Next Article in Journal
Spatiotemporal Dynamics of Land Cover and Their Driving Forces in the Yellow River Basin since 1990
Previous Article in Journal
The Spatial Pedigree in Traditional Villages under the Perspective of Urban Regeneration—Taking 728 Villages in Jiangnan Region, China as Cases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Risks Assessment of Potentially Toxic Elements’ Contamination in the Egyptian Red Sea Surficial Sediments

1
Geology Department, Faculty of Science, Suez University, El Salam City 43518, Egypt
2
Geology Department, Faculty of Science, Ain Shams University, Cairo 11566, Egypt
3
National Research Institute of Astronomy and Geophysics (NRIAG), Cairo 11421, Egypt
4
Department of Biological and Geological Sciences, Faculty of Education, Ain Shams University, Cairo 11341, Egypt
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Land 2022, 11(9), 1560; https://doi.org/10.3390/land11091560
Submission received: 16 August 2022 / Revised: 6 September 2022 / Accepted: 9 September 2022 / Published: 14 September 2022
(This article belongs to the Special Issue Advances in Coastline Evolution)

Abstract

:
The potential impact of tourism, industrial, and urban activities on Egypt’s Red Sea coastline, which is well-known for its economic and environmental importance, was investigated at fifteen coastal sites. In the present study, the concentration of cadmium (Cd), cobalt (Co), chromium (Cr), copper (Cu), iron (Fe), manganese (Mn), nickel (Ni), lead (Pb), and zinc (Zn) in marine sediments from these sites, was determined using Inductively Coupled Plasma Mass Spectrometers (ICP-MS). In detail, various pollution indices, statistical analyses, and spatial distribution patterns were used to assess the pollution status, impacts of human activities, ecological risks, and sources of potentially toxic elements (PTEs) in surface marine sediment. A detailed comparison with up-to-date data was conducted. These sediments were composed predominantly of fine and very fine sands. Mean grain size distribution typically depends on the source of the sediment from the following two prime sources: terrigenous (autochthonous) and biogenic (allochthonous). The detected PTE mean concentrations were as follows: Fe > Mn > Zn > Cr > Ni > Pb > Co > Cu > Cd. Multivariate statistical analysis results revealed their close distribution and association. Cd and Pb levels in the studied area have been slightly impacted by anthropogenic inputs. According to the calculated pollution indices, although a minimal or moderate contamination degree was detected in the study area, it was determined that there was a low to moderate ecological risk. The slightly high degree of contamination and risk centered in the middle of the study area around phosphate mining and related activities. More attention should be given to the concentrations and sources of Cd, Ni, and Pb as the main pollution factors.

1. Introduction

Potentially Toxic Elements (PTEs) are vital indicators for aquatic ecosystems since they originate from multiple sources and have bio-accumulative, persistent, and toxic characteristics [1,2]. PTEs are imported into aquatic ecosystems through both natural and anthropogenic processes. They are typically released through natural processes such as rock weathering and erosion, and they strike the water via various transport pathways. Anthropogenic activities, rapid industrial growth, and urbanization represent the critical factors contributing to enhanced PTE concentrations over recent decades [3,4,5,6]. For example, Christophoridis et al. [7] reported a hotspot of Cd, Cu, Ni, Pb, and Zn contamination in the northern part of Thessaloniki Bay, Greece, indicating the cumulative effect of the anthropogenic load from industrial and maritime activities on urban coastal zones. Looi et al. [8] linked PTE pollution in the Tanjung Dawai’s marine sediments, Malaysia, to fisheries, aquaculture, paddy cultivation, use of insecticides and arsenical herbicides, the leaching of antifouling paints, and the discharge of wastewater sludge. Hoai et al. [9] found that stations near river inflows, urban areas, and coal mining had greater As, Cu, and Pb concentrations than offshore stations in Vietnam’s Ha Long Bay. Radomirović et al. [10] ascribed elevated As, Cr, Cu, Ni, Pb, and Zn concentrations in Tivat bay (Montenegro) marine sediments to shipyard activities, grit excavation, port waste, and wastewater discharges. Apaydın et al. [11] noted that concentrations of Cu, Pb, and Zn were higher in marine sediments in the Middle and Eastern Black Sea region, Turkey, due to natural rock erosion as well as anthropogenic inputs (mining, agricultural activities, and traffic emissions). Oura et al. [12] attributed the elevated As, Cd, Cr, Sb, and Pb concentrations in marine sediments recorded along the Ivorian coast to maritime trade, offshore oil production, and petrochemical plants.
Currently, PTE contaminations in coastal environments are increasingly encountered and are considered one of the most pressing global issues [9,13,14,15,16,17]. PTEs are thought to be the main pollutants which directly affect the marine environment [18,19]. While PTEs such as Cu, Ni, and Zn are necessary for aquatic life, they can be harmful at certain concentrations, posing a threat to aquatic biological diversity. PTE-contaminated sediment and water have toxic effects on aquatic fauna and flora [8,20,21,22,23,24]. Because aquatic living organisms are part of the food chain, increasing PTE levels in the aquatic environment have toxic impacts not only on aquatic living organisms but also on humans. Humans’ exposure to PTEs occurs via the gastrointestinal tract (water drinking, food ingestion), inhalation, and dermal contact [25,26]. The main toxic impacts on human health associated with exposure to PTEs are cancers, dermatitis and muscular impairment, reproductive system damage, low infant birth weight, lactation problems, gastrointestinal and lung diseases, kidney pathology, immune system dysfunction, and neurological disorders [26]. Thus, the measurement of ecological and toxicological effects of some PTEs is meaningful for evaluating their severity degree [27,28]. Consequently, contamination by PTEs represents a considerable socioeconomic problem and has become a worldwide environmental concern with potential long-term effects on humans and the ecosystem [18,19,25].
Coastal sediment acts as an important sink of PTEs in the marine environment and plays a fundamental role in element transmission, complexation, and deposition [29]. In addition, coastal sediments are considered to be conveyors and potential secondary sources of PTEs in the marine environment [30]. PTE pollutants cannot be frequently settled in sediment; they may be released back into the water column because of possible changes in environmental conditions such as redox potential, temperature, pH, and microbial activity [31]. However, PTEs are expected to be more concentrated in the sediment in the aquatic system than in the water body [23,32]. Therefore, continual surveys of pollutant concentrations (e.g., PTEs) in the sediment of coastal areas are not only valuable for assessing pollution in the aquatic environment but moreover provide essential updated information on environmental health status and help to evaluate future risks [33].
The Red Sea has immense universal value due to its facilitation of important maritime transportation between six countries and as a connection between two continents (Africa and Asia). The Red Sea is characterized by a length of ≈1930 km with an average width of 280 km [34,35]. In the past two decades, the Red Sea has been regarded as one of the most important industrial, commercial, tourism and recreation centers in Egypt. The economic development activities on the Egyptian side of the Red Sea have been accompanied by ecological damage. Recently, the Red Sea coastal area has suffered from multiple environmental challenges such as coastal activities, shipping, maritime transport, spillage problems, industrial and sewage effluents, and natural oil seeps [29]. As a result, these continuous inputs have generated tremendous and diverse types of contaminants, including PTEs. Therefore, the present study attempted to provide the most recent picture of PTE levels along the Egyptian side of the Red Sea coastal area and to identify their potential sources. In addition, performing a multi-index risk assessment of PTEs in the studied marine sediment provides decision makers with data for a future control plan.

2. Materials and Methods

2.1. Sampling and Samples Preparation

Coastal Red Sea sediments are the product of a complex interplay of sedimentary material influx from aeolian, fluvial, and marine sources. The fluvial sediment is deposited by intermittently active wadies as a consequence of the weathering of Miocene and younger sedimentary succession and the Precambrian basement rocks of Red Sea mountains. In the current study, a total of fifteen coastal sites along the Egyptian side of the Red Sea coast were investigated (Figure 1 and Table 1). The investigated coastal locations were extended to include different natural and anthropogenic probable sources. From every sampling site, up to 15 subsamples were collected during the field survey (July 2020), utilizing the Van Veen Grab Sampling tool (thickness of about 10 cm) to represent a perfect sample that reflects chemical, physical and biological exchange processes. The sample collection strictly followed the standard methods reported by [36]. The collected sediment samples were preserved in airtight polyethylene bags, labeled, and transported to the Geology Department, Faculty of Science, Suez University Labs, for preparation. The Collected samples were air-dried to a constant weight, carefully mixed, homogenized, and divided into subsamples for further analyses.

2.2. Sediment Granulometry

In order to determine grain-size distribution, the air-dried sediment samples were analyzed using the dry sieving technique [37]. A total of 100 g of each sample was separated using a Ro-Tap shaking machine for 20 m, with a set of standard mesh sieves (−1, 0, 1, 2, 3, and 4 Ø). For samples that contained mud (<64 µm) portions of >5%, the pipetting technique was used to separate silt and clay fractions [38]. Using data obtained from cumulative curves, the textural parameters including mean size (Mz), kurtosis (KG), skewness (SkI), and sorting (σI) were calculated according to [39]. Samples preparation and granulometry analyses were conducted at the Geology Department, Faculty of Science, Suez University Laboratories.

