016. These statistical analyses yielded significant insights into the changes in temperature characteristics associated with the two El Niño events. To facilitate the interpretation of these comparisons, graphical representations were employed.

Data of MMD (temperature C), MJO, IOD, and ONI were compiled and synched based on date, from February 1993 until December 2020. The compiled data were imported into R programming via RStudio and transformed into a time-series data format. MMD was treated as the dependent variable, while MJO, IOD, and ONI were treated as co-variate. The time-series data were decomposed using ggsdc (ggseas V0.5.4) to observe the trend, seasonality variation, and noise frequency in each series. A normality test (Shapiro–Wilk test) was conducted on the MMD to examine the distribution pattern of the dependent variable (MMD). A generalized linear mixed model (GLMM) using Template Model Builder was constructed to test if the interactions of MJO, IOD, and ONI were significant in driving the fluctuations of MMD. MMD was set as a Gaussian distribution for GLMMTMB.

The research team employed the ISO Cluster classification method to develop land cover maps for the years 1998 and 2016, chosen due to the availability of the clearest satellite images with minimal cloud cover. This clustering algorithm grouped pixels with similar spectral characteristics into various land cover types, aiding in the identification and characterization of different land cover within the study area. The precision of these maps was assessed using a time-series analysis from Google Earth, following the methodology outlined by Kemarau & Eboy (2020). This analysis involved comparing the land cover maps with high-resolution satellite imagery and other reliable sources to ensure the accuracy and reliability of the classifications. The primary goal of the study was to categorize the data into four main surface cover types: urban areas, vegetation, wetlands, and water bodies. The validation process was crucial for assessing the reliability of RS data in identifying hotspot prone areas, particularly during dry and heatwave events. This endeavor required close collaboration with relevant departments to ensure effective assistance and intervention during such critical weather occurrences. The findings obtained through this comprehensive approach were expected to offer valuable insights into the intricate interactions between climate phenomena and LST, ultimately contributing to an enhanced comprehension of climate variability in the study area.

The correlation analysis yielded a correlation coefficient of 0.97, indicating a strong relationship between the LST values and the ground-based meteorological data. This result was consistent with the findings of Eboy & Kemarau (2023), who also reported a similar high correlation between satellite-derived LST and ground-based measurements. Notably, the pixel coordinates in the LST maps align with the location of meteorological stations, enabling the validation of LST data obtained from the Landsat satellite.

The patterns and influences of the ONI, MJO, and IOD on the mean monthly temperature (MMT) in the study area are examined in this study (Figure 2). The ONI values serve as indicators of the intensity of El Niño events during the specified years. Positive ONI values (above 0.5) indicate the presence of El Niño, which is associated with warmer sea surface temperatures. Conversely, negative ONI values (below −0.5) indicate the presence of La Niña, associated with cooler sea surface temperatures. The table shows that during the El Niño years (January 1997–December 1998 and January 2015–December 2016), the ONI values range from −1.6 to 2.6. Notably, there is a trend of increasing ONI values from negative to positive during the El Niño periods, signifying a transition from La Niña to El Niño conditions. This transition is accompanied by a gradual increase in MMT in the study area.

The MJO values in the table represent the intensity of the MJO, an atmospheric phenomenon that impacts tropical weather patterns. The MJO index indicates the presence of enhanced or suppressed rainfall in the region. Throughout the specified years, the MJO values vary. Some months exhibit higher MJO values (>1), indicating a stronger influence on weather patterns, while other months have lower values (<1), suggesting a weaker impact. The influence of MJO on MMD is not straightforward and can vary depending on the specific phase and intensity of the MJO. However, it is generally expected that during periods of enhanced rainfall (high MJO values), there may be a cooling effect on MMD due to increased cloud cover and precipitation.

The IOD values represent the IOD index, which measures the temperature gradient between the western and eastern parts of the equatorial Indian Ocean. Positive IOD values indicate a positive phase of the IOD, characterized by warmer waters in the western Indian Ocean and colder waters in the eastern Indian Ocean. Negative IOD values indicate a negative phase, with the temperature gradient reversed. The IOD values range from −0.76 to 1.28, and during the El Niño years, they vary, indicating different phases of the IOD. The influence of IOD on MMD can be complex, as it depends on the interaction between the IOD phase and other climate factors. Generally, a positive IOD phase, combined with El Niño conditions, may amplify the warming effect on MMD, while a negative IOD phase may have a weaker influence. The MJO index exhibits high variability, characterized by frequent, sharp fluctuations throughout the entire period. This represents the short-term cyclical nature of the MJO, which typically recurs every 30–60 days. There is no clear long-term trend in the MJO data, indicating stability in its cyclical pattern over the years (Figure 3).
Figure 3

The steps taken to accomplish the study's objective.

Figure 3

The steps taken to accomplish the study's objective.

