Next Article in Journal
The Open Question of How GPCRs Interact with GPCR Kinases (GRKs)
Next Article in Special Issue
Silymarin-Enriched Biostimulant Foliar Application Minimizes the Toxicity of Cadmium in Maize by Suppressing Oxidative Stress and Elevating Antioxidant Gene Expression
Previous Article in Journal
Subnormothermic Perfusion with H2S Donor AP39 Improves DCD Porcine Renal Graft Outcomes in an Ex Vivo Model of Kidney Preservation and Reperfusion
Previous Article in Special Issue
Exogenous 3,3′-Diindolylmethane Improves Vanadium Stress Tolerance in Brassica napus Seedling Shoots by Modulating Antioxidant Enzyme Activities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fractionation of Heavy Metals in Multi-Contaminated Soil Treated with Biochar Using the Sequential Extraction Procedure

1
Key Laboratory of Plant Nutrition and Fertilizer in South Region, Ministry of Agriculture, Institute of Agricultural Resources and Environment, Guangdong Academy of Agricultural Sciences, Guangzhou 510640, China
2
Department of Soils and Water, Faculty of Agriculture, Al-Azhar University, Assiut 71524, Egypt
3
Department of Botany and Plant Physiology, Faculty of Agrobiology, Food, and Natural Resources, Czech University of Life Sciences Prague, Kamycka 129, 165 00 Prague, Czech Republic
4
Department of Biology, College of Science, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
5
Department of Plant Physiology, Slovak University of Agriculture, Nitra, Tr. A. Hlinku 2, 949 01 Nitra, Slovakia
6
National Institute of Research in Rural Engineering, Water 13 and Forests (INRGREF). BP 10, Ariana 2080, Tunisia
7
Department of Ecology and Environmental Protection, Poznan University of Life Sciences, Piątkowska 94, 60-649 Poznan, Poland
8
Department of Agronomy, Faculty of Agriculture, Kafrelsheikh University, Kafrelsheikh 33516, Egypt
*
Authors to whom correspondence should be addressed.
Biomolecules 2021, 11(3), 448; https://doi.org/10.3390/biom11030448
Submission received: 13 February 2021 / Revised: 6 March 2021 / Accepted: 14 March 2021 / Published: 17 March 2021

Abstract

:
Heavy metals (HMs) toxicity represents a global problem depending on the soil environment’s geochemical forms. Biochar addition safely reduces HMs mobile forms, thus, reducing their toxicity to plants. While several studies have shown that biochar could significantly stabilize HMs in contaminated soils, the study of the relationship of soil properties to potential mechanisms still needs further clarification; hence the importance of assessing a naturally contaminated soil amended, in this case with Paulownia biochar (PB) and Bamboo biochar (BB) to fractionate Pb, Cd, Zn, and Cu using short sequential fractionation plans. The relationship of soil pH and organic matter and its effect on the redistribution of these metals were estimated. The results indicated that the acid-soluble metals decreased while the fraction bound to organic matter increased compared to untreated pots. The increase in the organic matter metal-bound was mostly at the expense of the decrease in the acid extractable and Fe/Mn bound ones. The highest application of PB increased the organically bound fraction of Pb, Cd, Zn, and Cu (62, 61, 34, and 61%, respectively), while the BB increased them (61, 49, 42, and 22%, respectively) over the control. Meanwhile, Fe/Mn oxides bound represents the large portion associated with zinc and copper. Concerning soil organic matter (SOM) and soil pH, as potential tools to reduce the risk of the target metals, a significant positive correlation was observed with acid-soluble extractable metal, while a negative correlation was obtained with organic matter-bound metal. The principal component analysis (PCA) shows that the total variance represents 89.7% for the TCPL-extractable and HMs forms and their relation to pH and SOM, which confirms the positive effect of the pH and SOM under PB and BB treatments on reducing the risk of the studied metals. The mobility and bioavailability of these metals and their geochemical forms widely varied according to pH, soil organic matter, biochar types, and application rates. As an environmentally friendly and economical material, biochar emphasizes its importance as a tool that makes the soil more suitable for safe cultivation in the short term and its long-term sustainability. This study proves that it reduces the mobility of HMs, their environmental risks and contributes to food safety. It also confirms that performing more controlled experiments, such as a pot, is a disciplined and effective way to assess the suitability of different types of biochar as soil modifications to restore HMs contaminated soil via controlling the mobilization of these minerals.

1. Introduction

Heavy metals (HMs) are considered severe worldwide problems that threaten the ecosystem and human life [1,2]. According to the Ministry of Environmental Protection and Ministry of Land and Resources, PRC, many HMs pollution sources and their high level in the soil cause hazardous effects on soil quality, fertility, food safety, and human health [3]. Among these sources, effluents discharges and mining activities for more than 50 years in northern Guangdong Province, China, are mainly implicated in the elevated levels with these pollutants. Shaoguan city is a representative area of these mining activities with high levels of heavy metals and low pH values [4]. However, studies on geochemical fractions of heavy metals in this area are limited.
In the past, various techniques were used to reduce HMs contamination by phytoremediation [5,6]. This idea may be useful in fairly polluted soils, but in highly contaminated soils, a gradual decrease in plant growth makes it ineffective. In this case, biochar serves as a faster, more efficient, and environmentally friendly alternative via immobilization of heavy metals in the agricultural soil [7,8,9,10,11,12]. The characteristics of biochar (surface area, pH, ash, elemental composition, and carbon content) make it one of the most suitable materials to handle and remediate different heavy metals [13]. Biochar properties are highly dependent on the temperature, duration of pyrolysis, feedstock from which the biochar is made, and final acidity [2].
Generally, the total heavy metal concentration is inexpressive of its toxicity, harm to plants and the environment. A low but mobile amount may be more harmful than a large immobile amount depending on the chemical form, species, and solid-phase characteristics, which significantly affect its redistribution [12,14,15]. These forms can be controlled by precipitation, ion exchange, water compounds, stabilization, and plant uptake [12]. Several other critical chemical factors also influence the mobility of soil HMs. One of the dominant factors is the high soil pH that immobilizes metals [16]. Additionally, soil organic matter content and its type, and the synergetic effects of soluble ions are limiting factors in stabilization and metal movement in the contaminated soils [13,17]. Therefore, we foster the hypothesis that biochar is characterized by being a tool capable of raising soil pH. The organic matter makes it a revolutionary solution to make these geochemical fractions more controlled, especially in the acidic soils of Shaoguan.
To fractionate these metals, many authors have recognized that sequential extraction is a crucial method that could provide useful information on the bioavailability of metals and their transformation forms [14,18,19,20]. Although several studies confirmed that biochar is an important material for adsorbing HMs from contaminated soil, the mechanism is still unclear [4,21]. The current study aims to (i) evaluate the effects of biochar types and their application rates on the geochemical form of lead (Pb), cadmium (Cd), zinc (Zn), and copper (Cu) in contaminated soil, and (ii) assess their effects on soil organic matter, soil pH and their relationship among the different forms of HMs.