2.3. Geochemical Analyses

After multi-acid solution digestion, PTE total concentrations were measured in the well-powdered sediment samples. An acid mixture (HF + HCl + HNO3; 1:4:5, in volume) was added to an accurate weight of 0.25 g of a sample in a pre-cleaned Teflon crucible. It was gently heated by using a heating plate. The digestion was continued until a small amount of the white residue was obtained. Then, the solution was heated until nearly dry and extracted with HNO3 [40]. Cadmium, Co, Cr, Cu, Fe, Mn, Ni, Pb, and Zn concentrations were determined with Inductively Coupled Plasma Mass Spectrometers (Varian 810/820-MS ICP, Varian Inc., Mulgrave, VIC, Australia). The geochemical analyses were carried out at the accredited Reference Laboratory for Drinking Water, Holding Company for Water and Wastewater, Shubra El-Khima Water Treatment Plant, Cairo, Egypt (ISO 17025).
The Teflon vessel and glassware were cleaned by soaking in a 10% (v/v) HNO3 solution for 15 min and then rinsed with deionized water about three times and dried before use. All solutions were carefully prepared by using double distilled water. Reagent-grade chemicals (HF, HCl, and HNO3; Merck, Darmstadt, Germany) were used for all purposes. The accuracy and precision of the analytical procedures were examined by analyzing a standard reference material (HISS-1, marine reference sediment, NRCC). The recovery of the measured elements fluctuated between 96.47% and 108.53%, with the uncertainty of the measured PTEs ranging from 1.9% to 5.0%. The detection limits of Cadmium, Co, Cr, Cu, Fe, Mn, Ni, Pb, and Zn were 0.53, 1.00, 0.60, 0.65, 2.00, 0.20, 0.60, 2.00, and 3.00 (µg/L), respectively.

2.4. Pollution Assessment

Environmental indices are useful mathematical tools for both analyzing and simplifying complex environmental data, and they are a valuable technique in evaluating environmental risk and sediment pollution [41]. In addition, indices make it easier to evaluate PTE pollution results in terms of their origin, availability, and toxicity [42]. Furthermore, contamination indices can be used to determine whether the accumulation of PTEs is caused by natural or anthropogenic sources. In data interpretation, selecting background values is critical. Background values are typically used to distinguish between the natural quantities of elements and their elevated concentrations due to anthropogenic influences [43]. As of yet, a national Egyptian background concentration for PTEs in sediment has not been determined, as well as sediment quality guidelines. The average element levels in the upper continental crust were utilized in the absence of local and regional geochemical background levels [7,8,14,44].
Single pollution indices such as the geo-accumulation index (Igeo) [45], and the contamination factor (Cf) [46] were calculated to quantify the anthropogenic inputs of every element above its natural levels in Red Sea sediments. Furthermore, integrated pollution indices such as degree of contamination (Cdeg) [46], the Ecological Risk Index (RI) [46,47], Contamination Security Index (CSI) [42], and mean ERM quotient (MERMQ) [48] were calculated to conduct an integrative assessment of sediment ecosystem quality at every sampling site and evaluate the potential ecological risks posed by the PTE contamination considering multi-element contamination. The categories of single and integrated pollution indices are summarized in Table S1 (in Supplementary Materials).

2.4.1. Geo-Accumulation Index (Igeo)

Igeo is widely used in studies concerned with the assessment of PTE pollution in coastal areas. According to [45], Igeo was calculated using Equation (1).
I g e o = log 2 ( C s 1.5   C r e f )
where Cs is the concentration of PTE in the investigated sample and Cref is the reference value.

2.4.2. Contamination Factor (Cf)

Cf is a direct and the most straightforward pollution indicator used to assess PTE contamination levels in coastal sediment [4,22,23,29]. Cf is calculated using Equation (2) [46].
C f i = C s i / C b i
where C s i is the element concentration in the analyzed samples, and C b i is the reference value of investigated PTEs.

2.4.3. Degree of Contamination (Cdeg)

Cdeg is considered a more comprehensive method for understanding the contamination level and was calculated for the Red Sea sediments according to Equation (3) [46].
C d e g = i = 1 n C f
where Cdeg is the degree of contamination of investigated sediment, Cf is the contamination factor, and n is the number of the examined PTEs.

2.4.4. Ecological Risk Index (RI)

Among the ecological risk indices, the individual ecological risk factor ( E i r ) and overall ecological risk index (RI) were investigated. ( E i r ) was suggested by Hakanson [46] for assessing individual ecological risk factors according to Equation (4).
( E i r ) = C f i × T i r
The overall ecological risk caused by all toxic elements (RI) was determined as the sum of E i r as shown in Equation (5).
RI = i = 1 n E i r
where C f i , T i r and n represent the contamination factor, toxicological response factor for each toxic element proposed by Hakanson [46] (1, 2, 5, 5, 5, and 30 for Zn, Cr, Pb, Cu, Ni, and Cd, respectively), and the number of investigated metals, respectively.

2.4.5. Contamination Security Index (CSI)

Although CSI is a relatively new index, it is considered valuable. It is used to evaluate the toxicity limit and both biological and ecological impacts for most PTEs [49]. CSI was adopted by Pejman et al. [42] and can be calculated using Equation (6).
CSI = i = 1 n W   [   ( C S ERM ) ² + ( C S ERL ) ½ ]
where CS, n, W, ERM and ERL are the metal concentration, number of investigated toxic elements, computed weight of element (0.25, 0.134, 0.075, 0.251, 0.215 and 0.075), effects range median ERM (9.6, 370, 270, 218, 51.6 and 410), and effects range low (1.2, 81, 34, 46.7, 20.9 and 150) for Cd, Cr, Cu, Pb, Ni and Zn, respectively [48].

2.4.6. Mean ERM Quotient (MERMQ)

MERMQ was implemented to assess the ecological impacts of PTEs with an accurate scale. The mean effects range-medium quotient (MERMQ) was calculated according to Equation (7) [50].
MERMQ = i = 1 n ( C S ERM ) n
where CS, ERM, and n were element concentration, effects range median, and number of investigated toxic elements, respectively [48].

2.5. Data Treatment

Sediment samples were classified using the ternary chart adopted by Folk [37]. Both cumulative frequency and cumulative weight-percentage curves were generated and studied to identify potential grain-size trends. PTEs concentration and the calculated integrated pollution indices were represented in a distribution pattern using the proportional/graduated symbols method (using Arc-Map 10.3, ESRI, Inc., Redlands, CA, USA). To determine if the PTE concentrations were normally distributed, the Kolmogorov–Smirnov (K-S) test was performed using OriginLab (version OriginPro 2021). Furthermore, statistical analyses including Pearson’s Correlation Coefficient (PCC), a Principal Component Analysis (PCA R-Mode), and Cluster Analysis (CA R-Mode and Q-Mode) were conducted using IBM SPSS (version 23, Armonk, NY, USA) and OriginLab (version OriginPro 2021) to investigate relations between the studied PTEs in terms of their geochemical affinity and their origins. CA was applied to show similarity relations between the studied metals, their association and to illustrate different geochemical groups in a dendrogram.

3. Results and Discussion

3.1. Textural Attributes

The results of the grain-size analysis of the collected samples are given in Table 2. The data reveal that Red Sea sediment is composed of gravelly sand (40%), muddy gravelly sand (33.33%), muddy sand (20%) and gravelly muddy sand (6.67%). Sand fractions were found to be dominant in all the studied sites (54.8–97.4%). On the other hand, gravel and mud fractions have no clear trend. It is evident that Red Sea sediments are mainly of terrigenous origins and are composed mainly of fine to very fine sands; they are also rich in terrigenous constituents that originate from surrounding areas. Gravel distribution depends on topographical features and benthic fauna shell fragments (biogenic fragments), leading to an increased gravel fraction in many sites. The limited distribution of fine fractions in the studied area may be due to a low fine sediment supply to coastal areas and/or the re-transportation of fine fractions to deeper areas as a result of the longshore current. Mud increased in areas of lagoons, where stagnant water makes it possible for mud (silt and clay fractions) to deposit [4,23,29,51].
The mean grain size varied between medium sand and very fine sand. Sorting fluctuated between well sorted and very poorly sorted. The skewness ranged from strongly coarse-skewed to strongly fine-skewed. Kurtosis varied from platy-kurtic to very leptokurtic (Table 2). The difference in the textural parameters implies that they are primarily influenced by dynamic processes. The variations in the particle size parameters in aquatic sediment depend on different factors (act on sediment together or separately), such as sedimentary processes, transportation of sediment, and the distance between source material and the deposition basin [52,53,54]. The studied sediments have two main sediment sources, namely terrigenous (autochthonous) and biogenic (allochthonous) sources, with different agents of transportation. Sorting indicates variations in kinetic energy, such as a decreasing velocity of the transporting agent [55]. The sediment sorting wide ranges point to turbulent conditions [53]. These sediments have negatively skewed values, indicating a high energy environment [56]. Kurtosis suggests the characteristics of deposition medium flow. Most of the studied sites’ sediment were lepto- and very leptokurtic in their nature, which suggests that the sediment is mainly sand with minor populations of other grain sizes.