Close modal

The IOD index also shows significant variability but with less frequency and amplitude compared with the MJO. Similar to the MJO, the IOD does not display a clear long-term trend, suggesting that the dipole effect remains consistent without significant changes in intensity over the years. The ONI, which measures the ENSO phenomena, shows periodic spikes that correlate with known El Niño events, particularly noticeable in years such as 1998, 2010, and 2016. These spikes represent the warm phases of ENSO, known as El Niño, which typically lead to higher global temperatures. There are dips corresponding to the La Niña events, which are cooler phases, though they are less pronounced in this graph. The temperature, represented by MMD (mean monthly temperature), generally trends upward over the observed period, with notable increases aligning with El Niño events (e.g., the peaks around 1998 and 2016). This suggests a correlation between high ONI values and increased regional temperatures. The overall upward trend in temperature across the years could be indicative of longer-term warming trends, possibly influenced by global climate change in addition to ENSO cycles.

The decomposed time-series data of the climate indices (MJO, IOD, and ONI) and MMT were conducted separately due to the different value ranges (−2 to 2, and 25–28) as displayed in Figure 4. There was no straightforward trend of IOD, MJO, and ONI (right graph). ONI was a relatively more varied trend as compared with the rest of the climate indices, with major peaks during 1997–1998 and 2015–2016, which represent the El Nino years. MJO did not show any observable trend, except for a small increase since 2015. Lastly, IOD saw a small but insignificant trend of increase. On the other hand, MMD saw an increase in temperature generally since 1993. However, there were two MMT spikes between January 1997 to December 1997 continuing to December 1998, and from January 2015 to December 2016. These two periods of temperature increase coincide with the IOD spikes. Seasonal changes of the time-series data showed only MMT and MJO showed seasonal fluctuation, specifically annual fluctuation. Irregular values of the time-series data showed that MJO had the highest irregular data, followed by MMT, ONI and the least variation is IOD. The MMT data series showed to be normally distributed (Shapiro–Wilk: W = 0.994, p = 0.244). Thus, the GLMMTMB model was constructed using a Gaussian distribution. GLMMTMB showed only IOD was significantly related to the fluctuation of MMT. Furthermore, a GLMM analysis showed that the IOD index had a significant relationship with fluctuations in MMT. The regression formula derived from this model was MMT = 0.38867 * IOD + 26.69622, indicating a strong correlation that is visually confirmed by a scatter plot, underscoring the influence of IOD on regional temperatures (Figure 5).
Figure 4

A comparative analysis of decomposed time-series data for two sets of variables.

Figure 4

A comparative analysis of decomposed time-series data for two sets of variables.

Close modal
Figure 5

Scatter plot correlation: MMT, IOD, and ONI.

Figure 5

Scatter plot correlation: MMT, IOD, and ONI.

Close modal

Table 5 from the GLMMTMB analysis provides detailed regression coefficients along with their statistical significance for various climate indices and their interactions, focusing on their effects on MMT. The intercept of 26.69622, with a z-value of 444.5 and a highly significant p-value (<2e − 16), indicates the base level of MMD when all other indices are at zero, showcasing the model's robust baseline for temperature predictions.

Table 5

Generalized Linear Mixed Models using Template Model Builder (glmmTMB) output of the regression constant for the co-variates with MMD

EstimateStandard Errorz valuep-value
(Intercept) 26.696 0.060 444.5 <0.001* 
IODa 0.389 0.194 2.0 0.046* 
ONIb 0.057 0.084 0.7 0.498 
MJOc −0.029 0.050 −0.6 0.555 
IOD: ONI −0.236 0.219 −1.1 0.281 
IOD: MJO −0.300 0.162 −1.9 0.064 
ONI: MJO 0.133 0.068 1.9 0.052 
IOD: ONI: MJO 0.117 0.197 0.6 0.551 
EstimateStandard Errorz valuep-value
(Intercept) 26.696 0.060 444.5 <0.001* 
IODa 0.389 0.194 2.0 0.046* 
ONIb 0.057 0.084 0.7 0.498 
MJOc −0.029 0.050 −0.6 0.555 
IOD: ONI −0.236 0.219 −1.1 0.281 
IOD: MJO −0.300 0.162 −1.9 0.064 
ONI: MJO 0.133 0.068 1.9 0.052 
IOD: ONI: MJO 0.117 0.197 0.6 0.551 

*p < 0.05

aIOD = Indian Ocean Dipole

bONI = El Niño-Southern Oscillation

cMJO= Madden-Julian Oscillation

The coefficient for the IOD stands at 0.38867, demonstrating a positive and statistically significant relationship (p-value = 0.0455) with MMT, suggesting that higher IOD values correlate with increased temperatures. In contrast, the ONI has a smaller, non-significant positive effect (coefficient = 0.057, p-value = 0.4978), indicating that ONI's influence on temperature, while present, does not consistently impact MMT. Similarly, the MJO shows a negative coefficient (−0.02927) with a non-significant p-value (0.5552), highlighting its minimal and inconsistent effect on temperature changes.