2. Materials and Methods

2.1. Soil Samples and Biochar

Surface soil samples (0–20 cm) were collected from Shaoquan city, Guangdong province, north China, which is a representative area for mining activities. In addition, this place has received liquid waste discharges for over 50 years, which causes a high level of heavy metals, especially lead and zinc. This soil is classified according to the soil survey staff [22] as Udult. The collected samples were air-dried, crushed, and sieved through a 2 mm sieve, and then kept for soil analysis. Paulownia (PB) and Bamboo (BB) (by-products of Huanyu Energy Technology Co., Ltd., Guangzhou city, China) were pyrolyzed at 700–800 °C in a EURO-Model CK-2-Charcoal Kiln. A stainless-steel mill crushed the produced aforesaid types of biochar to fixed grain size (<0.1 mm). An amount of 1.5 kg of the tested soil was homogeneously mixed with biochar materials at rates of 2, 4, and 6%, and replicated three times plus a control treatment with zero biochar addition (CK). Through experiment duration (two months), the soil treated with biochar was moist to 70% of its field capacity. The relative humidity during this period ranged between 60 and 70%, and the temperature was 25–30 °C. The biochar’s elemental composition was executed using energy dispersive X-ray spectrometry (EDS) elemental mapping (Figure 1). The functional groups involved in SD adsorption were analyzed by Fourier-transform infrared (FTIR) spectroscopy (Nicolet 6700, Thermo Scientific). After 60 days, the soil reaction (pH) and salinity (electrical conductivity-EC) of the suspension 1:2.5 (soil: water ratio) were determined by a pH- and EC-meter, respectively [23]. Soil organic matter (SOM) content was determined through the oxidation of organic matter by dichromate in an acid medium [24]. Some properties of the soil and biochar types were measured according to [25], which are shown in Table 1. In the end, the heavy metals (Pb, Cd, Zn, and Cu) in the soil sample were extracted using USEPA Toxicity Characteristic Leaching Procedure (TCLP), the method 1311 [26] was used to assess the available form. Their geochemical fractions were also determined according to the European Union Bureau of Reference Procedure (EUBCR) as mentioned below.

2.2. USEPA Extraction

In polypropylene centrifuge tubes, 20 mL of unbuffered glacial acetic acid (pH 2.88) were added to one gram soil sample. The mixture was shaken using an orbital shaker at 30 rpm for 18 h, centrifuged at 3000 rpm for 30 min. The supernatant was filtered and quantified with atomic absorption spectroscopy (AA 800, Perkin Elmer, Centennial, CO, USA).

2.3. Sequential Extraction Procedure

A sequential extraction procedure was applied according to EUBCR to determine heavy metals redistribution (Pb, Cd, Zn, and Cu) as described by [27]. One gram of soil sample was placed into a 100 mL polycarbonate tube then four sequential extractions were carried out as fractions of F1 (acid-soluble and carbonate bound), then 40 mL of acetic acid (0.11 M) were added, shaken for 16 h, centrifuged at 3000 rpm for 20 min, and finally filtered using Whatman-42 filter paper. F2 (bound to Fe-Mn oxides), to the residual from F1, add 40 mL of hydroxylamine hydrochloride (0.5 M) then the extraction procedure was performed as described in F1. F3 (oxidizable bound of organic matter), to the residual from the F2, add 10 mL H2O2 (8.8 M), evaporated to near dryness at 85 °C, cooled then 50 mL of 1 M NH4OAc (pH 2) was added, and the extraction procedure was followed as described in F1. The residual (remaining residue, bound to the mineral matrix) was determined using HNO3-HF-HClO4 acids on a hot plate without boiling, filtered, and diluted with pure deionized water to 15 mL. Before the next step, the residue from each stage was washed with 20 mL of de-ionized water and shaken again for 30 min and then centrifuged for 20 min at 3000 rpm. The supernatant liquid was also collected into a 100 mL polyethylene volumetric flask and stored in a refrigerator at 4 °C before analysis. The metal concentrations were determined with atomic absorption spectrophotometer (AA 800, Perkin Elmer Co., 14775 E Hinsdale Ave, Centennial, CO, USA).

3. Results

3.1. Biochars Description

Depending on SEM-EDS analysis, paulownia biochar (PB) had higher oxygen and lower carbon content than bamboo biochar (BB). The relative Si, K, and Ca contents in the PB were higher than those of the BB, while the BB had higher P and Mg content than PB (Figure 1). Additionally, the paulownia biochar had higher alkalinity than that of bamboo biochar (Table 1). The O/C ratios of PB and BB were 0.22 and 0.19, respectively (Figure 1). The results of the FTIR analysis of PB are shown in the FTIR spectra presented in Figure 2. Stretching of hydroxyl (-OH), carbonyl or carboxyl (C=O), aromatic (C-H) and Si-O functional groups for PB than BB indicate that the ability of paulownia makes it more than bamboo to absorb heavy metals in the soil.

3.2. Effects of Biochar on Soil Properties

Data in Table 2 indicated that biochar applications significantly increased the organic matter (OM) content, and the soil pH was raised along with increasing biochar levels compared to the control. The highest PB and BB rates increased the OM by 1.28 and 1.17 times, respectively, over the untreated one. The pH values were increased due to the biochar application, where it reaches the maximum value at 4% PB and 6% BB. The EC varied according to the types of biochar and their application rates. The PB application decreased soil salinity (EC), which was distinct from the low application level. On the other hand, the EC values gradually increased with the increasing BB levels.

3.3. Heavy Metals Mobility

The illustrated data in Table 3 shows that the biochar additions caused a decrease in HMs concentration compared with CK treatment. A significant reduction of the TCLP-extractable Pb, Cd, Zn, and Cu in the tested soil was observed using biochar. The magnitude reductions varied with the types of feedstock and their application rates. Generally, the PB material was more efficient in reducing the TCLP-extractable metals than that of BB. As compared to the control, application of 6% PB decreased the TCLP-extractable metals by 22, 17, 24, and 45% for the Pb, Cd, Zn, and Cu, respectively. On the contrary, the corresponding reductions were 20.45, 15, 15, and 12%, respectively, with 6% BB.

