Next Article in Journal
Impact of Wastewater Pollution on Antibiotic Resistance in an Algerian Waterway: A Preliminary Investigation
Previous Article in Journal
Spectral Analysis of Ocean Variability at Helgoland Roads, North Sea: A Time Series Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Atmospheric Processes over the Broader Mediterranean Region 1980–2024: Effect of Volcanoes, Solar Activity, NAO, and ENSO

by
Harry D. Kambezidis
1,2
1
Atmospheric Research Team, Institute of Environmental Research and Sustainable Development, National Observatory of Athens, GR-11810 Athens, Greece
2
Laboratory of Soft Energies and Environmental Protection, Department of Mechanical Engineering, University of West Attica, GR-12241 Athens, Greece
Earth 2025, 6(4), 138; https://doi.org/10.3390/earth6040138
Submission received: 22 August 2025 / Revised: 9 October 2025 / Accepted: 20 October 2025 / Published: 1 November 2025

Abstract

The Mediterranean region is regarded as a hot spot on Earth because of its placement at the junction of many aerosols. Numerous studies have demonstrated that the North Atlantic Oscillation (NAO), which is closely related to the El Niño–Southern Oscillation (ENSO) phenomenon, influences the weather in the area. However, a recent study by the same author examined the ENSO effect on atmospheric processes in this area and discovered a slight but notable influence. This study builds on that earlier work, but it divides the Mediterranean region into four smaller regions during the same time span as the previous study, which is extended by two years, from 1980 to 2024. The division is based on geographical, climatological, and atmospheric process features. The findings demonstrate that volcanic eruptions significantly affect the total amount of aerosols. Additionally, the current study reveals that the Granger-causality test of the physical phenomena of solar activity, ENSO, and NAO indicates that all have a significant impact, either separately or in combination, on the atmospheric process over the four Mediterranean regions, and this effect can last up to six months. Moreover, a taxonomy of the different forms of aerosols across the four subregions is given.

1. Introduction

Electromagnetic energy from the Sun makes up the atmospheric radiation that traverses the Earth’s atmosphere. This radiation is affected by clouds, particles, and gases in the atmosphere [1]. Atmospheric aerosols are solid or liquid particles that flow in the air for periods of hours to days and are crucial to the Earth’s climate system [2,3], and its energy balance because they scatter and absorb solar radiation (direct effect) at both short- and long-wavelengths [4]. They also indirectly affect precipitation and cloud albedo [5,6,7]. Additionally, depending on their size, concentration, chemistry, and height in the atmosphere, atmospheric aerosols have a substantial impact on global climate change [8]. Dust, sea salt, black carbon, and volcanic debris are examples of primary aerosols, whereas sulphates, nitrates, ammonium, and secondary organic compounds are examples of secondary aerosols, which are produced by chemical processes [9]. As a result, atmospheric aerosols come from human activities, natural sources, or a mix of both [10]. Black carbon is primarily an absorbing aerosol, while sulphates, nitrates, ammonium, and sea-salt particles are examples of solely scattering aerosols. The direct role of atmospheric aerosols is related to their capacity to scatter and absorb solar radiation, while the indirect impact is related to their role as cloud condensation nuclei [11], which affects cloud albedo and modifies precipitation patterns in a particular location.
Even though their optical properties are currently well understood, there are still many unknowns regarding the climatic effects of atmospheric aerosols due to their wide variety of types, physicochemical and optical properties, impact on meteorology through weather modification, and atmospheric mixing processes [12]. Additionally, a rise or fall in aerosol concentration can proportionately diminish or increase solar radiation, which is particularly visible in clear-sky conditions because air particles interact with solar radiation through the mechanisms of scattering and absorption. Because of this, atmospheric aerosols are a major source of uncertainty in climate projections [13], and knowledge of them is necessary to make precise predictions about the amount of solar energy available in a given place [14,15].
On Earth, there are numerous large-scale atmospheric and oceanic circulation phenomena that influence weather both locally and globally. Among them, the El Niño–Southern Oscillation (ENSO) and the North Atlantic Oscillation (NAO) play important roles in global and regional climates, respectively. The significance of the Arctic Oscillation (AO), ENSO, and NAO circulations was examined by Yoo et al. [16], who concurred with [17] that ENSO activity is significant since future changes in NAO are anticipated to be more reliant on ENSO. Abish and Mohanakumar [18] demonstrated that during El Niño years, aerosol concentrations over the Indian subcontinent rise noticeably; high aerosol index values (i.e., 1.7) over the area make this clear. European weather visibly and strongly responds to the peak phases of NAO (both positive and negative), according to Laken and Stordal’s [19] analysis of European weather data from the Hess–Brezowsky Large Weather Scenario. On the other hand, these authors discovered no meaningful connection between ENSO and/or solar cycle and European weather. On the contrary, Kirov and Georgieva [20] found a close connection between solar activity and NAO and ENSO circulations. Mezzina et al. [21] underlined that the ENSO phenomenon helps make surface weather forecasts over the North Atlantic’s mid-latitudes more predictable. In coupled climate models, Casselman et al. [22,23,24] also verified the teleconnection between ENSO and NAO circulation patterns. The timing of the link between ENSO and the tropical North Atlantic was re-examined by García-Serrano et al. [25]. In the spring (March–May) of the following year, Liu et al. [26] discovered a strong positive long-term correlation between the tropospheric CO levels over the European sector of the North Atlantic and the winter ENSO event. The South Asian monsoon has a significant impact on the Eastern and Central Mediterranean climates, according to Alpert et al. [27]. Throughout the past 120 years (1901–2020), Urdiales-Flores et al. [28] showed that the Mediterranean basin has warmed more quickly that the rest of the world; they ascribed the greater warmth to the combined impacts of the region’s lower soil moisture and decreased aerosol pollution. Park [29] found a remote influence of ENSO on Mediterranean sky conditions during late summer and autumn. A recent study by Alsubhi and Ali [30] demonstrated the effect of ENSO on the dust outbreaks over the Arabian Peninsula.
Although the teleconnection between the ENSO and NAO circulations has been established as previously shown, there has been little attention paid to the question of whether ENSO may have an impact on atmospheric processes over the Mediterranean (i.e., atmospheric radiation and atmospheric aerosols). The only work on this topic is a recent study by the same author [31] that found a statistically significant relationship between ENSO and Ångström’s exponent, and between ENSO and direct aerosol radiative forcing over the larger Mediterranean region. On the other hand, little research has been conducted on how the ENSO phenomenon affects aerosol particles in the wider area surrounding the event. However, Westervelt et al. [32] stated that regional decreases in aerosol emissions are usually responsible for a shift to the positive ENSO phase (i.e., an El Niño event). Huang et al. [33] found that ENSO-related sea-surface temperature anomalies in the tropical Eastern Pacific have a substantial effect on suspended dust particles in the Arabian Peninsula to the Central Asia region. Banerjee and Kumar [34] have shown that El Niño can change the intensity of dust subsidence across the Indian subcontinent by remotely adjusting the strength of convection over the India–Pacific region. According to Brönnimann et al. [35] and Toniazzo and Scaife [36], both phases of ENSO can influence NAO. Leamon [37] discovered a highly significant association between the occurrence of the termination of the last five solar cycles and the Pacific’s Ocean shift from El Niño to La Niña. Huo et al. [38] studied the influence of solar activity in the North Atlantic and the tropical Pacific regions, but found no solar effect in the simulated sea-level pressure/sea-surface temperature coupled mode.
As far as the Mediterranean is concerned, many researchers think that the region is a hot spot for aerosol events from a number of sources, including air pollution from big cities or industrial areas [39,40], dust from the Sahara Desert [41,42,43,44,45], eruptions from volcanoes [46,47], and wildfires [48]. Figure 1 in [49] provides valuable information about the transport of Sahara dust and air pollutants throughout the Mediterranean region.
There has been some hazy evidence of a direct ENSO influence on the Mediterranean climate, despite the fact that the impact of the ENSO phenomenon on atmospheric processes has not been thoroughly investigated for the Mediterranean region up to now [50]. Conversely, there are studies that look at how ENSO affects weather patterns in many parts of the world. For example, the El Niňo/La Niňa phenomena have been demonstrated to have a significant influence on worldwide weather patterns, including monsoons [51,52,53], and the North Atlantic Ocean close to the east coast of the United States and Western Europe [21,54].
The present study is a continuation of a recent work by the same author [31]. That study examined the Mediterranean region as a unified area; its main conclusions were as follows: (i) positive correlation was found for TAOD550, AE470–870 with ONI or NAOI, and a negative one for DARF, netSWRBOA,CS,A with either ONI or NAOI; (ii) a positive trend for netLWRBOA and a negative trend for COD, TAOD550, SAOD550, and AE470–870 in the period 1980–2024; (iii) the majority of the atmospheric aerosols over the entire Mediterranean region were classified as CMA and UIA types; (iv) there was a noticeable drop in DARF from cold to warm ONI events; and (v) DARF exposed a positive trend with time (1980–2022). (The explanation of the symbols used here are given in Table 2, Section 2). For further research, that article suggested that “a future study might perform analyses on parts of the Mediterranean region in order to achieve better spatial resolution”. This future research is therefore conducted here.
The questions to be investigated are, therefore, as follows: (i) Are the effects of solar activity, NAO, and ENSO on the atmospheric processes found in [31] valid in smaller areas of the Mediterranean too? (ii) How do volcanic eruptions affect the status of atmospheric aerosols over the Mediterranean? To do this, a preliminary condition is posed: the Mediterranean domain considered in [31] is split into four smaller sectors, which are configured in Section 2. Solar activity (in terms of the sunspot number, SSN) has also been added in the present study because some authors have shown a dependence of NAO events on SSN (e.g., [20,55]).
The work is divided into sections. Section 1 gives the necessary state of knowledge on the subject. Section 2 provides the data and the methodology used in this work. Section 3 deploys the results of the study in two parts: Part A. Physical approach; and Part B. Statistical approach. Section 4 provides the main findings of the work and gives thoughts for future research. Finally, Acknowledgements, Appendix A, and References follow.

2. Materials and Methods

As mentioned above, the present work is a logical extension to [31]. The difference between that work and the present study is the investigation of the effects from ENSO and volcanic eruptions on the atmospheric processes over four adjacent regions that make up the broader Mediterranean area used in [31]. These four subregions are shown in Figure 1 and have the following geographical coordinates: West Mediterranean (or WestMed in brevity, 6° W–10° E, 30° N–45° N); Central Mediterranean (CentMed, 10° E–0° E, 30° N–45° N); East Mediterranean (EastMed, 20° E–35° E, 30° N–41° N); and Balkans and Black Sea (BalBSea, 20° E–35° E, 41° N–45° N). The reasoning behind the division of the entire Mediterranean area into four subregions was dictated by the following considerations: (i) WestMed (Spain, South France, Northwest Italy, North Morocco, North Algeria) has distinct Atlantic influence, aerosol inflow from Iberia and Sahara, and high solar flux; (ii) CentMed (Italy, North Tunisia, Malta, West Greece, North Libya) lies in a transitional zone with complex aerosol mixing, volcanic activity, and high solar irradiance; (iii) EastMed (Greece, Turkey, Cyprus, Israel, North Egypt) is characterised by strong eastern aerosol transport, semi-arid climate, intense solar radiation; and (iv) BalBSea (Balkans, Bulgaria, Black Sea) has features of continental climate, unique aerosol pathways, and lower solar insolation. Table 1 provides more details on the above criteria.
To answer the research questions posed in Section 1, the Giovanni platform, which can be accessed without cost at https://giovanni.gsfc.nasa.gov/giovanni (accessed on 8–25 October 2023) [59], provided month-to-month data for the years 1980–2024 (45 years or 4.5 decades) at a spatial resolution of 0.5° × 0.625°. More specifically, for each subregion, the following data were downloaded: netSWR at the bottom-of-the-atmosphere (BOA) with and without aerosols in the atmosphere and under clear-sky conditions, netSWRBOA,CS,A and netSWRBOA,CS,NA, respectively, in Wm−2 (MERRA-2 model; Giovanni files M2TMNXRAD v5.12.4); netSWR at the top-of-the-atmosphere (TOA) with and without aerosols in the atmosphere and under clear-sky conditions, netSWRTOA,CS,A and netSWRTOA,CS,NA, respectively, in Wm−2 (MERRA-2 model; Giovanni files M2TMNXRAD v5.12.4); netLWR at BOA with and without aerosols in the atmosphere and under clear-sky conditions, netLWRBOA,CS,A and netLWRBOA,CS,NA, respectively, in Wm−2 (MERRA-2 model; Giovanni files M2TMNXRAD v5.12.4); netLWR radiation at TOA with and without aerosols in the atmosphere and under clear-sky conditions, netLWRTOA,CS,A and netLWRTOA,CS,NA, respectively, in Wm−2 (MERRA-2 model; Giovanni files M2TMNXRAD v5.12.4); scattering aerosol optical depth at 550 nm, SAOD550 (MERRA-2 model; Giovanni file M2TMNXAER v5.12.4); desert-dust aerosol optical depth at 550 nm, DDAOD550 (MERRA-2 model; Giovanni file M2TMNXAER v5.12.4); total aerosol optical depth at 550 nm, TAOD550 (MERRA-2 model; Giovanni file M2TMNXAER v5.12.4); black-carbon aerosol optical depth at 550 nm, BCAOD550 (MERRA-2 model; Giovanni file M2TMNXAER v5.12.4); stratospheric aerosol optical depth at 550 nm, StAOD550 (MERRA-2 model; Giovanni file M2TMNXRAD v5.12.4); sea-salt aerosol optical depth at 550 nm, SSAOD550 (MERRA-2 model; Giovanni file M2TMNXAER v5.12.4); cloud optical depth, COD (MERRA-2 model; Giovanni file M2TMNXRAD v5.12.4); and Ångström’s exponent in the spectral band 470–870 nm, AE470–870 (MERRA-2 model; Giovanni file M2TMNXAER v5.12.4). Although the MERRA-2 scientific team in charge evaluates the data for quality before publishing it on the Giovanni website, further quality checks were performed on the downloaded data to look for outliers, missing, and out-of-range values; however, none were discovered.
As far as the accuracy in the estimation of AOD by the MERRA-2 reanalysis is concerned, Shaheen et al. [60], in a study of AOD trends over the East Mediterranean in the period 2000–2018, concluded the following: (i) MERRA-2 AOD estimates show a systematic low bias compared to AERONET measurements, particularly under high-AOD conditions associated with desert dust; and (ii) MODIS observations perform better during episodic events but have larger spread in monthly mean values. On the other hand, Su et al. [61], analysing AOD data from MERRA-2, MODIS, and AERONET in the period 1980–2023, found that MERRA-2 AOD estimates had the greatest accuracy with an expected error ±0.05 ± 20% compared to AERONET measurements. Mukkavilli et al. [62] used AOD data over Australia in the period 2002–2012 taken from NASA’s MERRA-2 and the Monitoring Atmospheric Composition and Climate (MACC) programme of the European Centre for Medium-Range Weather Forecasts (ECMWF), evaluating them against concurrent ground-based measurements from the Aerosol characterisation via Sun photometry: Australian Network (AeroSpan), which is part of the global Aerosol Robotic Network (AERONET) and found that MERRA-2 reanalysis overestimates monthly AOD twice as much compared to AeroSpan/AERONET ground observations but corelates better with AeroSpan/AERONET than ECMWF/MACC. Khatibi and Krauter [63] evaluated MERRA-2 air temperature, wind speed, and solar radiation datasets against measured data from Meteonorm and Deutscher Wetterdienst (DWD, Germany) over eight stations around the world and found high correlation coefficients from 0.95 to 0.99 for solar radiation, 0.99 for air temperature, and from 0.81 to 0.99 for wind speed. They concluded that MERRA-2 datasets are valuable, considering their global coverage and availability.
In addition to the above datasets, monthly mean data was downloaded during the same time period for the sunspot number, SSN, the North Atlantic Oscillation Index, NAOI, and the Oceanic El Niño Index 3.4, ONI3.4. For more information about the indices, the reader is referred to Section 2 in [31] and Table 2 in the present work. Moreover, it is important to mention that the SSN, NAOI, and ONI data is not site-specific.
The netSWR and netLWR data was used to determine DARF in the atmosphere. This study has taken into account clear-sky (CS) conditions because, as said in the Introduction, “…a rise or fall in aerosol concentration can proportionally reduce or increase solar radiation, especially noticeable under clear-sky conditions”. The following are the net-radiation expressions [1]:
DARF = (QA − QNA)TOA − (QA − QNA)BOA,
QA = netSWRA − netLWRA,
QNA = netSWRNA − netLWRNA,
netSWR or netLWR = F↓ − F↑.
In the equations above, Q is the difference between netSWR and netLWR, whereas netSWR or netLWR is the difference between the downward (F↓) minus the upward (F↑) radiation fluxes (by convention, downward fluxes are positive and upward fluxes are negative). Aerosol-laden and aerosol-free atmospheres are denoted by the subscripts A and NA, respectively. In all computations, the four Mediterranean subregions shown in Figure 1 have been regarded as distinct entities; that is, all examined parameter values are time series of monthly data averaged over the selected Mediterranean domains between 1980 and 2024.
The absorption, scattering, and total aerosol optical depths at the wavelength of 550 nm are interrelated through the equation [1]:
TAOD550 = SAOD550 + AAOD550.
Since AAOD550 values are not available on the Giovanni platform, they were computed from Equation (5) by subtracting SAOD550 from TAOD550. Additionally, the single-scattering albedo, SSA, was estimated by the following equation:
SSA = SAOD550/TAOD550.
In Equation (5), light-scattering entities in the atmosphere are considered the desert-dust particles, volcano-related stratospheric aerosols, sea-salt, and, partially, clouds as they may have either a scattering or an absorbing role (e.g., [64,65,66]). Therefore, SAOD550 in Equation (5) can be replaced with the sum of the optical depths of desert-dust aerosols, stratospheric aerosols, sea-salt, and clouds. As far as the inclusion of COD in Equation (5) is concerned, this was performed by dividing COD by 1000 in order to have a compatible magnitude with those of other optical depths. Moreover, because of the clouds’ double role as scatterers or absorbers, one can consider that COD = CODsca + CODabs, where CODsca = 0.85∙COD refers to their scattering property and CODabs = 0.15∙COD to their absorbing role (e.g., [64]). These coefficients represent globally averaged fractions derived from radiative transfer studies; while they may vary depending on cloud microphysics and wavelength, they offer a reasonable first-order estimation for climatological analysis [67]. Thus, the scattering and absorbing CODs to be included in Equation (5) can be expressed as follows:
CODsca = 0.85∙(COD/1000),
CODabs = 0.15∙(COD/1000).
Therefore, SAOD550 and AAOD550 can be rewritten as follows:
SAOD550 = DDAOD550 + StAOD550 + SSAOD550 + CODsca + OSAOD550,
AAOD550 = BCAOD550 + CODabs + OAAOD550.
In Equation (8), the latter parameter implies the optical depth of other scattering particles not explicitly designated here (cf. nitrates, ammonium salts, secondary organic aerosols, secondary aerosols). In Equation (9), the absorbing aerosols are black carbon and clouds; the latter parameter means the optical depth of other absorbing particles not explicitly referred to here (cf. brown carbon, soot aggregates). An approximation of TAOD550 in Equation (5) can thus be rewritten as TAOD’550 by taking into consideration Equations (8) and (9):
TAOD’550 ≈ (DDAOD550 + StAOD550 + SSAOD550 + CODsca) + (BCAOD550 + CODabs).
In Equation (10), the first parenthesis refers to SAOD550 and the second to AAOD550; both OSAOD550 and OAAOD550 have been omitted. According to the above, the downloaded TAOD550 can be considered as the observed values, while TAOD’550 may refer to the estimated TAOD550 time series. This is said because TAOD550 comes directly from MERRA-2 reanalysis calculations (cf. observations) and TAOD’550 is the summation (cf. estimation) of the optical depths of atmospheric aerosols.
NOAA uses the ONI3.4 standard to classify El Niño (warm) and La Niña (cold) episodes in the East Pacific Ocean (https://cpc.ncep.noaa.gov/products/analysis_monitoring/ensostuff/ONI_v5.php (accessed on 8–25 October 2023)). This index [68] is used to monitor the oceanic part of the ENSO climate pattern. ONI3.4 is computed as a running 3-month mean sea-surface temperature (SST) anomaly for the El Niño region (5° S–5° N, 120° W–170° W). If the SST anomaly for five consecutive overlapping 3-month periods is >+0.5 °C, an event is considered warm; if it is <−0.5 °C, it is considered cold.
The sea-level pressure difference between the subpolar Low over Iceland and the subtropical High at the Azores serves as the basis for NAOI. When NAOI is positive, the central region of the North Atlantic Ocean, the East United States, and West Europe are experiencing above-normal pressures, whereas the high latitudes of the North Atlantic are experiencing below-normal pressures [69]. The Gulf Stream, which starts in the Gulf of Mexico and moves northward across the Atlantic, is solely linked to the NAO phenomenon. Temperatures above normal in the East United States and North Europe and below normal in Greenland and frequently in South Europe and the Middle East are associated with high positive NAOI phases. They provide below-normal precipitation patterns throughout South and Central Europe and above-normal precipitation over North Europe and Scandinavia [70]. Temperature and precipitation patterns during severe adverse (negative) NAOI phases are opposite of those during significant positive NAO occurrences. The NAOI calculation procedure is explained in [71].
The Earth’s energy budget is determined by COD, a fundamental cloud feature that comes from ground-based or satellite observations (e.g., [64,65,66]).
In the Ångström’s wavelength-dependent formula for characterising how atmospheric aerosols attenuate solar light, AE is the exponent, which can be calculated through the following equation [1]:
TAODλ = β·λAE
where λ is the wavelength (in nm). The present work has taken AE in the visible wavelength band, i.e., 470–870 nm. The value of AE = 1.3 is commonly considered in the international literature; above this threshold, aerosols are characterised as fine, and below this value, aerosol particles are characterised as coarse.
The SSN is associated with the 11-year solar activity. Number 25 is the current solar cycle, which began in December 2019 and is still running. The examined period from 1980 to 2024 includes solar cycles 22–25.
Table A1 (see Appendix A) gives a list of the energetic volcanoes in the period of the study with VEI (volcanic explosivity index) magnitudes 4–6 taken from the Smithsonian Institution, USA (https://volcano.si.edu). VEI is a measure of the size of explosive volcanic eruptions and was defined by [72]. The VEI scale spans between 0 and 8; VEIs up to 6 have the following definitions: VEI = 0, category: effusive, column height: <100 m, ejecta volume: <104 m3, periodicity: constant, tropospheric/stratospheric circulation: negligible/none; VEI = 1, category: gentle, column height: 100 m−1 km, ejecta volume: >104 m3, periodicity: daily, tropospheric/stratospheric circulation: minor/none; VEI = 2, category: explosive, column height: 1–5 km, ejecta volume: >106 m3, periodicity: 2 weeks, tropospheric/stratospheric circulation: moderate/none; VEI = 3, category: severe, column height: 3–15 km, ejecta volume: >107 m3, periodicity: 3 months, tropospheric/stratospheric circulation: substantial/possible; VEI = 4, category: catastrophic, column height: >10 km, ejecta volume: >0.1 km3, periodicity: 18 months, tropospheric/stratospheric circulation: substantial/definite; VEI = 5, category: cataclysmic, column height: >10 km, ejecta volume: >1 km3, periodicity: 12 years, tropospheric/stratospheric circulation: substantial/significant; VEI = 6, category: colossal, column height: >20 km, ejecta volume: >10 km3, periodicity: 50–100 years, tropospheric/stratospheric circulation: substantial/substantial. The selection of volcanoes with VEI ≥ 4 was based on their severity and the potential circulation of the debris in the troposphere/stratosphere. An exception was made for the volcanoes Etna and Stromboli in Italy [73], which showed VEIs between 2 and 3 in the period of the study. This exception is justified by the fact that both volcanoes are within the studied region; no matter how low their activity is, they may affect atmospheric loading and composition in the CentMed area [74].
Table 2 provides a summary of the parameters examined in this study, along with relevant commentary to provide an overview of the variables. It is important to note that this list includes factors like solar fluxes and clouds that do not represent aerosol properties and large-scale circulation patterns; nevertheless, they are directly related to aerosol properties, and it is therefore desirable to examine how they depend on SSN, NAO, and ENSO.

3. Results

This section is divided into two parts: the first part examines the physical properties of the aerosols and atmospheric parameters mentioned in Section 2, while the second part investigates the relationships among the parameters using statistical procedures.

3.1. Part A—Physical Approach

3.1.1. Observed and Estimated Aerosol Optical Depth

The definitions of the observed and estimated TAOD are provided and elucidated by Equations (5) and (10), respectively. To demonstrate the accuracy, Figure 2 illustrates the annual mean values of the estimated TAOD’550 and the observed TAOD550, along with their percentage difference, defined as follows:
ΔTAOD550 (%) = 100·(TAOD550 − TAOD’550)/TAOD550.
The plots in Figure 2 indicate that in all four Mediterranean subregions, the mean errors ΔTAOD550 are mainly within the band ±6% of the TAOD550 values or lower; more specifically, the mean ± std of ΔTAOD550 are as follows: −0.247% ± 2.218 (WestMed), −2.019% ± 4.121 (CentMed), −2.858% ± 2.303 (EastMed), and +0.488% ± 2.796 (BalBSea), which indicate that TAOD’550 overestimates TAOD550 in WestMed, CentMed, and EastMed, and underestimates it over BalBSea. Nevertheless, the error introduced is very low, as it lies in the band (−2.86%, +0.49%), and the approximation of TAOD’550 to TAOD550 can therefore be considered accurate. This means that the estimation of the total aerosol extinction factor can be derived from the summation of the solar radiation attenuations outlined in Equation (10), in case of TAOD550 unavailability (see also Figure 2). It is noteworthy that the ΔTAOD550 values are negative (i.e., TAOD550 < TAOD’550) in WestMed, CentMed, and EastMed, and positive (i.e., TAOD550 > TAOD’550) only over BalBSea. This observation can be attributed to the fact that TAOD550 intrinsically contains the effect of clouds (i.e., COD), which is found to have a lesser real impact (lower cloudiness) over the WestMed, CentMed, and EastMed subregions than the simulated TAOD’550. In fact, the central and eastern areas of the Mediterranean region present fewer clouds, according to a study on cloudiness over the Mediterranean [75]. Despite the procedure for estimating TAOD’550, the present work uses the observed TAOD550 values in all calculations.
The observed TAOD550 values from Figure 2 during the research period over all four Mediterranean subregions are reproduced in Figure 3. In this graph, linear regression fits to the TAOD550 annual time series are presented, having the following expressions, significant at the 99.9% confidence interval (CI): TAOD550 = −1.6 × 10−3·t + 3.439, R2 = 0.401 (WestMed), TAOD550 = −2.2 × 10−3·t + 4.268, R2 = 0.510 (CentMed), TAOD550 = −1.2 × 10−3·t + 2.715, R2 = 0.272 (EastMed), and TAOD550 = −3.6 × 10−3·t + 7.503, R2 = 0.738 (BalBSea). The EastMed’s lower R2 score indicates that the linear regression model poorly captures the relationship between t and TAOD550, at least in relation to the expressions for the other three subregions. However, it is noteworthy that the EastMed’s R2 is greater than those for the other three areas when considering the TAOD550,WVE values (see Figure 5, Section 3.1.2).