Regarding interaction terms, both IOD:ONI and IOD:MJO interactions display negative coefficients (−0.23575 and −0.29962, respectively) but lack statistical significance, hinting at possible moderating effects where the presence of one index might reduce the impact of another on temperatures. The ONI:MJO interaction, with a positive coefficient of 0.1325, suggests a potential synergistic effect that nearly reaches statistical significance (p-value = 0.0519), possibly indicating the combined influences of these indices under specific conditions. Lastly, the three-way interaction (IOD:ONI:MJO) shows a positive effect (0.11724), though it is not statistically significant (p-value = 0.5508), suggesting that the combined influence of these indices on temperature is complex and requires further investigation. Overall, this regression analysis underlines the nuanced relationships between climate indices and regional temperatures, with the IOD showing a clear impact, while the effects of ONI and MJO are more variable. These findings are critical for refining predictive models and enhancing the understanding of how these climate phenomena influence regional climate patterns. However, during extreme events, particularly El Niño 1997/1998 and 2015/2016, when the ONI values spiked significantly higher compared with IOD and MJO, these indices substantially influenced MMT (Kemarau & Eboy 2019) and LST (Eboy & Kemarau 2023; Kemarau & Eboy 2023).

The findings of the study differ from Lenka et al. (2022), who demonstrated a strong relationship between ENSO and the Indian monsoon, as well as the influence of the trans-Niño-3 pattern on the connection between typical ENSO measurements and overall rainfall in India (Schulte et al. 2021). In addition, the IOD also significantly affects the Indian climate system, particularly Indian monsoon rainfall (Ratna et al. 2021; Lenka et al. 2022). A positive relationship between the IOD and El Niño in the Pacific Ocean is illustrated in Figure 6. Changes in the surface temperature of the Pacific Ocean have indirect effects on global weather and climate, leading to temperature variations on the surface of the Indian Ocean. When there is an increase in surface temperature in the Indian Ocean, it indicates a positive value of the IOD index. This positive correlation is particularly evident in the charts for January, February, and March of 1998, as well as October and November of 2015, and January, February, and March of 2016.
Figure 6

The intensity and progression of the Madden–Julian oscillation (Source: Bureau of Meteorology Australia) and SST (Source: APEC Climate Centre, Busan) during El Niño events (ONI value >1) are indicated. The red box highlights the location of the study.

Figure 6

The intensity and progression of the Madden–Julian oscillation (Source: Bureau of Meteorology Australia) and SST (Source: APEC Climate Centre, Busan) during El Niño events (ONI value >1) are indicated. The red box highlights the location of the study.

Close modal

During positive IOD and El Niño periods, both the Asian continent and Western Australia experience drought-like conditions. Previous studies have highlighted that in drought years, the SST in the East Indian Ocean is significantly lower (Streten 1983; Lenka et al. 2022). It is well established that ENSO phases strongly influence Australian rainfall, with El Niño events leading to below-average precipitation (Ropelewski & Halpert 1987; Lenka et al. 2022). However, with the discovery of IOD, the understanding of annual Australian rainfall has evolved slightly. Observational studies have shown that despite being in an El Niño period, widespread and substantial anomalies occur due to overlapping positive IOD years (Nicholls 1989).

Several researchers have investigated the relationship between local MJO in the Pacific and ENSO SST indices, revealing strong correlations (Kessler & Kleeman 2000). However, in the global context, no such relationship has been reported, which is consistent with the findings of this study, as supported by Slingo et al. (1999) and Hendon et al. (1999). Due to the MJO, oceanic Kelvin waves and warm surface water travel eastward, enabling the MJO to extend further into the central Pacific. This phenomenon occurs concurrently with El Niño phases (Hendon et al. 1999; Lenka et al. 2022), but it does not significantly impact LST, as found in this study. Some studies suggest that intra-seasonal variability in MJO circulation induces tropospheric precipitation, which may directly influence deep convective conditions (Tompkins 2001) or indirectly affect moist stability (Myers & Waliser 2003). Similarly, SST also influences MJO, as mean SST determines MJO propagation speed and period by maintaining a low moisture level (Davey 1989). To analyze the influence of IOD phases on different regions, researchers have conducted studies using multi-source observed data and model simulations (Kayano et al. 2021; Sena & Magnusdottir 2021) to assess the combined processes.

According to Wang & Wang (2014), El Niño Modoki weakens the Walker circulation, resulting in reduced rainfall in the eastern tropical Indian Ocean, while a positive IOD tends to occur in the Indian Ocean. The impact is reversed in the case of El Niño Modoki II. Generally, IOD and ENSO phases are correlated, as the ENSO phase influences SST and circulation, leading to the occurrence of IOD (Polonsky & Tolonsky 2021). Previous case studies also supported the notion that IODs are independent of ENSO (Sun et al. 2015). The findings of this study indicate that the El Niño events of 1997/1998 and 2015/2016 significantly influenced the weather in Kuching, Sarawak. The impact of ENSO on hotspot areas (extreme heat) shows a more pronounced increase compared with natural conditions, as explained in the study (Figures 7 and 8).
Figure 7

LST between the El Niño events of 1997/1998 and 2015/2016.