3.4. Heavy Metals Redistribution

3.4.1. Acid-Soluble Fraction

Applying the biochar decreased the acid-soluble fractions of Pb, Cd, Zn and Cu compared to the control treatment (Table 4). The degree of reduction of these metal fractions was depended on the types of biochar, application rates, and metal species. Adding 6% PB or BB remarkably decreased the Cu content followed by Pb over the control treatment, and their reductions reached 37.89 and 31.22%, respectively. Despite Cd total content (2.5%), its acid-soluble fraction is very high since it reached 64 to 73% of its total content compared to other metals. The acid-soluble fraction varied from 12 to 14% for Pb, 14 to 17% for Zn, and 4 to 6% for Cu, and these variations were governed by the nature of biochar and application rates. The mobility of target metals due to the biochar application was in the decreasing order of Cd > Zn > Pb > Cu.

3.4.2. Bound to Fe and Mn Oxides

The biochar treatments realized a significant reduction in cadmium oxide bounds with few exceptions (Table 5). The PB reduced the Cu bound to Fe/Mn oxides, especially with the highest application rate, while the BB showed an insignificant increase. There was an irregular trend of decreasing Pb oxides bound fraction with either type or level of biochar application. The PB at a 4% level showed the highest Pb concentration bound to the Fe/Mn oxides, while BB applied at a level of 6% showed the lowest one (Table 5). The Pb was predominantly adsorbed by the Fe/Mn fractions (45 to 63%), while an insignificant amount of Cd, Zn and Cu was extracted from the Fe/Mn oxide bound fraction. In Zn case, all the biochar treatments significantly (p < 0.05) increased its oxide bound fraction.

3.4.3. Bound to Organic Matter

Biochar application tends to significantly increase in metal concentrations associated with soil organic matter (Table 6). This increase was mainly on the acid extractable (pore water ones) expense and partially on the oxides bound. The organic bound of Pb fraction, as relative change, increased over control by 62 and 61% at the highest PB and BB rates, respectively. Application of PB at 6% increased the OM bound fraction of Cd, Zn, and Cu to reach 62, 34, and 61%, respectively. BB highest addition also increased the organically bound fraction by 49, 42 and 22 % for Cd, Zn and Cu, respectively. The results indicated that Pb was adsorbed mainly by the organic fractions, while the large portion of the acid extractable bound to Cd. Meanwhile, Fe/Mn oxides bound represents the large portion associated with Zn and Cu.

3.4.4. Residual Fraction

The residual fraction of Pb, Cd, Zn, and Cu did not change significantly due to the application of any of the biochar tested with few exceptions for Pb (Table 7). The residual fraction of Pb, Cd, Zn, and Cu varied from 18 to 20%, 5 to 6%, 65 to 67% and 42 to 47%.

3.4.5. Correlation Analysis

As can be seen, there was a significant negative correlation between acid-soluble extractable metal and soil organic matter (SOM) and soil pH (Figure 3). There was also a significant negative correlation between the organic matter bound metal and SOM and pH, while the Fe and Mn oxide had an insignificant negative correlation for Cd and Cu. Regarding the Fe and Mn oxide for Pb, there was a significant negative correlation with SOM, while the Zn significantly positive correlation was observed with SOM and pH. In other words, based on the above results, in soil treated with biochar, acid-soluble extractable metal was decreased, and subsequently, the organic matter bound increased.

3.4.6. Principal Component Analysis (PCA)

The PCA was used to determine the approximate correlation of the studied TCLP-extractable metals, extractable acid metals (F1), Fe and Mn oxides (F2), organic matter bound (F3) forms, soil pH, and soil organic matter (SOM) versus different biochar additives (Figure 4). The PCA shows that the total variance represents 89.7% for the TCPL-extractable metals, heavy metal forms, and their relation to pH and SOM. Regarding the results mentioned in the PCA confirm those obtained in our current study. Additionally, the organically bound realized while reducing the acid- extractable metal demonstrates the positive effect of the PB and BB treatments on reducing the studied metals’ risk.

4. Discussion

4.1. Biochar Type Effect

Biochar types affected soil properties (increased pH and OM and varied EC). The high soil pH and organic matter content (Table 2) might be attributed to biochar application. Similar results were obtained by [28,29]. The PB decreased the EC values, whereas the BB increased it as compared to the control. This may be attributed to salt adsorption, and its preservation resulted in a lower EC value of soil solution and/or movement of saline water downward, which reflected in a reduction EC value in the surface layer [2,28,30]. Our results clearly showed that all feedstock biochar reduced the risk of toxic metals by stabilizing them in insoluble forms. The PB was more effective in immobilizing these metals, especially at the highest level. This may be due to its high pH and EC values, which means a high elemental composition content as well as functional groups (Table 1, Figure 1 and Figure 2). These results are compatible with those obtained by [31,32,33]. They indicated that biochar’s elemental composition forms HM compounds with cations and anions via ion exchange, especially soft acids such as Pb+2 and Cu+2.

4.2. Biochar Rate Effect

The different rates of biochar application caused an increase of OM and alkalinity with few exceptions. The soil salinity (EC) was gradually increased by increasing BB rate, while it regularly decreased in PB. Similar results were observed by several authors [2,28,34]. The mobility reduction of the tested metals as a result of increasing biochar application was evident. This may be because the gradual addition of biochar resulted in increased pH values, which, in turn, was negatively reflected in the concentration of metals, especially in acidic soils. The increasing rate of biochar addition of bamboo and rice straw decreased HMs extractability, which was significantly highly (p < 0.01) correlated with soil pH [34]. Increased biochar additives might increase the formation of stable complexes via an increase in ions and functional groups. Raising soil pH in acidic soils due to increased biochar applications led to enhanced adsorption and reduced metal mobility [21,35]. The concentration of HMs extracted with calcium chloride (CaCl2) or DTPA of soil treated by biochar at the rate of 1 and 5% was significantly reduced at 5% compared to that at 1% [4,36].