3.1.2. Effect of Volcanoes

Strong volcanic eruptions that release sulphur into the stratosphere are the main cause of stratospheric aerosols [76]. Because of the massive eruptions from Mt. El Chichón and Mt. Pinatubo volcanoes in 1982/1983 and 1992/1993, respectively, the impact of volcanic debris cycling in the atmosphere is clearly visible in Figure 3. Sulphating particles make up the majority of the debris left over during volcanic eruptions. The two large StAOD550 peaks in Figure 4 in 1982/1983 and 1992/1993, as a result of the eruptions from the aforementioned volcanoes, demonstrate the importance of these particles in the global atmospheric aerosol loading.
The dashed lines in Figure 4 are linear fits to the StAOD550 data series with the following expressions: StAOD550 = −0.19 × 10−2·t + 3.786, R2 = 0.434 (WestMed), StAOD550 = −0.26 × 10−2·t + 542042, R2 = 0.588 (CentMed), StAOD550 = −0.17 × 10−2·t + 3.565, R2 = 0.399 (EastMed), and StAOD550 = −0.38 × 10−2·t + 7.823, R2 = 0.750 (BalBSea). All regressions are significant at the 99.9% CI for the period of the study, where t is a year in the period 1980–2024. The lower R2 score for EastMed is again noticeable and may be attributed to two factors: (i) EastMed is a transitional zone between continental, maritime, and desert influences, it receives episodic inputs from Saharan dust and volcanic plumes, which vary irregularly over time [77]; (ii) stratospheric aerosol transport is modulated by Quasi-Biennial Oscillation (QBO), Brewer–Dobson circulation, and tropopause height variability, which may affect EastMed more erratically than other regions [78]. On the other hand, the subregion is subject to complex synoptic systems and shifting wind regimes (e.g., Etesians, Sirocco), making aerosol transport highly variable and less predictable by a simple time-based model [79].
As seen in Figure 4, StAOD550 falls between 0.03 and 0.3 in each of the four Mediterranean subregions. In a study on stratospheric aerosols from 1850 to 1990, Sato et al. [80] discovered global StAOD550 values ranging from 0.001 to 0.125 (see their Table 2), quite comparable with those of the present study. According to the authors, the significant eruption of the Indonesian Krakatau volcano in August 1883 (VEI = 6) caused the high StAOD550 value of 0.125 in 1885, which affected the globe for over two years.
The volcanic debris that circles the Earth at stratospheric heights and drastically scatters solar radiation obscures all other aerosol impacts. As a result, the sharp decline in TAOD550 in all four Mediterranean subregions is plasmatic (see Figure 2 and Figure 3). This issue becomes clear by removing StAOD550 from TAOD550:
TAOD550,WVE = TAOD550 − StAOD550,
where WVE = without volcanic eruptions. The result clearly demonstrates the impact of both natural and man-made aerosols; TAOD550,WVE then represents the aerosol loading in the atmosphere free of volcanic debris. To graphically show this result, Figure 3 is converted to Figure 5, which shows TAOD550,WVE in the four Mediterranean areas.
The annual average TAOD550 values in the entire 45-year period are as follows: 0.203 ± 0.034 (WestMed), 0.228 ± 0.037 (CentMed), 0.205 ± 0.031 (EastMed), and 0.234 ± 0.056 (BalBSea); the corresponding TAOD550,WVE values are as follows: 0.135 ± 0.013 (WestMed), 0.135 ± 0.017 (CentMed), 0.117 ± 0.014 (EastMed), and 0.084 ± 0.010 (BalBSea). From these values, one can easily conclude that the average TAOD550,WVE levels are 33.5%, 40.8%, 42.9%, and 64.1% lower than the corresponding TAOD550 in the four Mediterranean subregions, respectively. This result demonstrates unequivocally how volcanic particles outweigh the total atmospheric aerosol loading in the larger Mediterranean region. This phenomenon is thought to be global, but further relevant research is required to prove it. The comparable average values of TAOD550 (as well as of TAOD550,WVE) across the three Mediterranean subregions (WestMed, CentMed, and EastMed), and their almost half values over the BalBSea region, are another finding of the analysis, which indicates that global volcanic eruptions do not necessarily affect Mediterranean aerosols in a uniform manner.
The aforementioned observations immediately lead to the conclusion that there is no discernible trend in aerosols for a volcanic-debris-free atmosphere over the research period. In fact, TAOD550,WVE = 2 × 10−4·t − 0.346, R2 = 0.061 (WestMed), TAOD550,WVE = 5 × 10−4·t − 0.938, R2 = 0.179 (CentMed), TAOD550,WVE = 5 × 10−4·t − 0.924, R2 = 0.238 (EastMed), and TAOD550,WVE = 2 × 10−4·t − 0.321, R2 = 0.074 (BalBSea) are the linear regression fits to the annual TAOD550,WVE values (see Figure 5). All regressions are insignificant at the 95% CI with the exception for EastMed, which is significant at the 99.9% CI and has a higher R2 score; t denotes any year between 1980 and 2024. The slope is almost negligible (of the order of 10−4) in all linear regression expressions, resulting in a trend that is not significant throughout the 45-year period. Therefore, in regard to the climate change issue, it would be very interesting if climatologists could research how a (hypothetical) atmosphere devoid of volcanic debris might affect the Earth’s climate.

3.1.3. Effect of NAO and ENSO

In a recent study, Serykh [81] reaffirmed the teleconnection between the El Niño–Southern Oscillation and the tropics of the Indian and Atlantic Oceans in temperate and high geographical latitudes; this was also performed in other research works (e.g., [21,22,23]). Therefore, to show the effect of both the NAO and ENSO phenomena on TAOD550, AE470–870, DARF, SSA, netSWRBOA,CS,A, and netLWRBOA,CS,A, 3D contour plots have been prepared.
Figure 6 displays TAOD550 as a function of ONI and NAOI in the four Mediterranean areas. It is clear that WestMed, CentMed, and EastMed favour higher TAOD550 values in neutral/positive ONI and NAOI events (see Figure 6a, Figure 6b and Figure 6c, respectively). In contrast, the Balkans and Black Sea subregion may accommodate greater TAOD550 values for negative ONI and positive NAOI phases (see Figure 6d).
In South Europe, below-normal precipitation patterns are often linked to positive NAO episodes, as shown in Section 2 [24]. The decreased washout impact of the atmospheric aerosols over the Mediterranean region (South Europe) in positive NAO phases explains the TAOD550 patterns in Figure 6. The sole difference is that, unlike the other three areas, the BalBSea subregion has high TAOD550 values for ONI ≤ 0 rather than ONI ≥ 0. Thus, the question that arises is why the ENSO effect reverses over the BalBSea area.
Quite opposite patterns to those of TAOD550 were derived for DARF, almost symmetrical to the (NAOI, ONI) = (0, 0) point. Higher DARF values (higher warming effect) are shown over WestMed and CentMed under both negative ONI and NAOI phases (see Figure 7a and Figure 7b, respectively). The patterns over EastMed and BalBSea are more complex; EastMed has a secondary peak in the neutral zones of both ONI and NAOI (see Figure 7c), whereas BalBSea presents a spread pattern with high DARF values within the neutral zones of both ONI and NAOI, as well as in the neutral ONI/positive NAOI bands, and in the neutral NAOI/positive ONI zone (see Figure 7d). The positive DARF values (warming effect) vary between 1 Wm−2 and 7 Wm−2, being higher over WestMed and CentMed and lower in EastMed and BalBSea.
For all four Mediterranean subregions, including BalBSea, the AE470–870 versus ONI and NAOI patterns (not shown here) were found to be similar to those of TAOD550. The difference from TAOD550 is that greater AE470–870 values occur for ONI ≥ 0 and NAOI ≤ 0. This new result indicates that over all four Mediterranean subregions, positive ONI and positive/negative NAOI phases favour fine-mode aerosols (AE470–870 > 1.3).
In all four Mediterranean domains, the 3D plots for SSA versus ONI and NAOI (not shown here) display broad peaked values in the upper left corner, i.e., positive/neutral ONI and negative NAOI values. In general, an aerosol scattering profile over the larger Mediterranean region with SSA values greater than 0.90 emerges.
The netSWRBOA,CS,A parameter yielded results that were strikingly comparable to the SSA patterns (not displayed here). This suggests that the scattering aerosol particles are the main source of solar radiation attenuation.
Higher values for the netLWRBOA,CS,A parameter appear symmetrically to the NAOI = 0 axis compared with those for the netSWRBOA,CS,A parameter (not displayed here) for WestMed, CentMed, and EastMed. These high values are found in the contour plot’s lower left corner, specifically for both negative ONI/NAOI phases.
Figure 8 shows the annual TAOD550 values as a function of ONI in the four selected subregions of the Mediterranean. The graph shows that the majority of the TAOD550 data points in all four areas lies within the neutral ONI band, fewer cases in the cold ONI zone, and less in the warm band, i.e., 11, 27, and 7 years (or 155, 324, and 125 months) for cold, neutral, and warm ENSO phases, respectively. This result is true for all the parameters considered in this study. In other words, ENSO seems to regulate the atmospheric aerosol circulation over the broader Mediterranean region. Similar hints have been given in a study by Westervelt et al. [32], which showed that positive ENSO phases (i.e., El Niño events) are associated by regional reductions in aerosol emissions. Also, Huang et al. [33] showed that suspended dust particles over the region from the Arabian Peninsula to Central Asia are significantly impacted by ENSO-related SST anomalies in the tropical Eastern Pacific. Moreover, Banerjee and Kumar [34] demonstrated that, by remotely altering the strength of convection across the India–Pacific region, El Niño can modify the intensity of dust subsidence over the Indian subcontinent. Nevertheless, an immediate conclusion from Figure 8 is that the aerosol loading over the broader Mediterranean area is more frequent during neutral ENSO phases in the majority of the cases (27 years out of the 45-year period in each Mediterranean area).

3.1.4. Effect of Desert Dust and Black Carbon

One may wonder how anthropogenic black-carbon aerosols and desert-dust particles impact the overall aerosol loading over the wider Mediterranean region. This section answers this question. To give an initial overview of both parameters, Figure 9 shows the TAOD550 and TAOD550,WDDBC curves (WDDBC = without desert-dust and black-carbon aerosols). These graphs show how the temporal history of the TAOD550 across the research period is exactly mirrored by the TAOD550 and TAOD550,WDDBC values across the four Mediterranean subregions. The maxima in TAOD550 and StAOD550 in the years 1982/1983 and 1992/1993 in Figure 3 and Figure 4, respectively, provide confirmation that StAOD550, SSAOD550, and COD are the primary contributing elements to the total TAOD550 after desert-dust and black-carbon aerosols are excluded.
To work out the contribution of TAOD550,WDDBC to the total TAOD550, one could find the ratio of the areas under the TAOD550,WDDBC curves to those under TAOD550, expressed as a percentage. To make things easier, instead of computing the areas under both curves of TAOD550 and TAOD550,WDDBC, one can convert the areas under the curves into equivalent orthogonal parallelograms by calculating their temporal averages in the whole period of the study. Then, the ratios of the areas of these orthogonal parallelograms are equal to those of the areas under the real curves. For these calculations, the annual averages of TAOD550 and TAOD550,WDDBC are estimated at 0.203 ± 0.034 and 0.094 ± 0.035 (WestMed), 0.228 ± 0.037 and 0.128 ± 0.039 (CentMed), 0.205 ± 0.031 and 0.121 ± 0.034 (EastMed), and 0.234 ± 0.056 and 0.183 ± 0.057 (BalBSea), respectively. The expression 100∙(TAOD550 − TAOD550,WDDBC)/TAOD550) (in %) represents the contribution of all other aerosols (i.e., StAOD550, SSAOD550, and COD) to the total TAOD550, which is 53.7% (WestMed), 43.9% (CentMed), 41.0% (EastMed), and 21.8% (BalBSea). By subtracting these percentages from 100%, one can derive the contribution of DDAOD550/BCAOD550 to the total TAOD550, which is 46.3% (WestMed), 56.1% (CentMed), 59.0% (EastMed), and 78.2% (BalBSea). These figures clearly show that there is an almost 50%/50% balance between the desert-dust/black-carbon aerosols and the stratospheric/sea-salt/cloud effects in the WestMed, CentMed, and EastMed areas. At first glance, the BalBSea subregion seems to be affected more by desert-dust and black-carbon aerosols than by volcanic eruptions. Nevertheless, desert dust is not generally present over BalBSea; this justifies the much lower TAOD550,WVE average over BalBSea than those over the other three areas (see Section 3.2). This imbalance over BalBSea may assign its different patterns shown in Figure 6 and Figure 7 in comparison to those in the other three Mediterranean subregions.

3.1.5. Examination of Some Individual Aerosol Parameters

Figure 10 shows how TAOD550, AAOD550, SAOD550, and DDAOD550 relied on ONI during the study period across the four Mediterranean subregions. DDAOD550 is linked to the movement of desert dust, primarily from the Sahara Desert, throughout the Mediterranean region [82]. A lower concentration of absorbing aerosols over the Mediterranean during the studied period is inferred from the graph’s very low AAOD550 values, which indicate low attenuation of solar radiation by absorbing aerosols. This is in contrast to the higher concentration of scattering aerosols (higher SAOD550 values). Therefore, the main factor reducing solar radiation as it passes through the atmosphere over the larger Mediterranean region is the scattering process. This result suggests that the only factors lowering solar irradiance throughout the Mediterranean are scatterers such as nitrates, sea-salt, and sulphate particles [9]. In a similar investigation for the Mediterranean as a whole, conducted during nearly the same experimental period as the current work, Kambezidis [31] made the same observation. Moreover, in a study on Greece’s climate and solar radiation, the same author [83] came to the opposite conclusion for the country between 2005 and 2016. This discrepancy was highlighted in [31], indicating that the Mediterranean needs to be examined in smaller sections, as is the case in the current study, in order to completely comprehend the aerosol influence throughout the region. Additionally, because dust aerosols appear to both scatter and absorb solar radiation (depending on the size of their particles), DDAOD550, which is located between SAOD550 and AAOD550 (see Figure 10), displays a Janus face, that is, one scattering and one absorbing face [84]. The levels of DDAOD550 approach those of AAOD550 when one transitions from WestMed to BalBSea, suggesting that the absorbing “face” of DDAOD550 may be taking the central stage. The methodology described in Kambezidis [31] can easily yield some interesting results when considering the averages of TAOD550, AAOD550, SAOD550, and DDAOD550 over each of the four subregions in the 45-year period (WestMed: 0.203 ± 0.034, 0.012 ± 0.001, 0.191 ± 0.034, 0.103 ± 0.013; CentMed: 0.228 ± 0.037, 0.012 ± 0.001, 0.216 ± 0.037, 0.093 ± 0.013; EastMed: 0.205 ± 0.031, 0.011 ± 0.001, 0.195 ± 0.031, 0.080 ± 0.012; BalBSea: 0.234 ± 0.056, 0.010 ± 0.001, 0.224 ± 0.055, 0.042 ± 0.007, respectively).
  • WestMed. The expression TAOD550 = SAOD550 + AAOD550 could alternatively be written as TAOD550 = (DAOD550 + OSAOD550) + AAOD550 (see Equations (8)–(10)). In the event of a Sahara dust outbreak, the scattering particles could include dust particles as well as other particles (such as sea-salt, sulphates, and nitrates, denoted as OSAOD550 in the equation). By replacing the average values of TAOD550 = 0.203, AAOD550 = 0.012, SAOD550 = 0.191, and DDAOD550 = 0.103 in the last expression, it is easily found that the average OSAOD550 is 0.088, which is equal to ≈7·AAOD550. In the same way, DAOD550 ≈ 9·AAOD550. Therefore, TAOD550 = (9·AAOD550 + 7·AAOD550) + AAOD550 = 17·AAOD550. This outcome indicates that the total solar radiation attenuation in the WestMed subregion is 17 times more than the attenuation resulting from aerosol absorption alone.
  • CentMed. Following the same procedure, it is estimated that TAOD550 = 19·AAOD550. This implies that the total solar radiation is attenuated in the CentMed subregion 19 times by an atmosphere containing absorbing aerosols only.
  • EastMed. Here, TAOD550 = 19·AAOD550, as in CentMed. This result shows the resemblance of the aerosol profiles in these two Mediterranean subregions (cf. Figure 10b,c).
  • BalBSea. Now, TAOD550 = 23·AAOD550 having a higher coefficient than the one in the other three subregions (see the aerosol profiles in Figure 10d).
  • Kambezidis [31] found the AAOD550 coefficient to be 18 for the whole Mediterranean area in the same period, a value rather close to the average of the four coefficients in the present study (19.5).
  • The immediate conclusion from the above results is that the contribution of absorbing aerosols to the total solar radiation attenuation becomes less and less important from WestMed to BalBSea as its role is progressively taken up by the scattering effect. This is also confirmed by the progressively increasing average SSA values over the four subregions: 0.939 (WestMed), 0.947 (CentMed), 0.948 (EastMed), and 0.958 (BalBSea).
Figure 10 shows that the annual values in the TAOD550 and AAOD550 time series occur within the neutral ONI zone most of the time. A remarkable observation, though, is that the DDAOD550 and SAOD550 time series are always occurring during cold La Niňa events.
EastMed’s average TAOD550 value throughout the time period under study in this work is 0.206, which is quite consistent with the East Mediterranean’s 0.207 value discovered by Ozdemir et al. [85] between 1999 and 2018. The average TAOD550 value for the Mediterranean region from 2005 to 2013 was found to be 0.150, according to Chiapello et al. [86]. This value is much lower than the average value found in the current investigation for the same experimental period (0.219). According to a research by Kaskaoutis et al. [87] on the aerosol loading over the Athens region in Greece, the city’s pollution influence on solar radiation resulted in a high yearly average of TAOD550 = 0.35 between 2000 and 2005. Mallet et al. [88] found TAOD440 values over the Mediterranean region in the period 1996–2012 in the range 0.15–0.32, these values include those of the present work (0.204–0.236).
Ångström’s exponent, AE470–870, is a number that indicates aerosol size. A plot of AE470–870 versus ONI similar to that of TAOD550 in Figure 8 is shown in Figure 11a. For the four Mediterranean subregions, the distribution of the data pairs (AE470–870, ONI) is shown here. Every year, it is noted that coarse-mode particles are present throughout the Mediterranean, irrespective of the ENSO phase. Only two years, 1983 and 1992, are associated with fine-mode aerosols, and these are related to the volcanic eruptions of Mt. El Chichón and Mt. Pinatubo (Mt. El Chichón, Mexico, 28 March–4 April 1982; Mt. Pinatubo, Philippines, 12–15 June 1991). Mt. El Chichón injected 7 × 106 tons of SO2 and 20 × 106 tons of sulphuric acid aerosols into the atmosphere within a month after the eruption, causing a surface cooling in the northern hemisphere of ≈−0.2 °C [89,90]. The eruption occurred just as the 1982/1983 El Niňo event was in its starting phase; because of this, some scientists have supposed that the El Chichón eruption triggered a positive El Niňo phase [91]. Nevertheless, climate models and detailed studies of past volcano eruptions and El Niňo events have not shown a connecting line between the two events; therefore, the timing must be just a coincidence [91]. On the other hand, the Pinatubo eruption released a huge amount (17 × 106 tons) of SO2 into the stratosphere causing a global temperature decrease of ≈−0.5 °C and a reduction in solar light by 10% globally over 1992–1993 [92].
Figure 11b shows the temporal evolution of AE470–870 over the 45-year period, with peaks in 1983 and 1992 corresponding to the Mt. El Chichón and Mt. Pinatubo eruptions, respectively. Higher AE470–870 values are observed over the BalBSea subregion than the other three areas. The dashed lines are linear regression fits to the AE470–870 data series with the following expressions: AE470–870 = −8.1 × 10−3·t + 17.072, R2 = 0.331 (WestMed); AE470–870 = −7.4 × 10−3·t + 15.723, R2 = 0.336 (CentMed); AE470–870 = −8.2 × 10−3·t + 17.434, R2 = 0.366 (EastMed); and AE470–870 = −7.91 × 10−3·t + 17.002, R2 = 0.481 (BalBSea). All regression equations are significant at the 99.9% CI. In all four subregions the trends are negative with negligible slopes, as shown by the very low coefficients in the regression lines.
A last investigation concerns the types of aerosols present in the atmosphere of the four Mediterranean subregions in the period of the study. To categorise aerosols into groups, several researchers have developed varying ranges for the TAOD550 and AE470–870 values (e.g., [93,94]). The current investigation adopts the aerosol-type discrimination criteria from [94]. These are as follows: (i) clean-maritime aerosols (CMA) for TAOD550 < 0.15 and AE470–870 < 1.30; (ii) urban/industrial/biomass-burning aerosols (UIA) for TAOD550 > 0.20 and AE470–870 > 1.00; (iii) coarse-mode, including desert-dust aerosols (DDA) for TAOD550 > 0.25 and AE470–870 < 0.70; and (iv) mixed-type aerosols (MTA) elsewhere. The different colouring of the data points in Figure 12 indicates the various types of aerosols present over the Mediterranean subregions on a monthly basis between 1980 and 2024. Most of them fall into the MTA and UIA categories. CMA is due to the large mass of water of the Mediterranean Sea, whereas the UIA are related to air pollution over the large urban agglomerations along the Mediterranean Sea coast, industrial activities in the surrounding countries, and occasional summertime wildfires. Although Figure 10d shows low annual DDAOD550 values, it is interesting to note that there are no desert-dust aerosols over the BalBSea subregion (see Figure 12d). This discrepancy may be due to the categorisation criteria’s inability to detect low concentrations of desert-dust particles. For the purpose of comparison, the annual scales on both axes of the plots in Figure 12 have been purposefully left in place. However, Figure 12’s scatter plots demonstrate that MTA-type aerosols contribute significantly to the overall aerosol loading across the four Mediterranean subregions; in fact, the percentages of contributions from CMA, UIA, DDA, and MTA to the overall aerosol loading are 30.4%, 8.1%, 11.1%, and 50.4% (WestMed); 16.5%, 22.2%, 6.5%, and 54.8% (CentMed); 18.0%, 16.9%, 7.8%, and 57.4% (EastMed); and 12.4%, 52.2%, 0.00%, and 35.4% (BalBSea), respectively. The low percentages of desert-dust aerosols, in particular, are caused by the DDA months being solely spring (April, May, and occasionally March or June, not stated specifically here). Similar conclusions were reached by Kambezidis [31] for the entire Mediterranean region, Kaskaoutis et al. [42] for Solar Village in Saudi Arabia, and Kalapureddy et al. [94] for the Arabian Sea. This finding implies that the aerosols in question over the larger Mediterranean region are related to suspended dust aerosols from the Sahara. The prevalence of UIA over BalBSea (52.2%) in contrast to the other three regions (8.1%, 22.2%, and 16.9% for WestMed, CentMed, and EastMed, respectively) is another noteworthy finding. The reason for this is that while some Balkan counties (i.e., Bulgaria, Croatia, Romania, and Slovenia) have recently joined the European Union (EU), their adherence to relevant EU air pollution standards has not yet produced notable outcomes; on the other hand, other Balkan nations remain outside the EU. These facts may be ascribed to more intense air pollution and activities based on petrol and coal [95].
The DARF and AE470–870 time series’ annual average values (represented by red curves) are displayed in Figure 13 for the whole Mediterranean region (allMed). Interestingly, the CentMed curves resemble the allMed curves more than the WestMed, EastMed, and BalBSea curves. The amount of the aerosols and the warming/cooling effect over CentMed may, therefore, play a major role and influence the adjacent WestMed and EastMed regions (see AE470–870 in Figure 13a), as well as every other subregion (see DARF in Figure 13b).
The Pearson’s correlation coefficient, r, for the monthly mean values of DARF and AE470–870 is displayed in Table A2 (see Appendix A). As exhibited in Figure 13, the highest r values between DARF (CentMed) and DARF (allMed) and between AE470–870 (CentMed) and AE470–870 (allMed) are verified. In the Mediterranean region, it is especially fascinating to note how the r values of both parameters progressively increase as one moves from a left to a right domain. Regarding AE470–870, for instance, the correlation coefficients between WestMed and CentMed are 0.777, and between CentMed and EastMed 0.935. Regarding DARF, the correlation between WestMed and EastMed is 0.797, while the correlation between WestMed and BalBSea is 0.913.