Figure 7

LST between the El Niño events of 1997/1998 and 2015/2016.

Close modal
Figure 8

Map of LST and land cover during El Niño and La Niña.

Figure 8

Map of LST and land cover during El Niño and La Niña.

Close modal

Table 6 presents the spatial changes in LST between the El Niño events of 1997/1998 and 2015/2016, as shown in Figure 7. The results indicate notable differences between the two events. In terms of mean temperature, El Niño 2015/2016 recorded a higher value compared with El Niño 1997/1998. Similarly, the maximum temperature during the 2015/2016 El Niño event was also higher than that of the 1997/1998 El Niño event. However, the minimum temperature values varied depending on the ONI value. For instance, during El Niño 1997/1998, the minimum temperature occurred at an ONI value of 1.9, while for El Niño 2015/2016, it occurred at an ONI value of 1.

Table 6

The LST spatial changes between 1997/1998 and 2015/2016 El Niño

Statistic valueONIEl Niño 2015/2016ONIEl Niño 1997/1998
Max 2.4 33.19 1.9 30.42 
Min 26.19 24.93 
Mean 31.13 29.14 
Max 1.8 32.59 30.11 
Min 25.09 25.1 
Mean 30.87 28.77 
Max 30.89 1.9 30.42 
Min 24.27 24.93 
Mean 29.14 28.98 
Statistic valueONIEl Niño 2015/2016ONIEl Niño 1997/1998
Max 2.4 33.19 1.9 30.42 
Min 26.19 24.93 
Mean 31.13 29.14 
Max 1.8 32.59 30.11 
Min 25.09 25.1 
Mean 30.87 28.77 
Max 30.89 1.9 30.42 
Min 24.27 24.93 
Mean 29.14 28.98 

These findings align closely with the study conducted by Houk et al. (2020), which observed a higher temperature value of 3.5 °C during the El Niño 2015/2016 event due to warming in the central Pacific Ocean, surpassing the 1997/1998 event. The study also revealed that the degree of El Niño's effect depends on the different phases of El Niño formation, particularly the rate of increase and decrease of ONI. In addition, Thirumalai et al. (2016) noted that April 2016 recorded the highest LST in history, coinciding with the peak of the 1997/1998 El Niño event. They further highlighted the contribution of global warming to rising temperatures. These findings support the results of the current study.

Despite the significant influence of El Niño, it is important to consider the impact of urbanization activities on the increase in LST (Eboy & Kemarau 2023). This explains the higher mean and maximum LST values observed during the El Niño 2015/2016 event. Furthermore, Thirumalai et al. (2016) found that April 2016 set a record for the highest temperature in the history of El Niño. Table 7 illustrates LST data for different years under varying conditions of three key climate indices: ONI, IOD, and MJO. The years selected – 2016, 1998, 2018, and 2011 – represent different climatic anomalies as characterized by the values of ONI, IOD, and MJO.

Table 7

LST value for each land cover during La Nina and El Nino

YearClimate indicesLST typeUrban (°C)Industrial (°C)Water bodies (°C)Vegetation (°C)Wetlands (°C)
2016 ONI 2.4, IOD −0.04, MJO 1.3 Max 35.81 35.9 27.29 26.99 26.67 
Mean 33.1 33.16 26.3 25.54 25.44 
Min 30.12 31.14 23.22 24.29 24.6 
1998 ONI 1.9, IOD 0.8, MJO 0.8 Max 33.31 33.29 27.12 26.81 26.77 
Mean 32.11 31.13 26.56 26.31 26.56 
Min 30.89 31.01 26.19 26.2 26.34 
2018 ONI −0.8, IOD −0.5, MJO 1.2 Max 31.39 31.89 24.2 24.12 23.23 
Mean 29.73 29.82 23.15 23.12 23.15 
Min 27.47 27.86 22.12 22.11 22.89 
2011 ONI −0.8, IOD −0.66, MJO 0.22 Max 30.55 30.21 23.42 23.13 22.23 
Mean 28.89 29.12 23.11 23.11 23.12 
Min 27.86 28.91 19.42 19.46 20.11 
YearClimate indicesLST typeUrban (°C)Industrial (°C)Water bodies (°C)Vegetation (°C)Wetlands (°C)
2016 ONI 2.4, IOD −0.04, MJO 1.3 Max 35.81 35.9 27.29 26.99 26.67 
Mean 33.1 33.16 26.3 25.54 25.44 
Min 30.12 31.14 23.22 24.29 24.6 
1998 ONI 1.9, IOD 0.8, MJO 0.8 Max 33.31 33.29 27.12 26.81 26.77 
Mean 32.11 31.13 26.56 26.31 26.56 
Min 30.89 31.01 26.19 26.2 26.34 
2018 ONI −0.8, IOD −0.5, MJO 1.2 Max 31.39 31.89 24.2 24.12 23.23 
Mean 29.73 29.82 23.15 23.12 23.15 
Min 27.47 27.86 22.12 22.11 22.89 
2011 ONI −0.8, IOD −0.66, MJO 0.22 Max 30.55 30.21 23.42 23.13 22.23 
Mean 28.89 29.12 23.11 23.11 23.12 
Min 27.86 28.91 19.42 19.46 20.11 