4.3. Heavy Metals Mobility and Redistribution

The addition of biochar efficiently reduced the concentrations of soil TCLP and acid-soluble fractions of Pb, Cd, Zn, and Cu over untreated one (Table 3 and Table 4). This decrease may be due to the high pH and/or EC values of the applied biochar, as shown in Table 1, and their role in increasing the pH (especially in acidic soils) that reduced metal mobility [2,35]. The high surface adsorption capacity of biochar and/or formation of insoluble compounds could cause a reduction in the HMs solubility [2,37]. In the same context, [37] observed a significant reduction in the acid-soluble fraction of Pb, Cd, and Cu. In the current study, the BCR results indicated that the acid-soluble fraction for metals could be arranged in descending order of Cd > Zn >Pb > Cu. These results are in agreement with [38] who arranged the HMs according to acid-soluble fraction in descending order of Ni > Mn > Cd > Zn > Pb > Cr > Co > Cu. Soil amended biochar shows a large amount of Pb (60%, from the total amount of Pb) in Fe/Mn oxides bound fraction in comparison to other fractions. These results may be due to the affinity of Pb bound to Mn oxides. These results are compatible with those observed by [39], who extracted a significant amount of Pb (57.4%) from the Fe/Mn oxide bound fraction. Additionally, [40] found that Pb was the largest one associated with the oxides. However, the organically bound fraction increased induced biochar additions, while the soluble fraction decreased compared to the control. These decreases in the acid-soluble metals fraction were based on the increase of OM bound metal fraction. Moreover, the increase in the OM bound metal fraction was dependent on feedstock type and the level of addition. The biochar application caused an increase in the metal fraction bound to the organic matter, and the opposite trend was observed with soluble fraction [12,41]. Heavy metals could be adsorbed on biochar surfaces resulting in a reduction in their mobilization in biochar treated soils [42]. The data collected from correlation and PCA (Figure 3 and Figure 4) have further emphasized a strong relation between biochar additions and acid-soluble metals and SOM, pH, and metal bound to organic matter. Biochar additives resulted in a positive relationship with the organic matter, pH, and metals fraction associated with organic matter and a negative one with the acid-soluble fraction. This confirms the hypothesis that the increase in the organic matter caused an increase in the organic fraction (organic complexes) at the acid soluble fraction’s expense. Additionally, it caused an increase in pH, which leads us to confirm another hypothesis of metal precipitation as a normal result of increasing the pH.

4.4. Mechanisms of Heavy Metals Immobilization

Numerous studies, including our current study, have demonstrated the susceptibility of biochar to immobilize metals [2,12,41] but the mechanism is still unknown. One and/or more mechanisms are likely to be responsible for stabilizing metals in soils amended with biochar, among which are (1) natural adsorption on biochar surfaces [33], (2) chemical bonding with ions on its surface [32,33,43], (3) complexation with active functional groups [11,32], (4) precipitate via phosphates ions [11,31,44,45], and (5) precipitation by raising soil pH [13,46]. However, all these mechanisms are not compatible with all the metals studied due to the differences in the characteristics between these metals. While Cd and Zn tend to be physically adsorbed on the surface of biochar, Cu can bond chemically on the surface, and Pb can be complexed with a functional group and/or precipitate. In this study, the PB was more effective in HMs immobilization than BB. This is attributed mainly to the higher pH and EC values and/or functional groups of PB and its higher elemental composition content than BB one (Table 1, Figure 1 and Figure 2). It was found that the concentration of the TCLP-extractable metals was as follows: Cu (45%) >Zn (24%) > Pb (22%) > Cd (17%) due to the application of PB at the highest level. The decrease in portion extracted with acid and the increase in the organic matter bound fraction may explain the increase of organic matter bound metal fraction based on the reduction in acid extractable metals. PB had higher alkalinity, total dissolved salts, and O, Si, K, and Ca content than those of BB, which may explain the more significant reduction in HMs solubility. Organic amendments affect the mobility, and bioavailability of soil HMS depends on the specific metal, soil type, and the amendment characteristics such EC, CEC, and pH [30,47].

5. Conclusions

The risk of heavy metals increases in contaminated soil with the increase in the soluble form. Therefore, it became necessary to monitor the different changes of the metal forms and their relation to changes in soil properties after adding biochar. The results indicated that the redistribution of HMs mainly depends on the feedstock type (paulownia and bamboo), their application rate, soil pH, organic matter content, and metal type. The paulownia biochar was superior in terms of its effect on these metals’ solubility compared to bamboo biochar. The acid-soluble metals reduction was mostly at the expense of the organic matter and oxide pounds increases. High-significant negative correlations were noticed with both soil organic matter and pH from one side and the soluble fraction for all the studied metals from the other, while the opposite trend was observed with the organic matter bound fraction. The movement of HMs in soils and their transformations depends mainly on their geochemically related bond. This study proves that paulownia biochar is a good material for reducing heavy metals solubility to reduce hazardous environmental effects. Besides metal type, a reduction in HMs mobilization was dependent on raising the pH in acidic soils side by side with the increase in organic matter. Adding biochar is not a scientific curiosity but a useful technique to restrict metals’ mobility, consequently decreasing their uptake and improving soil properties.

Author Contributions

Conceptualization, M.A. and Z.L.; methodology, M.A. and Z.L.; software, M.A.; validation, M.A. and Z.L.; formal analysis, M.A., Z.L. and A.E.S.; investigation, M.A. and Z.L.; resources, M.A. and Z.L.; data curation, M.A. and Z.L.; writing—original draft preparation, M.A. and Z.L.; writing—review and editing, S.M., E.S.D., M.S., M.B., A.E.S., A.R.; funding acquisition, M.S., M.B., E.S.D. All authors have read and agreed to the published version of the manuscript.

Funding

This study was financially supported by the National Key Research and Development Program (SQ2017YFNC060046), the National Natural Science Foundation of China (41571313, 41401353), Guangdong Natural Science Foundation 2015A030313570, 2016A030313772), Department of Science and Technology of Guangdong Province (2016A020210034, 2017B020203002), Pearl River S&T Nova Program of Guangzhou, China (201610010131), and Talented Young Scientist Program (TYSP), Ministry of Science and Technology (MOST). The Taif University Researchers Supporting Project number (TURSP-2020/85), Taif University, Taif, Saudi Arabia, funded current work.

