Next Article in Journal
Photoluminescence of Ni(II), Pd(II), and Pt(II) Complexes [M(Me2dpb)Cl] Obtained from C‒H Activation of 1,5-Di(2-pyridyl)-2,4-dimethylbenzene (Me2dpbH)
Previous Article in Journal
Valorization of Tomato Waste as a Source of Carotenoids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Characterization Techniques as Supporting Tools for the Interpretation of Biochar Adsorption Efficiency in Water Treatment: A Critical Review

by
Michele Castiglioni
1,*,
Luca Rivoira
1,*,
Irene Ingrando
1,
Massimo Del Bubba
2 and
Maria Concetta Bruzzoniti
1,*
1
Department of Chemistry, University of Torino, Via Pietro Giuria 7, 10125 Torino, Italy
2
Department of Chemistry “Ugo Schiff”, University of Florence, Via della Lastruccia 3, 50019 Sesto Fiorentino, Italy
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(16), 5063; https://doi.org/10.3390/molecules26165063
Submission received: 26 July 2021 / Revised: 17 August 2021 / Accepted: 17 August 2021 / Published: 20 August 2021
(This article belongs to the Section Analytical Chemistry)

Abstract

:
Over the past decade, biochar (BC) has received significant attention in many environmental applications, including water purification, since it is available as a low-cost by-product of the energetic valorisation of biomass. Biochar has many intrinsic characteristics, including its porous structure, which is similar to that of activated carbon (AC), which is the most widely used sorbent in water treatment. The physicochemical and performance characteristics of BCs are usually non-homogenously investigated, with several studies only evaluating limited parameters, depending on the individual perspective of the author. Within this review, we have taken an innovative approach to critically survey the methodologies that are generally used to characterize BCs and ACs to propose a comprehensive and ready-to-use database of protocols. Discussion about the parameters of chars that are usually correlated with adsorption performance in water purification is proposed, and we will also consider the physicochemical properties of pollutants (i.e., Kow). Uniquely, an adsorption efficiency index BC/AC is presented and discussed, which is accompanied by an economic perspective. According to our survey, non-homogeneous characterization approaches limit the understanding of the correlations between the pollutants to be removed and the physicochemical features of BCs. Moreover, the investigations of BC as an adsorption medium necessitate dedicated parallel studies to compare BC characteristics and performances with those of ACs.

1. Introduction

Biochar (BC) is a low-cost solid by-product of the thermal conversion of feedstocks of a different nature, such as agricultural [1], wood residues [2], manure [3], and sludge [4].
The thermal conversion procedures available at this time to obtain biochar can be grouped into the following technologies, according to the processing parameters: fast [5] and slow pyrolysis [6], gasification [7], hydrothermal [8], and flash carbonization [9]. Biochar production mainly relies on the first two technologies. Valuable information for biochars produced through pyro-gasification processes is provided by the related literature [10]. We do not aim to exhaustively discuss the biochar production technologies, regardless of how useful they are, however it is useful to recognize that each requires varying temperature range, heating rate, pressure and residence time different one to the other. Accordingly, these technologies maximize different ratios and characteristics of the final products, i.e., gas (biogas), solid (char), and liquid (tar, oil). Further, gasification (a 700–1500 °C temperature range with residence time ranging from seconds to minutes) methods mainly converts biomass to biogas, and fast pyrolysis (a 400–600 °C temperature range with residence time of seconds) maximizes bio-oil formation, whereas slow pyrolysis (a 350–800 °C temperature range with residence time ranging from seconds to hours) generally favours the formation of biochar [11].
Over the years, biochar has attracted significant attention in many fields. Indeed, it shows great potential in addressing the issue of climate change through CO2 storage [12] and in establishing a circular economy model [13,14] in many sectors, such as agriculture, where it could be used to regulate the carbon, phosphorus, and nitrogen cycles in soils [15]. The first applications of biochar were directed toward agricultural purposes [16] since the addition of biochar improves soil properties. In Italy, the use of biochar as a soil amendment as well as the technical requirements for this use are governed by DL 75/2010 and further modifications. The “Standardized Product Definition and Product Testing Guidelines for Biochar That Is Used in Soil” was recently released by International Biochar Initiative (IBI) to prescribe the characteristics that biochar intended for sale must possess for safe use in soil.
In recent years, studies on biochar have been extended to include water purification issues due to the adsorption properties exhibited by this material and due to its low environmental impact [17]. This new interest in biochar has led to a progressive increase of publications in this field. The first two reviews on the application of biochar for water purification, dated 2014 [11,18], followed by more recent publications [19,20], have detailed the sorption of organic/inorganic contaminants grouped by classes to elucidate possible adsorption mechanisms based on the surface chemistry of biochars. The performance of biochar towards selected classes of compounds, mainly nutrients [21], antibiotics [22], and metals [23], was the main theme of the reviews that were subsequently published on water purification using biochar.
The concept of exploiting biochar for water treatment as a potential surrogate of activated carbon (AC), the commercial adsorbent that is the most widely used in the refining tertiary stage of water treatment technologies, has escalated the need to improve the sorption characteristics of biochar. These efforts are clearly highlighted by the literature and reviews that analyse strategies for the surface modification of biochar [24] or the production of biochar with the addition of nanoscale metals [25] to the biomass before thermal conversion. However, even if BCs are proposed to possibly replace ACs in water purification, only a limited number of studies that have already been published on this topic compare the performance of BCs with those of a commercial activated carbon, thus making it extremely difficult to highlight differences between both adsorbents and the advantages or disadvantages in the use of one over the other.
Although there is established information for the use of biochars in agricultural applications, there have been no publications produced by international organizations on biochar characterization for other applications, including for water purification. Despite authors studying the removal of compounds by BCs agree with the fact that physicochemical and performance characterization is of paramount importance to understand pollutant–BC interactions [26,27], a lack of homogenous scientific approaches is observed in the current literature due to the absence of any standardization.
On this premise, it becomes relevant to compare biochar and activated carbon as adsorbents for organic micropollutants in water purification. This review addresses the way in which the two media are characterized within the literature. To the best of our knowledge, this is the first review aiming to collect and discuss the most used physicochemical and performance characterization techniques and to propose a detailed and easy-to-use compendium. In addition, the adsorption capacities of unmodified BCs and ACs toward the organic compounds investigated in the literature are compared and discussed in terms of the physicochemical properties of pollutants (i.e., Kow) and an efficiency BCs/ACs ratio is derived and then discussed in terms of the economic value of the two adsorbents. Moreover, the possibility of a chemometric treatment of BC adsorption capacity towards organic compounds together with the main physicochemical properties of BCs is also discussed. To the best of our knowledge, all of these aspects are innovatively investigated in this review.

2. Activated Carbon and Biochar: The Different Approach to Characterization

When studied for water purification purposes, it seems reasonable to compare the adsorption performance and the physicochemical features of biochar with those of activated carbon, which is by far the most widely used regulated adsorbent that is commercially available. However, the overview of the literature currently available on biochar indicates that this is not the approach followed by authors, as biochar is synthesized and tested for the removal of selected single [28,29] or mixed [30,31] compounds.
To better compare the capabilities of biochar in respect to activated carbon in water filtration, we will describe the characterization methods that are currently available for activated carbon and those that are primarily used for biochar.

2.1. Activated Carbon Characterization

The characterization of activated carbon is necessary in order to identify the materials that are suitable for use in the tertiary potabilization treatments or wastewater treatment plants.
The characterization of activated carbons is obtained through two core alternative approaches, which are based on the determination of (i) adsorption performance and (ii) the physicochemical parameters. The first approach determines the structure-dependent indices that are estimated in respect to standardized compounds of the proper molecular dimensions that are directly correlated to the adsorption capabilities of the activated carbon itself [32]. The second approach allows the collection of information on the main structural properties, i.e., the morphology, porosity distribution, and nature of the chemical groups present on the surface.
The adsorption performance parameters (or performance indices) of activated carbons are strictly regulated when such materials are utilized in drinking water purification processes. Many official documents, which will be briefly reviewed, specify the requirements that active carbons should meet when used for water filtration together with the methods to evaluate such performance indices. Most of these documents were released by the American Society for Testing and Materials International (ASTM), the American Water Works Association (AWWA), the German Institute for Standardization (DIN), and by the International Organization for Standardization (ISO). Thanks to the work of the European Chemical Industry Council (CEFIC), many of these methods are collected in the publication “Test Methods for Activated Carbon” [33].
Experiments to be performed on activated carbons are divided into three groups: physical tests, adsorption tests, and chemical/physicochemical tests. In the next part of this review, a brief description of the regulated parameters that need to be tested will be provided, along with the documents in which each parameter is defined. These methods are collected in Table 1.

2.1.1. Physical Tests

Through physical tests, apparent or bulk density, absolute and particle density, particle size, pressure drop, and mechanical strength are determined. Apparent or bulk density (expressed as kg/m3 on dry basis) is defined as the mass of the unit volume of the sample in air, including both the pore system and the voids between the particles. In contrast, absolute density and particle density (also called He and Hg density, respectively) are defined as the mass of the unit volume of the solid carbon skeleton not accessible to He (for absolute density) or of the carbon particle (for particle density). Both variables are expressed as g/mL. The density parameters are necessary to evaluate the shape and size of the activated carbon particles as well as to evaluate the packing volume, bed porosity, and void fraction [46,47]. Methods to determine bulk density are described in ASTM standard [34], whereas methods to measure bulk density, absolute and particle density, particle size, pressure drop, and mechanical strength are collected by CEFIC standard [33]. Many authors agree that the monitoring of density parameters during the synthesis of activated carbons should be exploited to follow the thermal degradation of the raw precursor material [48,49].
Particle size (expressed in mm) in activated carbons influences both the adsorption of target compounds and certain mechanical properties, i.e., hydraulic conductivity and flow speed [50]. The official methods from ASTM [35] and CEFIC require the mechanical separation of the particles through sieves. It should be noted that the particle size as well as the surface morphology and porosity of the adsorbent surface [51] could also be evaluated by using physiochemical techniques, such as scanning electron microscopy (SEM) and transmission electron microscopy (TEM), as detailed in the following sections.
In addition to particle size, other characteristics such as pressure drop, resistance to flow, and mechanical strength should also be determined. The pressure drop of a gas flow over a packed bed gives information regarding a gas’ resistance to flow through the carbon layer, and the mechanical strength simulates the resistance to abrasion or friction under real conditions [36,37]. Both the particle size and mechanical strength parameters are determined for granular active carbons, which are defined by CEFIC as those having 90% of their particles larger than 0.18 mm, as determined by the aforementioned particle size test.