The year 2016, with ONI at 2.4, represents a strong El Niño phase, which typically results in higher temperatures globally (Figure 2). The maximum LSTs recorded in urban and industrial areas (35.81 and 35.9 °C, respectively) align with expected increases due to El Niño's influence on heat accumulation in built environments. Vegetation and wetlands show slightly lower temperatures, indicating their cooling effect even under El Niño conditions. This year also exhibits an El Niño condition, though with a positive IOD value (0.8). The combination of a positive IOD with El Niño can enhance warming effects in certain regions, particularly Southeast Asia and Australia (Weller & Cai 2013). The LSTs are generally high but slightly lower than in 2016, potentially due to the mitigating effects of a less intense El Niño and the regional impacts of the IOD.

With negative values for both ONI and IOD, 2018 reflects La Niña conditions, typically associated with cooler, wetter weather in the equatorial Pacific region. This is reflected in the generally lower maximum, mean, and minimum LSTs across all types of land cover compared with El Niño years. The cooling effect is particularly evident in water bodies and wetlands, which register the lowest temperatures, emphasizing the significant influence of La Niña on water-rich environments. Similar to 2018, 2011 was a year of La Niña conditions but with very weak MJO activity. The generally low LSTs across all categories, especially in water bodies and wetlands, corroborate the expected cooling effects of La Niña. The weak MJO might have contributed to less variability in temperature, resulting in consistently lower temperatures across the board.

The data show a clear pattern where El Niño years are associated with higher temperatures across most land covers, particularly in urban and industrial areas, which are more susceptible to heat due to the urban heat island effect. Conversely, La Niña years show a general cooling trend, especially noticeable in natural landscapes such as water bodies and wetlands. These observations are consistent with the known effects of these climate phenomena on global and regional climate patterns, as documented in the literature, where El Niño events have been shown to increase temperatures by 0.5 − 1.5 °C according to data provided by weather stations in Kuching (Panisset et al. 2017; Moura et al. 2019; Kemarau & Eboy 2023). Eboy & Kemarau (2023) also note that rainfall plays a critical role in shaping local temperature patterns, reporting that El Niño contributes to drought and dry conditions due to reduced rainfall, factors that contribute to temperature increases in Kuching during these events.

Furthermore, Mahmud (2018) reported that El Niño decreases rainfall by 25–30% compared with neutral conditions. Surface cover types such as wetlands, water bodies, vegetation, and urban, and industrial areas exhibit higher temperatures during El Niño compared with La Niña. Consistent with previous scientific reports, industrial areas have the highest temperatures due to heat generation and significant greenhouse gas emissions (Li et al. 2011; Kemarau & Eboy 2023). Figure 8 details changes in surface temperature patterns based on the type of land cover.

According to Figure 8 and Table 7, water bodies such as the Sarawak River and vegetation have lower temperatures compared with urban areas. For instance, water bodies are around 26 °C, and vegetation varies from 27.5 to 29 °C, whereas urban areas exceed 29 °C. Temperatures depend on the volume of vegetation, its quantity, and the extent of area it covers. Urban areas experience the highest temperatures due to the presence of man-made structures that absorb and release heat, thus raising temperatures (Estoque & Murayama 2017; Kemarau & Eboy 2023). The warming of urban areas, consisting of man-made materials, has become a significant contributor to urban heat islands, especially when indoor air conditioning is used (Parker 2010). Kemarau & Eboy (2023) also report that emissions of carbon dioxide (CO2) and carbon monoxide (CO) in industrial zones cause local temperature increases.

Figure 9 displays the spatial distribution of surface temperatures, indicating areas where LSTs exceed 30 °C (shown in red), during neutral and El Niño events. The data clearly show that the extent of hotspots during El Niño events surpasses that during neutral events. Figures 9 and 10 offer a comprehensive examination of the variations in hotspot areas between La Niña and El Niño occurrences.
Figure 9

Hotspot areas during the El Nino and neutral events.

Figure 9

Hotspot areas during the El Nino and neutral events.

Close modal
Figure 10

Hotspot area in km2 (>30 °C) during El Niño and neutral events.

Figure 10

Hotspot area in km2 (>30 °C) during El Niño and neutral events.