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Acknowledgments

The authors extend their appreciation to Taif University for funding current work by Taif University Researchers Supporting Project number (TURSP-2020/85), Taif University, Taif, Saudi Arabia. In addition, we acknowledge the financial contribution to EPPN2020-OPVaI-VA - ITMS313011T813.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shakoor, M.B.; Ali, S.; Rizwan, M.; Abbas, F.; Bibi, I.; Riaz, M.; Khalil, U.; Niazi, N.K.; Rinklebe, J. A review of biochar-based sorbents for separation of heavy metals from water. Int. J. Phytoremed. 2020, 22, 111–126. [Google Scholar] [CrossRef]
  2. Awad, M.; Moustafa-Farag, M.; Wei, L.; Huang, Q.; Liu, Z. Effect of garden waste biochar on the bioavailability of heavy metals and growth of Brassica juncea (L.) in a multi-contaminated soil. Arab. J. Geosci. 2020, 13, 439. [Google Scholar] [CrossRef]
  3. Ministry of Environmental Protection PRC and Ministry of Land and Resources PRC, Report on Soil Pollution in China. 2014. Available online: http://www.sdpc.gov.cn/fzgggz/ncjj/zhdt/201404/t20140418-607888.html (accessed on 6 March 2021).
  4. Lu, K.; Yang, X.; Gielen, G.; Bolan, N.; Ok, Y.S.; Niazi, N.K.; Xu, S.; Yuan, G.; Chen, X.; Zhang, X.; et al. Effect of bamboo and rice straw biochars on the mobility and redistribution of heavy metals (Cd, Cu, Pb and Zn) in contaminated soil. J. Environ. Manag. 2017, 186, 285–292. [Google Scholar] [CrossRef] [PubMed]
  5. Sarwar, N.; Imran, M.; Shaheen, M.R.; Ishaque, W.; Kamran, M.A.; Matloob, A.; Rehim, A.; Hussain, S. Phytoremediation strategies for soils contaminated with heavy metals: Modifications and future perspectives. Chemosphere 2017, 171, 710–721. [Google Scholar] [CrossRef]
  6. Awasthi, A.K.; Li, J.; Pandey, A.K.; Khan, J. An Overview of the Potential of Bioremediation for Contaminated Soil from Municipal Solid Waste Site. In Emerging and Eco-Friendly Approaches for Waste Management; Springer Singapore: Singapore, 2019; pp. 59–68. ISBN 9789811086694. [Google Scholar]
  7. Botté, S.E.; Freije, R.H.; Marcovecchio, J.E. Distribution of Several Heavy Metals in Tidal Flats Sediments within Bahía Blanca Estuary (Argentina). Water Air Soil Pollut. 2010, 210, 371–388. [Google Scholar] [CrossRef]
  8. Gonzaga, M.I.S.; de Matias, M.I.A.S.; Andrade, K.R.; de Jesus, A.N.; da Cunha, G.C.; de Andrade, R.S.; de Santos, J.C.J. Aged biochar changed copper availability and distribution among soil fractions and influenced corn seed germination in a copper-contaminated soil. Chemosphere 2020, 240, 124828. [Google Scholar] [CrossRef]
  9. LI, J.; XU, Y. Immobilization of Cd in paddy soil using moisture management and amendment. Environ. Sci. Pollut. Res. 2015, 22, 5580–5586. [Google Scholar] [CrossRef] [PubMed]
  10. Liu, D.; Li, S.; Islam, E.; Chen, J.; Wu, J.; Ye, Z.; Peng, D.; Yan, W.; Lu, K. Lead accumulation and tolerance of Moso bamboo (Phyllostachys pubescens) seedlings: Applications of phytoremediation. J. Zhejiang Univ. B 2015, 16, 123–130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Huang, M.; Zhu, Y.; Li, Z.; Huang, B.; Luo, N.; Liu, C.; Zeng, G. Compost as a Soil Amendment to Remediate Heavy Metal-Contaminated Agricultural Soil: Mechanisms, Efficacy, Problems, and Strategies. Water Air Soil Pollut. 2016, 227, 359. [Google Scholar] [CrossRef]
  12. Kabiri, P.; Motaghian, H.; Hosseinpur, A. Effects of Walnut Leaves Biochars on Lead and Zinc Fractionation and Phytotoxicity in a Naturally Calcareous Highly Contaminated Soil. Water Air Soil Pollut. 2019, 230, 263. [Google Scholar] [CrossRef]
  13. Wang, X.; Chang, V.W.-C.; Li, Z.; Chen, Z.; Wang, Y. Co-pyrolysis of sewage sludge and organic fractions of municipal solid waste: Synergistic effects on biochar properties and the environmental risk of heavy metals. J. Hazard. Mater. 2021, 412, 125200. [Google Scholar] [CrossRef]
  14. Awad, M.Y.M.; El-Desoky, M.A.; Ghallab, A.; Abdel-Mawly, S.E. Changes in Soil Zn and Mn Forms of Some Contaminated Egyptian Soils Treated with Organic Materials. Assiut J. Agric. Sci. 2017, 48, 269–285. [Google Scholar] [CrossRef]
  15. Zang, F.; Wang, S.; Nan, Z.; Ma, J.; Wang, Y.; Chen, Y.; Zhang, Q.; Li, Y. Influence of pH on the release and chemical fractionation of heavy metals in sediment from a suburban drainage stream in an arid mine-based oasis. J. Soils Sediments 2017, 17, 2524–2536. [Google Scholar] [CrossRef]
  16. Awad, M.Y.M.; El-Desoky, M.; Abdel-Mawly, S.E.; Ghallab, A. Mobility of Heavy Metals in Some Contaminated Egyptian Soils Treated with Certain Organic Materials. Ph.D. Thesis, Assiut University, Assiut, Egypt, 2007. [Google Scholar]
  17. Palansooriya, K.N.; Shaheen, S.M.; Chen, S.S.; Tsang, D.C.W.; Hashimoto, Y.; Hou, D.; Bolan, N.S.; Rinklebe, J.; Ok, Y.S. Soil amendments for immobilization of potentially toxic elements in contaminated soils: A critical review. Environ. Int. 2020, 134, 105046. [Google Scholar] [CrossRef] [PubMed]
  18. He, J.; Li, Y.; Qi, H.; Li, H.; Zhang, W. Biochar amendment changed soil-bound fractions of silver nanoparticles and ions but not their uptake by radish at an environmentally-relevant concentration. Biochar 2020, 2, 307–317. [Google Scholar] [CrossRef]