2.1.2. Adsorption Tests

While physical tests are necessary to design drinking water filters (e.g., open gravity types), adsorption tests are required to evaluate the removal performances of activated carbons. These approaches usually combine the evaluation of the adsorption isotherm of a given adsorbate–adsorbent system and the fit to a theoretical or empirical model of the adsorption process to estimate the adsorption characteristics of activated carbons. The Freundlich model is suggested by ASTM [38,39] and CEFIC [33] to determine the adsorptive capacity of activated carbons. It can be expressed by the formula, X/M = KF × Cf1/n, where X/M (mg/g) is the amount of the target pollutant removed per unit mass of carbon, Cf (mg/L) is the residual concentration of the pollutant after treatment, and KF ((mg/g)/(g/m3)1/n) and 1/n (adimensional) are the constants for a given adsorption system. The same equation is generally used in the linearized logarithmic form.
The adsorption performances of activated carbons are expressed through the determination of selected indices, namely iodine [33,40,41], phenol [33], methylene blue [33], molasses [43], and tannin [41] numbers. These parameters give information about the porous structure and the adsorption properties of the activated carbon towards target compounds with similar dimensions to the probe molecules.
To elaborate, the iodine number is specifically defined as the milligrams of iodine adsorbed from an aqueous solution by 1 g of activated carbon when the iodine concentration of the residual filtrate is 0.02 N. Procedures to evaluate this parameter have been proposed both by UNI [52] and by AWWA [37], only differing in the mathematical treatment of the results. In fact, in the UNI EN 12902 standard [52], the iodine number is extrapolated from the linear regression model obtained by plotting the mg of iodine adsorbed by three different amounts of carbons versus the residual iodine concentration, while in the AWWA B600-78 standard, the parameter is obtained by introducing a tabulated correction factor, which depends on the residual iodine normality of the filtrate.
When comparing both of these protocols to calculate the iodine number, the AWWA B600-78 standard is both less time consuming and less laborious. However, the correction factor is derived according to the residual concentrations typically obtained by activated carbons. This implies that, as a general consideration, the method cannot be applied to materials with adsorbing capacities that are substantially different to those of activated carbon since the residual iodine concentrations would be outside the range of those specified by AWWA. UNI EN 12902, alternately, does not require any mathematical treatment and has the advantage of being applicable to a wider range of adsorbent materials of different capacities. However, as UNI EN 12902 does require the determination of an adsorption isotherm, this approach is obviously more time consuming than that described by the AWWA method that was previously mentioned.
The phenol number is defined as the phenol adsorption for a singular weight unit of carbon when, after adsorption, the phenol concentration in a solution decreases from 10 to 1 mg/L.
The methylene blue index—defined as the methylene blue volume (in mL) of a 1.2 g/L solution adsorbed by a known amount of sorbent (0.1 g) within a prearranged contact time (5 min)—was demonstrated to be directly correlated to the specific surface area (SSA) and micropore volume of the adsorbent [53]. Official procedures to evaluate the aforementioned indices have been detailed by the ASTM and the Water Research Commission (WRC) [54,55].
The iodine and phenol indices are commonly related to the presence of micropores (<2 nm), and they are therefore considered to be informative of the effectiveness in removing small-sized organic water pollutants. On the other hand, the methylene blue index is linked to the abundance of mesopores (2–50 nm) and thus is a useful indicator of the adsorption capacities for medium–large-sized organic pollutants [56].
The tannin number is defined as the concentration of activated carbon (mg/L) that is required to reduce the concentration of a standard tannic acid solution from 20 to 2 mg/L. Tannins, a mix of medium- and large-sized molecules, can be efficiently adsorbed by the mesopores and macropores (>50 nm) of the activated carbons. The tannin values of different activated carbon substrates and the pore volume of mesopores are linearly correlated [57], and carbons with low tannin numbers exhibit the highest quality of removing high molecular weight impurities.
The molasses number is a measure of the degree of decolorization of a standard molasses solution that has been diluted and standardized against standardized activated carbon. The molasses number represents the potential pore volume that is available for larger adsorbing species.
Of these adsorption tests, the iodine index is the only one required by the UNI EN 12915-1 standard for products used for the treatment of water intended for human consumption and for which a specific requirement is set (>600 mg/g).
An additional adsorption test for activated carbons, also collected in the CEFIC report, is the phenazone index. Since it is not directly linked to the removal of pollutants from waters but is instead linked to pharmaceutical purposes [58,59,60], the phenazone index will not be further detailed.

2.1.3. Chemical and Physicochemical Tests

Chemical tests are expected when activated carbon must be used for drinking water filtration in potabilization plants, as they are utilised to assess the purity criteria for adsorbents. As an example, according to the EN 12915-1 standard [40], ashes, water-soluble material, and water-extractable substances (As, Cd, Cr, Hg, Ni, Pb, Sb, Se, CN-, fluoranthene, benzo[b]fluoranthene, benzo[k]fluoranthene, benzo[a]pyrene, benzo[g,h,i]perylene, indeno[1,2,3-cd]pyrene) must be evaluated for activated carbons. These substances represent the most probable impurities that may be present in the adsorbent as a result of both raw material composition [61] and thermal process conditions [62]. Limits are provided for these parameters. As far as the determination of water-extractable substances is concerned, extraction is performed in a solution containing sodium hydrogen carbonate, calcium chloride, and magnesium sulphate [40,52]. It is important to underline that users should be notified of the presence of other impurities not included in the standard should be, as defined by the same standard. Notably, the determination of the aforementioned polycyclic aromatic hydrocarbons (PAHs) and metal content in activated carbon is foreseen by the CEFIC procedure [33] under more drastic conditions, such as cyclohexane extraction and total oxidation, respectively.
Strictly related to the operating procedures and start-up operations of the refining stage inside the plant is the pH value of the solutions in contact with the adsorbent. The pH conditions strongly influence the duration of the washing procedures of the carbon before it reaches its full operability. Inorganic and chemically active groups on the carbon surface are responsible for the possible modification of the pH when the substrate is placed in contact with water. Among the procedures available for the determination of this parameter, the ASTM D6851-02 and D3838-05 standards [44,45] consider the contact pH and the pH of a boiled water extract, respectively, in a 1:10 w/v ratio.
In addition to the techniques detailing the characterization of activated carbons that have been presented above, the literature that is currently available uses other techniques and approaches with the aim of fully understanding surface and bulk chemistry and of speculating on possible interactions with the target compounds. As an example, the evaluation of the pH of zero-point charge (PZC), the pH value at which the net charge density of the material is equal to zero, could be important to define the adsorption behaviour of a material towards pollutants of different charges at different pH conditions [63].
Additional information could also be obtained through Boehm’s titration, by which the equivalent of surface acidic/basic functionalities (carboxyl, lactone, phenolic groups) per gram of carbon can be obtained [64]. This approach gives deeper insight into surface chemistry.
Directly connected to the surface charge of the material, the cationic exchange capacity (CEC) is defined as the total amount of the exchangeable cations of a sorbent [65,66]. Microscopy (both scanning and transmitting, SEM and TEM), nitrogen adsorption isotherms, and Fourier transform infrared spectroscopy (FTIR) are usually applied in order to derive a topographical morphology of the material and to define the particle dimension and porous structure as well as to study the surface functional groups and structure [67].
An additional characterization of activated carbons is based on the analysis of the elemental composition. From the amount of carbon, oxygen, and hydrogen, it is possible to derive the Van Krevelen diagram, in which the O/C ratio is plotted against the H/C ratio [68]. The resultant graph is used to classify carbons and their feedstock [69] and to fully understand the evolution of carbon during the heating treatment, thus predicting specific the properties of carbon adsorbents (degree of carbonization by H/C ratio, hydrophilicity by O/C ratio, polarity by (O + N)/C ratio) [70].

2.2. Biochar Characterization

2.2.1. Chemical and Physicochemical Tests

At this time, no official method for the evaluation of the performance and characteristics of biochars intended for water filtration has been published. It is of the authors’ opinion that, in the absence of indications by regulatory bodies, official procedures previously presented for activated carbons (EN, CEFIC, AWWA standards, see Table 1) should be homogenously applied to test the performance of biochars. This approach would ensure the coherence of the results presented in the literature and the straightforward comparison among the biochars produced and a comparison between biochar and activated carbon.
Studies dealing with the removal of contaminants from water using biochar show that biochar is synthesized from different types of feedstock and different thermal process conditions. The extent of adsorption of different pollutants is generally correlated with certain properties of the biochar that are specifically measured to characterize the biochar produced.
Table 2 collects the main approaches and instrumental techniques used for the characterization of biochar according to the literature of the last decade concerning the removal of organic pollutants using biochar. The following are indicative, though not exhaustive, examples of the characterization of biochars that have been tested for water filtration.
Many authors [71,72] apply nitrogen adsorption isotherms to derive information about surface area, pore volume, and average pore size. We have previously pointed out that these parameters are clearly correlated with the widely discussed adsorption indices for activated carbons, which have rarely been determined for biochar. A thorough search of the relevant literature yielded few studies presenting this correlation for biochar [56,73], which is in contrasty to what has been presented for activated carbon [32,74,75]. Of note, recent literature now measures the adsorption indices, i.e., the iodine index, as an indicator of the best adsorption performance of biochars obtained under different processing conditions (e.g., heating rate). The iodine and methylene blue indices have been proposed as comparative indicators for the performance of activated carbons and biochars [76].
The physicochemical methods used for the characterization of biochars mainly depend on the objectives of the study in question. However, these methods are used to define the surface properties of biochars and to highlight possible adsorption mechanisms that are responsible for biochar–pollutant affinity [71,77]. Among these methods, FTIR, XPS, and XRD are the most frequently applied to characterize surface chemistry.
FTIR spectroscopy is a widespread technique used to investigate the surface chemical functional groups (e.g., aliphatic or aromatic nature) in biochar [78]. Through the FTIR spectrum, vibration bands can be assigned to defined functional groups that enable speculation on possible interaction mechanisms with the pollutants of interest [11]. Additional information gained from FTIR is the understanding of the reactions that occurred during thermal treatment steps [70,79,80].
XPS analysis provides information regarding the chemical composition and bonds on the biochar surface to a surface depth < 10 nm and on the relative abundance of different species of certain elements on the surface, e.g., different C and N containing groups and bonds (C-C, C-H, C=O, -COOH, N-C, amino acid N, and ammonium-N) on the biochar surface [81,82].
The structure and phase composition of biochars can be derived by XRD analysis. The information that is gained can be important for determining if the successful preparation of modified biochar occurred [77] or to understand the evolution of the thermal process with respect to the initial biomass composition [83].
Van Krevelen plots, which were previously discussed in the activated carbon section, are extensively employed in biochar characterization to define biochar composition as a function of the thermal conditions used for biochar production [70,84,85].
Table 2. Main characterization methods employed for biochar according to the recent literature.
Table 2. Main characterization methods employed for biochar according to the recent literature.
TestRef Biochar
Adsorption Indices
Iodine number[56,76,86,87,88]
Methylene Blue number[56,76,89]
Chemical Tests
Total PAHs[90]
pH, ash[56,89]
Surface and Functional Groups Characterization
Nitrogen-adsorption isotherm[30,56,71,91,92]
FTIR[56,89,92,93,94]
SEM[92,94,95]
XRD[56,92,93,95]
Zero-point charge[56,93,94,96]
Boehm’s titration[89]
Cation exchange capacity[89]
Elemental Composition
H/C, O/C, (O + N)/C (Van Krevelen plots)[56,84,85,95]
Kinetic and isotherm studies[56,97,98,99,100]

2.2.2. Adsorption Tests

The careful study of the current literature provides strong indications that the assessment of the adsorption features of biochars is mainly performed and that the studies in which the adsorption performance of the biochars are simultaneously compared with that of activated carbon are extremely limited. This approach makes understanding the benefits or limitations of using biochar in place of activated carbon incredibly complex. In this regard, it is of the authors’ opinion that research on the use of biochar for water refining should be addressed a priori by always including activated carbon for a direct comparison.
Materials intended for the removal of pollutants from waters are usually described with reference to their retention performance through kinetics and adsorption isotherms [101].
The optimization of the contact time is essential for further adsorption studies to ensure the presence of equilibrium conditions within pollutant-adsorbent (biochar/activated carbon) systems. Kinetic studies have also provided additional information, including the identification of rate controlling steps such as (i) the transport of the solute molecules from the aqueous phase to the surface of the solid (film or external diffusion); (ii) the transfer of the solutes from the surface to the intra-particle sites (intra-particle diffusion); (iii) and the adsorption of the solutes on the interior surfaces of the adsorbent [102]. Pseudo-first order and pseudo-second order models are the most frequently computed models. However, there are some less common works that thoroughly investigate the adsorption performances of BCs through the Elovich model [103] (predicting the mass and surface diffusion and the activation/deactivation energy of the system) or through the intraparticle diffusion and Boyd models [104,105] (which determine the effect of internal and external diffusion on the adsorption mechanism).
Once the equilibrium conditions are identified, adsorption isotherm data are obtained to describe the type of interactions involved between adsorbates and adsorbents and to quantify the extent of the adsorption (i.e., the adsorbent capacity). The correlation of the isotherm data using theoretical or empirical equations is useful for practical operation. In fact, a proper understanding and interpretation of adsorption isotherms is crucial for the overall improvement of the adsorption mechanism pathways and the effective design of an adsorption system [106]. Because of its wide applicability, linear regression analysis is frequently used to fit experimental data and to assess adsorption performance in the biochar related literature [107,108]. The Langmuir, Freundlich, and Tempkin models are the most applied models [97,98,99,109,110,111,112,113], with limited works also investigating the Dubinin-Radushkevich [114] and Sips [115] models. Non-linear regression analysis [116], which has also been widely used by a number of researchers in an attempt to minimize the gap between the predicted and experimental data, is also applied to biochar studies [117]. The description of the aforementioned models [118,119] is out of the scope of the present review.
Our survey showed that recent publications on adsorption onto BCs dedicate significant effort to the study of adsorption kinetics and isotherms. In contrast, simultaneous studies on both BCs and ACs are only rarely considered for comparison [56,99,120], resulting in a consequent lack of information.