Close modal

According to Figure 9, it is clear that the hotspot area during the El Niño event exceeds that during the neutral event. Specifically, during the El Niño events of 2015 and 2016, the hotspot areas covered 95.55 and 97.8 km2, respectively. In contrast, during the neutral events in May 2014 and October 2017, the hotspot areas measured 66.8 and 73.3 km2, respectively. This consistent pattern of concentrated heat in the region persisted during the 1998 El Niño event, with a hotspot area of 89.32 km2 (Eboy & Kemarau 2023).

These findings provide evidence of the influence of ENSO on the alteration of hotspot patterns. The observed increase in a hotspot area, ranging from 34 to 36 km2, during the El Niño event, corresponds to a rise in electricity consumption of 20–30% above normal levels. This increase in electricity usage is aimed at balancing temperatures in residential and office areas in Kuching City, Sarawak. These results align with previous studies conducted by Lin et al. (2018) and Drosdowsky & Wheeler (2014), which have highlighted the contribution of El Niño events to global temperature rise. This global temperature increase indirectly impacts local temperatures, particularly in the study location. Furthermore, supporting evidence for this phenomenon can be found in the works of Tan et al. (2020) and Moura et al. (2019). Yu et al. concluded that the influence of local climate does not affect the impact of dominant El Niño warming, especially during the mature stage.

This study has elucidated the significant impacts of the IOD, MJO, and ENSO on LST in the MC, specifically in Kuching, Sarawak, Malaysia, during the severe El Niño events of 1997/1998 and 2015/2016. The key findings indicate that the ONI plays a pivotal role in modulating temperature extremes and dry spells, exerting a pronounced influence during peak El Niño years. In addition, the IOD also exhibits a considerable, albeit less pronounced, impact on regional temperatures. These findings have substantial implications for climate research, highlighting the need for integrated climate monitoring systems in the region to better predict and mitigate the effects of such climate phenomena. For policymakers, this information underscores the importance of developing and implementing more effective climate resilience and adaptation strategies, especially in areas highly vulnerable to ENSO-driven anomalies.

Considering the complexities observed in the interactions between these climate systems, future research should focus on several areas. Enhanced spatial analysis is needed to provide more detailed investigations that could help delineate the localized impacts of these phenomena, potentially uncovering microclimatic variations critical for urban planning and agricultural practices. Furthermore, longitudinal studies involving extended time-series analyses spanning multiple climate cycles could offer deeper insights into the long-term trends and shifts in climate patterns due to global climate change. Lastly, applying interdisciplinary approaches that integrate ecological, socio-economic, and climatological data could enhance the understanding of the broader impacts of climate variability on ecosystems and human communities in the MC. This study's approach and findings contribute valuable knowledge to the field of tropical climatology and provide a solid foundation for further explorative studies aimed at refining the predictive capabilities and response strategies to climate variability.

We acknowledge NASA for providing Landsat data and the Malaysia Meteorological Department for providing meteorological data.

Data cannot be made publicly available; readers should contact the corresponding author for details.

The authors declare there is no conflict.

