Journal Pre-proof Recovery of grape waste for the preparation of adsorbents for water treatment: Mercury removal ˜ N.M. Zu´ niga-Muro (Writing - original draft) (Conceptualization) (Methodology) (Investigation), A. Bonilla-Petriciolet (Supervision) (Project administration) (Writing - review and editing), D.I. Mendoza-Castillo (Investigation) (Writing - review and editing), H.E. ´ Reynel-Avila (Investigation) (Writing - review and editing), C.J. Duran-Valle (Investigation) (Writing - review and editing), H. Ghalla (Software) (Investigation) (Writing - review and editing), L. Sellaoui
Software, Investigation) (Writing review and editing)
PII:
S2213-3437(20)30086-5
DOI:
https://doi.org/10.1016/j.jece.2020.103738
Reference:
JECE 103738
To appear in:
Journal of Environmental Chemical Engineering
Received Date:
14 November 2019
Revised Date:
30 January 2020
Accepted Date:
1 February 2020
˜ Please cite this article as: Zu´ niga-Muro NM, Bonilla-Petriciolet A, Mendoza-Castillo DI, ´ Reynel-Avila HE, Duran-Valle CJ, Ghalla H, Sellaoui L, Recovery of grape waste for the preparation of adsorbents for water treatment: Mercury removal, Journal of Environmental Chemical Engineering (2020), doi: https://doi.org/10.1016/j.jece.2020.103738
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2019 Published by Elsevier.
RECOVERY OF GRAPE WASTE FOR THE PREPARATION OF ADSORBENTS FOR WATER TREATMENT: MERCURY REMOVAL 1
N.M. Zúñiga-Muro , A. Bonilla-Petriciolet 1*, D.I. Mendoza-Castillo 1,2, H.E. Reynel-Ávila 1,2, C.J. Duran-Valle 3, H. Ghalla 4, L. Sellaoui 4 1
Instituto Tecnológico de Aguascalientes, Aguascalientes, México, 20256 2
CONACYT, Cátedras Jóvenes Investigadores, 03940, México 3
Universidad de Extremadura, Badajoz, 06006, España Monastir University, Monastir, 5000, Tunisia
Jo
ur
na
lP
re
-p
ro of
4
1
Highlights
Recovery of grape bagasse biomass to prepare adsorbents was proposed New adsorbents for mercury removal were synthetized from grape bagasse Pyrolysis of grape bagasse was analyzed to improve mercury adsorption Mercury – grape bagasse char interactions were analyzed via DFT Oxygenated and silicon-based functional groups were involved in mercury adsorption Two-site Langmuir model correlated the mercury adsorption isotherms
ABSTRACT. This study has focused on the recovery of grape bagasse biomass to prepare new adsorbents for
ro of
the removal of mercury from aqueous solutions.
Keywords: Grape bagasse waste, Pyrolysis, Char, Water treatment, Adsorption, Mercury.
Corresponding author: 524499105002, [email protected]
Jo
ur
na
lP
re
-p
*
2
1. INTRODUCTION. Industrial activities generate significant amounts of waste and by-products that generally cause serious environmental problems. Recovery operations use these discarded materials as a feedstock to create valuable products, which is an alternative approach to reduce the waste generation and the extraction of natural resources [1]. In this context, the grapes are one of the most harvested fruits with an estimated world production of 75 million tons/year [2]. Statistical data indicate that more than 50 % of the grapes is used in the wine and juice industries and that approximately 7 % of the initial processed mass is grape bagasse (GB) [2-4]. GB consists of seeds and skins, its main chemical components are carbon (46.6 %), oxygen (45.5 %), hydrogen (6.3 %) and nitrogen (1.7 %) and is generally used as animal food, incinerated or just discarded [3,5-8]. Based on these
ro of
facts, previous studies have concluded that this biomass could be used as raw material to obtain a variety of value-added products, including new materials for environmental applications (e.g.,
for preparing carbon-based materials) [3,5,7,8,13,14]. Therefore, the preparation of GB-based adsorbents is an attractive alternative to develop low cost and ecofriendly water treatment processes. It is important to remark that wastewater generation and the need for water treatment
-p
have promoted the search for alternative technologies and the application of novel materials with outstanding properties [9-12].
re
In the industrial production of adsorbents (e.g., activated carbons, chars) for water treatment, a significant amount of the used precursors is related to lignocellulosic materials [15-17]. These adsorbents are obtained by pyrolysis and their physicochemical properties depend on the
lP
characteristics of the precursor and the preparation conditions [18,19]. Specifically, the pyrolysis of GB produces a gas phase formed mainly of CO, CO2, H2 and CH4; a liquid containing methanol, acetone, furfuryl alcohol and phenol; and a solid phase that can be classified as a char [3]. This
na
product distribution (i.e., the amount of solid, liquid and gas) depends on the experimental conditions applied in the pyrolysis [3,20]. The char generated in the GB pyrolysis is a porous carbon material, which can been employed as an adsorbent for heavy metals, dyes and other
ur
chemical compounds present in groundwater and wastewaters [8,21]. The use of GB-based chars as adsorbents in water treatment requires a tailored surface chemistry, which is the most important
Jo
property of the adsorbent compared to the textural parameters (e.g., surface area and pore distribution) [13,18,22]. The surface chemical properties of GB-based adsorbents can be modified via chemical and physical treatment methods to obtain a specific performance for the removal of the target pollutants from water. The possible modification method used to improve the adsorbent properties should be selected considering both technical and economic factors. According to the literature, the biomass of GB has been used in its natural form for the removal of heavy metals such as Cd(II) and Pb(II) in aqueous solutions showing adsorption capacities of 53.8 and 42.3 mg/g, respectively [5,7]. GB biomass has been also utilized to produce activated carbons for the adsorption of Cu(II) and dyes (basic blue 9 and acid yellow 36) with adsorption 3
capacities from 43 to 417 mg/g [8,14]. These results have shown that the GB-based adsorbents can be effectively used in water purification. But, to the best of author’s knowledge, this biomass has not been utilized to remove mercury from aqueous solutions. Mercury can be found in water resources due to anthropogenic and geogenic factors, such as the geological composition of aquifers or improper management of polluted industrial effluents [23]. This heavy metal is considered a hazardous chemical for human beings and ecosystems due to its toxicological profile and it has also been categorized as a priority pollutant in the context of water treatment [24-27]. Mercury adsorption is a challenge and adsorbents with specific surface functionalities are required to obtain a competitive removal performance [19,28]. The aim of this study was to analyze the preparation of GB-based chars via pyrolysis and its
ro of
application in the mercury removal from aqueous solutions as an alternative for the recovery of this residual biomass. A full experimental factorial design was defined to analyze the effects of the pyrolysis conditions of the GB biomass on the physicochemical properties and the mercury
adsorption capacity of the corresponding chars. The best adsorbent was selected and the mercury
adsorption thermodynamics was studied. Density functional theory (DFT) calculations were used
-p
to characterize the adsorbent-adsorbate interatomic interactions and to understand the mercury
re
adsorption mechanism.
