Next Article in Journal
Effect of CaCl2 on Microstructure of Calciothermic Reduction Products of Ti2O3 to Prepare Porous Titanium
Next Article in Special Issue
One Step Hydrothermal Synthesis of Magnesium Silicate Impregnated Palm Shell Waste Activated Carbon for Copper Ion Removal
Previous Article in Journal
Study on Fracture-Split Performance of 36MnVS4 and Analysis of Fracture-Split Easily-Induced Defects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Removal of Pb2+ in Wastewater via Adsorption onto an Activated Carbon Produced from Winemaking Waste

by
Francisco José Alguacil
,
Lorena Alcaraz
,
Irene García-Díaz
and
Félix Antonio López
*
Centro Nacional de Investigaciones Metalúrgicas (CENIM), Consejo Superior de Investigaciones Científicas (CSIC), Avda. Gregorio del Amo 8, 28040 Madrid, Spain
*
Author to whom correspondence should be addressed.
Metals 2018, 8(9), 697; https://doi.org/10.3390/met8090697
Submission received: 24 July 2018 / Revised: 27 August 2018 / Accepted: 30 August 2018 / Published: 5 September 2018

Abstract

:
This work describes the adsorption of Pb2+ in aqueous solution onto an activated carbon (AC) produced from winemaking waste (cluster stalks). After characterizing the AC using Fourier transform infrared spectroscopy (FTIR) and micro-Raman spectroscopy, the influence of different physico-chemical factors (stirring rate, temperature, pH, adsorbent concentration, etc.) on its capacity to adsorb Pb2+ was examined. Kinetic and thermodynamic studies showed that the adsorption of the Pb2+ follows a pseudo-second-order kinetic model and fits the Langmuir isotherm model, respectively. The maximum adsorption capacity of the AC was 58 mg/g at 288 K temperature and pH of 4. In conclusion, ACs made from waste cluster stalks could be successfully used to remove Pb2+ from polluted water.

1. Introduction

Heavy metals, such as mercury, chromium, cadmium, arsenic, nickel and lead, are among the most dangerous industrial pollutants of water [1]. Their toxicity, persistence, and accumulation in living organisms have negative environmental and health effects [2,3] and are a source of increasing concern in many countries [4].
Lead is one of the most common and most toxic heavy metals found in industrial wastewater. It is released into the environment through mining, melting, galvanizing, and industrial metallurgical processes and from batteries, paints, ceramics, munitions, lead piping, etc. [5,6]. This can have serious effects on the human nervous, reproductive, and circulatory systems, kidneys, and liver, with children the most susceptible to intoxication [2,7,8].
The World Health Organization recommends the concentration of lead in drinking water to be under 0.1 mg/L [9], and different technologies have been developed to minimize its presence, including those based on precipitation [10], ion exchange [11], inverse osmosis [12], coagulation [13], electrodialysis [14], ultrafiltration [15], and supported liquid membranes [2]. However, all of these require large energy inputs, which is inefficient, and none is able to completely remove it [16,17]. However, adsorption methods for removing heavy metals continue to spur interest given their selectivity, low cost, ease of use, high efficiency (even when these metals are in low concentration), and the possibility of reusing the materials involved [6,18,19]. Polymers [20], silica [21], bionanocomposites [6], metal-organic frameworks (MOFs) supported on the nanofibrous membrane [22], and other materials such as activated carbon (AC) have all been studied as possible adsorbents. ACs are now used as adsorbents in many applications due to their high specific surface area and porous structure (a consequence of their large mesopore and micropore contents) [18]. They also have different functional groups, e.g., carboxyl, carbonyl, phenol, quinone, lactone groups, etc.) on their graphite layers [23,24,25], giving them a wide range of adsorption targets.
ACs can be obtained via the hydrothermal carbonization of organic substances, e.g., saccharides (glucose, sucrose, starch, etc.) or directly from biomass. This treatment brings the organic material used into contact with water at a temperature of 423–623 K under autogenic pressure (i.e., produced through heating), producing a hydrothermal carbon (HTC)—a solid that is rich in carbon [26,27]—with many functional oxygenated groups. By means of its chemical or thermal activation, this HTC can be turned into an AC [26]. In recent years, it has become possible to produce ACs from old carpets [28], peach stones [29], apple peel [30], grape seeds [31], and other materials.
Spain produces huge quantities of agricultural waste, including winemaking waste. Currently, winemaking waste has to be taken away from wineries, generating transport and storage costs. It could, however, provide an abundant and cheap source of material for producing AC, which would solve an environmental problem and also return a profit.
The present work examines the capacity of an AC—derived from a KOH-activated [32] HTC produced through the hydrothermal pyrolysis of waste cluster stalks (a winemaking waste)—to adsorb Pb2+ from aqueous solution. The influence of temperature, stirring rate, pH, contact time, quantity of AC, and concentration of Pb2+ on the adsorption process was investigated.