Artis
D. A.
&
Carnahan
W. H.
1982
Survey of emissivity variability in thermography of urban areas
.
Remote Sensing of Environment
12
(
4
),
313
329
.
Australia Government (Bureau of Meteorology)
Carlson
T. N.
&
Ripley
D. A.
1997
On the relation between NDVI, fractional vegetation cover, and leaf area index
.
Remote Sensing of Environment
62
,
241
252
.
Chander
G.
,
Markham
B. L.
&
Helder
D. L.
2009
Summary of current radiometric calibration coefficients for Landsat MSS, TM, ETM + , and EO-1 ALI 14.Sensors
.
Remote Sensing of Environment
113
(
5
),
893
903
.
Davey
M. K.
1989
A simple tropical moist model applied to the ‘40-day’ wave
.
Quarterly Journal of the Royal Meteorological Society
115
(
489
),
1071
1107
.
Estoque
R. C.
&
Murayama
Y.
2017
Monitoring surface urban heat island formation in a tropical mountain city using Landsat data (1987–2015)
.
ISPRS Journal of Photogrammetry and Remote Sensing
2017
(
133
),
18
29
.
Garreaud
R. D.
,
Boisier
J. P.
,
Rondanelli
R.
,
Montecinos
A.
,
Sepúlveda
H. H.
&
Veloso-Aguila
D.
2019
The central Chile mega drought (2010–2018): A climate dynamics perspective
.
International Journal of Climatology, Royal Meteorological Society
https://doi.org/10.1002/joc.6219
.
Hendon
H. H.
,
Zhang
C.
&
Glick
J. D.
1999
Interannual variation of the Madden–Julian oscillation during austral summer
.
Journal of Climate
12
(
8
),
2538
2550
.
Houk
P.
,
Yu
J.-Y.
&
Qian
C.
2017
Why were 2015/2016 and 1997/1998 extreme El Niños different?
Geophysical Research Letters
44
,
1848
1856
.
doi:10.1002/2016GL071515
.
Houk
P.
,
Yalon
A.
,
Maxin
S.
,
Starsinic
C.
,
McInnis
A.
,
Gouezo
M.
,
Golbuu
Y.
&
Van Woesik
R.
2020
Predicting coral-reef futures from El Niño and Pacific Decadal Oscillation events
.
Scientific Reports
,
10
(
1
),
7735
.
Huang
Z.
,
Zhang
W.
,
Liu
C.
&
Stuecker
M. F.
2022
Extreme Indian Ocean dipole events associated with El Niño and Madden–Julian oscillation
.
Climate Dynamics
59
(
7–8
),
1953
1968
.
Huang
B.
,
L'Heureux
M.
,
Hu
Z. Z.
&
Zhang
H. M.
2016
Ranking the strongest ENSO events while incorporating SST uncertainty
.
Geophysical Research Letters
43
(
17
),
9165
9172
.
Hua
T. M.
,
Hui
R. C. Y.
&
Husen
R.
2013
Trends of rainfall in Sarawak from 1999 to 2008
.
Proc., The International Conf. on Social Science Research
,
Penang, WorldConferences.net
,
261
269
.
Ibadullah
W. M. W.
,
Tangang
F.
,
Juneng
L.
&
Jamaluddin
A. F.
2019
Practical predictability of the 17 December 2014 heavy rainfall event over East Coast of Peninsular Malaysia using WRF model
.
Sains Malaysiana
,
48
(
11
),
2297
-
2306
.
Kayano
M. T.
,
Cerón
W. L.
,
Andreoli
R. V.
,
Souza
R. A.
,
Souza
I. P.
&
Canchala
T.
2021
El Niño-Southern Oscillation and Indian Ocean Dipole modes: Their effects on South American rainfall during austral spring
.
Atmosphere
12
(
11
),
1437
.
Kemarau
R. A.
&
Eboy
O. V.
2019
Landscape Indices and It Impacts To Land Surface Temperature On Small Medium Size City For the Year 1991, 2011 and 2018: Case Study Kota Kinabalu
.
Journal of Built Environment, Technology, and Engineering
7
.
Kemarau
R. A.
&
Eboy
O. V.
2020
Analyses water bodies effect in mitigation of urban heat effect: Case study small size cities Kuching, Sarawak
.
In IOP Conference Series: Earth and Environmental Science
(Vol.
540
, No.
1
, p.
012010
).
IOP Publishing.
Kemarau
R. A.
&
Eboy
O. V.
2021b
Spatial-temporal distribution of malaria risk and its association with El Niño Southern Oscillation (ENSO)
.
Malaysian Journal of Social Sciences and Humanities (MJSSH)
6
(
4
),
276
286
.
Kemarau
R. A.
&
Eboy
O. V.
2021c
Land cover change detection in Kuching, Malaysia using satellite imagery
.
Borneo Journal of Sciences & Technology
3
(
1
),
61
65
.
Kessler
W. S.
&
Kleeman
R.
2000
Rectification of the Madden–Julian oscillation into the ENSO cycle
.
Journal of Climate
13
(
20
),
3560
3575
.
Kovats
R. S.
,
Bouma
M. J.
,
Hajat
S.
,
Worrall
E.
&
Haines
A.
2003
El Niño and health
.
The Lancet
362
(
9394
),
1481
1489
.
Lenka
S.
,
Devi
R.
,
Joseph
C. M.
&
Gouda
K. C.
2022
Effect of large-scale oceanic and atmospheric processes on the Indian summer monsoon
.
Theoretical and Applied Climatology
147
(
3
),
1561
1576
.
Li
J.
,
Song
C.
,
Cao
L.
,
Zhu
F.
,
Meng
X.
&
Wu
J.
2011
Impacts of landscape structure on surface urban heat islands: A case study of Shanghai, China
.
Remote Sensing of Environment
2011
(
115
),
3249
3263
.
Mahmud
M.
2018
Peristiwa El Nino dan pengaruh IOD terhadap hujan di Malaysia
.
e-BANGI
13
(
2
),
166
177
.
Moura
M. M.
,
Dos Santos
A. R.
,
Pezzopane
J. E. M.
,