  19. Patel, A.K.; Das, N.; Goswami, R.; Kumar, M. Arsenic mobility and potential co-leaching of fluoride from the sediments of three tributaries of the Upper Brahmaputra floodplain, Lakhimpur, Assam, India. J. Geochem. Explor. 2019, 203, 45–58. [Google Scholar] [CrossRef]
  20. Baran, A.; Mierzwa-Hersztek, M.; Gondek, K.; Tarnawski, M.; Szara, M.; Gorczyca, O.; Koniarz, T. The influence of the quantity and quality of sediment organic matter on the potential mobility and toxicity of trace elements in bottom sediment. Environ. Geochem. Health 2019, 41, 2893–2910. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Ahmad, M.; Ok, Y.S.; Rajapaksha, A.U.; Lim, J.E.; Kim, B.-Y.; Ahn, J.-H.; Lee, Y.H.; Al-Wabel, M.I.; Lee, S.-E.; Lee, S.S. Lead and copper immobilization in a shooting range soil using soybean stover- and pine needle-derived biochars: Chemical, microbial and spectroscopic assessments. J. Hazard. Mater. 2016, 301, 179–186. [Google Scholar] [CrossRef]
  22. Soil Survey Staff. Keys to Soil Taxonomy, 12th ed.; U.S. Department of Agriculture, Natural Resources Conservation Service: Washington, DC, USA, 2014.
  23. Richards, L.A. Diagnosis and Improvement of Saline and Alkali Soils. Soil Sci. 1954, 78, 154. [Google Scholar] [CrossRef]
  24. Nelson, D.; Sommers, L.E. Total carbon, organic carbon, and organic matter. In Methods of Soil Analysis, Part 2: Chemical and Microbiological Properties, 2nd ed.; Page, A.L., Ed.; John Wiley & Sons: New York, NY, USA, 1982; pp. 539–579. [Google Scholar]
  25. Burt, R. Soil Survey Laboratory Methods Manual; Soil survey investigations report no. 42, version 4.0; Natural Resources Conservation Service, US Department of Agriculture: Washington, DC, USA, 2004.
  26. USEPA. Method 1311 TCLP-Toxicity Characteristic Leaching Procedure. In Test Methods for Evaluating Solid Waste, 3rd ed.; Environmental Protection Agency: Washington, DC, USA, 1992. [Google Scholar]
  27. Žemberyová, M.; Barteková, J.; Hagarová, I. The utilization of modified BCR three-step sequential extraction procedure for the fractionation of Cd, Cr, Cu, Ni, Pb and Zn in soil reference materials of different origins. Talanta 2006, 70, 973–978. [Google Scholar] [CrossRef]
  28. Awad, M.Y.M. Effect of Some Organic Compounds on Soil Properties and Plant Growth. Master’s Thesis, Menoufia University, Al Minufya, Egypt, 2001. [Google Scholar]
  29. Rekaby, S.A.; Awad, M.Y.M.; Hegab, S.A.; Eissa, M.A. Effect of some organic amendments on barley plants under saline condition. J. Plant Nutr. 2020, 43, 1840–1851. [Google Scholar] [CrossRef]
  30. Dahlawi, S.; Naeem, A.; Rengel, Z.; Naidu, R. Biochar application for the remediation of salt-affected soils: Challenges and opportunities. Sci. Total Environ. 2018, 625, 320–335. [Google Scholar] [CrossRef]
  31. Senthilkumar, R.; Prasad, D.M.R. Sorption of Heavy Metals onto Biochar. In Applications of Biochar for Environmental Safety; IntechOpen: Rijeka, Croatia, 2020. [Google Scholar]
  32. Lin, H.; Li, G.; Dong, Y.; Li, J. Effect of pH on the release of heavy metals from stone coal waste rocks. Int. J. Miner. Process. 2017, 165, 1–7. [Google Scholar] [CrossRef]
  33. Lu, H.; Zhang, W.; Yang, Y.; Huang, X.; Wang, S.; Qiu, R. Relative distribution of Pb2+ sorption mechanisms by sludge-derived biochar. Water Res. 2012, 46, 854–862. [Google Scholar] [CrossRef]
  34. Yang, X.; Liu, J.; McGrouther, K.; Huang, H.; Lu, K.; Guo, X.; He, L.; Lin, X.; Che, L.; Ye, Z.; et al. Effect of biochar on the extractability of heavy metals (Cd, Cu, Pb, and Zn) and enzyme activity in soil. Environ. Sci. Pollut. Res. 2016, 23, 974–984. [Google Scholar] [CrossRef] [PubMed]
  35. Ali, A.; Guo, D.; Zhang, Y.; Sun, X.; Jiang, S.; Guo, Z.; Huang, H.; Liang, W.; Li, R.; Zhang, Z. Using bamboo biochar with compost for the stabilization and phytotoxicity reduction of heavy metals in mine-contaminated soils of China. Sci. Rep. 2017, 7, 2690. [Google Scholar] [CrossRef] [Green Version]
  36. Zhang, Z.; Solaiman, Z.M.; Meney, K.; Murphy, D.V.; Rengel, Z. Biochars immobilize soil cadmium, but do not improve growth of emergent wetland species Juncus subsecundus in cadmium-contaminated soil. J. Soils Sediments 2013, 13, 140–151. [Google Scholar] [CrossRef]
  37. Jiang, J.; Xu, R.; Jiang, T.; Li, Z. Immobilization of Cu(II), Pb(II) and Cd(II) by the addition of rice straw derived biochar to a simulated polluted Ultisol. J. Hazard. Mater. 2012, 229–230, 145–150. [Google Scholar] [CrossRef]
  38. Ozcan, N.; Altundag, H. Speciation of heavy metals in street dust samples from sakarya I. Organized industrial district using the bcr sequential extraction procedure by ICP-OES. Bull. Chem. Soc. Ethiop. 2013, 27, 205–212. [Google Scholar] [CrossRef] [Green Version]
  39. Golui, D.; Datta, S.P.; Dwivedi, B.S.; Meena, M.C.; Trivedi, V.K.; Jaggi, S.; Bandyopadhyay, K.K. Assessing Geoavailability of Zinc, Copper, Nickel, Lead and Cadmium in Polluted Soils Using Short Sequential Extraction Scheme. Soil Sediment Contam. Int. J. 2021, 30, 74–91. [Google Scholar] [CrossRef]
  40. Feng, X.H.; Zhai, L.M.; Tan, W.F.; Liu, F.; He, J.Z. Adsorption and redox reactions of heavy metals on synthesized Mn oxide minerals. Environ. Pollut. 2007, 147, 366–373. [Google Scholar] [CrossRef] [PubMed]