2. Materials and Methods

2.1. Synthesis of HTC and AC

Waste cluster stalks, which are generated during the production of Albariño wine (Denomination of Origin ‘Rías Baixas’, Galicia) and supplied by the Misió Biológica de Galicia (CSIC), were used to produce an AC. These cluster stalks were ground to a grain size of <2 mm diameter in an SK 100 Cross Beater Mill (Retsch, Haan, Germany). An aqueous suspension (75 g·L−1) of this milled material was then introduced into a Berghof BR300 high pressure reactor (Berghof GmbH, Eningen, Germany) at 523 K and 30 bars for 3 h, with constant stirring at 1400 rpm [31]. The resulting suspension was filtered using a Millipore YT30 pressure filter (Merck KGaA, Darmstadt, Germany) to extract the HTC produced, which was dried in an oven at 353 K for 12 h. AC was produced from this HTC via exposure to KOH (weight ratio 1:2) at 1073 K for 2 h in a Carbolite STF 15 tubular oven (Carbolite Gero Limited, Hope Valley, United Kingdom) under a N2 atmosphere (150 mL·min−1). After cooling to room temperature in the oven, the AC was repeatedly washed with MilliQ water until a neutral pH was achieved for the run-off water. The washed AC was than dried at 353 K for 12 h.

2.2. Characterization of the Activated Carbon

The elemental analysis (C, H, N, and S) was determined with an elemental analyzer (Leco CHNS 932, Leco Corporation, Saint Joseph, MI, USA). The proximate analysis was conducted according to ASTM D 3172–3175 test standards and the results were given as moisture, volatile matter, ash, and fixed carbon contents.
The Z potential was measured using a Zetasizer Malvern Nano ZS (Malvern Panalytical Ltd., Worcestershire, United Kingdom) at 25°. To study the influence of pH, aqueous suspensions were prepared in pH solutions between 1 and 13 using solutions of HCl and 0.5 M NaOH. An aqueous 10−3 M (1mM) KCl solution was used as the electrolyte. The concentration of activated carbon was adjusted to a value of 18 g L−1.The suspensions were dispersed with the aid of a sonicator Bandelin Electconic Sonopuls HD 3100, (Bandelin electronic GmbH & Co. KG, Berlin, Germany) with amplitude of 60% for 150 s.
The structural characteristics of the AC were examined by Fourier transform infrared spectroscopy (FTIR) using a Varian 670 FTIR spectrometer (Varian Inc., Palo Alto, CA, USA) with a spectral range of 1600–600 cm−1 and a spectral resolution of 4 cm−1 in transmittance mode. Micro-Raman spectroscopy was performance using a Horiba Jobin Yvon LabRAM HR800 system (Horiba Scientific, Kyoto, Japan), For the Raman measurements, the sample was excited using a He-Cd laser (633 nm) under a confocal Olympus BX 41 microscope (Olympus, Tokyo, Japan) with a 40× objective lens. A charge-coupled detector was used to collect the scattered light dispersed by a 2400 lines·mm−1 grating (micro-Raman). The spectral resolution of the system used was 1.5 cm−1.
The porous structure of the AC was characterized by its N2 adsorption–desorption isotherm at 77 K using a Micromeritics ASAP 2010 Accelerated Surface Area and Porosimetry System (Micromeritics, Norcross, GA, USA). The specific surface area was determined by analyzing the adsorption isotherm using the Brunauer-Emmett-Teller (BET) equation and density functional theory (DFT) models, with calculations performed using Micromeritics and Quantachrome software (Version 1.01, Quantacrome Instruments, Boynton Beach, FL, USA). The BET and DFT results were compared using Kaneco and Dubinin equations.

2.3. Batch Adsorption Experiments

Loading and elution batch experiments were carried out in a 250 mL glass reactor provided for mechanical shaking via a 27-mm diameter, four-blade glass impeller. Metal content in the solution was analyzed by flame atomic adsorption spectrometry using a Varian SpectrAA 220G (Varian Inc., Palo Alto, CA, USA), and the metal content in the carbon was estimated by the mass balance. The metal uptake capacity onto AC was obtained by the mass balance Equation (1):
q t = ( c 0 c t ) · V m
where qt is the amount of adsorbed metal ions (mg Pb2+ g−1 sorbent), while c0 and ct are metal ion concentrations (mg Pb2+ L−1) in the aqueous solution initially and that at time t, respectively. V is the volume of the solution in (L) and m is the mass of the sorbent (g).

3. Results and Discussion

3.1. Characterization of the Activated Carbon

3.1.1. Chemical Composition and Z Potential

The proximate and ultimate analysis results related to the AC are given in Table 1. Figure 1 shows the variation of Z potential as a function of pH in 10−3 M KCl solution. The value of the isoelectric point for activated carbon is 1.42.

3.1.2. Fourier Transform Infrared and Raman Spectroscopy

Figure 2 shows the FTIR spectrum of the activated carbon. Different absorption bands can be seen in the range of 1600–600 cm−1.
Absorption bands at around 800 cm−1 are usually attributed to C–H and CH=CH2 stretching vibration modes in aromatic structures [33,34]. The band centered at 1023 cm−1 can be assigned to alcohol groups (R–OH) [35]. The absorption band at around 1100 cm−1 indicates the presence of C–O groups characteristic of esters, ether, phenols, organic acids, and C–O bonds within functional groups [34,36,37]. Finally, the band at around 1200 cm−1 is associated with tension in the C–O bonds of phenol groups [38].
Figure 3 shows the Raman spectrum of the AC, which has two predominant bands at about 1340 and 1600 cm−1. These are attributable to the D and G bands typical of carbonaceous materials.
The D band is related to disordered carbon atoms and thus highlights defects in the crystalline structure. The G band reflects the crystallinity and graphitic structure of carbon materials [39,40]. The D band is more intense than the G band, indicating that AC has a large amount of highly disordered graphite [41].