Alexandre
R. S.
,
da Silva
S. F.
,
Pimentel
S. M.
,
de Andrade
M. S. S.
,
Silva
F. G. R.
,
Branco
E. R. F.
,
Moreira
T. R.
,
da Silva
R. G.
&
de Carvalho
J. R.
2019
Relation of El Niño and La Niña phenomena to precipitation, evapotranspiration, and temperature in the Amazon basin
.
Science of the Total Environment
651
,
1639
1651
.
National Oceanic Atmospheric Administration
.
Nicholls
N.
1989
Sea surface temperatures and Australian winter rainfall
.
Journal of Climate
2
(
9
),
965
973
.
Panisset
J. S.
,
Gouveia
C. M.
,
Libonati
R.
,
Peres
L.
&
Machado-silva
F.
2017
Evaluation of last extreme drought events in the Amazon basin using remote sensing data
. In:
EGU 19
, p.
16144
.
Parker
D. E.
2010
Urban heat island effects on estimates of observed climate change
.
Wiley Interdisciplinary Reviews: Climate Change
1
(
1
),
123
133
.
http://dx.doi.org/10.1002/wcc.021
.
Qin
Z. H.
,
Zhang
M. H.
,
Karnieli
A.
&
Berliner
P.
2001
Mono-window algorithm for retrieving land surface temperature from Landsat TM6 data
.
Acta Geographica Sinica
56
(
4
),
456
466
.
Ratna
S. B.
,
Cherchi
A.
,
Osborn
T. J.
,
Joshi
M.
&
Uppara
U.
2021
The extremely positive Indian Ocean dipole of 2019 and associated Indian summer monsoon rainfall response
.
Geophysical Research Letters
48
(
2
),
e2020GL091497
.
Ropelewski
C. F.
&
Halpert
M. S.
1987
Global and regional scale precipitation patterns associated with the El Niño/Southern Oscillation
.
Monthly Weather Review
115
(
8
),
1606
1626
.
Sa'adi
Z.
,
Shahid
S.
,
Ismail
T.
,
Chung
E. S.
&
Wang
X. J.
2019
Trends analysis of rainfall and rainfall extremes in Sarawak, Malaysia using modified Mann–Kendall test
.
Meteorology and Atmospheric Physics
131
,
263
277
.
Schulte
J.
,
Policelli
F.
&
Zaitchik
B.
2021
A continuum approach to understanding changes in the ENSO–Indian monsoon relationship
.
Journal of Climate
34
(
4
),
1549
1561
.
Sena
A. C.
&
Magnusdottir
G.
2021
Influence of the Indian Ocean dipole on the large-scale circulation in South America
.
Journal of Climate
34
(
15
),
6057
6068
.
Sharma
K. V.
,
Khandelwal
S.
&
Kaul
N.
2020
Comparative assessment of vegetation indices in downscaling of MODIS satellite land surface temperature
.
Remote Sensing in Earth Systems Sciences
3
,
156
167
.
Sharma
K. V.
,
Kumar
V.
,
Gautam
L.
,
Choudhary
S.
&
Mathew
A.
2023
Geo-physical seasonal deviations of land use, terrain analysis, and water cooling effect on the surface temperature of Pune city
.
Journal of Water and Climate Change
14
(
12
),
4802
4820
.
Slingo
J. M.
,
Rowell
D. P.
,
Sperber
K. R.
&
Nortley
F.
1999
On the predictability of the interannual behavior of the Madden-Julian Oscillation and its relationship with El Niño
.
Quarterly Journal of the Royal Meteorological Society
125
(
554
),
583
609
.
Sun
X.
,
Renard
B.
,
Thyer
M.
,
Westra
S.
&
Lang
M.
2015
A global analysis of the asymmetric effect of ENSO on extreme precipitation
.
Journal of Hydrology
530
,
51
65
.
Tan
M. L.
,
Juneng
L.
,
Tangang
F. T.
,
Chung
J. X.
&
Radin Firdaus
R. B.
2020
Changes in temperature extremes and their relationship with ENSO in Malaysia from 1985 to 2018
.
International Journal of Climatology
.
https://doi.org/10.1002/joc.6864
.
Tangang
F. T.
,
Juneng
L.
,
Salimun
E.
,
Sei
K.
&
Halimatun
M.
2012
Climate change and variability over Malaysia: Gaps in science and research information
.
Sains Malaysiana
41
(
11
),
1355
1366
.
Tavakol
A.
,
Rahmani
V.
&
Harrington
J.
Jr
2020
Evaluation of hot temperature extremes and heat waves in the Mississippi River Basin
.
Atmospheric Research
239
,
104907
.
Thirumalai
P. N.
,
DiNunzio
Y. O.
&
Deser
C.
2016
Extreme temperatures in Southeast Asia are caused by El Niño and worsened by global warming
.
Nature Communications
.
doi:10.1038/ncomms15531
.
Tompkins
A. M.
2001
Organization of tropical convection in low vertical wind shears: The role of cold pools
.
Journal of the Atmospheric Sciences
58
(
13
),
1650
1672
.
Trenberth
K. E.
&
Stepaniak
D. P.
2001
Indices of El Niño evolution
.
Journal of Climate
14
(
8
),
1697
1701
.
https://doi.org/10.1175/1520
.
Yik
D. J.
,
Sang
Y. W.
,
Mat Adam
M. K.
,
Chang
N. K.
,
Yunus
F.
&
Abdullah
M. H.
2015
The Definitions of the Southwest Monsoon Climatological Onset and Withdrawal over Malaysian Region
.
JMM Research Publication
, No.
3
/
2015
,
30
pp.
Young
S. J.
,
Johnson
B. R.
&
Hackwell
J. A.
2002
An in-scene method for atmospheric compensation of thermal hyperspectral data
.
Journal of Geophysical Research: Atmospheres
107
(
D24
),
ACH-14
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).