3.1.6. General Conclusions from Section 3.1.1 to Section 3.1.5

  • Volcanic eruptions, depending on the amount of debris outflown in the atmosphere, can mask and modulate the optical properties of the natural and anthropogenic aerosols. In the present study, volcanoes from the (continuously updating) catalogue of the Global Volcanism Programme, Smithsonian Institute (https://volcano.si.edu), were selected in the period 1980–2024. They are characterised by VEIs in the range 4–6, except for Etna and Stromboli volcanoes in CentMed, that were “allowed” to have VEIs of 2 and 3.
  • The ΔTAOD550 values were found to be almost negative (i.e., TAOD550 < TAOD’550) in the inner part of the Mediterranean region (CentMed and EastMed), while ΔTAOD550 varied between negative and positive values over WestMed and BalBSea in the whole period of the study. This indicates the smaller effect of clouds on the CentMed and EastMed areas than the other two subregions.
  • In an old study, Dutton et al. [96] examined TAOD500 at the Mauna Loa Observatory (MLO, 19.54° N) and found that the maximum values of the parameter were 0.28 and 0.23 from the El Chichón (1982) and Pinatubo (1991) eruptions, respectively. These values are comparable with the peaked values in Figure 2a–d of the present study. Indeed, the maximum annual TAOD550 values for the El Chichón and Pinatubo eruptions are 0.289 and 0.319 (WestMed), 0.319 and 0.335 (CentMed), 0.274 and 0.303 (EastMed), and 0.354 and 0.354 (BalBSea), respectively. These higher maximum values are due to the fact that the MLO is located on the remote island of Hawaii, thus recording the background aerosol loading in the atmosphere. Similar observations to those at the MLO were made at some Antarctic stations by Herber et al. [97]. Sato et al. [80], in a study to investigate the stratospheric aerosol loading from volcanic eruptions in the period 1850–1990, found that El Chichón mostly affected regions in the northern hemisphere at latitudes 0° N–30° N just after its eruption in March/April 1982, and regions at latitudes 60° N–90° N at the end of the year; one can remark that these low-latitude areas in the northern hemisphere include most of the broader region of the Mediterranean in this study, which were affected by those eruptions (see Figure 3, Figure 4, Figure 11b and Figure 13a). A similar study by Long and Stowe [98] about the Pinatubo eruption found that the stratospheric aerosol loading was particularly high over areas with latitudes 20° S–20° N just after its eruption in June 1991. This confirms the lower TAOD550 values recorded at MLO due to the Pinatubo eruption in comparison to those over the four Mediterranean subregions (see Figure 3, Figure 4, Figure 11b and Figure 13a). A recent work by Toohey and Sigl [99] reconstructed StAODs in the period from 500 BC to 1900 AD; they found peak StAOD550 values ranging between 0.2 and 0.4, quite comparable to the two peaks shown in Figure 2 in the four Mediterranean subregions due to the El Chichón and Pinatubo eruptions.
  • For the first time, it was demonstrated that the TAOD550,WVE time series, i.e., atmospheric aerosols devoid of volcanic materials but including anthropocentric-origin ones, show rather constant levels (no significant trend) in any of the four Mediterranean subregions. This outcome indicates that a volcanic-debris-free atmosphere may have a lesser impact on the climate change phenomenon. Such a simulation has never been attempted to the knowledge of the author, and it would be interesting to examine the characteristics and evolution of a volcanic-debris-free climate in the past and the future.
  • The desert-dust and black-carbon aerosols themselves account for half of the aerosol loading in the broader area of the Mediterranean region, especially over the Mediterranean Sea and the surrounding countries; an exception is the Balkans and Black Sea areas where the contribution from desert dust has a lesser or negligible effect.
  • The free-of-volcanic-debris aerosols show lower (almost half) AOD values than those with volcanic debris over the Mediterranean (cf. TAOD550 and TAOD550,WVE in Figure 2 and Figure 4).
  • The global volcanism seems to have had a greater influence over the Balkans and Black Sea (see Figure 4), and the least aerosol loading without volcanoes (see Figure 5) in the period 1980–2024.
  • The above differentiation among the four Mediterranean subregions may have a major contribution to the patterns of TAOD550 from the ENSO and NAO phases (see Figure 6a–d), especially over the BalBSea subregion (see Figure 6d).
  • The majority of the annual TAOD550 values are favoured by negative and neutral ENSO conditions (see Figure 8).
  • The linear regression fits to TAOD550 as a function of time (see Figure 3) give the following trends: −0.017/decade (WestMed); −0.024/decade (CentMed); −0.013/decade (EastMed); and −0.037/decade (BalBSea). All trends are negative with a greater slope over BalBSea and a smaller slope over EastMed. Kambezidis [31] found a trend over the entire Mediterranean in the period 1980–2022 equal to −0.018/decade, close to the average of the above four TAOD550 slopes (−0.023/decade). Chiapello et al. [86] found a TAOD550 trend of −0.06/decade over the Western Mediterranean in the period 2005–2013, which is almost three times higher than the corresponding average value in the present work. Shaheen et al. [60] found a TAOD550 trend over EastMed of ≈−0.002/decade from MERRA-2 in the period 1980–2018, ten times lower than the average value in the present work. These varying slopes among different studies are due to the different periods examined, the source of the data used, and the areas covered; they, therefore, reflect the need to standardise the whole procedure for a more comparable way among various works.
  • The linear regression fits to TAOD550,WVE as a function of time (see Figure 5) derived the following trends: +0.003/decade (WestMed); +0.003/decade (CentMed); +0.006/decade (EastMed); and +0.002/decade (BalBSea). These trends are all positive, greater in EastMed, smaller over BalBSea, equal in the other two areas, and an order of magnitude lower than those for TAOD550.
  • The above result shows that the atmospheric aerosol loading over the entire Mediterranean in the period 1980–2024 has a negligible trend or is rather constant in the absence of global volcanic eruptions.
  • There is an almost 50%/50% balance between the desert-dust/black-carbon aerosols and the stratospheric/sea-salt/cloud effect on the total aerosol loading over the WestMed, CentMed, and EastMed areas (see Figure 9). The BalBSea subregion is affected more by black-carbon aerosols than volcanic eruptions, probably because of its remote location.
  • To include the cloud effect as a component of SAOD550, the COD values can be divided by 1000 in order to derive comparable values with SAOD550 (see Equation (9)).
  • TAOD550 was found as a function of AAOD550 over the four areas: TAOD550 ≈ 17·AAOD550 (WestMed), TAOD550 ≈ 19·AAOD550 (CentMed, EastMed), and TAOD550 ≈ 23·AAOD550 (BalBSea). The higher weighting factor for BalBSea (i.e., 23) refers to the higher contribution of absorbing aerosols to the total aerosol loading over the region in comparison to that over the other three areas. This was justified in Figure 10 and Figure 12.
  • Regardless of the ENSO phase, all AE470–870 values indicate coarse-mode particles across the Mediterranean on a yearly basis. Fine-mode aerosols are linked to only two years: 1983 and 1992, related to the volcanic eruptions of Mt. El Chichón and Mt. Pinatubo, respectively.
  • Most of the aerosol types fall into the MTA and UIA categories over the entire Mediterranean region. There is an absence of or at least a contribution to desert-dust aerosols over BalBSea.

3.2. Part B—Statistical Approach

All statistical tests were implemented by the freeware JASP 0.19.3 statistical software (see https://jasp-stat.org).

3.2.1. Nexus Between the Aerosol Parameters and SSN/NAOI/ONI

This section investigates the nexus among all the parameters considered in the present study with the SSN, NAO, and ENSO phenomena. This is performed by calculating Pearson’s correlation coefficient, r, and the coefficient of determination, R2, to the monthly averaged values of the parameters. The statistical results are shown in Table A3 (see Appendix A) for the four Mediterranean subregions. It is interesting to note that Kambezidis [31] found significant nexus in the AE470–870/NAOI and DARF/ONI time series pairs for the broader Mediterranean area in almost the same experimental period with that of the present work. From Table A3, the following outcomes can be drawn: (i) there are significant relations at 95%, or 99.9% CI (indicated by *, ***, respectively); (ii) high correlations with ONI exist for AAOD550, DDAOD550, StAOD550, AE470–870, and DARF over various parts of the Mediterranean; (iii) high correlations with NAOI occur for StAOD550, AE470–870, SSA, DARF, and COD over certain areas of the Mediterranean. In summary, there is a significant effect of (i) the ENSO phenomenon on some aerosol types (absorbing, desert-dust, and stratospheric aerosols), and some atmospheric parameters (AE and DARF), and (ii) the NAO phenomenon on stratospheric aerosols and some atmospheric entities (AE, DARF, SSA, and COD), mainly over WestMed and CentMed. Table A8 (see Appendix A) gives a summary of the significant correlations of the parameters with SSN, NAOI, and ONI for |r| > 0.2.
Linear regression fits to TAOD550 over the four Mediterranean subregions in the period of the study as a function of ONI (not imprinted in Figure 8), show increasing trends of TAOD550 with the following expressions: 4.8 × 10−3·ONI + 0.203, R2 = 0.001 (WestMed), 8.0 × 10−3·ONI + 0.228, R2 = 0.022 (CentMed), 8.0 × 10−3·ONI + 0.205, R2 = 0.032 (EastMed), and 13.2 × 10−3·ONI + 0.233, R2 = 0.027, (BalBSea), all insignificant at the 95% CI. Nevertheless, these trends are considered quite low or negligible; they are therefore compatible with the small positive r values for TAOD550/ONI and SAOD550/ONI, and small negative r values for DDAOD550/ONI and AAOD550/ONI (see Table A3). In conjunction with this, Urdiales-Flores et al. [28] failed to associate a declining trend of TAOD550 over the Mediterranean region with a significant effect of ENSO.
A more careful observation of the statistical results in Table A3 (see Appendix A) shows that all three natural phenomena (SSN, NAOI, and ONI) are linked with AE470–870, SAOD550, SSA, and StAOD550 at a CI of 95% to 99.9% in all four areas or parts of the Mediterranean region, no matter of whether the correlation coefficient is positive or negative. In other words, there is a combined effect of the three mechanisms on the aerosol-particle size, their scattering capability, their scattering albedo, and the stratospheric aerosol concentration. Actually, the latter three parameters are related directly or indirectly to AE470–870, because the scattering or absorption effects of aerosols on solar radiation depend on their size.
The relationship between the various aerosol properties could also be shown in Table A3. However, in order to save space and because the primary objective of the current study is to identify any (in)direct association between the ENSO phenomenon and the aerosol parameters, this was left out. It was preferred that a description of some statistical findings for the aerosol parameters be provided in the sections that follow.

3.2.2. Nexus Between the Aerosol Parameters and the Four Mediterranean Areas

This section investigates the nexus of TAOD550 and AE470–870 with the four selected areas that make up the broader Mediterranean region. The deeper meaning of this investigation is to examine statistically the possible dominance of the various aerosol types over the four subregions, or, in other words, to support the hypothesis that each subregion may exhibit distinct aerosol regimes; this was performed in Section 3.1.5 in a graphical way (see Figure 12). To implement this, the statistical process of Cluster Analysis (CA) was invoked; the analysis was performed on the monthly values of the above aerosol parameters over the four areas in the whole period of the study. The hierarchical analysis gave the following silhouette scores (SS) for four clusters: Cluster 1 SS = 0.194, Cluster 2 SS = 0.391, Cluster 3 SS = 0.352, and Cluster 4 SS = 0.562. It should be noted that SS is delimited in the range [−1, +1]; when SS → +1 means perfectly clustered, 0.5 ≤ SS ≤ 0.7 means reasonably well-clustered, 0.3 ≤ SS < 0.5 implies moderate structure, and SS < 0.3 refers to weak or overlapping clusters. Therefore, the adoption of SS = 0.562 represents well-separated and cohesive clusters. The next step was to associate clusters with subregions. For this reason, a Decision-Tree Classification (DTC) was applied between the subregions and TAOD550 and AE470–870. The results in Table 3 give the frequency (in %) of the predicted (modelled) TAOD550 and AE470–870. Their observed diagonal values sum up to 49%, which means that the model correctly predicts the subregion about half the time, which is modest but meaningful given the complexity of aerosol dynamics.
The interpretation of the results in Table 3 are as follows.
  • BalBSea. Correctly classified: 21%; often misclassified as CentMed (6%) and EastMed (7%); suggests overlap in aerosol signatures with neighbouring subregions.
  • CentMed. Correctly classified: 7%; frequently confused with EastMed (1%) and WestMed (4%); indicates transitional aerosol behaviour.
  • EastMed. Correctly classified: 9%; misclassified as BalBSea (1%) and WestMed (3%); suggests shared fine-particle regimes.
  • WestMed. Highest correct rate: 12%; still confused with EastMed (6%) and CentMed (6%); likely reflects seasonal dust variability.
The scientific takeaways of the above results are that (i) TAOD550 and AE470–870 alone offer partial discrimination between subregions, and (ii) an overlap in regimes (e.g., dust vs. air pollution) reduces classification clarity. By considering the average TAOD550 and AE470–870 values over the study period, there is a connection between the four clusters and the subregions shown in Table 4.
The inter-relation of the aerosol types with the four clusters and the subregions can justify the geographical determination of the four Mediterranean areas given in Section 2 (see Table 1 and Figure 1).

3.2.3. Descriptive Statistics

Table 5 gives the averages, standard deviations, medians, and trend values of the parameters considered in this study. The trend, α (in units/decade), was estimated from the linear regression fit to the data points of the parameter q in the form q (t) = a·t + b; q = any parameter in the headings of Table 5 (see columns 3–15); a, b = regression constants; t = any year within the period 1980–2024; α = 10·[q (2024) − q (1980)]/45 (in units/decade).
The statistical results in the above Table can give some valuable information.
  • The magnitudes of the μ, σ, and m values for the parameters of TAOD550, AAOD550, SAOD550, SSAOD550, BCAOD550, and SSA are similar throughout the four Mediterranean subregions. This suggests that the subregions have comparable aerosol loading patterns. The significance of large water bodies (Mediterranean Sea, Black Sea) must be acknowledged here, particularly for the SSAOD550 and BCAOD550 parameters, as these aquatic surfaces contribute to TAOD550 with relatively high sea-salt aerosol loads. On the other hand, numerous large cities, industrial areas, and wildfires around the Mediterranean contribute to TAOD550 with black-carbon aerosols, particularly over BalBSea (see µ in Table 5).
  • In contrast, the DDAOD550, StAOD550, and AE470–870 parameters exhibit a declining, rising, and increasing trend, respectively, from WestMed to BalBSea. This suggests that the overall particle sizes of the aerosols of volcanic debris and desert dust vary across the four Mediterranean subregions. While stratospheric aerosols appear to finally influence the Balkan peninsula and Black Sea area more (see Figure 4 and µ in Table 5), desert-dust outbreaks from the Sahara Desert dominate the West Mediterranean [100]. According to AE470–870, fine-mode aerosols gradually take over from WestMed to BalBSea (see Figure 10a and µ’s in Table 5).
  • As DARF decreases from WestMed to BalBSea, a comparatively greater warming effect was observed over the West Mediterranean than over the Balkans and Black Sea regions (see µ’s in Table 5). In contrast, COD is marginally rising in each of the subregions listed, which is consistent with [31], who noted it for the Mediterranean region as a whole. However, between 1979 and 2012, Kambezidis et al. [101] discovered rising trends in COD for mid- and high- level clouds and falling trends for low-level clouds. A modest increasing trend in COD was observed from WestMed to BalBSea (see COD μ’s in Table 5), taking into account all clouds in the current study. This discovery is consistent with the increasing COD in [101] and could be explained by the fact that MERRA-2 reanalysis data “see better” upper clouds.
  • Table 5’s average yearly COD values are in line with those from other studies. For instance, in the Atacama Desert in Northern Chile, Luccini et al. [102] discovered COD levels of about 15 (at Arica) and about 11 (near Poconchile). It is important to keep in mind that the COD values indicate how frequently cloudiness occurs in a given area; they are lower in deserts and greater in the Mediterranean regions that include vegetation and water surfaces.
  • For the majority of parameters (except from DDAOD550 and COD in all four areas and for AAOD550 and DARF over BalBSea), linear regression fits to the data points of the parameters as a function of time, t (years in the period 1980–2024), demonstrated significance at the 99.9%, or 95% CI. One can, therefore, forecast the parameter level shortly after 2024 with a reasonable degree of accuracy thanks to the high CIs (the accuracy assumes no additional high volcanic activity at the global level). For instance, according to the Giovanni platform (accessed on 27 January 2025), the μTAOD550 values for January 2025 for the four locations are 0.087, 0.118, 0.105, and 0.098, respectively. The January 2025 values, also calculated using the linear regression formulas (not displayed here for spacing reasons), were found to be 0.070, 0.100, 0.115, and 0.098, respectively, quite close to the observed (Giovanni) ones.
  • Except for the DARF and COD parameters, which displayed positive trends (see α’s in Table 5), all of the aerosol and atmospheric metrics have trends that are marginally negative or even close to zero.
  • The four Mediterranean subregions’ average yearly AE470–870 values for the study period were determined. For the entire Mediterranean, Kambezidis [31] obtained an average AE470–870 value of 0.939, which is very similar to the average of the four subregions (0.989). However, Ozdemir et al. [85] observed higher AE470–870 values for the East Mediterranean between 1999 and 2018, in the range 1.15–1.66. From 2000 to 2015, Sharafa et al. [100] discovered that the average annual AE470–870 values from AERONET stations in three sub-Saharan African nations ranged from 0.555 to 1.722. AE440–870 fluctuated between 0.92 and 1.57 over the Mediterranean between 1996 and 2012, according to Mallet et al. [88]. These values are in line with the 0.874–1.217 values obtained in the current investigation.

3.2.4. Deseasonalisation; Stationarity; Cross-Correlation Analysis; Durbin–Watson Test; Granger-Causality Test

The statistics given in Table A2 and Table A3 provide information of a rather strong (i.e., CI = 99.9%) relationship between SSN/NAOI/ONI and some of the parameters examined. Nevertheless, these outcomes do not show if any of the NAOI, ONI, or SSN time series causes (influences) the existence of any of the aerosol parameters. In other words, one would like to explore the possibility that NAOI, ONI, or SSN can affect (cause) any of the aerosol parameters, at least those which were found to have a higher r and R2 at the 99.9% CI level in Table A3. To implement this statistical procedure, the monthly mean values of all the parameters in Table 5 (those in the headings of columns 3–15) together with those of NAOI, ONI, and SSN were considered. This was performed because the annual values used in the preceding sections can mask any “fine-resolution” connection among them. To resolve this issue, all the monthly time series were deseasonalised using the seasonal decomposition of time series (STL) method in order to remove cyclical (seasonal) effects on the time series of the data [103,104]. It should be reminded here that all the parameters in Table 5 and SSN/NAOI/ONI do have a fixed (for SSN and aerosols) or a variable (for NAOI and ONI) cycle. By performing the STL process on all-time series of the variables, new (deseasonalised) time series were derived and used in the analysis of the current section (see Table A4 and Table A5, Appendix A).
Finding a cause-and-effect link between SSN/NAOI/ONI and the aerosol/atmospheric parameters is the ultimate goal of this study. This can be shown by applying the Granger-causality (GC) test to two time series, i.e., between any of the aerosol parameters as the effect (the dependent variable) and SSN/NAOI/ONI as the cause (the independent variable). The following requirements for the time series in question must apply to the GC test.
  • Stationarity. It is necessary for the time series to be stationary, meaning that its statistical characteristics, such as μ and σ, must remain constant over time. Numerous statistical tests can be used to analyse stationarity. The present study decided to use the Kwiatkowski–Phillips–Schmidt–Shin (KPSS) test [105]. The Ho hypothesis makes the assumption that the time series are stationary during the test. The time series are said to be stationary if the corresponding p-values are less than or equal to α (0.05 for 95% CI or 0.01 for 99% CI), which means that Ho applies. The stationarity results for all variables examined for the four Mediterranean subregions between 1980 and 2024 are displayed in Table A4 (see Appendix A). Before moving on to the GC application, the deaseasonalised time series in Table A5 that were determined to be non-stationary received a first-order difference statistical modification to become stationary (see Table A6, Appendix A).
  • Linearity. There are several statistical measures available to assess whether the independent and dependent time series are linear. The rather low R2 values indicate a weak linearity between the paired parameters, although this does not exclude the use of the GC statistic. In this case, the R2 values were adopted in the regression analysis between SSN/NAOI/ONI and an aerosol/atmospheric variable given in Table A3.
  • Lag length. In the GC analysis, it is essential to select a suitable lag length for the dependent aerosol/atmospheric variables with respect to the independent variables of SSN/NAOI/ONI time series. Six (months) was determined to be the ideal lag period for the independent variables under investigation. This was accomplished by computing the Bayesian Information Criterion (BIC) and the Akaike Information Criterion (AIC); it was discovered that both BIC and AIC values became stabilised following lag 6. The (ideal) lag 6 in Table A4 further supports this.
  • No omitted variables. The results may be skewed if important (independent) variables are left out of the analysis; therefore, all pertinent variables should be included. For this reason, all (independent) SSN/NAOI/ONI time series were used in this study with a lag of six months.
  • No multi-collinearity. The independent variables should not be perfectly multi-collinear; high correlation coefficients (around +1 or −1) suggest that multi-collinearity may exist. However, the low (near-to-zero) r values among the SSN/NAOI/ONI time series in Table A3 (Appendix A) indicate that there is no such collinearity for them.
  • No auto-correlation of the error terms. Any dependent or independent variable was subjected to a regression analysis against time; the auto-correlation plot of the residuals from each fit revealed no discernible trend. Consequently, no auto-correlation was discovered. By using the Durbin–Watson (DW) test on the regression residuals, the same conclusion was reached: all DW values fell between 1 and 2, suggesting that there was little to no auto-correlation in the residuals. It should be noted that DW values range from 0 to 4, and an auto-correlation is minimal or non-existent when the DW value falls between 1.5 and 2.5.
The above criteria for the implementation of the GC test were, therefore, met for the time series of the examined variables. The detailed statistical results of the Granger-causality test are shown in Table A7 (see Appendix A). Table 6 shows a summary of the cause–effect of the SSN/NAOI/ONI variables on the aerosol/atmospheric time series. The Table consists of “Yes”/”No” for a cause–effect/no cause–effect of the SSN/NAOI/ONI time series on the examined aerosol/atmospheric parameter. If a lagged component of any of the SSN/NAOI/ONI time series is present in the regression expression (see Table A7), a “Yes” is marked in Table 6; otherwise, a “No”.
The following conclusions can be drawn from the above GC statistical analysis.
  • Acceptable regression results in the form q = f(SSN,NAOI,ONI) occur for TAOD550, AAOD550 (all domains except ONI in BalBSea), SAOD550 (all domains), DDAOD550, (WestMed, CentMed), StAOD550 (WestMed, BalBSea), SSAOD550 (CentMed, EastMed except NAOI), SSA (BalBSea), DARF (all domains except ONI in BalBSea), netSWRBOA,CS,A, netLWRBOA,CS,A (all domains except ONI in WestMed), and COD (all domains).
  • Solar activity (through SSN) does not show a cause–effect on (i) SSAOD550, BCAOD550, AE470–870, and SSA over WestMed; (ii) StAOD550, BCAOD550, AE470–870, and SSA over CentMed; (iii) DDAOD550, StAOD550, BCAOD550, AE470–870, and SSA over EastMed; and (iv) DDAOD550, SSAOD550, BCAOD550, AE470–870, and COD over BalBSea.
  • From the above statement, it is obvious that solar activity does not possess a statistically significant cause–effect on BCAOD550 and AE470–870 over the entire Mediterranean region.
  • The North Atlantic Oscillation (through NAOI) shows a cause–effect on almost the same parameters and Mediterranean areas with those of SSN.
  • The El Niño–Southern Oscillation (through ONI) influences all aerosol parameters mentioned above.
  • From Table A7, the SSN/NAOI/ONI time series in the regression analyses include a prevailing retarded six-month influence on the Mediterranean atmosphere. This is almost ½ of the 11-year solar activity cycle and exactly half a full year. In the latter case, one may consider a warm and a cold season of six months duration each; perhaps this occurs because nature favours these two main seasons.
  • In the analyses for the cause–effect of the SSN/NAOI/ONI time series on the aerosol/atmospheric parameters in Table A7, only the expressions with overall α ≥ 0.4 (a 60% CI) are shown. Nevertheless, the regression expressions with α < 0.4 still contain certain lagged values of the independent variables.
  • The last observation comes in agreement with the initial outcome of a rather synchronous fluctuation in DARF and SSN (see periods 1988–1992 and 1998–2004 in Figure 9 of [31]), and of DARF and ONI (see marked rectangles in Figure A3 of [31]). The characterisation “synchronous” was given by Kambezidis [31], who used annual mean values to show an almost synchronous temporal evolution of the DARF/SSN and DARF/ONI time series. To demonstrate this synchronous fluctuation in ONI and DARF over the four Mediterranean subregions in the case of progressively decreasing positive ENSO phases or turning from positive to negative events, Figure 14 was prepared. The ellipses show the above-mentioned events; the red arrows inside the ellipses indicate the effect of a decreasing ENSO phase on DARF, while DARF follows a similar decreasing tendency with that for ENSO with a retardation of about six months, according to the Granger-causality effect shown in this section. It is, therefore, clearly seen that there exists a signalling effect of the ENSO phase on the DARF levels over all four Mediterranean areas.

3.2.5. Normal Distribution of the Examined Variables

The distribution patterns of the parameters under investigation are the focus of this section. P-P (probability–probability) plots were created as a result. With the same μ and σ as the variable’s time series, a P-P plot displays the probability of a variable against its theoretical (normal) one. A normal probability distribution’s fit to the observed data can be evaluated using a P-P plot. To put it another way, a P-P plot indicates which ranges of the parameter values follow a normal distribution curve; data points that fall away from the x = y line indicate that the observed data deviates from the normal distribution, while data points that fall along this line indicate that all or some of the observed data matches the theoretical distribution. In a P-P plot, the x-axis represents the cumulative probabilities of the reference (normal) distribution, and the y-axis the cumulative distribution of the actual data.
In the present study, P-P plots for all parameters in the four Mediterranean subregions were derived using the monthly mean values of their time series in the period 1980–2024. For space saving, selected P-P plots are only shown here. Figure 15 presents the P-P plots for the three physical phenomena (i.e., SSN, NAOI, and ONI) as well as for AAOD550 (BalBSea), SAOD550 (CentMed), and SSA (WestMed).
The P-P patterns of AAOD550 (BalBSea), SAOD550 (CentMed), and SSA (WestMed) appear to be significantly influenced by those of SSN, NAOI, and ONI, respectively. The SSA, SAOD550, and AAOD550 graphs are examples of the three patterns discovered among all analysed parameters in this work and over the four Mediterranean subregions. It is astounding how similar the PDFs are for the cause (impact) of SSN, NAOI, and ONI on SSA, SAOD550, or AAOD550, respectively. As a consequence, this result has the value of a rule and provides more evidence of how the three physical phenomena affect the atmospheric aerosols in the larger Mediterranean region. This rule is consistent with the study’s findings that the patterns of atmospheric aerosols over the Mediterranean region show a signal (effect), no matter how strong or small of any one of the three natural phenomena or a combination of them.