  41. Liu, Y.; Wang, Y.; Lu, H.; Lonappan, L.; Brar, S.K.; He, L.; Chen, J.; Yang, S. Biochar application as a soil amendment for decreasing cadmium availability in soil and accumulation in Brassica chinensis. J. Soils Sediments 2018, 18, 2511–2519. [Google Scholar] [CrossRef]
  42. Lucchini, P.; Quilliam, R.S.; DeLuca, T.H.; Vamerali, T.; Jones, D.L. Increased bioavailability of metals in two contrasting agricultural soils treated with waste wood-derived biochar and ash. Environ. Sci. Pollut. Res. 2014, 21, 3230–3240. [Google Scholar] [CrossRef] [PubMed]
  43. Lei, S.; Shi, Y.; Qiu, Y.; Che, L.; Xue, C. Performance and mechanisms of emerging animal-derived biochars for immobilization of heavy metals. Sci. Total Environ. 2019, 646, 1281–1289. [Google Scholar] [CrossRef] [PubMed]
  44. Karami, N.; Clemente, R.; Moreno-Jiménez, E.; Lepp, N.W.; Beesley, L. Efficiency of green waste compost and biochar soil amendments for reducing lead and copper mobility and uptake to ryegrass. J. Hazard. Mater. 2011, 191, 41–48. [Google Scholar] [CrossRef]
  45. Park, J.H.; Choppala, G.; Lee, S.J.; Bolan, N.; Chung, J.W.; Edraki, M. Comparative Sorption of Pb and Cd by Biochars and Its Implication for Metal Immobilization in Soils. Water Air Soil Pollut. 2013, 224, 1711. [Google Scholar] [CrossRef]
  46. Cantrell, K.B.; Hunt, P.G.; Uchimiya, M.; Novak, J.M.; Ro, K.S. Impact of pyrolysis temperature and manure source on physicochemical characteristics of biochar. Bioresour. Technol. 2012, 107, 419–428. [Google Scholar] [CrossRef] [PubMed]
  47. Houben, D.; Evrard, L.; Sonnet, P. Mobility, bioavailability and pH-dependent leaching of cadmium, zinc and lead in a contaminated soil amended with biochar. Chemosphere 2013, 92, 1450–1457. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The energy dispersive X-ray spectra (EDS) collected from the scanning electron microscopy (SEM) regions of paulownia biochar (a), and bamboo biochar (b).
Figure 1. The energy dispersive X-ray spectra (EDS) collected from the scanning electron microscopy (SEM) regions of paulownia biochar (a), and bamboo biochar (b).
Biomolecules 11 00448 g001
Figure 2. Fourier-transform infrared (FTIR) spectroscopy of tested biochar.
Figure 2. Fourier-transform infrared (FTIR) spectroscopy of tested biochar.
Biomolecules 11 00448 g002
Figure 3. Values representing the correlation coefficient of the acid soluble extractable (F1), Fe and Mn oxides (F2), and organic matter bound (F3) in relation to soil organic matter (SOM) and soil pH. ns, Not significant, * significant, ** highly significant, *** very highly significant.
Figure 3. Values representing the correlation coefficient of the acid soluble extractable (F1), Fe and Mn oxides (F2), and organic matter bound (F3) in relation to soil organic matter (SOM) and soil pH. ns, Not significant, * significant, ** highly significant, *** very highly significant.
Biomolecules 11 00448 g003
Figure 4. Principal component analysis of TCLP-extractable and acid soluble extractable metals in contaminated soil treated with paulownia biochar and bamboo biochar at the rate of (0-control, 2, 4, and 6%). PC: Principal Components.
Figure 4. Principal component analysis of TCLP-extractable and acid soluble extractable metals in contaminated soil treated with paulownia biochar and bamboo biochar at the rate of (0-control, 2, 4, and 6%). PC: Principal Components.
Biomolecules 11 00448 g004
Table 1. Some characteristics of the tested soil and the utilized biochar.
Table 1. Some characteristics of the tested soil and the utilized biochar.
PropertyUnitSoilBiochar
PaulowniaBamboo
Clay(g kg−1)260--
Silt(g kg−1)320--
Sand(g kg−1)420--
Texture loam--
Organic matter(g kg−1)24.40840732
EC (1:5)(dS m‒1)0.901.800.75
pH (1:2.5) 5.5010.5010.00
Total Cd(mg kg−1)2.50ND *ND *
Total Pb(mg kg−1)9704.00ND *
Total Zn(mg kg−1)101040.0030.50
Total Cu(mg kg−1)3615.503.88
* Not detected, each value represents the mean of three replicates.
Table 2. Effect of biochar application on soil organic matter (%), soil pH (1:2.5), and electrical conductivity (EC, dSm−1).
Table 2. Effect of biochar application on soil organic matter (%), soil pH (1:2.5), and electrical conductivity (EC, dSm−1).
Biochar TypeBiochar RateOMpHEC
CK02.40 ± 0.05 d5.54 ± 0.10 d0.64 ± 0.10 b
PB23.58 ± 0.04 c5.75 ± 0.07 c0.34 ± 0.09 c
44.31 ± 0.30 b6.21 ± 0.11 a0.38 ± 0.11 c
65.47 ± 0.04 a6.048 ± 0.7 ab0.39 ± 0.08 c
BB23.51 ± 0.16 c5.58 ± 0.10 d0.65 ± 0.10 b
44.28 ± 0.23 b5.92 ± 0.08 b0.68 ± 0.10 b
65.21 ± 0.16 a6.05 ± 0.10 b0.76 ± 0.08 a
All values are the means of triplicate analysis. Within each column, values followed by the same letter are not significantly different at p < 0.05 level. CK, control (no biochar applied); PB, Paulownia biochar; BB, Bamboo biochar.
Table 3. Effects of biochar addition on concentration (mg kg−1) of the TCLP-extractable metals in the contaminated soil.
Table 3. Effects of biochar addition on concentration (mg kg−1) of the TCLP-extractable metals in the contaminated soil.
Biochar TypeBiochar RatePbCdZnCu
CK049.95 ± 0.63 a1.45 ± 0.20 a132.03 ± 1.77 a0.850 ± 0.08 a
PB242.06 ± 0.58 c1.26 ± 0.08 c113.51 ± 1.88 d0.698 ± 0.05 c
440.01 ± 0.76 d1.23 ± 0.09 d107.63 ± 1.67 f0.546 ± 0.08 d
638.90 ± 0.35 d1.21 ± 0.10 e99.94 ± 1.26 g0.470 ± 0.07 e