3.1.3. N2 adsorption-Desorption Isotherms

Figure 4 shows the N2 adsorption-desorption isotherms at 77 K for the AC. A large amount of N2 was adsorbed at low pressure of approximately 979 cm3·g−1 at a p/p0 ~1. According to the IUPAC, the shape of the isotherm is type I for very low partial pressures (p/p0) and type II for very high pressures. The information in Table 2 reveals the simultaneous presence of a large number of micropores (<2 nm diameter) and fewer mesopores (2–50 nm diameter) [40].

3.2. Batch Adsorption Studies

3.2.1. Effect of the Stirring Rate

The effect of the stirring rate was investigated in order to optimize AC lead loading using an aqueous solution of 0.01 g·L−1 Pb (II) at pH 4 and 0.25 g·L−1 AC (Table 3). Figure 5 shows that metal load increased with the stirring rate, reaching a maximum at 500–750 rpm. Thus, in this range, the thickness of the aqueous film boundary layer was at its minimum, and maximum Pb2+ loading could therefore be achieved. The reduction in Pb2+ uptake at higher stirring rates may be due to (i) local equilibrium between the solution and the carbon particles and/or (ii) the agglomeration of carbon particles and the consequent reduction in the active surface area of the adsorbent.

3.2.2. Effect of Temperature

The influence of temperature (288–333 K) on Pb2+ adsorption by the AC was investigated using the same aqueous solutions and AC concentration as above. The results showed temperature to have no influence on Pb2+ adsorption (always 38 mg·g−1), although it did influence the rate at which the system reached equilibrium (Figure 6).
In the 288–333 K temperatures range, the system fits a pseudo-second-order kinetic model well (Figure 7):
t [ Pb ] t = 1 k 2 [ Pb ] e 2 + t [ Pb ] e
with constant kinetic rates of 4 × 10−3 (R2 = 0.9993), 8 × 10−3 (R2 = 0.9999), and 3 × 10−2 (R2 = 0.9992) g·mg−1·min−1 at 288, 313, and 333 K, respectively. In Equation (2), t is the time elapsed; [Pb]t and [Pb]e are the lead concentrations on the AC at that time and at equilibrium, respectively; and k2 is the rate constant.

3.2.3. Influence of pH

Table 4 and Figure 8 show the effect of the aqueous solution’s pH on the loading of Pb2+ onto the AC. In this experiment, an aqueous solution of 0.01 g/L Pb(II) with different pH and 0.25 g/L AC was used; the pH experiments were also done with higher AC concentrations. Pb2+ adsorption fell sharply between pH 4 and pH 2; the low adsorption values at low pH were generally due to the surface charge. In the AC, the pHpzc was 1.4 and at pH around the zero point, the surface charge was neutral; therefore, the cation (Pb2+) adsorption was lower. However, at pH values higher than the pHpzc, such as pH 4, the surface charge was negative and the adsorption of the cations was greater. The same behavior was obtained when a higher AC concentration was used.
The results obtained at pH 4 were used to determine the probable rate law governing the adsorption of the Pb2+ onto the AC. Three possible adsorption mechanisms were contemplated: (i) the diffusion of Pb2+ from the aqueous solution to the AC surface (film diffusion), (ii) the diffusion of ions within the AC (particle diffusion), and (iii) the moving boundary process. Figure 9 summarizes the results obtained.
film-diffusion   controlled   process :   ln ( 1 F ) = k t
particle-diffusion   controlled   process :   ln ( 1 F 2 ) = k t
moving   boundary   process :   3 3 · ( 1 F ) 2 3 2 F = k t
The best fit (R2 = 0.9969) for the present system corresponded to the particle diffusion controlled process, as outlined in Equation (4), where k is the rate constant (0.042 min−1), t the time elapsed, and F the fractional approach to the equilibrium, defined as:
F = [ Pb ] t [ Pb ] e
where [Pb]t and [Pb]e are the Pb2+ concentrations in the solution at the elapsed time t and at equilibrium, respectively.

3.2.4. Influence of AC Concentration

Table 5 shows the effect of AC concentration on Pb2+ removal from the aqueous solution (0.01 g L−1 Pb(II), pH 4, 288 K) over 3 h. Increasing the AC concentration increased the percentage of Pb2+ loaded. Figure 10 shows the metal loading isotherm generated by plotting the Pb2+ concentration of the AC against that of the aqueous solution. Figure 11 shows the fit of the obtained isotherm to the Freundlich—Equation (7), Langmuir—Equation (8), and Temkin—Equation (9) linear isotherms.
Freundlich isotherm :   ln q e = ln K F + 1 n ln C e
Langmuir isotherm :   C e q e = 1 q m b + 1 q m c e
Temkin isotherm :   q e = B · ln A T + B · ln c e   with   B = R T b T
where qe (mg·g−1) is the quantity of metal adsorbed per mass of AC at equilibrium; KF (L·g−1) and n are adsorption constants; 1/n is a measure of the intensity of adsorption; qm (mg·g−1) is the maximum adsorption capacity of the AC; b (L·mg−1) the Langmuir equilibrium constant related to the adsorption energy; Ce the concentration of Pb2+ in the aqueous solution at equilibrium (mg·L−1); AT is the Temkin isotherm equilibrium binding constant (L·g−1); B (R·T/bT) is a constant related to the heat of sorption (J·mol−1); bT is the Temkin isotherm constant; and R is the universal gas constant (8.314 × 103 kJ·K−1·mol−1).
Table 6 shows the calculated parameter values obtained from the linear fitting of Equations (7)–(9). The results best fitted the Langmuir equation, with an R2 value of 0.9989. Accordingly, the maximum theoretical adsorption of Pb by the AC should be 58 mg·g−1, which is similar to the 55–57 mg·g−1 obtained experimentally. Pb adsorption occurs on a homogeneous adsorbent surface, forming a monolayer in which each adsorption site can take a single molecule of adsorbate with the same adsorption energy [42]. Table 7 shows the qm values in Pb adsorption onto AC obtained from other materials. Although the experimental conditions were not the same, the qm value obtained was similar to that found in the literature.