2. METHODOLOGY.
2.1 Recovery of the precursor and its usage in the preparation of chars for mercury adsorption
lP
GB was collected from a local wine factory, washed with deionized water, dried, crushed and sieved to obtain a particle size of 0.84 - 1.0 mm. This precursor was physicochemically characterized at room temperature using FTIR spectroscopy and XRD analysis.
na
The GB biomass was pyrolyzed to obtain the adsorbents in a Carbolite Eurotherm CTF 12165/550 tubular furnace using a quartz tube with loads of 20 g. A full factorial experimental design was established for the analysis of the impact of the pyrolysis conditions (i.e., pyrolysis
ur
temperature and dwell time) on the mercury adsorption properties of the GB chars. In this experimental design, the adsorbents were prepared with pyrolysis temperatures of 600, 800 and
Jo
1000 °C and dwell times of 1, 2 and 3 h. Table 1 shows the corresponding experimental design utilized in the preparation of the chars. A heating rate of 10 °C/min and a N2 flow of 200 mL/min were used during the pyrolysis. The adsorbent samples obtained from the experimental design were weighted and washed with deionized water under constant stirring at 120 rpm until to obtain a constant pH in the washing solution. Then, the adsorbents were dried at 110 °C for 24 h and the particles with size of 0.50 - 0.84 mm were selected and used in the mercury adsorption tests.
2.2 Mercury adsorption studies
4
All mercury adsorption studies were carried out by triplicate using batch adsorbers under constant stirring of 120 rpm. In particular, 0.02 g of the adsorbents and 10 mL of the mercury solutions (which were prepared from reactive grade HgCl2) were used in all experiments. The mercury adsorption performance of the adsorbents obtained with the experimental design was analyzed at 30 °C, pH 1.5 and 24 h, where the pH of the mercury solutions was adjusted using HNO3. Mercury concentration was quantified using an atomic absorption spectrophotometer (Ice 3000 Thermo Scientific) with a linear calibration curve (i.e., 10 - 300 mg/L). The response variable of the experimental design was the mercury adsorption capacity (q, mg/g), which was calculated using the adsorbate mass balance for the batch adsorber 𝑞 = 𝑉(
𝐶0 −𝐶𝑓 𝑚
)
(1)
ro of
where V is the volume of mercury solution (L), C0 and Cf are the initial and final mercury
concentrations (mg/L) in the aqueous solutions utilized in the adsorption experiments and m is the adsorbent mass (g).
The results of the adsorption capacity of the experimental design (Table 1) were used to
identify the best preparation conditions for GB-based char. The best adsorbent was utilized in
-p
additional experiments to determine kinetic and equilibrium thermodynamic parameters of the mercury removal. Adsorption kinetic studies were performed at pH 4 and 30 °C using mercury
re
solutions with initial concentrations of 100 and 200 mg/L. Mercury adsorption kinetic profiles were quantified at operating times from 5 to 1440 min. The same adsorbent dosage was also
lP
utilized in both kinetic and equilibrium adsorption studies. Pseudo second order model was applied for kinetic data modeling and for the calculation of the adsorption rate constants via a nonlinear data regression. Mercury adsorption isotherms were measured at different conditions of pH (1.5, 3 and 4) and temperature (30 and 40 ºC) with adsorbate concentrations from 10 to 900
na
mg/L and an equilibrium time of 24 h. These experimental isotherms were fitted by a Langmuirtype adsorption model. Enthalpy change for the mercury removal using the best GB-based char
ur
was calculated using the distribution coefficient approach and Van’t Hoff’s plot [29].
2.3 Physicochemical characterization of the GB-based chars
Jo
Physicochemical characteristics of GB-based adsorbents were studied via N2 adsorption-
desorption isotherms, elemental analysis, X-ray fluorescence (XRF), Boehm titration, Fourier Transform Infrared (FTIR) and X-ray diffraction (XRD) techniques. N2 adsorption-desorption isotherm studies were carried out at 77 K with an equipment Quadrasorb Evo de Quantachrome Instruments, where the samples were previously degassed for 24 h at 120 ºC. Elemental analysis (C, H, O and N) was performed with an equipment LECO CHNS 932, while a Bruker S8 Tiger analyzer was utilized in XRF.
5
Boehm-based method titrations were performed with 0.25 g of the adsorbents, which were kept in contact with 25 mL of solutions of NaOH (0.1 M), Na2CO3 (0.05 M) and NaHCO3 (0.1 M) to estimate the amount of acidic groups in the adsorbent surface. These tests were carried out at 30 °C for 48 h. Subsequently, these aqueous solutions were titrated with HCl (0.1 M) until reaching the neutralization that was verified with a pH-meter. The amount of acidic sites was estimated under the assumptions that NaOH neutralized carboxylic, phenolic and lactonic groups, Na2CO3 was utilized to neutralize the carboxylic and lactonic functionalities, while NaHCO3 neutralized only the carboxylic functional groups [30]. Surface chemistry analysis was carried out with FTIR spectra of the GB-based chars. These spectra were obtained with a Nicolet iS10 Thermo Scientific spectrometer equipped with a DTGS KBr detector using KBr pellets where
ro of
FTIR analysis was done in the range of 4000 - 600 cm-1. XRD patterns of selected samples of adsorbents were recorded in a PANalytical Empyrean X-ray diffractometer equipped with a
PIXcel1D detector and the Bragg-Brentano geometry, on a range of 5° ≤ 2θ ≤ 100° with CuKα
nickel filtered radiation (λ = 1.5406 Å, 45 kV, 30 mA). HighScore Plus software and the PDF 2
-p
database were employed for the analysis and interpretation of diffraction patterns.