3.2. PTE Distribution

Geochemical analyses were carried out with surface sediment samples of the Egyptian Red Sea coast. Their analytical values of PTEs and the background values of the upper continental crust [57] and average global sediment [58] are listed in Table 3. Their distribution was determined using the proportional/graduated symbols method and is shown in Figure 2. PTE distributions in the Red Sea sediments indicated a wide variation between not only the considered sites but also the metals themselves. The recorded maximum value of the investigated PTEs seemed to be 21, 18, 6, 11, 16, 8, 12, 9, and 4 orders higher than the corresponding minimum value of Cd, Co, Cr, Cu, Fe, Mn, Ni, Pb, and Zn, respectively, suggesting the heterogeneity of their distribution. The results indicate that mean concentrations were in the order of Fe: 2923.85 ppm > Mn: 145.85 ppm > Zn: 29.10 ppm > Cr: 21.47 ppm > Ni: 18.42 ppm > Co: 4.45 ppm > Pb: 4.06 ppm > Cu: 2.51 ppm > Cd: 0.16 ppm. Fe and Mn are regarded as major elements in sediments because of their natural abundance, and their concentration is often higher than that of other metals/metalloids.
The recorded values are comparable with the background values. In the Red Sea sediments, the mean concentrations of all PTEs were lower than those generally found in sediments globally [58]; however, the sites affected by intensive human activities showed concentrations of Cd, Co, Cr and Ni higher than those of the upper continental crust concentrations [57] (Table 3). This indicates that Red Sea sediments are dominated by a large amount of natural sediment with a low PTE content and are slightly affected by human activities.
Elements such as Fe and Mn possess a lithogenic origin. Fe and Mn mean concentrations in the Red Sea sediments are much lower than upper continental crust [57] and global sediments [58], most likely due to their leaching during weathering and the close association within the geochemical cycle. The lower abundances of Fe and Mn at all locations relative to the crustal average could be attributed to their increased attraction to the dissolved phases due to mineral ion exchange with salt water [59]. Cr, Cu, Ni, Pb, and Zn, which are naturally or anthropogenically mobilized elements, are associated with sedimentary phases such as carbonate, exchangeable Fe/Mn oxyhydroxides, and organic matter [60,61]. Accumulation effects in the deep seabed associated with mud particles, the initial PTEs contents of rock and parent materials, different weathering processes, and varying sources of anthropogenic input at specific sampling sites are some of the key factors impacting the relative abundance of PTEs in marine sediment [12,14,62]. Moreover, a significant source of PTEs is maritime transport that is linked to the density of traffic in the marine environment [7]. Cu and Pb are often used as active ingredients in antifouling paints used on ships, whereas Zn is commonly used as a binding agent in paint formulation [20,63].
Zn, Cu, Ni, Cr, and Co distributions have the same trend in all the studied sites (Figure 2). Thus, they increase in the middle areas and decrease both northwards and southwards, while Pb and Co distributions have no clear trends. Various affinities of PTEs for sediments and their mixing sources could explain the different distribution patterns found among the measured PTEs in the Red Sea sediments.
In this study, PTE concentrations in sediment samples from Egypt’s Red Sea Coast were compared to up-to-date data from other similar areas in Egypt and throughout the world (Table 4). The findings suggest that the PTE concentrations in the study area are comparable to those in the other regions listed. The mean values of Co, Cr, Ni, and Zn are exhibited to be higher than their corresponding values reported in earlier studies in Egypt, except for Co and Cr, which were lower than those reported by Badawy et al. [4] in the Red Sea sediments, and for Zn in the Mediterranean Sea Coast sediments reported by Badawy et al. [64]. On the other hand, Cd, Cu, Fe, Mn, and Pb mean values are lower than those reported in most areas on the Red Sea coast and Mediterranean Sea coast in Egypt.
Overall, the PTE concentrations evaluated in this study revealed significant variability compared to the recent literature regarding the coastal locations shown in Table 4. The mean concentration of Cd in the Red Sea sediments is lower than all other locations in the world except Daya Bay (China) [70], Ha Long Bay (Vietnam) [9], the West Coast of Malaysia [8] and Asturias (Spain) [72]. The Co mean concentration in this study is higher than that in other countries except for the Gulf of Aqaba (Saudi Arabia) [68] and Diu coast (India) [71]. The Cr mean concentration in the present study is higher than in some locations such as Jeddah and Ras Tanura (Saudi Arabia) [15,67], the Puducherry coast (India) [71], Asaluyeh port (Iran) [59], Nigeria [14], and Asturias (Spain) [72]. The Cu and Fe are lower than all other locations in the world except Nigeria [14] and the Puducherry coast (India) [71]. The mean concentration of Ni is higher than all other locations in the world except Daya Bay (China) [70], Diu coast (India) [71], Asaluyeh port (Iran) [59], and Black Sea (Turkey) [11]. The mean concentration of Pb in the Red Sea sediments is lower than all other locations in the world except Ras Tanura (Saudi Arabia) [15], Puducherry coast (India) [71], Asaluyeh port (Iran) [59], and Nigeria [14]. The Zn mean concentration in the present study is higher than some locations such as Jeddah, Ras Tanura, and Gulf of Aqaba (Saudi Arabia) [15,67,68], Puducherry and Diu coast (India) [71], Asaluyeh port (Iran) [59], Nigeria [14], Asturias (Spain) [72], and Ivorian coastal zone [12].

3.3. Statistical Analyses

The K-S test was used to evaluate the normality of the PTE distributions, and the distribution was considered normal (not showing a significant statistical difference) if the p-value was >0.05 [3,11,71]. As noted from the K–S test (p-value) (Table 3), the measured PTEs in the Red Sea marine sediment showed a normal distribution and did not show any significant statistical difference (p-value ranged from 0.20 to 1.00 > 0.05). The close relationship between the mean and standard deviation of Cd, Co, Cr, Cr, Ni, and Pb, as well as the high value of the coefficient of variation (CV), indicates the presence of temporal and spatial changes likely induced by pollution sources (Table 3). Natural resources are denominated in the inputs of metals with a low CV value, whereas human-induced activities are denominated in the inputs of metals with a considerable CV value [11]. In contrast, the CV values for Zn (33%) indicate that this element’s spatial distribution was more uniform among different stations.
Pearson’s correlation coefficients were calculated to interpret the relationship among the studied PTEs (Table 5). Correlations values of (0.20–0.39), (0.40–0.59), (0.60–0.79), and (0.80–1.00) can be considered as weak, moderate, strong, and very strong correlations; respectively [35,73,74]. The results specify that Fe has a strong positive correlation with Co (r = 0.679; p-value = 0.005), Mn (r = 0.736; p-value = 0.002), Ni (r = 0.622; p-value = 0.013), and Zn (r = 0.647; p-value = 0.009), and a moderate positive correlation with Cr (r = 0.593; p-value = 0.02) and Cu (r = 0.570; p-value = 0.027). Likewise, Mn has a strong positive correlation with Cr (r = 0.710; p-value = 0.003) and Ni (r = 0.798; p-value = 0.0003), and a moderate positive correlation with Co (r = 0.548; p-value = 0.035) and Cu (r = 0.487; p-value = 0.065). Additionally, Zn has a moderate positive correlation with Co (r = 0.459; p-value = 0.086), Cr (r = 0.479; p-value = 0.071), and Pb (r = 0.555; p-value = 0.032). Moreover, Cu has a very strong positive correlation with Co (r = 0.904; p-value = 0.000003), a strong positive correlation with Ni (r = 0.715; p-value = 0.003), and a moderate positive correlation with Cr (r = 0.459; p-value = 0.085). Additionally, Ni was strongly positively correlated with Co (r = 0.707; p-value = 0.003) and Cr (r = 0.792; p-value = 0.0004), and Cr was moderately positively correlated with Cd (r = 0.457; p-value = 0.087) and Co (r = 0.584; p-value = 0.022).
These various positive correlations may suggest a shared source or mode of regulation by similar geochemical processes that properly control their evolution in the environment. Under certain physicochemical conditions, this significant positive correlation indicates that PTEs can share sources and comparable behaviors during transportation, transformation, and migration [12,51]. The direct linear correlation between Co, Cr, Cu, Ni, Fe and Mn indicates the considerable influence of the scavenging effect of Fe and Mn on the distribution of other elements [43,60,75]. Fe/Mn oxides and oxyhydroxides would lead to the coprecipitation of PTEs from the water column, resulting in elevated element concentrations in sediments [61]. Ni has a strong ability to bond to metals, particularly sulphides to Fe (pyrite). In addition, Mn and Ni combine to form a less-mobile species that settles to the bottom sediments [59,76]. Strong and moderate significant correlations among the studied elements indicate a close distribution and association, which is in line with the findings of Nour and El-Sorogy [23], El Nemr et al. [29], Palleyi et al. [52], Zhang et al. [77], and Oura et al. [12].
On the other hand, there was no observed association between Cd and Pb and the other elements, except that between Cd-Cr and Pb-Zn. As a result, Cd and Pb concentrations were unaffected by a sole factor. However, they could be influenced by various natural and anthropogenic activities and their potential existence may have been affected by many human activities. Furthermore, only weak positive correlations were noted between mud content and Cd, Cr, and Zn, indicating that granulometry is not the dominant factor that controls the distribution of PTEs into Egyptian Red Sea surficial sediments. Jones et al. [78] observed a lack of strong positive association between fine-sized sediments and PTEs in Kembla Harbour, New South Wales, Australia. They hypothesized that PTEs are also associated with the coarser components, such as fugitive metal-bearing ore particles, marine invertebrate faecal pellets, concentrated in Fe oxyhydroxide coatings on sand-sized particles or affiliated with coarse diagenetic pyrite.
In the present study, PCA (Figure 3) reduced data into three components, and clearly shows that the eigenvalues of the components were 82.03% in the extracted data. The first component (PC1), with an eigenvalue of 4.72, accounted for 52.53% of the total variance and was highly loaded with all of the studied PTEs. The second component (PC2) had an eigenvalue of 1.53, accounted for 17.02% of the total variance and was positively loaded with Cd, Cr, Fe, Pb, and Zn and negatively loaded with Co, Cu, Mn, and Ni. Finally, the third component (PC3) had an eigenvalue of 1.12, accounted for 12.48% of the total variance and was positively loaded with Co, Cu, Fe, Pb, and Zn and negatively loaded with Cd, Cr, Mn, and Ni.
CA was applied in the R-Mode (variables) and Q-Mode (observations) to produce a dendrogram. The R-Mode dendrogram confirms the deduced association between the measured element (Figure 4a). The Q-Mode dendrogram (Figure 4b) contains two main clusters, subdivided, in turn, into two sub-clusters (4 sub-clusters). The first main cluster includes (a) sites 2, 4, 10, 11, 13, 14, and 15 (represent the tourist activity), and (b) sites 3, 5, and 8 (represent commercial and tourist port activity), while the second cluster was subdivided into (c) sites 1 and 7 (represent phosphate and petroleum industrial activity), and (d) sites 6, 9 and 12 (represent different activities). Generally, the PCA and CA results were strongly matched with the results of Pearson’s correlation.