3.2.5. Influence of Ionic Strength

It happens very often that the aqueous solution containing the metal to be removed contains other metallic and nonmetallic components that affect its ionic strength. The effect of this variable on Pb2+ adsorption onto the AC was investigated by adding different amounts of lithium nitrate to the aqueous solution (Pb2+ 0.01 g·L−1, AC concentration 0.25 g·L−1). Table 8 shows that increasing the ionic strength dramatically reduced Pb2+ adsorption capacity, which indicates that the electrostatic forces between the AC surface and Pb2+ are attractive [49]. This variable must therefore be taken into account in any scaling-up of the process.

3.2.6. Performance of the AC Compared to Other Adsorbents/Ion Exchangers

Table 9 shows how the present AC performed against other adsorbents/ion exchangers (aqueous solution at pH 4, Pb2+ concentration 0.01 g·L−1, AC concentration 0.25 g·L−1; other adsorbent/ion exchanger concentration 0.25 g·L−1). In terms of Pb2+ adsorbed, the AC performed better than multiwalled carbon nanotubes (MWCNT) or Lewatit EP-63 resin. The Lewatit K2621 and Lewatit TP 208 resins performed as well as the AC in this respect but were much slower (30%, 27%, and 81%, at 30 min, respectively, and 65%, 60%, and 94% at 2 h, respectively).

3.2.7. Comparative Adsorption of Pb2+ and Others Metals in Solution

The removal of other metals from the aqueous solution by the AC was compared to its adsorption of Pb2+. For this, monoelemental solutions containing 0.01 g·L−1 of each metal at pH 4 were used. Table 10 shows that at maximum Pb2+ uptake, the load of the other metals on the AC was quite low. These results in no way suggest that this AC is unsuitable for the removal of other metals in solution; they simply mean that under the present experimental conditions, Pb2+ adsorption was higher than the other metals.

3.2.8. Pb2+ Elution

Elution experiments were performed with the AC loaded with 2 mg Pb2+/0.05 g activated carbon. The influence of temperature (293–333 K) on Pb2+ elution was investigated using 0.1 M nitric acid solutions as the eluent at a ratio of 4000 mL per mg AC. Temperature was seen to have negligible influence (with around 65% Pb2+ elution recorded across the temperature range); equilibrium was always reached within 15 min.
The effect of the concentration of the nitric acid eluent (0.1–0.25 M) was also investigated under the same experimental conditions but at 293 K. As the nitric acid concentration increased, so did the amount of Pb2+ eluted; after 3 h with 0.25 M nitric acid, 94% of the Pb2+ had been eluted compared to 65% when 0.1 M nitric acidic was used.
The effect of varying the eluent (nitric acid 0.1 M) volume/AC concentration ratio (at 293 K) was also investigated. Table 11 shows that at all the ratios tested, the quantity of Pb2+ eluted was constant; the remaining lead concentration on the AC was 14 mg·g−1. Therefore, the maximum lead concentration—130 mg L−1—was reached for a volume eluent/AC concentration ratio of 200 mL g−1.

4. Conclusions

HTC was satisfactorily obtained from cluster stalks and activated via exposure to KOH to produce an AC. The FTIR and Raman spectra for this AC showed bands typical of carbonaceous materials. The adsorption-desorption N2 isotherm for the AC showed it had a porous structure mainly dominated by micropores. Although temperature was found not to influence Pb2+ absorption by the AC, equilibrium was reached more quickly at higher temperatures. The adsorption kinetics fitted a pseudo-second-order kinetic model. A direct relationship was seen between the AC concentration and the percentage of Pb2+ removed from the aqueous solution. The Pb2+ adsorption isotherm fitted the Langmuir model well, with a maximum adsorption capacity value of 58 mg/g. AC made from winemaking waste, such as cluster stalks, could therefore be successfully used to remove Pb2+ from polluted water. Making such AC from this material not only provides a means of dealing with the environmental issues caused by Pb2+ but also solves the winemaking industry’s problem of how to dispose of its waste cluster stalks by actually turning them into a profitable material.

Author Contributions

Investigation, Methodology, Writing-Review: F.J.A.; Investigation, Supervision, Writing-Review & Editing: F.A.L.; Formal analysis, Writing-original draft, Validation: L.A.; Formal analysis, Writing-Review: I.G.-D.

Funding

This research received no external funding.