2.4 Theoretical analysis of the mercury adsorption mechanism via DFT calculations
re
DFT calculations were performed to understand the mercury adsorption mechanism using GB chars. These calculations were done at the ground state using Gaussian 09 program and GaussView as a visualization package [31,32]. A simplified molecular structure for the GB-based
lP
adsorbent surface (M) was proposed employing the results of physicochemical characterization, see Figure 1. The molecular structures of M and the complex M Hg2+ were optimized in the context of DFT method using the B3LYP exchange correlation hybrid functional. For all atoms,
na
except Hg, the polarized 6-311G(d,p) basis set was utilized. ECP LanL2DZ basis set was employed for the heavy atom Hg. Harmonic vibrational modes were computed at the same levels of theory and the absence of any imaginary frequency confirmed that the optimized structures
ur
were stable.
Jo
3. RESULTS AND DISCUSSION. 3.1 Analysis of preparation conditions of grape bagasse-based adsorbents for mercury
removal and their physicochemical properties The yields for the preparation of the GB-based chars via pyrolysis are reported in Table 1.
These yields ranged from 33.7 to 38.2 % where the highest value corresponded to the adsorbent prepared at 600 °C for 1 h. Demiral and Ayan [3] obtained similar results in the GB pyrolysis at the same temperature, reporting a char yield of 36.8 %. In general, the adsorbent yield decreased slightly with pyrolysis conditions (i.e., temperature and dwell time), where the lowest char yield (33.7 %) was obtained for the biomass pyrolyzed at 1000 °C and 3 h. Note that the GB pyrolysis 6
showed char yields higher than those reported for other adsorbents produced from lignocellulosic biomasses, for example, corncob, almond shell and olive stone [33,34]. This result could be attributed to the lignin content of GB (~ 42 %), which has been recognized as the most stable compound of these lignocellulosic materials [8,33,34]. On the other hand, the results of mercury adsorption for the chars prepared via the full factorial experimental design are also reported in Table 1. The maximum mercury adsorption capacity was 5.3 mg/g for the adsorbent obtained at 600 °C and 2 h, while the char prepared at the pyrolysis conditions of 1000 °C and 3 h showed the lowest adsorption capacity of 0.5 mg/g. Mercury adsorption capacity decreased with the pyrolysis conditions utilized to obtain the chars and this trend agreed with other studies concerning to the preparation of chars from lignocellulosic
ro of
biomasses [35]. This result can be associated to the thermal degradation of surface functional groups that could be involved in the mercury adsorption [35,36]. Note that it was expected that increments on the pyrolysis temperature and dwell time decreased the number of surface functionalities of GB-based chars.
N2 adsorption-desorption isotherms of selected GB chars are reported in Figure 2, while their
-p
textural parameters are shown in Table 2. According to the IUPAC classification, these adsorbents exhibited a type III isotherm, which corresponds to materials with low porosity that can be
re
characterized by a weak adsorbate-adsorbent interaction [37]. Specific surface area of char samples calculated with Brunauer-Emmett-Teller (SBET) model ranged from 5.7 m2/g (adsorbent obtained at 1000 °C for 2 h) to 21.9 m2/g (adsorbent obtained at 600 °C and 1 h). Some studies
lP
have reported SBET of 2.0 - 280.0 m2/g for different lignocellulosic-based materials and the GBbased chars can be considered as adsorbents with low surface area [13,18,38,39]. Density Functional Theory was also utilized to estimate the specific surface areas (SDFT), see Table 2. The
na
highest value of SDFT was 12.8 m2/g and corresponded to the adsorbent with the maximum mercury adsorption capacity (i.e., 5.3 mg/g). Overall, the specific surface area decreased with pyrolysis temperature. Similar results have been reported by other researchers in the preparation
ur
of adsorbents from lignocellulosic biomasses [40-42]. For example, Paethanom et al. [42] reported that the specific surface areas of rice husk chars varied from 141 m2/g (pyrolysis at 600
Jo
°C) to 117 and 46 m2/g (pyrolysis at 800 and 1000 °C). These results can be attributed to the adsorbent structural ordering and to the fact that a high pyrolysis temperature can develop the porous structure in the char, which can reach the point where the walls of the pores become so thin that they can collapse, thus reducing the available char surface area [40,42]. Total pore volume (VTotal) of char samples ranged from 0.010 to 0.028 cm3/g and they also exhibited a low microporosity (i.e., 0.001 - 0.005 cm3/g). It is convenient to remark that the GB char with the highest mercury adsorption capacity showed the best textural parameters. Specific surface area and total pore volume of this char decreased after the mercury adsorption (i.e., 9.49 m2/g and 0.0225 cm3/g, respectively). These results indicated that textural properties of the GB chars 7
affected their mercury adsorption performance, which was consistent with the results reported by other authors [42,43]. Table 3 shows the results of elemental composition and XRF analysis of the GB adsorbents. In general, the percentage of C of these chars increased from 63.6 to 74.9 wt% with both pyrolysis temperature and dwell time, while the content of H, N and O decreased. These results can be explained taking into account that the changes in the pyrolysis conditions caused the loss of surface functional groups, the release of volatile matter and the loss of H, N and O atoms attached to C due to the degradation of the structural core. At the same time, the increment of pyrolysis conditions (i.e., temperature and/or dwell time) resulted in a higher content of fixed carbon, thus implying that the chars became carbonaceous-based materials especially at the highest pyrolysis
ro of
temperature [42,44,45]. Results also showed that the inorganic phase of the adsorbents was composed of several elements (e.g., Al, Ca, Si, P, K, Fe), where Si was the most abundant. The highest quantity of Si (i.e., 3.58 wt%) was found in the adsorbent with the highest mercury
adsorption capacity, which was prepared at the pyrolysis conditions of 600 °C and 2 h. This result
suggested the participation of Si moieties in the mercury adsorption mechanism. The best
-p
adsorbent was further analyzed after the adsorption experiments and mercury was identified in its structure ( 1.11 wt%).