4. Conclusions and Discussion

The current study was an extension of a former study by the same author [31] that examined the impact of three physical phenomena on atmospheric processes over the Mediterranean: solar activity, the North Atlantic Oscillation, and the El Niňo–Southern Oscillation. The current study divided the Mediterranean region into four smaller areas (West Mediterranean, 6° W–10° E, 30° N–45° N; Central Mediterranean 10° E–20° E, 30° N–45° N; East Mediterranean, 20° E–35° E, 30° N–41° N; Balkans and Black Sea, 20° E–35° E, 41° N–45° N), whereas the earlier work considered the entire region into account. The current study expanded the data to 2024 (45 years), whereas the previous study used data from 1980 to 2022 (43 years). These distinctions were thought to be crucial for a more thorough investigation of the effects of the three physical events on atmospheric aerosols (smaller areas and longer time period).
The present work was articulated in a different way; it was divided into two parts: a physical approach and a statistical approach. In the first part, the effects of SSN, NAOI, and ONI were examined on the physical and optical properties of aerosols, solar radiation, and clouds. The second part was devoted to analysing the above effects in a statistical manner. Both approaches were used to give a more in-depth answer on the significant effects of the physical phenomena on the aerosol processes over the four Mediterranean areas. The main conclusions from the study are deployed below.
  • Part A. Physical approach
Volcanic eruptions have been shown to have the ability to alter and/or conceal the optical characteristics of both man-made and natural aerosols. Only worldwide volcanic eruptions with VEIs between 4 and 6 were taken into consideration in the analysis; however, the Etna and Stromboli volcanoes, which are situated in the CentMed subregion and have VEIs of 2 and 3, were also taken into consideration.
Compared to EastMed and BalBSea, a more significant warming effect (higher positive DARF values) was observed over WestMed and CentMed.
Over the wider Mediterranean region, an aerosol scattering profile with SSA values higher than 0.90 was generally observed. This implies that the primary cause of solar radiation attenuation is the scattering aerosol particles.
For the first time, a modelling approach to the estimation of TAOD550 was made as TAOD’550 = DDAOD550 + StAOD550 + SSAOD550 + CODsca (the scattering atmospheric components) + BCAOD550 + CODabs (the absorbing atmospheric components), where CODsca = 0.85·(COD/1000) and CODabs = 0.15·(COD/1000). The difference between TAOD550 and TAOD’550 annual values varied in the extreme range ±6% in all four Mediterranean areas.
Maximum TAOD500 values at the Mauna Loa Observatory, Hawaii, were found to be 0.28 and 0.23 from the El Chichón (1982) and Pinatubo (1991) eruptions, respectively. These values are comparable to the maximum values found in the present study, i.e., 0.289 and 0.319 (WestMed), 0.319 and 0.335 (CentMed), 0.274 and 0.303 (EastMed), and 0.354 and 0.354 (BalBSea), from the El Chichón and Pinatubo eruptions, respectively.
For the first time, it was demonstrated that the TAOD550,WVE time series, i.e., atmospheric aerosols without volcanic materials but including anthropogenic activities, showed rather constant levels (no significant trends) over any of the four Mediterranean subregions.
The desert-dust and black-carbon aerosols account for half of the aerosol loading in the broader area of the Mediterranean region, especially over the Mediterranean Sea and the surrounding countries, except the Balkans and the Black Sea.
It was discovered that the DDAOD550 and SAOD550 time series consistently occurred during cold La Niňa occurrences. Additionally, the levels of DDAOD550 approached those of AAOD550 as one moved from WestMed to BalBSea, suggesting that the absorbing “face” of DDAOD550 may be taking a central role. Accordingly, from WestMed to BalBSea, the absorption aerosols’ contribution to the overall solar radiation attenuation decreases significantly as the scattering effect gradually replaces it. Furthermore, in the whole Mediterranean, scatterers such as nitrates, sea-salt, and sulphate particles are essentially the only elements reducing solar radiation.
In agreement with observations made by Sannino et al. [73] in Naples, Italy, who discovered that the AOD500 levels due to Saharan dust transport over the area on 25 February 2021 increased because of a coincident arrival of volcanic debris from Etna on the same day, the volcanic-debris-free aerosols showed lower (almost half) AOD values than those when volcanic debris was present over the Mediterranean.
Compared to the other three Mediterranean subregions, the Balkans and Black Sea appear to have been more affected by global volcanism between 1980 and 2024, and they also saw the least amount of aerosol loading in the absence of volcanoes. Accordingly, over the Mediterranean, the free-of-volcanic-debris aerosols displayed lower (almost half) AOD values than those with volcanic debris.
The dependence of TAOD550 on the ENSO and NAO phases (see Figure 6), particularly over the BalBSea subregion (see Figure 6d), may be largely explained by the foregoing divergence among the four Mediterranean subregions.
Several patterns demonstrated how each of the factors being studied are influenced by ENSO and NAO. Table 7 summarises the NAO and ENSO phases that have supported the highest parameter values in the four Mediterranean subregions between 1980 and 2024. The highest value of an atmospheric parameter occurred across adjacent subregions for the same NAO/ENSO phase pair if it occurred over a subregion when NAO and ENSO were in the same phase (same colour in Table 7). The same is true when the ENSO and NAO phenomena are at different stages (seen in Table 7 as distinct colours). For instance, over WestMed, neutral and/or positive ENSO phases may be present when the highest COD values occurred under negative NAO. This can be linked to the occurrence of COD during the transition from positive to neutral ENSO phase, while NAO is in a negative phase. Figure 14 shows such instances for DARF. Both observational and model data suggest that there are more instances of negative NAOI occurrences during warm ENSO and positive NAOI occurrences during cold ENSO events, according to Li and Lau’s [106] study on the effects of ENSO on atmospheric parameters (precipitation and surface air temperature) over the North Atlantic during late winter. Their findings occasionally coincide with those of the current study, albeit being for various parameters and seasons (cf. AAOD550, BCAOD550, and AE470–870 in BalBSea; SSA in all four subregions; COD in WestMed and CentMed).
Negative and neutral ENSO periods were found to benefit most of the annual TAOD550 values.
A TAOD550 trend of −0.017/decade (WestMed), −0.024/decade (CentMed), −0.013/decade (EastMed), and −0.037/decade (BalBSea) was discovered.
Nearly an order of magnitude lower than that of TAOD550, the trends in TAOD550,WVE were found to be +0.003/decade (WestMed), +0.003/decade (CentMed), +0.006/decade (EastMed), and +0.002/decade (BalBSea).
In the absence of worldwide volcanic eruptions, the atmospheric aerosol loading over the whole Mediterranean region from 1980 to 2024 was shown to exhibit a very slight but negligible trend. Additionally, it was discovered that in most circumstances, the aerosol loading over the larger Mediterranean region “prefers” neutral ENSO phases.
The stratospheric/sea-salt/cloud effect and the desert-dust/black-carbon aerosols were shown to have a nearly 50%/50% balance on the overall aerosol loading in WestMed, CentMed, and EastMed.
In the four subregions (TAOD550 ≈ 17·AAOD550, WestMed; TAOD550 ≈ 19·AAOD550, CentMed, EastMed; and TAOD550 ≈ 23·AAOD550, BalBSea), TAOD550 was expressed as a function of AAOD550 after [38].
All of the AE470–870 results showed coarse-mode particles throughout the Mediterranean on an annual basis, regardless of the ENSO phase. Only the years 1983 and 1992, which corresponded to the volcanic eruptions of Mt. El Chichón and Mt. Pinatubo, respectively, were associated with fine-mode aerosols. Furthermore, in all four Mediterranean subregions, fine-mode aerosols were preferred by positive ONI and positive/negative NAOI phases.
Over the Mediterranean region, the majority of aerosol types are classified as MTA or UIA. The BalBSea region appears to be less impacted by desert-dust particles.
  • Part B. Statistical approach
The Pearson’s correlation coefficient, r, between any of the studied parameters and any of the three physical phenomena was examined for |r| ≥ 0.2. It is interesting to observe that the correlations occurred at higher CI over WestMed and CentMed for the parameter pairs indicated, thus dictating a greater accuracy in these relationships. Nevertheless, it is seen that SSN, NAOI, and ONI influence EastMed more than the other three subregions, regardless of the CI level (see column 4, Table A8). Moreover, SSN also influences CentMed and BalBSea. These observations may result in a general conclusion that all three physical phenomena “prefer” the broader Eastern Mediterranean region. On the other hand, these r values cannot be considered high; therefore, the indicated relationships are rather weak.
The μ, σ, and m values of TAOD550, AAOD550, SAOD550, SSAOD550, BCAOD550, and SSA across the four Mediterranean subregions have comparable magnitudes. This indicates the existence of similar aerosol loading patterns over the four areas. On the contrary, DDAOD550, StAOD550, and AE470–870 show respective decreasing, increasing, and increasing trends from WestMed to BalBSea.
Trends, α (units/decade) > |0.2| (see Table 5) were found to be exhibited by DARF and netLWRBOA,CS,A over all four Mediterranean subregions, netSWRBOA,CS,A over WestMed, CentMed, BalBSea, and COD over EastMed and BalBSea. The other parameters showed negligible trends in the period 1980–2024.
The Balkan Peninsula and Black Sea region appear to be more affected by stratospheric aerosols, whereas West Mediterranean is mostly impacted by desert-dust outbreaks from the Sahara Desert. As indicated by increasing AE470–870 readings, fine-mode aerosols gradually take over from WestMed to BalBSea.
It was shown that any one of the three physical phenomena, or a combination of them, had a signalling influence on the patterns of the atmospheric aerosols over the Mediterranean region, regardless of how powerful or weak.
In the Mediterranean atmosphere, a Granger-causality test of the SSN/NAOI/ONI time series on those of the aerosol parameters revealed a predominantly, albeit delayed, six-month influence of the former on the latter. This delay is precisely half a year and nearly ½ of the 11-year solar activity cycle. This result is shown in Figure 14.
P-P plots between any of the three physical phenomena and one of the parameters under study demonstrated that the latter had an immediate impact on the former. It is true that the distribution of the corresponding values of the parameters is dominated by the monthly values of SSN, NAOI, and ONI. Regardless of the strength of the relationship, this fact determines the dominance of the physical phenomena on the parameters.
The absolute impact of the volcanic material released into the atmosphere on the optical characteristics of the aerosols across the larger Mediterranean region was one of the work’s key findings. This finding was validated by Figure 5. The current inquiry concerns the extent to which volcanoes impact the world’s climate. Volcanic eruptions were not taken into account in the most recent Intergovernmental Panel on Climate Change (IPCC, Sixth Assessment Report—IPCC (https://www.ipcc.ch/assessment-report/ar6/ (accessed on 10 February 2025))) 6th Assessment Report (AR6), which was partially released between August 2021 and April 2022. Instead, AR6 makes the assumption that volcanic activity between 2015 and 2100 will be comparable to that which was recorded between 1850 and 2014. This assumption is vague in the sense that, as shown by comparing Figure 3 and Figure 5, the volcanic activity is crucial in modulating TAOD550 levels and trends; different TAOD550 levels in the atmosphere may result in differing scattering and absorption mechanisms of solar radiation and, consequently, in reforming DARF. This can be demonstrated with basic computations. Examine the four Mediterranean subregions’ inter-period average values of TAOD550 and DARF in Table 5 and TAOD550,WVE in Table 8. The formula DARFWVE = DARF·(TAOD550,WVE/TAOD550) can be used to predict DARFWVE (DARF without volcanic eruptions). Table 8 shows that DARFWVE levels can rise dramatically, suggesting that aerosols play a more significant influence in the global climate. The following is the explanation: a decrease in (stratospheric) aerosols in the atmosphere implies a rise in the downward flux of solar radiation on Earth’s surface, which raises netSWRBOA. Thus, DARFWVE rises in accordance with Equations (1)–(4), so netLWRBOA may be regarded as nearly constant in the equations.
The dispersion of the different types of aerosols throughout the larger Mediterranean region was the second result of this study. Due to the region’s closeness to the African deserts, desert-dust aerosols are either non-existent or extremely uncommon over the BalBSea region; however, they are more common over WestMed. Because WestMed is bordered by the Atlantic Ocean and the Mediterranean, westerlies have the ability to carry sea-spray aerosols from the Atlantic to the region. This is also true for aerosols of marine origin. However, CentMed has the disadvantage of having a higher concentration of aerosols of urban and industrial origin due to its urbanisation and industrialisation. However, a number of factors (coal-fired plants and low-quality solid fuels, obsolete industrial facilities, an ageing fleet of vehicles, inefficient heating systems, and complex topography; see www.blue-europe.eu and https://westernbalkans-infohub.eu) contribute to the higher levels of UIA over the Balkans and Black Sea.
The influence of solar activity, the North Atlantic Oscillation, and the El Niňo–Southern Oscillation on atmospheric aerosols across the larger Mediterranean region, particularly with their six-month delayed effect, can be summarised as the study’s final significant finding. These postponed impacts of the physical phenomena, either separately or collectively, in the climate models could provide fresh perspectives on the projections of climate change. The interconnectedness of these three phenomena has been examined by numerous researchers. According to Vencloviene et al. [55], the La Niňa phase has a statistically significant effect on NAOI. Leamon [37] discovered a very statistically significant association between the occurrence of the last five solar cycle termination and the Pacific Ocean’s shift from El Niño to La Niña. The impact of solar modulation on the relationship between NAO and ENSO has been investigated by Huo et al. [38].
As regards the research questions posed in the end of Section 1, the analysis of the present study can answer them quite confidently. They are given below.
  • Are the effects of solar activity, NAO, and ENSO on the atmospheric processes found in [31] valid in smaller areas of the Mediterranean too? The answer to this question is positive. Table A2 and Table A3 (see Appendix A) provide sufficient correlations (positive or negative) at high CIs over almost all four Mediterranean domains considered.
  • How do volcanic eruptions affect the status of atmospheric aerosols over the Mediterranean? Extensive analysis provided in Section 3.1.2 showed that the absence of volcanic activity on Earth would result in rather constant levels of the atmospheric processes over all four Mediterranean domains.
  • A third significant conclusion related to the first research question was that El Niňo–Southern Oscillation, North Atlantic Oscillation, and solar activity all have an impact on the atmospheric processes over the larger Mediterranean region, either separately or in combination, and this effect can last up to six months. The three events have a weak but significant cause-and-effect relationship that drives several aerosol characteristics, even at a 99.9% significance level.
From the above conclusions, some areas that may require more research and are deserving of future study are, therefore, brought to light.
  • Global volcanism.
    • Look closely at how global volcanism affects the characteristics of aerosols in regions other than the Mediterranean.
    • In the absence of volcanic eruptions, study again the Earth’s climate. Forecast the future and model the past and present world climates devoid of volcanic debris. Since the current study at least showed that large volcanoes have an impact on global aerosol levels and their concentrations as well as on the climate, and that they should not be ignored, there should be alternatives to the “as was in the past” scenario for the years 2015–2100 in light of the IPCC’s upcoming AR7. Thus, the following volcanic scenarios should be considered: (i) unchanged; (ii) absence of dramatic volcanoes; (iii) presence of drastic volcanoes by including extreme volcanism of VEIs of 5 or more (cataclysmic category). Furthermore, these scenarios could potentially be broken down and examined into smaller timeframes within the future timeframe of 2015–2100.
  • TAOD550 levels.
    • In BalBSea, they are higher for ONI < 0 than for ONI > 0 compared to the other three Mediterranean domains. Then, examine whether this phenomenon holds true for other bordering regions worldwide.
    • Examine why they prefer to occur during neutral ENSO phases.
  • ENSO, NAO effects.
    • Examine how ENSO regulates the circulation of atmospheric aerosols outside of the Mediterranean region.
    • Examine how solar activity, NAO, and ENSO affect atmospheric processes in various regions of the planet.
    • Examine why black-carbon aerosols have a greater impact over BalBSea than volcanic debris. The sporadic occurrence of the latter type of aerosols across the area could be one factor.

Funding

This research received no external funding.

Data Availability Statement

Monthly data of netSWR, netLWR, AOD550, AE470–870, and COD was downloaded from the Giovanni platform free-of-charge (access at https://giovanni.gsfc.nasa.gov/giovanni in the period 8–25 October 2023). Monthly data of NOAI and ONI was downloaded from the NOAA website free-of-charge (access at https://www.cpc.ncep.noaa.gov/products/precip/CWlink/pna/norm.nao.monthly.b5001.current.ascii.table for NOAI, and: https://cpc.ncep.noaa.gov/products/analysis_monitoring/ensostuff/ONI_v5.php for ONI, in the period 8–25 October 2023). Monthly data of SSN was retrieved from the Royal Observatory of Belgium (ROB) free-of-charge (access at https://www.sidc.be/SILSO/datafiles on 18 October 2023).

Acknowledgments

The author is thankful to the personnel of GSFC/NASA, NPC/NCEP/NOAA, and ROB for preparing the databases used.

Conflicts of Interest

The author declares no conflicts of interest.

Appendix A

This section provides information which was not considered crucial to be embedded in the main body of the paper.

Appendix A.1. Energetic Volcanoes

Table A1 gives a list of the energetic volcanoes in the period of the study.
Table A1. List of volcanoes with volcanic explosivity index, VEI, in the period 1980–2024. Note that various volcanoes have been active through several years.
Table A1. List of volcanoes with volcanic explosivity index, VEI, in the period 1980–2024. Note that various volcanoes have been active through several years.
YearVolcano, CountryVEI
1980St. Helens, USA
Etna, Italy
Stromboli, Italy
5
2
2
1981Pagan, USA
Alaid, Russia
St. Helens, USA
Etna, Italy
Stromboli, Italy
4
4
5
2
2
1982Galunggung, Indonesia
El Chichón, Mexico
Pagan, Mexico
St. Helens, USA
Etna, Italy
Stromboli, Italy
4
5
4
5
2
2
1983Colo, Indonesia
Galunggung, Indonesia
Pagan, USA
St. Helens, USA
Etna, Italy
Stromboli, Italy
4
4
4
5
2
2
1984Pagan, USA
St. Helens, USA
Etna, Italy
Stromboli, Italy
4
5
2
2
1985Pagan, USA
St. Helens, USA
Etna, Italy
Stromboli, Italy
4
5
2
2
1986Chikurachki, Russia
Klyuchevskoy, Russia
St. Helens, USA
Augustine, USA
Etna, Italy
Stromboli, Italy
4
4
5
4
2
2
1987Klyuchevskoy, Russia
Etna, Italy
Stromboli, Italy
4
2
2
1988Klyuchevskoy, Russia
Etna, Italy
Stromboli, Italy
4
2
2
1989Klyuchevskoy, Russia
Etna, Italy
Stromboli, Italy
4
2
2
1990Kelud, Indonesia
Klyuchevskoy, Russia
Etna, Italy
Stromboli, Italy
4
4
2
2
1991Cerro Hudson, Chile
Pinatubo, Philippines
Etna, Italy
Stromboli, Italy
5
6
2
2
1992Spurr, USA
Etna, Italy
Stromboli, Italy
4
2
2
1993Lascar, Chile
Etna, Italy
Stromboli, Italy
4
1
2
1994Rabaul, Papua N. Guinea
Etna, Italy
Stromboli, Italy
4
3
2
1995Rabaul, Papua N. Guinea
Etna, Italy
Stromboli, Italy
4
3
2
1996Etna, Italy
Stromboli, Italy
3
2
1997Etna, Italy
Stromboli, Italy
3
2
1998Etna, Italy
Stromboli, Italy
3
2
1999Sheveluch, Russia
Etna, Italy
Stromboli, Italy
4
3
2
2000Ulawun, Papua N. Guinea
Sheveluch, Russia
Etna, Italy
Stromboli, Italy
4
4
3
2
2001Sheveluch, Russia
Etna, Italy
Stromboli, Italy
4
3
2
2002Reventador, Ecuador
Ruang, Indonesia
Sheveluch, Russia
4
4
4
2003Reventador, Ecuador
Sheveluch, Russia
4
4
2004Manam, Papua N. Guinea
Sheveluch, Russia
4
4
2005Manam, Papua N. Guinea
Sheveluch, Russia
4
4
2006Rabaul, Papua N. Guinea
Manam, Papua N. Guinea
Sheveluch, Russia
4
4
4
2007Rabaul, Papua N. Guinea
Manam, Papua N. Guinea
Sheveluch, Russia
4
4
4
2008Kasatochi, USA
Okmok, USA
Chaiten, Chile
Rabaul, Papua N. Guinea
Manam, Papua N. Guinea
Sheveluch, Russia
4
4
4
4
4
4
2009Sarychev Peak, Russia
Chaiten, Chile
Rabaul, Papua N. Guinea
Manam, Papua N. Guinea
Sheveluch, Russia
4
4
4
4
4
2010Merapi, Indonesia
Eyjafjallajokull, Iceland
Chaite, Chile
Rabaul, Papua N. Guinea
Sheveluch, Russia
4
4
4
4
4
2011Nabro, Eritrea
Puyehue-Cordon Caulle, Chile
Grimsvotn, Iceland
Chaite, Chile
Sheveluch, Russia
4
5
4
4
4
2012Nabro, Eritrea
Puyehue-Cordon Caulle, Chile
Sheveluch, Russia
4
5
4
2013Sinabung, Indonesia
Etna, Italy
Sheveluch, Russia
Stromboli, Italy
4
3
4
2
2014Manam, Papua N. Guinea
Semeru, Indonesia
Kelud, Indonesia
Sinabung, Indonesia
Etna, Italy
Sheveluch, Russia
4
4
4
4
3
4
2015Wolf, Ecuador
Calbuco, Chile
Manam, Chile
Semeru, Indonesia
Sinabung, Indonesia
Sheveluch, Russia
Etna, Italy
4
4
4
4
4
4
3
2016Manam, Papua N. Guinea
Semeru, Indonesia
Sinabung, Indonesia
Etna, Italy
Sheveluch, Russia
4
4
4
3
4
2017Semeru, Indonesia
Manam, Papua N. Guinea
Sinabung, Indonesia
Etna, Italy
Sheveluch, Russia
4
4
4
3
4
2018Manam, Papua N. Guinea
Semeru, Indonesia
Sinabung, Indonesia
Etna, Italy
Sheveluch, Russia
Stromboli, Italy
4
4
4
3
4
2
2019Ulawun, Papua N. Guinea
Sinabung, Indonesia
Manam, Papua N. Guinea
Etna, Italy
Semeru, Indonesia
Sheveluch, Russia
Stromboli, Italy
4
4
4
3
4
4
2
2020Soufriere, Saint Vincent and the Grenadines
Taal, Philippines
Manam, Papua N. Guinea
Semeru, Indonesia
Etna, Italy
Sheveluch, Russia
Stromboli, Italy
4
4
4
4
3
4
2
2021Hunga Tonga-Hunga Ha’apai, Tonga
Fukutu-Oka-no-Ba, Japan
Soufriere, Saint Vincent and the Grenadines
Manam, Papua N. Guinea
Semeru, Indonesia
Sheveluch, Russia
Etna, Italy
Stromboli, Italy
5
4
4
4
4
4
3
2
2022Hung Tonga-Hunga Ha’apai, Tonga
Manam, Papua N. Guinea
Semeru, Indonesia
Sheveluch, Russia
Etna, Italy
Stromboli, Italy
5
4
4
4
3
2
2023Manam, Papua N. Guinea
Semeru, Indonesia
Sheveluch, Russia
Etna, Italy
Stromboli, Italy
4
4
4
2
2
2024Manam, Papua N. Guinea
Semeru, Indonesia
Sheveluch, Russia
Etna, Italy
Stromboli, Italy
4
4
4
2
2