2.2.3. Adsorption Tests in Dynamic Bench Scale and Pilot Scale Conditions

In addition to what has previously been presented, where the adsorption performances of BCs were reported for batch conditions, the evaluation of kinetic and isotherm models in dynamic conditions by bench scale and pilot scale systems is of paramount importance in order to simulate real BC applications in filtering systems and to achieve technology readiness levels (TRL) that are higher than an experimental proof of concept.
Despite this premise, few studies evaluate the retention of pollutants from water matrices on unmodified BCs using small lab-scaled columns [121], and even fewer studies investigate the adsorption isotherms [122]
In dynamic testing conditions, kinetic models should be replaced by breakthrough curves, in which the profile of the effluent adsorptive concentration is continuously measured at the outlet of a fixed bed adsorber. It should be noted that these curves can be considered as the last of the essential characterizations of an activated carbon since they simulate performance in an industrial application [123]. Whilst it seems reasonable to adopt the same curves to evaluate the performances of BCs as adsorbents for aqueous solutions, most studies work at laboratory scale conditions [124]. This is most likely due to an undeveloped biochar technology.
When considering pilot or pre-commercial scaled systems, in which higher volumes of water are treated using biochar packed columns, the calculation of adsorption performance is only obtained through the transposition of isotherm models from batch tests to pilot scale conditions [121].
Finally, it is interesting to highlight a lack of adsorption performance studies in bench- or pilot-scaled conditions towards organic pollutants, which is the object of our review. In comparison, greater attention is devoted to the adsorption of metal ions [125,126,127,128]. It is therefore evident that further research and insights are necessary for TRL improvement.

2.2.4. Leaching Tests

As previously described for ACs (Section 2.1.3), leaching tests for organic (PAHs) and inorganic compounds (metals) are necessary to evaluate the safety of the supports employed for the removal of pollutants in water treatment [52]. Therefore, for BCs that are intended to be used for water filtration, considering that these adsorbents are produced by thermal treatment similar to ACs, then the same characterizations should be suggested. Despite this assumption, few papers have evaluated the leaching of some classes of compounds in aqueous medium from biochars studied for water treatment [56,129]. Specifically, Hong and co-workers [129] extensively describe the release of selected heavy metals from biochar in ultrapure water together with nutrients and total organic carbon as representative indicators of organic matter without focusing on any specific class of compounds. In a pioneering vision, Del Bubba and co-workers [56] followed the analytical protocols imposed by UNI EN 12915 [40] for evaluating the leaching characteristics of metals and PAHs while comparing BC vs. AC performance.
Apart from the limited works that have been previously noted, the actual literature shows a disproportionate number of papers exploring the use of leaching tests in BCs used as an amendment in agricultural applications [130,131] over water treatment. This highlights the different maturity stages of BCs used in agriculture compared to those used in water treatment.

3. Biochar as Adsorbent for Organic Micropollutants

3.1. Typical Target Pollutants and Performance Capabilities

The application of biochar as an adsorbent for the removal of organic pollutants from water and wastewater has been extensively studied. A search of the Web of Science database for the words “biochar” and “adsorption” yields approximately 5800 studies, with about being 40% focused on the removal of heavy metals (research refinement using “metal”) and the other 40% being focused on organic compounds (research refinement using both “organic” and “pollutants”). However, despite the volume, a critical approach investigating the possible correlation between biochar adsorption capacities and both the intrinsic properties of target compounds (i.e., polar character) and/or the physicochemical features through statistic tools is missing. Therefore, the following paragraphs will focus on a critical investigation on this topic, limiting the discussion to the literature of the last ten to fifteen years relating to the removal of organic pollutants by unmodified biochars. As previously mentioned, since activated carbon is the standard material for water treatment, ACs will be considered in parallel.
From a subset of this literature, a representative list of the organic compounds that are the most frequently removed by biochars was retrieved and summarized in Table 3. Papers were only selected if adsorption capacity data at equilibrium conditions were presented. Each analyte is clustered depending on its ionic form (undissociated, anionic or cationic) at a neutral pH value, which is typical of raw wastewater. The logKow and the pka (if available) values are reported. Within each reviewed paper the comparison of BC performances with those of AC is reported if present.
The data collected in Table 3 allow several considerations that will be discussed.
Most of the organic pollutants tested for removal by biochars are dyes, herbicides, or drugs, as they account for the most detected species in wastewater.
Regarding the ionizability properties, the reported data clearly show that at the pH value typical for wastewater and natural water (pH ~ 7), most of the organic molecules that are typically investigated are present in neutral form. However, it should be noted that, among the neutral compounds that are considered, many of them are ionizable species with pKa values higher than 7, thus suggesting that the chemistry of their functional groups will be altered at different pH conditions with consequent variation in the adsorption mechanism [164] and in BC removal performance.
The analysis of the reviewed literature shows that most of the studies do not consider ACs in their investigations, limiting the discussion to the selected BCs. This approach makes it difficult to assess BC performance and its advantages over more traditional adsorbents, i.e., AC. Therefore, to understand and to critically discuss BC capabilities compared to AC, the range of the typical adsorption capacities of BCs and ACs towards the most studied organic pollutants was derived through a distributional statistical approach. The maximum adsorption capacity at equilibrium conditions (qe expressed as mg/kg) derived from data from adsorption isotherm tests (see Section 2.2.2) was chosen as a representative parameter for adsorption efficiency. Statistical treatment was not limited to the data in Table 3 (restricted to the last 10–15 years of literature) but was extended to less recent literature to achieve a representative population of 50 studies. The collected qe data for BCs were treated using a boxplot tool (Figure 1a) that enabled an evaluation of the data dispersion and the main statistical indices (i.e., average, median, etc). In a similar manner, a representative data population of qe for ACs was reported (Figure 2b, n = 50 as for BCs). It is important to highlight that due to the scarcity of studies that simultaneously include BCs and ACs, to achieve the same data numerosity (n = 50), it was necessary to also include studies that were only related to ACs.
Graphical results demonstrate that the inter quartile range (IQR-green box), which is between 25th (Q1) and 75th quartile (Q3), obtained for the qe values of ACs is wider than the one obtained for BCs. Indeed, for BCs, the Q1 and Q3 values are 10,000 mg/kg and 150,000 mg/kg, respectively, with an IQR area (Q3-Q1) of 140,000 mg/kg, while for ACs, the Q1 and Q3 values are 21,400 mg/kg and 266,400 mg/kg, respectively, with an IQR area of 245,000 mg/kg. These data clearly highlight that activated carbons show a wider range of adsorption capacities in respect to BCs of about 50% and higher qe values. A similar gap is also observed when considering the average capacity (95,000 mg/kg and 176,800 mg/kg for BCs and ACs, respectively).
As proposed in the Introduction section, these observations support of the need to increase research on properly modified BCs.

3.2. Effect of Hydrophobic Character of Analytes on Biochar Adsorption Efficiencies

When considering the intrinsic properties of the compounds under investigation, one of the most important elements to be accounted for is the hydrophobic/hydrophilic characteristics, since they strongly influence the type and the strength of the interaction occurring between the biochar surface groups and the target pollutants. Therefore, to produce evidence of a possible correlation trend between the adsorption capacity of BCs and the hydrophobic characteristics of the organic compounds that are being removed, a distribution graph was chosen. For this purpose, the distribution of BC qe (mg/kg) data taken from the literature (n = 50) was plotted as a function of the logKow values of the target pollutants to be removed (Figure 2A). It should be specified that, from all of the data that was previously retrieved (Table 3), only the molecules having a logKow ranging from 0.5 (±0.25) to 3.5 (±0.25) were chosen for statistical treatment since those outside of those values have a reduced frequency, rendering their processing insignificant. The logKow values of the selected target compounds are referred to slightly hydrophobic to medium-high hydrophobic analytes.
The same distribution was also represented for the same molecules adsorbed on activated carbons (n = 50) in Figure 2.
To elaborate, in both graphs (Figure 2A,B), the bars represent the min–max qe frequency range for each 0.5 unit of the logKow cluster (±0.25) in which all of the target compounds were divided. The black squares represent the average qe, and the black lines represent the average qe trend line.
The data representation in Figure 2 leads to interesting observations. When focusing on the average qe values, a Gaussian-similar trend could be observed for BCs (Figure 2A). In fact, molecules characterized by logKow ranging from 1.5 to 2 are more retained from BCs than those characterized by lower and higher hydrophobic characteristics. Conversely, similar behaviour is not observed when studying the removal of the same compounds using ACs (Figure 2B) since adsorption increases with the increase of logKow, reaching a plateau at logKow that is approximately equal to 3.0 (medium–high hydrophobicity).
Such results could be explained by examining the different interactions between the organic pollutants and the biochars or activated carbons as a result of the different surface properties of the two adsorbents. In fact, it can be determined that the surface of ACs is characterized by a higher degree of aromaticity with respect to the surface of BCs since their production temperature and activation processes are demonstrated to convert aliphatic carbons into aromatic carbons [164,166]. This feature is reflected in the preferred π-π and/or hydrophobic interactions in ACs [167] rather than BCs. Consequently, the higher the aromaticity of the molecule and the higher the hydrophobicity of the target pollutants, the stronger their interactions, and, therefore, the higher the adsorption capacities of the activated carbons will be.
Alternatively, due to their typical process production conditions, BC is characterized by a surface containing both aliphatic ionizable moieties (i.e., -COOH), and, to a lesser extent, aromatic carbons. In fact, the lower pyrolysis temperature conditions in BC production reduce the conversion of the aliphatic C to aromatic C, so the H/C ratio is typically higher in biochars [168] than in activated carbons [169]. Therefore, a wider range of mechanisms, such as electrostatic, π-π/hydrophobic, and hydrogen bond interactions, are involved within the pollutant and the BC surface [164]. Consequently, organic pollutants characterized by a medium range hydrophobic character, i.e., logKow of about 1.5–2 (Figure 2B, which displays both polar and nonpolar functional groups) are expected to interact with the biochar surface, synergically exploiting all of the previously mentioned mechanisms. Conversely, target compounds with lower logKow values are expected to interact with BC via a reduced number of interactions (electrostatic, if ionizable, or hydrogen bond interactions), thus resulting in lower adsorption on BCs. In contrast, pollutants with higher logKow values are expected to interact with BC via π-π or hydrophobic interactions, which only limit the extent of retention.
The distribution plots reported here show that the use of biochar in water filtration and the activated carbon replacement could be suggested for the removal of medium non-polar compounds (i.e., herbicides), for which the affinity is similar to activated carbons (average qe about 2.5 × 105 mg/kg for both sorbents). In this regard, the use of highly energy demanding activation processes typical of ACs could be avoided.

4. Considerations on the Correlations of Physicochemical and Performance Indices with Biochar Adsorption Efficiencies through Chemometric Approaches

As described in Section 2 of this review, several physicochemical and performance indices of biochars strictly influence adsorption performances. Selected authors have tried to correlate only one biochar property at a time (i.e., pyrolysis temperature [6], feedstock [170], etc.) with the adsorption efficiencies. Nevertheless, an overall evaluation that considers the broadest possible physicochemical parameters and indices is still missing. A principal component analysis (PCA), a chemometric approach frequently applied to directly highlight the correlation status of several variables, would be highly desirable to evaluate the possible correlation between the main biochar features described in Section 2.1 and Section 2.2 and the adsorption capabilities expressed, for instance, as the maximum adsorption capacities derived from the Langmuir models (see Section 2.2.2) or as the KF derived from the Freundlich models (see Section 2.1.2). Despite the volume of papers devoted to the removal of organic compounds by biochars, a lack of homogenous data on physicochemical and performance characterization is observed since most of the studies only evaluate a few of the parameters that were previously mentioned (Section 2) depending on each author’s particular requirement. For this reason, when trying to set-up a PCA matrix by retrieving data from the available literature, the absence for relevant data makes chemometric treatment impossible [171]. To overcome this obstacle, a possible solution to perform chemometric PCA treatment could be to either work with the data from a single study, in which, for example several biochars are tested, or to reduce the variables to be computed to those few that are in common among the selected works. However, in both cases, the number of treated samples or variables to be considered would be so reduced that significant results from PCA could not be obtained.