3.4. Pollution Assessment

The first group of indices was used to assess the contamination level for each measured PTE separately (including Igeo and Cf). The Igeo results (Table S2; Figure 5) of all studied PTEs were found between 0 and 1, which classified them into class (1) [45], suggesting uncontaminated to moderately contaminated sediments at all of the studied sites (Table S2; Figure 5). The Cf results indicated that most of the studied sites were within the low contamination degree category (unpolluted sediment with Cf < the unite) [46], which was in parallel with the trend in Igeo values. Some sites recorded Cf values for Cd, Co, Cr, and Ni of between 1 and 3 and were classified as moderately polluted sediment. However, only at sites 5 (Gemsha) and 7 (El-Hamrawein) were cadmium levels within the category of a considerable contamination factor (Cf values were 3.627 and 4.118, respectively) (Table S3; Figure 5). The highest Igeo and Cf values were recorded for Cd in the Gemsha, North Safaga, and El-Hamrawein sites, which are affected by petroleum and phosphate industries.
The present results of Cf-based E i r indicate that most of the studied site’s sediments are within the low-risk category ( E i r < 40), as a consequence of the toxicity response factors of the toxic elements [46]. The calculated E i r of Cd shows a moderate risk (Table S4; Figure 5). The results indicate that the measured PTEs were probably a result of some natural sources. These results agreed with those reported by Badawy et al. [4] and El Nemr et al. [29].
The calculated Cf-based Cdeg values proposed that the studied sites had a low and moderate degree of contamination [46] (Table 6). To evaluate the degree of the severity, the overall ecological risk index (RI), contamination security index (CSl), and Mean EMR quotient (MERMQ) were calculated (Table 6). Results of RI show a wide range of low to high ecological risks. The CSI results indicated that the studied sites were within an uncontaminated to very low degree of severity (CSI < 0.5; 0.5 > CSI < 1, respectively) [42]. MERMQ of the middle area indicates a medium ecological risk level with a 21% probability of toxicity (0.1 > MERMQ < 0.5). The recorded values were 0.17, 0.18, 0.35, 0.10, and 0.33 for sites 7, 8, 9, 10 and 12, respectively. Meanwhile, the rest of the studied sites were within a low ecological risk level with a 9% probability of toxicity (MERMQ < 0.1) (Table 6). The distribution pattern maps of Cdeg, RI, CSI, and MERMQ, in the Red Sea coast sediment samples are shown in Figure 6. Despite the degrees of assessment for these four indicators not being entirely analogous, their distribution pattern maps clearly show that the slightly high degree of contamination and risk is concentrated in the middle of the study area around phosphate mining and related activities. The results of the ecological risks indices of E i r , RI, CSI, and MERMQ were in line with those published by Badawy et al. [4] and El Nemr et al. [29] on the Egyptian Red Sea coast. However, the previously published results may vary slightly due to the variation in sampling sites and different analytical methods used. In addition, the results of indices in the present paper are in agreement with statistical analyses, thereby indicating that the studied PTEs are generally of natural origin, and their concentrations were enhanced in some sites by human activities.
The economic importance of the Red Sea region in Egypt’s strategic planning has increased dramatically in the last few years. The region has become an attractive area for international investments in several fields, including the development of seaports and transportation hubs, mining industries, and tourism. The possible expansion in these activities may be accompanied by a continuous increase in the amount of hazardous waste materials and the concomitant environmental hazards, necessitating the periodic assessment and monitoring of PTE concentrations.

4. Conclusions

In recent decades, PTE pollutants have become a serious environmental concern in the aquatic ecosystem. The granulometric analysis demonstrated a diversity of facies represented by gravelly sand, muddy gravelly sand, muddy sand, and gravelly muddy sand. The Red Sea coastal sediment is composed mainly of fine to very-fine sand. There were two main sediment source, namely biogenic (allochthonous) and terrigenous (autochthonous), with different agents of transportation. Wide ranges of sediment sorting values indicate turbulent conditions. The results showed that Fe and Mn were the most abundant elements in all investigated samples. The mean concentrations of all PTEs were lower than those usually found in sediments globally. Few sites showed concentrations of Cd, Co, Cr, and Ni of higher than those of the upper continental crust. A strong significant correlation among the studied elements indicates their close distribution and association. The direct linear correlation between Co, Cr, Cu, Ni, and Fe and Mn indicates the considerable influence of the scavenging effect of Fe and Mn on the distribution of other elements. Cd and Pb concentrations were unaffected by a sole factor and could be influenced by a variety of natural and anthropogenic activities. According to the computed pollution indices, Red Sea sediment shows minimal to moderate pollution levels. The highest Igeo and Cf values were recorded for Cd sites affected by petroleum and phosphate industries. The distribution pattern maps of Cdeg, RI, CSI, and MERMQ clearly show that the slightly high degree of contamination and risk is concentrated in the middle of the study area around phosphate mining and related activities. The results, including correlations and pollution indices, suggest that the PTE contents in the Red Sea sediments are of natural origin and are slightly affected by anthropogenic sources, especially petroleum and phosphate mining and related activities. More attention should be paid to Cd, Ni, and Pb concentrations as the main pollution factors. Finally, the findings of this study will be relevant and informative for future research and economic development because they provide updated data on PTE contamination levels in Red Sea marine sediments. In addition, the present results may be considered as a good base for establishing local guidelines. PTE concentrations in the Red Sea coastline must be adequately monitored on a regular basis to avoid any environmental hazards.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/land11091560/s1, Table S1: Categories of single and integrated pollution indices, References [42,45,46,50]; Table S2: Geo-accumulation index (Igeo) values; Table S3: Contamination factor (Cf) values; Table S4: Ecological risk factors ( E i r ) values.

