Next Article in Journal
DIGIT: An In Situ Experiment for Studying the Diffusion of Water and Solutes under Thermal Gradient in the Toarcian Clay Rock at the Tournemire Underground Research Laboratory: Part 1—Goals, Scoping Calculations, Installation and First Results under Unheated Conditions
Next Article in Special Issue
Geochemical Signature and Risk Assessment of Potential Toxic Elements in Intensively Cultivated Soils of South-West Punjab, India
Previous Article in Journal
Timing and Tectonic Implications of the Development of the Orosirian Qianlishan Ductile Shear Zones in the Khondalite Belt, North China Craton
Previous Article in Special Issue
Source Analysis and Bioavailability of Soil Cadmium in Poyang Lake Plain of China Based on Principal Component Analysis and Positive Definite Matrix Factor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multi-Elemental Characterization of Soils in the Vicinity of Siderurgical Industry: Levels, Depth Migration and Toxic Risk

1
INPOLDE Research Center, Department of Chemistry, Physics and Environment, Faculty of Sciences and Environment, Dunarea de Jos University of Galati, 47 Domneasca St., 800008 Galati, Romania
2
Joint Institute for Nuclear Research, 6 Joliot Curie St., 141980 Dubna, Russia
3
Geological Institute of Romania, 1 Caransebes St., 012271 Bucharest, Romania
4
Department of Structure of Matter, Earth and Atmospheric Physics, Astrophysics, Faculty of Physics, University of Bucharest, 405 Atomistilor St., 077125 Magurele, Romania
*
Author to whom correspondence should be addressed.
Minerals 2024, 14(6), 559; https://doi.org/10.3390/min14060559
Submission received: 8 April 2024 / Revised: 24 May 2024 / Accepted: 25 May 2024 / Published: 29 May 2024

Abstract

:
The assessment of soil contamination in the vicinity of integrated siderurgical plants is of outmost importance for agroecosystems and human health, and sensitive techniques should be employed for accurate assessment of chemical elements (metals, potential toxic elements, rare earths, radioelements) in soil and further evaluation of potential ecological and safety risk. In this paper a total of 45 major, minor and trace elements (Al, As, Au, Ba, Br, Ca, Cd, Ce, Co, Cr, Cs, Cu, Dy, Eu, Fe, Hf, Hg, I, K, La, Mg, Mn, Mo, Na, Nd, Ni, Pb, Rb, Sb, Sc, Sm, Sn, Sr, Ta, Tb, Th, Ti, Tm, U, V, W, Y, Yb, Zn and Zr) were quantified in soils located around a large siderurgical works (Galati, SE Romania) using instrumental neutron activation analysis (INAA) in combination with X-ray fluorescence (XRF) and inductively coupled plasma mass spectrometry (ICP–MS). The statistical analysis results and vertical distribution patterns for three depths (0–5 cm, 5–20 cm, 20–30 cm) indicate inputs of toxic elements in the sites close to the ironmaking and steelmaking facilities and industrial wastes dumping site. For selected elements, a comparison with historical, legislated and world reported concentration values in soil was performed and depth migration, contamination and toxic risk indices were assessed. The distribution of major, rock forming elements was closer to the Upper Continental Crust (UCC), and to the Dobrogea loess, a finding confirmed by the ternary diagram of the incompatible trace elements Sc, La and Th, as well as by the La to Th rate. At the same time, the La/Th vs. Sc and Th/Sc vs. Zr/Sc bi-plots suggested a felsic origin and a weak recycling of soils’ mineral components.

1. Introduction

Soil is a vital resource which provides essential support to humans and ecosystems. Industrial activities, such as metal smelting, steel production, mining, the chemical industry and coal-fired power production, may cause soil quality degradation and serious pollution of the environment with toxic trace elements which could negatively impact human health due to their accumulation, persistence and transfer in the compartments of the food chain [1,2,3,4,5,6,7]. The elemental analysis at trace level is an important analytical task in various laboratories and high precision techniques, such as Instrumental Neutron Activation Analysis (INAA) [6,7,8,9,10], X-ray Fluorescence (XRF) [4,8,10,11] and inductively coupled plasma mass spectrometry (ICP–MS) [2,8,12,13,14], have shown to be powerful tools for multi-elemental analysis of a variety of environmental matrices, including soil and rock material, as well as for the assessment of mineralogical and geochemical features and spatio-temporal contamination trends of a certain region.
This work aimed to apply INAA, XRF and ICP–MS in order to examine the soil multi-element load, level of soil pollution with Potential Contaminant Elements (PCEs) and their ecological impact around the largest integrated ferrous (iron and steel) metallurgical plant (ISP) in Romania located at Galati town, SE Romania, which is one of the most important integrated complexes in the South-East of Europe (Figure 1) [11,15,16], with a great potential for environmental contamination and health damage (neurological disorders, asthma, cancer, cardiopulmonary disease, allergies, organ damage, hormonal imbalance, anemia, etc.) due to specific emissions from iron smelting, processing of ores, coal and auxiliary materials, agglomeration and sintering, minerals and scrap transportation, casting, molding, steelmaking, etc. [3,5,17]. The activity of this large ferrous metallurgical facility started in 1965, is continuing today, but in the last decade some of the units (blast furnaces, steelworks and mills) were closed and slag dumping has stopped [18].
The aim of this paper was two-fold: (1) to study the soil mineralogy, spatial and depth distribution pattern of 45 major and trace elements in soils adjacent to the siderurgical industry; and (2) to assess the ecological risk of selected PCEs, which might produce serious health disorders.
The work was carried out as a continuation and extension of pollution and risk assessment in previous studies limited to soil contamination with heavy metals [6,11,15] and persistent organic pollutants (POPs) [16] in the Galati industrial area and SE Romania in the period 2005–2009, and in connection with other international projects focused on the determination of toxic compounds and assessment of ecological and human health risk in the Black Sea Basin [3,14,19,20].

2. Materials and Methods

2.1. Sampling and Sample Processing

Composite soil samples of about 1.5 kg were collected at nine sites belonging to several territorial-administrative units (TAUs) of Galati (GL) and Braila (BR) counties in the South–East Development Region of Romania, around a large steel plant in SE Romania (Figure 1), from three different layers at depth intervals of 0–5 cm, 5–20 cm and 20–30 cm. The samples labeled as G1 were taken from the village of Vadeni (TAU Vadeni), Braila County, and the other samples (G2, G3, G4) from Galati County, from the localities of Sendreni and Smardan TAUs and Galati town. A control sample (GC) was collected from Vanatori village, in the northern part of Galati town (Figure 1). The site description is presented in Table 1. The importance of the targeted region also derives from its closeness to the Danube River, the confluence of Danube with its main tributaries the Siret and Prut Rivers, and proximity to the ecologically valuable natural sites and special protected areas in Romania, Republic of Moldova and Ukraine in the Lower Danube River and Black Sea basins [16,19,21].
The soil samples were taken from homogeneous, undisturbed soils, belonging to the classes: (a) Cernisols, Chernozems type, calcareous subtype—developed on parental materials of aeolian origin (loess and loess deposits), and (b) Protisols, Alluviosol type, calcareous and gleic mollic limestone subtypes—developed on carbonate fluvial parent materials. The general characteristics of soils in the target region are given in Table A1, Appendix A [11,16,21,22,23,24,25,26]. In the studied area, the soils do not show variability within very wide limits, regarding their delimitation at the level of type and subtype. On the territories of Smardan, Sendreni and Galati, located in the Covurlui Plain, the soil samples were taken from agricultural lands located in the interfluvial area, without considerable slope variations. On the Vadeni territory, geographically located in the Lower Siret Plain, the samples were collected from the alluvial plain. From the climatic point of view, the territories mentioned are located in the temperate-continental climate zone, with excessive nuances and variations during the year regarding the distribution of temperatures and precipitation. The sampling was performed from three layers in the first 30 cm of soil in order to observe the migration of elements in deeper layers and compare the actual contamination status with historical data.
The collected soil samples were prepared for various analyses at INPOLDE research center, Dunarea de Jos University, Galati, Romania, cleaned from stones, vegetal material and other debris, dried at room temperature, then grounded and sieved to a granulation of 0.01 mm.