5. Economic Evaluation on Biochar Use over Activated Carbon

For the compounds included in this review, the consideration reported in Section 3.1 showed that the adsorption capacities for ACs are superior to BCs. However, removal effectiveness is not the only parameter to consider when evaluating the applicability of adsorbents in wastewater treatment. Production costs should also be considered.
We have derived an effectiveness/cost index, weighting the adsorption performances with the production costs, evaluated according to the literature survey. To further elaborate, it has been recently reported that the AC production costs range from 1.35 (China) [172] to 2.6 EUR/Kg (Italy) [56]. For BCs, these values range from 0.15 (Italy) [56] to 0.42 EUR/Kg (U.S.) [173,174], with average values of 1.98 EUR/Kg and 0.29 EUR/Kg, respectively. On the basis of the mean qe values discussed above in Section 3.1, (95,000 mg/kg and 176,800 mg/kg for BCs and ACs, respectively), an effectiveness/cost index has been calculated for both BCs and ACs, dividing the average maximum equilibrium adsorption capacity (qe, expressed as mg/kg) by the average cost per Kg of adsorbent. The results (3.3 × 105 mg removed for each EUR of BC and 8.9 × 104 mg/EUR for ACs) show that although BCs have lower adsorption capabilities, their effectiveness/cost index is 3.7 times higher than that of activated carbons, which means that for the same capital investment, a removal of organic pollutants that is approximately 4-fold higher (in terms of mg) is obtained using BCs compared to ACs.

6. Conclusions

In the last ten–fifteen years, biochars intended for water remediation have become a trending topic, with thousands of published works studying their possible application for the removal of both organic and inorganic contaminants. The data reviewed and depicted using boxplot statistics indicate that BCs exhibit slightly lower adsorption capacities for organic compounds, especially for those of medium–high hydrophobicity, in comparison with standard adsorption materials (activated carbons). Despite the reduced adsorption capacities of BCs, the production cost is significantly lower, with a calculated effectiveness/cost index that is about 4 times higher for BCs in respect to ACs.
BC removal efficiencies could be correlated with the main properties of the target compounds, such as the hydrophobic characteristics as expressed by the logKow values. According to our survey, for medium non-polar compounds, the replacement of ACs with BCs in tertiary treatments appears feasible since average BC adsorption efficiencies are in the same range as those of AC. This finding supports the implementation of column experiments for establishing the maximum loading capacity of BCs in experimental conditions that are more similar to a real-life scale.
The correlation of the intrinsic physicochemical and performance parameters with biochar removal capacity is more challenging. The presented review highlighted a lack of homogenous physicochemical and performance characterization techniques, which, in our opinion, are mandatory for an overall detailed understanding of the core interactions and correlations between the pollutants and the BC surface. A correlation of this type is absent in the current literature.
The comprehensive survey on the activated carbon characterization protocols (most of them proposed by regulatory organisations), as reviewed in the first part of this paper, proves that by using these procedures, a full understanding of the capability and safe use of ACs in the removal of pollutants from water can be achieved. It is consequently desirable that, because of the lack of regulatory methods specific to biochars, all protocols for ACs should be also homogenously applied for the characterization of BCs intended for water purification, thereby avoiding fragmentary and subjective characterization that only depends on the predilection of the authors. This approach would allow the verification of whether the biochars prepared in the various experimental conditions comply with the requirements set out in the aforementioned standards relating to the absorbent materials intended to be used for the filtration of drinking water. A direct comparison between the capabilities of BCs and ACs is also recommended to explore the scale up of BC applications (bench and pilot scales) towards higher TRL levels.
Additionally, through this overview, an easy-to-use compendium on the main physicochemical and performance characterization techniques of both BCs and ACs is provided.
Finally, the currently available literature data for BCs are non-homogeneous which makes it impossible to extract detailed information through the main chemometric tools, such as PCA. Developing a body of knowledge of the main physicochemical and performance characteristics of the BC that is as complete as possible is therefore of fundamental importance to statistically evaluate the possible correlations between the removal capacity of BCs and intrinsic BC physicochemical properties.

Author Contributions

Conceptualization, M.C.B.; software, M.C. and L.R.; validation, M.C., L.R. and M.C.B.; investigation, M.C., L.R., M.D.B. and M.C.B.; data curation, M.C. and L.R.; writing—original draft preparation, M.C., L.R., M.D.B. and M.C.B.; writing—review and editing, M.C., L.R., M.C.B.; visualization, I.I.; supervision, M.C.B.; funding acquisition, M.C.B. All authors have read and agreed to the published version of the manuscript.