Author Contributions

Conceptualization, H.I.F., A.G. and S.M.A.E.B.; methodology, H.I.F., A.G. and S.M.A.E.B.; software, H.I.F. and A.G.; validation, H.I.F. and A.G.; formal analysis, H.I.F., A.G. and S.M.A.E.B.; investigation, H.I.F., A.G. and S.M.A.E.B.; resources, H.I.F., A.G. and A.S.; data curation, H.I.F., A.G. and S.M.A.E.B.; writing—original draft preparation, H.I.F., A.G. and S.M.A.E.B.; writing—review and editing, H.I.F., A.G. and A.S.; supervision, H.I.F. and A.G.; project administration, A.S.; funding acquisition, A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This paper is part of the work supported by the Science, Technology & Innovation Funding Authority (STDF), Egypt, Under Grant Number: 30116 and 37233.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This paper is part of the work supported by Science, Technology & Innovation Funding Authority (STDF), Egypt, under Grant Number: 30116 and 37233. The authors would like to acknowledge the laboratory facilities offered by the Geology Department, Faculty of Science, Suez University, Egypt. The authors thank El Sayed O. (Ain Shams University, Cairo, Egypt) for providing Landsat data and helping in the application of the GIS technique. Thanks are also extended to Smillie, Z. (University of Stirling, Stirling, UK) for her valuable suggestions and proofreading of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Abu El-Magd, S.A.; Taha, T.H.; Pienaar, H.H.; Breil, P.; Amer, R.A.; Namour, P. Assessing heavy metal pollution hazard in sediments of Lake Mariout, Egypt. J. Afr. Earth Sci. 2021, 176, 104116. [Google Scholar] [CrossRef]
  2. Calmuc, V.A.; Calmuc, M.; Arseni, M.; Topa, C.M.; Timofti, M.; Burada, A.; Iticescu, C.; Georgescu, L.P. Assessment of Heavy Metal Pollution Levels in Sediments and of Ecological Risk by Quality Indices, Applying a Case Study: The Lower Danube River, Romania. Water 2021, 13, 1801. [Google Scholar] [CrossRef]
  3. Doufexi, M.; Gamvroula, D.E.; Alexakis, D.E. Elements’ Content in Stream Sediment and Wildfire Ash of Suburban Areas in West Attica (Greece). Water 2022, 14, 310. [Google Scholar] [CrossRef]
  4. Badawy, W.M.; El-Taher, A.; Frontasyeva, M.V.; Madkour, H.A.; Khater, A.E. Assessment of anthropogenic and geogenic impacts on marine sediments along the coastal areas of Egyptian Red Sea. Appl. Radiat. Isot. 2018, 140, 314–326. [Google Scholar] [CrossRef] [PubMed]
  5. Farhat, H.I. Impact of Drain Effluent on Surficial Sediments in the Mediterranean Coastal Wetland: Sedimentological Characteristics and Metal Pollution Status at Lake Manzala, Egypt. J. Ocean Univ. China 2019, 18, 834–848. [Google Scholar] [CrossRef]
  6. Osman, R.; Dawood, Y.H.; Melegy, A.; El-Bady, M.S.; Saleh, A.; Gad, A. Distributions and Risk Assessment of the Natural Radionuclides in the Soil of Shoubra El Kheima, South Nile Delta, Egypt. Atmosphere 2022, 13, 98. [Google Scholar] [CrossRef]
  7. Christophoridis, C.; Bourliva, A.; Evgenakis, E.; Papadopoulou, L.; Fytianos, K. Effects of anthropogenic activities on the levels of heavy metals in marine surface sediments of the Thessaloniki Bay, Northern Greece: Spatial distribution, sources and contamination assessment. Microchem. J. 2019, 149, 104001. [Google Scholar] [CrossRef]
  8. Looi, L.J.; Aris, A.Z.; Yusoff, F.M.; Isa, N.M.; Haris, H. Application of enrichment factor, geoaccumulation index, and ecological risk index in assessing the elemental pollution status of surface sediments. Environ. Geochem. Health 2019, 41, 27–42. [Google Scholar] [CrossRef]
  9. Hoai, N.D.; Manh, H.N.; Duc, T.T.; Cong, T.D.; Dinh, L.T.; Johnstone, R.; Kim, D.N. An assessment of heavy metal contamination in the surface sediments of Ha Long Bay, Vietnam. Environ. Earth Sci. 2020, 79, 436. [Google Scholar] [CrossRef]
  10. Radomirović, M.; Tanaskovski, B.; Pezo, L.; Ceccotto, F.; Cantaluppi, C.; Onjia, A.; Stanković, S. Spatial and temporal distribution of pollution indices in marine surface sediments—A chemometric approach. Environ. Sci. Pollut. Res. 2021, 28, 42496–42515. [Google Scholar] [CrossRef]
  11. Apaydın, A.; Kabaoğlu, H.; Apaydın, G.; Şirin, M.; Cengiz, E.; Köksal, O.K.; Baltaş, H.; Tıraşoğlu, E. Evaluation of ecological risk, source, and spatial distribution of some heavy metals in marine sediments in the Middle and Eastern Black Sea region, Turkey. Environ. Sci. Pollut. Res. 2022, 29, 7053–7066. [Google Scholar] [CrossRef] [PubMed]
  12. Oura, L.E.; Kouassi, K.E.; Konan, A.T.S.; Koné, H.; Kouakou, A.R.; Boa, D.; Yao, K.B. Spatial distribution of heavy metals in sediments of the Ivory Coastal zone (Toukouzou Hozalem-Assinie) in correlation with anthropic activities. Chem. Ecol. 2022, 38, 72–94. [Google Scholar] [CrossRef]
  13. Rizzo, A.; De Giosa, F.; Di Leo, A.; Lisco, S.; Moretti, M.; Scardino, G.; Scicchitano, G.; Mastronuzzi, G. Geo-Environmental Characterisation of High Contaminated Coastal Sites: The Analysis of Past Experiences in Taranto (Southern Italy) as a Key for Defining Operational Guidelines. Land 2022, 11, 878. [Google Scholar] [CrossRef]
  14. Tunde, O.L.; Oluwagbenga, A.P. Assessment of heavy metals contamination and sediment quality in Ondo coastal marine area, Nigeria. J. Afr. Earth Sci. 2020, 170, 103903. [Google Scholar] [CrossRef]
  15. El-Taher, A.; Massoud, E.E. Distribution of heavy metal and pollution assessment in environmental samples along the coastal area of Ras Tanura. Int. J. Environ. Anal. Chem. 2021, 1–10. [Google Scholar] [CrossRef]
  16. Tnoumi, A.; Angelone, M.; Armiento, G.; Caprioli, R.; Crovato, C.; De Cassan, M.; Montereali, M.R.; Nardi, E.; Parrella, L.; Proposito, M.; et al. Assessment of Trace Metals in Sediments from Khnifiss Lagoon (Tarfaya, Morocco). Earth 2021, 2, 16–31. [Google Scholar] [CrossRef]
  17. Yang, C.; Yu, G.; Liu, Y.; Shan, B.; Wang, L.; Sun, D.; Huang, Y. Heavy Metal Distribution in Surface Sediments of the Coastal Pearl Bay, South China Sea. Processes 2022, 10, 822. [Google Scholar] [CrossRef]
  18. Kuang, Z.; Gu, Y.; Rao, Y.; Huang, H. Biological Risk Assessment of Heavy Metals in Sediments and Health Risk Assessment in Marine Organisms from Daya Bay, China. J. Mar. Sci. Eng. 2021, 9, 17. [Google Scholar] [CrossRef]
  19. Abdolvand, S.; Kaedinejadesfahani, S. Measurement of Some Heavy Metals in the Sediments and Sole Fish Euryglossa orientalis from the Musa Creek. Thalassas 2022, 38, 675–685. [Google Scholar] [CrossRef]
  20. Lim, Y.C.; Chen, C.F.; Tsai, M.L.; Wu, C.H.; Lin, Y.L.; Wang, M.H.; Albarico, F.P.J.B.; Chen, C.W.; Dong, C.D. Impacts of Fishing Vessels on the Heavy Metal Contamination in Sediments: A Case Study of Qianzhen Fishing Port in Southern Taiwan. Water 2022, 14, 1174. [Google Scholar] [CrossRef]
  21. Abd El Naby, A.E.; Al Menoufy, S.; Gad, A. Benthic foraminiferal assemblages as bio-indicators of metals contamination in sediments, Qarun Lake as a case study, Egypt. J. Afr. Earth Sci. 2018, 139, 96–112. [Google Scholar] [CrossRef]
  22. Zhuang, M.; Sanganyado, E.; Li, P.; Liu, W. Distribution of microbial communities in metal-contaminated nearshore sediment from Eastern Guangdong, China. Environ. Pollut. 2019, 250, 482–492. [Google Scholar] [CrossRef] [PubMed]
  23. Nour, H.E.; El-Sorogy, A.S. Heavy metals contamination in seawater, sediments and seashells of the Gulf of Suez, Egypt. Environ. Earth Sci. 2020, 79, 274. [Google Scholar] [CrossRef]
  24. El-Kahawy, R.; El-Shafeiy, M.; Helal, S.; Aboul-Ela, N.; Abd El-Wahab, M. Benthic ostracods (crustacean) as a nearshore pollution bio-monitor: Examples from the Red Sea Coast of Egypt. Environ. Sci. Pollut. Res. 2021, 28, 31975–31993. [Google Scholar] [CrossRef]
  25. Gad, A.; Abd El Bakey, S.M.; Sakr, S. Concentrations of heavy metals and associated human health risk in unrefined salts of inland hypersaline lakes, Egypt. Int. J. Environ. Anal. Chem. 2022, 102, 1278–1291. [Google Scholar] [CrossRef]
  26. Witkowska, D.; Słowik, J.; Chilicka, K. Heavy Metals and Human Health: Possible Exposure Pathways and the Competition for Protein Binding Sites. Molecules 2021, 26, 6060. [Google Scholar] [CrossRef]
  27. Wang, S.L.; Xu, X.R.; Sun, Y.X.; Liu, J.L.; Li, H.B. Heavy metal pollution in coastal areas of South China: A review. Mar. Pollut. Bull. 2013, 76, 7–15. [Google Scholar] [CrossRef]
  28. Lu, J.; Li, A.; Huang, P. Distribution, sources and contamination assessment of heavy metals in surface sediments of the South Yellow Sea and northern part of the East China Sea. Mar. Pollut. Bull. 2017, 124, 470–479. [Google Scholar] [CrossRef]
  29. El Nemr, A.; El-Said, G.F.; Khaled, A.; Ragab, S. Distribution and ecological risk assessment of some heavy metals in coastal surface sediments along the Red Sea, Egypt. Int. J. Sediment. Res. 2016, 31, 164–172. [Google Scholar] [CrossRef]
  30. Gopal, V.; Krishnamurthy, R.R.; Sreeshma, T.; Chakraborty, P.; Nathan, C.S.; Kalaivanan, R.; Anshu, R.; Magesh, N.S.; Jayaprakash, M. Effect of a tropical cyclone on the distribution of heavy metals in the marine sediments off Kameswaram, Southeast coast of India. Mar. Pollut. Bull. 2021, 171, 112741. [Google Scholar] [CrossRef]
  31. Li, J.; Zuo, Q.; Feng, F.; Jia, H. Occurrence and Ecological Risk Assessment of Heavy Metals from Wuliangsuhai Lake, Yellow River Basin, China. Water 2022, 14, 1264. [Google Scholar] [CrossRef]
  32. Darwish, D.H.; Basiony, A.I.; El-Alfy, M.A. Mapping heavy metals contamination and eco-risk along Mediterranean Sea coast, Egypt. Geocarto Int. 2022, 1–27. [Google Scholar] [CrossRef]
  33. Liang, J.; Liu, J.; Yuan, X.; Zeng, G.; Lai, X.; Li, X.; Wu, H.; Yuan, Y.; Li, F. Spatial and temporal variation of heavy metal risk and source in sediments of Dongting Lake wetland, mid-south China. J. Environ. Sci. Health-Part A 2015, 50, 100–108. [Google Scholar] [CrossRef] [PubMed]
  34. Fahmy, M. Water quality in the Red Sea coastal waters (Egypt): Analysis of spatial and temporal variability. Chem. Ecol. 2010, 19, 67–77. [Google Scholar] [CrossRef]
  35. Al-Mur, B.; Gad, A. Radiation Hazard from Natural Radioactivity in the Marine Sediment of Jeddah Coast, Red Sea, Saudi Arabia. J. Mar. Sci. Eng. 2022, 10, 1145. [Google Scholar] [CrossRef]
  36. IAEA. Soil Sampling for Environmental Contaminants; International Atomic Energy Agency: Vienna, Austria, 2004. [Google Scholar]
  37. Folk, R.L. Petrology of Sedimentary Rocks; Hemphill Pub. Co.: Austin, TX, USA, 1980. [Google Scholar]
  38. Ingram, R.L. Sieve analysis. In Procedures in Sedimentary Petrology; Carver, R.E., Ed.; Jhon Wiley Intersciences: New York, NY, USA, 1971; pp. 49–67. [Google Scholar]
  39. Folk, R.L.; Ward, W.C. Brazos River bar [Texas]; a study in the significance of grain size parameters. J. Sediment. Res. 1957, 27, 3–26. [Google Scholar] [CrossRef]
  40. Loring, D.H.; Rantala, R.T.T. Manual for the geochemical analyses of marine sediments and suspended particulate matter. Earth-Sci. Rev. 1992, 32, 235–283. [Google Scholar] [CrossRef]
  41. Caeiro, S.; Costa, M.H.; Ramos, T.B.; Fernandes, F.; Silveira, N.; Coimbra, A.; Medeiros, G.; Painho, M. Assessing heavy metal contamination in Sado Estuary sediment: An index analysis approach. Ecol. Indic. 2005, 5, 151–169. [Google Scholar] [CrossRef]
  42. Pejman, A.; Bidhendi, G.N.; Ardestani, M.; Saeedi, M.; Baghvand, A. A new index for assessing heavy metals contamination in sediments: A case study. Ecol. Indic. 2015, 58, 365–373. [Google Scholar] [CrossRef]
  43. Saleh, A.; Dawood, Y.H.; Gad, A. Assessment of Potentially Toxic Elements’ Contamination in the Soil of Greater Cairo, Egypt Using Geochemical and Magnetic Attributes. Land 2022, 11, 319. [Google Scholar] [CrossRef]
  44. Baghdady, A.; Awad, S.; Gad, A. Assessment of metal contamination and natural radiation hazards in different soil types near iron ore mines, Bahariya Oasis, Egypt. Arab. J. Geosci. 2018, 11, 506. [Google Scholar] [CrossRef]
  45. Müller, G. Index of geoaccumulation in sediments of the Rhine River. Geo. J. 1969, 2, 108–118. [Google Scholar]
  46. Hakanson, L. An ecological risk index for aquatic pollution control: A sedimentological approach. Water Res. 1980, 14, 975–1001. [Google Scholar] [CrossRef]
  47. Tomlinson, D.C.; Wilson, D.J.; Harris, C.R.; Jeffrey, D.W. Problem in heavy metals in estuaries and the formation of pollution index. Helgol. Meeresunters 1980, 33, 566–575. [Google Scholar] [CrossRef]
  48. Long, E.R.; Macdonald, D.D.; Smith, S.L.; Calder, F.D. Incidence of adverse biological effects within ranges of chemical concentrations in marine and estuarine sediments. Environ. Manag. 1995, 19, 81–97. [Google Scholar] [CrossRef]
  49. Ipeaiyeda, A.R.; Onianwa, P.C. Monitoring and assessment of sediment contamination with toxic heavy metals: Case study of industrial effluent dispersion in Alaro River, Nigeria. Appl. Water Sci. 2018, 8, 161. [Google Scholar] [CrossRef]
  50. Long, E.R.; Ingersoll, C.G.; MacDonald, D.D. Calculation and uses of mean sediment quality guideline quotients: A critical review. Environ. Sci. Technol. 2006, 40, 1726–1736. [Google Scholar] [CrossRef]
  51. Ibrahim, M.I.A.; Mohamed, L.A.; Mahmoud, M.G.; Shaban, K.S.; Fahmy, M.A.; Ebeid, M.H. Potential ecological hazards assessment and prediction of sediment heavy metals pollution along the Gulf of Suez, Egypt. Egypt. J. Aquat. Res. 2019, 45, 329–335. [Google Scholar] [CrossRef]
  52. Palleyi, S.; Banoo, S.; Kar, R.N.; Panda, C.R. Textural and geochemical characteristics of off shore sediment of North Bay of Bengal: A statistical approach for marine metal pollution. Int. J. Sediment. Res. 2015, 30, 208–222. [Google Scholar] [CrossRef]
  53. Farhat, H.I. Textural features and transportation mode of Nile Delta coastal lagoon surficial sediments (Lake Burullus, Ramsar Site). SN Appl. Sci. 2019, 1, 1013. [Google Scholar] [CrossRef]
  54. Farhat, H.I.; Aly, W. Effect of site on sedimentological characteristics and metal pollution in two semi-enclosed embayments of great freshwater reservoir: Lake Nasser, Egypt. J. Afr. Earth Sci. 2018, 141, 194–206. [Google Scholar] [CrossRef]
  55. Sahu, B.K. Depositional Mechanisms from the Size Analysis of Clastic Sediments. J. Sediment. Petrol. 1964, 34, 73–83. [Google Scholar] [CrossRef]
  56. Duane, D.B. Significance of skewness in recent sediments, western pamlico sound, North California. J. Sediment. Petrol. 1964, 34, 864–874. [Google Scholar] [CrossRef]
  57. Wedepohl, K.H. The composition of the continental crust. Geochim. Cosmochim. Acta 1995, 59, 1217–1239. [Google Scholar] [CrossRef]
  58. Bowen, H.M. Environmental Chemistry of the Elements; Academic Press: London, UK, 1979. [Google Scholar]
  59. Delshab, H.; Farshchi, P.; Keshavarzi, B. Geochemical distribution, fractionation and contamination assessment of heavy metals in marine sediments of the Asaluyeh port, Persian Gulf. Mar. Pollut. Bull. 2017, 115, 401–411. [Google Scholar] [CrossRef]