2.2. Analytical Techniques

Two non-destructive multi-elemental nuclear analytical techniques commonly applied in geochemistry studies and materials certification programs were employed in this research: epithermal neutron activation analysis (INAA) and X-ray fluorescence (XRF) analysis. The two analysis techniques were used to determine with high accuracy and precision the total concentrations of 44 major, minor and trace elements (INAA—Na, Mg, Al, K, Ca, Ti, V, Cr, Mn, Fe, Ni, Co, Zn, As, Br, Rb, Sr, Zr, Nb, Mo, Sb, I, Cs, Ba, La, Ce, Nd, Sm, Eu, Tb, Dy, Tm, Yb, Hf, Ta, W, Au, Hg, Th and U; XRF—Cu, Pb, Sn, Y) in the soil samples taken around a large steel plant in SE Romania. QA/QC was performed using certified reference materials having a similar matrix as the samples. Due to unsuitability of INAA and XRF for quantification of Cd at very low concentrations in soil, the ICP–MS technique was employed in this work for Cd analysis.
The instrumental neutron activation analysis (INAA) analytical technique was applied at the REGATA installation of the IBR–2M reactor within the Frank Laboratory of Neutron Physics (FLNP), of the Joint Institute for Nuclear Research (JINR) Dubna, Russian Federation. Several channels of the IBR–2M reactor were used, which allowed irradiation of the samples with thermal and epithermal neutrons. Soil samples weighing around 100 mg, together with certified reference materials, were wrapped in polyethylene bags and aluminium foils for short and long-term irradiation, respectively, according to the analytical scheme described in other work [27,28,29]. The induced radioactivity in the samples was measured with the aid of gamma spectrometric chains equipped with an automatic switch system for soil and standard material samples [27]. The gamma spectra were processed using the software developed at JINR [28] and the analysis was performed based on the radionuclides formed in the samples during activation with neutrons with various energies, listed in [27].
For the application of XRF spectrometric analysis with energy dispersion (EDXRF), the prepared samples were put in specific capsules coated with a Myler foil. The encapsulated samples were irradiated for 120 s in SOIL mode using a portable Genius XRF spectrometer manufactured by Skyray Instruments Inc., equipped with a large area and detector with Be window and an excitation source of 40 kV/100 µA miniature X-ray tube with an Ag-target [14,18].
ICP–MS was applied for cadmium determination with the aid of an Agilent 7700Χ ICP–MS (Agilent, Santa Clara, CA, USA) spectrometer from the International Hellenic University (IHU), Kavala, Greece, a partner laboratory of the INPOLDE center within the framework of the MONITOX network [20], as described in a previous study [14]. The solid homogenized soil samples were weighed via the analytical balance (≈3 g), and 30 mL of strong HCl and HNO3 mixture at a ratio of 1:3 (aqua regia) was added. The samples were put in a Berghof MWS-2 microwave oven (P = 2000 W) using the following ramping parameters: T = 200 °C, P = 90%, t = 25 min. After digestion, the obtained solutions were filtered, transferred into graduated flasks (100 mL) and diluted with 5% HNO3 solution to the filling mark. A calibration curve with five points (R2 = 0.997) was built with the aid of stock standard solutions of different concentrations. All samples were analyzed in triplicate. Using an independent certified standard, a 105% recovery of Cd was obtained.

2.3. Depth Migration Index

In order to qualitatively assess the degree of migration/mobility in soil of the determined chemical elements, the depth migration index (DMI) was calculated for each element using the Equation (1), adapted by us after [30], for all the depths from which samples were taken (from 0 to 30 cm):
DMI   = i = 1 n c i c T D i
where n = 3 represents the number of layers from which the samples were taken, ci is the value of the chemical element concentration in each sampled layer i, c T = i = 1 n c i is the sum of the values of the chemical element concentrations from all the depths of the sampled layers, and Di is the depth (in cm) of the lower limit of each sampled layer (5, 20 or 30 cm). The value of this index ranges from 1 (if the element accumulated entirely in the first cm) to 30 (if the element accumulated entirely between 29 and 30 cm). The migration/mobility potential is classified into four categories, labelled as: A (DMI < 5), very low; B (5 < DMI < 10), moderate; C (10 < DMI < 20), high; D (DMI > 20), very high [15].

2.4. Contamination and Toxic Risk Indices

In order to describe the pollution extent and risk in a region impacted by multi-pollutants, several single and complex risk indices have been developed [2,12,31,32,33,34,35].
In this paper, we only used single and multi-element contamination indices, as well as the ecological risk index to highlight the impact of selected PCEs on the industrial soil quality and surrounding ecosystem state. Nevertheless, to assess the toxicity of PCEs in environmental compartments and for better understanding the ecotoxicological danger, these evaluations could be completed with bioassays [4,35] and health risk assessments [5].
Contamination factor (CF) for an individual chemical element is defined by Equation (2) as the ratio of each element concentration in the soil sample to the element concentration in background soil [33]:
CF = c sample c background
where csample and cbackground are the measured concentration and the background concentration value of metal i, respectively. The contamination classification of soils using this index is: CF < 1, low; 1 ≤ CF < 3, moderate; 3 ≤ CF < 6, considerable; CF ≥ 6, very high [2,31,32,33].
To evaluate the pollution levels of various contaminants specific to a site, the site Pollution Load Index (PLI) was employed, utilizing Equation (3) as introduced by Tomlinson et al. (1980) [34]. This equation involves computing the nth root of the product of the highest individual contamination factors CF i , calculated for specific metals (PCEs) which might result from siderurgical industry activity (n = number of elements/contamination factors). In this work we considered n = 13 PCEs, chosen from those specified in Romanian legislation [36] and probably emitted during the iron and steel making and steel products manufacturing processes, e.g., Hg, Cd, As, Pb, Cu, Co, Ni, Cr, Mn, Zn, Fe, V and Sb. The site PLI categories are the following: PLI ≤ 1, unpolluted (Background Pollution); PLI ≥ 1, polluted.
P L I = CF 1 ×   CF 2   ×     × CF n 1 / n
Similar to the definition of the site PLI and classification, a Regional Pollution Load Index (RPLI) is defined in this work with the aid of Equation (4), as the mth root of the product of the individual site average PLIs [34], calculated for the m = 9 industrial sites (except for the control site GC) in the Galati ISP area,
R P L I = PLI 1   ×   PLI 2 ×     ×   PLI m 1 / m
Ecological risk factor ( E r i ) for an element i is defined with the aid of Equation (5):
E r i = T r i ×   CF i
where the toxic response factors ( T r i ) for the selected toxic trace elements, indicated in the literature studies, are the following: Hg–40 [31,32,33,37], Cd–30 [31,32,33,37], As–10 [2,31,32,33,37], Pb–5 [2,31,32,33,37], Cu–5 [2,31,32,33,37], Co–5 [32], Ni–5 [2,31,32,37], Cr–2 [2,31,32,33,37], Mn–1 [2,32], Zn–1 [2,31,32,33,37]. The risk classification using this individual ecotoxicological index is: E r i < 40, low potential ecological risk; 40 ≤ E r i < 80, moderate potential ecological risk; 80 ≤ E r i < 160, considerable potential ecological risk; 160 ≤ E r i < 320, high potential ecological risk; E r i ≥ 320, very high potential ecological risk [31,37].
The risk index RI is the sum of calculated ecological risk factors corresponding to all n = 10 hazardous elements analyzed, according to Equation (6):
R I = i = 1 n E r i
The risk classification using this complex ecotoxicological index is: RI < 90, low; 90 ≤ RI < 180, moderate; 180 ≤ RI < 360, strong; 360 ≤ RI < 720, very strong; RI ≥ 720, highly strong [31].