Funding

Funding by Finpiemonte-Regione Piemonte is gratefully acknowledged (BIO-ENPRO4TO. SMART SOLUTIONS FOR SMART COMMUNITIES: bio-energies and sustainable products of Green Chemistry (output) from the integrated valorisation of urban/industrial wastes, biomasses and wastewaters (input), through functional intelligent modules-333-201).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The Authors acknowledge Susanna Robert-Randazzo and Domenico Randazzo for the language revision.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Colantoni, A.; Evic, N.; Lord, R.; Retschitzegger, S.; Proto, A.R.; Gallucci, F.; Monarca, D. Characterization of biochars produced from pyrolysis of pelletized agricultural residues. Renew. Sustain. Energy Rev. 2016, 64, 187–194. [Google Scholar] [CrossRef] [Green Version]
  2. Wang, Y.; Hu, Y.; Zhao, X.; Wang, S.; Xing, G. Comparisons of Biochar Properties from Wood Material and Crop Residues at Different Temperatures and Residence Times. Energy Fuels 2013, 27, 5890–5899. [Google Scholar] [CrossRef]
  3. Cao, X.; Harris, W. Properties of dairy-manure-derived biochar pertinent to its potential use in remediation. Bioresour. Technol. 2010, 101, 5222–5228. [Google Scholar] [CrossRef]
  4. Méndez, A.; Cárdenas-Aguiar, E.; Paz-Ferreiro, J.; Plaza, C.; Gascó, G. The effect of sewage sludge biochar on peat-based growing media. Biol. Agric. Hortic. 2017, 33, 40–51. [Google Scholar] [CrossRef]
  5. DeSisto, W.J.; Hill, N.; Beis, S.H.; Mukkamala, S.; Joseph, J.; Baker, C.; Ong, T.-H.; Stemmler, E.A.; Wheeler, M.C.; Frederick, B.G. Fast pyrolysis of pine sawdust in a fluidized-bed reactor. Energy Fuels 2010, 24, 2642–2651. [Google Scholar] [CrossRef]
  6. Ahmad, M.; Lee, S.S.; Dou, X.; Mohan, D.; Sung, J.-K.; Yang, J.E.; Ok, Y.S. Effects of pyrolysis temperature on soybean stover-and peanut shell-derived biochar properties and TCE adsorption in water. Bioresour. Technol. 2012, 118, 536–544. [Google Scholar] [CrossRef]
  7. Brewer, C.E.; Schmidt-Rohr, K.; Satrio, J.A.; Brown, R.C. Characterization of biochar from fast pyrolysis and gasification systems. Environ. Prog. Sustain. Energy 2009, 28, 386–396. [Google Scholar] [CrossRef]
  8. Rillig, M.C.; Wagner, M.; Salem, M.; Antunes, P.M.; George, C.; Ramke, H.-G.; Titirici, M.-M.; Antonietti, M. Material derived from hydrothermal carbonization: Effects on plant growth and arbuscular mycorrhiza. Appl. Soil Ecol. 2010, 45, 238–242. [Google Scholar] [CrossRef]
  9. Antal, M.J.; Mochidzuki, K.; Paredes, L.S. Flash carbonization of biomass. Ind. Eng. Chem. Res. 2003, 42, 3690–3699. [Google Scholar] [CrossRef]
  10. Meyer, S.; Glaser, B.; Quicker, P. Technical, economical, and climate-related aspects of biochar production technologies: A literature review. Environ. Sci. Technol. 2011, 45, 9473–9483. [Google Scholar] [CrossRef]
  11. Mohan, D.; Sarswat, A.; Ok, Y.S.; Pittman, C.U., Jr. Organic and inorganic contaminants removal from water with biochar, a renewable, low cost and sustainable adsorbent—A critical review. Bioresour. Technol. 2014, 160, 191–202. [Google Scholar] [CrossRef] [PubMed]
  12. Woolf, D.; Amonette, J.E.; Street-Perrott, F.A.; Lehmann, J.; Joseph, S. Sustainable biochar to mitigate global climate change. Nat. Commun. 2010, 1, 56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Hu, Q.; Jung, J.; Chen, D.; Leong, K.; Song, S.; Li, F.; Mohan, B.C.; Yao, Z.; Prabhakar, A.K.; Lin, X.H. Biochar industry to circular economy. Sci. Total Environ. 2021, 757, 143820. [Google Scholar] [CrossRef] [PubMed]
  14. Khawkomol, S.; Neamchan, R.; Thongsamer, T.; Vinitnantharat, S.; Panpradit, B.; Sohsalam, P.; Werner, D.; Mrozik, W. Potential of Biochar Derived from Agricultural Residues for Sustainable Management. Sustainability 2021, 13, 8147. [Google Scholar] [CrossRef]
  15. Pan, S.-Y.; Dong, C.-D.; Su, J.-F.; Wang, P.-Y.; Chen, C.-W.; Chang, J.-S.; Kim, H.; Huang, C.-P.; Hung, C.-M. The Role of Biochar in Regulating the Carbon, Phosphorus, and Nitrogen Cycles Exemplified by Soil Systems. Sustainability 2021, 13, 5612. [Google Scholar] [CrossRef]
  16. Atkinson, C.J.; Fitzgerald, J.D.; Hipps, N.A. Potential mechanisms for achieving agricultural benefits from biochar application to temperate soils: A review. Plant Soil 2010, 337, 1–18. [Google Scholar] [CrossRef]
  17. Thompson, K.A.; Shimabuku, K.K.; Kearns, J.P.; Knappe, D.R.; Summers, R.S.; Cook, S.M. Environmental comparison of biochar and activated carbon for tertiary wastewater treatment. Environ. Sci. Technol. 2016, 50, 11253–11262. [Google Scholar] [CrossRef] [PubMed]
  18. Ahmad, M.; Rajapaksha, A.U.; Lim, J.E.; Zhang, M.; Bolan, N.; Mohan, D.; Vithanage, M.; Lee, S.S.; Ok, Y.S. Biochar as a sorbent for contaminant management in soil and water: A review. Chemosphere 2014, 99, 19–33. [Google Scholar] [CrossRef]
  19. Abbas, Z.; Ali, S.; Rizwan, M.; Zaheer, I.E.; Malik, A.; Riaz, M.A.; Shahid, M.R.; ur Rehman, M.Z.; Al-Wabel, M.I. A critical review of mechanisms involved in the adsorption of organic and inorganic contaminants through biochar. Arab. J. Geosci. 2018, 11, 448. [Google Scholar] [CrossRef]
  20. Dai, Y.; Zhang, N.; Xing, C.; Cui, Q.; Sun, Q. The adsorption, regeneration and engineering applications of biochar for removal organic pollutants: A review. Chemosphere 2019, 223, 12–27. [Google Scholar] [CrossRef]
  21. Gong, H.; Tan, Z.; Zhang, L.; Huang, Q. Preparation of biochar with high absorbability and its nutrient adsorption–desorption behaviour. Sci. Total Environ. 2019, 694, 133728. [Google Scholar] [CrossRef] [PubMed]
  22. Li, C.; Zhu, X.; He, H.; Fang, Y.; Dong, H.; Lü, J.; Li, J.; Li, Y. Adsorption of two antibiotics on biochar prepared in air-containing atmosphere: Influence of biochar porosity and molecular size of antibiotics. J. Mol. Liq. 2019, 274, 353–361. [Google Scholar] [CrossRef]
  23. Ni, B.-J.; Huang, Q.-S.; Wang, C.; Ni, T.-Y.; Sun, J.; Wei, W. Competitive adsorption of heavy metals in aqueous solution onto biochar derived from anaerobically digested sludge. Chemosphere 2019, 219, 351–357. [Google Scholar] [CrossRef]
  24. Cheng, N.; Wang, B.; Wu, P.; Lee, X.; Xing, Y.; Chen, M.; Gao, B. Adsorption of emerging contaminants from water and wastewater by modified biochar: A review. Environ. Pollut. 2021, 273, 116448. [Google Scholar] [CrossRef] [PubMed]
  25. Hasan, M.S.; Geza, M.; Vasquez, R.; Chilkoor, G.; Gadhamshetty, V. Enhanced heavy metal removal from synthetic stormwater using nanoscale zerovalent iron–modified biochar. Water Air Soil Pollut. 2020, 231, 1–15. [Google Scholar] [CrossRef]
  26. Yaashikaa, P.; Kumar, P.S.; Varjani, S.; Saravanan, A. A critical review on the biochar production techniques, characterization, stability and applications for circular bioeconomy. Biotechnol. Rep. 2020, 28, e00570. [Google Scholar] [CrossRef]
  27. Yargicoglu, E.N.; Sadasivam, B.Y.; Reddy, K.R.; Spokas, K. Physical and chemical characterization of waste wood derived biochars. Waste Manag. 2015, 36, 256–268. [Google Scholar] [CrossRef] [PubMed]
  28. Chen, T.; Luo, L.; Deng, S.; Shi, G.; Zhang, S.; Zhang, Y.; Deng, O.; Wang, L.; Zhang, J.; Wei, L. Sorption of tetracycline on H3PO4 modified biochar derived from rice straw and swine manure. Bioresour. Technol. 2018, 267, 431–437. [Google Scholar] [CrossRef]
  29. Qian, W.C.; Luo, X.P.; Wang, X.; Guo, M.; Li, B. Removal of methylene blue from aqueous solution by modified bamboo hydrochar. Ecotoxicol. Environ. Saf. 2018, 157, 300–306. [Google Scholar] [CrossRef] [PubMed]
  30. Li, Y.; Meas, A.; Shan, S.; Yang, R.; Gai, X.; Wang, H.; Tsend, N. Hydrochars from bamboo sawdust through acid assisted and two-stage hydrothermal carbonization for removal of two organics from aqueous solution. Bioresour. Technol. 2018, 261, 257–264. [Google Scholar] [CrossRef] [PubMed]
  31. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Guo, W.; Johir, M.A.H.; Sornalingam, K. Single and competitive sorption properties and mechanism of functionalized biochar for removing sulfonamide antibiotics from water. Chem. Eng. J. 2017, 311, 348–358. [Google Scholar] [CrossRef]
  32. Wei, Z.; Qi-Gang, C.; Wan-Dong, L.; Bing-Jing, L.; Wen-Xin, J.; Li-Jun, F.; Wei-Chi, Y. Selecting activated carbon for water and wastewater treatability studies. Environ. Prog. 2007, 26, 289–298. [Google Scholar]
  33. European Chemical Industry Council (CEFIC). Test Methods for Activated Carbon; European Council of Chemical Manufacturers’ Federation (CEFIC): Bruxelles, Belgium, 1986. [Google Scholar]
  34. ASTM D2854–09, Standard Test Method for Apparent Density of Activated Carbon; ASTM International: West Conshohocken, PA, USA, 2014; Available online: www.astm.org (accessed on 13 April 2021).
  35. ASTM D2862–16, Standard Test Method for Particle Size Distribution of Granular Activated Carbon; ASTM International: West Conshohocken, PA, USA, 2016; Available online: www.astm.org (accessed on 13 April 2021).
  36. ASTM D3802–16, Standard Test Method for Ball-Pan Hardness of Activated Carbon; ASTM International: West Conshohocken, PA, USA, 2016; Available online: www.astm.org (accessed on 13 April 2021).
  37. AWWA B604, Standard for Granular Activated Carbon; American Water Works Association: Denver, CO, USA, 2012; Available online: www.awwa.org (accessed on 13 April 2021).
  38. ASTM D3860-98, Standard Practice for Determination of Adsorptive Capacity of Activated Carbon by Aqueous Phase Isotherm Technique; ASTM International: West Conshohocken, PA, USA, 2014; Available online: www.astm.org (accessed on 13 April 2021).
  39. ASTM D 5919-96, Standard Practice for Determination of Adsorptive Capacity of Activated Carbon by a Micro-Isotherm Technique for Adsorbates at ppb Concentrations; ASTM International: West Conshohocken, PA, USA, 2017; Available online: www.astm.org (accessed on 13 April 2021).
  40. EN 12915-1, Products Used for the Treatment of Water Intended for Human Consumption-Granular Activated Carbon-Part 1: Virgin Granular Activated Carbon; European Committee for Standardization: Brussels, Belgium, 2009; Available online: www.cen.eu (accessed on 13 April 2021).
  41. AWWA B600-16, Powdered Activated Carbon; American Water Works Association: Denver, CO, USA, 2016; Available online: www.awwa.org (accessed on 13 April 2021).
  42. ASTM D4607-14, Standard Test Method for Determination of Iodine Number of Activated Carbon; ASTM International: West Conshohocken, PA, USA, 2014; Available online: www.astm.org (accessed on 13 April 2021).
  43. EPA 625171002A, Process Design Manual for Carbon Adsorption; U.S. Environmental Protection Agency: Washington, DC, USA, 1973. Available online: www.epa.gov (accessed on 13 April 2021).
  44. ASTM D3838-05, Standard Test Method for pH of Activated Carbon; ASTM International: West Conshohocken, PA, USA, 2017; Available online: www.astm.org (accessed on 13 April 2021).
  45. ASTM D6851-02, Standard Test Method for Determination of Contact pH with Activated Carbon; ASTM International: West Conshohocken, PA, USA, 2011; Available online: www.astm.org (accessed on 13 April 2021).
  46. Plantard, G.; Goetz, V.; Py, X. A direct method for porous particle density characterization applied to activated carbons. Adv. Powder Technol. 2010, 21, 592–598. [Google Scholar] [CrossRef]
  47. Ahmedna, M.; Johns, M.; Clarke, S.; Marshall, W.; Rao, R. Potential of agricultural by-product-based activated carbons for use in raw sugar decolourisation. J. Sci. Food Agric. 1997, 75, 117–124. [Google Scholar] [CrossRef]
  48. Caturla, F.; Molina-Sabio, M.; Rodriguez-Reinoso, F. Preparation of activated carbon by chemical activation with ZnCl2. Carbon 1991, 29, 999–1007. [Google Scholar] [CrossRef]
  49. Rodriguez-Reinoso, F.; Molina-Sabio, M. Activated carbons from lignocellulosic materials by chemical and/or physical activation: An overview. Carbon 1992, 30, 1111–1118. [Google Scholar] [CrossRef]