  60. Tessier, A.; Campbell, P.G.C.; Bisson, M. Sequential extraction procedure for the speciation of particulate trace metals. Anal. Chem. 1979, 51, 844–851. [Google Scholar] [CrossRef]
  61. Anbuselvan, N.; Senthil, D.N.; Sridharan, M. Heavy metal assessment in surface sediments off Coromandel Coast of India: Implication on marine pollution. Mar. Pollut. Bull. 2018, 131, 712–726. [Google Scholar] [CrossRef]
  62. Zhao, L.; Mi, D.; Chen, Y.; Wang, L.; Sun, Y. Ecological risk assessment and sources of heavy metals in sediment from Daling River basin. Environ. Sci. Pollut. Res. 2015, 22, 5975–5984. [Google Scholar] [CrossRef]
  63. Jeon, S.k. Assessment of metal contamination in sediments induced by small ships around lighters wharf in Southwest Korea. Environ. Earth Sci. 2021, 80, 627. [Google Scholar] [CrossRef]
  64. Badawy, W.; Elsenbawy, A.; Dmitriev, A.; El Samman, H.; Shcheglov, A.; El-Gamal, A.; Kamel, N.H.M.; Mekewi, M. Characterization of major and trace elements in coastal sediments along the Egyptian Mediterranean Sea. Mar. Pollut. Bull. 2022, 177, 113526. [Google Scholar] [CrossRef]
  65. Soliman, N.; Thabet, W.M.; El-Sadaawy, M.M.; Morsy, F.A.M. Contamination Status and Potential Risk of Metals in Marine Sediments of Shalateen Coast, the Red Sea. Soil. Sediment. Contam. 2022, 31, 40–56. [Google Scholar] [CrossRef]
  66. Nour, H.E.; El-Sorogy, A.S.; Abdel-Wahab, M.; Almadani, S.; Alfaifi, H.; Youssef, M. Assessment of sediment quality using different pollution indicators and statistical analyses, Hurghada area, Red Sea coast, Egypt. Mar. Pollut. Bull. 2018, 133, 808–813. [Google Scholar] [CrossRef] [PubMed]
  67. Halawani, R.F.; Wilson, M.E.; Hamilton, K.M.; Aloufi, F.A.; Taleb, M.A.; Al-Zubieri, A.G.; Quicksall, A.N. Spatial Distribution of Heavy Metals in Near-Shore Marine Sediments of the Jeddah, Saudi Arabia Region: Enrichment and Associated Risk Indices. J. Mar. Sci. Eng. 2022, 10, 614. [Google Scholar] [CrossRef]
  68. El-Sorogy, A.S.; Youssef, M.; Al-Kahtany, K.; Saleh, M.M. Distribution, Source, Contamination, and Ecological Risk Status of Heavy Metals in the Red Sea-Gulf of Aqaba Coastal Sediments, Saudi Arabia. Mar. Pollut. Bull. 2020, 158, 111411. [Google Scholar] [CrossRef] [PubMed]
  69. Al-Edresy, M.A.M.; Wasel, S.O.; Al-Hagibi, H.A. Ecological risk assessment of heavy metals in coastal sediments between Al-Haymah and Al-Mokha, south red sea, Yemen. Int. J. Hydrol. 2019, 3, 159–173. [Google Scholar] [CrossRef]
  70. Tang, H.; Ke, Z.; Yan, M.; Wang, W.; Nie, H.; Li, B.; Zhang, J.; Xu, X.; Wang, J. Concentrations, Distribution, and Ecological Risk Assessment of Heavy Metals in Daya Bay, China. Water 2018, 10, 780. [Google Scholar] [CrossRef]
  71. Jha, D.K.; Dharani, G.; Verma, P.; Ratnam, K.; Kumar, R.S.; Rajaguru, S. Evaluation of factors influencing the trace metals in Puducherry and Diu coasts of India through multivariate techniques. Mar. Pollut. Bull. 2021, 167, 112342. [Google Scholar] [CrossRef]
  72. Sanz-Prada, L.; Garcia-Ordiales, E.; Flor-Blanco, G.; Roqueñí, N.; Álvarez, R. Determination of heavy metal baseline levels and threshold values on marine sediments in the Bay of Biscay. J. Environ. Manag. 2022, 303, 114250. [Google Scholar] [CrossRef]
  73. Evans, J.D. Straight Forward Statistics for the Behavioral Sciences; Brook/Cole Publishing: Pacific Grove, CA, USA, 1996. [Google Scholar]
  74. Gad, A.; Saleh, A.; Farhat, H.I.; Dawood, Y.H.; Abd El Bakey, S.M. Spatial Distribution, Contamination Levels, and Health Risk Assessment of Potentially Toxic Elements in Household Dust in Cairo City, Egypt. Toxics 2022, 10, 466. [Google Scholar] [CrossRef]
  75. Khalifa, M.; Gad, A. Assessment of heavy metals contamination in agricultural soil of southwestern Nile Delta, Egypt. Soil. Sediment. Contam. 2018, 27, 619–642. [Google Scholar] [CrossRef]
  76. Laxen, D.P.; Harrison, R.M. Physico-chemical speciation of selected metals in the treated effluent of a lead-acid battery manufacturer and in the receiving river. Water Res. 1983, 17, 71–80. [Google Scholar] [CrossRef]
  77. Zhang, M.; He, P.; Qiao, G.; Huang, J.; Yuan, X.; Li, Q. Heavy metal contamination assessment of surface sediments of the Subei Shoal, China: Spatial distribution, source apportionment and ecological risk. Chemosphere 2019, 223, 211–222. [Google Scholar] [CrossRef] [PubMed]
  78. Jones, B.G.; Alyazichi, Y.M.; Low, C.; Goodfellow, A.; Chenhall, B.E.; Morrison, R.J. Distribution and sources of trace element pollutants in the sediments of the industrialised Port Kembla Harbour, New South Wales, Australia. Environ. Earth Sci. 2019, 78, 357. [Google Scholar] [CrossRef]
Figure 1. Map displaying the study area and sampling sites.
Figure 1. Map displaying the study area and sampling sites.
Land 11 01560 g001
Figure 2. The distribution pattern of PTEs (As illustrated in legend, circles with varying radiuses represent different concentration classes).
Figure 2. The distribution pattern of PTEs (As illustrated in legend, circles with varying radiuses represent different concentration classes).
Land 11 01560 g002
Figure 5. Boxplots of Igeo, Cf, and E i r values.
Figure 5. Boxplots of Igeo, Cf, and E i r values.
Land 11 01560 g005
Figure 6. Distribution pattern of Cdeg, RI, CSI, and MERMQ values.
Figure 6. Distribution pattern of Cdeg, RI, CSI, and MERMQ values.
Land 11 01560 g006
Figure 3. R-mode PCA ordination diagram: (a) PC1 vs. PC2; (b) PC1 vs. PC3; (c) 3D variables loading.
Figure 3. R-mode PCA ordination diagram: (a) PC1 vs. PC2; (b) PC1 vs. PC3; (c) 3D variables loading.
Land 11 01560 g003
Figure 4. CA dendrogram (a) R-mode (Variables); (b) Q-mode (Sampling sites).
Figure 4. CA dendrogram (a) R-mode (Variables); (b) Q-mode (Sampling sites).
Land 11 01560 g004
Table 1. Location and contributing sources of pollution of Red Sea sampling sites.
Table 1. Location and contributing sources of pollution of Red Sea sampling sites.
Site No.CoordinatesSample SitesDepth (m)Contributing Sources of Pollution
LatitudeLongitude
127°38′15.11″33°31′ 26.28″Gemsha8.0Petroleum exploration, sulphur mining, fishing port
227°26′13.95″33°39′51.05″North Hurghada6.5Tourism
327°10′33.53″33°50′34.32″Hurghada6.0Urban, tourism, fishing port
427°06′50.21″33°49′54.79″South Hurghada11.0Tourism
526°45′26.39″33°57′18.23″North Safaga9.0Urban, commercial harbor
626°33′47.67″34°02′24.15″South Safaga8.0Urban, phosphate seaport
726°15′12.01″34°12′04.66″El-Hamrawein9.0Urban, phosphate seaport
826°10′19.84″34°15′42.15″Qusier14.0Urban, phosphate mining, fishing port
925°40′08.02″34°34′10.28″N. Marsa Umm El-Griufat6.0Tourism, mangrove ecosystem
1025°36′01.69″34°36′48.74″N. Sharm El Bahria5.0Tourism, mangrove ecosystem
1125°31′39.23″34°38′40.02″Abu Dabab4.0Tourism
1225°28′23.90″34°40′4821″Sharm El Naaba6.0Tourism
1325°03′45.50″34°54′07.20″Marsa Alam10.0Urban, tourism, fishing port
1424°11′57.44″35°25′56.30″Wadi Lahmi5.0Tourism
1523°08′13.99″35°40′17.67″Marsa Humira8.0Tourism
Table 2. Grain size data, textural nomenclature, and Statistical grain size parameters of the studied sediments.
Table 2. Grain size data, textural nomenclature, and Statistical grain size parameters of the studied sediments.
Sample
No.
Gravel
%
Sand %Mud %NomenclatureMean SizeSortingSkewnessKurtosis
VC
Sand
C
Sand
M
Sand
F
Sand
VF
Sand
Total
Sand
SiltClay
13.57.011.431.331.515.496.50.00.0GSF SandPSNSLK
218.76.26.013.050.25.981.30.00.0GSM SandPSSCSPK
33.71.52.746.737.18.496.30.00.0GSM SandMSNSMK
