Next Article in Journal
Effectiveness of Natural Language Processing Based Machine Learning in Analyzing Incident Narratives at a Mine
Next Article in Special Issue
Influence of Amino Acids on the Mobility of Iodide in Hydrocalumite
Previous Article in Journal
Characterization of the Physical, Chemical, and Adsorption Properties of Coal-Fly-Ash–Hydroxyapatite Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Contamination of the Soil–Groundwater–Crop System: Environmental Risk and Opportunities

by
Maria Economou-Eliopoulos
* and
Ifigeneia Megremi
Department of Geology and Geoenvironment, National University of Athens, GR-15784 Athens, Greece
*
Author to whom correspondence should be addressed.
Submission received: 8 June 2021 / Revised: 4 July 2021 / Accepted: 12 July 2021 / Published: 16 July 2021

Abstract

:
The increasing development of industries, resulting in a large volume of mining, smelting, and combustion wastes, and intense agricultural activities, due to demand for food and energy, have caused environmental hazards for food quality and ecosystems. This is a review on the contamination of the soil–groundwater–crop system and a potential reduction of the contamination by a gradual shift towards green economy within the European Union and on a worldwide scale. Available mineralogical and geochemical features from contaminated Neogene basins have shown a diversity in the contamination sources for soil and groundwater, and highlighted the need to define the contamination sources, hot spots, degree/extent of contamination, and provide ways to restrict the transfer of heavy metals/metalloids into the food chain, without the reduction of the agricultural and industrial production. Among harmful elements for human health and ecosystems, the contamination of groundwater (thousands of μg/L Cr(VI)) by industrial activities in many European countries is of particular attention. Although Cr(VI) can be reduced to Cr(III) and be completely attenuated in nature under appropriate pH and Eh conditions, the contamination by Cr(VI) of coastal groundwater affected by the intrusion of seawater often remains at the hundreds μg/L level. A positive trend between B and Cr(VI) may provide insights on the role of the borate [B(OH)4] ions, a potential buffer, on the stability of Cr(VI) in coastal groundwater. Efforts are needed towards reducing toxic metal(loids) from the industrial wastewaters prior to their discharge into receptors, as well as the transformation of hazardous mining/industrial wastes to new products and applications to the optimization of agricultural management strategies.

1. Introduction

During past decades, the increasing development of industries, intense use of chemicals, the large volume of mining/smelting residues, industrial wastes, and traffic in many urban areas have resulted in environmental hazards for terrestrial and aquatic ecosystems, and food quality and socioeconomic problems [1,2,3,4]. Soils provide plants with certain trace elements that, in small amounts, are considered to be critical for the healthy growth of humans, plants, and animals [5,6]. Potential sources of the soil/groundwater contamination, such as natural processes (weathering/alteration of rocks and raw materials) and human activities (mining, smelting, industrial, and agricultural activities), often result in elevated contents of harmful elements (Cr, Cu, Hg, Pb, Zn, Sb, Co, Ni, Cd, and As) [1,7,8,9,10,11]. The salinization at contaminated coastal areas, where groundwater is mixed with seawater, such as the Mediterranean, threatens plant growth [10]. Changes in agricultural production due to population growth, including expansion of agricultural land and irrigated areas and greater use of agricultural inputs (mostly fertilizers and pesticides), are used more intensively than ever before [9,12,13,14]. The increasing food demand, coupled with the intensive use of cultivated lands, have resulted in vital problems for the agricultural economy, societies, and increasing healthcare cost [8,15]. Recently, there has been significant progress on environmental protection, in particular, waste prevention and management. In this frame, the European community is focused on the transition to a circular economy (green economy) in order to make it more sustainable [16,17,18]. The investigation of raw materials in the frame of research projects, including the ProMine databases, provides information on the mineral resources for mineral deposits, mines, mining wastes, potential products, and their importance for the EU economy [19,20,21,22]. The present review study is focused on the current status and challenges on the contamination by metal(loid)s, either by natural processes (weathering of rocks, raw material deposits, salinization) and/or human activities, (mining and smelting, energy production, agricultural), with emphasis to the soil–groundwater–plants/crops system in Greece, aiming to contribute to potential ways forward to protect and better manage soil and groundwater for human health and ecosystems.

2. Methodology of Investigation

Study Area and Sampling Procedure

The Neogene Assopos and Thiva basins, with an extension of approximately 700 km2 in the former and 150 km2 in the latter, located north of the Assopos, are composed of alternations of alluvial, marls, marl limestones, and continental sediments consisting of conglomerates derived from carbonates, mafic/ultramafic rocks, and Fe-Ni laterites [23,24].
Polished sections from ultramafic rocks, Fe-Ni-laterite deposits from C. Greece and Kastoria, and bauxite processing residue or red mud were carbon coated and examined with a scanning electron microscope (SEM) using energy-dispersive spectroscopy (EDS). SEM images and EDS analyses were carried out at the University of Athens, National and Kapodistrian University of Athens (NKUA) (Department of Geology and Geoenvironment) using a JEOL JSM 5600 scanning electron microscope (JEOL, Tokyo, Japan), equipped with the ISIS 300 OXFORD automated energy-dispersive X-ray analysis system. Analytical conditions were 20 kV accelerating voltage and 0.5 nA beam current.
Representative surface (up to 20 cm) soil samples from cultivated and uncultivated areas were collected during the year 2009 and spring of 2014, covering some sites of the basin from the rhizosphere of plants. They were air-dried, crumbled mechanically, and then passed through a sieve with a 2-mm mesh. The elements Se, Cu, Pb, Fe, Mn, Zn, K, Na, Ni, Ca, Mg, P, and S were analyzed by inductively coupled plasma mass spectroscopy (ICP/MS) after aqua regia digestion at ACME Laboratories Ltd., Vancouver, BC, Canada. Platinum-group element (PGE) analyses were carried out using Ni sulfide fire-assay preconcentration technique, with the nickel fire-assay technique from large (30 g) samples at Genalysis Laboratory, Perth Services, Australia. This method allows for complete dissolution of samples. Detection limits were 1 ppb for Pd, 10 ppb for Pt, and 5 ppb Au. CDN-PGMS-23 was used as standard.
Plant samples were analyzed after cleaning and drying at 70 °C. They were powdered in an agate mortar and analyzed by inductively coupled plasma mass spectroscopy (ICP/MS), after aqua regia digestion, at the ACME Analytical Laboratories, Vancouver, BC, Canada.
Average values were calculated for groundwater samples collected from domestic and irrigation wells, and municipality of the Assopos–Thiva and Evia basins (during both wet and dry seasons), which were analyzed for many elements (Al, As, B, Ba, Cu, Zn, Fe, K, Li, Mn, Na, Ni, P, S, Se, Si, and V), hexavalent chromium, Cr(VI), and physical and chemical parameters of the water samples, including total dissolved solids (TDS). Although the analytical methods applied for the determination of the groundwater composition, including chromium stable isotopes, are provided in previous publications, a brief outline is given here. Over 200 coastal groundwater samples from domestic and irrigation wells of C. Evia and Assopos-Thiva basins, collected during the period from 2007 to 2017, were analyzed for major and trace elements by inductively coupled plasma mass spectroscopy (ICP/MS) [25,26,27,28,29,30]. Detection limits, quality control samples, and the precision of the analyses are in agreement with international standards (~10%). Physical and chemical parameters (pH and total dissolved solids) of the water samples were measured in the field using a portable Consort 561 Multiparameter Analyzer. The analyses of total chromium were performed by GFAAS (Perkin Elmer 1100B system), with an estimated detection limit of ~1 μg/L. The chemical analyses for Cr(VI) were performed by the 1,5-diphenylcarbohydrazide colorimetric method, using a HACH DR/4000 spectrophotometer. The estimated detection limit of the method was determined at ~4 μg/L.
Leaching experiments for soils, rocks, and ores were carried out in order to study the one week leaching responses of Cr under atmospheric conditions. Water and water leachates in an appropriate amount to have approximately 1 μg of Crtotal were used for the determination of the Cr-isotope composition following the method described by [31]. Both Cr concentrations and isotope ratios were analyzed using an IsotopX/GV IsoProbe T thermal ionization mass spectrometer (TIMS) equipped with eight Faraday cups at the University of Copenhagen, Denmark. Four Cr beams (50Cr+, 52Cr+, 53Cr+, and 54Cr+) were analyzed simultaneously with 49Ti+, 51V+, and 56Fe+ beams, which were used to monitor interfering ions. The final isotope composition of a sample was determined as the average of the repeated analyses and reported relative to the certified SRM 979 standard as: δ53Cr(‰) = [(53Cr/52Crsample/53Cr/52CrSRM979) − 1] × 1000.
The raw data were corrected for naturally and instrumentally induced isotope fractionation using the double spike routine. To assess the precision of the analyses, a double-spike-treated, certified standard reference material (NIST SRM 979) was used.

3. Potential Sources of Soil and Groundwater Contamination

3.1. Weathering of Rocks and Raw Materials

Arsenic (As) contamination by weathering of volcanic rocks, thermal springs, or deposits has been well described in Greece [32,33,34,35]. Recently, elevated As contents (61–210 mg/kg As) were found in a limestone quarry in Attica (Greece) that is exploited for a popular multicolor building material, and in the associated soil (33 to 430 mg/kg As) [36]. Applying a geographical information system (GIS), geostatistical techniques, and mapping software, a digitized geological–geomorphological map and an assessment of the extent and intensity of the As and other harmful elements in those lacustrine and fluvial sediments revealed a significant contamination in the uppermost travertine limestone, which may give rise to a significant risk to human health and ecosystems [36,37].
The exploitation of primary raw materials worldwide, as well as their metallurgical residues, is a potential source of soil and groundwater contamination by metal(loid)s [1,10]. Substantial volumes of groundwater, used for mining operations, may result in the oxidation and acid mine drainage in the case of sulfide ores [38,39]. The weathering of rocks and raw material deposits is a source of the transfer of potentially harmful elements into soil and shallow aquifer groundwater (chemical degradation), related to the type of the ore deposit, for example, mixed polymetallic sulfide deposits, which are related to calc-alkaline intrusions and are enriched in Pb, Zn, Fe, Cu, As, Sb, Ag, Bi, and Au, like the mines at the famous ancient Lavrion and Chalkididi Peninsula (Greece) [40,41] and the Olympias-Stratoni mines at Chalkidiki [42,43]. That type of sulfide, with a long exploitation history (mining of Ag-bearing galena at Lavrion began before 3000 BC), is characterized by the production of large volumes of metallurgical wastes transferred into alluvial soil, which has caused contamination by Pb, Cd, and As of cultivated areas and groundwater [32,44,45]. In particular, at Lavrion, contaminants like Pb are still verified in blood and deciduous teeth, and As in urine of children [46].
The major type of sulfide mineralization associated with ophiolite complexes is a massive Cyprus-type, composed mainly by pyrite and Cu-pyrite, and lesser amounts of sphalerite, characterized by an increasing content of Au, Ag, As, Se, Sb, Mo, and Hg [47,48,49]. The Cyprus-type massive sulfide ores are associated with a volcanic/subvolcanic sequence composed mainly of basalts and basaltic andesite pillow lavas, such as the Othrys, Ermioni (Peloponnesos) to Pindos, and Mirdita (Albania) ophiolites [49,50]. In general, ophiolite complexes are a type of mafic–ultramafic complex covering more than 1% of orogenetic zones, including several European countries which are strongly enriched in heavy metals, such as Cr, Fe, Mn, Co, and Mg [51,52]. Besides the sulfide mineralization, the major types of raw minerals are chromites and Fe-Ni laterites derived by weathering/alteration processes of ultramafic/mafic rocks [53,54,55]. Recently, an Ultramafic Index of Alteration (UMIA) was used, applying the equation UMIA = (Al2O3 + Fe2O3)/(Al2O3 + Fe2O3 + MgO + SiO2) × 100, based on major components in ultramafic rocks and their behavior during the weathering process; the Mg and Si are mobile and they are depleted, while Fe and Al are less mobile and enriched during the weathering process [56,57,58]. Using analytical data provided in previous publications [50,59], the calculated values (Table 1) showed a wide variation, ranging between 1.4 and 20 in saprolite and highly weathered peridotites, 18 to 24 in soils, and from 40 to 98 in laterites of Greece, throughout the contaminated C. Evia and Assopos–Thiva basins (Greece).
The best pronounced relationship between the UIMA index and Fe, Cr, Mn, and Co, which are characteristic components of ophiolites, is that with Cr (Figure 1a) and MnO (Figure 1b), especially for weathered peridotites and saprolite zones.
The positive correlation between UIMA and Cr is of particular importance for human health and ecosystems because, during weathering of ultramafic rocks, the availability of Cr increases and may be transported in soil, groundwater (drinking and irrigation), and plants/crops (food chain).
Plots of Cr, Fe, Co, and Mn versus Ni contents in contaminated soils affected by the transfer of weathered ophiolite rocks, in addition to a good positive correlation, showed an increasing trend from the Assopos basin (southern part) towards north Thiva basin to C. Evia, where ultramafic rocks are dominant, while the highest contents of these heavy metals were found in laterites (Figure 2).
In addition, elevated contents of these elements have been determined in plants/crops grown in such soils (Table 2).
The groundwater from Neogene shallow aquifers (10–100 m) throughout central Evia (Messapia), characterized by the extensive presence of ultramafic rocks and Fe-Ni-laterite deposits, contains more than 10 μg/L (up to 360 μg/L) Cr(VI) into irrigation water from Evia. Moreover, at the Assopos basin, deposits as high as 950 μg/L Cr(VI) exceed the acceptable limit for Crtotal in drinking water (50 μg/L), while up to 8000 μg/L Cr(VI) have been recorded at the area of Oinofyta, with intense industrial activity in that basin [29,60,64,65]. Although there is a wide variation in Cr concentrations throughout Evia and Assopos–Thiva basins, the Cr(VI) variations in groundwater wells during the dry and wet seasons have been recorded as relatively small [30,60]. Thus, in contrast to the lowest Cr content in the soil samples from the Assopos basin, the highest Cr concentrations in groundwater were recorded at the area of Oinofyta, Assopos basin, with an intense industrial activity [64]. Such a spatial diversity between the Cr distribution in soil and groundwater throughout the C. Evia and Assopos–Thiva basins (Figure 3) is a salient feature, suggesting their major sources of contamination.

3.2. Influence of the Groundwater Salinity

A major difference between freshwater and seawater is their salinity, expressed by the total dissolved solids (TDS), being a characteristic feature of sea and geothermal water that can be deduced from measurements of the electrical conductivity of groundwater [66]. Groundwater flowing along the land–sea margin shows that subsurface salinity distribution can extend tens to hundreds of kilometers offshore beneath continental shelves, and is often accompanied by relatively high Cr(VI) concentrations, reaching hundreds μg/L in many countries of Europe [28,29,30,59,66,67,68,69,70]. An overview of literature data on the composition of coastal groundwater and seawater has shown a positive trend between Cr(VI) concentrations with TDS (Figure 4a) and B (Figure 4b), suggesting that seawater components significantly inhibit Cr(VI) reduction into Cr(III). In addition, in faulted regions, circulated meteoric water underground can be heated by magma or hot rocks, and geothermal water ascends back to the surface, as it is known in Greece along the Hellenic Volcanic Arc, in the Chalkidiki Peninsula, in Serres, Thrace, C. Greece, other countries in Europe (Czech Republic, Hungary, and France), and worldwide [71,72,73,74,75,76,77]. Contaminated groundwater in geothermal systems is commonly characterized by relatively high TDS (up to 10,000 mg/L TDS), pH ranging from 8.5 to 5.5, and As (up to 2000 μg/L As) [78,79,80,81].
Coastal groundwater and geothermal water are similar in terms of the high salinity, while they differ in the wide pH range (5 to 7.5), much higher As concentrations, and much lower Cr concentrations in saline deep geothermal water compared to coastal and sea water [83,84,85].

3.3. Smelting/Combustion of Raw Materials

3.3.1. Bauxites and Fe-Ni-Laterites

Currently, applying the Bayer process, aluminum is extracted from bauxite ore by the production of high-grade metallurgical alumina and, subsequently, the electrolytic reduction of alumina to aluminum [86]. Residues of the metallurgical process of bauxite or red mud are relatively toxic due to its high alkalinity (pH 10–12.5) and the presence of heavy and radioactive metals (V, Cr, Cd, Ni, Zn, Pb, Ba, Sr, Hf, Nd, U, Th) [2,65,86]. The bauxite deposit at the Parnassos–Ghiona mountains of Greece is the 11th largest bauxite producer in the world [87]. Alumina (Al2O3) has been produced for more than 100 years, and up to 120 million tons of bauxite processing residue or red mud are produced annually during Al extraction via the Bayer process, causing critical environmental problems [88,89]. Unexpectedly high (2100 μg/L) Cr(VI) concentrations have been measured in water leachates from a representative red mud sample from the Aspra spitia plant, Greece (May 2008, mining company S & B Industrial Minerals) [65].
The Fe-Ni laterite deposits are a major source of Ni in Greece (Lokris, C. Evia, and Kastoria), mined by Larco, and produce some 2–3% of the world’s total, in Serbia, Turkey, Albania (Bitinca), and elsewhere [90]. Although elevated levels of Cr have been identified in Ni laterites, Cr(VI) derived from that source, and its availability and contribution to natural waters remain poorly understood. Specifically, it is well known that the Fe-Ni-laterite ores from Kastoria are characterized by higher Cr, Co, and Mn, and lower Al2O3 and TiO2 contents compared to those from the Lokris and Evia [50,91]. Water leaching (one week), using Cr-free natural water, has shown that leachates for the Kastoria laterite samples from the lower goethite zone, containing considerable amounts of Mn from Kastoria (Figure 1b), reached levels up to 1300 μg/L Cr(VI), in contrast to those from the Lokris and Evia Fe-Ni laterites with less than 10 μg/L [65].

3.3.2. Chromite Deposits and Ferrochromium (FeCr)

Ferrochromium (FeCr), which is a critical alloy in the production of stainless steel, is produced in several European countries (mainly in Finland, France, Italy, Norway, Sweden and former Yugoslavia, Germany, Italy, Switzerland, and the U.K.) and is a potential source of soil and groundwater aquifer contamination by Cr(VI). Specifically, chromite ore processing residue (COPR), a waste byproduct during the chromate production that is generated in the U.S. and European countries, is highly alkaline (pH > 12) and contains high levels of Cr(VI) under the form of water-soluble sodium chromate (Na2CrO4) [92,93,94].

3.3.3. Coal Mining/Combustion (Fly Ash)

Τwo main lignite-mining districts, at the Ptolemais–Amynteon (N. Greece) and Megalopolis in the Peloponnese (S. Greece), are in production, currently [95,96]. Mineralogical and geochemical data on fly and bottom ash from the lignite combustion have shown that many trace elements, including Cr, Cu, Pb, Zn, Cd, Mo, Ag, As, Au, Se, Te, U, and W, are associated with minerals in the peat and the organic matter [97,98,99]. In the Sarigkiol basin, a sub-basin of the Ptolemais basin, in addition to the influence from the open pit of the nearby lignite mines and fly ash produced by a power station, there is a contribution of heavy metals due to the presence of weathered ultramafic rocks and agricultural activities [100,101].

4. Discussion

Global climate changes (air quality, sea level), the increasing population, and demand for food and energy have resulted in intense agricultural activities and exploitation of raw material deposits to supply and use metals in high-tech applications and have caused environmental problems [12,102]. In general, the contamination of soil and groundwater by metal/metalloids related to either natural processes (weathering of rocks and ore deposits, salinization) and/or human activities (industrial, mining, smelting, extensive use of fertilizers and pesticides in the agricultural activities, urban pollution) are critical problems for the water quality and food chain [12,14,103]. The consequence of plant/crop irrigation with contaminated water and bioaccumulation of toxic elements by plants (transfer into food chain) are potential threats for human health and ecosystems [1,6,9,15,104].

4.1. Implications of the Green Economy

Recently, there has been significant progress on industrial waste prevention and management at a worldwide scale, including the European community, on the transition to a circular economy (green economy) in order to make it more sustainable [16,17,18]. The investigation of raw materials in the frame of research projects, including the ProMine databases, provides valuable information in a geographic information system (GIS) for the distribution of the major raw material deposits, mining wastes, potential products, and their importance for the economy in all European countries, as well as the creation of national and European cultivation strategies for the optimization of agricultural management strategies [19,21,22,105].

4.1.1. Smelting Residues (Red Mud and Slag)

With respect to the hundreds of million tons of bauxite processing toxic red mud, produced during Al extraction via the Bayer process, several applications have been proposed, such as a source for the recovery of critical metals and ceramics [106,107,108]. A new application proposed the use of the red mud as a component in cements, in building materials, or synthesis of inorganic polymers with excellent properties [17]. Thus, the Bayer process, causing critical environmental problems, can be effectively transformed into a zero-waste process [17]. On the basis of Cr Kedge XANES and TEM analysis of the speciation of chromium in red mud samples, it has been suggested that Cr in red mud is mostly present as Cr(III), substituted into hematite and goethite [2,109]. However, unexpectedly high Cr(VI) concentrations (2100 μg/L) have been measured in water leachates of red mud [65], suggesting that the reduction of the oxidation stage of the Cr(VI) is required before the applications of red mud. Similarly, the reduction of Cr(VI) in the Cr(V)-bearing wastes during the production of FeCr from chromite [92,93,94] is necessary.
The reductive smelting plant for the production of ferronickel (FeNi) from Fe-Ni-laterites of Central Greece (Lokris and Evia) and Kastoria at Larymna (Greece) allows the formation of one metallic phase containing all of the nickel and part of the iron and one slag residue [110]. The residue of the metallurgical process (slag) contains a glassy phase with inclusions of spherical Fe-Ni alloy of varying size [65]. Water leaching experiments on Fe-Ni laterite ores from Lokris and Kastoria deposits and metallurgical residues (slag) have shown that the Cr(VI) concentrations in leachates for the slag and the Lokris Fe-Ni laterites are very low (2 to <1 μg/L), but those from the Kastoria laterites reach as high as 1200 μg/L Cr(VI) concentrations [65]. Therefore, the release of Cr(VI) from the use of that type of slag is not facilitated under ambient conditions, but examination of potential Cr(VI) leaching during mining is required.

4.1.2. Bioaccumulation of Harmful Metal(loids)

Special attention has been paid to the soil–groundwater–plant/crop system, especially to the accumulation or transfer factor, which is the ratio of the element content in plants relative to the total element content in the relative soil, showing a wide variation as a function of the oxidation state of metals/metalloids in soil and irrigation water, the content of organic matter, and the pH and Eh parameters [111,112]. Review of the available database/literature highlights the importance of water quality on human health/ecosystems and the need to differentiate between Cr(III) and Cr(VI) discharge limits to soil and/or drinking/irrigation water for all European countries [113]. Mineralogical and geochemical data on fly and bottom ash from the lignite combustion in northern Greece, coupled with leaching experiments, suggested that applying these materials on agricultural land via the addition of a small proportion of 5 wt% fly ash in acid soil can increase the pH to 7.3–7.8, and the Ca, Mg, K, and Na content of the soil [98,99]. Although these elements can be classified as essential nutrients, the coexisting trace elements, such as As, Cr, Cu, Ni, Zn, and Hg, are considered to have some environmental or public health impacts, including potential groundwater contamination, and should be used with caution for agricultural applications [114]. In addition, organic carbon may play a key role in the mobility of added Cr(VI) to soil via irrigation water [115]. The use of leonardite, which is an oxidized form of lignite, due to its high content of humic acid, is considered to be a useful organic fertilizer that provides possibilities for combining food production with soil [116,117]. Since Europe (Germany, Greece, Poland, and the Czech Republic) accounts for roughly 40% of global lignite reserves [118], the application of the natural organic material, leonardite, as a land management technique seems to be a cost-effective method consistent to related protocols for the protection of the soil quality. In addition, several land uses in the exploited lignite quarries have been proposed in the frame of the circular economy, such as agriculture activities, greenhouses, livestock farming, forests, and photovoltaic parks, which can contribute to the energy economy cycle [96].
It is well known that selenium (Se) is an essential micronutrient for humans and animals, but both Se excess and deficiency can cause various health risks. The potential presence of selenates (SeO42−), due to industrial activities, mines, or smelting plants, and the Se level in coastal groundwater affected by the intrusion of seawater in costal groundwater, occasionally exceeding the guidelines [103] for Se (10μg/L) in drinking water [27,28], may cause contamination problems in cultivated plants/crops [119,120]. However, soils commonly exhibit Se deficiency, and people have insufficient levels of Se-serum [121,122]. More research on the Se distribution in crops/plants and relative soil coming from their rhizosphere is required to define potential Se source(s), factors controlling Se bioaccumulation, and ways for its enrichment in food.

4.2. Application of the Stable Chromium Isotopes to the Groundwater Contamination by (Cr(VI)

Oxidative weathering and the presence of Mn oxides as a catalyst on the surface of Cr-bearing ultramafic rocks facilitate the oxidation of Cr(III) into water-soluble Cr(VI) [123], eventually being deposited in the seabed sediments [124,125,126,127,128]. Subsequently, during the reduction of the Cr(VI) to Cr(III), the lighter chromium isotopes (52Cr) are preferentially reduced, resulting in an enrichment of 53Cr relative to 52Cr values (δ53Cr values in units per mil (‰) relative to a standard) [31] and redox processes produce significant Cr isotope fractionation [129]. Moreover, Cr(III) forms in igneous rocks with an average δ53Cr value of −0.124 ± 0.101‰ [130] can be oxidized to Cr(VI) in the presence of Mn oxides (in oxidation stage +4) [51,131]. Contaminated groundwater (4500 μg/L) by industrial activities (1997) at the area of Friuli Venezia Giulia (northern Italy) was completely attenuated, subsequently (2003), under pH conditions of 7.5–8.2 and Eh 0.36–0.41V [132]. However, the contamination by Cr(VI) of coastal groundwater in the Assopos basin remains at the hundreds μg/L level [65,82]. Given the lack of seawater effect in groundwater in the case of Friuli Venezia Giulia Region, northern Italy, as is exemplified by the relatively low TDS and Na concentrations in those aquifers [133], it seems likely that the reduction of Cr(VI) to Cr(III) is not inhibited in that case. In contrast, in the coastal groundwater from C. Evia and Assopos basins, the presence of borate [B(OH)4] ions (Figure 4a,b) [134,135], a potential buffer, may indicate that the reduction of Cr(VI) to Cr(III) is inhibited [136]. In addition, there is a difference between the Kastoria Fe-Ni laterites leaching 1300 μg/L Cr(VI) and showing δ53Cr values ranging between +0.03 ±0.06‰ and −0.21± 0.08‰, and the laterite from Lokris leaching <2 μm/L Cr(VI) and showing δ53Cr values of +1.01 ± 0.05‰ (Figure 5). The Cr stable isotope data of soil profiles developed on serpentinized, chromite-bearing ultramafic rocks at La Cabana have shown negative δ53Cr values (−0.089 to −0.320 ‰, average of −0.178‰) [137]. According to these authors, the degree of weathering of the La Cabana soils corresponds to an early to intermediate stage of weathering compared to the traditional tropical laterite deposits.
Thus, available analytical data on groundwater contamination by Cr(VI) (either by natural processes and/or industrial activities) show a well-pronounced negative trend between chromium isotopes (δ53Cr values) versus Cr(VI) concentrations, suggesting that the dominant cause of Cr isotope fractionation is post-mobilization reduction of Cr(VI) [82,132,138,141,142]. However, the abundance of Cr(VI) in coastal groundwater may be related to the presence of [B](OH)4] ions, which significantly inhibit Cr(VI) reduction to Cr(III).

4.3. Knowledge Gaps

The consequences of the cultivation of plants/crops in contaminated soil and their irrigation with contaminated water, resulting in the bioaccumulation of toxic elements (transfer into food chain), are important threats for human health and ecosystems [1,6,9,15,104]. A review of the available database/literature highlights the need for the protection of soil and water quality. Although many cities in Europe have been geochemically mapped [143], application of GIS (geographical information system) and risk analysis (mapping) for groundwater and soil geochemistry for every city in Europe is required in order to define contamination/pollution sources, the presence of hotspots, and the degree and extent of groundwater contamination. Efforts need to assess heavy metal and/or metalloid contamination, and the creation of national and European cultivation strategies are needed in the cultivated land. Free or low-cost access for everybody to the information obtained could facilitate optimization of agricultural management strategies.
According to recent legislation [113], Cr discharge to the aquatic environment is regulated by EU member states at maximum discharge limits of 1 mg/L Cr(VI) and 5 mg/L Crtotal for EU member states [144]. Efforts need to differentiate between Cr(III) and Cr(VI) discharge limits to soil and/or aquatic environments, as well as in the drinking water for all European countries.
Recently, there has been significant progress on industrial waste prevention and management on a worldwide scale, including the European community on the transition to a circular economy (green economy), making it more sustainable [16,17,18]. However, further research on the transformation of mining, smelting, and combustion wastes to new products and the creation of cultivation strategies for the optimization of agricultural management strategies [19,21,22] are required.

5. Conclusions

Natural processes (weathering of rocks and ore), the intense exploitation of raw material deposits to supply metals in high-tech applications, and agricultural activities have caused soil and groundwater contamination by heavy metals and metalloids, which, in turn, are transferred into the food chain. The present review of literature data suggests that, despite the environmental risk, potential opportunities can contribute towards the development of new products and new ways for groundwater management and land use:
Efforts need to provide ways to restrict the transfer of heavy metals/metalloids from soil to plants/crops and groundwater, without the reduction of agricultural and industrial production.
Mining wastes from the mining of coal/lignite (leonardite), and combustion byproducts (fly ash) may be used to optimize agricultural production.
Efforts are needed towards reducing Cr(VI) from industrial wastewaters, prior to their discharge into receptors.
Soil and groundwater contamination, due to mining, smelting, or combustion, can be reduced by applying transformations of harmful wastes into new products with potential applications towards a circular economy.
At the abandoned mines, quarries, and surroundings, changes in land use may allow the development of a new land management and the production of green energy.

Author Contributions

Both authors, M.E.-E. and I.M., contribute to the conceptualization, methodology, software and validation of data, writing—original draft preparation. All authors have read and agreed to the published version of the manuscript.

Funding

The Municipality of Oropos, Greece (Grand No AK 70/3/9997) and the University of Athens (Grant No ΚΕ 11078).

Acknowledgments

Many thanks are expressed once again to the University of Athens, the Municipalities of Oropos and of Dirfys-Messapia (Greece) for the financial support of the assessment of soil and groundwater contamination in Greece, and to the anonymous reviewers for their criticism and suggestions on an earlier draft of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Panagos, P.; van Liedekerke, M.; Jones, A.; Montanarella, L. European Soil Data Centre: Response to European policy support and public data requirements. Land Use Policy 2012, 29, 329–338. [Google Scholar] [CrossRef]
  2. Burke, I.T.; Mayes, W.M.; Peacock, C.L.; Brown, A.P.; Jarvis, A.P.; Gruiz, K. Speciation of Arsenic, Chromium, and Vanadium in Red Mud Samples from the Ajka Spill Site, Hungary. Environ. Sci. Technol. 2012, 46, 3085–3092. [Google Scholar] [CrossRef]
  3. Balderacchi, M.; Filippini, M.; Gemitzi, A.; Klöve, B.; Petitta, M.; Trevisan, M.; Wachniew, P.; Witczak, S.; Gargini, A. Does groundwater protection in Europe require new EU-wide environmental quality standards? Front. Chem. 2014, 2, 32. [Google Scholar] [CrossRef] [Green Version]
  4. Tiwari, A.K.; Orioli, S.; De Maio, M. Assessment of groundwater geochemistry and diffusion of hexavalent chromium contamination in an industrial town of Italy. J. Contam. Hydrol. 2019, 225, 103503. [Google Scholar] [CrossRef] [PubMed]
  5. Losi, E.; Amrhein, C.; Frankenberger, W.T.J. Environmental biochemistry of chromium. Rev. Environ. Contam. Toxicol. 1994, 136, 91–131. [Google Scholar]
  6. Massoud, M.A.; Scrimshaw, M.D.; Lester, J.N. Qualitative assessment of the effectiveness of the Mediterranean action plan: Wastewater management in the Mediterranean Geofluids 3 region. Ocean Coast. Manag. 2003, 46, 875–899. [Google Scholar] [CrossRef]
  7. ATSDR (Agency for Toxic Substances, Disease Registry). Toxicological Profile for Chromium; Agency for Toxic Substances and Disease Registry, Public Health Service, U.S. Department of Health and Human Services: Washington, DC, USA, 2000. [Google Scholar]
  8. Fawell, J.; Nieuwenhuijsen, M.J. Contaminants in drinking water: Environmental pollution and health. Br. Med. Bull. 2003, 68, 199–208. [Google Scholar] [CrossRef]
  9. Welsh Government. River Basin Districts Surface Water and Groundwater Classification (Water Framework Directive) (England and Wales) Direction 2009 (2009 No.53). Available online: https://gov.wales/sites/default/files/publications/2019-07/091222direct53.pdf (accessed on 30 May 2021).
  10. Libutti, A.; Monteleone, M. Soil vs. groundwater: The quality dilemma. Managing nitrogen leaching and salinity control under irrigated agriculture in Mediterranean conditions. Agric. Water Manag. 2017, 186, 40–50. [Google Scholar] [CrossRef]
  11. Wang, X.; Li, L.; Yan, X.; Meng, X.; Chen, Y. Processes of chromium (VI) migration and transformation in chromate production site: A case study from the middle of China. Chemosphere 2020, 257, 127282. [Google Scholar] [CrossRef] [PubMed]
  12. USEPA. Literature Review of Contaminants in Livestock and Poultry Manure and Implications for Water Quality; USEPA: Washington, DC, USA, 2013. [Google Scholar]
  13. Keesstra, S.D.; Bouma, J.; Wallinga, J.; Tittonell, P.; Smith, P.; Cerdà, A.; Montanarella, L.; Quinton, J.N.; Pachepsky, Y.; van der Putten, W.H.; et al. The significance of soils and soil science towards realization of the United Nations Sustainable Development Goals. Soil 2016, 2, 111–128. [Google Scholar] [CrossRef] [Green Version]
  14. Mateo-Sagasta, J.; Marjani, S.; Turral, H.; Burke, J. Water Pollution from Agriculture: A Global Review Executive Summary; FAO IWMI: Rome, Italy, 2017; pp. 1–35. [Google Scholar]
  15. Van der Putten, W.; Ramirez, K.; Poesen, J.; Lenka, L.; Šimek, M.; Anne, W.; Mari, M.; Philippe, L.; Heikki, S.; Andrey, Z.; et al. Opportunities for Soil Sustainability in Europe; European Academies’ Science Advisory Council: Halle, Germany, 2018; p. 41. [Google Scholar]
  16. Loiseau, E.; Saikku, L.; Antikainen, R.; Droste, N.; Hansjürgens, B.; Pitkänen, K.; Leskinen, P.; Kuikman, P.; Thomsen, M. Green economy and related concepts: An overview. J. Clean. Prod. 2016, 139, 361–371. [Google Scholar] [CrossRef]
  17. Hertel, T.; Blanpain, B.; Pontikes, Y. A Proposal for a 100 % Use of Bauxite Residue Towards Inorganic Polymer Mortar. J. Sustain. Metall. 2016, 2, 394–404. [Google Scholar] [CrossRef] [Green Version]
  18. European Commission. Critical Raw Materials Resilience: Charting a Path towards Greater Security and Sustainability. Communication from the Commission to the European Parliament, the Council, the European Economic and Social Committee 2 and the Committee of the Regions COM(2020) 474 Final; European Commission: Brussels, Belgium, 2020. [Google Scholar]
  19. Cassard, D.; Bertrand, G.; Billa, M.; Serrano, J.J.; Tourlière, B.; Angel, J.M.; Gaál, G. ProMine Mineral Databases: New Tools to Assess Primary and Secondary Mineral Resources in Europe. In 3D, 4D and Predictive Modelling of Major Mineral Belts in Europe; Weihed, P., Ed.; Springer: Cham, Switzerland, 2015. [Google Scholar]
  20. Cassard, D.; Tertre, F. Improvement of KDPs’ applications and interaction with the RMIS and the GeoERA Information Platform. 2021. Available online: https://geoera.eu/wp-content/uploads/2021/03/D5.5-KDP%E2%80%99s-applications-delivery-to-RMIS.pdf (accessed on 30 May 2021).
  21. EUROPE. A Strategy for Smart, Sustainable and Inclusive Growth. 2020. Available online: https://ec.europa.eu/eu2020/pdf/COMPLET%20EN%20BARROSO%20%20%20007%20-%20Europe%202020%20-%20EN%20version.pdf (accessed on 30 May 2021).
  22. Mintell4EU. Mineral Intelligence for Europe. Available online: https://geoera.eu/projects/mintell4eu7/ (accessed on 30 May 2021).
  23. Papanikolaou, D.; Lykoysis, V.; Chronis, G.; Pavlakis, P. A comparative study of neotectonic basins across the Hellenic arc: The Messiniakos, Argolikos, Saronikos and Southern Evoikos gulfs. Basin Res. 1988, 1, 167–176. [Google Scholar] [CrossRef]
  24. Chatoupis, T.; Fountoulis, I. The neotectonic deformation of N. Parnis Mt, (Attica, Greece). Bull. Geol. Soc. Greece 2004, 36, 1588–1597. [Google Scholar]
  25. Giannoulopoulos, P. Hydrogeological-Hydro-Chemical Survey of Groundwater Quality in the Wider Area of Assopos River Basin, Viotia Prefecture; Institute of Geology and Mineral Exploration (IGME): Athens, Greece, 2008; pp. 1–74. (In Greek) [Google Scholar]
  26. Vasilatos, C.; Megremi, I.; Economou-Eliopoulos, M.; Mitsis, I. Hexavalent chromium and other toxic elements in natural waters in the Thiva-Tanagra-Malakasa Basin, Greece. Hell. J. Geosci. 2008, 43, 57–66. [Google Scholar]
  27. Moraki, A. Assessment of groundwater contamination by hexavalent chromium and its remediation at Avlida area, Central Greece. Hell. J. Geosci. 2010, 45, 175–183. [Google Scholar]
  28. Pavlopoulos, K.; Chrisanthaki, I.; Economou-Eliopoulos, M.; Lekkas, S. Hydrochemical study of metals in the groundwater of the wider area of Koropi. In Advances in the Research of Aquatic Environment; Springer: Berlin/Heidelberg, Germany, 2011; pp. 169–176. [Google Scholar]
  29. Megremi, I.; Vasilatos, C.; Atsarou, A.; Theodoratou, C.; Economou-Eliopoulos, M.; Mitsis, I. Geochemical evidences for the sources of the Cr(VI) contamination in groundwater in central Euboea and Assopos-Thiva basins, Greece. Natural versus anthropogenic origin. Eur. Water 2013, 41, 23–34. [Google Scholar]
  30. Megremi, I.; Vasilatos, C.; Vassilakis, E.; Economou-Eliopoulos, M. Spatial diversity of Cr distribution in soil and groundwater sites in relation with land use management in a Mediterranean region: The case of C. In Evia and Assopos-Thiva basins, Greece. Sci. Total Environ. 2019, 651, 656–667. [Google Scholar] [CrossRef]
  31. Ellis, A.S.; Johnson, T.M.; Bullen, T.D. Chromium isotopes and the fate of hexavalent chromium in the environment. Science 2002, 295, 2060–2062. [Google Scholar] [CrossRef] [Green Version]
  32. Demetriades, A.; Stavrakis, P.; Vergou-Vichou, K. Contamination of surface soil of the Lavreotiki peninsula (Attiki, Greece) by mining and smelting activities. Min. Wealth 1996, 98, 7–15. [Google Scholar]
  33. Korre, A.; Durucan, S.; Koutroumani, A. Quantitative-spatial assessment of the risks associated with high Pb loads in soils around Lavrio, Greece. Appl. Geochem. 2002, 17, 1029–1045. [Google Scholar] [CrossRef]
  34. Komnitsas, K.; Modis, K. Soil risk assessment of As and Zn contamination in a coal mining region using geostatistics. Sci. Total Environ. 2006, 371, 190–196. [Google Scholar] [CrossRef]
  35. Gamaletsos, P.; Godelitsas, A.; Dotsika, E.; Tzamos, E.; Göttlicher, J.; Filippidis, A. Geological sources of as in the environment of Greece: A review. Threat. Qual. Groundw. Resour. 2013, 77–113. [Google Scholar] [CrossRef]
  36. Kampouroglou, E.; Economou-Eliopoulos, M. Natural contamination by As and heavy metals in soil, their bio-accumulation and potential sources: The case of a travertine limestone quarry, Greece. Cent. Eur. J. Geosci. 2013, 5, 174–188. [Google Scholar] [CrossRef] [Green Version]
  37. Kampouroglou, E.E.; Economou-Eliopoulos, M. Assessment of arsenic and associated metals in the soil-plant-water system in Neogene basins of Attica, Greece. Catena 2017, 150, 206–222. [Google Scholar] [CrossRef]
  38. Morris, B.L.; Lawrence, A.R.; Chilton, P.J.; Adams, B.; Calow, R.; Klinck, B.A. Groundwater and Its Susceptibility to Degradation: A Global Assessment of the Problems and Options for Management. Early Warning and Assessment Report Series, RS, 03–3; United Nations Environment Programme: Nairobi, Kenya, 2003. [Google Scholar]
  39. Schmoll, O.; Howard, G.; Chilton, P.J.; Chorus, I. (Eds.) Protecting Groundwater for Health: Managing the Quality of Drinking Water Sources; WHO/IWA: London, UK, 2006. [Google Scholar]
  40. Marinos, G.; Petrascheck, W.E. Lavrion. Institute for geology and subsurface research. Geol. Geophys. Res. 1956, 4, 1–247. [Google Scholar]
  41. Skarpelis, N.; Ardyraki, A. Geology and origin of supergene ore at the Lavrion Pb-Ag-Zn deposit, Attica, Greece. Resour. Geol. 2008, 59, 1–14. [Google Scholar] [CrossRef]
  42. Kalogeropoulos, S.I.; Kilias, S.P.; Bitzios, D.C.; Nicolaou, M.; Both, R.A. Genesis of the Olympias carbonate-hosted Pb-Zn (Au, Ag) sulfide ore deposit, eastern Chalkidiki peninsula, northern Greece. Econ. Geol. 1989, 84, 1210–1234. [Google Scholar] [CrossRef]
  43. Eldorado Gold Corporation. Resources and Reserves. Available online: www.eldoradogold.com/assets/resources-and-reserves (accessed on 29 October 2017).
  44. Conophagos, E.C. Le Laurium Antique et la Technique Grecque de la Production de L’argen; National Technical University: Athens, Greece, 1980; p. 458. (In French) [Google Scholar]
  45. Stamatis, G.; Lambrakis, N.; Alexakis, D.; Zagana, E. Groundwater quality in Mesogea basin in eastern Attica (Greece). Hydrol. Process 2006, 20, 2803–2818. [Google Scholar]
  46. Maravelias, C.; Hatzakis, A.; Katsouyanni, K.; Trichopoulos, D.; Koutselinis, A.; Ewers, U.; Brockhaus, A. Exposure to lead and cadmium of children living near a lead smelter at Lavrion, Greece. Sci. Total Environ. 1989, 84, 61–70. [Google Scholar] [CrossRef]
  47. Hannington, M.; Herzig, P.; Scott, S.; Thompson, G.; Rona, P. Comparative mineralogy and geochemistry of gold-bearing sulfide deposits on the mid-ocean ridges. Mar. Geol. 1991, 101, 217–248. [Google Scholar] [CrossRef]
  48. Franklin, J.M.; Sangster, D.M.; Lydon, J.W. Volcanic-associated massive sulfide deposits. Econ. Geol. 1991, 75, 485–627. [Google Scholar]
  49. Economou-Eliopoulos, M.; Eliopoulos, D.G.; Chryssoulis, S. A comparison of high-Au massive sulfide ores hosted in ophiolite complexes of the Balkan Peninsula with modern analogues: Genetic significance. Ore Geol. Rev. 2008, 33, 81–100. [Google Scholar] [CrossRef]
  50. Eliopoulos, D.; Economou-Eliopoulos, M. Geochemical and mineralogical characteristics of Fe-Ni and bauxitic-laterite deposits of Greece. Ore Geol. Rev. 2000, 16, 41–58. [Google Scholar] [CrossRef]
  51. Oze, C.; Bird, K.D.; Fendorf, S. Genesis of hexavalent chromium from natural sources in soil and groundwater. Proc. Natl. Acad. Sci. USA 2007, 104, 6544–6549. [Google Scholar] [CrossRef] [Green Version]
  52. Vithanage, M.; Kumarathilaka, P.; Oze, C.; Karunatilake, S.; Seneviratne, M.; Hseu, Z.Y.; Gunarathne, V.; Dassanayake, M.; Ok, Y.S.; Rinklebe, J. Occurrence and cycling of trace elements in ultramafic soils and their impacts on human health: A critical review. Environ. Int. 2019, 131, 104974. [Google Scholar] [CrossRef] [PubMed]
  53. Golightly, J.P. Progress in Understanding the Evolution of Nickel Laterites. In The Challenge of Finding New Mineral Resources: Global Metallogeny, Innovative Exploration, and New Discoveries; Richard, J.G., Erin, E.M., Thomas, M., Eds.; Special Publication of the Society of Economic Gelogists: Littleton, CO, USA, 2010; Volume 15. [Google Scholar]
  54. Valeton, I.; Biermann, M.; Reche, R.; Rosenberg, F. Genesis of nickel laterites and bauxites in Greece during the Jurassic and Cretaceous, and their relation to ultrabasic parent rocks. Ore Geol. Rev. 1987, 2, 359–404. [Google Scholar] [CrossRef]
  55. Alevizos, G. Mineralogy, Geochemistry and Origin of the Sedimentary Fe-Ni Ores of Lokris. Ph.D. Thesis, Technical University, Crete, Greece, 1997; p. 245. [Google Scholar]
  56. Babechuk, M.; Widdowson, M.; Kamber, B.S. Quantifying chemical weathering intensity and trace element release from two contrasting basalt profiles, Deccan Traps, India. Chem. Geol. 2014, 363, 56–75. [Google Scholar] [CrossRef]
  57. Aiglsperger, T.; Proenza, J.A.; Lewis, J.F.; Labrador, M.; Svojtka, M.; Rojas-Purón, A.; Longo, F.; Ďurišová, J. Critical metals (REE, Sc, PGE) in Ni laterites from Cuba and the Dominican Republic. Ore Geol. Rev. 2016, 73, 127–147. [Google Scholar] [CrossRef]
  58. Bolaños-Benítez, V.; van Hullebusch, E.D.; Birck, J.L.; Garnier, J.; Lens, P.N.L.; Tharaud, M.; Quantin, C.; Sivry, Y. Chromium mobility in ultramafic areas affected by mining activities in Barro Alto massif, Brazil: An isotopic study. Chem. Geol. 2021, 561, 120000. [Google Scholar] [CrossRef]
  59. Papazotos, P.; Vasileiou, E.; Perraki, M. The synergistic role of agricultural activities in groundwater quality in ultramafic environments: The case of the Psachna basin, Central Euboea. Greece. Environ. Monit. Assess. 2019, 191, 317. [Google Scholar] [CrossRef] [PubMed]
  60. Megremi, I. Controlling Factors of the Mobility and Bioavailability of Cr and Other Metals at the Environment of Ni-Laterites. Ph.D Thesis, University of Athens, Athens, Greece, 2010; p. 316. [Google Scholar]
  61. Atsarou, C. Distribution of Chromium and Other Heavy Metals in Groundwater, Soil and Crops at the Avlona Area, Attica: Factors Controlling Their Bioavailability. Master’s Thesis, University of Athens, Athens, Greece, 2011; p. 204. (In Greek). [Google Scholar]
  62. Theodoratou, C. Estimation of Contamination Due to Hexavalent Chromium and Other Heavy Metals in the Oropos Area: Interaction in the Soil-Plant-Water System. Master’s Thesis, University of Athens, Athens, Greece, 2011; p. 198. (In Greek). [Google Scholar]
  63. Economou-Eliopoulos, M.; Antivachi, D.; Vasilatos, C.; Megremi, I. Evaluation of the Cr(VI) and other toxic element contamination and their potential sources: The case of the Thiva basin (Greece). Geosci. Front. 2012, 3, 523–539. [Google Scholar] [CrossRef] [Green Version]
  64. Michalakis, I.B. Hydrogeological Research in the Industrial Plain of ELVAL Ltd. (Oinofita, Viotia): Implication to the Contamination Source Report; Institute of Geology and Mineral Exploration (IGME): Athens, Greece, 2015; pp. 1–161. (In Greek) [Google Scholar]
  65. Economou-Eliopoulos, M.; Frei, R.; Megremi, I. Potential leaching of Cr(VI) from laterite mines and residues of metallurgical products (red mud and slag): An integrated approach. J. Geochem. Explor. 2016, 262, 40–49. [Google Scholar] [CrossRef]
  66. Post, V.E.A.; Eichholz, M.; Brentführer, R. Groundwater Management in Coastal Zones; Bundesanstalt für Geowissenschaften und Rohstoffe (BGR): Hannover, Germany, 2018; p. 107. [Google Scholar]
  67. Rose, W.; Hawkes, E.; Webb, S. Geochemistry in Mineral Exploration, 2nd ed.; Academic Press: New York, NY, USA, 1979. [Google Scholar]
  68. Fantoni, D.; Brozzo, G.; Canepa, M.; Cipolli, F. Natural hexavalent chromium in groundwaters interacting with ophiolitic rocks. Environ. Geol. 2002, 42, 871–882. [Google Scholar] [CrossRef]
  69. Purnama, S.; Marfai, M.A. Saline Water Intrusion Toward Groundwater: Issues and Its Control. J. Nat. Resour. Dev. 2012, 2, 25–32. [Google Scholar]
  70. DiGiulio, D.C.; Jackson, R.B. Impact to underground sources of drinking water and domestic wells from production well stimulation and completion practices in the Pavillion, Wyoming, field. Environ. Sci. Technol. 2016, 50, 4524–4536. [Google Scholar] [CrossRef]
  71. Lund, J.W. Geothermal Spas in Czechoslovakia. Geo-Heat Center Quart. Bull. 1990, 12, 20–24. [Google Scholar]
  72. Fytikas, M.; Andritsos, N.; Karydakis, G.; Kolios, N.; Mendrinos, D.; Papachristou, M. Geothermal exploration and development activities in Greece during 1995–1999. In Proceedings of the World Geothermal Congress, Kyushu-Tohoku, Japan, 28 May–10 June 2000; pp. 199–208. [Google Scholar]
  73. Smedley, P.L.; Kinniburgh, D.G. A review of the source, behaviour and distribution of arsenic in natural waters. Appl. Geochem. 2002, 17, 517–568. [Google Scholar] [CrossRef] [Green Version]
  74. Tyrovola, K.; Nikolaidis, N.P.; Veranis, N.; Kallithrakas-Kontos, N.; Koulouridakis, P.E. Arsenic removal from geothermal waters with zero-valent iron—Effect of temperature, phosphate and nitrate. Water Res. 2006, 40, 2375–2386. [Google Scholar] [CrossRef]
  75. Lund, J.W.; Freeston, D.H.; Boyd, T.L. Direct utilization of geothermal energy 2010 worldwide review. In Proceedings of the World Geothermal Congress 2010 (CD-ROM), Bali, Indonesia, 25–29 April 2010. [Google Scholar]
  76. Katsoyiannis, I.A.; Mitrakas, A.; Zouboulis, A.I. Arsenic occurrence in Europe: Emphasis in Greece and description of the applied full-scale treatment plants. Desalination Water Treat. 2015, 54, 2100–2107. [Google Scholar] [CrossRef]
  77. Kanellopoulos, C. Influence of ultramafic rocks and hot springs with travertine depositions on geochemical composition and baseline of soils. Application to eastern central Greece. Geoderma 2020, 380, 114649. [Google Scholar] [CrossRef]
  78. Wilkie, J.A.; Hering, J.G. Rapid oxidation of geothermal arsenic (III) in stream waters of the eastern Sierra Nevada. Environ. Sci. Technol. 1998, 32, 657–662. [Google Scholar] [CrossRef]
  79. Langner, H.W.; Jackson, C.R.; McDermott, T.R.; Inskeep, W.P. Rapid oxidation of arsenite in a hot spring ecosystem, Yellowstone National Park. Environ. Sci. Technol. 2001, 35, 3302–3309. [Google Scholar] [CrossRef]
  80. Katsoyiannis, I.A.; Tzollas, N.M.; Tolkou, A.K.; Mitrakas, M.; Ernst, M.; Zouboulis, A.I. Use of novel composite coagulants for arsenic removal from waters—Experimental insight for the application of polyferric sulfate (PFS). Sustainability 2017, 9, 590. [Google Scholar] [CrossRef] [Green Version]
  81. Golfinopoulos, S.K.; Varnavas, S.P.; Alexakis, D.E. The Status of Arsenic Pollution in the Greek and Cyprus Environment: An Overview. Water 2021, 13, 224. [Google Scholar] [CrossRef]
  82. Economou-Eliopoulos, M.; Frei, R.; Atsarou, C. Application of chromium stable isotopes to the evaluation of Cr (VI) contamination in groundwater and rock leachates from central Euboea and the Assopos basin (Greece). Catena 2014, 122, 216–228. [Google Scholar] [CrossRef]
  83. Kanellopoulos, C.; Mitropoulos, P.; Valsami-Jones, E.; Voudouris, P. A new terrestrial active mineralising hydrothermal system associated with ore-bearing travertines in Greece (northern Euboea Island and Sperchios area). J. Geochem. Explor. 2017, 179, 9–24. [Google Scholar] [CrossRef]
  84. Sanjuan, B.; Negrel, G.; Le Lous, M.; Poulmarch, E.; Gal, F.; Damy, P.C. Main geochemical characteristics of the deep geothermal brine at Vendenheim (Alsace, France) with constraints on temperature and fluid circulation. In World Geothermal Congress; CCSD: Rekjavik, Iceland, 2020. [Google Scholar]
  85. Sanjuan, B.; Millot, R.; Innocent, C.; Dezayes, C.; Scheiber, J.; Brach, M. Main geochemical characteristics of geothermal brines collected from the granite basement in the upper Rhine Graben and constraints on their deep temperature and circulation. Chem. Geol. 2016, 428, 27–47. [Google Scholar] [CrossRef]
  86. Kumar, M. An approach towards green alumina refinery and sustainable development. Travaux 2011, 36, 11–22. [Google Scholar]
  87. Bray, E.L. Bauxite and alumina. In Min. Eng.; 2010; 62, pp. 40–41, USGS Publications Warehouse. Available online: http://pubs.er.usgs.gov/publication/70044852 (accessed on 30 May 2021).
  88. Balomenos, E.; Gianopoulou, I.; Panias, D.; Paspaliaris, I.; Perry, K.; Boufounos, D. Efficient and complete exploitation of the bauxite residue (red mud) produced in the Bayer process. In Proceedings of the EMC, Düsseldorf, Germany, 26–29 June 2011; pp. 745–758. [Google Scholar]
  89. Power, G.; Graefe, M.; Klauber, C. Bauxite residue issues: I. Current management, disposal and storage practices. Hydrometallurgy 2011, 108, 33–45. [Google Scholar] [CrossRef]
  90. Eliopoulos, D.G.; Economou-Eliopoulos, M.; Apostolikas, A.; Golightly, J.P. Geochemical features of nickel-laterite deposits from the Balkan Peninsula and Gordes, Turkey: The genetic and environmental significance of arsenic. Ore Geol. Rev. 2012, 48, 413–427. [Google Scholar] [CrossRef]
  91. Alevizos, G. Mineralogische und erzmikroskopische Untersuchung an Eisennickelerzen des Vorkommens von Marmeiko (Lokris-Griechenland). Ber. Deutsch. Mineral. Ges. Beih. Eur. J. Mineral. 1999, 11, 18. [Google Scholar]
  92. Zevgolis, M.N.; Georgakellos, N.; Spyropoulos, P. Ferrous Metals and Ferroalloys in Greece. Present State and Perspective. 1999. In Proceedings of the Conference, Metallurgy of Balkan Countries, Metallurgy of Ferous Metals of Balkan Countries, Volume No. 1 Sofia, Bulgari. Available online: https://www.researchgate.net/profile/Emmanouil-Zevgolis/publication/320979810_Ferrous_Metals_and_Ferroalloys_in_Greece_Present_State_and_Perspective/links/5a057cbf0f7e9bc40794e2e6/Ferrous-Metals-and-Ferroalloys-in-Greece-Present-State-and-Perspective.pdf (accessed on 30 May 2021).
  93. Wazne, M.; Jagupilla, S.C.; Moon, D.H.; Christodoulatos, C.; Koutsospyros, A. Leaching mechanisms of Cr(VI) from chromite ore processing residue. J. Environ. Qual. 2008, 37, 2125–2134. [Google Scholar] [CrossRef] [PubMed]
  94. Chrysochoou, M.; Dermatas, D.; Moon, D.H.; Christodoulatos, C. Reductive treatment of Chromite Ore Processing Residue (COPR): Lessons from a field study. In Proceedings of the GeoCongress 2008: Geotechnics of Waste Management and Remediation, New Orleans, LA, USA, 9–12 March 2008; pp. 748–755. [Google Scholar]
  95. Tsioptsias, C.; Samiotis, G.; Lefteri, L.; Amanatidou, E. Cr(VI) Leached from Lignite Fly Ash—Assessment of Groundwater Contamination Risk. Water Air Soil Pollut. 2020, 231, 373. [Google Scholar] [CrossRef]
  96. Pavloudakis, F.; Roumpos, C.; Karlopoulos, E.; Koukouzas, N. Sustainable Rehabilitation of Surface Coal Mining Areas: The Case of Greek Lignite Mines. Energies 2020, 13, 3995. [Google Scholar] [CrossRef]
  97. Christanis, K.; Georgakopoulos, A.; Fernandez-Turiel, J.L.; Bouzinos, A. Geological factors influencing the concentration of trace elements in the Philippi peatland, eastern Macedonia, Greece. Int. J. Coal Geol. 1998, 36, 295–313. [Google Scholar] [CrossRef]
  98. Georgakopoulos, A.; Filippidis, A.; Kassoli-Fournaraki, A. Leachability of major and trace elements of fly ash from Ptolemais power station, Northern Greece. Energy Sources 2002, 24, 103–113. [Google Scholar] [CrossRef]
  99. Veranis, N.; Nimfopoulos, M.K.; Gertsis, A.; Gerouki, F. Agricultural and industrial applications of the hellenic fly ash and environmental impacts. In Proceedings of the 19th International Congress Industrial Minerals, Athens, Greek, 30 March–2 April 2008. [Google Scholar]
  100. Kazakis, N.; Mattas, C.; Pavlou, A.; Patrikaki, O.; Voudouris, K. Multivariate statistical analysis for the assessment of groundwater quality under different hydrogeological regimes. Environ. Earth Sci. 2017, 76, 349. [Google Scholar] [CrossRef]
  101. Vasileiou, E.; Papazotos, P.; Dimitrakopoulos, D.; Perraki, M. Expounding the origin of chromium in groundwater of the Sarigkiol basin, Western Macedonia, Greece: A cohesive statistical approach and hydrochemical study. Environ. Monit. Assess. 2019, 191, 509. [Google Scholar] [CrossRef] [PubMed]
  102. Blagoeva, D.T.; Alves Dias, P.; Marmier, A.; Pavel, C.C. Assessment of Potential Bottlenecks Along the Materials Supply Chain for the Future Deployment of Low-Carbon Energy and Transport Technologies in the EU. Wind Power, Photovoltaic and Electric Vehicles Technologies, Time Frame: 2015–2030; JRC Science for Policy Report; Publications Office of the European Union: Luxembourg, 2016. [Google Scholar] [CrossRef]
  103. World Health Organization (WHO). Guidelines for Drinking Water Quality, 4th ed.; World Health Organization: Geneva, Switzerland, 2017. [Google Scholar]
  104. McGrane, S.J. Impacts of urbanisation on hydrological and water quality dynamics, and urban water management: A review. Hydrol. Sci. J. Des. Sci. 2016, 61, 2295–2311. [Google Scholar] [CrossRef]
  105. Cassard, D.; Lips, A.L.W.; Leistel, J.M.; Itard, Y.; Debeglia-Marchand, N.; Guillou-Frottier, L.; Spakman, W.; Stein, G.; Husson, Y. Understanding and assessing European mineral resources—A new approach using GIS Central Europe. Schweiz. Mineral. Petrogr. Mitt. 2004, 84, 3–24. [Google Scholar]
  106. Pontikes, Y.; Rathossi, C.; Nikolopoulos, P.; Angelopoulos, G.N.; Jayaseelan, D.D.; Lee, W.E. Effect of firing temperature and atmosphere on sintering of ceramics made from Bayer process bauxite residue. Ceram. Int. 2009, 35, 401–407. [Google Scholar] [CrossRef]
  107. Binnemans, K.; Jones, P.T.; Blanpain, B.; van Gerven, T.; Pontikes, Y. Towards zero-waste valorisation of rare-earth-containing industrial process residues: A critical review. J. Clean. Prod. 2015, 99, 17–38. [Google Scholar] [CrossRef] [Green Version]
  108. Borra, C.R.; Pontikes, Y.; Binnemans, K.; van Gerven, T. Leaching of rare earths from bauxite residue (red mud). Min. Eng. 2015, 76, 20–27. [Google Scholar] [CrossRef] [Green Version]
  109. Gamaletsos, P. Mineralogy and Geochemistry of Bauxites from Parnassos-Ghiona Mines and the Impact on the Origin of the Deposits. Ph.D. Thesis, University of Athens, Athens, Greece, 2014; p. 347. [Google Scholar]
  110. Bergman, R.A. Nickel production from low-iron laterite ores: Process descriptions. CIM Bull. 2003, 96, 127–138. [Google Scholar]
  111. Sharma, S.; Singh, R.; Nielson, G.G. Selenium in soil, plant, and animal systems. Crit. Rev. Environ. Sci. Technol. 1983, 13, 23–50. [Google Scholar] [CrossRef]
  112. Tabelin, C.B.; Igarashi, T.; Villacorte-Tabelin, M.; Park, I.; Opiso, E.M.; Ito, M.; Hiroyoshi, N. Arsenic, selenium, boron, lead, cadmium, copper, and zinc in naturally contaminated rocks: A review of their sources, modes of enrichment, mechanisms of release, andmitigation strategies. Sci. Total Environ. 2018, 645, 1522–1553. [Google Scholar] [CrossRef] [PubMed]
  113. Tumolo, M.; Ancona, V.; De Paola, D.; Losacco, D.; Campanale, C.; Massarelli, C.; Uricchio, V.F. Chromium Pollution in European Water, Sources, Health Risk, and Remediation Strategies. Int. J. Environ. Res. Public Health 2020, 17, 5438. [Google Scholar] [CrossRef] [PubMed]
  114. Wang, T.; Wang, J.; Tang, Y.; Shi, H.; Ladwig, K. Leaching Characteristics of Arsenic and Selenium from Coal Fly Ash: Role of Calcium. Energy Fuels 2009, 23, 2959–2966. [Google Scholar] [CrossRef]
  115. Dixit, R.; Malaviya, D.; Pandiyan, K.; Singh, U.B.; Sahu, A.; Shukla, R.; Singh, B.P.; Rai, J.P.; Sharma, P.K.; Lade, H. Bioremediation of heavy metals from soil and aquatic environment: An overview of principles and criteria of fundamental processes. Sustainability 2015, 7, 2189–2212. [Google Scholar] [CrossRef] [Green Version]
  116. Kang, J.; Zhang, Z.; Wang, J.J. Influence of humic substances on bio-availability of Cu and Zn during sewage sludge composting. Bioresour. Technol. 2011, 102, 8022–8026. [Google Scholar] [CrossRef]
  117. Raptis, S.; Gasparatos, D.; Economou-Eliopoulos, M.; Petridis, A. Chromium uptake by lettuce as affected by the application of organic matter and Cr (VI)-irrigation water: Implications to the land use and water management. Chemosphere 2018, 210, 597–606. [Google Scholar] [CrossRef]
  118. Cavani, L.; Ciavatta, C.; Gessa, C. Identification of organic matter from peat, leonardite and lignite fertilisers using humification parameters and electro-focusing. Bioresour. Technol. 2003, 86, 45–52. [Google Scholar] [CrossRef]
  119. Levy, D.; Fogelman, E.; Itzhak, Y. The effect of water salinity on potatoes (Solanum tuberosum L.): Physiological indices and yielding capacity. Potato Res. 1988, 31, 601–610. [Google Scholar] [CrossRef]
  120. Rhoades, S.A. The efficiency effects of horizontal bank mergers. J. Bank. Financ. 1993, 17, 411–422. [Google Scholar] [CrossRef]
  121. Kabata-Pendias, A.; Mukherjee, A.B. Trace Elements from Soil to Human; Springer: Berlin/Heidelberg, Germany, 2007; p. 550. [Google Scholar]
  122. Gupta, M.; Gupta, S. An Overview of Selenium Uptake, Metabolism and Toxicity in Plants. Front. Plant Sci. 2017, 7, 2074. [Google Scholar] [CrossRef] [Green Version]
  123. Rai, D.; Eary, L.E.; Zachara, J.M. Environmental chemistry of chromium. Sci. Total Environ. 1989, 86, 15–23. [Google Scholar] [CrossRef]
  124. Frei, R.; Gaucher, C.; Poulton, S. Canfield, D.E. Fluctuations in Precambrian atmospheric oxygenation recorded by chromium isotopes. Nature 2009, 461, 250–253. [Google Scholar] [CrossRef]
  125. Døssing, L.N.; Dideriksen, K.; Stipp, S.L.S.; Frei, R. Reduction of hexavalent chromium by ferrous iron: A process of chromium isotope fractionation and its relevance to natural environments. Chem. Geol. 2011, 285, 157–166. [Google Scholar] [CrossRef]
  126. Frei, R.; Gaucher, C.; Døssing, L.N.; Sial, A.N. Chromium isotopes in carbonates—A tracer for climate change and for reconstructing the redox state of ancient seawater. Earth Planet. Sci. Lett. 2011, 312, 114–125. [Google Scholar] [CrossRef]
  127. Bauer, K.W.; Cole, D.B.; Asael, D.; Francois, R.; Calvert, S.E.; Poulton, S.W.; Planavsky, N.J.; Crowe, S.A. Chromium isotopes in marine hydrothermal sediments. Chem. Geol. 2019, 529, 119286. [Google Scholar] [CrossRef]
  128. Liang, J.; Huang, X.; Yan, J.; Li, Y.; Zhao, Z.; Liu, Y.; Ye, J.; Wei, Y. A review of the formation of Cr(VI) via Cr(III) oxidation in soils and groundwater. Sci. Total Environ. 2021, 774, 145762. [Google Scholar] [CrossRef]
  129. Schauble, E.A. Applying Stable Isotope Fractionation Theory to New Systems. Rev. Mineral. Geochem. 2004, 55, 65–111. [Google Scholar] [CrossRef]
  130. Schoenberg, R.; Zink, S.; Staubwasser, M.; von Blanckenburg, F. The stable Cr isotope inventory of solid Earth reservoirs determined by double spike MC-ICP-MS. Chem. Geol. 2008, 249, 294–306. [Google Scholar] [CrossRef]
  131. Eary, L.E.; Rai, D. Kinetics of chromium(III) oxidation to chromium(VI) by reaction with manganese dioxide. Environ. Sci. Technol. 1987, 21, 1187–1193. [Google Scholar] [CrossRef]
  132. Slejko, F.F.; Petrini, R.; Lutman, A.; Forte, C.; Ghezzi, L. Chromium isotopes tracking the resurgence of hexavalent chromium contamination in a past-contaminated area in the Friuli Venezia Giulia Region, northern Italy. Isot. Environ. Health Stud. 2019, 55, 56–59. [Google Scholar] [CrossRef]
  133. Cucchi, F.; Franceschini, G.; Zini, L. Hydrochemical investigations and groundwater provinces of the Friuli Venezia Giulia Plain aquifers, northeastern Italy. Environ. Geol. 2008, 55, 985–999. [Google Scholar] [CrossRef]
  134. Pettine, M.; Millero, F.J.; La Noce, T. Chromium(III) interactions in seawater through its oxidation kinetics. Mar. Chem. 1991, 34, 29–46. [Google Scholar] [CrossRef]
  135. Xie, B.; Shan, C.; Xu, Z.; Li, X.; Zhang, X.; Chen, J.; Pan, B. One-step Removal of Cr(VI) at Alkaline pH by UV/sulfiteProcess: Reduction to Cr(III) and in situ Cr(III) Precipitation. Chem. Eng. J. 2017, 308, 791–797. [Google Scholar] [CrossRef]
  136. Eliopoulos, I.P.; Eliopoulos, G.D.; Economou-Eliopoulos, M. The Cr(VI) stability in contaminated coastal groundwater: Salinity as a driving force. Minerals 2021, 11, 160. [Google Scholar] [CrossRef]
  137. Riveraa, J.; Reicha, M.; Schoenbergb, R.; Gonzalez-Jimenezc, J.M.; Barraa, F.; Aiglspergerd, T.; Proenzae, J.A.; Carretier, S. Platinum-group element and gold enrichment in soils monitored by chromium stable isotopes during weathering of ultramafic rocks. Chem. Geol. 2018, 499, 84–99. [Google Scholar] [CrossRef]
  138. Novak, M.; Chrastny, V.; Cadkova, E.; Farkas, J.; Bullen, T.D.; Tylcer, J.; Szurmanova, Z.; Cron, M.; Prechova, E.; Curik, J.; et al. Common occurrence of a positive δ53Cr shift in Central European waters contaminated by geogenic/industrial chromium relative to source values. Environ. Sci. Technol. 2014, 48, 6089–6096. [Google Scholar] [CrossRef] [PubMed]
  139. Novak, M.; Martinkova, E.; Chrastny, V.; Stepanova, M.; Sebek, O.; Andronikov, A.; Curik, J.; Veselovsky, F.; Prechova, E.; Houskova, M. The fate of Cr(VI) in contaminated aquifers 65 years after the first spillage of plating solutions: A δ53Cr study at four Central European sites. Catena 2017, 158, 371–380. [Google Scholar] [CrossRef]
  140. Koilakos, D. Aspects of hexavalent chromium pollution of Thebes Plain Aquifer, Boeotia, Greece. Water 2017, 9, 611. [Google Scholar] [CrossRef] [Green Version]
  141. Economou-Eliopoulos, M.; Megremi, I.; Vasilatos, C.; Frei, R.; Mpourodimos, I. Geochemical constraints on the sources of Cr(VI) contamination in waters of Messapia (Central Evia) Basin. Appl. Geochem. 2017, 84, 13–25. [Google Scholar] [CrossRef]
  142. Izbicki, J.A.; Wright, M.T.; Seymour, W.A.; McCleskey, R.B.; Fram, M.S.; Belitz, K.; Esser, B.K. Cr(VI) occurrence and geochemistry in water from public-supply wells in California. Appl. Geochem. 2015, 63, 203–217. [Google Scholar] [CrossRef] [Green Version]
  143. Johnson, C.C.; Demetriades, A. Urban geochemical mapping: A review of case studies in this volume. Chapter 2. In Mapping the Chemical Environment of Urban Areas; Johnson, C.C., Demetriades, A., Locutura, J., Ottesen, R.T., Eds.; John Wiley & Sons, Ltd.: Chichester, UK, 2011; pp. 7–27. [Google Scholar]
  144. Vaiopoulou, E.; Gikas, P. Regulations for chromium emissions to the aquatic environment in Europe and elsewhere. Chemosphere 2020, 254, 126876. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Plots of the UMIA (calculated ultramafic index of alteration) versus Cr contents in weathered peridotites and saprolite zone, laterites from Kastoria and W. Vermion lying on peridotites, laterites of karst type from Lokris, and contaminated soils from Evia. Data from Table 1 and [50].
Figure 1. Plots of the UMIA (calculated ultramafic index of alteration) versus Cr contents in weathered peridotites and saprolite zone, laterites from Kastoria and W. Vermion lying on peridotites, laterites of karst type from Lokris, and contaminated soils from Evia. Data from Table 1 and [50].
Minerals 11 00775 g001
Figure 2. Variation of (a) Cr, (b) Fe, (c) Co, and (d) Mn versus Ni contents in soils from the Assopos–Thiva basins, C. Evia., and, in laterite deposits. Data from [50,60,61,62,63].
Figure 2. Variation of (a) Cr, (b) Fe, (c) Co, and (d) Mn versus Ni contents in soils from the Assopos–Thiva basins, C. Evia., and, in laterite deposits. Data from [50,60,61,62,63].
Minerals 11 00775 g002
Figure 3. Maps showing spatial distribution of Cr in soils due to natural processes and mining of Fe-Ni laterites (a), and in groundwater from the Evia and Assopos–Thiva basins due to natural processes, mining, and industrial activity (b) [30].
Figure 3. Maps showing spatial distribution of Cr in soils due to natural processes and mining of Fe-Ni laterites (a), and in groundwater from the Evia and Assopos–Thiva basins due to natural processes, mining, and industrial activity (b) [30].
Minerals 11 00775 g003
Figure 4. Plots of the Cr(VI) and total dissolved solids (TDS) versus B concentrations (a,b) for coastal groundwater from C. Evia, Assopos, and Thiva basins, seawater, and geothermal water. Data from [25,64,65,82,83,84].
Figure 4. Plots of the Cr(VI) and total dissolved solids (TDS) versus B concentrations (a,b) for coastal groundwater from C. Evia, Assopos, and Thiva basins, seawater, and geothermal water. Data from [25,64,65,82,83,84].
Minerals 11 00775 g004
Figure 5. Plot of the δ53Cr values versus Cr(VI) in contaminated by Cr(VI) groundwater by industrial activities in the Friuli Venezia Giulia (Italy), the Czech Republic (Central Europe), the Assopos, Thiva basin affected by both industrial activities and natural processes, and Central Evia of geogenic origin. Data from [82,132,138,139,140].
Figure 5. Plot of the δ53Cr values versus Cr(VI) in contaminated by Cr(VI) groundwater by industrial activities in the Friuli Venezia Giulia (Italy), the Czech Republic (Central Europe), the Assopos, Thiva basin affected by both industrial activities and natural processes, and Central Evia of geogenic origin. Data from [82,132,138,139,140].
Minerals 11 00775 g005
Table 1. Calculated Ultramafic Index of Alteration (UMIA) for representative samples of weathered peridotites Fe–Ni-laterites, bauxite laterites, and contaminated soils from Greece. Data from [50,59].
Table 1. Calculated Ultramafic Index of Alteration (UMIA) for representative samples of weathered peridotites Fe–Ni-laterites, bauxite laterites, and contaminated soils from Greece. Data from [50,59].
Indexwt% mg/kg Indexwt% mg/kg
LocationUMIAMnOCrNiCoLocationUMIAMnOCrNiCo
Kastoria C. Greece, Lokris520.18335013,300280
Weathered peridotites Laterites of karst type69.40.1226,9604790220
Ka-21.40.0512017,00050 77.40.2740,5705400280
Ka-36.90.08150023,000180 70.20.1320,19058,900280
Fe-Ni-Laterites 83.60.3320,1902360280
Ka-454.70.9516,00085001600 85.11.5720,3406120283
Ka-4b69.30.8823,00010,5001200 87.80.5826,9605180219
Ka-598.40.2520,0008000330 95.10.1738,8004320280
Ka-693.41.715,0009100820 77.40.532,30013,500281
Ka-793.00.7923,0007100420 67.80.2419,8408280286
Ka-894.60.317,0004400330 77.10.3114,6503690280
W. Vermion 83.20.2727,8003700280
Weathered peridotites 59.60.0719,0004630278
PR-1200.2458005500300 76.20.1323,10013,400280
PR-2120.1125002200200C. Evia (Psachna)
PR-3180.2335004100200Soils19.70.11108065041
PR-4180.1833004700200 18.70.09118067040
P-p140.138003100240 18.80.112090112062
Laterites 20.20.12150081052
PR-5870.9721,90015,0001300 13.60.11850130066
PR-6790.6112,9006900700 23.90.152200128076
WV4510.4111,50010,200500 24.10.161760163088
WV5360.3311,4009500400 20.90.12153070045
WV6780.6211,0007800820 22.20.13123082052
Bauxite laterites 18.10.1164073045
P-1470.3755005200400 21.10.13165078052
P-2800.4144005200320 17.90.11165065041
P-3810.2438006400400 21.60.12146073048
P-4810.3641004200400 18.40.11164086050
P-5820.2636004200400 23.70.13133072052
P-6170.3119002000200 16.70.09172060038
Table 2. Chromium, Ni, Mn, Co, Zn, and Fe range and mean in plants/crops from the C. Evia and Assopos–Thiva basins.
Table 2. Chromium, Ni, Mn, Co, Zn, and Fe range and mean in plants/crops from the C. Evia and Assopos–Thiva basins.
mg/kg
MetalsRangeMeanNormalDeficientExcesive
Cr0.3–200270.1–0.5_5–210
Ni4–210500.1–5_10–100
Mn8–1906530–3000–30400–1000
Co0.5–15710_50–69
Zn28–3805027–1500–20100–400
Fe28–3000780__2000
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Economou-Eliopoulos, M.; Megremi, I. Contamination of the Soil–Groundwater–Crop System: Environmental Risk and Opportunities. Minerals 2021, 11, 775. https://0-doi-org.brum.beds.ac.uk/10.3390/min11070775

AMA Style

Economou-Eliopoulos M, Megremi I. Contamination of the Soil–Groundwater–Crop System: Environmental Risk and Opportunities. Minerals. 2021; 11(7):775. https://0-doi-org.brum.beds.ac.uk/10.3390/min11070775

Chicago/Turabian Style

Economou-Eliopoulos, Maria, and Ifigeneia Megremi. 2021. "Contamination of the Soil–Groundwater–Crop System: Environmental Risk and Opportunities" Minerals 11, no. 7: 775. https://0-doi-org.brum.beds.ac.uk/10.3390/min11070775

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