2.5. Mapping and Statistical Data Analysis

Principal Component Analysis (PCA), Discriminant Analysis (DA), and Cluster Analysis (CA), i.e., Joining or Tree Clustering, were applied to examine the relationship among the sampling sites and elemental load for each soil layer. The inputs consisted of 45 variables representing the concentrations for all 45 chemical elements, which were determined in soils collected from nine industrial sites surrounding the enterprise and one control site (GC).
Given the high number of elements of which mass fractions were determined by INAA, we have used Principal Component Analysis (PCA) [38] due to its capacity to perform a linear dimensionality reduction. This technique permits the organizing and plotting of a great volume of independent data in more two-dimensional scatterplots, thus visualizing the clusters of closely related data. Consequently, all data were classified into cases, in our case the sampling points, and variables which, in the present study, represent the mass fractions of the considered elements distributed in each case, i.e., sampling point.
PCA can be performed in Q or R mode, depending on what needs to be grouped, variables, or cases in clusters. In the case of R mode, samples, in our case sampling points, are classified using the mass fractions of all 45 elements as variables.
Closer to PCA, DA also represents a classification technique [39] used when a priori the number of clusters or populations in our case the soil levels, are known. In this way, the variables that consist of independent variables, i.e., the PCE mass fraction values and the dependent variables, in this case, the soil layers, generate a bi-plot that illustrates several clusters, whose number is equal to the number of dependent variables, more or less distinctly separated.
The experimental data concerning the spatial distribution of RI as well as PLI values for selected PCEs in all three investigated soil layers were organized in a geospatial data base, further processed by ArcMap 10.4 Spatial Analyst [40]. This permitted the generating of six maps representing the spatial distribution of both RI and site PLI indices.
Due to the fact that the sampling points did not follow a regular network, the entire considered area was decomposed in Thiessen–Voronoi polygons [41,42]. In this case, an inverse distance weighting algorithm [43,44] gave the best results. Due to a reduced number of sampling points and in the absence of confident data concerning the distribution function of PLI values, the nonparametric ANOVA Tukey, Mann–Whitney, Dunnett post hoc and Spearman correlation were used to calculate the probabilities that the PLI distributions corresponding to the considered three layers are closer.
Further, Discriminant Analysis (DA) [39] was used to evidence to which extent the PLI values of each layer are correlated.
PAST 4.1 [45], TIBCO® Data Science/Statistica™ 10 [46], and OriginLab® OriginProTM 2021 [47] were used for statistical data analysis, and MS Excel 2019 was employed for experimental data processing and graphing.

3. Results and Discussion

3.1. Concentrations of Elements in Industrial Soils

The results obtained by the combined analytical methods INAA, XRF and ICP–MS highlighted the existence of 45 elements in the soils. Figure 2 shows the spatial and depth wise distribution of the element concentrations and Table 2 presents the concentration ranges and average of values obtained in this paper, along with the limits allowed by Romanian norms for trace elements in soil (normal values, low alert and intervention for sensitive use of soil, and high alert and intervention for less sensitive use of soil) [36] and literature data for Earth’s crust composition [48].
From Figure 2, it can be seen that the highest concentrations of most of the trace elements occur in sites G.2.2. G2.3 and G3.2 which are closest to the industrial complex, as well as in location G1.1, situated within the predominant wind direction, a trend similar to previous findings for the same industrial area [11,15].
Compared with norms stipulated by Romanian law [36], the concentration values obtained in this work for several PCEs (Table 2)—As, Ba, Co, Cr, Cu, Hg, Mn, Mo, Ni, Pb and Zn—exceed the normal levels, while in the case of Ba, Cr, Pb and V the alert levels are surpassed. With the exception of Al, Ba, Co, K, Mg, Na Sc, Sr and Y, all other major and trace elements are present in the investigated soils in concentrations higher than the Upper Continental Crust (UCC) values [48].
The variation of the depth migration index DMI of identified chemical elements for each sampling location is illustrated in Figure 3.
DMI showed very high values (DMI > 20 cm) for the following elements at specified locations: Ca at G2.1 and G4.1; Co and Ni at G2.1 and G4.1; Cr and Cu at G4.1; As and Sb at G4.1; Mo at G1.1, G1.3 and G4.1; Rb, Sr and Zn at G4.1; Br at G2.1 and G4.1; Cs at G4.1; Th, Tm, U, and Yb at G4.1; Sc at G2.1; Ce and Zn at G4.1; Ba at G2.1; Dy at G2.3; La and Sm at G4.1; Tb at G2.1 and G4.1; Nd at G1.2, G2.1 and G4.1; Au at G2.1, G3.1, G3.2, G4.1; Hf, W, Y and Ta at G4.1. For the other chemical elements, the DMI values point to an accumulation in the second soil layer (5–20 cm). Thus, a tendency can be noticed for element mobilization in the deeper layers of soil, most of the sites being characterized by high (class C) and very high (class D) vertical migration potential. The urban site G4.1. is different from other locations, being classified with very high migration potential of trace elements; this trend is also visible in Figure 2, as the elemental levels in the first soil layer are lower than in the second and third layers of this sublocation.
The results for PCA and CA of elemental concentrations are presented in Figures S1 and S2, Supplementary Material, respectively, for 10 cases (soil sampling sites—9 industrial sites around Galati integrated siderurgical plant and 1 control site) and 45 variables (concentrations of chemical elements).
PCA results modeled on the basis of the first two most significant principal components (PC1, PC2) specified in Figure S1 are depicted in Figure 4.
The PCA shows that as the depth increases the sites are better grouped, with the exception of sites G4.1 for the depth 0–5 cm, and G1.1 for the depths 5–20 cm and 20–30 cm, respectively. In all cases, it is noticed that the control site GC is distant from the main group, as expected. Similar results were found by applying CA for site clustering (Figure S2, Supplementary Material).

3.2. Soil Mineralogy

In assessing the origin of the soil mineral component, the incompatible trace elements, such as Sc, Co, Zr, rare earth elements (REE), Hf, or Th, are very useful [7,10]. In this regard, felsic rocks are depleted in Sc but enriched in Th and light REE, while the mafic rocks appear enriched in Sc, Cr, Co, Ni and heavy REE [49]. On this subject, the Sc mass fraction in felsic rocks is less than 20 mg/kg but exceeds 20–40 mg/kg in mafic rocks so that Sc is one of the most appropriate proxies for this kind of study [50]. In the case of the investigated soils, Sc mass fraction was 9.8 ± 2 mg/kg. In this regard, the discriminating bi-plot Th/Co vs. La/Sc (Figure 5a) [51] points with clarity towards a felsic origin of the mineralogy of the investigated soil.
The presence and distribution of Sc, Zr and Th in the investigated material permitted inference of the past processes that shaped the current situation. Indeed, the mineral zircon, due to an extreme resilience to abrasion explained by its hardness on Mohs scale as greater than 7.5, showed extreme resilience during recycling, so that the higher the Zr mass fraction, the larger the sedimentary material for sorting and recycling [52].
In this case, the biplot Th/Sc vs. Zr/Sc showed a relatively reduced recycling when compared with the Dobrogea loess (Figure 5b) [53], a fact explained by the predominance of clastic material in the considered soils. More information concerning the affinities of the mineral constituents of the soil with the other mineral systems can be furnished by reciprocal distribution of Sc, La and Th of La and Th [54]. In this regard, both discriminating La-Sc-Th ternary (Figure 5c) and La vs. Th biplot (Figure 5d) showed a good correlation between inorganic soil, minerals, components and the Upper Continental Crust (UCC) [48,54] and the North American Shale Composite (NASC) [55].
In the case of the Sc-La-Th ternary diagram (Figure 5c), all the experimental points corresponding to the Galati soils form a cluster around UCC [48,54] and NASC [55], suggesting a composition consisting of clay, silt or gravels from mixed sources. In its turn, the the La vs. Th bi-plot (Figure 5d) shows a good correlation between these incompatible elements, of which the ratio of 3.3 ± 0.35 is close to the UCC ratio of 2.95 [54] or the NASC value of 2.65 [55]. This approach to the UCC [48,54] and NASC [55] also demonstrates an obtained La/Sm mass fractions ratio of 5.47 ± 0.17, close to the UCC value of 6.6 [48] and NASC of 5.7 [55].