re
In terms of the char surface chemistry, Boehm titration results showed that the acidic groups in the char surface decreased with respect to the conditions used in the pyrolysis of GB, which ranged from 0.11 to 0.27 mmol/g, see Table 1. Herein, it is important to highlight that the acidic
lP
functional groups have been considered as the main functionalities involved in the heavy metal adsorption on chars obtained from lignocellulosic feedstocks [22]. Therefore, this trend on the loss of functional groups can be partially associated with the mercury adsorption properties of the
na
GB chars prepared via the experimental design. However, the best GB adsorbent did not show the highest amount of acidic functional groups. These findings indicated that the adsorption performance of this char was the result of its surface chemistry (in terms of acidic functional
ur
groups and Si moieties) and its textural parameters. Surface chemistry of the best adsorbent demonstrated the predominant presence of phenolic groups (0.15 mmol/g) in comparison to
Jo
carboxylic and lactonic functionalities (0.02 and 0.03 mmol/g, respectively). With illustrative purposes, Figure 3 displays the FTIR spectra of the GB precursor, the best
char for mercury adsorption and the adsorbent obtained at the maximum pyrolysis temperature and dwell time (i.e., 1000 °C and 3 h). Spectrum of the precursor showed absorption bands that corresponded to O-H of alcohols, phenols and carboxylic acids (3600 - 3000 cm-1), C-H from aliphatic and aromatic structures (2920 - 2853, 1420 cm-1), C=C of aromatic rings from oxygenated groups as lactones (1630 cm-1) and C=O from aldehydes and ketones (1127 cm-1) [5,8,14]. The lignin in the precursor structure was also confirmed by the strong band of C-O at 1051 cm-1, while the Si-OH absorption band was identified at 787 cm-1 [5,46]. FTIR spectra of 8
the GB biomass and the char samples were similar. The spectra of the chars showed the reduction of different absorption bands depending of the pyrolysis conditions used in their preparation, thus confirming the loss of surface functional groups. FTIR spectrum of the best adsorbent after mercury removal is also given in Figure 3. A loss of intensity in the absorption bands associated to O-H, C=C, C=O and Si-OH groups was observed indicating that these functionalities could be interacting with the metallic adsorbate. XRD patterns of the GB biomass, the best adsorbent and the char obtained at 1000 °C and 3 h are shown in Figure 4. These diffractograms indicated that the crystal structure of these samples contained SiO2. However, it was also confirmed that the content of SiO2 on the adsorbent surface was partially lost due to the pyrolysis conditions where these results agreed with XRF analyses
ro of
and other studies reported in literature [5]. Two broad peaks at 22.9 and 44.3 ° 2 were identified in the XRD patterns of char samples indicating the presence of a graphene-like structure, whose
intensity increased with the pyrolysis conditions due to they promoted a structural ordering in the adsorbent [47,48]. Figure 4 also displays the XRD pattern of the best adsorbent loaded with
mercury where it was possible to identify that the signals associated to SiO2 shifted drastically.
-p
This result was an evidence of the participation of Si moieties in the mercury adsorption process.
In the literature, some studies have reported the surface chemistry modification of adsorbents
re
using silicon in order to improve the mercury adsorption capacity, where the affinity of Si to this adsorbate has been pointed out [49]. In summary, the best GB char for mercury removal was obtained via pyrolysis at 600 °C and 2 h using N2 flow of 200 mL/min and a heating rate of 10
lP
°C/min. This adsorbent was selected for further adsorption studies including the theoretical understanding of the mercury removal mechanism via DFT calculations.
na
3.2 Experimental mercury adsorption performance of the best grape bagasse char. Kinetic and thermodynamic mercury adsorption studies were performed using the GB char with the highest adsorption capacity. Results of adsorption kinetics at pH 4 and 30 °C are reported
ur
in Figure 5. The adsorption equilibrium was reached with 20 h at tested experimental conditions and the pseudo second order model was used to estimate the adsorption kinetic parameters, see
Jo
Table 4. Mercury adsorption rate for GB char was 5.47E-03 g/mg min at 100 mg/L and 1.10E-03 g/mg min at 200 mg/L where the calculated equilibrium adsorption capacities were 6.86 and 13.73 mg/g, respectively; while the experimental adsorption capacities were 7.8 and 14.3 mg/g for the same initial concentrations. This kinetic model showed a R2 > 0.964, see Table 4. Mercury adsorption isotherms at different pH are given in Figure 6a-c. These results showed that the maximum adsorption capacity of the GB char increased with pH varying from 20.8 to 32.3 mg/g. This adsorption behavior can be attributed to the fact that the increment in solution pH caused the deprotonation of the different surface functional groups of GB char. Therefore, the amount of available adsorption sites for mercury binding increased under these experimental 9
conditions. Note that there was also a reduction of repulsion forces between the adsorbent surface and mercury ions and the competition of protons for the binding acidic sites also decreased [19,50]. Adsorption isotherms at pH 4, 30 and 40 °C (see Figures 6c and 6d) indicated that the mercury removal with this GB char was an endothermic process. Mercury removal increased from 32.3 to 45.9 mg/g at tested temperatures. It is convenient to remark that the temperature of the solution showed a greater impact on the mercury adsorption than the increment of pH. Estimated enthalpy for mercury adsorption on GB char was 24.48 kJ/mol, which could be associated to an adsorption process involving both physical and chemical interactions [51]. An endothermic nature of mercury adsorption has been reported by different authors [39,52,53]. Mercury adsorption capacities of the best GB char were higher than those reported for several adsorbents e.g.: multi-
ro of
walled carbon nanotubes (10.9 mg/g), Brazilian pepper biochars (24.2 mg/g) and bone charcoal (21.0 mg/g) [38,54,55].
Results of the adsorbent surface characterization showed that the GB char comprised oxygenated and silicon functionalities, which could be responsible of mercury binding.
Considering this adsorbent surface heterogeneity, the isotherm data were fitted to a two-site
-p
Langmuir model [56]. This isotherm model was developed based on the traditional Langmuir equation to handle the presence of different adsorption sites on the adsorbent surface [56]. The
re
two-site Langmuir equation has been successfully applied to represent the equilibrium adsorption of different metals (not mercury) on biomasses, bone char, activated carbon and goethite [56-59]. 𝑞′ =
𝑏1 𝑞1 𝐶𝑒 1+𝑏1 𝐶𝑒
+
𝑏2 𝑞2 𝐶𝑒 1+𝑏2 𝐶𝑒
lP
The equation of this two-adsorption site isotherm is given by
(2)
where q’ is the total adsorption capacity of the adsorbent in mg/g, q1 and q2 are the maximum
na
adsorption capacities for the two different adsorption sites (i.e., oxygenated and silicon functionalities for GB char) also given in mg/g, b1 and b2 are the adsorption equilibrium constants in L/mg and Ce is the mercury equilibrium concentration in the aqueous solution given in mg/L,
ur
respectively.