Acknowledgments

The authors thank the Biological Mission of Galicia (CSIC) for the supply of vine waste necessary to obtain the activated carbon.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yu, X.L.; He, Y. Optimal ranges of variables for an effective adsorption of lead(II) by the agricultural waste pomelo (Citrus grandis) peels using Doehlert designs. Sci. Rep. 2018, 8, 729. [Google Scholar] [CrossRef] [PubMed]
  2. De Agreda, D.; Garcia-Diaz, I.; López, F.A.; Alguacil, F.J. Supported liquid membranes technologies in metals removal from liquid effluents. Rev. Metal. 2011, 47, 146–168. [Google Scholar] [Green Version]
  3. Has-Schön, E.; Bogut, I.; Strelec, I. Heavy Metal Profile in Five Fish Species Included in Human Diet, Domiciled in the End Flow of River Neretva (Croatia). Arch. Environ. Contam. Toxicol. 2006, 50, 545–551. [Google Scholar] [CrossRef] [PubMed]
  4. Edathil, A.A.; Shittu, I.; Hisham Zain, J.; Banat, F.; Haija, M.A. Novel magnetic coffee waste nanocomposite as effective bioadsorbent for Pb(II) removal from aqueous solutions. J. Environ. Chem. Eng. 2018, 6, 2390–2400. [Google Scholar] [CrossRef]
  5. Wongrod, S.; Simon, S.; Guibaud, G.; Lens, P.N.L.; Pechaud, Y.; Huguenot, D.; van Hullebusch, E.D. Lead sorption by biochar produced from digestates: Consequences of chemical modification and washing. J. Environ. Manag. 2018, 219, 277–284. [Google Scholar] [CrossRef] [PubMed]
  6. Seema, K.M.; Mamba, B.B.; Njuguna, J.; Bakhtizin, R.Z.; Mishra, A.K. Removal of lead (II) from aqeouos waste using (CD-PCL-TiO2) bio-nanocomposites. Int. J. Biol. Macromol. 2018, 109, 136–142. [Google Scholar] [CrossRef] [PubMed]
  7. Beltrame, K.K.; Cazetta, A.L.; de Souza, P.S.C.; Spessato, L.; Silva, T.L.; Almeida, V.C. Adsorption of caffeine on mesoporous activated carbon fibers prepared from pineapple plant leaves. Ecotoxicol. Environ. Saf. 2018, 147, 64–71. [Google Scholar] [CrossRef] [PubMed]
  8. Sone, H.; Fugetsu, B.; Tanaka, S. Selective elimination of lead(II) ions by alginate/polyurethane composite foams. J. Hazard. Mater. 2009, 162, 423–429. [Google Scholar] [CrossRef] [PubMed]
  9. World Health Organisation. Guidelines for Drinking-Water Quality, 3rd ed.; WHO Library Cataloguing-in-Publication Data; World Health Organisation: Geneva, Switzerland, 2004; Volume 1, pp. 10–330. [Google Scholar]
  10. Kavak, D. Removal of lead from aqueous solutions by precipitation: Statistical analysis and modeling. Desalin. Water Treat. 2013, 51, 1720–1726. [Google Scholar] [CrossRef]
  11. Murray, A.; Örmeci, B. Use of polymeric sub-micron ion-exchange resins for removal of lead, copper, zinc, and nickel from natural waters. J. Environ. Sci. 2018. [Google Scholar] [CrossRef]
  12. Mlayah, A.; Jellali, S. Study of continuous lead removal from aqueous solutions by marble wastes: Efficiencies and mechanisms. Int. J. Environ. Sci. Technol. 2015, 12, 2965–2978. [Google Scholar] [CrossRef]
  13. Pang, F.M.; Kumar, P.; Teng, T.T.; Mohd Omar, A.K.; Wasewar, K.L. Removal of lead, zinc and iron by coagulation–flocculation. J. Taiwan Inst. Chem. Eng. 2011, 42, 809–815. [Google Scholar] [CrossRef]
  14. Gherasim, C.-V.; Křivčík, J.; Mikulášek, P. Investigation of batch electrodialysis process for removal of lead ions from aqueous solutions. Chem. Eng. J. 2014, 256, 324–334. [Google Scholar] [CrossRef]
  15. Jana, D.K.; Roy, K.; Dey, S. Comparative assessment on lead removal using micellar-enhanced ultrafiltration (MEUF) based on a type-2 fuzzy logic and response surface methodology. Sep. Purif. Technol. 2018, 207, 28–41. [Google Scholar] [CrossRef]
  16. Fu, F.; Wang, Q. Removal of heavy metal ions from wastewaters: A review. J. Environ. Manag. 2011, 92, 407–418. [Google Scholar] [CrossRef] [PubMed]
  17. Eccles, H. Treatment of metal-contaminated wastes: Why select a biological process? Trends Biotechnol. 1999, 17, 462–465. [Google Scholar] [CrossRef]
  18. Xu, J.; Cao, Z.; Zhang, Y.; Yuan, Z.; Lou, Z.; Xu, X.; Wang, X. A review of functionalized carbon nanotubes and graphene for heavy metal adsorption from water: Preparation, application, and mechanism. Chemosphere 2018, 195, 351–364. [Google Scholar] [CrossRef] [PubMed]
  19. Taraba, B.; Bulavová, P. Adsorption enthalpy of lead(II) and phenol on coals and activated carbon in the view of thermodynamic analysis and calorimetric measurements. J. Chem. Thermodyn. 2018, 116, 97–106. [Google Scholar] [CrossRef]