  50. Malusis, M.A.; Barben, E.J.; Evans, J.C. Hydraulic conductivity and compressibility of soil-bentonite backfill amended with activated carbon. J. Geotech. Geoenviron. Eng. 2009, 135, 664–672. [Google Scholar] [CrossRef]
  51. Achaw, O.-W. A study of the porosity of activated carbons using the scanning electron microscope. In Scanning Electron Microscopy; InTech: London, UK, 2012. [Google Scholar]
  52. EN 12902, Products Used for Treatment of Water Intended for Human Consumption. Inorganic Supporting and Filtering Materials; European Committee for Standardization: Brussels, Belgium, 2004.
  53. Nunes, C.A.; Guerreiro, M.C. Estimation of surface area and pore volume of activated carbons by methylene blue and iodine numbers. Química Nova 2011, 34, 472–476. [Google Scholar] [CrossRef] [Green Version]
  54. ASTM D 1783-70, Phenolic Compounds in Water; ASTM International: West Conshohocken, PA, USA, 2017; Available online: www.astm.org (accessed on 13 April 2021).
  55. Water Research Commission. WRC Report No. K5/1124; Water Research Commission: Pretoria, South Africa, 2003. [Google Scholar]
  56. Del Bubba, M.; Anichini, B.; Bakari, Z.; Bruzzoniti, M.C.; Camisa, R.; Caprini, C.; Checchini, L.; Fibbi, D.; El Ghadraoui, A.; Liguori, F. Physicochemical properties and sorption capacities of sawdust-based biochars and commercial activated carbons towards ethoxylated alkylphenols and their phenolic metabolites in effluent wastewater from a textile district. Sci. Total Environ. 2020, 708, 135217. [Google Scholar] [CrossRef]
  57. Wu, Y.; Xia, C.; Cai, L.; Shi, S.Q. Controlling pore size of activated carbon through self-activation process for removing contaminants of different molecular sizes. J. Colloid Interface Sci. 2018, 518, 41–47. [Google Scholar] [CrossRef]
  58. Wu, M.; Guo, Q.; Fu, G. Preparation and characteristics of medicinal activated carbon powders by CO2 activation of peanut shells. Powder Technol. 2013, 247, 188–196. [Google Scholar] [CrossRef]
  59. Ronowicz, J.; Kupcewicz, B.; Pałkowski, L.; Krysiński, J. Development and optimization of the activated charcoal suspension composition based on a mixture design approach. Acta Pharm. 2015, 65, 83–90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Flament, M.; M’bisi, R.; Leterme, P.; Gayot, A. Formulation of activated charcoal for per os administration to addicted subjects. Drug Dev. Ind. Pharm. 2000, 26, 731–736. [Google Scholar] [CrossRef] [PubMed]
  61. Freddo, A.; Cai, C.; Reid, B.J. Environmental contextualisation of potential toxic elements and polycyclic aromatic hydrocarbons in biochar. Environ. Pollut. 2012, 171, 18–24. [Google Scholar] [CrossRef] [PubMed]
  62. Rafiq, M.; Bachmann, R.; Rafiq, M.T.; Shang, Z.; Joseph, S.; Long, R. Influence of Pyrolysis Temperature on Physico-Chemical Properties of Corn Stover (Zea mays L.) Biochar and Feasibility for Carbon Capture and Energy Balance. PLoS ONE 2016, 11, e0156894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. To, M.H.; Hadi, P.; Hui, C.W.; Lin, C.S.K.; McKay, G. Mechanistic study of atenolol, acebutolol and carbamazepine adsorption on waste biomass derived activated carbon. J. Mol. Liq. 2017, 241, 386–398. [Google Scholar] [CrossRef]
  64. Contescu, A.; Contescu, C.; Putyera, K.; Schwarz, J.A. Surface acidity of carbons characterized by their continuous pK distribution and Boehm titration. Carbon 1997, 35, 83–94. [Google Scholar] [CrossRef]
  65. Youssef, A.; El-Nabarawy, T.; Samra, S. Sorption properties of chemically-activated carbons: 1. Sorption of cadmium (II) ions. Colloids Surf. A Physicochem. Eng. Asp. 2004, 235, 153–163. [Google Scholar] [CrossRef]
  66. Munera-Echeverri, J.L.; Martinsen, V.; Strand, L.T.; Zivanovic, V.; Cornelissen, G.; Mulder, J. Cation exchange capacity of biochar: An urgent method modification. Sci. Total Environ. 2018, 642, 190–197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Yang, J.; Qiu, K. Preparation of activated carbons from walnut shells via vacuum chemical activation and their application for methylene blue removal. Chem. Eng. J. 2010, 165, 209–217. [Google Scholar] [CrossRef]
  68. Van Krevelen, D. Graphical-statistical method for the study of structure and reaction processes of coal. Fuel 1950, 29, 269–284. [Google Scholar]
  69. Benstoem, F.; Becker, G.; Firk, J.; Kaless, M.; Wuest, D.; Pinnekamp, J.; Kruse, A. Elimination of micropollutants by activated carbon produced from fibers taken from wastewater screenings using hydrothermal carbonization. J. Environ. Manag. 2018, 211, 278–286. [Google Scholar] [CrossRef]
  70. Tan, X.; Liu, Y.; Zeng, G.; Wang, X.; Hu, X.; Gu, Y.; Yang, Z. Application of biochar for the removal of pollutants from aqueous solutions. Chemosphere 2015, 125, 70–85. [Google Scholar] [CrossRef]
  71. Fan, S.; Wang, Y.; Wang, Z.; Tang, J.; Tang, J.; Li, X. Removal of methylene blue from aqueous solution by sewage sludge-derived biochar: Adsorption kinetics, equilibrium, thermodynamics and mechanism. J. Environ. Chem. Eng. 2017, 5, 601–611. [Google Scholar] [CrossRef]
  72. Li, R.; Wang, Z.; Guo, J.; Li, Y.; Zhang, H.; Zhu, J.; Xie, X. Enhanced adsorption of ciprofloxacin by KOH modified biochar derived from potato stems and leaves. Water Sci. Technol. 2018, 77, 1127–1136. [Google Scholar] [CrossRef] [Green Version]
  73. Zhang, J.; Tian, Y.; Yin, L.; Zhang, J.; Drewes, J.E. Insight into the effects of biochar as adsorbent and microwave receptor from one-step microwave pyrolysis of sewage sludge. Environ. Sci. Pollut. Res. 2018, 25, 18424–18433. [Google Scholar] [CrossRef] [PubMed]
  74. Gu, H.; Bergman, R.; Anderson, N.; Alanya-Rosenbaum, S. Life cycle assessment of activated carbon from woody biomass. Wood Fiber Sci. 2018, 50, 229–243. [Google Scholar] [CrossRef] [Green Version]
  75. Gu, J.; Zhou, J.; Ma, H.; Ma, M.; Xing, M. Characteristics of camellia shell pyrolysis products and optimization of preparation parameters of activated carbon. Nongye Gongcheng Xuebao Trans. Chin. Soc. Agric. Eng. 2015, 31, 233–239. [Google Scholar]
  76. Strubinger, A.; Oliveros, A.R.; Araque, M.A.; Guerra, J. Assessment of the energy recovery of Aloe Vera solid residues by pyrolysis and hydrothermal conversion. Chem. Eng. Trans. 2017, 57, 19–24. [Google Scholar]
  77. Yin, Z.; Liu, Y.; Liu, S.; Jiang, L.; Tan, X.; Zeng, G.; Li, M.; Liu, S.; Tian, S.; Fang, Y. Activated magnetic biochar by one-step synthesis: Enhanced adsorption and coadsorption for 17β-estradiol and copper. Sci. Total Environ. 2018, 639, 1530–1542. [Google Scholar] [CrossRef]
  78. Igalavithana, A.D.; Mandal, S.; Niazi, N.K.; Vithanage, M.; Parikh, S.J.; Mukome, F.N.D.; Rizwan, M.; Oleszczuk, P.; Al-Wabel, M.; Bolan, N.; et al. Advances and future directions of biochar characterization methods and applications. Crit. Rev. Environ. Sci. Technol. 2017, 47, 2275–2330. [Google Scholar] [CrossRef]
  79. Maiti, S.; Purakayastha, S.; Ghosh, B. Production of Low-Cost Carbon Adsorbents from Agricultural Wastes and Their Impact on Dye Adsorption. Chem. Eng. Commun. 2007, 195, 386–403. [Google Scholar] [CrossRef]
  80. Long, L.; Xue, Y.; Zeng, Y.; Yang, K.; Lin, C. Synthesis, characterization and mechanism analysis of modified crayfish shell biochar possessed ZnO nanoparticles to remove trichloroacetic acid. J. Clean. Prod. 2017, 166, 1244–1252. [Google Scholar] [CrossRef]
  81. Lin, Y.; Munroe, P.; Joseph, S.; Ziolkowski, A.; van Zwieten, L.; Kimber, S.; Rust, J. Chemical and structural analysis of enhanced biochars: Thermally treated mixtures of biochar, chicken litter, clay and minerals. Chemosphere 2013, 91, 35–40. [Google Scholar] [CrossRef] [PubMed]
  82. Yang, F.; Gao, Y.; Sun, L.; Zhang, S.; Li, J.; Zhang, Y. Effective sorption of atrazine by biochar colloids and residues derived from different pyrolysis temperatures. Environ. Sci. Pollut. Res. 2018, 25, 18528–18539. [Google Scholar] [CrossRef]
  83. Komnitsas, K.A.; Zaharaki, D. Morphology of Modified Biochar and Its Potential for Phenol Removal from Aqueous Solutions. Front. Environ. Sci. 2016, 4, 26. [Google Scholar] [CrossRef] [Green Version]
  84. Vithanage, M.; Mayakaduwa, S.S.; Herath, I.; Ok, Y.S.; Mohan, D. Kinetics, thermodynamics and mechanistic studies of carbofuran removal using biochars from tea waste and rice husks. Chemosphere 2016, 150, 781–789. [Google Scholar] [CrossRef]
  85. Oliveira, F.R.; Patel, A.K.; Jaisi, D.P.; Adhikari, S.; Lu, H.; Khanal, S.K. Environmental application of biochar: Current status and perspectives. Bioresour. Technol. 2017, 246, 110–122. [Google Scholar] [CrossRef]
  86. Tan, Z.; Wang, Y.; Kasiulienė, A.; Huang, C.; Ai, P. Cadmium removal potential by rice straw-derived magnetic biochar. Clean Technol. Environ. Policy 2017, 19, 761–774. [Google Scholar] [CrossRef]
  87. Sun, J.; He, F.; Pan, Y.; Zhang, Z. Effects of pyrolysis temperature and residence time on physicochemical properties of different biochar types. Acta Agric. Scand. Sect. B-Soil Plant Sci. 2017, 67, 12–22. [Google Scholar] [CrossRef]
  88. Lawal, A.A.; Hassan, M.A.; Zakaria, M.R.; Yusoff, M.Z.M.; Norrrahim, M.N.F.; Mokhtar, M.N.; Shirai, Y. Effect of oil palm biomass cellulosic content on nanopore structure and adsorption capacity of biochar. Bioresour. Technol. 2021, 332, 125070. [Google Scholar] [CrossRef]
  89. Carrier, M.; Hardie, A.G.; Uras, Ü.; Görgens, J.; Knoetze, J. Production of char from vacuum pyrolysis of South-African sugar cane bagasse and its characterization as activated carbon and biochar. J. Anal. Appl. Pyrolysis 2012, 96, 24–32. [Google Scholar] [CrossRef]
  90. Hilber, I.; Blum, F.; Leifeld, J.; Schmidt, H.-P.; Bucheli, T.D. Quantitative Determination of PAHs in Biochar: A Prerequisite to Ensure Its Quality and Safe Application. J. Agric. Food Chem. 2012, 60, 3042–3050. [Google Scholar] [CrossRef] [PubMed]
  91. Yao, Y.; Gao, B.; Chen, H.; Jiang, L.; Inyang, M.; Zimmerman, A.R.; Cao, X.; Yang, L.; Xue, Y.; Li, H. Adsorption of sulfamethoxazole on biochar and its impact on reclaimed water irrigation. J. Hazard. Mater. 2012, 209–210, 408–413. [Google Scholar] [CrossRef]
  92. Shaaban, A.; Se, S.-M.; Mitan, N.M.M.; Dimin, M. Characterization of biochar derived from rubber wood sawdust through slow pyrolysis on surface porosities and functional groups. Procedia Eng. 2013, 68, 365–371. [Google Scholar] [CrossRef] [Green Version]
  93. Essandoh, M.; Kunwar, B.; Pittman, C.U., Jr.; Mohan, D.; Mlsna, T. Sorptive removal of salicylic acid and ibuprofen from aqueous solutions using pine wood fast pyrolysis biochar. Chem. Eng. J. 2015, 265, 219–227. [Google Scholar] [CrossRef]
  94. KS, A.S.; Shyang, O.; Iberahim, N.; Mohamed, A. Kinetic and Isotherm Studies Biochar on Ammonium Solution. Preprint 2020. [Google Scholar] [CrossRef]
  95. Suliman, W.; Harsh, J.B.; Abu-Lail, N.I.; Fortuna, A.-M.; Dallmeyer, I.; Garcia-Perez, M. Influence of feedstock source and pyrolysis temperature on biochar bulk and surface properties. Biomass Bioenergy 2016, 84, 37–48. [Google Scholar] [CrossRef]
  96. Lu, H.; Zhang, W.; Wang, S.; Zhuang, L.; Yang, Y.; Qiu, R. Characterization of sewage sludge-derived biochars from different feedstocks and pyrolysis temperatures. J. Anal. Appl. Pyrolysis 2013, 102, 137–143. [Google Scholar] [CrossRef]
  97. Zhang, B.; Wu, Y.; Cha, L. Removal of methyl orange dye using activated biochar derived from pomelo peel wastes: Performance, isotherm, and kinetic studies. J. Dispers. Sci. Technol. 2020, 41, 125–136. [Google Scholar] [CrossRef]
  98. Zhang, Y.P.; Adi, V.S.K.; Huang, H.-L.; Lin, H.-P.; Huang, Z.-H. Adsorption of metal ions with biochars derived from biomass wastes in a fixed column: Adsorption isotherm and process simulation. J. Ind. Eng. Chem. 2019, 76, 240–244. [Google Scholar] [CrossRef]