Appendix A.2. Pearson’s Correlation Coefficient

Table A2 and Table A3 give a list of the Pearson’s correlation coefficients among the subregions and among the various parameters considered in the present study.
Table A2. Pearson’s correlation coefficient, r, between the monthly mean AE470–870 and the monthly mean DARF values over the four Mediterranean domains in the period 1980–2024. All correlations are significant at the 99.9% CI. DARF has units of Wm−2.
Table A2. Pearson’s correlation coefficient, r, between the monthly mean AE470–870 and the monthly mean DARF values over the four Mediterranean domains in the period 1980–2024. All correlations are significant at the 99.9% CI. DARF has units of Wm−2.
ParameterPair of Mediterranean Areasr
AE470–870WestMed, CentMed0.777
WestMed, EastMed0.706
WestMed, BalBSea0.678
CentMed, EastMed0.935
CentMed, BalBSea0.895
EastMed, BalBSea0.932
WestMed, allMed0.887
CentMed, allMed0.968
EastMed, allMed0.944
BalBSea, allMed0.895
DARFWestMed, CentMed0.899
WestMed, EastMed0.797
WestMed, BalBSea0.871
CentMed, EastMed0.932
CentMed, BalBSea0.920
EastMed, BalBSea0.913
WestMed, allMed0.950
CentMed, allMed0.983
EastMed, allMed0.938
BalBSea, allMed0.939
Table A3. Pearson’s correlation coefficient, r, and coefficient of determination, R2, for any parameter considered in the present study with either SSN, NAOI, or ONI over the four Mediterranean domains in the period 1980–2024. Both r and R2 statistical indices are applied to the monthly averaged values of the parameters. R2 refers to the linear regression fits between the independent variable (any aerosol or irradiance parameter) as a function of the dependent variable (SSN, NAOI, or ONI). The annual averages of DARF, netSWR, and netLWR are expressed in Wm−2. Note that the r or R2 values with asterisk(s) indicate the significance level (* for CI = 95%, or *** for CI = 99.9%). All values are rounded to the third decimal digit.
Table A3. Pearson’s correlation coefficient, r, and coefficient of determination, R2, for any parameter considered in the present study with either SSN, NAOI, or ONI over the four Mediterranean domains in the period 1980–2024. Both r and R2 statistical indices are applied to the monthly averaged values of the parameters. R2 refers to the linear regression fits between the independent variable (any aerosol or irradiance parameter) as a function of the dependent variable (SSN, NAOI, or ONI). The annual averages of DARF, netSWR, and netLWR are expressed in Wm−2. Note that the r or R2 values with asterisk(s) indicate the significance level (* for CI = 95%, or *** for CI = 99.9%). All values are rounded to the third decimal digit.
Parameters in PairrR2
WestMedCentMedEastMedBalBSeaWestMedCentMedEastMedBalBSea
NAOI-ONI+0.0110.000
NAOI-SSN+0.0610.004
ONI-SSN+0.0180.000
TAOD550-NAOI−0.059−0.017+0.205+0.2310.0030.0000.0420.053
TAOD550-ONI+0.084+0.090 *+0.201+0.1930.0070.008 *0.0400.037
TAOD550-SSN+0.173 ***+0.206 ***+0.293+0.368 *0.030 ***0.043 ***0.0860.135 *
AAOD550-NAOI−0.152 ***−0.146 ***−0.285−0.0310.023 ***0.021 ***0.0820.001
AAOD550-ONI−0.053−0.072−0.306 *−0.1220.0030.0050.093 *0.015
AAOD550-SSN+0.019+0.013−0.224+0.1060.0000.0000.0500.011
SAOD550-NAOI−0.052−0.009−0.215+0.2320.0030.0000.0460.054
SAOD550-ONI+0.092 *+0.098 *+0.211+0.1950.008 *0.010 *0.0450.038
SAOD550-SSN+0.181 ***+0.215 ***+0.299 *+0.367 *0.033 ***0.046 ***0.089 *0.135 *
DDAOD550-NAOI−0.144 ***−0.108 *−0.284−0.1910.021 ***0.012 *0.0810.037
DDAOD550-ONI−0.059−0.098 *−0.284 *−0.333 *0.0040.010 *0.110 *0.111 *
DDAOD550-SSN+0.052+0.014−0.011−0.0430.0030.0000.0000.002
StAOD550-NAOI+0.101 *+0.121 *+0.295 *+0.2310.10 *0.015 *0.087 *0.053
StAOD550-ONI+0.225 ***+0.183 ***+0.295 *+0.2130.051 ***0.034 ***0.087 *0.046
StAOD550-SSN+0.260 ***+0.236 ***+0.323 *+0.353 *0.068 ***0.056 ***0.104 *0.124 *
BCAOD550-NAOI−0.108 *−0.117 *−0.188+0.0260.012 *0.014 *0.0350.001
BCAOD550-ONI−0.034−0.025−0.144+0.0210.0010.0010.0210.000
BCAOD550-SSN−0.110 *−0.057−0.271+0.1150.012 *0.0030.0730.013
SSAOD550-NAOI−0.107 *−0.061−0.212−0.2780.012 *0.0040.0450.077
SSAOD550-ONI−0.074+0.081−0.082−0.1330.0050.0070.0070.018
SSAOD550-SSN−0.165 ***+0.126 *−0.306 *−0.401 *0.027 ***0.016 *0.093 *0.160 *
AE470–870-NAOI+0.197 ***+0.160 ***+0.324 *+0.314 *0.039 ***0.026 ***0.105 *0.098 *
AE470–870-ONI+0.219 ***+0.233 ***+0.331 *+0.2790.048 ***0.054 ***0.109 *0.078
AE470–870-SSN+0.131 *+0.167 ***+0.268+0.2920.017 *0.028 ***0.0720.085
SSA-NAOI+0.152 ***+0.213 ***+0.301 *+0.2020.023 ***0.045 ***0.090 *0.041
SSA-ONI+0.182 ***+0.186 ***+0.287+0.2100.033 ***0.035 ***0.0820.044
SSA-SSN+0.246 ***+0.239 ***+0.306 *+0.2930.060 ***0.057 ***0.094 *0.086
DARF-NAOI−0.161 ***−0.150 ***−0.294 *−0.0250.026 ***0.022 ***0.086 *0.001
DARF-ONI−0.023−0.037−0.334 *−0.1830.0010.0010.111 *0.034
DARF-SSN+0.016+0.012−0.228+0.0810.0000.0000.0520.007
netSWRBOA,CS,A-NAOI−0.194 ***−0.195 ***−0.027−0.0460.138 ***0.038 ***0.0010.002
netSWRBOA,CS,A-ONI−0.013+0.013−0.098−0.1430.0000.0000.0100.020
netSWRBOA,CS,A-SSN+0.019+0.018−0.088−0.2720.0000.0000.0080.074
netLWRBOA,CS,A-NAOI+0.145 ***−0.183 ***−0.154−0.2660.021 ***0.033 ***0.0240.071
netLWRBOA,CS,A-ONI+0.031+0.069+0.106+0.0740.0010.0050.0110.005
netLWRBOA,CS,A-SSN−0.064+0.013−0.237−0.1560.0020.0000.0560.004
COD-NAOI+0.076+0.235 ***−0.042−0.316 *0.0060.055 ***0.0020.100 *
COD-ONI−0.023−0.030+0.200+0.0050.0010.0010.0400.000
COD-SSN−0.080−0.072−0.113−0.0830.0060.0050.0130.007

Appendix A.3. Time Series Stationarity

The level/trend-stationarity results obtained by the KPSS test on the investigated aerosol/atmospheric variables are shown in Table A4. For those variables, stationarity is satisfied (hypothesis Ho) if the p-value is less than the chosen CI level, that is, p ≤ α = 0.1 for 90% CI, p ≤ 0.05 for CI = 95%, p ≤ 0.025 for 97.5% CI, and p ≤ 0.01 for CI = 99%. The KPSS statistic (either KPSSl for level stationarity or KPSSt for trend stationarity) is less than corresponding critical values, KPSScl and KPSSct, respectively, for KPSSl and KPPSt. According to the CI levels listed above, the KPSScl values for level stationarity are 0.347, 0.463, 0.574, and 0.739, and the KPSSct values for trend stationarity are 0.119, 0.146, 0.176, and 0.216 [105]. Unless a warning for “first-order difference” is noted, the findings demonstrate stationarity for the time series of the variables under investigation.
Table A4. KPSS statistics on the aerosol/atmospheric parameter, q, over the four Mediterranean domains. Unprocessed (non-dimensionalised) monthly mean values of the parameters in the period 1980–2024 were used. The values presented in the Table are KPSSl, KPSSt, truncation lag, and stationarity conclusion at the prescribed CI level. Level stationarity exists for KPSSl < 0.347, 0.463, 0.574, and 0.739, and trend stationarity for KPSSt < 0.119, 0.146, 0.176, 0.216 at the 90%, 95%, 97.5%, and 99% CI, respectively.
Table A4. KPSS statistics on the aerosol/atmospheric parameter, q, over the four Mediterranean domains. Unprocessed (non-dimensionalised) monthly mean values of the parameters in the period 1980–2024 were used. The values presented in the Table are KPSSl, KPSSt, truncation lag, and stationarity conclusion at the prescribed CI level. Level stationarity exists for KPSSl < 0.347, 0.463, 0.574, and 0.739, and trend stationarity for KPSSt < 0.119, 0.146, 0.176, 0.216 at the 90%, 95%, 97.5%, and 99% CI, respectively.
Parameter, qWestMedCentMedEastMedBalBSea
NAOI0.2327, 0.1148, 6, level/trend stationarity at 90%/90% CI
ONI0.1073, 0.0507, 6, level/trend stationarity at 90%/90% CI
SSN1.2061, 0.1620, 6, level/trend non-stationarity at 99%/96.3% CI (needs deseasonalisation)
TAOD5501.4767, 0.0558, 6, trend stationarity at 90% CI2.196, 0.1359, 6, trend stationarity at 93.1% CI1.4006, 0.1408, 6, trend stationarity at 94% CI4.8332, 0.4187, 6 (needs deseasonalisation)
AAOD5500.3933, 0.0480, 6, level/trend stationarity at 92%/90% CI0.4635, 0.0556, 6, level/trend stationarity at 95%/90% CI1.2862, 0.1582, 6, (needs deseasonalisation)0.2097, 0.2160, 6, level stationarity at 99% CI
SAOD5501.6464, 0.0628, 6, trend stationarity at 90% CI2.6988, 0.1472, 6 (needs deseasonalisation)1.5666, 0.1461, 6, trend stationarity at 95% CI4.9290, 0.4260, 6 (needs deseasonalisation)
DDAOD5500.0895, 0.0780, 6, level/trend stationarity at 90%/90% CI0.1610, 0.0790, 6, level/trend stationarity at 90%/93.4% CI0.2178, 0.0669, 6, level/trend stationarity at 90%/90% CI0.1313, 0.1017, 6, level stationarity at 90%/90% CI
StAOD5502.9796, 0.1888, 6 (needs deseasonalisation)3.9768, 0.1826, 6 (needs deseasonalisation)2.7110, 0.2011, 6 (needs deseasonalisation)5.5435, 0.5583, 6 (needs deseasonalisation)
SSAOD5503.8573, 0.2145, 6 (needs deseasonalisation)1.6125, 0.4173, 6 (needs deseasonalisation)3.6875, 0.0992, 6, trend stationarity at 90% CI3.8994, 0.1983, 6 (needs deseasonalisation)
BCAOD5503.8221, 0.1895, 6 (needs deseasonalisation)2.5526, 0.3679, 6 (needs deseasonalisation)5.4606, 0.2255, 6 (needs deseasonalisation)0.4781, 0.4447, 6 (needs deseasonalisation)
AE470–8702.4305, 0.3405, 6 (needs deseasonalisation)2.5526, 0.3679, 6 (needs deseasonalisation)2.6150, 0.2978, 6 (needs deseasonalisation)3.6170, 0.4681, 6 (needs deseasonalisation)
SSA5.2424, 0.3160, 6 (needs deseasonalisation)5.2485, 0.3705, 6 (needs deseasonalisation)
4.5733, 0.3567, 6 (needs deseasonalisation)5.8728, 0.3991, 6 (needs deseasonalisation)
DARF0.2874, 0.0318, 6, level/trend non-stationarity at 90%/90% CI0.2355, 0.0266, 6, trend stationarity at 90%/90% CI0.9382, 0.0671, 6, trend stationarity at 99%/99% CI0.1046, 0.0553, 6, level/trend stationarity at 90%/90% CI
netSWRBOA,CS,A0.0037, 0.0037, 6, level/trend stationarity at 90%/90% CI0.0040, 0.0038, 6, trend stationarity at 90%/90% CI0.0014, 0.0037, 6, level/trend stationarity at 90%/90% CI0.0071, 0.0040, 6, level/trend stationarity 90%/90% CI
netLWRBOA,CS,A0.1023, 0.0138, 6, level/trend stationarity at 90%/90% CI0.3624, 0.0520, 6, level/trend stationarity at 90.7%/90% CI2.4677, 0.0491, 6, trend stationarity at 90% CI1.5676, 0.0808, 6, trend stationarity at 90% CI
COD0.0935, 0.0832, 6, level/trend stationarity at 90%/90% CI0.0905, 0.0879, 6, level/trend stationarity at 90%/90% CI0.0460, 0.0341, 6, level/trend stationarity at 90%/90% CI0.1542, 0.1033, 6, level/trend stationarity at 90%/90% CI

Appendix A.4. Time Series Stationarity After Deseasonalisation

According to the results in Table A5, some time series need deseasonalisation. This was performed by using the STL method mentioned in Section 3.2.4. Here, stationarity results are given for the required parameters.
Table A5. KPSS statistics for those aerosol/atmospheric parameters, q, that are mentioned in Table A4. The STL method was applied to only those time series before the KPSS test.
Table A5. KPSS statistics for those aerosol/atmospheric parameters, q, that are mentioned in Table A4. The STL method was applied to only those time series before the KPSS test.
Parameter, qWestMedCentMedEastMedBalBSea
SSN1.2061, 0.1620, 6, level/trend non-stationarity at 99%/96.3% CI (needs 1st-order differencing)
TAOD550no further processing is necessaryno further processing is necessaryno further processing is necessary5.3726, 0.5543, 6 (needs 1st-order differencing)
AAOD550no further processing is necessaryno further processing is necessary2.5479, 0.3662, 6 (needs 1st-order differencing)no further processing is necessary
SAOD550no further processing is necessary3.3870, 0.1956, 6 (needs 1st-order differencing)no further processing is necessary5.3992, 0.5460, 6 (needs 1st-order differencing)
StAOD5502.7938, 0.1749, 6 (needs 1st-order differencing)3.8396, 0.1733, 6 (needs 1st-order differencing)2.5888, 0.1616, 6 (needs 1st-order differencing)5.5774, 0.5635, 6 (needs 1st-order differencing)
SSAOD5504.6631, 0.2990, 6 (needs deseasonalisation)1.5834, 0.4151, 6 (needs 1st-order differencing)no further processing is necessary4.7182, 0.2903, 6 (needs 1st-order differencing)
BCAOD5505.5608, 0.4434, 6 (needs 1st-order differencing)4.2254, 0.5087, 6 (needs 1st-order differencing)4.9921, 0.5311, 6 (needs 1st-order differencing)0.8177, 0.7104, 6 (needs 1st-order differencing)
AE470–8702.5659, 0.3821, 6 (needs 1st-order differencing)2.6225, 0.3724, 6 (needs 1st-order differencing)2.7621, 0.3072, 6 (needs 1st-order differencing)3.8136, 0.5049, 6 (needs 1st-order differencing)
SSA5.3253, 0.3308, 6 (needs 1st-order differencing)5.3962, 0.3967, 6 (needs 1st-order differencing)4.7512, 0.3834, 6 (needs 1st-order differencing)6.0146, 0.4352, 6 (needs 1st-order differencing)

Appendix A.5. Time Series Stationarity After Deseasonalisation and First-Order Differencing

Some time series required first-order differencing, as indicated by the results in Table A5. To do this, each value in the time series was subtracted from the one before it. Here, the parameters and the Mediterranean, for which such a modification was indicated in Table A5, have stationarity results in Table A6.
Table A6. KPSS statistics for those aerosol/atmospheric parameters, q, that required first-order differencing (see Table A2).
Table A6. KPSS statistics for those aerosol/atmospheric parameters, q, that required first-order differencing (see Table A2).
Parameter, qWestMedCentMedEastMedBalBSea
SSN0.1369, 0.0463, 6, level/trend stationarity at 90%/90% CI
TAOD550no further processing is necessaryno further processing is necessaryno further processing is necessary0.0101, 0.0103, 6 level/trend stationarity at 90%/90% CI
AAOD550no further processing is necessaryno further processing is necessary0.0117, 0.0113, 6 level/trend stationarity at 90%/90% CIno further processing is necessary
SAOD550no further processing is necessary0.0086, 0.0083, 6 level/trend stationarity at 90%/90% CIno further processing is necessary0.0104, 0.0106, 6 level/trend stationarity at 90%/90% CI
StAOD5500.0145, 0.0139, 6, level/trend stationarity at 99%/97.4% CI0.0183, 0.0112, 6 level/trend stationarity at 90%/90% CI0.0160, 0.0150, 6 level/trend stationarity at 90%/90% CI0.0157, 0.0159, 6 level/trend stationarity at 90%/90% CI
SSAOD5500.0088, 0.0087, 6 level/trend stationarity at 90%/90% CI0.0369, 0.0131, 6 level/trend stationarity at 90%/90% CIno further processing is necessary0.0083, 0.0083, 6 level/trend stationarity at 90%/90% CI
BCAOD5500.0160, 0.0127, 6 level/trend stationarity at 90%/90% CI0.0109, 0.0095, 6 level/trend stationarity at 90%/90% CI0.0150, 0.0116, 6 level/trend stationarity at 90%/90% CI0.0185, 0.0120, 6 level/trend stationarity at 90%/90% CI
AE470–8700.0220, 0.02171, 6 level/trend stationarity at 90%/90% CI0.0161, 0.0158, 6 level/trend stationarity at 90%/90% CI0.0169, 0.0166, 6 level/trend stationarity at 90%/90% CI0.0158, 0.0160, 6 level/trend stationarity at 90%/90% CI
SSA0.0290, 0.0235, 6 level/trend stationarity at 90%/90% CI0.0254, 0.0253, 6 level/trend stationarity at 90%/90% CI0.0239, 0.0238, 6 level/trend stationarity at 90%/90% CI0.0232, 0.0231, 6 level/trend stationarity at 90%/90% CI

Appendix A.6. Granger-Causality Test

A linear regression analysis between each parameter q and the first six lagged components of the independent variables was carried out following deseasonalisation and first-order differencing to the time series shown in Table A4, Table A5 and Table A6, while retaining the first six lags of the independent time series of SSN/NAOI/ONI. Only those lagged components that had p-values > α = 0.4 (significant at the 60% CI) were considered in each regression expression. To investigate the impact of the El Niño phases on the dependent parameter q, the unlagged ONI time series was employed as a categorial factor in the regression analysis. Table A7 displays the findings. Only regression results with an overall α ≥ 0.4 (or 60% CI) were considered in this table; also, only parameter components that yielded a significance level ≥60% were kept in the regression analyses. The H1 hypothesis, according to the regression analysis, that the physical phenomena Granger-cause parameter q is justified by an overall R2 > 0.6 at a significance level ≥60% CI.
Table A7. Application of the GC test: linear regression analysis between the (dependent) parameter, q, and the (independent) lagged SSN/NAOI/ONI time series with categorial factor the unlagged ONI time series. The regression expression for each q is in the form as follows: q = a + b1·SSNt−1 + b2·SSNt−2 + b3·SSNt−3 + b4·SSNt−4 + b5·SSNt−5 + b6·SSNt−6 + c1·NAOIt−1 + c2·NAOIt−2 + c3·NAOIt−3 + c4·NAOIt−4 + c5·NAOIt−5 + c6·NAOIt−6 + d1·ONIt−1 + d2·ONIt−2 + d3·ONIt−3 + d4·ONIt−4 + d5·ONIt−5 + d6·ONIt−6 + e·t (t = month number from 1 to 540, i.e., from January 1980 to December 2024, respectively). The Ho hypothesis assumes no Granger-cause–effect of the lagged SSN/NAOI/ONI time series on the parameter q; the alternative hypothesis H1 is just the opposite. In the regression expression, the significant level for each component is shown in parenthesis. Deseasonalised monthly mean values of the time series in the period 1980–2024 were used in all four Mediterranean domains, and first-order differenced time series for the indicated dependent variables; a deseasonalised SSN time series was used throughout all GC tests.
Table A7. Application of the GC test: linear regression analysis between the (dependent) parameter, q, and the (independent) lagged SSN/NAOI/ONI time series with categorial factor the unlagged ONI time series. The regression expression for each q is in the form as follows: q = a + b1·SSNt−1 + b2·SSNt−2 + b3·SSNt−3 + b4·SSNt−4 + b5·SSNt−5 + b6·SSNt−6 + c1·NAOIt−1 + c2·NAOIt−2 + c3·NAOIt−3 + c4·NAOIt−4 + c5·NAOIt−5 + c6·NAOIt−6 + d1·ONIt−1 + d2·ONIt−2 + d3·ONIt−3 + d4·ONIt−4 + d5·ONIt−5 + d6·ONIt−6 + e·t (t = month number from 1 to 540, i.e., from January 1980 to December 2024, respectively). The Ho hypothesis assumes no Granger-cause–effect of the lagged SSN/NAOI/ONI time series on the parameter q; the alternative hypothesis H1 is just the opposite. In the regression expression, the significant level for each component is shown in parenthesis. Deseasonalised monthly mean values of the time series in the period 1980–2024 were used in all four Mediterranean domains, and first-order differenced time series for the indicated dependent variables; a deseasonalised SSN time series was used throughout all GC tests.
Parameter, qWestMedCentMedEastMedBalBSea
TAOD550overall regression
RMSE = 0.0706, R2 = 0.6510, p-value = 0.0002 (≈100% CI)
 
+0.1815 (96.5%)
+0.0002·SSNt−1 (63.9%)
−0.0004·SSNt−6 (98%)
−0.0072·ΝΑΟΙt (88.8%)
−0.0099·ΝΑΟΙt−1 (96.1%)
+0.0068·ΝΑΟΙt−3 (86.1%)
+0.0062·ΝΑΟΙt−4 (81.2%)
+0.0050·ΝΑΟΙt−5 (72%)
+0.0082·ΝΑΟΙt−6 (94.3%)
−0.0387·ONI t−1 (70.1%)
+0.0452·ONI t−2 (72.4%)
−9.2806 × 10−5·t (99.7%)
overall regression
RMSE = 0.0430, R2 = 0.5714, p-value = 0.1947 (80.5% CI)
 
+0.0001·SSNt−3 (67.4%)
+0.0002·SSNt−4 (96.8%)
+0.0037·ΝΑΟΙt (83.4%)
+0.0150·ΝΑΟΙt−3 (94%)
−0.0075·ΝΑΟΙt−4 (65%)
−0.0110·ΝΑΟΙt−5 (81.6%)
−0.0008·ONI t−2 (61.8%)
−0.0004·ONI t−4 (85%)
overall regression
RMSE = 0.0573, R2 = 0.5969, p-value = 0.0440 (95.6% CI)
 
+0.1216 (93.4%)
−0.0002·SSNt−6 (89.5%)
+0.0103·ΝΑΟΙt−3 (66.7%)
+0.0011·ONI t−2 (65.5%)
−0.0007·ONI t−3 (62.1%)
−0.0009·ONI t−4 (73.3%)
+0.0007·ONI t−6 (73.3%)
−8.4831 × 10−5·t (99.9%)
overall regression
RMSE = 0.0405, R2 = 0.5710, p-value = 0.1819 (81.8% CI)
 
−0.0438 (65.4%)
+0.0002·SSNt−4 (95.2%)
+0.0001·SSNt−5 (76.5%)
+0.0103·ΝΑΟΙt−1 (82%)
−0.0085·ΝΑΟΙt−4 (74.6%)
−0.0160·ΝΑΟΙt−5 (96%)
+0.0144·ΝΑΟΙt−6 (94.9%)
+0.0006·ONIt−1 (62.6%)
−0.0019·ONIt−2 (97.9%)
+0.0015·ONIt−3 (99.2%)
−0.0011·ONIt−4 (97.8%)
+0.0006·ONIt−4 (70.2%)
−2.6653 × 10−5·t (86.9%)
AAOD550overall regression
RMSE = 0.0044, R2 = 0.6508, p-value = 0.0002 (≈100% CI)
 
+0.0089 (90.1%)
+1.1025 × 10−5·SSNt (62.8%)
+2.5971 × 10−5·SSNt−1 (95.8%)
+1.5868 × 10−5·SSNt−2 (73.3%)
−2.1539 × 10−5·SSNt−6 (94.1%)
−0.0004·SSNt−6 (98%)
−0.0006·ΝΑΟΙt (94.1%)
−0.0008·ΝΑΟΙt−1 (93.6%)
+0.0004·ΝΑΟΙt−3 (84.2%)
+0.0004·ΝΑΟΙt−4 (83.6%)
+0.0004·ΝΑΟΙt−5 (86.5%)
+0.0004·ΝΑΟΙt−6 (89.9%)
−0.0022·ONI t−1 (64.4%)
+0.0029·ONI t−2 (72.7%)
+4.9699 × 10−6·t (98.9%)
overall regression
RMSE = 0.0042, R2 = 0.6930, p-value = 0.0575 (94.3% CI)
 
+1.3261 × 10−5·SSNt (75.2%)
+2.1955 × 10−5·SSNt−1 (93.7%)
+1.4242 × 10−5·SSNt−2 (71.4%)
−2.2672 × 10−5·SSNt−6 (96.3%)
−0.0004·ΝΑΟΙt (90.4%)
+0.0010·ΝΑΟΙt−1 (80.2%)
+0.0011·ΝΑΟΙt−3 (86%)
−0.0007·ΝΑΟΙt−5 (61.6%)
−7.6620 × 10−5·ONI t−1 (70.4%)
+0.0002·ONI t−2 (97.5%)
overall regression
RMSE = 0.0024, R2 = 0.5601, p-value = 0.3127 (68.7% CI)
 
−0.0054 (94.6%)
+7.7984 × 10−6·SSNt−3 (68.7%)
+1.1722 × 10−6·SSNt−4 (88.7%)
−5.4475 × 10−6·SSNt−6 (60.9%)
+0.0007·ΝΑΟΙt−3 (86.6%)
−4.4589 × 10−5·ONI t−2 (63%)
+3.7864 × 10−5·ONI t−5 (75.7%)
overall regression
RMSE = 0.0031, R2 = 0.5616, p-value = 0.2955 (70.5% CI)
 
+0.0064 (92.9%)
+7.3445 × 10−6·SSNt−1 (60.2%)
+9.5150 × 10−6·SSNt−5 (74.7%)
−8.4786 × 10−6·SSNt−6 (71.1%)
−0.0004·SSNt−6 (98%)
−0.0003·ΝΑΟΙt (89.8%)
+0.0012·ΝΑΟΙt−1 (95.6%)
+0.0006·ΝΑΟΙt−2 (71.1%)
+0.0008·ΝΑΟΙt−3 (82.8%)
−0.0008·ΝΑΟΙt−5 (81.8%)
−0.0005·ΝΑΟΙt−6 (67.3%)
SAOD550overall regression
RMSE = 0.0666, R2 = 0.6551, p-value = 0.0001 (≈100% CI)
 
+0.1726 (96.7%)
−0.0004·SSNt−6 (94.7%)
−0.0066·ΝΑΟΙt (88.1%)
−0.0091·ΝΑΟΙt−1 (95.6%)
+0.0064·ΝΑΟΙt−3 (86%)
+0.0058·ΝΑΟΙt−4 (66.7%)
+0.0046·ΝΑΟΙt−5 (70.4%)
+0.0078·ΝΑΟΙt−6 (64.8%)
−0.0366·ONIt−1 (70.2%)
+0.0424·ONIt−2 (72.1%)
−9.7801 × 10−5·t (99.9%)
overall regression
RMSE = 0.0406, R2 = 0.5699, p-value = 0.2090 (79.1% CI)
 
+0.0001·SSNt−3 (67.1%)
+0.0002·SSNt−4 (94.2%)
−0.0066·ΝΑΟΙt (88.1%)
+0.0136·ΝΑΟΙt−3 (92.8%)
−0.0071·ΝΑΟΙt−4 (65.8%)
−0.0106·ΝΑΟΙt−5 (82.6%)
+0.0008·ONIt−2 (63.6%)
−0.0007·ONIt−4 (85.1%)
overall regression
RMSE = 0.0540, R2 = 0.6094, p-value = 0.0167 (98.3% CI)
 
+0.1173 (94.4%)
−0.0001·SSNt (61.4%)
−0.0002·SSNt−6 (89.7%)
+0.0095·ΝΑΟΙt−3 (65.7%)
−0.0366·ON t−1 (70.2%)
+0.0010·ONI t−2 (61.2%)
−0.0007·ONI t−3 (62.5%)
−0.0008·ONI t−4 (80.3%)
+0.0007·ONI t−4 (74.3%)
−8.9362 × 10−5·t (≈100%)
overall regression
RMSE = 0.0389, R2 = 0.5761, p-value = 0.1547 (84.5% CI)
 