  20. Kaya, I.G.B.; Duranoglu, D.; Beker, U.; Senkal, B.F. Development of Polymeric and Polymer-Based Hybrid Adsorbents for Chromium Removal from Aqueous Solution. CLEAN-Soil Air Water 2011, 39, 980–988. [Google Scholar] [CrossRef]
  21. Walcarius, A.; Mercier, L. Mesoporous organosilica adsorbents: Nanoengineered materials for removal of organic and inorganic pollutants. J. Mater. Chem. 2010, 20, 4478. [Google Scholar] [CrossRef]
  22. Efome, J.E.; Rana, D.; Matsuura, T.; Lan, C.Q. Metal-organic frameworks supported on nanofibers to remove heavy metals. J. Mater. Chem. A 2018, 6, 4550–4555. [Google Scholar] [CrossRef]
  23. Ravulapalli, S.; Kunta, R. Removal of lead (II) from wastewater using active carbon of Caryota urens seeds and its embedded calcium alginate beads as adsorbents. J. Environ. Chem. Eng. 2018, 6, 4298–4309. [Google Scholar] [CrossRef]
  24. Renu, M.A.; Singh, K.; Upadhyaya, S.; Dohare, R.K. Removal of heavy metals from wastewater using modified agricultural adsorbents. Mater. Today Proc. 2017, 4, 10534–10538. [Google Scholar] [CrossRef]
  25. Pawar, R.R.; Kim, M.; Kim, J.-G.; Hong, S.-M.; Sawant, S.Y.; Lee, S.M. Efficient removal of hazardous lead, cadmium, and arsenic from aqueous environment by iron oxide modified clay-activated carbon composite beads. Appl. Clay Sci. 2018, 162, 339–350. [Google Scholar] [CrossRef]
  26. Jain, A.; Balasubramanian, R.; Srinivasan, M.P. Hydrothermal conversion of biomass waste to activated carbon with high porosity: A review. Chem. Eng. J. 2016, 283, 789–805. [Google Scholar] [CrossRef]
  27. Sevilla, M.; Fuertes, A.B. The production of carbon materials by hydrothermal carbonization of cellulose. Carbon 2009, 47, 2281–2289. [Google Scholar] [CrossRef] [Green Version]
  28. Hassan, A.F.; Elhadidy, H. Production of activated carbons from waste carpets and its application in methylene blue adsorption: Kinetic and thermodynamic studies. J. Environ. Chem. Eng. 2017, 5, 955–963. [Google Scholar] [CrossRef]
  29. Tsoncheva, T.; Mileva, A.; Tsyntsarski, B.; Paneva, D.; Spassova, I.; Kovacheva, D.; Velinov, N.; Karashanova, D.; Georgieva, B.; Petrov, N. Activated carbon from Bulgarian peach stones as a support of catalysts for methanol decomposition. Biomass Bioenergy 2018, 109, 135–146. [Google Scholar] [CrossRef]
  30. Enniya, I.; Rghioui, L.; Jourani, A. Adsorption of hexavalent chromium in aqueous solution on activated carbon prepared from apple peels. Sustain. Chem. Pharm. 2018, 7, 9–16. [Google Scholar] [CrossRef]
  31. Jimenez-Cordero, D.; Heras, F.; Alonso-Morales, N.; Gilarranz, M.A.; Rodriguez, J.J. Porous structure and morphology of granular chars from flash and conventional pyrolysis of grape seeds. Biomass Bioenergy 2013, 54, 123–132. [Google Scholar] [CrossRef]
  32. Alcaraz, L.; López Fernández, A.; García-Díaz, I.; López, F.A. Preparation and characterization of activated carbons from winemaking wastes and their adsorption of methylene blue. Adsorpt. Sci. Technol. 2018, 36, 1331–1351. [Google Scholar] [CrossRef] [Green Version]
  33. Unur, E. Functional nanoporous carbons from hydrothermally treated biomass for environmental purification. Microporous Mesoporous Mater. 2013, 168, 92–101. [Google Scholar] [CrossRef]
  34. Li, M.; Li, W.; Liu, S. Hydrothermal synthesis, characterization, and KOH activation of carbon spheres from glucose. Carbohydr. Res. 2011, 346, 999–1004. [Google Scholar] [CrossRef] [PubMed]
  35. Altintig, E.; Kirkil, S. Preparation and properties of Ag-coated activated carbon nanocomposites produced from wild chestnut shell by ZnCl2 activation. J. Taiwan Inst. Chem. Eng. 2016, 63, 180–188. [Google Scholar] [CrossRef]
  36. Liu, S.; Sun, J.; Huang, Z. Carbon spheres/activated carbon composite materials with high Cr(VI) adsorption capacity prepared by a hydrothermal method. J. Hazard. Mater. 2010, 173, 377–383. [Google Scholar] [CrossRef] [PubMed]
  37. Salman, J.M.; Njoku, V.O.; Hameed, B.H. Adsorption of pesticides from aqueous solution onto banana stalk activated carbon. Chem. Eng. J. 2011, 174, 41–48. [Google Scholar] [CrossRef]
  38. Foo, K.Y.; Hameed, B.H. Factors affecting the carbon yield and adsorption capability of the mangosteen peel activated carbon prepared by microwave assisted K2CO3 activation. Chem. Eng. J. 2012, 180, 66–74. [Google Scholar] [CrossRef]