  99. Shin, J.; Lee, S.-H.; Kim, S.; Ochir, D.; Park, Y.; Kim, J.; Lee, Y.-G.; Chon, K. Effects of physicochemical properties of biochar derived from spent coffee grounds and commercial activated carbon on adsorption behavior and mechanisms of strontium ions (Sr2+). Environ. Sci. Pollut. Res. 2020, 28, 40623–40632. [Google Scholar] [CrossRef] [PubMed]
  100. Choudhary, V.; Patel, M.; Pittman, C.U., Jr.; Mohan, D. Batch and Continuous Fixed-Bed Lead Removal Using Himalayan Pine Needle Biochar: Isotherm and Kinetic Studies. ACS Omega 2020, 5, 16366–16378. [Google Scholar] [CrossRef] [PubMed]
  101. Bruzzoniti, M.C.; de Carlo, R.M.; Sarzanini, C.; Caldarola, D.; Onida, B. Novel insights in Al-MCM-41 precursor as adsorbent for regulated haloacetic acids and nitrate from water. Environ. Sci. Pollut. Res. 2012, 19, 4176–4183. [Google Scholar] [CrossRef] [Green Version]
  102. Bruzzoniti, M.C.; Appendini, M.; Onida, B.; Castiglioni, M.; Del Bubba, M.; Vanzetti, L.; Jana, P.; Sorarù, G.D.; Rivoira, L. Regenerable, innovative porous silicon-based polymer-derived ceramics for removal of methylene blue and rhodamine B from textile and environmental waters. Environ. Sci. Pollut. Res. 2018, 25, 10619–10629. [Google Scholar] [CrossRef]
  103. Zand, A.D.; Abyaneh, M.R. Adsorption of Lead, manganese, and copper onto biochar in landfill leachate: Implication of non-linear regression analysis. Sustain. Environ. Res. 2020, 30, 1–16. [Google Scholar] [CrossRef]
  104. Zhao, S.; Ta, N.; Wang, X. Absorption of Cu (II) and Zn (II) from Aqueous Solutions onto Biochars Derived from Apple Tree Branches. Energies 2020, 13, 3498. [Google Scholar] [CrossRef]
  105. Choudhary, B.; Paul, D. Isotherms, kinetics and thermodynamics of hexavalent chromium removal using biochar. J. Environ. Chem. Eng. 2018, 6, 2335–2343. [Google Scholar] [CrossRef]
  106. Aygün, A.; Yenisoy-Karakaş, S.; Duman, I. Production of granular activated carbon from fruit stones and nutshells and evaluation of their physical, chemical and adsorption properties. Microporous Mesoporous Mater. 2003, 66, 189–195. [Google Scholar] [CrossRef]
  107. Yao, H.; Lu, J.; Wu, J.; Lu, Z.; Wilson, P.C.; Shen, Y. Adsorption of fluoroquinolone antibiotics by wastewater sludge biochar: Role of the sludge source. Water Air Soil Pollut. 2013, 224, 1370. [Google Scholar] [CrossRef]
  108. Cheng, G.; Sun, L.; Jiao, L.; Peng, L.-x.; Lei, Z.-h.; Wang, Y.-x.; Lin, J. Adsorption of methylene blue by residue biochar from copyrolysis of dewatered sewage sludge and pine sawdust. Desalin. Water Treat. 2013, 51, 7081–7087. [Google Scholar] [CrossRef]
  109. Park, J.-H.; Wang, J.J.; Kim, S.-H.; Kang, S.-W.; Cho, J.-S.; Delaune, R.D.; Ok, Y.S.; Seo, D.-C. Lead sorption characteristics of various chicken bone part-derived chars. Environ. Geochem. Health 2019, 41, 1675–1685. [Google Scholar] [CrossRef]
  110. Wu, L.; Zhang, S.; Wang, J.; Ding, X. Phosphorus retention using iron (II/III) modified biochar in saline-alkaline soils: Adsorption, column and field tests. Environ. Pollut. 2020, 261, 114223. [Google Scholar] [CrossRef]
  111. Shao, F.; Zhang, X.; Sun, X.; Shang, J. Antibiotic removal by activated biochar: Performance, isotherm, and kinetic studies. J. Dispers. Sci. Technol. 2020, 42, 1–12. [Google Scholar] [CrossRef]
  112. Solanki, A.; Boyer, T.H. Physical-chemical interactions between pharmaceuticals and biochar in synthetic and real urine. Chemosphere 2019, 218, 818–826. [Google Scholar] [CrossRef] [PubMed]
  113. Lam, Y.Y.; Lau, S.S.; Wong, J.W. Removal of Cd (II) from aqueous solutions using plant-derived biochar: Kinetics, isotherm and characterization. Bioresour. Technol. Rep. 2019, 8, 100323. [Google Scholar] [CrossRef]
  114. Taghlidabad, R.H.; Sepehr, E.; Khodaverdiloo, H.; Samadi, A.; Rasouli-Sadaghiani, M.H. Characterization of cadmium adsorption on two cost-effective biochars for water treatment. Arab. J. Geosci. 2020, 13, 1–10. [Google Scholar]
  115. Jung, K.-W.; Jeong, T.-U.; Choi, J.-W.; Ahn, K.-H.; Lee, S.-H. Adsorption of phosphate from aqueous solution using electrochemically modified biochar calcium-alginate beads: Batch and fixed-bed column performance. Bioresour. Technol. 2017, 244, 23–32. [Google Scholar] [CrossRef]
  116. Ayawei, N.; Ebelegi, A.N.; Wankasi, D. Modelling and Interpretation of Adsorption Isotherms. J. Chem. 2017, 2017, 11. [Google Scholar] [CrossRef]
  117. Ahmad, M.; Lee, S.S.; Rajapaksha, A.U.; Vithanage, M.; Zhang, M.; Cho, J.S.; Lee, S.-E.; Ok, Y.S. Trichloroethylene adsorption by pine needle biochars produced at various pyrolysis temperatures. Bioresour. Technol. 2013, 143, 615–622. [Google Scholar] [CrossRef]
  118. Kajjumba, G.W.; Emik, S.; Öngen, A.; Özcan, H.K.; Aydın, S. Modelling of adsorption kinetic processes—Errors, theory and application. In Advanced Sorption Process Applications; IntechOpen: London, UK, 2018; pp. 187–206. [Google Scholar]
  119. Qiu, H.; Lv, L.; Pan, B.-C.; Zhang, Q.-J.; Zhang, W.-M.; Zhang, Q.-X. Critical review in adsorption kinetic models. J. Zhejiang Univ. Sci. A 2009, 10, 716–724. [Google Scholar] [CrossRef]
  120. Jung, C.; Park, J.; Lim, K.H.; Park, S.; Heo, J.; Her, N.; Oh, J.; Yun, S.; Yoon, Y. Adsorption of selected endocrine disrupting compounds and pharmaceuticals on activated biochars. J. Hazard. Mater. 2013, 263, 702–710. [Google Scholar] [CrossRef] [PubMed]
  121. Ashoori, N.; Teixido, M.; Spahr, S.; LeFevre, G.H.; Sedlak, D.L.; Luthy, R.G. Evaluation of pilot-scale biochar-amended woodchip bioreactors to remove nitrate, metals, and trace organic contaminants from urban stormwater runoff. Water Res. 2019, 154, 1–11. [Google Scholar] [CrossRef] [PubMed]
  122. Ulrich, B.A.; Im, E.A.; Werner, D.; Higgins, C.P. Biochar and activated carbon for enhanced trace organic contaminant retention in stormwater infiltration systems. Environ. Sci. Technol. 2015, 49, 6222–6230. [Google Scholar] [CrossRef]
  123. Marsh, H.; Rodríguez-Reinoso, F. CHAPTER 4-Characterization of Activated Carbon. In Activated Carbon; Marsh, H., Rodríguez-Reinoso, F., Eds.; Elsevier Science Ltd.: Oxford, UK, 2006; pp. 143–242. [Google Scholar]
  124. Kearns, J.; Dickenson, E.; Knappe, D. Enabling organic micropollutant removal from water by full-scale biochar and activated carbon adsorbers using predictions from bench-scale column data. Environ. Eng. Sci. 2020, 37, 459–471. [Google Scholar] [CrossRef]
  125. Park, J.-H.; Cho, J.-S.; Ok, Y.S.; Kim, S.-H.; Kang, S.-W.; Choi, I.-W.; Heo, J.-S.; DeLaune, R.D.; Seo, D.-C. Competitive adsorption and selectivity sequence of heavy metals by chicken bone-derived biochar: Batch and column experiment. J. Environ. Sci. Health Part A 2015, 50, 1194–1204. [Google Scholar] [CrossRef]
  126. Mahdi, Z.; Qiming, J.Y.; El Hanandeh, A. Removal of lead (II) from aqueous solution using date seed-derived biochar: Batch and column studies. Appl. Water Sci. 2018, 8, 1–13. [Google Scholar] [CrossRef] [Green Version]
  127. Mahdi, Z.; Yu, Q.J.; El Hanandeh, A. Investigation of the kinetics and mechanisms of nickel and copper ions adsorption from aqueous solutions by date seed derived biochar. J. Environ. Chem. Eng. 2018, 6, 1171–1181. [Google Scholar] [CrossRef] [Green Version]
  128. Boni, M.R.; Chiavola, A.; Marzeddu, S. Application of biochar to the remediation of Pb-contaminated solutions. Sustainability 2018, 10, 4440. [Google Scholar] [CrossRef] [Green Version]
  129. Hong, N.; Cheng, Q.; Goonetilleke, A.; Bandala, E.R.; Liu, A. Assessing the effect of surface hydrophobicity/hydrophilicity on pollutant leaching potential of biochar in water treatment. J. Ind. Eng. Chem. 2020, 89, 222–232. [Google Scholar] [CrossRef]
  130. Yuan, H.; Lu, T.; Wang, Y.; Chen, Y.; Lei, T. Sewage sludge biochar: Nutrient composition and its effect on the leaching of soil nutrients. Geoderma 2016, 267, 17–23. [Google Scholar] [CrossRef]
  131. Chen, X.; Yang, L.; Myneni, S.C.; Deng, Y. Leaching of polycyclic aromatic hydrocarbons (PAHs) from sewage sludge-derived biochar. Chem. Eng. J. 2019, 373, 840–845. [Google Scholar] [CrossRef]
  132. ChemAxon. Chemicalize. 2021. Available online: https://www.chemicalize.org (accessed on 15 July 2021).
  133. Inyang, M.; Gao, B.; Zimmerman, A.; Zhang, M.; Chen, H. Synthesis, characterization, and dye sorption ability of carbon nanotube–biochar nanocomposites. Chem. Eng. J. 2014, 236, 39–46. [Google Scholar] [CrossRef]
  134. Mahmoud, D.K.; Salleh, M.A.M.; Karim, W.A.W.A.; Idris, A.; Abidin, Z.Z. Batch adsorption of basic dye using acid treated kenaf fibre char: Equilibrium, kinetic and thermodynamic studies. Chem. Eng. J. 2012, 181, 449–457. [Google Scholar] [CrossRef]
  135. Park, J.-H.; Wang, J.J.; Meng, Y.; Wei, Z.; DeLaune, R.D.; Seo, D.-C. Adsorption/desorption behavior of cationic and anionic dyes by biochars prepared at normal and high pyrolysis temperatures. Colloids Surf. A Physicochem. Eng. Asp. 2019, 572, 274–282. [Google Scholar] [CrossRef]
  136. Jian, X.; Zhuang, X.; Li, B.; Xu, X.; Wei, Z.; Song, Y.; Jiang, E. Comparison of characterization and adsorption of biochars produced from hydrothermal carbonization and pyrolysis. Environ. Technol. Innov. 2018, 10, 27–35. [Google Scholar] [CrossRef]
  137. Parra-Marfíl, A.; Ocampo-Pérez, R.; Collins-Martínez, V.H.; Flores-Vélez, L.M.; Gonzalez-Garcia, R.; Medellín-Castillo, N.A.; Labrada-Delgado, G.J. Synthesis and characterization of hydrochar from industrial Capsicum annuum seeds and its application for the adsorptive removal of methylene blue from water. Environ. Res. 2020, 184, 109334. [Google Scholar] [CrossRef]
  138. Nguyen, D.H.; Tran, H.N.; Chao, H.-P.; Lin, C.-C. Effect of nitric acid oxidation on the surface of hydrochars to sorb methylene blue: An adsorption mechanism comparison. Adsorpt. Sci. Technol. 2019, 37, 607–622. [Google Scholar] [CrossRef] [Green Version]
  139. Xu, R.-K.; Xiao, S.-C.; Yuan, J.-H.; Zhao, A.-Z. Adsorption of methyl violet from aqueous solutions by the biochars derived from crop residues. Bioresour. Technol. 2011, 102, 10293–10298. [Google Scholar] [CrossRef]
  140. Hameed, B.; Rahman, A. Removal of phenol from aqueous solutions by adsorption onto activated carbon prepared from biomass material. J. Hazard. Mater. 2008, 160, 576–581. [Google Scholar] [CrossRef]
  141. Liu, C.-H.; Chuang, Y.-H.; Li, H.; Teppen, B.J.; Boyd, S.A.; Gonzalez, J.M.; Johnston, C.T.; Lehmann, J.; Zhang, W. Sorption of lincomycin by manure-derived biochars from water. J. Environ. Qual. 2016, 45, 519. [Google Scholar] [CrossRef] [Green Version]
  142. Inyang, M.; Gao, B.; Zimmerman, A.; Zhou, Y.; Cao, X. Sorption and cosorption of lead and sulfapyridine on carbon nanotube-modified biochars. Environ. Sci. Pollut. Res. 2015, 22, 1868–1876. [Google Scholar] [CrossRef] [PubMed]
  143. Liu, P.; Liu, W.-J.; Jiang, H.; Chen, J.-J.; Li, W.-W.; Yu, H.-Q. Modification of bio-char derived from fast pyrolysis of biomass and its application in removal of tetracycline from aqueous solution. Bioresour. Technol. 2012, 121, 235–240. [Google Scholar] [CrossRef] [PubMed]
  144. Xie, M.; Chen, W.; Xu, Z.; Zheng, S.; Zhu, D. Adsorption of sulfonamides to demineralized pine wood biochars prepared under different thermochemical conditions. Environ. Pollut. 2014, 186, 187–194. [Google Scholar] [CrossRef]
  145. Ni, J.; Pignatello, J.J.; Xing, B. Adsorption of aromatic carboxylate ions to black carbon (biochar) is accompanied by proton exchange with water. Environ. Sci. Technol. 2011, 45, 9240–9248. [Google Scholar] [CrossRef]
  146. Qiu, Y.; Zheng, Z.; Zhou, Z.; Sheng, G.D. Effectiveness and mechanisms of dye adsorption on a straw-based biochar. Bioresour. Technol. 2009, 100, 5348–5351. [Google Scholar] [CrossRef]
  147. Lin, Y.-C.; Ho, S.-H.; Zhou, Y.; Ren, N.-Q. Highly efficient adsorption of dyes by biochar derived from pigments-extracted macroalgae pyrolyzed at different temperature. Bioresour. Technol. 2018, 259, 104–110. [Google Scholar]