−0.0417 (64.9%)
+0.0003·SSNt−4 (97.2%)
+0.0001·SSNt−5 (83.1%)
+0.0001·SSNt−6 (70.3%)
+0.0100·ΝΑΟΙt−1 (82.6%)
−0.0077·ΝΑΟΙt−4 (71.2%)
−0.0159·ΝΑΟΙt−5 (96.6%)
+0.0148·ΝΑΟΙt−6 (96.2%)
+0.0007·ONIt−1 (66.8%)
−0.0018·ONIt−2 (97.9%)
+0.0015·ONIt−3 (99.3%)
−0.0011·ONIt−4 (98.2%)
+0.0006·ONIt−5 (74.8%)
−2.5458 × 10−5·t (86.6%)
DDAOD550overall regression
RMSE = 0.0501, R2 = 0.6388, p-value = 0.0009 (≈100% CI)
 
+0.0860 (84.3%)
+0.0002·SSNt−1 (91.3%)
+0.0002·SSNt−2 (67.9%)
−0.0002·SSNt−6 (94.5%)
−0.0065·ΝΑΟΙt (95.7%)
−0.0086·ΝΑΟΙt−1 (98.9%)
+0.0054·ΝΑΟΙt−3 (90.3%)
+0.0045·ΝΑΟΙt−4 (81.9%)
+0.0039·ΝΑΟΙt−5 (75.6%)
+0.0038·ΝΑΟΙt−6 (78.1%)
−0.0279·ONI t−1 (70.9%)
+0.0417·ONI t−2 (84.3%)
overall regression
RMSE = 0.0482, R2 = 0.5741, p-value = 0.1713 (82.9% CI)
 
+0.0001·SSNt−1 (68.7%)
−0.0001·SSNt−4 (66.6%)
−0.0003·SSNt−6 (97.6%)
−0.0041·ΝΑΟΙt (83.7%)
+0.0084·ΝΑΟΙt−1 (64.5%)
+0.0144·ΝΑΟΙt−3 (89.3%)
−0.0009·ONI t−1 (71.1%)
+0.0025·ONI t−2 (99%)
−0.0006·ONI t−3 (63.4%)
−0.0008·ONI t−4 (81.6%)
overall regression
RMSE = 0.0516, R2 = 0.4895, p-value = 0.9653 (3.5% CI)
overall regression
RMSE = 0.0298, R2 = 0.4923, p-value = 0.9577 (4.2% CI)
StAOD550overall regression
RMSE = 0.0149, R2 = 0.5941, p-value = 0.0534 (94.7% CI)
 
−6.3900 × 10−5·SSNt−1 (86.7%)
+4.4276 × 10−5·SSNt−3 (64.9%)
+8.9313 × 10−5·SSNt−4 (95%)
+0.0023·ΝΑΟΙt (98.4%)
−0.0040·ΝΑΟΙt−1 (99.9%)
−0.0034·ΝΑΟΙt−5 (99.9%)
+0.0008·ΝΑΟΙt−6 (62.5%)
−0.0094·ONI t−1 (77%)
−0.0080·ONI t−4 (62.7%)
+0.0082·ONI t−5 (67.4%)
overall regression
RMSE = 0.0220, R2 = 0.4865, p-value = 0.9722 (2.8% CI)
overall regression
RMSE = 0.0170, R2 = 0.5279, p-value = 0.7142 (28.6% CI)
overall regression
RMSE = 0.0263, R2 = 0.5944, p-value = 0.0522 (94.8% CI)
 
+0.0002·SSNt−4 (95.9%)
+0.0001·SSNt−5 (95.9%)
+0.0001·SSNt−6 (90.8%)
+0.0106·ΝΑΟΙt−1 (96.6%)
−0.0049·ΝΑΟΙt−4 (68.7%)
−0.0077·ΝΑΟΙt−5 (87%)
+0.0084·ΝΑΟΙt−6 (91.8%)
+0.0005·ONI t−1 (68.5%)
−0.0013·ONI t−2 (98.1%)
+0.0010·ONI t−3 (99.2%)
−0.0006·ONI t−4 (95.3%)
+0.0003·ONI t−5 (67.3%)
SSAOD550overall regression
RMSE = 0.0029, R2 = 0.5269, p-value = 0.7252 (27.5% CI)
overall regression
RMSE = 0.0053, R2 = 0.6028, p-value = 0.0286 (97.1% CI)
 
−0.0094 (87.7%)
+2.3837 × 10−5·SSNt−3 (84%)
−1.5358 × 10−5·SSNt−5 (71.2%)
−1.2409 × 10−5·SSNt−6 (62.8%)
+0.0006·NAOI t (92.8%)
+0.0009·NAOI t−5 (62.1%)
+0.0010·NAOI t−6 (68.3%)
−0.0001·ONIt−2 (78.6%)
+0.0003·ONIt−3 (≈100%)
−0.0001·ONIt−4 (96.7%)
+0.0002·ONIt−5 (98%)
−6.4949 × 10−5·ONIt−6 (74.2%)
overall regression
RMSE = 0.0053, R2 = 0.6084, p-value = 0.0182 (98.2% CI)
 
+0.0182 (99.7%)
+1.3741 × 10−6·SSNt−1 (64.4%)
+1.3953 × 10−5·SSNt−4 (61.8%)
−1.2409 × 10−5·SSNt−6 (62.8%)
+0.0002·ONIt−1 (96.5%)
−0.0001·ONIt−2 (76.1%)
+0.0001·ONIt−3 (94.3%)
+9.5472 × 10−6·t (≈100%)
overall regression
RMSE = 0.0029, R2 = 0.5386, p-value = 0.5862 (41.4% CI)
BCAOD550overall regression
RMSE = 0.0008, R2 = 0.5239, p-value = 0.7564 (24.4% CI)
overall regression
RMSE = 0.0009, R2 = 0.5208, p-value = 0.7876 (21.2% CI)
overall regression
RMSE = 0.0008, R2 = 0.5240, p-value = 0.7560 (24.4% CI)
overall regression
RMSE = 0.0016, R2 = 0.4916, p-value = 0.9599 (4% CI)
AE470–870overall regression
RMSE = 0.1117, R2 = 0.5461, p-value = 0.4890 (51.1% CI)
overall regression
RMSE = 0.1478, R2 = 0.5188, p-value = 0.8056 (19.4% CI)
overall regression
RMSE = 0.1482, R2 = 0.5050, p-value = 0.9050 (9.5% CI)
overall regression
RMSE = 0.1327, R2 = 0.4475, p-value = 0.9993 (0.1% CI)
SSAoverall regression
RMSE = 0.0038, R2 = 0.5187, p-value = 0.8067 (19.3% CI)
overall regression
RMSE = 0.0051, R2 = 0.4881, p-value = 0.9687 (3.1% CI)
overall regression
RMSE = 0.0052, R2 = 0.5233, p-value = 0.7632 (23.7% CI)
overall regression
RMSE = 0.0051, R2 = 0.5758, p-value = 0.1571 (84.3% CI)
 
−0.0087 (86.3%)
+1.4672 × 10−5·SSNt (70.4%)
+2.8446 × 10−5·SSNt−1 (95.1%)
+2.5693 × 10−5·SSNt−3 (88.8%)
+1.7178 × 10−5·SSNt−4 (73.4%)
+1.7562 × 10−5·SSNt−5 (79.7%)
+3.0049 × 10−5·SSNt−6 (97.6%)
+0.0004·NAOIt (81.9%)
+0.0015·NAOI t−2 (89%)
−0.0013·NAOI t−4 (81.7%)
+7.0650 × 10−5·ONIt−5 (70.3%)
DARFoverall regression
RMSE = 2.4927, R2 = 0.6433, p-value = 0.0005 (≈100% CI)
 
+0.0070·SSNt (68.9%)
+0.0136·SSNt−1 (94.2%)
+0.0100·SSNt−2 (78.8%)
−0.0116·SSNt−6 (92.9%)
−0.3498·NAOIt (97.2%)
−0.3853·ΝΑΟΙt−1 (97.8%)
+0.2425·ΝΑΟΙt−3 (86.5%)
+0.3482·ΝΑΟΙt−4 (96.4%)
+0.3333·ΝΑΟΙt−5 (95.6%)
+0.2160·ΝΑΟΙt−6 (84.3%)
+1.3390·ONIt−2 (63.9%)
+0.0026·t (98.3%)
overall regression
RMSE = 2.2501, R2 = 0.5875, p-value = 0.0819 (91.8% CI)
 
−0.2267 (87.7%)
+0.0075·SSNt (77.3%)
+0.0101·SSNt−1 (88.6%)
+0.0068·SSNt−2 (65.2%)
−0.0125·SSNt−6 (96.5%)
−0.2632·NAOIt (94.3%)
+0.5943·ΝΑΟΙt−1 (83.7%)
+0.7397·ΝΑΟΙt−3 (92.4%)
+0.3510·ΝΑΟΙt−4 (60%)
+0.0714·ONI t−2 (87.8%)
+0.0018·t (93.6%)
overall regression
RMSE = 1.8977, R2 = 0.5922, p-value = 0.0604 (94% CI)
 
+0.0092·SSNt (92%)
+0.0103·SSNt−1 (94.5%)
+0.0080·SSNt−2 (81%)
+0.0056·SSNt−3 (64.3%)
+0.0053·SSNt−5 (69.2%)
−0.0065·SSNt−6 (81%)
−0.2250·NAOIt (94.6%)
+0.5957·ΝΑΟΙt−1 (90.2%)
+0.4686·ΝΑΟΙt−3 (81.7%)
+0.3951·ΝΑΟΙt−6 (74.5%)
+0.0705·ONIt−2 (92.9%)
−0.0258·ONIt−3 (66.9%)
+0.0023·t (99.5%)
overall regression
RMSE = 1.8606, R2 = 0.5822, p-value = 0.1116 (88.8% CI)
 
+0.0056·SSNt (72%)
+0.0063·SSNt−1 (77%)
+0.0041·SSNt−2 (69.8%)
+0.0062·SSNt−5 (77.7%)
−0.0069·SSNt−6 (84.7%)
−0.2535·NAOIt (97.3%)
+0.5968·ΝΑΟΙt−1 (90.9%)
+0.3334·ΝΑΟΙt−2 (67.5%)
+0.4841·ΝΑΟΙt−3 (84%)
−0.4100·ΝΑΟΙt−5 (74.7%)
−0.3822·ΝΑΟΙt−6 (73.9%)
netSWRBOA,CS,Aoverall regression
RMSE = 60.8066 W−2, R2 = 0.6209, p-value = 0.0060 (99.4% CI)
+130.6428 (92.3%)
 
+0.2024·SSNt (76.9%)
+0.2629·SSNt−1 (86.9%)
+0.2234·SSNt−2 (74.6%)
−0.1976·SSNt−6 (79.5%)
−10.3043·NAOIt (99.2%)
−4.8354·ΝΑΟΙt−1 (76.2%)
+6.7794·ΝΑΟΙt−3 (91.3%)
+10.2190·ΝΑΟΙt−4 (98.8%)
+8.8681·ΝΑΟΙt−5 (97.2%)
overall regression
RMSE = 67.8601 W−2, R2 = 0.5954, p-value = 0.0488 (95.1% CI)
 
+98.0516 (79.1%)
+0.2576·SSNt (83%)
+0.2743·SSNt−1 (84.5%)
−0.2643·SSNt−6 (86.3%)
−12.0709·NAOIt (99.6%)
+19.5626·ΝΑΟΙt−3 (88%)
+10.2190·ΝΑΟΙt−4 (98.8%)
−0.9205·ONIt−3 (66.8%)
overall regression
RMSE = 65.3489 W−2, R2 = 0.5971, p-value = 0.0435 (95.7% CI)
 
+107.7180 (84.8%)
+0.2525·SSNt (83.9%)
+0.2743·SSNt−1 (85.2%)
−0.2606·SSNt−6 (87.2%)
−11.5600·NAOIt (99.6%)
+18.9829·ΝΑΟΙt−3 (88.3%)
+10.2190·ΝΑΟΙt−4 (98.8%)
−0.8794·ONIt−3 (66.4%)
overall regression
RMSE = 78.0674 W−2, R2 = 0.5655, p-value = 0.0485 (95.2% CI)
 
+0.3042·SSNt (84.1%)
+0.3203·SSNt−1 (85.1%)
+0.2123·SSNt−2 (60.2%)
−0.3213·SSNt−6 (88.4%)
−14.0072·NAOIt (99.6%)
+21.7551·ΝΑΟΙt−3 (86.7%)
−1.0693·ONIt−3 (67.2%)
+0.0299·t (61.9%)
netLWRBOA,CS,Aoverall regression
RMSE = 5.9091 W−2, R2 = 0.5927, p-value = 0.0586 (94.1% CI)
 
−96.7403 (99.9%)
−0.0168·SSNt (69.3%)
+0.0253·SSNt−1 (86.5%)
−0.0234·SSNt−2 (78%)
−0.0250·SSNt−3 (81.5%)
−0.0161·SSNt−5 (68%)
+0.0150·SSNt−6 (67.8%)
+0.5868·NAOIt (88%)
+0.4046·ΝΑΟΙt−2 (70%)
−0.9142·ΝΑΟΙt−4 (98%)
−1.0876·ΝΑΟΙt−5 (99.4%)
+8.8681·ΝΑΟΙt−5 (97.2%)
−3.1520·ONIt−3 (61.9%)
+3.5934·ONIt−4 (68.5%)
−3.4192·ONIt−5 (69.4%)
+1.9677·ONIt−6 (70.6%)
overall regression
RMSE = 4.3512 W−2, R2 = 0.5972, p-value = 0.1321 (86.8% CI)
 
−103.6478 (≈100%)
+0.0124·SSNt (69.7%)
+0.0147·SSNt−1 (76.6%)
−0.0161·SSNt−4 (77.7%)
−0.0104·SSNt−6 (64.1%)
−0.9157·NAOIt (99.9%)
−0.9835·ΝΑΟΙt−1 (76.8%)
−1.3072·ΝΑΟΙt−2 (90.1%)
−1.0363·ΝΑΟΙt−5 (78.3%)
−0.1569·ONIt−1 (95.8%)
+0.1914·ONIt−2 (96.7%)
−0.1014·ONIt−5 (91.8%)
+0.1151·ONIt−6 (98.5%)
+0.0027·t (84.2%)
overall regression
RMSE = 3.3649 W−2, R2 = 0.5811, p-value = 0.1191 (88.1% CI)
 
−107.7368 (≈100%)
+0.0091·SSNt−1 (65.8%)
+0.0093·SSNt−2 (61.1%)
−0.5049·NAOIt (98.5%)
−0.9429·ΝΑΟΙt−2 (87.6%)
−1.3072·ΝΑΟΙt−2 (90.1%)
−0.0961·ONIt−1 (89.4%)
+0.1211·ONIt−2 (92%)
+0.0550·ONIt−3 (75.7%)
−0.0732·ONIt−4 (93%)
+0.0666·ONIt−6 (93.3%)
+0.0058·t (≈100%)
overall regression
RMSE = 4.4422 W−2, R2 = 0.5679, p-value = 0.2283 (77.2% CI)
 
−95.0348 (≈100%)
−0.0167·SSNt (82.5%)
−1.0330·ΝΑΟΙt−2 (79.8%)
+0.8875·ΝΑΟΙt−3 (71.9%)
+0.7096·ΝΑΟΙt−6 (61.8%)
+0.1469·ONIt−2 (89.3%)
−0.0670·ONIt−4 (79.2%)
+0.0528·ONIt−6 (72.9%)
+0.0085·t (≈100%)
CODoverall regression
RMSE = 5.8037, R2 = 0.5935, p-value = 0.0554 (94.5% CI)
 
+15.6029 (97.3%)
−0.0170·SSNt−3 (64.2%)
+0.5651·ΝΑΟΙt−1 (85.1%)
−0.8495·ΝΑΟΙt−4 (97.2%)
−0.8446·ΝΑΟΙt−5 (97.2%)
−1.3957·ΝΑΟΙt−6 (≈100%)
−3.0681·ONIt−4 (61.8%)
overall regression
RMSE = 8.9935, R2 = 0.6002, p-value = 0.0346 (96.5% CI)
 
+25.36029 (98.5%)
−0.0388·SSNt (88%)
−0.0307·SSNt−1 (77%)
−0.0304·SSNt−2 (71%)
+0.0267·SSNt−6 (74.4%)
+2.1672·ΝΑΟΙt (≈100%)
−1.7244·ΝΑΟΙt−1 (68.9%)
−2.3902·ΝΑΟΙt−3 (84.8%)
−0.1864·ΝΑΟΙt−4 (84.4%)
+1.6426·ΝΑΟΙt−5 (65.7%)
+2.9777·ΝΑΟΙt−6 (92.9%)
+0.1864·ONIt−3 (86.1%)
−0.1703·ONIt−4 (88.6%)
+0.1215·ONIt−5 (68.8%)
overall regression
RMSE = 10.0918, R2 = 0.6140, p-value = 0.0113 (98.9% CI)
 
+20.85789 (92.7%)
−0.0457·SSNt (89.8%)
−0.0286·SSNt−1 (68.1%)
+0.0351·SSNt−4 (74.8%)
+0.0512·SSNt−6 (94.7%)
+3.0228·ΝΑΟΙt (≈100%)
−2.8010·ΝΑΟΙt−3 (86.6%)
−3.5270·ΝΑΟΙt−4 (94%)
+2.4207·ΝΑΟΙt−6 (81%)
−0.1072·ONIt−4 (74.4%)
overall regression
RMSE = 6.6790, R2 = 0.5492, p-value = 0.4486 (55.1% CI)

Appendix A.7. Summary of Correlations

This section provides the Pearson’s correlation coefficient, r, between any parameter, q, and one of the three physical phenomena (SSN, NAOI, or ONI). Values of |r|≥0.2 have been chosen from Table A3 to be included here. This was performed in order to promote those subregions and parameters that have an enhanced relationship with the physical phenomena.
Table A8. Summary of the Pearson’s correlation coefficient, r, between parameter q and one of the three physical phenomena (SSN, NAOI, ONI) over the four Mediterranean domains in the period 1980–2024. Annual mean values have been used. The asterisks denote the significant level of the relationship (* for 95% CI, *** for 99.9% CI). The absence of an r value in a parameter pair implies insignificant level, i.e., a p-value > 0.05 (< 95% CI). DARF has units of Wm−2.
Table A8. Summary of the Pearson’s correlation coefficient, r, between parameter q and one of the three physical phenomena (SSN, NAOI, ONI) over the four Mediterranean domains in the period 1980–2024. Annual mean values have been used. The asterisks denote the significant level of the relationship (* for 95% CI, *** for 99.9% CI). The absence of an r value in a parameter pair implies insignificant level, i.e., a p-value > 0.05 (< 95% CI). DARF has units of Wm−2.
Parameters in PairWestMedCentMedEastMedBalBSea
TAOD550–SSN +0.206 *** +0.368 *
SAOD550–SSN +0.215 ***+0.299 *+0.367 *
StAOD550–SSN+0.260 ***+0.236 ***+0.323 *+0.353 *
SSAOD550–SSN −0.306 *−0.401 *
SSA–SSN+0.246 ***+0.239 ***+0.306 *
StAOD550–NAOI +0.295 *
AE470–870–NAOI +0.324 *+0.314 *
SSA–NAOI+0.246 ***+0.239 ***+0.306 *
DARF–NAOI −0.294 *
COD–NAOI +0.235 *** −0.316 *
AAOD550–ONI −0.306 *
DDAOD550–ONI −0.284 *−0.333 *
StAOD550–ONI+0.225 *** +0.295 *
AE470–870–ONI+0.219 ***+0.233 ***+0.331 *
DARF–ONI −0.334 *