3.3. Soil Contamination and Ecological Risk

The second aim of this study consisted of a detailed investigation of 13 PCEs’ distribution around the Galati integrated steel plant (ISP) given its status as a potential contamination source. Accordingly, based on the mass fractions of V, Cr, Mn, Fe, Co, Ni, Cu, Zn, As, Cd, Hg, Sb and Pb determined in soil cores covering the 0–5, 5–20 and 20–30 depth layers, the corresponding PLI (series PLI-13 in Figure 6) was calculated for each sampling point and each layer as one of the most representative contamination indices (Figure 6). In the calculations for Contamination Factors (CF) (Table A2, Appendix A) with the aid of Equation (2), the UCC element concentrations [48] were considered as background values.
Based on the average values of 9 site PLIs, a regional index RPLI of 1.20 was calculated for the metallurgical industrial area of Galati, SE Romania, emphasizing a polluted zone (RPLI >1).
Comparing the PLI values for 13 PCEs with the PLI calculated for all 45 identified elements (series PLI-45 in Figure 6), it can be stated that the use of pollutant elements gives rise to a relative higher PLI than in the case of considering the elements present at near baseline levels [34], suggesting a real deterioration in soil quality in the region, which is the correct approach.
Final results showing the PLI distribution for each sampling point, including the remote one GC, are illustrated by the violin diagram reproduced in Figure 7a. Here, it can be remarked that, for all three soil layers, the reference example appears less contaminated, a fact sustained by PLI values varying between 0.45 and 0.78, i.e., lower than the threshold equal to one, which is necessary to prove local contamination.
To deepen the study of local contamination, we have performed a DA of the PCE values by considering the sampling points as cases and the mass fractions of each of the 13 contaminating elements as independent variables, while the dependent variables are designated as soil layers. The result is reproduced in Figure 7b by a Root 2 vs. Root 1 biplot. This graph shows with clarity the presence of two distinct clusters, one consisting of a partial superposition of the 0–5 and 5–20 cm points and the other containing only the 20–30 cm data. In our opinion, this graph suggests that the contamination process is spread onto the 0-20 cm layer while the third layer, between 20 and 30 cm seems less affected.
This observation is confirmed also by several samples in the ANOVA, Tukey, Mann–Whitney (Table 3a) and Dunnett post hoc tests, as well as Spearman rank correlation (Table 3b), which showed a certain degree of similitude only between the 0–5 and 5–20 cm layers.
Another peculiarity can be remarked by analyzing the spatial distribution of PLI corresponding to each of the considered soil layers and illustrated by the maps in Figure 8a–c. Here, it can be observed that the maximum contamination corresponds to two well evidenced maxima in the sites G2.2 (close to the slag dump) and G3.2 (at the north gate of the enterprise, close to the iron scrap deposit, lime factory, steel plant and plate product mills), for which the PLI reaches closer values of 1.55 and 1.51, respectively, for the 0–5 cm layer, closely followed by the G2.3 site (at the south gate of the enterprise, close to agglomeration, sintering and coking plants, blast furnaces and ironmaking facilities) with a PLI value of 1.46.
With the depth, these areas enlarge in the second layer (5–20 cm) until they merge in the third layer (20–30 cm), while the maximum PLI values decrease to 1.41 in all three places in the second layer, and 1.10, 1.15, and 1.18, respectively, in the third layer. In our opinion, this finding illustrates a 3D diffusion of contaminating elements into soil starting from the two hotspots better evidenced in Figure 8a.
Three maps were generated representing the spatial distribution of the risk index RI for each soil layer (Figure 8d–f), calculated in Table 4 with the aid of Equation (6) for 10 selected PCEs based on the individual ecological risk factors E r i computed with Equation (5).
The RI distribution maps highlight the depth diffusion trend of hazardous elements shown by the PLI maps (Figure 8a–c).
The obtained values for the individual ecological risk factors E r i and the risk index RI (Table 4) show that highly ecological risk in the identified industrial hot spots is due to Cd and Hg.
The percentage contribution of each of the ten PCEs at the total value of RI depicted in Figure 9 shows the considerable risk due to highly toxic elements Hg and Cd, followed by As, Pb, Cu, Ni, Co, Cr, Zn and Mn.
Due to the negative impact of PCEs emitted as a result of steel production on soil quality, terrestrial and aquatic ecosystems’ state and people’s health [13,17,56], it is very important to regularly monitor these elements in soils neighboring industrial factories and waste deposits and assess the environmental damage.
At the most contaminated sites, it is recommended to apply soil depollution strategies and metallogenic area remediation techniques, such as phytoremediation, microbe-assisted remediation, constructed wetlands or extraction of polluted soil material [32,57,58,59], depending on the local weather conditions, contamination degree and depth, and soil physico-chemical properties, which could influence the mobility of elements in soil [58].

4. Conclusions

The current study aimed to bring new insights into the quality of soils in the neighborhood of a large siderurgical complex located in Galati, SE region of Romania, regarding the compositional scheme, spatial and depth migration of 45 chemical elements (metals, radioelements, rare earth elements and other trace elements), contamination trends, mineral assemblage, ecological risk of selected potential contaminant elements and regional contamination level due to industrial activity related to ferrous metallurgy.
The results obtained for the concentrations and environmental and safety risk indices of elements of concern demonstrate a present pollution of surface layer of soils adjacent to the siderurgical industry, although the industrial production and activity had been restrained in the last decade, along with a vertical migration of most of the analyzed elements in deeper soil layers.
The geochemical and contamination features suggest a similarity with results obtained in previous surveys performed in the Galati industrial area, regarding the location of the impacted sites with regard to the distance from the enterprise and dominant direction of wind. Our results can constitute a useful database for further investigations in the area and elaboration of land management strategies and exploitation of terrains, as well as decommissioning of the slag dump territory and valorization of metallic materials in various purposes.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/min14060559/s1, Figure S1: Explained variance for the principal components and plot of the PC2 loadings as a function of PC1 loadings; Figure S2: Dendrograms obtained with the Tree Clustering algorithm for the 10 soil samples and 45 variables in three soil layers.

Author Contributions

Conceptualization, A.E. and O.G.D.; Data curation, A.E. and M.V.F.; Formal analysis, F.S. and D.P.; Investigation, A.E., F.S., M.V.F., A.S., S.G. and D.P.; Methodology, A.E., F.S., M.V.F., O.G.D., A.S. and S.G.; Project administration, A.E. and M.V.F.; Resources, A.E.; Software, O.G.D., S.G. and D.P.; Supervision, A.E.; Validation, A.E. and M.V.F.; Visualization, M.V.F. and A.S.; Writing—original draft, A.E., F.S., M.V.F., O.G.D., A.S., S.G. and D.P.; Writing—review and editing, A.E., M.V.F. and O.G.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by 2017-2022 JINR-Romania program (JINR Theme no. 03-4-1128-2017/2022; Protocol no. 4613-4-17/22 between JINR and Dunarea de Jos University of Galati), and EC through JOP Black Sea Basin 2014-2020, project code BSB27-MONITOX (2018-2021).

Data Availability Statement

Data supporting reported results are available from the corresponding author upon request.

Acknowledgments

We acknowledge the support given by the JINR and IHU Kavala laboratory teams for technical support during INAA and ICP–MS analyses.

Conflicts of Interest

The authors declare no conflicts 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.

Appendix A

Table A1. General characteristics of soils in the investigated region.
Table A1. General characteristics of soils in the investigated region.
Location pHOC (%) *Clay (%)Soil Type
(WRB and SRTS) [23,24]
Soil Class (SRTS) [23]
TAU Smardan [21]8.00–8.370.86–2.7412.26–19.97Calcaro-calcic ChernozemCernisols
TAU Sendreni [21,22]7.88–8.440.91–1.661.15–27.98Calcaro-calcic ChernozemCernisols
TAU Vadeni [21,22]8.18–8.210.69–1.4519.20–33.87Calcaric and Calcaric Mollic Gleic FluvisolsProtisols
Galati town [11,16]8.40–8.841.37–2.4713.00–20.00 **Calcaro-calcic ChernozemCernisols
* Considering OC (%) = OM (%)/1.7241 (OC—organic carbon, OM—organic matter); ** References [25,26].
Table A2. Contamination Factors (CF) for 13 PCEs in investigated soils for three layers.
Table A2. Contamination Factors (CF) for 13 PCEs in investigated soils for three layers.
SiteD (cm)AsCrCuCdCoHgMnNiPbZnFeVSb
G1.10–51.491.060.863.560.672.141.260.912.111.960.881.062.03
5–201.651.030.683.220.890.691.181.362.771.761.111.052.45
20–302.210.841.072.780.650.240.970.851.531.020.710.572.60
G1.20–52.011.280.713.110.820.500.981.031.732.331.120.762.83
5–201.441.000.772.890.654.110.990.991.511.780.870.772.60
20–301.650.940.752.560.710.690.860.981.411.260.780.631.95
G1.30–51.541.090.792.670.533.040.920.651.341.410.760.653.33
5–201.660.961.962.560.552.141.090.641.281.751.100.652.33
20–301.450.990.662.440.690.850.910.871.471.210.750.711.80
G2.10–51.680.800.493.780.403.750.910.351.441.120.560.683.23
5–201.900.940.573.110.602.320.780.651.461.320.770.602.50
20–301.821.100.582.890.730.570.740.920.641.070.810.532.25
G2.20–51.711.200.624.670.6016.071.370.761.801.690.930.722.48
5–201.761.280.764.330.627.501.050.711.621.410.890.722.33
20–301.681.050.693.890.700.600.971.021.601.070.810.592.08
G2.30–52.001.300.845.440.620.501.500.713.113.611.100.644.18
5–201.760.790.845.890.584.161.080.752.192.220.660.852.38
20–301.920.971.045.560.640.491.050.821.971.380.740.562.78
G3.10–52.031.080.792.890.680.500.860.711.601.520.810.692.63
5–201.730.980.782.780.661.130.830.851.371.070.710.582.30
20–302.041.070.822.560.730.710.810.871.411.000.770.592.13
G3.20–52.311.100.836.220.722.681.140.862.021.911.010.753.05
5–202.041.100.935.670.722.141.100.861.701.580.920.782.53
20–302.100.930.865.220.680.650.901.001.771.060.790.582.10
G4.10–50.780.430.803.000.240.500.940.292.480.790.840.701.40
5–201.900.880.532.780.572.320.950.611.871.580.720.753.33
20–301.850.851.282.560.620.530.760.811.521.350.690.562.50
GC0–51.030.600.271.890.450.500.560.450.640.890.450.451.60
5–200.980.480.311.780.422.500.610.361.120.750.430.541.53
20–300.830.460.091.560.390.350.450.260.530.450.350.321.38

References

  1. Gonçalves, D.A.M.; Pereira, W.V.d.S.; Johannesson, K.H.; Pérez, D.V.; Guilherme, L.R.G.; Fernandes, A.R. Geochemical Background for Potentially Toxic Elements in Forested Soils of the State of Pará, Brazilian Amazon. Minerals 2022, 12, 674. [Google Scholar] [CrossRef]
  2. Zafeiriou, I.; Gasparatos, D.; Megremi, I.; Ioannou, D.; Massas, I.; Economou-Eliopoulos, M. Assessment of Potentially Toxic Element Contamination in the Philippi Peatland, Eastern Macedonia, Greece. Minerals 2022, 12, 1475. [Google Scholar] [CrossRef]
  3. Ene, A.; Pantelică, A.; Sloată, F.; Zakaly, H.M.H.; Tekin, H.O. Gamma spectrometry analysis of natural and man-made radioactivity and assessment of radiological risk in soils around steel industry. Rom. J. Phys. 2023, 68, 803. [Google Scholar] [CrossRef]
  4. Andreu-Sánchez, Ó.; García-Lorenzo, M.L.; Esbrí, J.M.; Sánchez-Donoso, R.; Iglesias-Martínez, M.; Arroyo, X.; Crespo-Feo, E.; Ruiz-Costa, N.; Roca-Pérez, L.; Castiñeiras, P. Soil and Freshwater Bioassays to Assess Ecotoxicological Impact on Soils Affected by Mining Activities in the Iberian Pyrite Belt. Toxics 2022, 10, 353. [Google Scholar] [CrossRef]
  5. Mousavian, N.A.; Mansouri, N.; Nezhadkurki, F. Estimation of heavy metal exposure in workplace and health risk exposure assessment in steel industries in Iran. Measurement 2017, 102, 286–290. [Google Scholar] [CrossRef]
  6. Pantelica, A.; Freitas, M.C.; Ene, A.; Steinnes, E. Soil pollution with trace elements at selected sites in Romania studied by instrumental neutron activation analysis. Radiochim. Acta 2013, 101, 45–50. [Google Scholar] [CrossRef]
  7. Zinicovscaia, I.; Vergel, K.; Duliu, O.G.; Grozdov, D.; Yushin, N.; Chaligava, O. Assessment of Soil Pollution with Presumably Contaminating Elements in Moscow Recreational Areas Using Instrumental Neutron Activation Analysis. Sustainability 2023, 15, 7886. [Google Scholar] [CrossRef]
  8. Balaram, V. Advances in Analytical Techniques and Applications in Exploration, Mining, Extraction, and Metallurgical Studies of Rare Earth Elements. Minerals 2023, 13, 1031. [Google Scholar] [CrossRef]
  9. 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]
  10. Badawy, W.; Silachyov, I.; Dmitriev, A.; Lennik, S.; Saleh, G.; Mitwalli, M.; El-Farrash, A.; Sallah, M. Elemental distribution patterns in rock samples from Egypt using neutron activation and complementary X-ray fluorescence analyses. Appl. Radiat. Isot. 2023, 202, 111063. [Google Scholar] [CrossRef]
  11. Ene, A.; Bosneaga, A.; Georgescu, L. Determination of heavy metals in soils using XRF technique. Rom. Journ. Phys. 2010, 55, 815–820. Available online: https://rjp.nipne.ro/2010_55_7-8/0815_0820.pdf (accessed on 14 December 2023).
  12. Di Duca, F.; Montuori, P.; De Rosa, E.; De Simone, B.; Russo, I.; Nubi, R.; Triassi, M. Assessing Heavy Metals in the Sele River Estuary: An Overview of Pollution Indices in Southern Italy. Toxics 2024, 12, 38. [Google Scholar] [CrossRef]
  13. Zhou, T.; Bo, X.; Qu, J.; Wang, L.; Zhou, J.; Li, S. Characteristics of PCDD/Fs and metals in surface soil around an iron and steel plant in North China Plain. Chemosphere 2019, 216, 413–418. [Google Scholar] [CrossRef]
  14. Sloata, F.; Ene, A.; Bogdevici, O.; Spanos, T. Characterization of soils around a former chemical plant in Braila, SE Romania, using high performance atomic techniques (EDXRF, AAS, ICP-MS). Ann. Dunarea De Jos Univ. Galati Fasc. II Math. Phys. Theor. Mech. 2022, 45, 23–32. [Google Scholar] [CrossRef]
  15. Bosneaga (Sion), A. Quantification of the Soil Pollution Level (Cuantificarea Gradului de Poluare a Solului—In Romanian). PhD Thesis, Dunarea de Jos University of Galati, Galați, Romania, 2011. [Google Scholar]
  16. Ene, A.; Bogdevich, O.; Sion, A. Levels of organochlorine pesticides (OCPs) and polycyclic aromatic hydrocarbons (PAHs) in topsoils from SE Romania. Sci. Total Environ. 2012, 439, 76–86. [Google Scholar] [CrossRef]
  17. IARC Working Group on the Evaluation of Carcinogenic Risks to Humans. Occupational Exposures during Iron and Steel Founding; IARC Monographs on the Evaluation of Carcinogenic Risks to Humans No. 100F; International Agency for Research on Cancer: Lyon, France, 2012. Available online: https://www.ncbi.nlm.nih.gov/books/NBK304429/ (accessed on 29 April 2024).
  18. Sloată, F. Performant Analytical Techniques Used for Toxic Substances Monitoring and Industrial Waste Management (Tehnici Analitice Performante Utilizate pentru Monitorizarea Substanțelor Toxice și Managementul Deșeurilor Industriale—In Romanian). PhD Thesis, Dunarea de Jos University of Galati, Galati, Romania, 2023. [Google Scholar]
  19. Chiţescu, C.L.; Ene, A.; Geana, E.-I.; Vasile, A.M.; Ciucure, C.T. Emerging and Persistent Pollutants in the Aquatic Ecosystems of the Lower Danube Basin and North West Black Sea Region—A Review. Appl. Sci. 2021, 11, 9721. [Google Scholar] [CrossRef]
  20. Ene, A.; Zubcov, E.; Spanos, T.; Bogdevich, O.; Teodorof, L. MONITOX international network for monitoring of environmental toxicants and risk assessment in the Black Sea Basin: Research and interdisciplinary cooperation dimensions. In Proceedings of the 10th International Conference “Sustainable Use and Protection of Animal World in the Context of Climate Change”, Chisinau, Moldova, 16–17 September 2021; pp. 11–17. [Google Scholar] [CrossRef]
  21. Moraru, S.-S.; Ene, A.; Badila, A. Physical and Hydro-Physical Characteristics of Soil in the Context of Climate Change. A Case Study in Danube River Basin, SE Romania. Sustainability 2020, 12, 9174. [Google Scholar] [CrossRef]
  22. Arbanas (Moraru), S.S. Research on Iron and Steel Works Industry Impact on Soil Edaphic and Vegetal Potential in the Adjacent Areas (Cercetări Privind Impactul Activităţilor Industriei Siderurgice Asupra Potenţialului Edafic şi Vegetal al Solurilor din Zonele Adiacente—In Romanian). Ph.D. Thesis, Dunarea de Jos University of Galati, Galati, Romania, 2022. [Google Scholar]
  23. Florea, N.; Munteanu, I. (Eds.) Romanian Soil Taxonomy System (Sistemul Român de Taxonomie a Solurilor (SRTS)); Sitech Publisher: Craiova, Romania, 2012. (In Romanian) [Google Scholar]
  24. FAO/UNESCO Soil Map of the World. Available online: https://www.fao.org/soils-portal/data-hub/soil-maps-and-databases/faounesco-soil-map-of-the-world/en/ (accessed on 28 April 2024).
  25. Barac, M.; Kalmár, J.; Kuti, L.; Vatai, J. Soil Stability and Groundwater Chemistry in Urban Areas Covered by Loess (Investigation Site—Galati City, Romania). Earth Sci. Res. 2013, 2, 214–220. [Google Scholar] [CrossRef]
  26. Dobrescu, C.F.; Calarasu, E.A.; Craifaleanu, I.G. Ground settlement in urban structures exposed to geo-environmental and anthropic hazards: A case study for Galati. Procedia Eng. 2017, 190, 611–618. [Google Scholar] [CrossRef]
  27. Frontasyeva, M.V. Epithermal Neutron Activation Analysis at the IBR-2 reactor of the Frank Laboratory of Neutron Physics at the Joint Institute for Nuclear Research (Dubna). Phys. Atom. Nucl. 2008, 71, 1684–1693. [Google Scholar] [CrossRef]
  28. Pavlov, S.S.; Dmitriev, A.Y.; Frontasyeva, M.V. Automation system for neutron activation analysis at the reactor IBR-2, Frank Laboratory of Neutron Physics, Joint Institute for Nuclear Research, Dubna, Russia. J. Radioanal Nucl. Chem. 2016, 309, 27–38. [Google Scholar] [CrossRef] [PubMed]
  29. Badawy, W.M.; Ghanim, E.H.; Duliu, O.G.; El Samman, H.; Frontasyeva, M.V. Major and trace element distribution in soil and sediments from the Egyptian central Nile Valley. J. Afr. Earth Sci. 2017, 131, 53–61. [Google Scholar] [CrossRef]
  30. Dorronsoro, C.; Martin, F.; Garcia, I.; Simon, M.; Fernandez, E.; Aguilar, J.; Fernandez, J. Migration of trace elements from pyrite tailings in carbonate soils. J. Environ. Qual. 2002, 31, 829–835. [Google Scholar] [CrossRef] [PubMed]
  31. Kowalska, J.B.; Mazurek, R.; Gasiorek, M.; Zaleski, T. Pollution indices as useful tools for the comprehensive evaluation of the degree of soil contamination–A review. Environ. Geochem. Health 2018, 40, 2395–2420. [Google Scholar] [CrossRef]