  148. Li, Y.; Tsend, N.; Li, T.; Liu, H.; Yang, R.; Gai, X.; Wang, H.; Shan, S. Microwave assisted hydrothermal preparation of rice straw hydrochars for adsorption of organics and heavy metals. Bioresour. Technol. 2019, 273, 136–143. [Google Scholar] [CrossRef]
  149. Li, Y.; Meas, A.; Shan, S.; Yang, R.; Gai, X. Production and optimization of bamboo hydrochars for adsorption of Congo red and 2-naphthol. Bioresour. Technol. 2016, 207, 379–386. [Google Scholar] [CrossRef]
  150. Kearns, J.P.; Wellborn, L.S.; Summers, R.; Knappe, D. 2,4-D adsorption to biochars: Effect of preparation conditions on equilibrium adsorption capacity and comparison with commercial activated carbon literature data. Water Res. 2014, 62, 20–28. [Google Scholar] [CrossRef]
  151. Peng, B.; Chen, L.; Que, C.; Yang, K.; Deng, F.; Deng, X.; Shi, G.; Xu, G.; Wu, M. Adsorption of antibiotics on graphene and biochar in aqueous solutions induced by π-π interactions. Sci. Rep. 2016, 6, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Jang, H.M.; Kan, E. A novel hay-derived biochar for removal of tetracyclines in water. Bioresour. Technol. 2019, 274, 162–172. [Google Scholar] [CrossRef] [PubMed]
  153. Dai, Y.; Li, J.; Shan, D. Adsorption of tetracycline in aqueous solution by biochar derived from waste Auricularia auricula dregs. Chemosphere 2020, 238, 124432. [Google Scholar] [CrossRef] [PubMed]
  154. Shang, J.; Kong, X.; He, L.; Li, W.; Liao, Q. Low-cost biochar derived from herbal residue: Characterization and application for ciprofloxacin adsorption. Int. J. Environ. Sci. Technol. 2016, 13, 2449–2458. [Google Scholar] [CrossRef] [Green Version]
  155. Zhang, L.; Tong, L.; Zhu, P.; Huang, P.; Tan, Z.; Qin, F.; Shi, W.; Wang, M.; Nie, H.; Yan, G. Adsorption of chlortetracycline onto biochar derived from corn cob and sugarcane bagasse. Water Sci. Technol. 2018, 78, 1336–1347. [Google Scholar] [CrossRef] [PubMed]
  156. Chen, B.; Chen, Z.; Lv, S. A novel magnetic biochar efficiently sorbs organic pollutants and phosphate. Bioresour. Technol. 2011, 102, 716–723. [Google Scholar] [CrossRef]
  157. Sun, K.; Ro, K.; Guo, M.; Novak, J.; Mashayekhi, H.; Xing, B. Sorption of bisphenol A, 17α-ethinyl estradiol and phenanthrene on thermally and hydrothermally produced biochars. Bioresour. Technol. 2011, 102, 5757–5763. [Google Scholar] [CrossRef]
  158. Cao, X.; Ma, L.; Gao, B.; Harris, W. Dairy-manure derived biochar effectively sorbs lead and atrazine. Environ. Sci. Technol. 2009, 43, 3285–3291. [Google Scholar] [CrossRef]
  159. Chen, B.; Chen, Z. Sorption of naphthalene and 1-naphthol by biochars of orange peels with different pyrolytic temperatures. Chemosphere 2009, 76, 127–133. [Google Scholar] [CrossRef]
  160. Kasozi, G.N.; Zimmerman, A.R.; Nkedi-Kizza, P.; Gao, B. Catechol and humic acid sorption onto a range of laboratory-produced black carbons (biochars). Environ. Sci. Technol. 2010, 44, 6189–6195. [Google Scholar] [CrossRef]
  161. Zhang, P.; Sun, H.; Yu, L.; Sun, T. Adsorption and catalytic hydrolysis of carbaryl and atrazine on pig manure-derived biochars: Impact of structural properties of biochars. J. Hazard. Mater. 2013, 244, 217–224. [Google Scholar] [CrossRef]
  162. Karakoyun, N.; Kubilay, S.; Aktas, N.; Turhan, O.; Kasimoglu, M.; Yilmaz, S.; Sahiner, N. Hydrogel–Biochar composites for effective organic contaminant removal from aqueous media. Desalination 2011, 280, 319–325. [Google Scholar] [CrossRef]
  163. Reddy, K.R.; Xie, T.; Dastgheibi, S. Evaluation of biochar as a potential filter media for the removal of mixed contaminants from urban storm water runoff. J. Environ. Eng. 2014, 140, 04014043. [Google Scholar] [CrossRef] [Green Version]
  164. Inyang, M.; Dickenson, E. The potential role of biochar in the removal of organic and microbial contaminants from potable and reuse water: A review. Chemosphere 2015, 134, 232–240. [Google Scholar] [CrossRef] [PubMed]
  165. Upton, G.; Cook, I. Understanding Statistics; Oxford University Press: Oxford, UK, 1996. [Google Scholar]
  166. Park, J.; Hung, I.; Gan, Z.; Rojas, O.J.; Lim, K.H.; Park, S. Activated carbon from biochar: Influence of its physicochemical properties on the sorption characteristics of phenanthrene. Bioresour. Technol. 2013, 149, 383–389. [Google Scholar] [CrossRef] [PubMed]
  167. Ania, C.; Cabal, B.; Parra, J.; Pis, J. Importance of the hydrophobic character of activated carbons on the removal of naphthalene from the aqueous phase. Adsorpt. Sci. Technol. 2007, 25, 155–167. [Google Scholar] [CrossRef] [Green Version]
  168. Ding, Z.; Zhang, F.; Sun, N.; Chi, J. Sorption of Polycyclic Aromatic Hydrocarbons by Biochars of Wheat Straw with Different Pyrolysis Temperatures; IOP Conference Series: Earth and Environmental Science; IOP Publishing: Bristol, UK, 2019; p. 052008. [Google Scholar]
  169. Ravichandran, P.; Sugumaran, P.; Seshadri, S.; Basta, A.H. Optimizing the route for production of activated carbon from Casuarina equisetifolia fruit waste. R. Soc. Open Sci. 2018, 5, 171578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Gai, X.; Wang, H.; Liu, J.; Zhai, L.; Liu, S.; Ren, T.; Liu, H. Effects of feedstock and pyrolysis temperature on biochar adsorption of ammonium and nitrate. PLoS ONE 2014, 9, e113888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Grung, B.; Manne, R. Missing values in principal component analysis. Chemom. Intell. Lab. Syst. 1998, 42, 125–139. [Google Scholar] [CrossRef]
  172. Varies Authors. Global and China Activated Carbon Industry Report, 2018–2023. 2018. Available online: https://www.researchandmarkets.com/reports/4478751/global-and-chinese-spherical-activated-carbon#rela0-4605170 (accessed on 13 April 2021).
  173. Gopinath, A.; Divyapriya, G.; Srivastava, V.; Laiju, A.; Nidheesh, P.; Kumar, M.S. Conversion of sewage sludge into biochar: A potential resource in water and wastewater treatment. Environ. Res. 2021, 194, 110656. [Google Scholar] [CrossRef]
  174. Karim, A.A.; Kumar, M.; Singh, E.; Kumar, A.; Kumar, S.; Ray, A.; Dhal, N.K. Enrichment of primary macronutrients in biochar for sustainable agriculture: A review. Crit. Rev. Environ. Sci. Technol. 2020, 1–42. [Google Scholar]
Figure 1. Boxplot of adsorption capacities at equilibrium for BCs (a) and ACs (b). n = 50 for both series. Inter quartile range (IQR), median, means, and ranges were calculated through OriginLab software. Ranges were calculated as 1.5-fold the IQR, according to Apton and Cook [165].
Figure 1. Boxplot of adsorption capacities at equilibrium for BCs (a) and ACs (b). n = 50 for both series. Inter quartile range (IQR), median, means, and ranges were calculated through OriginLab software. Ranges were calculated as 1.5-fold the IQR, according to Apton and Cook [165].
Molecules 26 05063 g001
Figure 2. Distribution plot of equilibrium adsorption capacities (qe, mg/kg) as a function of the logKow of organic target pollutants removed by biochars (A) and activated carbons (B). n = 50 for both adsorbents. The clusters considered the log Kow range from 0.5 to 3.5.
Figure 2. Distribution plot of equilibrium adsorption capacities (qe, mg/kg) as a function of the logKow of organic target pollutants removed by biochars (A) and activated carbons (B). n = 50 for both adsorbents. The clusters considered the log Kow range from 0.5 to 3.5.
Molecules 26 05063 g002
Table 1. Standard methods for the characterization of activated carbon.
Table 1. Standard methods for the characterization of activated carbon.
TestMethodReference
Physical Tests
Bulk densityASTM D2854, CEFIC[33,34]
Absolute densityCEFIC[33]
Particle densityCEFIC[33]
Particle sizeASTM D2862, CEFIC[33,35]
Pressure dropCEFIC[33]
Mechanical strengthASTM D3802, AWWA B604, CEFIC[33,36,37]
Adsorption Tests and Indices
Adsorption isothermASTM: D3860-98, 5919-96, CEFIC [33,38,39]
Iodine numberAWWA B600-16, EN 12915-1, ASTM D4607-14, CEFIC[33,40,41,42]
Phenol numberCEFIC[33]
Methylene Blue numberCEFIC[33]
Molasses numberEPA625171002A[43]
Tannin numberAWWA B600[41]
Chemical Tests
Ashes, water soluble material, and water-extractable substances (As, Cd, Cr, Hg, Ni, Pb, Sb, Se, CN-, fluoranthene, benzo[b]fluoranthene, benzo[k]fluoranthene, benzo[a]pyrene, benzo[g,h,i]perylene, indeno[1,2,3-cd]pyrene)EN 12915-1[40]
pHASTM: D6851-02, D3838-05[44,45]
Table 3. List of the most investigated organic pollutants removed by biochars in the last ten years’ worth of literature. Compounds are grouped following their ionizable properties. The values of their logKow and pKa (if applicable), as calculated using Chemicalize Software [132], are reported.
Table 3. List of the most investigated organic pollutants removed by biochars in the last ten years’ worth of literature. Compounds are grouped following their ionizable properties. The values of their logKow and pKa (if applicable), as calculated using Chemicalize Software [132], are reported.
CATIONClasslogKowpKa (Ionic Group)Comparison with ACRef
Methylene blueDye0.75/No[133,134,135,136,137,138]
Methyl violetDye0.439.17No[139]
Malachite greenDye0.8/No[140]
LincomycinAntibiotic−0.37.97No[141]
ANIONClasslogKowpKa (Ionic Group)Comparison with ACRef
SulphapyridineAntibiotic0.356.24Yes[142,143]
SulfamethoxazoleAntibiotic0.796.16No[144]
p-coumaric acidDrug1.463.81No[145]
Reactive brilliant blueDye−1.33−2.69No[146]
Congo RedDye2.63/No[135,147,148,149]
tris(2-carboxyethyl)phosphineFlame retardant1.783.22; 4.38No[121]
2,4-dichlrophenenoxiacetic acidHerbicide/Pesticide2.612.81Yes[150]
t-Cinnamic acidPrecursor2.134.32No[145]
NEUTRALClasslogKowpKa (Ionic Group)Comparison with ACRef
AmoxicillinAntibiotic−2.37.22No[151]
TetracyclineAntibiotic−3.47.36No[22,152,153]
CiprofloxacinAntibiotic−0.85.56; 8.77No[154]
SulfadiazineAntibiotic0.387No[22]
ChlortetracyclineAntibiotic−1.982.99No[155]
1H-benzotriazoleCorrosion Inhibitor1.449.04No[121]
p-nitrotolueneDye2.37/No[156]
Bisphenol-aEndocrine disruptors3.329.78; 10.39No[157]
AtrazineHerbicide 2.61/No/Yes[121,158]
DiuronHerbicide2.6813.18No[121]
1-naphtolHerbicide2.859.6Yes[159]
CatecholHerbicide0.99.34; 12.79No[160]
CarbarylHerbicide0.9/No[161]
17α-ethinyl estradiolEstrogen3.6710.33No[157]
PhenolPlastic production1.4610.02No[162]
PhenanthrenePolycyclic Aromatic Hydrocarbon3.71/No[157,163]
NaphthalenePolycyclic Aromatic Hydrocarbon2.96/No/Yes/No[156,159,163]
TrichloroethyleneSolvent2.42/Yes[6]
4-tert-OctylphenolNon-ionic Surfactant5.1810.23Yes[56]
TRITONTMX-45 (mixture of 4-t-octylphenol polyethoxylated)Non-ionic Surfactant3.49–4.9015.10 aYes[56]
4-(1-Ethyl-1,4-dimethylpentyl)-phenolSurfactant5.7910.22Yes[56]
IGEPAL®CO-520 (mixture of branched 4-nonylphenol polyethoxylated oligomers)Surfactant4.26–5.50/Yes[56]
a Referred to the compound 2-[4-(2,4,4-trimethylpentan-2-yl)phenoxy]ethan-1-ol, as retrieved from Chemicalize.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Castiglioni, M.; Rivoira, L.; Ingrando, I.; Del Bubba, M.; Bruzzoniti, M.C. Characterization Techniques as Supporting Tools for the Interpretation of Biochar Adsorption Efficiency in Water Treatment: A Critical Review. Molecules 2021, 26, 5063. https://doi.org/10.3390/molecules26165063

AMA Style

Castiglioni M, Rivoira L, Ingrando I, Del Bubba M, Bruzzoniti MC. Characterization Techniques as Supporting Tools for the Interpretation of Biochar Adsorption Efficiency in Water Treatment: A Critical Review. Molecules. 2021; 26(16):5063. https://doi.org/10.3390/molecules26165063

Chicago/Turabian Style

Castiglioni, Michele, Luca Rivoira, Irene Ingrando, Massimo Del Bubba, and Maria Concetta Bruzzoniti. 2021. "Characterization Techniques as Supporting Tools for the Interpretation of Biochar Adsorption Efficiency in Water Treatment: A Critical Review" Molecules 26, no. 16: 5063. https://doi.org/10.3390/molecules26165063

Article Metrics

Back to TopTop