References

  1. Iqbal, M. An Introduction to Solar Radiation; Academic Press: Cambridge, MA, USA, 1983; p. 408. ISBN 978-0-323-15181-8. [Google Scholar]
  2. Donateo, A.; Feudo, T.L.; Marinoni, A.; Dinoi, A.; Avolio, E.; Merico, E.; Calidonna, C.R.; Contini, D.; Bonasoni, P. Characterization of in Situ Aerosol Optical Properties at Three Observatories in the Central Mediterranean. Atmosphere 2018, 9, 369. [Google Scholar] [CrossRef]
  3. Myhre, G.; Myhre, C.E.L.; Samset, B.H.; Storelvmo, T. Aerosols and Their Relation to Global Climate and Climate Sensitivity. Nat. Educ. 2015, 4, 7. [Google Scholar]
  4. Kaskaoutis, D.G.; Kambezidis, H.D.; Adamopoulos, A.D.; Kassomenos, P.A. On the Characterization of Aerosols Using the Ångström Exponent in the Athens Area. J. Atmos. Sol.-Terr. Phys. 2006, 68, 2147–2163. [Google Scholar] [CrossRef]
  5. Chen, Y.; Li, Q.; Kahn, R.A.; Randerson, J.T.; Diner, D.J. Quantifying Aerosol Direct Radiative Effect with Multiangle Imaging Spectroradiometer Observations: Top-of-Atmosphere Albedo Change by Aerosols Based on Land Surface Types. J. Geophys. Res. Atmos. 2009, 114, D02109. [Google Scholar] [CrossRef]
  6. Yin, Y.; Wurzler, S.; Levin, Z.; Reisin, T.G. Interactions of Mineral Dust Particles and Clouds: Effects on Precipitation and Cloud Optical Properties. J. Geophys. Res. Atmos. 2002, 107, AAC 19-1–AAC 19-14. [Google Scholar] [CrossRef]
  7. Wurzler, S.; Reisin, T.G.; Levin, Z. Modification of Mineral Dust Particles by Cloud Processing and Subsequent Effects on Drop Size Distributions. J. Geophys. Res. Atmos. 2000, 105, 4501–4512. [Google Scholar] [CrossRef]
  8. Metangley, S.; Middey, A.; Kadaverugu, R. Modern Methods to Explore the Dynamics between Aerosols and Convective Precipitation: A Critical Review. Dyn. Atmos. Ocean. 2024, 106, 101465. [Google Scholar] [CrossRef]
  9. Li, J.; Carlson, B.E.; Yung, Y.L.; Lv, D.; Hansen, J.; Penner, J.E.; Liao, H.; Ramaswamy, V.; Kahn, R.A.; Zhang, P.; et al. Scattering and Absorbing Aerosols in the Climate System. Nat. Rev. Earth Environ. 2022, 3, 363–379. [Google Scholar] [CrossRef]
  10. Di Iorio, T.; di Sarra, A.; Junkermann, W.; Cacciani, M.; Fiocco, G.; Fuà, D. Tropospheric Aerosols in the Mediterranean: 1. Microphysical and Optical Properties. J. Geophys. Res. Atmos. 2003, 108, 4316. [Google Scholar] [CrossRef]
  11. Martin, R.V.; Jacob, D.J.; Yantosca, R.M.; Chin, M.; Ginoux, P. Global and Regional Decreases in Tropospheric Oxidants from Photochemical Effects of Aerosols. J. Geophys. Res. Atmos. 2003, 108, 4097–4105. [Google Scholar] [CrossRef]
  12. Kulmala, M.; Arola, A.; Nieminen, T.; Riuttanen, L.; Sogacheva, L.; de Leeuw, G.; Kerminen, V.-M.; Lehtinen, K.E.J. The First Estimates of Global Nucleation Mode Aerosol Concentrations Based on Satellite Measurements. Atmos. Chem. Phys. 2011, 11, 10791–10801. [Google Scholar] [CrossRef]
  13. Sinha, P.R.; Kaskaoutis, D.G.; Manchanda, R.K.; Sreenivasan, S. Characteristics of Aerosols over Hyderabad in Southern Peninsular India: Synergy in the Classification Techniques. Ann. Geophys. 2012, 30, 1393–1410. [Google Scholar] [CrossRef]
  14. Choobari, O.A.; Zawar-Reza, P.; Sturman, A. The Global Distribution of Mineral Dust and Its Impacts on the Climate System: A Review. Atmos. Res. 2014, 138, 152–165. [Google Scholar] [CrossRef]
  15. Kambezidis, H.D. The Solar Resource. In Comprehensive Renewable Energy; Letcher, T.M., Ed.; Elsevier: Oxford, UK, 2022; Volume 3, pp. 26–117. ISBN 978-0-12-819727-1. [Google Scholar]
  16. Yoo, J.M.; Won, Y.I.; Jeong, M.J.; Kim, K.M.; Shin, D.B.; Lee, Y.R.; Cho, Y.J. Intensity of Climate Variability Derived from the Satellite and MERRA Reanalysis Temperatures: AO, ENSO, and QBO. J. Atmos. Sol.-Terr. Phys. 2013, 95–96, 15–27. [Google Scholar] [CrossRef]
  17. Müller, W.A.; Roeckner, E. ENSO Impact on Midlatitude Circulation Patterns in Future Climate Change Projections. Geophys. Res. Lett. 2006, 33, L05711. [Google Scholar] [CrossRef]
  18. Abish, B.; Mohanakumar, K. Absorbing Aerosol Variability over the Indian Subcontinent and Its Increasing Dependence on ENSO. Glob. Planet. Change 2013, 106, 13–19. [Google Scholar] [CrossRef]
  19. Laken, B.A.; Stordal, F. Are There Statistical Links between the Direction of European Weather Systems and ENSO, the Solar Cycle or Stratospheric Aerosols? R. Soc. Open Sci. 2016, 3, 150320. [Google Scholar] [CrossRef]
  20. Kirov, B.; Georgieva, K. Long-Term Variations and Interrelations of ENSO, NAO and Solar Activity. Phys. Chem. Earth 2002, 27, 441–448. [Google Scholar] [CrossRef]
  21. Mezzina, B.; García-Serrano, J.; Bladé, I.; Kucharski, F. Dynamics of the ENSO Teleconnection and NAO Variability in the North Atlantic-European Late Winter. J. Clim. 2020, 33, 907–923. [Google Scholar] [CrossRef]
  22. Casselman, J.W.; Jiménez-Esteve, B.; Domeisen, D.I.V. Modulation of the El Niño Teleconnection to the North Atlantic by the Tropical North Atlantic During Boreal Spring and Summer. Weather Clim. Dyn. 2022, 3, 1077–1096. [Google Scholar] [CrossRef]
  23. Casselman, J.W.; Taschetto, A.S.; Domeisen, D.I.V. Nonlinearity in the Pathway of El Ninõ-Southern Oscillation to the Tropical North Atlantic. J. Clim. 2021, 34, 7277–7296. [Google Scholar] [CrossRef]
  24. Casselman, J.W.; Lübbecke, J.F.; Bayr, T.; Huo, W.; Wahl, S.; Domeisen, D.I.V. The Teleconnection of Extreme El Niño-Southern Oscillation (ENSO) Events to the Tropical North Atlantic in Coupled Climate Models. Weather Clim. Dyn. 2023, 4, 471–487. [Google Scholar] [CrossRef]
  25. García-Serrano, J.; Cassou, C.; Douville, H.; Giannini, A.; Doblas-Reyes, F.J. Revisiting the ENSO Teleconnection to the Tropical North Atlantic. J. Clim. 2017, 30, 6945–6957. [Google Scholar] [CrossRef]
  26. Liu, Y.; Liu, J.; Xie, M.; Fang, K.; Tarasick, D.W.; Wang, H.; Meng, L.; Cheng, X.; Han, H.; Zhang, X. ENSO Teleconnection to Interannual Variability in Carbon Monoxide Over the North Atlantic European Region in Spring. Front. Environ. Sci. 2022, 10, 894779. [Google Scholar] [CrossRef]
  27. Alpert, P.; Baldi, M.; Ilani, R.; Krichak, S.; Price, C.; Rodó, X.; Saaroni, H.; Ziv, B.; Kishcha, P.; Barkan, J.; et al. Chapter 2 Relations Between Climate Variability in the Mediterranean Region and the Tropics: ENSO, South Asian and African Monsoons, Hurricanes and Saharan Dust. In Developments in Earth and Environmental Sciences; Lionello, P., Malanotte-Rizzoli, P., Boscolo, R., Eds.; Mediterranean; Elsevier: Amsterdam, The Netherlands, 2006; Volume 4, pp. 149–177. [Google Scholar]
  28. Urdiales-Flores, D.; Zittis, G.; Hadjinicolaou, P.; Osipov, S.; Klingmüller, K.; Mihalopoulos, N.; Kanakidou, M.; Economou, T.; Lelieveld, J. Drivers of Accelerated Warming in Mediterranean Climate-Type Regions. npj Clim. Atmos. Sci. 2023, 6, 97. [Google Scholar] [CrossRef]
  29. Park, S. Remote ENSO Influence on Mediterranean Sky Conditions During Late Summer and Autumn: Evidence for a Slowly Evolving Atmospheric Bridge. Q. J. R. Meteorol. Soc. 2004, 130, 2409–2422. [Google Scholar] [CrossRef]
  30. Alsubhi, Y.; Ali, G. Impact of El Niño-Southern Oscillation on Dust Variability during the Spring Season over the Arabian Peninsula. Atmosphere 2024, 15, 1060. [Google Scholar] [CrossRef]
  31. Kambezidis, H.D. Atmospheric Processes over the Broader Mediterranean Region: Effect of the El Niño–Southern Oscillation? Atmosphere 2024, 15, 268. [Google Scholar] [CrossRef]
  32. Westervelt, D.M.; Conley, A.J.; Fiore, A.M.; Lamarque, J.F.; Shindell, D.T.; Previdi, M.; Mascioli, N.R.; Faluvegi, G.; Correa, G.; Horowitz, L.W. Connecting Regional Aerosol Emissions Reductions to Local and Remote Precipitation Responses. Atmos. Chem. Phys. 2018, 18, 12461–12475. [Google Scholar] [CrossRef]
  33. Huang, Y.; Liu, X.; Yin, Z.Y.; An, Z. Global Impact of ENSO on Dust Activities with Emphasis on the Key Region from the Arabian Peninsula to Central Asia. J. Geophys. Res. Atmos. 2021, 126, e2020JD034068. [Google Scholar] [CrossRef]
  34. Banerjee, P.; Kumar, S.P. ENSO Modulation of Interannual Variability of Dust Aerosols over the Northwest Indian Ocean. J. Clim. 2016, 29, 1287–1303. [Google Scholar] [CrossRef]
  35. Brönnimann, S.; Luterbacher, J.; Staehelin, J.; Svendby, T.M.; Hansen, G.; Svenøe, T. Extreme Climate of the Global Troposphere and Stratosphere in 1940–42 Related to El Niño. Nature 2004, 431, 971–974. [Google Scholar] [CrossRef]
  36. Toniazzo, T.; Scaife, A.A. The Influence of ENSO on Winter North Atlantic Climate. Geophys. Res. Lett. 2006, 33, L24704. [Google Scholar] [CrossRef]
  37. Leamon, R.J. The Triple-Dip La Niña of 2020–22: Updates to the Correlation of ENSO with the Termination of Solar Cycles. Front. Earth Sci. 2023, 11, 1204191. [Google Scholar] [CrossRef]
  38. Huo, W.; Xiao, Z.; Zhao, L. Modulation of the Solar Activity on the Connection between the NAO and the Tropical Pacific SST Variability. Front. Earth Sci. 2023, 11, 1147582. [Google Scholar] [CrossRef]
  39. Im, U.; Kanakidou, M. Impacts of East Mediterranean Megacity Emissions on Air Quality. Atmos. Chem. Phys. 2012, 12, 6335–6355. [Google Scholar] [CrossRef]
  40. Stohl, A.; Aamaas, B.; Amann, M.; Baker, L.H.; Bellouin, N.; Berntsen, T.K.; Boucher, O.; Cherian, R.; Collins, W.; Daskalakis, N.; et al. Evaluating the Climate and Air Quality Impacts of Short-Lived Pollutants. Atmos. Chem. Phys. 2015, 15, 10529–10566. [Google Scholar] [CrossRef]
  41. Abdelkader, M.; Metzger, S.; Mamouri, R.E.; Astitha, M.; Barrie, L.; Levin, Z.; Lelieveld, J. Dust-Air Pollution Dynamics over the Eastern Mediterranean. Atmos. Chem. Phys. 2015, 15, 9173–9189. [Google Scholar] [CrossRef]
  42. Kaskaoutis, D.G.; Rashki, A.; Dumka, U.C.; Mofidi, A.; Kambezidis, H.D.; Psiloglou, B.E.; Karagiannis, D.; Petrinoli, K.; Gavriil, A. Atmospheric Dynamics Associated with Exceptionally Dusty Conditions over the Eastern Mediterranean and Greece in March 2018. Atmos. Res. 2019, 218, 269–284. [Google Scholar] [CrossRef]
  43. Amiridis, V.; Wandinger, U.; Marinou, E.; Giannakaki, E.; Tsekeri, A.; Basart, S.; Kazadzis, S.; Gkikas, A.; Tayler, M.; Baldasano, J.; et al. Optimizing Saharan Dust CALIPSO Retrievals. Atmos. Chem. Phys. Discuss. 2013, 13, 14749–14795. [Google Scholar] [CrossRef]
  44. Kaskaoutis, D.G.; Kambezidis, H.D.; Nastos, P.T.; Kosmopoulos, P.G. Study on an Intense Dust Storm over Greece. Atmos. Environ. 2008, 42, 6884–6896. [Google Scholar] [CrossRef]
  45. Rizza, U.; Barnaba, F.; Marcello Miglietta, M.; Mangia, C.; Di Liberto, L.; Dionisi, D.; Costabile, F.; Grasso, F.; Paolo Gobbi, G. WRF-Chem Model Simulations of a Dust Outbreak over the Central Mediterranean and Comparison with Multi-Sensor Desert Dust Observations. Atmos. Chem. Phys. 2017, 17, 93–115. [Google Scholar] [CrossRef]
  46. Sicard, M.; Córdoba-Jabonero, C.; Barreto, A.; Welton, E.J.; Gil-Díaz, C.; Carvajal-Pérez, C.V.; Comerón, A.; García, O.; García, R.; López-Cayuela, M.Á.; et al. Volcanic Eruption of Cumbre Vieja, La Palma, Spain: A First Insight to the Particulate Matter Injected in the Troposphere. Remote Sens. 2022, 14, 2470. [Google Scholar] [CrossRef]
  47. Robock, A.; Matson, M. Circumglobal Transport of the El Chichón Volcanic Dust Cloud. Science 1983, 221, 195–197. [Google Scholar] [CrossRef]
  48. Liu, Y.; Kahn, R.A.; Chaloulakou, A.; Koutrakis, P. Analysis of the Impact of the Forest Fires in August 2007 on Air Quality of Athens Using Multi-Sensor Aerosol Remote Sensing Data, Meteorology and Surface Observations. Atmos. Environ. 2009, 43, 3310–3318. [Google Scholar] [CrossRef]
  49. Mishra, A.K.; Klingmueller, K.; Fredj, E.; Lelieveld, J.; Rudich, Y.; Koren, I. Radiative Signature of Absorbing Aerosol over the Eastern Mediterranean Basin. Atmos. Chem. Phys. 2014, 14, 7213–7231. [Google Scholar] [CrossRef]
  50. King, M.P.; Keenlyside, N.; Li, C. ENSO Teleconnections in Terms of Non-NAO and NAO Atmospheric Variability. Clim. Dyn. 2023, 61, 2717–2733. [Google Scholar] [CrossRef]
  51. Khadgarai, S.; Kumar, V.; Pradhan, P.K. The Connection between Extreme Precipitation Variability over Monsoon Asia and Large-Scale Circulation Patterns. Atmos. 2021, 12, 1492. [Google Scholar] [CrossRef]
  52. Yadav, M. South Asian Monsoon Extremes and Climate Change. In Extremes in Atmospheric Processes and Phenomenon: Assessment, Impacts and Mitigation, Disaster Resilience and Green Growth; Saxena, P., Shukla, A., Gupta, A.K., Eds.; Springer Nature: Singapore, 2022; pp. 59–86. ISBN 978-981-16-7727-4. [Google Scholar]
  53. Kaskaoutis, D.G.; Kalapureddy, M.C.R.; Devara, P.C.S.; Kosmopoulos, P.G.; Nastos, P.T.; Krishna Moorthy, K.; Kambezidis, H.D. Spatio-Temporal Aerosol Optical Characteristics over the Arabian Sea during the PreMonsoon Season. Atmos. Chem. Phys. Discuss. 2009, 9, 22223–22269. [Google Scholar] [CrossRef]
  54. Hochet, A.; Dodet, G.; Ardhuin, F.; Hemer, M.; Young, I. Sea State Decadal Variability in the North Atlantic: A Review. Climate 2021, 9, 173. [Google Scholar] [CrossRef]
  55. Vencloviene, J.; Kiznys, D.; Zaltauskaite, J. Statistical Associations Between Geomagnetic Activity, Solar Wind, Solar Proton Events, and Winter NAO and AO Indices. Earth Space Sci. 2022, 9, e2021EA002179. [Google Scholar] [CrossRef]
  56. Rojas, M.; Li, L.Z.; Kanakidou, M.; Hatzianastassiou, N.; Seze, G.; Le Treut, H. Winter Weather Regimes over the Mediterranean Region: Their Role for the Regional Climate and Projected Changes in the Twenty-First Century. Clim. Dyn. 2013, 41, 551–571. [Google Scholar] [CrossRef]
  57. Bitsa, E.; Flocas, H.A.; Hatzaki, M.; Kouroutzoglou, J.; Rudeva, I.; Simmonds, I. Climatological Study of Frontal Precipitation over the Mediterranean. In Proceedings of the EGU General Assembly 2022, Vienna, Austria, 23–27 May 2022. [Google Scholar] [CrossRef]
  58. Drugé, T.; Nabat, P.; Mallet, M.; Somot, S. Future Evolution of Aerosols and Implications for Climate Change in the Euro-Mediterranean Region Using the CNRM-ALADIN63 Regional Climate Model. Atmos. Chem. Phys. 2021, 21, 7639–7669. [Google Scholar] [CrossRef]
  59. Acker, J.G.; Leptoukh, G. Online Analysis Enhances Use of NASA Earth Science Data. Eos Trans. Am. Geophys. Union. 2007, 88, 14–17. [Google Scholar] [CrossRef]
  60. Shaheen, A.; Wu, R.; Aldabash, M. Long-Term AOD Trend Assessment over the Eastern Mediterranean Region: A Comparative Study Including a New Merged Aerosol Product. Atmos. Environ. 2020, 238, 117736. [Google Scholar] [CrossRef]
  61. Su, X.; Huang, G.; Wang, L.; Wei, Y.; Ma, X.; Wang, L.; Feng, L. Validation and Comparison of Long-Term Accuracy and Stability of Global Reanalysis and Satellite Retrieval AOD. Remote Sens. 2024, 16, 3304. [Google Scholar] [CrossRef]
  62. Mukkavilli, S.K.; Prasad, A.A.; Taylor, R.A.; Huang, J.; Mitchell, R.M.; Troccoli, A.; Kay, M.J. Assessment of Atmospheric Aerosols from Two Reanalysis Products over Australia. Atmos. Res. 2019, 215, 149–164. [Google Scholar] [CrossRef]
  63. Khatibi, A.; Krauter, S. Validation and Performance of Satellite Meteorological Dataset MERRA-2 for Solar and Wind Applications. Energies 2021, 14, 882. [Google Scholar] [CrossRef]
  64. Chiu, J.C.; Huang, C.H.; Marshak, A.; Slutsker, I.; Giles, D.M.; Holben, B.N.; Knyazikhin, Y.; Wiscombe, W.J. Cloud Optical Depth Retrievals from the Aerosol Robotic Network (AERONET) Cloud Mode Observations. J. Geophys. Res. Atmos. 2010, 115, D14202. [Google Scholar] [CrossRef]
  65. Min, Q.; Joseph, E.; Duan, M. Retrievals of Thin Cloud Optical Depth from a Multifilter Rotating Shadowband Radiometer. J. Geophys. Res. Atmos. 2004, 109, D02201. [Google Scholar] [CrossRef]
  66. Eldering, A.; Kulawik, S.S.; Worden, J.; Bowman, K.; Osterman, G. Implementation of Cloud Retrievals for TES Atmospheric Retrievals: 2. Characterization of Cloud Top Pressure and Effective Optical Depth Retrievals. J. Geophys. Res. Atmos. 2008, 113, D16S37. [Google Scholar] [CrossRef]
  67. Li, Z. On the Solar Radiation Budget and the Cloud Absorption Anomaly Debate. In World Scientific Series on Asia-Pacific Weather and Climate; World Scientific: Singapore, 2004; Volume 3, pp. 437–456. ISBN 978-981-238-704-2. [Google Scholar] [CrossRef]
  68. Huang, B.; Thorne, P.W.; Banzon, V.F.; Boyer, T.; Chepurin, G.; Lawrimore, J.H.; Menne, M.J.; Smith, T.M.; Vose, R.S.; Zhang, H.-M. Extended Reconstructed Sea Surface Temperature, Version 5 (ERSSTv5): Upgrades, Validations, and Intercomparisons. J. Clim. 2017, 30, 8179–8205. [Google Scholar] [CrossRef]
  69. Van Den Dool, H.M.; Saha, S.; Johansson, Å. Empirical Orthogonal Teleconnections. J. Clim. 2000, 13, 1421–1435. [Google Scholar] [CrossRef]
  70. Chen, W.Y.; Van den Dool, H. Sensitivity of Teleconnection Patterns to the Sign of Their Primary Action Center. Mon. Weather Rev. 2003, 131, 2885–2899. [Google Scholar] [CrossRef]
  71. Barnston, A.G.; Livezey, R.E. Classification, Seasonality and Persistence of Low-Frequency Atmospheric Circulation Patterns. Mon. Weather Rev. 1987, 115, 1083–1126. [Google Scholar] [CrossRef]
  72. Newhall, C.G.; Self, S. The Volcanic Explosivity Index (VEI): An Estimate of Explosive Magnitude for Historical Volcanism. J. Geophys. Res. Oceans 1982, 87, 1231–1238. [Google Scholar] [CrossRef]
  73. Sannino, A.; Amoruso, S.; Damiano, R.; Scollo, S.; Sellitto, P.; Boselli, A. Optical and Microphysical Characterization of Atmospheric Aerosol in the Central Mediterranean During Simultaneous Volcanic Ash and Desert Dust Transport Events. Atmos. Res. 2022, 271, 106099. [Google Scholar] [CrossRef]
  74. Moretti, S.; Salmatonidis, A.; Querol, X.; Tassone, A.; Andreoli, V.; Bencardino, M.; Pirrone, N.; Sprovieri, F.; Naccarato, A. Contribution of Volcanic and Fumarolic Emission to the Aerosol in Marine Atmosphere in the Central Mediterranean Sea: Results from Med-Oceanor 2017 Cruise Campaign. Atmosphere 2020, 11, 149. [Google Scholar] [CrossRef]
  75. Sanchez-Lorenzo, A.; Enriquez-Alonso, A.; Calbó, J.; González, J.-A.; Wild, M.; Folini, D.; Norris, J.R.; Vicente-Serrano, S.M. Fewer Clouds in the Mediterranean: Consistency of Observations and Climate Simulations. Sci. Rep. 2017, 7, 41475. [Google Scholar] [CrossRef]
  76. Sigl, M.; Toohey, M.; McConnell, J.R.; Cole-Dai, J.; Severi, M. Volcanic Stratospheric Sulfur Injections and Aerosol Optical Depth During the Holocene (Past 11 500 Years) from a Bipolar Ice-Core Array. Earth Syst. Sci. Data 2022, 14, 3167–3196. [Google Scholar] [CrossRef]
  77. Kallos, G.; Solomos, S.; Kushta, J.; Mitsakou, C.; Spyrou, C.; Bartsotas, N.; Kalogeri, C. Natural and Anthropogenic Aerosols in the Eastern Mediterranean and Middle East: Possible Impacts. Sci. Total Environ. 2014, 488–489, 389–397. [Google Scholar] [CrossRef]
  78. Dai, Y.; Hitchcock, P.; Mahowald, N.M.; Domeisen, D.I.V.; Hamilton, D.S.; Li, L.; Marticorena, B.; Kanakidou, M.; Mihalopoulos, N.; Aboagye-Okyere, A. Stratospheric Impacts on Dust Transport and Air Pollution in West Africa and the Eastern Mediterranean. Nat. Commun. 2022, 13, 7744. [Google Scholar] [CrossRef]
  79. Hochman, A.; Marra, F.; Messori, G.; Pinto, J.G.; Raveh-Rubin, S.; Yosef, Y.; Zittis, G. Extreme Weather and Societal Impacts in the Eastern Mediterranean. Earth Syst. Dyn. 2022, 13, 749–777. [Google Scholar] [CrossRef]
  80. Sato, M.; Hansen, J.E.; McCormick, M.P.; Pollack, J.B. Stratospheric Aerosol Optical Depths, 1850–1990. J. Geophys. Res. Atmos. 1993, 98, 22987–22994. [Google Scholar] [CrossRef]
  81. Serykh, I.V. El Niño–Southern Oscillation Prediction Based on the Global Atmospheric Oscillation in CMIP6 Models. Climate 2025, 13, 25. [Google Scholar] [CrossRef]
  82. Marinou, E.; Amiridis, V.; Binietoglou, I.; Solomos, S.; Proestakis, E.; Konsta, D.; Tsikerdekis, A.; Papagiannopoulos, N.; Vlastou, G.; Zanis, P.; et al. 3D Evolution of Saharan Dust Transport towards Europe Based on a 9-Year EARLINET-Optimized CALIPSO Dataset. Atmos. Chem. Phys. Discuss. 2016, 17, 5893–5919. [Google Scholar] [CrossRef]
  83. Kambezidis, H.D. The Solar Radiation Climate of Greece. Climate 2021, 9, 183. [Google Scholar] [CrossRef]
  84. Huang, G.; Zhou, Y.; Guo, Z.; Liu, X.; Chen, Y.; Liu, Q.; Ta, Z.; Wang, P.; He, Q.; Gao, J.; et al. The Influence of Dust Aerosols on Solar Radiation and Near-Surface Temperature during a Severe Duststorm Transport Episode. Front. Environ. Sci. 2023, 11, 1126302. [Google Scholar] [CrossRef]
  85. Ozdemir, E.; Tuna Tuygun, G.; Elbir, T. Application of Aerosol Classification Methods Based on AERONET Version 3 Product over Eastern Mediterranean and Black Sea. Atmos. Pollut. Res. 2020, 11, 2226–2243. [Google Scholar] [CrossRef]
  86. Chiapello, I.; Formenti, P.; Mbemba Kabuiku, L.; Ducos, F.; Tanré, D.; Dulac, F. Aerosol Optical Properties Derived from POLDER-3/PARASOL (2005–2013) over the Western Mediterranean Sea—Part 2: Spatial Distribution and Temporal Variability. Atmos. Chem. Phys. 2021, 21, 12715–12737. [Google Scholar] [CrossRef]
  87. Kaskaoutis, D.G.; Kosmopoulos, P.; Kambezidis, H.D.; Nastos, P.T. Aerosol Climatology and Discrimination of Different Types over Athens, Greece, Based on MODIS Data. Atmos. Environ. 2007, 41, 7315–7329. [Google Scholar] [CrossRef]
  88. Mallet, M.; Dubovik, O.; Nabat, P.; Dulac, F.; Kahn, R.; Sciare, J.; Paronis, D.; Leon, J.F. Absorption Properties of Mediterranean Aerosols Obtained from Multi-Year Ground-Based and Satellite Remote Sensing Observations. Atmos. Chem. Phys. Discuss. 2013, 13, 9195–9210. [Google Scholar] [CrossRef]
  89. Rose, W.I.; Durant, A.J. El Chichón Volcano, April 4, 1982: Volcanic Cloud History and Fine Ash Fallout. Nat. Hazards 2009, 51, 363–374. [Google Scholar] [CrossRef]
  90. Thomas, M.A.; Timmreck, C.; Giorgetta, M.A.; Graf, H.-F.; Stenchikov, G. Simulation of the Climate Impact of Mt. Pinatubo Eruption Using ECHAM5—Part 1: Sensitivity to the Modes of Atmospheric Circulation and Boundary Conditions. Atmos. Chem. Phys. 2009, 9, 757–769. [Google Scholar] [CrossRef]
  91. Robock, A. Climate Model Simulations of the Effects of the El Chichon Eruption. Geofis. Int. 1984, 23, 403–414. [Google Scholar] [CrossRef]
  92. Shi-Jun, S.; Shand, L.; Li, B. Tracing the Impacts of Mount Pinatubo Eruption on Global Climate Using Spatially-Varying Changepoint Detection. Ann. Appl. Stat. 2025, 19, 465–484. [Google Scholar] [CrossRef]
  93. Gharibzadeh, M.; Alam, K.; Abedini, Y.; Bidokhti, A.A. Classification of Aerosols Using Multiple Clustering Techniques over Zanjan, Iran, During 2010–2014. E3S Web Conf. 2019, 99, 02007. [Google Scholar] [CrossRef]
  94. Kalapureddy, M.C.R.; Kaskaoutis, D.G.; Ernest Raj, P.; Devara, P.C.S.; Kambezidis, H.D.; Kosmopoulos, P.G.; Nastos, P.T. Identification of Aerosol Type over the Arabian Sea in the Premonsoon Season During the Integrated Campaign for Aerosols, Gases and Radiation Budget (ICARB). J. Geophys. Res. Atmos. 2009, 114, D17203. [Google Scholar] [CrossRef]
  95. Banja, M.; Đukanović, G.; Belis, C.A. Status of Air Pollutants and Greenhouse Gases in the Western Balkans: Benchmarking the Accession Process Progress on Environment; Publications Office of the European Union: Luxemburg, 2020; p. 53. [Google Scholar]
  96. Dutton, E.G.; Reddy, P.; Ryan, S.; Deluisi, J.J. Features and Effects of Aerosol Optical Depth Observed at Mauna Loa, Hawaii: 1982–1992. J. Geophys. Res. Atmos. 1994, 99, 8295–8306. [Google Scholar] [CrossRef]
  97. Herber, A.; Thomason, L.W.; Radionov, V.F.; Leiterer, U. Comparison of Trends in the Tropospheric and Stratospheric Aerosol Optical Depths in the Antarctic. J. Geophys. Res. Atmos. 1993, 98, 18441–18447. [Google Scholar] [CrossRef]
  98. Long, C.S.; Stowe, L.L. Using the NOAA/AVHRR to Study Stratospheric Aerosol Optical Thicknesses Following the Mt. Pinatubo Eruption. Geophys. Res. Lett. 1994, 21, 2215–2218. [Google Scholar] [CrossRef]
  99. Toohey, M.; Sigl, M. Volcanic Stratospheric Sulfur Injections and Aerosol Optical Depth from 500ĝ€BCE to 1900ĝ€CE. Earth Syst. Sci. Data 2017, 9, 809–831. [Google Scholar] [CrossRef]
  100. Sharafa, S.B.; Aliyu, R.; Ibrahim, B.B.; Tijjani, B.I.; Darma, T.H.; Gana, U.M.; Ayedun, F.; Sulu, H.T. Model Prediction and Climatology of Aerosol Optical Depth (τ550) and Angstrom Exponent (α470-660) over Three Aerosol Robotic Network Stations in Sub-Saharan Africa Using Moderate Resolution Imaging Spectroradiometer Data. Niger. J. Technol. 2020, 39, 255–268. [Google Scholar] [CrossRef]
  101. Kambezidis, H.D.; Kaskaoutis, D.G.; Kalliampakos, G.K.; Rashki, A.; Wild, M. The Solar Dimming/Brightening Effect over the Mediterranean Basin in the Period 1979–2012. J. Atmos. Sol.-Terr. Phys. 2016, 150–151, 31–46. [Google Scholar] [CrossRef]
  102. Luccini, E.; Rivas, M.; Rojas, E. Cloud Optical Depth from Total and UV Solar Irradiance Measurements at Two Sites of the Atacama Desert in Chile. Atmos. Res. 2016, 174–175, 18–30. [Google Scholar] [CrossRef]
  103. Saragih, Y.F.; Darnius, O. Accuracy of the Moving Averages and Deseasonalizing Methods for Trend, Cyclical and Seasonal Data Forecasting. JMEA J. Math. Educ. Appl. 2023, 2, 143–151. [Google Scholar] [CrossRef]
  104. Montanari, A. Deseasonalisation of Hydrological Time Series through the Normal Quantile Transform. J. Hydrol. 2005, 313, 274–282. [Google Scholar] [CrossRef]
  105. Kwiatkowski, D.; Phillips, P.C.B.; Schmidt, P.; Shin, Y. Testing the Null Hypothesis of Stationarity against the Alternative of a Unit Root. J. Econom. 1992, 54, 159–178. [Google Scholar] [CrossRef]
  106. Li, Y.; Lau, N.-C. Impact of ENSO on the Atmospheric Variability over the North Atlantic in Late Winter—Role of Transient Eddies. J. Clim. 2012, 25, 320–342. [Google Scholar] [CrossRef]
Figure 1. The study’s four domains comprising the broader Mediterranean region: the left, pink-framed panel shows the West Mediterranean (WestMed); the middle, yellow-framed panel shows the Central Mediterranean (CentMed), the lower right, green-framed panel shows the Eastern Mediterranean (EastMed), and the upper right, black-framed panel shows the Balkans and Black Sea (BalBSea). The lower left corner of the entire domain has geographical coordinates (6° W, 30° N), and the upper right one (35° E, 45° N).
Figure 1. The study’s four domains comprising the broader Mediterranean region: the left, pink-framed panel shows the West Mediterranean (WestMed); the middle, yellow-framed panel shows the Central Mediterranean (CentMed), the lower right, green-framed panel shows the Eastern Mediterranean (EastMed), and the upper right, black-framed panel shows the Balkans and Black Sea (BalBSea). The lower left corner of the entire domain has geographical coordinates (6° W, 30° N), and the upper right one (35° E, 45° N).
Earth 06 00138 g001
Figure 2. Annual mean (observed) TAOD550, (estimated) TAOD’550, and difference ΔTAOD550 values over the four selected domains of the Mediterranean in the period 1980–2024: (a) WestMed, (b) CentMed, (c) EastMed, and (d) BalBSea. The dashed horizontal lines indicate ∆TAOD550 = 0%.
Figure 2. Annual mean (observed) TAOD550, (estimated) TAOD’550, and difference ΔTAOD550 values over the four selected domains of the Mediterranean in the period 1980–2024: (a) WestMed, (b) CentMed, (c) EastMed, and (d) BalBSea. The dashed horizontal lines indicate ∆TAOD550 = 0%.
Earth 06 00138 g002
Figure 3. Annual mean TAOD550 values over the four selected domains of the Mediterranean in the period 1980–2024. The dashed lines are linear regression fits to the TAOD550 curves.
Figure 3. Annual mean TAOD550 values over the four selected domains of the Mediterranean in the period 1980–2024. The dashed lines are linear regression fits to the TAOD550 curves.
Earth 06 00138 g003
Figure 4. Annual mean StAOD550 values over the four selected domains of the Mediterranean in the period 1980–2024: (a) WestMed, (b) CentMed, (c) EastMed, and (d) BalBSea. The dashed lines are linear regression fits to the StAOD550 curves.
Figure 4. Annual mean StAOD550 values over the four selected domains of the Mediterranean in the period 1980–2024: (a) WestMed, (b) CentMed, (c) EastMed, and (d) BalBSea. The dashed lines are linear regression fits to the StAOD550 curves.
Earth 06 00138 g004
Figure 5. Annual mean TAOD550,WVE values over the four selected domains of the Mediterranean in the period 1980–2024; WVE = without volcanic eruptions. The dashed lines are linear regression fits to the TAOD550,WVE curves.
Figure 5. Annual mean TAOD550,WVE values over the four selected domains of the Mediterranean in the period 1980–2024; WVE = without volcanic eruptions. The dashed lines are linear regression fits to the TAOD550,WVE curves.
Earth 06 00138 g005
Figure 6. Contour plots of the annual mean TAOD550 values as a function of ONI and NAOI in the period 1980–2024: (a) WestMed, (b) CentMed, (c) EastMed, and (d) BalBSea. The zone between the vertical blue and red solid lines shows neutral NAO events, while that between the horizontal blue and red dashed ones indicates neutral ENSO phases.
Figure 6. Contour plots of the annual mean TAOD550 values as a function of ONI and NAOI in the period 1980–2024: (a) WestMed, (b) CentMed, (c) EastMed, and (d) BalBSea. The zone between the vertical blue and red solid lines shows neutral NAO events, while that between the horizontal blue and red dashed ones indicates neutral ENSO phases.
Earth 06 00138 g006
Figure 7. Contour plots of the annual mean DARF (in Wm−2) values as a function of ONI and NAOI in the period 1980–2024: (a) WestMed, (b) CentMed, (c) EastMed, and (d) BalBSea. The vertical blue and red solid lines show neutral NAO events, while the horizontal blue and red dotted ones show neutral ENSO phases.
Figure 7. Contour plots of the annual mean DARF (in Wm−2) values as a function of ONI and NAOI in the period 1980–2024: (a) WestMed, (b) CentMed, (c) EastMed, and (d) BalBSea. The vertical blue and red solid lines show neutral NAO events, while the horizontal blue and red dotted ones show neutral ENSO phases.
Earth 06 00138 g007
Figure 8. Annual mean TAOD550 values as a function of ONI over the four selected domains of the Mediterranean in the period 1980–2024.
Figure 8. Annual mean TAOD550 values as a function of ONI over the four selected domains of the Mediterranean in the period 1980–2024.
Earth 06 00138 g008
Figure 9. Areas under the TAOD550 and TAOD550,WDDBC curves over the four selected domains of the Mediterranean in the period 1980–2024: (a) WestMed, (b) CentMed, (c) EastMed, and (d) BalBSea.
Figure 9. Areas under the TAOD550 and TAOD550,WDDBC curves over the four selected domains of the Mediterranean in the period 1980–2024: (a) WestMed, (b) CentMed, (c) EastMed, and (d) BalBSea.
Earth 06 00138 g009
Figure 10. Annual mean values of TAOD550, AAOD550, SAOD550, DDAOD550, and ONI in the period 1980–2024: (a) WestMed, (b) CentMed, (c) EastMed), and (d) BalBSea. The zone constrained by the blue and red horizontal lines indicates neutral ENSO events. ONI values ≥ +0.5 denote (warm) El Niňo cases and ≤−0.5 (cold) La Niňa events.
Figure 10. Annual mean values of TAOD550, AAOD550, SAOD550, DDAOD550, and ONI in the period 1980–2024: (a) WestMed, (b) CentMed, (c) EastMed), and (d) BalBSea. The zone constrained by the blue and red horizontal lines indicates neutral ENSO events. ONI values ≥ +0.5 denote (warm) El Niňo cases and ≤−0.5 (cold) La Niňa events.
Earth 06 00138 g010
Figure 11. Annual mean values of AE470–870 over the four Mediterranean domains in the period 1980–2024. (a) AE470–870 versus ONI; the zone constrained by the blue and red vertical lines include neutral ENSO phases. ONI ≥ +0.5 denote (warm) El Niňo cases and ONI ≤ −0.5 (cold) La Niňa events. The horizontal dashed line denotes the threshold of AE470–870 = 1.3; aerosol particles with AE470–870 > 1.3 are characterised as fine-mode and with AE470–870 < 1.3 as coarse-mode. (b) Temporal evolution of AE470–870; the dashed lines are linear regression fits to the AE470–870 time series.
Figure 11. Annual mean values of AE470–870 over the four Mediterranean domains in the period 1980–2024. (a) AE470–870 versus ONI; the zone constrained by the blue and red vertical lines include neutral ENSO phases. ONI ≥ +0.5 denote (warm) El Niňo cases and ONI ≤ −0.5 (cold) La Niňa events. The horizontal dashed line denotes the threshold of AE470–870 = 1.3; aerosol particles with AE470–870 > 1.3 are characterised as fine-mode and with AE470–870 < 1.3 as coarse-mode. (b) Temporal evolution of AE470–870; the dashed lines are linear regression fits to the AE470–870 time series.
Earth 06 00138 g011
Figure 12. Classification of the aerosol types over (a) WestMed, (b) CentMed, (c) EastMed, and (d) BalBSea in the period 1980–2024. Monthly mean values of TAOD550 and AE470–870 are plotted. CMA = clean-maritime aerosols; UIA = urban/industrial/biomass-burning aerosols; DDA = desert-dust aerosols; MTA = mixed-type aerosols. The vertical dashed lines indicate the threshold AE470–870 = 1.3.
Figure 12. Classification of the aerosol types over (a) WestMed, (b) CentMed, (c) EastMed, and (d) BalBSea in the period 1980–2024. Monthly mean values of TAOD550 and AE470–870 are plotted. CMA = clean-maritime aerosols; UIA = urban/industrial/biomass-burning aerosols; DDA = desert-dust aerosols; MTA = mixed-type aerosols. The vertical dashed lines indicate the threshold AE470–870 = 1.3.
Earth 06 00138 g012
Figure 13. Inter-annual variation in (a) the Ångström’s exponent, AE470–870, and (b) the direct aerosol radiative forcing, DARF (in Wm−2), over the four Mediterranean domains (1980–2024) and the entire Mediterranean (allMed, 1980–2022 [39]). The horizontal dashed line in (a) indicates the threshold AE470–870 = 1.3.
Figure 13. Inter-annual variation in (a) the Ångström’s exponent, AE470–870, and (b) the direct aerosol radiative forcing, DARF (in Wm−2), over the four Mediterranean domains (1980–2024) and the entire Mediterranean (allMed, 1980–2022 [39]). The horizontal dashed line in (a) indicates the threshold AE470–870 = 1.3.
Earth 06 00138 g013
Figure 14. Six-month mean values of ONI and DARF in the period 1980–2024 over (a) WestMed, (b) CentMed, (c) EastMed), and (d) BalBSea. The horizontal blue and red solid lines define the band of neutral ENSO events. The ellipses enclose decreases in ONI and retarded DARF ones, both indicated by red arrows. The distance between two adjacent tick marks on the x-axis corresponds to the duration of a semester (six-month period).
Figure 14. Six-month mean values of ONI and DARF in the period 1980–2024 over (a) WestMed, (b) CentMed, (c) EastMed), and (d) BalBSea. The horizontal blue and red solid lines define the band of neutral ENSO events. The ellipses enclose decreases in ONI and retarded DARF ones, both indicated by red arrows. The distance between two adjacent tick marks on the x-axis corresponds to the duration of a semester (six-month period).
Earth 06 00138 g014
Figure 15. P-P plots of SSN (a), NAOI (c), ONI (e) with influence on SSA, WestMed (b), SAOD550, CentMed (d), and AAOD550, BalBSea (f). All data points in the plots are monthly mean values in the period 1980–2024. The x-axis shows the values of the variable calculated by the normal distribution function having the same μ and σ with those of the variable’s time series. The y-axis is the actual probability curve of the variable. The solid black line represents the function x = y.
Figure 15. P-P plots of SSN (a), NAOI (c), ONI (e) with influence on SSA, WestMed (b), SAOD550, CentMed (d), and AAOD550, BalBSea (f). All data points in the plots are monthly mean values in the period 1980–2024. The x-axis shows the values of the variable calculated by the normal distribution function having the same μ and σ with those of the variable’s time series. The y-axis is the actual probability curve of the variable. The solid black line represents the function x = y.
Earth 06 00138 g015aEarth 06 00138 g015b
Table 1. Justification of criteria for dividing the entire Mediterranean into four segments. The criteria refer to the climatology, atmospheric aerosols, geography, and solar irradiance levels in the four areas.
Table 1. Justification of criteria for dividing the entire Mediterranean into four segments. The criteria refer to the climatology, atmospheric aerosols, geography, and solar irradiance levels in the four areas.
A. ClimatologyB. AerosolsC. GeographyD. Solar radiationReferences
Western and eastern basins have different precipitation regimes.
Northern areas (Atlantic, Black Sea) experience more continental weather patterns.
Western: Atlantic and Saharan dust.
Central: mixed influence.
Eastern: Middle East and Anatolian sources.
Northern: European industrial and biomass-burning aerosols.
Each subregion clusters seas and coasts with shared topography and oceanographic features.Highest in East and Central Mediterranean.
Lower in the northern and western parts due to higher cloudiness and latitude.
Table 2. List of the parameters used in the present work with a short description.
Table 2. List of the parameters used in the present work with a short description.
ParameterRemarks
ONIIndex of the El Niño–Southern Oscillation: a standard for tracking the oceanic part of the ENSO climate pattern. The current standard is ONI3.4. In the rest of the text, Figures and Tables ONI implies ONI3.4. The NOAA’s scientists have determined the (−0.5, +0.5) zone as neutral ENSO occurrences; further, they have defined warm El Niňo events for ONI ≥ +0.5 and cold La Niňa ones for ONI ≤ −0.5 (see https://cpc.ncep.noaa.gov/products/analysis_monitoring/ensostuff/ONI_v5.php (accessed on 8–25 October 2023)).
NAOIIndex of the North Atlantic Oscillation: a measure of the surface sea-level pressure difference between the subtropical High at Azores and the subpolar Low over Iceland. A neutral zone similar to that for ONI is established here. Therefore, it makes sense to use the same delimitation as ONI; neutral NAO events in the band (−0.5, +0.5), positive NAO cases for NAOI ≥ +0.5, and negative NAO events for NAOI ≤ −0.5.
SSNSunspot number: the number of sunspots on the surface of the Sun associated with the 11-year solar activity.
TAOD550Total aerosol optical depth at 550 nm: a measure of the total solar radiation extinction by aerosols at 550 nm.
SAOD550Scattering aerosol optical depth at 550 nm: a measure of the solar radiation extinction by scattering aerosols at 550 nm.
AAOD550Absorbing aerosol optical depth at 550 nm: a measure of the solar radiation extinction by absorbing aerosols at 550 nm.
DDAOD550Desert-dust aerosol optical depth at 550 nm: a measure of the solar radiation extinction by scattering desert-dust aerosols at 550 nm.
StAOD550Stratospheric aerosol optical depth at 550 nm: a measure of the solar radiation extinction by scattering stratospheric aerosols at 550 nm (mainly sulphates erupted by volcanoes).
SSAOD550Sea-salt aerosol optical depth at 550 nm: a measure of the solar radiation extinction by sea-salt particles at 550 nm.
BCAOD550Black-carbon aerosol optical depth at 550 nm: a measure of the solar radiation extinction by absorbing black-carbon aerosols at 550 nm.
AE470–870Ångström’s exponent: a measure of the aerosols size in the visible band of 470–870 nm. A typical value of 1.3 is often used.
SSASingle-scattering albedo: a measure of the scattering property of aerosols; if SSA = 1, the particles are purely scatterers; if SSA ≈ 0, the particles only absorb.
DARFDirect aerosol radiative forcing: it refers to the scattering and/or absorption of solar radiation by atmospheric aerosols. Positive DARF values indicate local/regional warming, while negative values indicate local/regional cooling.
CODCloud optical depth: a measure of the solar radiation extinction by clouds. It plays significant role in the Earth’s energy budget.
netSWRBOA,CS,ANet SW radiation at BOA, CS, and A: difference between the incoming and reflected radiation at the surface of the Earth under clear skies and presence of aerosols in the atmosphere.
netSWRBOA,CS,NANet SW radiation at BOA, CS, and NA: difference between the incoming and reflected radiation at the surface of the Earth under clear skies and absence of aerosols in the atmosphere.
netSWRTOA,CS,ANet SW radiation at TOA, CS, and A: difference between the incoming and outgoing radiation at 100 km from the surface of the Earth under clear skies and presence of aerosols in the atmosphere.
netSWRTOA,CS,NANet SW radiation at TOA, CS, and A: difference between the incoming and outgoing radiation at 100 km from the surface of the Earth under clear skies and absence of aerosols in the atmosphere.
netLWRBOA,CS,ANet LW radiation at BOA, CS, and A: difference between the incoming and emitted infra-red (IR) radiation at the surface of the Earth under clear skies and presence of aerosols in the atmosphere.
netLWRBOA,CS,NANet LW radiation at BOA, CS, and A: difference between the incoming and emitted IR radiation at the surface of the Earth under clear skies and absence of aerosols in the atmosphere.
netLWRTOA,CS,ANet LW radiation at TOA, CS, and A: difference between the incoming and outgoing IR radiation at 100 km from the surface of the Earth under clear skies and presence of aerosols in the atmosphere.
netLWRTOA,CS,NANet LW radiation at TOA, CS, and A: difference between the incoming and outgoing IR radiation at 100 km from the surface of the Earth under clear skies and absence of aerosols in the atmosphere.
Table 3. Confusion matrix from a DTC analysis of the frequencies of the monthly mean modelled and observed TAOD550 and AE470–870 values.
Table 3. Confusion matrix from a DTC analysis of the frequencies of the monthly mean modelled and observed TAOD550 and AE470–870 values.
Modelled ValuesObserved Values
WestMedCentMedEastMedBalBSea
WestMed12660
CentMed4714
EastMed3891
BalBSea46721
Table 4. Inter-period 1980–2024 mean TAOD550 and AE470–870 values for the four clusters and characteristic aerosol regimes over the four subregions. The symbols in parentheses denote the four dominant aerosol types set up in Section 3.1.5 (see Figure 12).
Table 4. Inter-period 1980–2024 mean TAOD550 and AE470–870 values for the four clusters and characteristic aerosol regimes over the four subregions. The symbols in parentheses denote the four dominant aerosol types set up in Section 3.1.5 (see Figure 12).
ClusterAerosol RegimeTAOD550AE470–870Dominant
Subregion
1Air pollution (UIA)0.1971.151BalBSea, EastMed
2Maritime/dust (DDA)0.2560.760WestMed, CentMed
3Secondary aerosols (MTA)0.3481.444BalBSea
4Clean-maritime (CMA)0.1370.844WestMed, EastMed
Table 5. Annual averages, μ, standard deviations, σ, medians, m, and trends of the parameters considered, q, in the present work over the four Mediterranean domains as a function of time (years in the examined period 1980–2024). All AODs are considered at 550 nm; AE in the band 470–870 nm; netSWR and netLWR under the BOA, CS, and A conditions; DARF, netSWR, and netLWR statistics have units in Wm−2. In the trend rows (fourth row in each Mediterranean area), the parameters downwards indicate the slope α (in units/decade), R2, and CI (* for 95% CI, and *** for 99.9% CI), respectively; absence of * in the trend indicates significance at CI < 95%. All numbers have been rounded to the third decimal digit. Σ = statistic parameter.
Table 5. Annual averages, μ, standard deviations, σ, medians, m, and trends of the parameters considered, q, in the present work over the four Mediterranean domains as a function of time (years in the examined period 1980–2024). All AODs are considered at 550 nm; AE in the band 470–870 nm; netSWR and netLWR under the BOA, CS, and A conditions; DARF, netSWR, and netLWR statistics have units in Wm−2. In the trend rows (fourth row in each Mediterranean area), the parameters downwards indicate the slope α (in units/decade), R2, and CI (* for 95% CI, and *** for 99.9% CI), respectively; absence of * in the trend indicates significance at CI < 95%. All numbers have been rounded to the third decimal digit. Σ = statistic parameter.
AreaΣAerosol/Atmospheric Parameter, q
TAODAAODSAODDDAODStAODSSAODBCAODAESSADARFnetSWRnetLWRCOD
WestMedμ0.2030.0120.1910.1030.0680.0110.0060.8740.9394.587194.532−99.37515.144
σ0.0340.0010.0340.0130.0370.0010.0010.1850.0100.4570.9071.5471.354
m0.1980.0120.1840.1020.0530.0110.0060.8130.9354.557194.721−99.50515.108
α
R2
sign.
−0.020
0.402
***
+4 × 10−4
0.263
***
−0.020
0.415
***
−4 × 10−4
0.006
 