  32. Kumar, V.; Pandita, S.; Setia, R. A meta-analysis of potential ecological risk evaluation of heavy metals in sediments and soils. Gondwana Res. 2022, 103, 487–501. [Google Scholar] [CrossRef]
  33. Håkanson, L. An Ecological Risk Index for Aquatic Pollution Control: A Sedimentological Approach. Water Res. 1980, 14, 975–1101. [Google Scholar] [CrossRef]
  34. Tomlinson, D.L.; Wilson, J.G.; Harris, C.R.; Jeffrey, D.W. Problems in the assessment of heavy-metal levels in estuaries and the formation of a pollution index. Helgol. Meeresunters 1980, 33, 566–575. [Google Scholar] [CrossRef]
  35. Ferri-Moreno, I.; Barquero-Peralbo, J.I.; Andreu-Sánchez, O.; Higueras, P.; Roca-Pérez, L.; García-Lorenzo, M.L.; Esbrí, J.M. Categorization of Mining Materials for Restoration Projects by Means of Pollution Indices and Bioassays. Minerals 2023, 13, 492. [Google Scholar] [CrossRef]
  36. Order No. 756 of 3 November 1997 for the Approval of the Regulation on Environmental Pollution Assessment. Eminent: Ministry of Waters, Forests and Environmental Protection. (Published in: Official Gazette No 303 bis of 6 November 1997). Available online: http://legislatie.just.ro/Public/DetaliiDocumentAfis/151788 (accessed on 10 January 2024). (In Romanian).
  37. Sur, I.M.; Micle, V.; Polyak, E.T.; Gabor, T. Assessment of soil quality status and the ecological risk in the Baia Mare, Romania Area. Sustainability 2022, 14, 3739. [Google Scholar] [CrossRef]
  38. Jolliffe, I.T. Principal Component Analysis (Springer Series in Statistics); Springer: Berlin/Heidelberg, Germany, 2002; ISBN 978-0387954424. [Google Scholar]
  39. McLachlan, G.J. Discriminant Analysis and Statistical Pattern Recognition; Wiley Interscience: Hoboken, NJ, USA, 2004; ISBN 978-0-471-69115-0. [Google Scholar]
  40. Spatial Analyst in ArcGIS for Desktop 10.4. Available online: https://www.esri.com/arcgis-blog/products/analytics/analytics/spatial-analyst-in-arcgis-for-desktop-10-4/ (accessed on 16 February 2024).
  41. Burrough, P.A.; McDonnell, R.A.; Lloyd, C.D. Principles of Geographical Information Systems; Oxford University Press: Oxford, UK, 2015; ISBN 9780198742845. [Google Scholar]
  42. Blebea-Apostu, A.-M.; Margineanu, R.M.; Persa, D.; Dumitras, D.-G.; Gomoiu, M.C.; Duliu, O.G. The distribution of natural radionuclides 40K, 228Ac, and 226Ra on Romanian Territory: A radiometric study. Environ. Monit. Assess. 2024, 196, 186. [Google Scholar] [CrossRef]
  43. Shepard, D. A two-dimensional interpolation function for irregularly-spaced data. In Proceedings of the 1968 ACM National Conference; Association for Computing Machinery: New York, NY, USA, 1968; pp. 517–524. Available online: https://dl.acm.org/doi/pdf/10.1145/800186.810616 (accessed on 22 February 2024).
  44. Spatial Autocorrelation (Global Moran’s I) (Spatial Statistics). Available online: https://pro.arcgis.com/en/pro-app/latest/tool-reference/spatial-statistics/spatial-autocorrelation.htm (accessed on 10 February 2024).
  45. PAST 4.03 Statistical analysis app for Windows. Available online: https://past.en.lo4d.com/windows (accessed on 10 February 2024).
  46. StatSoft, Inc. STATISTICA 11. Available online: www.statsoft.com (accessed on 10 January 2024).
  47. OriginLab Corporation. Available online: https://www.originlab.com/2021 (accessed on 16 January 2024).
  48. Rudnick, R.L.; Gao, S. Composition of the continental crust. In Treatise on Geochemistry; Holland, H.D., Turekian, K.K., Eds.; Elsevier-Pergamon: Amsterdam, The Netherlands, 2004; Volume 3, pp. 1–64. ISBN 978-0-08-098300-4. [Google Scholar]
  49. Marshak, S. Essentials of Geology, 6th ed.; W.W. Norton & Company: New York, NY, USA, 2019; ISBN 978-0393667523. [Google Scholar]
  50. Norman, J.C.; Haskin, L.A. The geochemistry of Sc: A comparison to the rare earths and Fe. Geochim. Et Cosmochim. Acta 1968, 32, 93–108. [Google Scholar] [CrossRef]
  51. Armstrong-Altrin, J.S.; Nagarajan, R.; Madhavaraju, J.; Rosalez-Hoz, L.; Lee, Y.I.; Balaram, V.; Cruz-Martínez, A.; Avila-Ramírez, G. Geochemistry of the Jurassic and Upper Cretaceous shales from the Molango Region, Hidalgo, eastern Mexico: Implications for source-area weathering, provenance, and tectonic setting. Comp. Ren. Geosci. 2013, 45, 185–202. [Google Scholar] [CrossRef]
  52. McLennan, S.M.; Hemming, S.; McDaniel, D.K.; Hanson, G.N. Geochemical Approaches to Sedimentation, Provenance, and Tectonics; Special Paper; Geological Society of America: Boulder, CO, USA, 1993; Volume 284, p. 20. [Google Scholar]
  53. Tugulan, L.; Duliu, O.; Bojar, A.-V.; Dumitras, D.; Zinicovskaia, I.; Culicov, O.A.; Frontasyeva, M.V. On the geochemistry of the Late Quaternary loess deposits of Dobrogea (Romania). Quat. Int. 2016, 399, 100–110. [Google Scholar] [CrossRef]
  54. Taylor, S.R.; McLennan, S.M. The Continental Crust: Its Composition and Evolution; Blackwell: Oxford, UK, 1985. [Google Scholar]
  55. Gromet, L.; Haskin, L.A.; Korotev, R.L.; Dymek, R.F. The North American shale composite: Its compilation, major and trace element characteristics. Geochim. Cosmochim. Acta 1984, 48, 2469–2482. [Google Scholar] [CrossRef]
  56. Cui, X.; Geng, Y.; Sun, R.; Xie, M.; Cui, Z. Distribution, speciation and ecological risk assessment of heavy metals in Jinan Iron & Steel Group soils from China. J. Clean. Prod. 2021, 295, 126504. [Google Scholar] [CrossRef]
  57. Lianwen, L.; Wei, L.; Weiping, S.; Mingxin, G. Remediation techniques for heavy metal-contaminated soils: Principles and applicability. Sci. Total Environ. 2018, 633, 206–219. [Google Scholar] [CrossRef]
  58. Zhang, X.; Yang, H.; Cui, Z. Evaluation and analysis of soil migration and distribution characteristics of heavy metals in iron tailings. J. Clean. Prod. 2018, 172, 475–480. [Google Scholar] [CrossRef]
  59. Khalid, S.; Shahid, M.; Niazi, N.K.; Murtaza, B.; Bibi, I.; Dumat, C. A comparison of technologies for remediation of heavy metal contaminated soils. J. Geochem. Explor. 2017, 182 Pt B, 247–268. [Google Scholar] [CrossRef]
Figure 1. The map of sampling points around the Galati integrated siderurgical plant (ISP) and the ISP scheme. The inset illustrates the position of the target area with respect to Romanian territory.
Figure 1. The map of sampling points around the Galati integrated siderurgical plant (ISP) and the ISP scheme. The inset illustrates the position of the target area with respect to Romanian territory.
Minerals 14 00559 g001
Figure 2. Spatial and depth wise distribution of the element concentrations determined in the investigated soil samples (concentrations on vertical axis are expressed in mg·kg−1).
Figure 2. Spatial and depth wise distribution of the element concentrations determined in the investigated soil samples (concentrations on vertical axis are expressed in mg·kg−1).
Minerals 14 00559 g002aMinerals 14 00559 g002bMinerals 14 00559 g002c
Figure 3. Depth migration index, DMI (cm), of elements in soils around Galati siderurgical plant.
Figure 3. Depth migration index, DMI (cm), of elements in soils around Galati siderurgical plant.
Minerals 14 00559 g003
Figure 4. Plot of the PC2 scores as a function of PC1 scores for the soils collected from three depths at the sites around the siderurgical plant and control site: (a) 0–5 cm; (b) 5–20 cm and (c) 20–30 cm.
Figure 4. Plot of the PC2 scores as a function of PC1 scores for the soils collected from three depths at the sites around the siderurgical plant and control site: (a) 0–5 cm; (b) 5–20 cm and (c) 20–30 cm.
Minerals 14 00559 g004
Figure 5. The Th/Co vs. La/Sc discriminating diagrams illustrating the felsic nature of the inorganic/mineral components of investigated soils (a), the Th/Sc vs. Zr/Sc diagram suggesting a reduced recycling of sedimentary material which enters into the soil composition (b), while ternary Sc-La-Th (c) and La vs Th (d) graphics illustrate the closeness of soil mineral components to the Upper Continental Crust (UCC) and the North American Shale Composite (NASC).
Figure 5. The Th/Co vs. La/Sc discriminating diagrams illustrating the felsic nature of the inorganic/mineral components of investigated soils (a), the Th/Sc vs. Zr/Sc diagram suggesting a reduced recycling of sedimentary material which enters into the soil composition (b), while ternary Sc-La-Th (c) and La vs Th (d) graphics illustrate the closeness of soil mineral components to the Upper Continental Crust (UCC) and the North American Shale Composite (NASC).
Minerals 14 00559 g005
Figure 6. PLI variation with site and depth, calculated for 13 PCEs and 45 elements.
Figure 6. PLI variation with site and depth, calculated for 13 PCEs and 45 elements.
Minerals 14 00559 g006
Figure 7. (a) The violin diagram of the distribution of PLI alone with the depth for contaminated and uncontaminated soil showing the difference between them, as well as (b) the DA showing the presence of three clusters, each of them corresponding to the investigated soil layers, i.e., 0–5, 50-20 and 20–30 cm. The partial superposition of the 0–5 and 5–20 cm clusters is visible.
Figure 7. (a) The violin diagram of the distribution of PLI alone with the depth for contaminated and uncontaminated soil showing the difference between them, as well as (b) the DA showing the presence of three clusters, each of them corresponding to the investigated soil layers, i.e., 0–5, 50-20 and 20–30 cm. The partial superposition of the 0–5 and 5–20 cm clusters is visible.
Minerals 14 00559 g007
Figure 8. The spatial distribution of the PLI (ac) and the RI (df) corresponding to the 0–5 cm layer (a,d), 5–20 cm layer (b,e), and the 20–30 cm layer (c,f).
Figure 8. The spatial distribution of the PLI (ac) and the RI (df) corresponding to the 0–5 cm layer (a,d), 5–20 cm layer (b,e), and the 20–30 cm layer (c,f).
Minerals 14 00559 g008
Figure 9. The percentage contribution of each of the ten PCEs at the total value of RI for each site and soil layer.
Figure 9. The percentage contribution of each of the ten PCEs at the total value of RI for each site and soil layer.
Minerals 14 00559 g009
Table 1. Site description.
Table 1. Site description.
CodeLongitudeLatitudeDescription
G1.145°23′00.4″27°57′46.2″TAU Vadeni, BR, rural, S of ISP
G1.245°22′22.0″27°56′40.6″TAU Vadeni, BR, rural
G1.345°23′03.7″27°54′56.8″TAU Vadeni, BR, agricultural
G2.145°25′53.8″27°55′28.4″TAU Sendreni, GL, agricultural, W of slag dump
G2.245°25′02.0″27°56′31.6″TAU Sendreni, Movileni, GL, rural, SW of slag dump
G2.345°24′18.2″27°58′26.1″GL town limit, peri-urban, S of ISP
G3.145°29′39.6″27°55′17.0″TAU Smardan, Mihail Kogalniceanu, GL, rural, N of ISP
G3.245°28′27.0″27°57′13.7″TAU Smardan, GL, agricultural, N of ISP
G4.145°25′14.1″28°01′31.3″GL town, urban, E of ISP
GC45°31′20.5″27°59′41.6″TAU Vanatori, GL, control rural site
Table 2. Elemental levels (range, average) in investigated industrial soil samples (except for the control site), norms and Upper Continental Crust (UCC) values, expressed in mg·kg−1 (except for the major elements, whose concentrations are given in g·kg−1, and Au, in µg·kg−1).
Table 2. Elemental levels (range, average) in investigated industrial soil samples (except for the control site), norms and Upper Continental Crust (UCC) values, expressed in mg·kg−1 (except for the major elements, whose concentrations are given in g·kg−1, and Au, in µg·kg−1).
ElementLiterature DataThis Work
Normal
[36]
Low; High-Alert [36]Low; High-Intervention [36]Upper Continental Crust Mean [48]Galati ISP,
min–max
Galati ISP, Average
Al, g·kg−1 79.2438–55.644.68
As515; 2525; 504.83.72–11.18.55
Au, µg·kg−1 1.53.53–69.410
Ba200400; 1000625; 2000628143–430355.6
Br 50; 100100; 3001.63.19–13.26.87
Ca, g·kg−1 24.9310.23–47.8229.26
Cd13; 55; 100.090.22–0.560.33
Ce 6328.6–80.866.65
Co1530; 10050; 25017.34.23–15.411.08
Cr30100; 300300; 6009240–12092.11
Cs 4.91.88–7.894.69
Cu20100;250200;5002813.79–54.9323.13
Dy 3.9n.d.–6.914.20
Eu 1n.d.–1.620.377
Fe, g·kg−1 38.0621.5–42.831.89
Hf 5.33.43–11.67.85
Hg0.11; 42; 100.05n.d.–0.90.13
I 1.44.82–10.66.62
K, g·kg−1 25.1416.73–24.4216.8
La 3113.6–40.933.85
Mg, g·kg−1 14.562.04–3.62.7
Mn9001500; 20002500; 4000753554–1130749.5
Mo25; 1510; 401.10.37–2.041.06
Na, g·kg−1 23.575.19–8.276.88
Nd 278.9–36.425.46
Ni2075; 200150; 5004713.5–63.737.99
Pb2050; 250100; 10001710.84–52.9029.40
Rb 8432.4–11982.53
Sb512.5; 2020; 400.40.51–1.620.95
Sc 147.31–13.910.29
Sm 4.72.54–7.676.23
Sn2035; 10050; 3002.1n.d.–9.160.89
Sr 32043–173117.6
Ta 0.90.42–1.140.925
Tb 0.70.31–0.910.725
Tm 0.30.28–3.430.88
Th 10.54.65–13.210.17
Ti, g·kg−1 3.693.19–4.763.84
U 2.71.05–3.342.651
V50100; 200200; 4009751.2–10367.25
W 1.90.84–2.742.14
Y 213.04–6.584.47
Yb 1.961.14–3.772.76
Zn100300; 700600; 15006752.6–242102.26
Zr 193105–482300.63
n.d.—not detected.
Table 3. The probabilities that the distribution of the 13 considered PCE are closer according to Tukey’s test (a, lower diagonal), Mann–Whitney (a, upper diagonal), Dunnette post hoc (b, lower diagonal), and Spearman correlation coefficient (b, upper diagonal).
Table 3. The probabilities that the distribution of the 13 considered PCE are closer according to Tukey’s test (a, lower diagonal), Mann–Whitney (a, upper diagonal), Dunnette post hoc (b, lower diagonal), and Spearman correlation coefficient (b, upper diagonal).
ab
Depth (cm)0–55–2020–30depth (cm)0–55–2020–30
0–5 0.7910.0630–5 0.8830.672
5–200.871 0.0045–200.573 0.672
20–300.0650.022 20–300.0300.006
Table 4. The ecological risk factors and risk index of the 10 considered PCEs.
Table 4. The ecological risk factors and risk index of the 10 considered PCEs.
Site CodeDepth (cm) E r i RI
HgAsPbCdCuCrZnCoNiMn
G1.10–585.7114.9010.54106.674.322.111.963.354.531.26235.35
5–2027.7116.5413.8496.673.412.051.764.456.781.18174.39
20–309.4322.087.6383.335.351.691.023.274.270.97139.03
G1.20–520.0020.088.6393.333.572.572.334.105.140.98160.73
5–20164.2914.427.5586.673.852.001.783.274.960.99289.75
20–3027.5716.487.0676.673.731.891.263.534.890.86143.93
G1.30–5121.4315.406.6880.003.952.171.412.663.260.92237.88
5–2085.7116.656.4176.679.811.931.752.753.201.09205.96
20–3033.9314.487.3373.333.291.981.213.474.340.91144.27
G2.10–5150.0016.797.19113.332.461.601.122.011.740.91297.15
5–2092.8618.987.2993.332.861.881.322.983.260.78225.54
20–3022.6418.233.1986.672.882.201.073.674.610.74145.88
G2.20–5642.8617.069.02140.003.092.391.693.013.811.37824.28
5–20300.0017.568.08130.003.792.571.413.093.541.05471.09
20–3024.0016.818.01116.673.432.091.073.505.100.97181.64
G2.30–520.0020.0415.56163.334.212.613.613.123.571.50237.56
5–20166.4317.6010.97176.674.221.592.222.923.731.08387.44
20–3019.7919.199.86166.675.191.931.383.214.091.05232.35
G3.10–520.0020.277.9986.673.962.161.523.383.570.86150.39
5–2045.2117.336.8383.333.911.971.073.324.270.83168.08
20–3028.3620.407.0376.674.082.131.003.674.340.81148.48
G3.20–5107.1423.1310.08186.674.142.201.913.584.301.14344.29
5–2085.7120.448.52170.004.662.201.583.584.311.10302.09
20–3026.0721.048.87156.674.311.861.063.385.020.90229.19
G4.10–520.007.7512.3890.003.980.860.791.221.440.94139.35
5–2092.8619.009.3483.332.661.751.582.843.030.95217.34
20–3021.3618.487.5976.676.391.711.353.124.040.76141.47
G.C.0–520.0010.293.2156.671.331.190.892.232.260.5698.63
5–20100.009.775.6053.331.530.950.752.091.820.61176.46
20–3014.078.312.6746.670.460.910.451.951.290.4577.22
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

Ene, A.; Sloată, F.; Frontasyeva, M.V.; Duliu, O.G.; Sion, A.; Gosav, S.; Persa, D. Multi-Elemental Characterization of Soils in the Vicinity of Siderurgical Industry: Levels, Depth Migration and Toxic Risk. Minerals 2024, 14, 559. https://doi.org/10.3390/min14060559

AMA Style

Ene A, Sloată F, Frontasyeva MV, Duliu OG, Sion A, Gosav S, Persa D. Multi-Elemental Characterization of Soils in the Vicinity of Siderurgical Industry: Levels, Depth Migration and Toxic Risk. Minerals. 2024; 14(6):559. https://doi.org/10.3390/min14060559

Chicago/Turabian Style

Ene, Antoaneta, Florin Sloată, Marina V. Frontasyeva, Octavian G. Duliu, Alina Sion, Steluta Gosav, and Diana Persa. 2024. "Multi-Elemental Characterization of Soils in the Vicinity of Siderurgical Industry: Levels, Depth Migration and Toxic Risk" Minerals 14, no. 6: 559. https://doi.org/10.3390/min14060559

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