40.00.00.00.055.323.278.518.33.2MSVF SandPSSFSVLK
51.90.00.00.050.343.794.00.00.0MGSVF SandMWSNSMk
67.61.31.226.059.14.892.40.00.0GSF SandMSSCSVLK
70.00.00.01.16.247.554.841.33.9MSVF SandPSFSLK
80.05.27.523.043.814.994.44.51.2MSF SandPSNSLK
911.35.79.122.029.517.984.10.00.0MGSM SandPSCSMK
109.826.25.818.551.58.290.20.00.0GSM SandPSSCSLK
111.910.00.05.628.363.697.40.00.0GSF SandWSSCSLK
127.10.00.016.547.825.189.40.00.0MGSF SandPSCSVLK
134.740.00.011.033.914.459.425.910.0GMSVF SandVPSSFSPK
148.320.00.041.441.44.887.50.00.0MGSF SandPSCSVLK
156.412.212.016.145.215.390.70.00.0MGSF SandPSSCSLK
VC = Very Coarse, C = Coarse, M = Medium, F = Fine, VF = Very Fine, GS = Gravelly Sand, MS = Muddy Sand, MGS = Muddy Gravelly Sand, GMS = Gravelly Muddy Sand, PS = Poorly sorted, MS = Moderately sorted, MWS = Medium well sorted, WS = Well sorted, VPS = Very poorly sorted, NS = Near symmetrical, SCS = Strongly coarse skewed, SFS = Strongly fine skewed, FS = Fine skewed, CS = Coarse skewed, LK = Lepto-kurtic, PK = Platy-kurtic, MK = Meso-kurtic, VLK = Very lepto-kurtic.
Table 3. Descriptive statistics of PTEs (ppm) in Red Sea sediments.
Table 3. Descriptive statistics of PTEs (ppm) in Red Sea sediments.
Sample
No.
AreaCdCoCrCuFeMnNiPbZn
1Gemsha0.272.0711.741.984266.8278.044.331.0021.74
2North Hurghada0.102.0310.061.061011.0162.846.243.0018.92
3Hurghada0.022.659.891.342190.8847.0911.269.0530.56
4South Hurghada0.061.679.701.871621.4759.659.994.0326.93
5North Safaga0.372.6716.962.662387.05155.0012.117.4342.13
6South Safaga0.176.5621.731.626402.44256.0210.486.4542.10
7El-Hamrawein0.427.9856.472.974972.29177.4029.833.9644.73
8Qusier0.183.6435.830.953342.87277.9532.883.8828.98
9N. Marsa Umm El-Griufat0.096.9737.975.077050.76354.1051.055.9831.94
10N. Sharm El Bahria0.144.9328.943.74652.93188.4722.943.4812.74
11Abu Dabab0.201.0212.860.84501.98108.249.932.2919.74
12Sharm El Naaba0.0718.0830.978.976032.35221.3041.733.0536.93
13Marsa Alam0.051.8310.621.941673.9384.816.822.2328.83
14Wadi Lahmi0.113.0519.421.821304.6549.9316.823.9232.92
15Marsa Humira0.101.678.870.83446.3266.839.921.0817.39
Minimum0.021.028.870.83446.3247.094.331.0012.74
Maximum0.4218.0856.478.977050.76354.1051.059.0544.73
Mean0.164.4521.472.512923.85145.8518.424.0629.10
Standard Deviation0.124.3413.942.132275.5796.6614.202.289.72
Coefficient of Variation (CV) (%)75.1397.4764.9584.9877.8366.2877.0656.3033.41
K-S test (p-value)0.540.300.540.200.580.510.180.321.00
Upper continental crust [57]0.10211.6035.0014.3030,890.00527.0018.6017.0052.00
Average global sediment [58]-14.0072.0033.0041,000.00770.0052.0019.0095.00
Table 4. Comparison of PTEs concentration in marine sediments reported in the various regions of Egypt and worldwide.
Table 4. Comparison of PTEs concentration in marine sediments reported in the various regions of Egypt and worldwide.
LocationNCdCoCrCuFeMnNiPbZnRef.
Egypt
Red Sea Coast150.164.4521.472.512923.85145.8518.424.0629.10Present study
Red Sea Coast32-4.8153.847.7014,562.50291.9415.374.8927.55[4]
Red Sea Coast (Abu Zenima)200.55--5.072384.00-2.8717.3022.40[23]
Red Sea Coast (Ras Gharib, Quseir)181.052.1018.3025.407094.4136.3011.508.5524.00[24]
Red Sea Coast (Shalateen)180.392.5919.174.17--10.193.7625.17[65]
Red Sea Coast (Hurghada)300.141.66-1.26355.4451.951.7442.387.77[66]
Mediterranean Sea Coast (Kafr Elsheikh)120.364.32-2.997284.58132.608.364.8413.52[32]
Mediterranean Sea Coast (North Nile Delta)99-11.70160.00-25,465636.2024.80-42.5[64]
Worldwide
Saudi Arabia (Jeddah)80--9.569.18-36.523.6877.3418.02[67]
Saudi Arabia (Ras Tanura)1532.74 0.53 3.27---1.951.790.60[15]
Saudi Arabia (Gulf of Aqaba)330.914.539303374184146.624[68]
Yemen (Red Sea)370.76--17.34--8.986.4736.81[69]
China (Daya Bay)190.09-108.724.1--26.835.3108.9[70]
China (Pearl Bay)200.55-35.7824.16---31.2848.53[17]
India (Puducherry coast)720.42.88.12.525209232.44.3[71]
India (Diu coast)721.41324.22715,290502254.225.3[71]
Vietnam (Ha Long Bay)480.08--14.53---30.3950.87[9]
Malaysia (West Coast)480.114.0627.5816.96--14.2121.3454.65[8]
Iran (Asaluyeh port)48-2.2016.1315.434800168.7119.043.3921.08[59]
Nigeria (Abereke)240.55-0.522.42369.34-7.350.6714.33[14]
Nigeria (Awoye)240.85-0.741.28286.17-7.842.0725.15[14]
Nigeria (Ayetoro)240.32-0.241.79417.50-3.730.082.44[14]
Spain (Asturias)150.06-9.784.15--7.806.3220.91[72]
Ivorian coastal zone723.08-43137.5719,144249.1217.378.3928.82[12]
Turkey (Black Sea)16--87.31104.0646,000565.3834.3832.31109.88[11]
Morocco (Khnifiss Lagoon)60.163.5726.66.60--16.506.1351.70[16]
Table 5. Pearson’s correlation coefficient for all variables (n = 15).
Table 5. Pearson’s correlation coefficient for all variables (n = 15).
Sand %Mud %CdCoCrCuFeMnNiPbZn
Gravel %0.055−0.420−0.4030.108−0.1350.161−0.0110.0490.062−0.065−0.320
(p-value)0.8470.1190.1360.7020.6300.5660.9680.8610.8250.8180.245
Sand % −0.921−0.128−0.089−0.367−0.074−0.0790.053−0.0900.177−0.282
(p-value) 0.0000010.6500.7520.1790.7940.7810.8510.7490.5280.309
Mud % 0.2540.0010.355−0.0510.047−0.0930.00009−0.1510.323
(p-value) 0.3600.9980.1940.8570.8680.7401.0000.5900.240
Cd −0.0090.457−0.0820.1900.167−0.0030.0050.399
(p-value) 0.9750.0870.7710.4980.5520.9920.9870.141
Co 0.5840.9040.6790.5480.7070.0720.459
(p-value) 0.0220.0000030.0050.0350.0030.8000.086
Cr 0.4590.5930.7100.7920.1220.479
(p-value) 0.0850.0200.0030.00040.6640.071
Cu 0.5700.4870.7150.0400.298
(p-value) 0.0270.0650.0030.8870.281
Fe 0.7360.6220.2860.647
(p-value) 0.0020.0130.3010.009
Mn 0.7980.2590.367
(p-value) 0.00030.3520.179
Ni 0.1650.312
(p-value) 0.5570.257
Pb 0.555
(p-value) 0.032
WeakModerateStrongVery Strong
Table 6. Cdeg, RI, CSI, and MERMQ values.
Table 6. Cdeg, RI, CSI, and MERMQ values.
Sample
No.
AreaCdegRICSIMERMQ
1Gemsha4.29Low83.73moderate risk0.35uncontaminated0.03low
2North Hurghada2.54Low34.43low risk0.34uncontaminated0.04low
3Hurghada2.68Low14.84low risk0.40uncontaminated0.06low
4South Hurghada2.59Low24.42low risk0.38uncontaminated0.05low
5North Safaga6.79Moderate118.67high risk0.54uncontaminated0.08low
6South Safaga5.40Low60.57moderate risk0.48uncontaminated0.07low
7El-Hamrawein9.81Moderate142.59high risk0.73very low0.17medium
8Qusier6.34Moderate68.93moderate risk0.66very low0.18medium
9N. Marsa Umm El-Griufat7.51Moderate51.85moderate risk0.86very low0.35medium
10N. Sharm El Bahria4.94Low54.73moderate risk0.55very low0.10medium
11Abu Dabab3.74Low64.47moderate risk0.41uncontaminated0.05low
12Sharm El Naaba7.49Moderate47.88low risk0.73very low0.33medium
13Marsa Alam2.35Low20.13low risk0.33uncontaminated0.04low
14Wadi Lahmi3.93Low42.48low risk0.48uncontaminated0.08low
15Marsa Humira2.50Low34.70low risk0.35uncontaminated0.05low
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Farhat, H.I.; Gad, A.; Saleh, A.; Abd El Bakey, S.M. Risks Assessment of Potentially Toxic Elements’ Contamination in the Egyptian Red Sea Surficial Sediments. Land 2022, 11, 1560. https://doi.org/10.3390/land11091560

AMA Style

Farhat HI, Gad A, Saleh A, Abd El Bakey SM. Risks Assessment of Potentially Toxic Elements’ Contamination in the Egyptian Red Sea Surficial Sediments. Land. 2022; 11(9):1560. https://doi.org/10.3390/land11091560

Chicago/Turabian Style

Farhat, Hassan I., Ahmed Gad, Ahmed Saleh, and Sahar M. Abd El Bakey. 2022. "Risks Assessment of Potentially Toxic Elements’ Contamination in the Egyptian Red Sea Surficial Sediments" Land 11, no. 9: 1560. https://doi.org/10.3390/land11091560

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

Article Metrics

Back to TopTop