BB248.34 ± 0.44 b1.40 ± 0.08 b128.10 ± 1.23 b0.774 ± 0.08 b
442.51 ± 0.44 c1.39 ± 0.11 b123.98 ± 1.78 c0.774 ± 0.03 b
639.73 ± 0.51 d1.24 ± 0.09 d112.00 ± 1.43 e0.749 ± 0.07 b
All values are the means of triplicate analysis. Within each column, values followed by the same letter are not significantly different at p < 0.05 level. CK, control (no biochar applied); PB, Paulownia biochar; BB, Bamboo biochar.
Table 4. Concentration of the acid extractable fraction of soil heavy metals (mg kg−1) determined with the BCR sequential extraction method.
Table 4. Concentration of the acid extractable fraction of soil heavy metals (mg kg−1) determined with the BCR sequential extraction method.
Biochar TypeBiochar RatePbCdZnCu
CK0137.46 ± 0.05 a1.85 ± 0.04 a168.01 ± 1.00 a2.05 ± 0.08 a
PB2126.45 ± 0.034 b1.71 ± 0.04 b145.15 ± 3.00 c1.65 ± 0.7 b
4125.46 ± 0.05 c1.64 ± 0.03 bc144.92 ± 4.00 c1.53 ± 0.5 b
6123.48 ± 1.0 d1.60 ± 0.02 c141.75 ± 1.00 c1.27 ± 0.05 c
BB2129.13 ± 0.58 e1.68 ± 0.05 b155.84 ± 1.00 b1.66 ± 0.08 b
4124.50 ± 0.07 f1.66 ± 0.06 bc153.59 ± 1.00 b1.66 ± 0.07 b
6120.14 ± 0.55 g1.64 ± 0.04 bc151.79 ± 2.00 b1.41 ± 0.06 bc
All values are the means of triplicate analysis. Within each column, values followed by the same letter are not significantly different at p < 0.05 level. CK, control (no biochar applied); PB, Paulownia biochar; BB, Bamboo biochar.
Table 5. Concentration of the soil heavy metal fraction bound to Fe/Mn oxides (mg kg−1) determined with the BCR sequential extraction method.
Table 5. Concentration of the soil heavy metal fraction bound to Fe/Mn oxides (mg kg−1) determined with the BCR sequential extraction method.
Biochar TypeBiochar RatePbCdZnCu
CK0609.33 ± 16 a0.491 ± 0.05 ab85.16 ± 2.00 c6.71 ± 0.13 bc
PB2549.45 ± 6 d0.487 ± 0.06 ab89.04 ± 1.00 b6.45 ± 0.35 cd
4592.63 ± 5 ab0.498 ± 0.03 a97.89 ± 2.00 a6.25 ± 0.05 d
6557.78 ± 9 cd0.477 ± 0.02 ab97.76 ± 3.00 a5.85 ± 0.07 e
BB2576.00 ± 3 bc0.481 ± 0.04 ab90.77 ± 2.00 b7.14 ± 0.22 a
4572.67 ± 1 bc0.488 ± 0.05 ab94.60 ± 1.00 a7.04 ± 0.05 ab
6526.18 ± 15 e0.463 ± 0.05 b90.25 ± 1.00 b7.05 ± 0.08 ab
All values are the means of triplicate analysis. Within each column, values followed by the same letter are not significantly different at p < 0.05 level. CK, control (no biochar applied); PB, Paulownia biochar; BB, Bamboo biochar.
Table 6. Concentration of the soil heavy metal fraction bound to organic matter (mg kg−1) determined with the BCR sequential extraction method.
Table 6. Concentration of the soil heavy metal fraction bound to organic matter (mg kg−1) determined with the BCR sequential extraction method.
Biochar TypeBiochar RatePbCdZnCu
CK061.70 ± 2 c0.104 ± 0.02 d140.82 ± 1.00 f7.34 ± 0.08 d
PB291.33 ± 2 c0.147 ± 0.05 bc165.67 ± 2.00 e9.05 ± 0.22 c
498.62 ± 2 a0.164 ± 0.06 a185.11 ± 3.00 c10.45 ± 0.38 b
699.96 ± 4 a0.168 ± 0.07 a189.35 ± 2.00 b11.80 ± 0.58 a
BB298.96 ± 2 a0.144 ± 0.08 c172.41 ± 1.00 d8.49 ± 0.31 c
499.49 ± 3 a0.147 ± 0.08 bc189.90 ± 1.00 b8.79 ± 0.22 c
699.63 ± 3 a0.155 ± 0.05 b200.01 ± 1.00 a8.93 ± 0.08 c
All values are the means of triplicate analysis. Within each column, values followed by the same letter are not significantly different at p < 0.05 level. CK, control (no biochar applied); PB, Paulownia biochar; BB, Bamboo biochar.
Table 7. Concentration of the residual fraction of soil heavy metals (mg kg−1) determined with the BCR sequential extraction method.
Table 7. Concentration of the residual fraction of soil heavy metals (mg kg−1) determined with the BCR sequential extraction method.
Biochar TypeBiochar RatePbCdZnCu
CK0195.95 ± 2.00 a0.127 ± 0.05 b658.00 ± 15 a16.89 ± 0.10 a
PB2186.60 ± 2.00 bcd0.146 ± 0.06 a645.60 ± 7 a15.86 ± 0.11 a
4193.53 ± 3.00 ab0.146 ± 0.08 a667.15 ± 2 a15.79 ± 0.11 a
6180.81 ± 5.00 d0.146 ± 0.07 a649.75 ± 2 a15.25 ± 1.57 a
BB2192.06 ± 4.00 abc0.145 ± 0.04 a629.78 ± 3 a15.92 ± 0.33 a
4185.10 ± 4.00 cd0.141 ± 0.06 a667.77 ± 10 a15.31 ± 0.81 a
6173.99 ± 2.00 e0.140 ± 0.03 a664.73 ± 11 a15.63 ± 0.88 a
All values are the means of triplicate analysis. Within each column, values followed by the same letter are not significantly different at p < 0.05 level. CK, control (no biochar applied); PB, Paulownia biochar; BB, Bamboo biochar.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Awad, M.; Liu, Z.; Skalicky, M.; Dessoky, E.S.; Brestic, M.; Mbarki, S.; Rastogi, A.; EL Sabagh, A. Fractionation of Heavy Metals in Multi-Contaminated Soil Treated with Biochar Using the Sequential Extraction Procedure. Biomolecules 2021, 11, 448. https://doi.org/10.3390/biom11030448

AMA Style

Awad M, Liu Z, Skalicky M, Dessoky ES, Brestic M, Mbarki S, Rastogi A, EL Sabagh A. Fractionation of Heavy Metals in Multi-Contaminated Soil Treated with Biochar Using the Sequential Extraction Procedure. Biomolecules. 2021; 11(3):448. https://doi.org/10.3390/biom11030448

Chicago/Turabian Style

Awad, Mahrous, Zhongzhen Liu, Milan Skalicky, Eldessoky S. Dessoky, Marian Brestic, Sonia Mbarki, Anshu Rastogi, and Ayman EL Sabagh. 2021. "Fractionation of Heavy Metals in Multi-Contaminated Soil Treated with Biochar Using the Sequential Extraction Procedure" Biomolecules 11, no. 3: 448. https://doi.org/10.3390/biom11030448

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