−0.020
0.434
***
+8 × 10−4
0.629
***
+6 × 10−4
0.711
***
−0.078
0.331
***
−6 × 10−3
0.702
***
+0.205
0.376
***
+0.137
0.042
***
+0.401
0.119
*
+0.049
0.003
 
CentMedμ0.2280.0120.2160.0930.0930.0200.0060.9280.9463.980209.044−95.65420.408
σ0.0370.0010.0370.0130.0440.0100.0010.1670.0090.4041.0841.2381.780
m0.2200.0120.2070.0940.0750.0180.0070.8820.9423.995209.298−95.69220.241
α
R2
sign.
−0.020
0.510
***
0.000
0.208
*
−0.020
0.520
***
−0.002
0.000
 
−0.029
0.588
***
+0.003
0.196
*
0.000
0.336
***
−0.078
0.348
***
−0.006
0.694
***
+0.156
0.265
***
+0.303
0.137
*
+0.342
0.139
*
+0.020
0.000
 
EastMedμ0.2050.0110.1950.0800.0890.0180.0060.9350.9483.258214.748−99.22018.337
σ0.0310.0010.0310.0120.0360.0020.0010.1790.0090.5711.0451.4841.776
m0.1980.0110.1880.0810.0730.0180.0060.8690.9433.385214.751−99.48318.214
α
R2
sign.
−0.010
0.271
***
+0.001
0.319
***
−0.010
0.292
***
+0.002
0.073
 
−0.020
0.399
***
+0.001
0.719
***
+0.000
0.592
***
−0.078
0.366
***
−0.005
0.601
***
+0.274
0.426
***
−0.098
0.017
 
+0.743
0.447
***
+0.176
0.017
 
Bal
BSea
μ0.2340.0100.2240.0420.1500.0110.0081.2170.9583.286196.987−91.25321.024
σ0.0560.0010.0550.0070.0580.0010.0010.1490.0090.4181.6081.8182.161
m0.2050.0100.1950.0420.1180.0110.0081.1650.9533.386197.569−91.08720.602
α
R2
sign.
−0.039
0.738
***
0.000
0.001
 
−0.039
0.741
***
+0.001
0.015
 
−0.039
0.750
***
+0.001
0.627
***
0.000
0.020
 
−0.078
0.481
***
−0.006
0.745
***
+0.059
0.036
 
+0.841
0.494
***
+0.694
0.262
***
+0.147
0.008
 
Table 6. Summary of the SSN/NAOI/ONI GC test on the time series of the aerosol/atmospheric parameters examined in the present work over the four Mediterranean domains in the period 1980–2024. DARF, netSWR, and netLWR statistics have units in Wm−2. “Yes” implies a cause–effect of any of the SSN/NAOI/ONI time series on the examined variable. “No” means no cause–effect. All parameters in the time series consist of their monthly mean deseasonalised or deseasonalised/first-order differenced values. A “Yes” under the heading of SSN, NAOI, or ONI refers to any lagged value of the phenomenon between one and six months.
Table 6. Summary of the SSN/NAOI/ONI GC test on the time series of the aerosol/atmospheric parameters examined in the present work over the four Mediterranean domains in the period 1980–2024. DARF, netSWR, and netLWR statistics have units in Wm−2. “Yes” implies a cause–effect of any of the SSN/NAOI/ONI time series on the examined variable. “No” means no cause–effect. All parameters in the time series consist of their monthly mean deseasonalised or deseasonalised/first-order differenced values. A “Yes” under the heading of SSN, NAOI, or ONI refers to any lagged value of the phenomenon between one and six months.
ParameterWestMedCentMedEastMedBalBSea
SSNNAOIONISSNNAOIONISSNNAOIONISSNNAOIONI
TAOD550YesYesYesYesYesYesYesYesYesYesYesYes
AAOD550YesYesYesYesYesYesYesYesYesYesYesNo
SAOD550YesYesYesYesYesYesYesYesYesYesYesYes
DDAOD550YesYesYesYesYesYesNoNoNoNoNoNo
StAOD550YesYesYesNoNoNoNoNoNoYesYesYes
SSAOD550NoNoNoYesYesYesYesNoYesNoNoNo
BCAOD550NoNoNoNoNoNoNoNoNoNoNoNo
AE470–870NoNoNoNoNoNoNoNoNoNoNoNo
SSANoNoNoNoNoNoNoNoNoYesYesYes
DARFYesYesYesYesYesYesYesYesYesYesYesNo
netSWRBOA,CS,AYesYesNoYesYesYesYesYesYesYesYesYes
netLWRBOA,CS,AYesYesNoYesYesYesYesYesYesYesYesYes
CODYesYesYesYesYesYesYesYesYesNoNoNo
Table 7. Phase of the NAO- and ENSO-circulation phenomena favouring maximum values of the atmospheric parameter investigated. Red colour = positive phase; green colour = neutral phase; blue colour = negative phase. DARF, netSWRBOA,CS,A, and netLWRBOA,CS,A have units in Wm−2.
Table 7. Phase of the NAO- and ENSO-circulation phenomena favouring maximum values of the atmospheric parameter investigated. Red colour = positive phase; green colour = neutral phase; blue colour = negative phase. DARF, netSWRBOA,CS,A, and netLWRBOA,CS,A have units in Wm−2.
ParameterWestMedCentMedEastMedBalBSea
TAOD550NAO/ENSONAO/ENSONAO/ENSONAO/ENSO
AAOD550NAO/ENSONAO/ENSONAO/ENSONAO/ENSO
NAO/ENSO
SAOD550NAO/ENSONAO/ENSONAO/ENSONAO/ENSO
DDAOD550NAO/ENSONAO/ENSONAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
BCAOD550NAO/ENSONAO/ENSONAO/ENSONAO/ENSO
NAO/ENSO
AE470–870NAO/ENSONAO/ENSONAO/ENSONAO/ENSO
NAO/ENSO
SSANAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
DARFNAO/ENSONAO/ENSONAO/ENSOvarious peaks
netSWRBOA,CS,ANAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
netLWRBOA,CS,ANAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
CODNAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
NAO/ENSO
Table 8. Estimation of DARF without volcanic activity in the atmosphere, DARFWVE, from simple calculations taking into account TAOD550, TAOD550,WVE, and DARF as average values over the four Mediterranean subregions in the period 1980–2024.
Table 8. Estimation of DARF without volcanic activity in the atmosphere, DARFWVE, from simple calculations taking into account TAOD550, TAOD550,WVE, and DARF as average values over the four Mediterranean subregions in the period 1980–2024.
ParameterWestMedCentMedEastMedBalBSea
TAOD5500.2030.2280.2050.234
TAOD550,WVE0.1350.1400.1170.084
DARF (Wm−2)4.5873.9803.2583.286
DARFWVE (Wm−2)6.8976.4825.7089.154
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kambezidis, H.D. Atmospheric Processes over the Broader Mediterranean Region 1980–2024: Effect of Volcanoes, Solar Activity, NAO, and ENSO. Earth 2025, 6, 138. https://doi.org/10.3390/earth6040138

AMA Style

Kambezidis HD. Atmospheric Processes over the Broader Mediterranean Region 1980–2024: Effect of Volcanoes, Solar Activity, NAO, and ENSO. Earth. 2025; 6(4):138. https://doi.org/10.3390/earth6040138

Chicago/Turabian Style

Kambezidis, Harry D. 2025. "Atmospheric Processes over the Broader Mediterranean Region 1980–2024: Effect of Volcanoes, Solar Activity, NAO, and ENSO" Earth 6, no. 4: 138. https://doi.org/10.3390/earth6040138

APA Style

Kambezidis, H. D. (2025). Atmospheric Processes over the Broader Mediterranean Region 1980–2024: Effect of Volcanoes, Solar Activity, NAO, and ENSO. Earth, 6(4), 138. https://doi.org/10.3390/earth6040138

Article Metrics

Back to TopTop