  39. Glonek, K.; Wróblewska, A.; Makuch, E.; Ulejczyk, B.; Krawczyk, K.; Wróbel, R.J.; Koren, Z.C.; Michalkiewicz, B. Oxidation of limonene using activated carbon modified in dielectric barrier discharge plasma. Appl. Surf. Sci. 2017, 420, 873–881. [Google Scholar] [CrossRef]
  40. Chen, W.F.; Pan, L.; Chen, L.F.; Yu, Z.; Wang, Q.; Yan, C.C. Comparison of EDTA and SDS as potential surface impregnation agents for lead adsorption by activated carbon. Appl. Surf. Sci. 2014, 309, 38–45. [Google Scholar] [CrossRef]
  41. Pereira, L.; Pereira, E.; Cremades, A.; Piqueras, J.; Jiménez, J.; Bielza, J. Characterisation of different habits in torch-flame-grown diamond and diamond-like films. Diam. Relat. Mater. 1999, 8, 1333–1341. [Google Scholar] [CrossRef]
  42. Basu, M.; Guha, A.K.; Ray, L. Adsorption of Lead on Cucumber Peel. J. Clean. Prod. 2017, 151, 603–615. [Google Scholar] [CrossRef]
  43. Kobya, M.; Demirbas, E.; Senturk, E.; Ince, M. Adsorption of heavy metal ions from aqueous solutions by activated carbon prepared from apricot stone. Bioresour. Technol. 2005, 96, 1518–1521. [Google Scholar] [CrossRef] [PubMed]
  44. Ho, Y.S.; Wase, D.A.J.; Forster, C.F. Removal of Lead Ions From Aqueous Solution Using Sphagnum Moss Peat as Adsorbent. Water SA 1996, 22, 219–224. [Google Scholar]
  45. Largitte, L.; Gervelas, S.; Tant, T.; Couespel Dumesnil, P.; Lodewyckx, P. Influence of the surface properties of the Bois Carré seeds activated carbon on the removal of lead from aqueous solutions. Eurasian Chem. J. 2013, 14, 201–210. [Google Scholar] [CrossRef]
  46. Machida, M.; Aikawa, M.; Tatsumoto, H. Prediction of simultaneous adsorption of Cu(II) and Pb(II) onto activated carbon by conventional Langmuir type equations. J. Hazard. Mater. 2005, 120, 271–275. [Google Scholar] [CrossRef] [PubMed]
  47. Kikuchi, Y.; Qian, Q.; Machida, M.; Tatsumoto, H. Effect of ZnO loading to activated carbon on Pb(II) adsorption from aqueous solution. Carbon 2006, 44, 195–202. [Google Scholar] [CrossRef]
  48. Ricordel, S. Heavy metals removal by adsorption onto peanut husks carbon: Characterization, kinetic study and modeling. Sep. Purif. Technol. 2001, 24, 389–401. [Google Scholar] [CrossRef]
  49. Al-Degs, Y.S.; El-Barghouthi, M.I.; El-Sheikh, A.H.; Walker, G.M. Effect of solution pH, ionic strength, and temperature on adsorption behavior of reactive dyes on activated carbon. Dyes Pigments 2008, 77, 16–23. [Google Scholar] [CrossRef]
Figure 1. Z potential of activated carbon.
Figure 1. Z potential of activated carbon.
Metals 08 00697 g001
Figure 2. Fourier transform infrared (FTIR) spectrum of the activated carbon.
Figure 2. Fourier transform infrared (FTIR) spectrum of the activated carbon.
Metals 08 00697 g002
Figure 3. Raman spectrum of the activated carbon.
Figure 3. Raman spectrum of the activated carbon.
Metals 08 00697 g003
Figure 4. Nitrogen adsorption-desorption isotherms for the activated carbon.
Figure 4. Nitrogen adsorption-desorption isotherms for the activated carbon.
Metals 08 00697 g004
Figure 5. Influence of the stirring rate on Pb2+ loading.
Figure 5. Influence of the stirring rate on Pb2+ loading.
Metals 08 00697 g005
Figure 6. Influence of temperature on Pb2+ removal from the aqueous solution. Stirring rate: 750 rpm.
Figure 6. Influence of temperature on Pb2+ removal from the aqueous solution. Stirring rate: 750 rpm.
Metals 08 00697 g006
Figure 7. Influence of temperature on Pb2+ removal from the aqueous solution. Stirring rate: 750 rpm.
Figure 7. Influence of temperature on Pb2+ removal from the aqueous solution. Stirring rate: 750 rpm.
Metals 08 00697 g007
Figure 8. Influence of temperature on Pb2+ removal from the aqueous solution. Stirring rate: 750 rpm.
Figure 8. Influence of temperature on Pb2+ removal from the aqueous solution. Stirring rate: 750 rpm.
Metals 08 00697 g008
Figure 9. Adsorption mechanisms of Pb2+on activated carbon.
Figure 9. Adsorption mechanisms of Pb2+on activated carbon.
Metals 08 00697 g009
Figure 10. Pb2+ loading isotherm. Temperature: 288 K. Time: 3 h. Stirring rate: 750 rpm.
Figure 10. Pb2+ loading isotherm. Temperature: 288 K. Time: 3 h. Stirring rate: 750 rpm.
Metals 08 00697 g010
Figure 11. The linear isotherms of (a) Freundlich, (b) Langmuir, and (c) Temkin.