The two-site Langmuir model correlated satisfactorily the adsorption data with determination coefficients (R2) > 0.99 where the parameters of this model are given in Table 5. The equilibrium
Jo
constant b2 was higher than b1 being an indicator of the heterogeneous adsorbent surface with at least two different mercury-adsorption site affinities. Note that the mercury adsorption sites with high equilibrium constants presented high binding energies [58]. The value of the two-site Langmuir parameters increased with pH and temperature because these operating variables promoted the adsorption process as already discussed. Overall, the active sites 1 (i.e., oxygenated functional groups) presented a major contribution to the mercury removal in comparison to active sites 2 (i.e., Si moieties), see Figure 6.
10
3.3 DFT calculations and its relationship with the mercury adsorption mechanism Results from char characterization were utilized to analyze the role of surface functionalities on the mercury adsorption via DFT calculations. Figure 7 illustrates the two oxygenated functional groups (carboxylic acid and lactone functional groups) and Si-based functionality that were selected for this theoretical analysis with DFT calculations. Table 6 reports the results of DFT calculations for the proposed atomic interactions between GB char surface functionalities and mercury ions. Calculated binding energies ranged from -1239 to -1233 kJ/mol and atomic bond distances varied from 2.813 to 3.653 Å. DFT results indicated that the interaction between carboxylic acid functionalities and mercury ions was the strongest, followed by the interaction of Si moieties and lactone functional groups. These results confirmed that the oxygenated functional
ro of
groups could be the main active sites involved in the mercury adsorption from aqueous solution using the GB-based chars. DFT calculations were also consistent with the characterization results
of the surface chemistry of the chars, where the best adsorbent contained the highest amount of
Si moieties, which contributed to the mercury removal. These simulations also supported the assignation of active sites 1 (oxygenated functional groups) and 2 (silicon moieties) in the two-
-p
site Langmuir isotherm model. Consequently, the adsorption of mercury from aqueous solution
with GB-based chars was associated to the presence of both, oxygenated and silicon
re
functionalities on the adsorbent surface.
4. CONCLUSIONS
lP
This study has discussed and analyzed the preparation of alternative adsorbents for mercury removal via the pyrolysis of grape bagasse biomass. Grape bagasse chars were synthetized using different pyrolysis conditions, which were improved to maximize the mercury adsorption from
na
aqueous solution. Mercury adsorption capacities of grape bagasse adsorbents were determined by their heterogeneous surfaces that contained phenolic, carboxylic, lactonic and silicon functionalities. Mercury adsorption with the best grape bagasse char was endothermic with a
ur
maximum adsorption capacity up to 45.9 mg/g. DFT calculations confirmed that carboxylic functional groups and Si moieties were the main active sites for mercury adsorption using grape
Jo
bagasse chars. Results of this study contribute to the recovery of the grape bagasse biomass to obtain value-added materials for water pollution control and environmental protection.
CONFLICT OF INTEREST STATEMENT
The authors declare no conflict of interest.
11
Credit author statement N.M. Zúñiga-Muro: Writing – Original draft, Conceptualization, Methodology, Investigation A. Bonilla-Petriciolet: Supervision, Project administration, Writing – Review and Editing D.I. Mendoza-Castillo: Investigation, Writing – Review and Editing H.E. Reynel-Ávila: Investigation, Writing – Review and Editing C.J. Duran-Valle: Investigation, Writing – Review and Editing H. Ghalla: Software, Investigation, Writing – Review and Editing
ro of
L. Sellaoui: Software, Investigation, Writing – Review and Editing
REFERENCES
[1] E. Iacovidou, J. Millward-Hopkins, J. Busch, P. Purnell, C.A. Velis, J.N. Hahladakis, O. Zwirner, A. Brown, A pathway to circular economy: Developing a conceptual framework for
-p
complex value assessment of resources recovered from waste. Journal of Cleaner Production 168 (2017) 1279-1288.
re
[2] Food and Agriculture Organization of the United Nations (FAO). Production: Crops (2018) [http://www.fao.org/faostat/en/?#data/QC] [12/02/2019].
[3] I. Demiral, E.A. Ayan, Pyrolysis of grape bagasse: effect of pyrolysis conditions on the
lP
product yields and characterization of the liquid product. Bioresource Technology 102 (2011) 3946-3951.
[4] Organization of Vine and Wine (OIV). Table and Dried Grapes: World data available [http://www.oiv.int/en/oiv-life/table-and-dried-grapes-world-data-available]
[12/02/2019].
na
(2017)
[5] N.V. Farinella, G.D. Matos, M.A.Z. Arruda, Grape bagasse as a potential biosorbent of
ur
metals in effluent treatments. Bioresource Technology 98 (2007) 1940-1946. [6] M. Al Bahri, L. Calvo, M.A. Gilarranz, J.J. Rodriguez, Activated carbon from grape seeds
Jo
upon chemical activation with phosphoric acid: Application to the adsorption of diuron from water. Chemical Engineering Journal 203 (2012) 348-356. [7] M. Antunes, I.V.I. Esteves, R. Guégan, J.S. Crespo, A.N. Fernandes, M. Giovanela,
Removal of diclofenac sodium from aqueous solution by Isabel grape bagasse. Chemical Engineering Journal 192 (2012) 114-121. [8] H. Demiral, C Güng, Adsorption of copper(II) from aqueous solutions on activated carbon prepared from grape bagasse. Journal of Cleaner Production 124 (2016) 103-113. [9] G. Chen, Electrochemical technologies in wastewater treatment. Separation and Purification Technology 38 (2004) 11-41. 12
[10] L. Ali, M. Asim, T.A. Khan, Low cost adsorbents for the removal of organic pollutants from wastewater. Journal of Environmental Management 113 (2012) 170-183. [11] P. Xu, G.M. Zeng, D.L. Huang, C.L. Feng, S. Hu, M.H. Zhao, C. Lai, Z. Wei, C. Huang, G.X. Xie, Z.F. Liu, Use of iron oxide nanomaterials in wastewater treatment: A review. Science of the Total Environment 424 (2012) 1-10. [12] A. Sepehri, M.H. Sarrafzadeh, M. Avateffazeli, Interaction between Chlorella vulgaris and nitrifying-enriched activated sludge in the treatment of wastewater with low C/N ratio. Journal of Cleaner Production (2019) In press. [13] D. Jimenez-Cordero, F. Heras, N. Alonso-Morales, M.A. Gilarranz, J.J. Rodriguez, Porous structure and morphology of granular chars from flash and conventional pyrolysis of grape
ro of
seeds. Biomass and Bioenergy 54 (2013) 123-132. [14] H. Saygili, F. Güzel, Y. Onal, Conversion of grape industrial processing waste to activated carbon sorbent and its performance in cationic and anionic dyes adsorption. Journal of Cleaner Production 30 (2015) 1-10.
[15] J. Pastor-Villegas, J.F. Pastor-Valle, J.M.M Rodríguez, M.G. García, Study of
-p
commercial wood charcoals for the preparation of carbon adsorbents. Journal of Analytical and Applied Pyrolysis 76 (2006) 103-108.
re
[16] M. Wang, Z.H. Huang, G. Liu, F. Kang, Adsorption of dimethyl sulfide from aqueous solution by a cost-effective bamboo charcoal. Journal of Hazardous Materials 190 (2011) 10091015.
lP
[17] J.P. Kearns, L.S. Wellborn, R.S. Summers, D.R.U. Knappe, 2,4-D adsorption to biochars: Effect of preparation conditions on equilibrium adsorption capacity and comparison with commercial activated carbon literature data. Water Research 62 (2014) 20-28.
na
[18] J.F. González, S. Román, J.M. Encinar, G. Martínez, Pyrolysis of various biomass residues and char utilization for the production of activated carbons. Journal of Analytical and Applied Pyrolysis 85 (2009) 134-141.
ur
[19] P. Hadi, M.H. To, C.W. Hui, C.S.K. Lin, G. McKay, Aqueous mercury adsorption by activated carbons. Water Research 73 (2015) 37-55.
Jo
[20] R. Labied, O. Benturki, A.Y.E. Hamitouche, A. Donnot, Adsorption of hexavalent chromium by activated carbon obtained from a waste lignocellulosic material (ziziphus jujube cores): kinetic, equilibrium and thermodynamic study. Adsorption Science Technology 36 (2018) 1066-1099.
[21] R.S. Juang, F.C. Wu, R.L. Tseng, Characterization and use of activated carbons prepared from bagasses for liquid-phase adsorption. Colloids and Surfaces A: Physicochemical and Engineering Aspects 201 (2002) 191-199.
13
[22] M. Marciniak, J. Goscianska, M. Frankoswki, R. Pietrzak, Optimal synthesis of oxidized mesoporous carbons for the adsorption of heavy metal ions. Journal of Molecular Liquids 276 (2019) 630-637. [23] C.T. Driscoll, R.P. Mason, H.M. Chan, D.J. Jacob, N. Pirrone, Mercury as a global pollutant: sources, pathways, and effects. Environmental and Science Technology 47 (2013) 4967-4983. [24] World Health Organization (WHO) Guidelines for drinking-water quality, Geneva (2011). [25] A. Wahby, Z. Abdelouahab-Reddam, R. El Mail, M. Stitou, J. Silvestre-Albero, A. Sepúlveda-Escribano, F. Rodríguez-Reinoso, Mercury removal from aqueous solution by
ro of
adsorption on activated carbons prepared from olive stones. Adsorption 17 (2011) 603-609. [26] A.A. Ismaiel, M.K. Aroua, R. Yusoff, Palm shell activated carbon impregnated with taskspecific ionic-liquids as a novel adsorbent for the removal of mercury from contaminated water. Chemical Engineering Journal 225 (2013) 306-314.
[27] K. Vikrant, K.H. Kim, Nanomaterials for the removal of Hg(II) ions from water. Chemical
-p
Engineering Journal 358 (2019) 264-282.
[28] Y.O. Al-Ghamdi, K.A. Alamry, M.A. Hussein, H.M. Marwani, A.M. Asiri, Sulfone-
Science Technology 37 (2019) 139-159.
re
modified chitosan as selective adsorbent for the extraction of toxic Hg(II) metal ions. Adsorption
[29] H.N. Tran, S.J. You, A. Hosseini-Bandegharaei, H.P. Chao, Mistakes and inconsistencies
120 (2017) 88-116.
lP
regarding adsorption of contaminants from aqueous solutions: A critical review. Water Research
[30] A. Pawlicka, B. Doczekalska, Determination of surface oxygen functional groups of
(2013) 11-14.
na
active carbons according to the Boehm's titration method. Forestry and Wood Technology 84
[31] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, H.
ur
Nakatsuji, Gaussian 09, Revision D. 01, Gaussian, Inc., Wallingford, CT, 2010. [32] R.I. Dennington, T. Keith, J. Millam, GaussView, Version 5.0.8, Semichem. Inc.
Jo
Shawnee Mission, KS, 2008.
[33] A. Demirbas, Effects of temperature and particle size on bio-char yield from pyrolysis of
agricultural residues. Journal of Analytical and Applied Pyrolysis 72 (2004) 243-248. [34] T. Kan, V. Strezov, T.J. Evans, Lignocellulosic biomass pyrolysis: A review of product
properties and effects of pyrolysis parameters. Renewable and Sustainable Energy Reviews 57 (2016) 1126-1140. [35] X. Gai, H. Wang, J. Liu, L. Zhai, S. Liu, T. Ren, H. Liu, Effects of feedstock and pyrolysis temperature on biochar adsorption of ammonium and nitrate. Plos One 9 (2014) 1-19.
14
[36] F. Lu, P. Lu, Experiment study on adsorption characteristics of SO2, NOx by biomass chars. International Conference on Digital Manufacturing and Automation (2010) 682-685. [37] S.J. Gregg, K.S.W. Sing, Adsorption, Surface Area and Porosity, Academic Press, NY, 1982. [38] A. Dawlet, D. Talip, H.Y. Mi, MaLiKeZhaTi, Removal of mercury from aqueous solution using sheep bone charcoal. Procedia Environmental Sciences 18 (2013) 800-808. [39] G. Li, S. Wang, Q. Wu, F. Wang, D. Ding, B. Shen, Mechanism identification of temperature influence on mercury adsorption capacity of different halides modified bio-chars. Chemical Engineering Journal 315 (2017) 251-261. [40] M. Guerrero, M.P. Ruiz, A. Millera, M.U. Alzueta, R. Bilbao, Characterization of
ro of
biomass chars formed under different devolatilization conditions: Differences between rice husk and eucalyptus. Energy Fuels 22 (2008) 1275-1284.
[41] M.A. Nahil, P.T. Williams, Pore characteristics of activated carbons from the phosphoric acid chemical activation of cotton stalks. Biomass Bioenergy 37 (2012) 142-149.
[42] A. Paethanom, K. Yoshikawa, Influence of pyrolysis temperature on rice husk char
-p
characteristics and its tar adsorption capability. Energies 5 (2012) 4941-4951.
[43] K. Johari, N. Saman, S.T. Song, S.C. Cheu, H. Kong, H. Mat, Development of coconut
re
pith chars towards high elemental mercury adsorption performance: Effect of pyrolysis temperatures. Chemosphere 156 (2016) 56-68.
[44] I. Lima, C. Steiner, K.C. Das, Characterization of designer biochar produced at different
lP
temperatures and their effects on a loamy sand. Annals of Environmental Science 3 (2009) 195206.
[45] D. Angın, Effect of pyrolysis temperature and heating rate on biochar obtained from
na
pyrolysis of safflower seed press cake. Bioresource Technology 128 (2013) 593-597. [46] R. Mohammad-Rezaei, H. Razmi, Preparation and characterization of reduced graphene oxide doped in sol-gel derived silica for application in electrochemical double-layer capacitors.
ur
International Journal of Nanoscience and Nanotechnology 12 (2016) 233-241. [47] A. Kumar, H.M. Jena, High surface area microporous activated carbons prepared from
Jo
Fox nut (Euryale ferox) shell by zinc chloride activation. Applied Surface Science 356 (2015) 753-761.
[48] L.D. Mafu, H.W.J.P. Neomagus, R.C. Everson, C.A. Strydom, M. Carrier, G.N. Okolo,
J.R. Bunt, Chemical and structural characterization of char development during lignocellulosic biomass pyrolysis. Bioresource Technology 243 (2017) 941-948. [49] T.A. Saleh, Mercury sorption by silica/carbon nanotubes and silica/activated carbon: A comparison study. Journal of Water Supply: Research and Technology-AQUA 64 (2015) 892903.
15
[50] N. Asasian, T. Kaghazchi, Optimization of activated carbon sulfurization to reach adsorbent with the highest capacity for mercury adsorption. Separation Science and Technology 48 (2013) 2059-2072. [51] Y. Liu, Y.J. Liu, Biosorption isotherms, kinetics and thermodynamics. Separation Purification Technology 61 (2008) 229-242. [52] A. Ahmadpour, M. Zabihi, T.R. Bastami, M. Tahmasbi, A. Ayati, Rapid removal of mercury ion (II) from aqueous solution by chemically activated eggplant hull adsorbent. Journal of Applied Research in Water and Wastewater 6 (2016) 236-240. [53] A.A. Taha, A.H.E. Moustafa, H.H. Abdel-Rahman, Comparative biosorption study of Hg (II) using raw and chemically activated almond shell. Adsorption Science Technology 36 (2017)
ro of
521-548. [54] X. Dong, L.Q. Ma, Y. Zhu, Y. Li, B. Gu, Mechanistic investigation of mercury sorption
by Brazilian Pepper biochars of diff erent pyrolytic temperatures based on X‑ ray Photoelectron
Spectroscopy and Flow Calorimetry, Environmental Science Technology 47 (2013) 1215612164.
-p
[55] K. Yaghmaeian, R.K. Mashizi, S. Nasseri, A.H. Mahvi, M. Alimohammadi, S. Nazmara,
Removal of inorganic mercury from aquatic environments by multi-walled carbon nanotubes.
re
Journal of Environmental Health Science and Engineering (2015) 13-55.
[56] C.W. Cheung, C.K. Chan, J.F. Porter, G. McKay, Film-pore diffusion control for the
Science 234 (2001) 328-336.
lP
batch sorption of cadmium ions from effluent onto bone char. Journal of Colloid and Interface
[57] D.P. Rodda, B.B. Johnson, J.D. Wells, Modeling the effect of temperature on adsorption
(1996) 365-77.
na
of lead(II) and zinc onto goethite at constant pH. Journal of Colloid and Interface Science 184
[58] M. Machida, Y. Kikuchi, M. Aikawa, H. Tatsumoto, Kinetics of adsorption and desorption of Pb(II) in aqueous solution on activated carbon by two-site adsorption model.
ur
Colloids and Surfaces A: Physicochemical Engineering Aspects 240 (2004) 179-186. [59] R. Balasubramanian, S.V. Perumal, K. Vijayaraghavan, Equilibrium isotherm studies for
Jo
the multicomponent adsorption of lead, zinc, and cadmium onto Indonesian peat. Industrial and Engineering Chemistry Research 48 (2009) 2093-2099.
16
Figure captions Figure 1. Proposed surface structure of the grape bagasse-based char (M) for the analysis of mercury adsorption via DFT. a) Optimized structure calculated with the B3LYP/6-311G(d,p) theory level. b) ESP map of M with isovalue = 0.0004 where ESP value ranged from -133.5 kJ/mol (red) to +133.5 kJ/mol (blue). Figure 2. N2 adsorption-desorption isotherms at 77 K for the chars obtained from the pyrolysis of grape bagasse biomass. Figure 3. FTIR spectra of grape bagasse biomass and selected char samples. Figure 4. X-ray diffraction patterns of grape bagasse biomass and selected char samples. Figure 5. Adsorption kinetics for the mercury removal from aqueous solution using the best
ro of
grape bagasse-based char. Experimental conditions: pH 4 and 30 °C. Figure 6. Adsorption isotherms for the mercury removal from aqueous solution using the best grape bagasse-based char. Experimental conditions: a) pH 1.5 and 30 °C, b) pH 3.0 and 30 °C, c) pH 4.0 and 30 °C, d) pH 4 and 40 °C.
Figure 7. Surface functionalities of grape bagasse-based char analyzed in the mercury
Jo
ur
na
lP
re
Lactone functional group, c) Silicon functional group.
-p
adsorption via DFT calculations. Type of adsorption site: a) Carboxylic acid functional group, b)
17
Figure 1.
b)
ro of
-p
re
lP
na
ur
Jo a)
18
ro of
-p
re
lP
na
ur
Jo Figure 2.
19
1000 °C, 3 h
Transmittance, %
600 °C, 2 h
Precursor Si-OH C-H C=C O-H 3800
3000
2200 Wavenumber, cm-1
C=O C-O
ro of
C-H
1400
600
After mercury adsorption
-p
Before mercury adsorption
Jo
ur
na
lP
re
Figure 3.
20
Graphene-like structure
SiO2
900 0
1000 °C, 3 h
Intensity, a.u.
700 0
500 0
300 0
100 0
600 °C, 2 h
-1000
-3000
30
10
90
70
50
2θ,
ro of
Precursor
110
After mercury adsorption
Before mercury adsorption
Jo
ur
na
lP
re
-p
Figure 4.
21
Mercury adsorption capacity, mg/g
Pseudo second order model 200 mg/L 100 mg/L
16
14.3 mg/g
12 7.8 mg/g
8
4
0 200
400
600
800 1000 Time, min
1400
Jo
ur
na
lP
re
-p
Figure 5.
1200
22
1600
ro of
0
a) two-site Langmuir model Adsorption capacity of oxygenated groups Adsorption capacity of Si moieties
60
40 20.8 mg/g
20
0 0
200
400
600
800
1000
b) two-site Langmuir model Adsorption capacity of oxygenated groups Adsorption capacity of Si moieties
ro of
60
25.0 mg/g
0
na
20
400
600
800
1000
two-site Langmuir model Adsorption capacity of oxygenated groups Adsorption capacity of Si moieties
60 40
200
re
0
-p
20
lP
c)
Mercury adsorption capacity, mg/g
40
32.3 mg/g
0
0
ur
d)
400
600
800
1000
two-site Langmuir model Adsorption capacity of oxygenated groups Adsorption capacity of Si moieties
60
Jo
200
45.9 mg/g
40 20 0 0
200
400
600
800
1000
Mercury equilibrium concentration, mg/L Figure 6. 23
a)
re
-p
ro of
b)
Jo
ur
na
lP
c)
Figure 7.
24
Table captions Table 1. Experimental design utilized for the preparation of grape bagasse-based chars for the mercury adsorption from water. Table 2. Textural parameters of the grape bagasse-based chars obtained from pyrolysis. Table 3. Elemental composition and XRF analysis of the grape bagasse-based chars obtained from pyrolysis. Table 4. Kinetic adsorption parameters for the mercury removal from aqueous solution using the best grape bagasse-based char. Experimental conditions: pH 4 and 30 °C. Table 5. Results of the two-site Langmuir model for the mercury adsorption isotherms using the best grape bagasse-based char.
ro of
Table 6. Results of DFT calculations for the atomic interactions of the proposed adsorbent
Jo
ur
na
lP
re
-p
surface structure (M) and mercury ion.
25
Table 1. Pyrolysis conditions
Char performance
Temperature, ºC
Dwell time, h
Yield, %
q, mg/g
Total acidity, mmol/g
1
600
1
38.2
4.2
0.27
2
600
2
37.7
5.3
0.20
3
600
3
37.5
4.9
0.24
4
800
1
36.0
2.8
0.21
5
800
2
35.5
2.1
0.17
6
800
3
35.3
1.8
0.13
7
1000
1
34.0
1.3
0.18
8
1000
2
33.8
0.7
0.18
9
1000
3
33.7
0.5
Jo
ur
na
lP
re
-p
ro of
Sample
26
0.12
Table 2. Textural parameters Sample
SBET, m /g
SDFT, m2/g
VTotal, cm3/g
VMicropore, cm3/g
1
21.91
9.94
0.0252
0.003
2
15.50
12.84
0.0281
0.005
3
9.48
5.20
0.0145
0.001
4
15.64
5.40
0.0147
0.001
5
-
3.29
0.0105
0.001
6
6.83
4.45
0.0136
0.004
7
6.30
4.56
0.0133
0.002
8
5.73
4.36
0.013
0.002
9
12.57
5.40
0.0147
Jo
ur
na
lP
re
-p
Symbol “-“ indicates a poor adjustment of BET model.
ro of
2
27
0.001
Table 3. Elemental analysis, wt% Sample
C
H
N
O*
XRF, wt% Si
Al
Ca
P
K
Fe
Others
69.40 2.46 3.13 19.58 2.69 0.81 0.66 0.47 0.30 0.24
0.26
2
63.60 3.99 3.76 21.88 3.58 1.00 0.71 0.56 0.30 0.28
0.34
3
71.10 3.24 3.00 18.24 2.16 0.67 0.60 0.37 0.20 0.17
0.24
4
67.10 1.91 3.04 21.74 3.33 0.86 0.63 0.56 0.35 0.20
0.28
5
69.80 1.99 2.27 20.88 2.67 0.68 0.63 0.37 0.23 0.22
0.26
6
71.10 2.11 2.28 19.51 2.51 0.69 0.61 0.42 0.28 0.23
0.26
7
72.90 1.68 1.80 18.44 2.54 0.74 0.68 0.46 0.28 0.23
0.24
8
72.90 1.49 1.27 18.81 2.80 0.81 0.68 0.42 0.31 0.23
0.28
9
74.90 1.39 1.59 17.10 2.52 0.72 0.60 0.43 0.28 0.23
ro of
1
Jo
ur
na
lP
re
-p
* This value was calculated by difference.
28
0.24
Table 4. Mercury initial concentration, mg/L Adsorption kinetic model
100
200
R2
0.964
0.982
k, g/mg min
5.47E-03
1.10E-03
qte, mg/g
6.86
13.73
Jo
ur
na
lP
re
-p
ro of
Pseudo second order
Parameter
29
Table 5. Adsorption conditions
Parameter
T, °C
q1, mg/g
b1, L/mg
q2, mg/g
b2, L/mg
q’, mg/g
R2
1.5
30
26.92
8.08E-04
17.92
3.00E-03
20.9
0.994
3.0
30
26.21
8.94E-04
19.70
3.30E-03
25.89
0.997
4.0
30
30.23
1.02E-03
24.70
3.98E-03
33.03
0.992
4.0
40
45.10
1.06E-03
32.78
4.23E-03
46.76
0.993
Jo
ur
na
lP
re
-p
ro of
pH
30
Table 6. Char surface functionality
Estimated atomic distance for the
Energy, kJ/mol
interaction M Hg, Å 3.447
-1239
Lactone
3.653
-1233
Silicon moiety
2.813
-1237
Jo
ur
na
lP
re
-p
ro of
Carboxylic acid
31