Figure 11. The linear isotherms of (a) Freundlich, (b) Langmuir, and (c) Temkin.
Metals 08 00697 g011
Table 1. Chemical characteristics of activated carbon (AC).
Table 1. Chemical characteristics of activated carbon (AC).
ParameterValue
Proximate analysis a
Moisture4.50
Ash11.50
Volatile matter15.00
Fixed Carbon69.00
Ultimate analysis a
C82.00
H2.70
N0.99
O b14.5
S0.10
a Dry basis, wt %; b by difference.
Table 2. Porosity features of the activated carbon.
Table 2. Porosity features of the activated carbon.
Total Pore Volume (Vp) (cm3·g−1)1.40
Micropore Volume (W0) (cm3·g−1)0.95
Average Micropore Size (L0) (nm)1.71
Microporous Surface Area (Smi) (m2·g−1)1111
Non-Microporous Surface Area (Se) (m2·g−1)19
Total Surface Area (Stotal) (m2·g−1)1194
SBET (m2·g−1)2662
Table 3. Influence of the stirring rate on Pb2+ loading.
Table 3. Influence of the stirring rate on Pb2+ loading.
Stirring Rate (rpm)Lead Loading (mg·g−1)
25031
50037
75038
87530
100028
Temperature: 288 K. Time: 5 h.
Table 4. Influence of aqueous solution pH on Pb2+ uptake.
Table 4. Influence of aqueous solution pH on Pb2+ uptake.
pH of Aqueous SolutionPb2+ Load a (mg·g−1)Pb2+ Load b (mg·g−1)
43820
345
20.8<0.5
a AC concentration: 0.25 g·L−1. b AC concentration: 0.5 g·L−1. Temperature: 288 K. Time: 5 h. Stirring rate: 750 rpm.
Table 5. Influence of AC concentration on percentage Pb2+ removed from the aqueous solution.
Table 5. Influence of AC concentration on percentage Pb2+ removed from the aqueous solution.
AC Concentration (g·L−1)Pb2+ Removed (%)
0.0528
0.0950
0.1371
0.2596
0.5099
Temperature: 288 K. Time: 3 h. Stirring rate: 750 rpm.
Table 6. Parameter values for the different linear models.
Table 6. Parameter values for the different linear models.
FreundlichLangmuirTemkin
KF (L·g−1)1/nqm (mg·g−1)b (L·mg−1)RLATbT
42.040.2858.314.750.0277.84235.01
Table 7. Parameter values for the different linear models.
Table 7. Parameter values for the different linear models.
qm, exp (mol/g)T (K)pH ValueMaterial TypeReference
22.82985AC from apricot stone[43]
253135Sphagnum moss peat[44]
37.93035AC from Bois carre seeds[45]
47.23035Commercial activated carbon[46]
76.42985.5AC from Coconut shell[47]
312936Peanuts husks carbon[48]
Table 8. Influence of the ionic of the strength aqueous solution on Pb2+ adsorption onto the AC.
Table 8. Influence of the ionic of the strength aqueous solution on Pb2+ adsorption onto the AC.
Ionic Strength (M)Pb2+ Uptake (mg·g−1)
<0.0138
0.137
0.2531
0.5023
Aqueous solution at pH 4. Temperature: 288 K. Time: 3 h. Stirring rate: 750 rpm.
Table 9. Pb2+ uptake by different adsorbents/ion exchangers.
Table 9. Pb2+ uptake by different adsorbents/ion exchangers.
Adsorbent-ExchangerActive GroupLead Load (mg·g−1)
Activated carbon-38
Lewatit EP-63-<2
MWCNT-<1
Lewatit TP208weakly acidic, Na+ form34
Lewatit K-2621strongly acidic, H+ form36
Temperature: 288 K. Time: 3 h. Stirring rate: 750 rpm. - Without group active.
Table 10. Metals uptake in the active carbon.
Table 10. Metals uptake in the active carbon.
MetalMetal Uptake (mg·g−1)
Pb(II)38
Ni(II)12
Mn(II)8
Co(II)9
Cu(II)12
Cr(III)11
Cd(II)10
AC concentration: 0.25 g·L−1. Temperature: 288 K. Time: 3 h. Stirring rate: 750 rpm.
Table 11. Pb2+ concentration in the eluent solution at various eluent volume/AC concentration ratios.
Table 11. Pb2+ concentration in the eluent solution at various eluent volume/AC concentration ratios.
Volume Eluent/AC Concentration (mL·g−1)Pb2+ in the Solution (mg·L−1)
200130
50052
100026
200013
40006.5
Time: 3 h.

Share and Cite

MDPI and ACS Style

Alguacil, F.J.; Alcaraz, L.; García-Díaz, I.; López, F.A. Removal of Pb2+ in Wastewater via Adsorption onto an Activated Carbon Produced from Winemaking Waste. Metals 2018, 8, 697. https://doi.org/10.3390/met8090697

AMA Style

Alguacil FJ, Alcaraz L, García-Díaz I, López FA. Removal of Pb2+ in Wastewater via Adsorption onto an Activated Carbon Produced from Winemaking Waste. Metals. 2018; 8(9):697. https://doi.org/10.3390/met8090697

Chicago/Turabian Style

Alguacil, Francisco José, Lorena Alcaraz, Irene García-Díaz, and Félix Antonio López. 2018. "Removal of Pb2+ in Wastewater via Adsorption onto an Activated Carbon Produced from Winemaking Waste" Metals 8, no. 9: 697. https://doi.org/10.3390/met8090697

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop