Next Article in Journal
Nomilin and Its Analogues in Citrus Fruits: A Review of Its Health Promotion Effects and Potential Application in Medicine
Next Article in Special Issue
Removal of Methylene Blue from Water Using Magnetic GTL-Derived Biosolids: Study of Adsorption Isotherms and Kinetic Models
Previous Article in Journal
Green Synthesized Zinc Oxide Nanoparticles Based on Cestrum diurnum L. of Potential Antiviral Activity against Human Corona 229-E Virus
Previous Article in Special Issue
Possibility of New Active Substrates (ASs) to Be Used to Prevent the Migration of Heavy Metals to the Soil and Water Environments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Lead Pb(II) Removal with Chemically Modified Nostoc commune Biomass

by
Carmencita Lavado-Meza
1,*,
Leonel De la Cruz-Cerrón
1,
Carmen Lavado-Puente
2,
Julio Angeles-Suazo
3 and
Juan Z. Dávalos-Prado
4
1
Facultad de Ingeniería, Universidad Continental, Huancayo 12003, Peru
2
Escuela Profesional de Ingeniería Ambiental, Universidad Nacional Intercultural de la Selva Central, Chanchamayo 12856, Peru
3
Facultad de Ingeniería Industrial, Universidad Tecnológica del Perú, Lima 15046, Peru
4
Instituto de Química Física “Rocasolano”, CSIC, 28006 Madrid, Spain
*
Author to whom correspondence should be addressed.
Submission received: 3 December 2022 / Revised: 23 December 2022 / Accepted: 23 December 2022 / Published: 28 December 2022
(This article belongs to the Special Issue Innovative Adsorbents for Water Treatment)

Abstract

:
A new biosorbent based on Nostoc commune (NC) cyanobacteria, chemically modified with NaOH (NCM), has been prepared, characterized and tested as an effective biomass to remove Pb(II) in aqueous media. The adsorption capacity of NCM was determined to be qe = 384.6 mg g−1. It is higher than several other biosorbents reported in the literature. Structural and morphological characterization were performed by FTIR, SEM/EDX and point zero of charge pH (pHPZC) measurements. NCM biosorbent showed more porous surfaces than those NC with heterogeneous plates including functional adsorption groups such as OH, C = O, COO, COH or NH. Optimal Pb(II) adsorption occurred at pH 4.5 and 5.5 with a biomass dose of 0.5 g L−1. The experimental data of the adsorption process were well fitted with the Freundlich-isotherm model and pseudo-2nd order kinetics, which indicated that Pb(II) adsorption was a chemisorption process on heterogeneous surfaces of NCM. According to the thermodynamic parameters, this process was exothermic (∆H0 < 0), feasible and spontaneous (∆G0 < 0). NCM can be regenerated and efficiently reused up to 4 times (%D > 92%). NCM was also tested to remove Pb (%R~98%) and Ca (%R~64%) from real wastewater.

1. Introduction

Nowadays, worldwide, there has been an increase in the demand for drinking water. However, there is general neglect regarding the control and prevention of various forms of contamination that affect the potential sources of water supply and storage. One of the contaminants of concern are heavy metals, due to their toxic, persistent and bioaccumulative nature, which seriously affect human health and the environment [1]. This is the case of Pb(II), which accumulates in the bones (half-life of 20 years), affects the nervous [2] and reproductive systems, red blood cells and kidneys [3], in addition to being a potential carcinogen, given its high affinity for the thio (–SH), oxo (–O) and phosphate (PO43−) groups of some enzymes, ligands and biomolecules.
The main removal techniques for heavy metals include methods such as coagulation, chemical precipitation, electrodialysis, evaporative recovery, flotation, flocculation, ion exchange, nanofiltration, reverse osmosis, ultrafiltration, etc. [4]. Although these methods are effective, they are not economical due to the high costs in energy and production reagents. These can also generate considerable amounts of sludge and toxic by-products. Therefore, it is important to develop efficient, economical and, above all, environmentally friendly methods, capable of effectively removing toxic products [5,6,7]. Within this category are the so-called biosorbents, which by biosorption processes eliminate metals and other toxic products from an aqueous solution. Biosorption includes absorption, adsorption, ion exchange, surface complexation and precipitation processes.
Biosorbents can be microbial, such as bacteria (e.g., cyanobacteria), fungi, yeasts, algae [8,9] and also plant materials and agro-industrial waste [10], that due to the presence on their surfaces of functional groups, such as hydroxyls, carbonyls, phosphates, amines or thiols, can efficiently remove a wide variety of metallic ions [11].
Cyanobacteria are microorganisms that inhabit a wide variety of environments, perform photosynthesis and form a thick microbial layer even in extreme habitat settings [12]. Used as biosorbents, they have interesting properties, such as a large surface area and a larger volume of mucilage with high binding affinity and simple nutrient requirements [13]. These microorganisms can produce considerable amounts of extracellular polymeric compounds including proteins, exopolysaccharides, uronic acids and humic and lipid compounds [14]. Among these components, extracellular exopolysaccharides stand out as powerful chelating agents capable of trapping metallic ions in solution [5,15].
Nostoc commune (NC), or blue-green algae, is a cyanobacterium that has gelatinous matter on its surface, contain little protein but a lot of dietary fiber with minerals such as Ca or Fe [16] and reproduces in large numbers on ponds, ditches, waterlogged places and nutrient-poor limestone soils [17]. NC is widely distributed throughout the world, from polar to tropical regions, making it an abundant and cheap biomass. Previous studies revealed the effectiveness of NC for the elimination of several contaminants in wastewater [18]. Proper treatment prevents its leaching and increases the concentration of active sites, which leads to an improvement in its metallic biosorption capacity. Treatments with acids (HNO3, HCl, H3PO4, citric acid, etc.) and bases (NaOH, KOH) have been reported, the latter being the ones that gave the best biosorption results [19,20,21]. In this work we report the removal capacity of Pb(II), in aqueous solutions, of the Nostoc commune biomass, chemically modified with NaOH and coming from Mantaro valley of the Junin Region (Peru).

2. Results and Discussion

2.1. Effect of Alkaline Treatment, Concentration of Acidic and Basic Sites, pHPZC Determination

Table 1 shows the concentration of acidic and basic sites on the NC (untreated) and NCM (treated) biomasses. The alkaline treatment of NC is reflected in a significant increase in the concentration of basic sites (almost six times) and a decrease in acid sites on the NCM sample. According to Bulgariu et al. [22,23], the increase in basic sites in biomasses—such as Ulva lactuca algae—treated with NaOH would be due to hydrolysis reactions that would increase, on the cell surfaces, the formation of carboxylic (–COO) and hydroxyl (–OH) groups. The presence of these groups on NCM has been confirmed by our FTIR analyses.
The point zero of charge pH values (pHPZC) were deduced from the ΔpH (=pH0 − pHf) vs. pH0 plot (See Figure 1). The two curves belonging to the NC and NCM intersect the pH0–axis at pHPZC = 1.3 and 2.5, respectively. It is important to mention that pHPZC provides valuable information on the electrostatic interactions between sorbents and sorbed species, thus for pH < pHPZC, the sorbent–surface charge is positive, therefore it can interact with anionic sorbates. On the contrary, at pH > pHPZC the sorbent–surface charge is negative and it can interact more with cationic sorbates. As can be seen in Figure 2, the pHPZC for NCM (2.5) is higher than for NC (1.3). This result indicates an increase in the superficial alkalinity of NCM compared to NC, which is consistent with the increase in the concentration of its basic sites, described above. In this regard, Šoštarić et al. [11] reported similar results, with the alkaline treatment of the apricot shell biomass.
An interesting consequence of the alkaline treatment of NC that increases adsorption basic sites and pHPZC value is the significant increase in its Pb(II) removal capacity, qe. Taking into account optimal conditions, described below, the qe of NCM is 1.6 times higher than for untreated biomass NC (See Figure 2).

2.2. SEM/EDX Morphological and Structural Characterization, FTIR Analysis

Figure 3 shows the morphologies of NC, NCM and the last one loaded with Pb(II) (NCM-Pb). NCM shows a more porous and cracked surface than NC. According to Zafar et al. [24], the treatment with NaOH would eliminate the protein and lipid fractions of the biomass, generating a greater contact area and thus increasing the adsorption capacity of Pb(II) from the treated surface. Iddou et al. [19] reported similar results, showing more porous surfaces on brown algae treated with NaOH, with respect to the precursor biomass. Pb(II) adsorption changes the morphology of NCM, making the surface less rough and porous (Figure 3c).
Figure 3 (right) shows the EDX spectra of NC, NCM and NMC-Pb. We can appreciate the presence of common peaks associated with C, O, N, Al and Si. NC and NCM also show the presence of Mg, and only in NCM is Na present. In NCM-Pb, the presence of Pb peaks is notable, as is the absence of Na and Mg peaks. El-Naggar et al. [25] reported similar results after loading with Pb(II) the biomass of the marine algae Gelidium amansii. The disappearance of the Mg and Na peaks on the surface of NCM-Pb (Figure 4c) would be due to the interaction of ion exchange and/or complexation of Pb(II) with monovalent (H+, Na+) or divalent (Mg2+) cations of the biomass, which would be released by Pb(II) [3,26].

FTIR Analysis

FTIR spectra (Figure 4 left) of the NCM show band positions at:
1/3427.3 cm−1 associated with the OH groups of polymeric compounds (such as sucrose) and the NH groups of proteins [27].
2/2921 cm−1 attributable to the symmetric stretching of the aliphatic chains of C–H bonds of alkyl functional groups in carbohydrates [27].
3/1639.4 cm−1 associated with the stretching of carbonyl group bonds (C=O), primary and secondary amides of protein peptide bonds [9].
4/1537.2 cm−1 associated with carboxylates (COO) [15,28].
5/1417.6 and 1035.5 cm−1 would be due to the stretching of C–H and C–OH bonds respectively in pyranose units [29,30].
FTIR spectra of NCM loaded with Pb(II), NCM-Pb (Figure 4 right) show changes, with respect to NCM, in the intensity and position of some adsorption peaks. Thus, the positions of the 3/, 4/ and 6/ bands are significantly displaced at Δ3 = 4.0, Δ4 = 9.6, Δ6CH = 3.8 and Δ6COH = −9.4 cm−1. The displacement of these bands proves the participation of carbonyl, amide and carboxylic groups in the Pb(II) adsorption process [31].

2.3. Influence of pH Solution

pH is an important factor in the biosorption of heavy metals, given that the speciation and removal of heavy metals in aqueous solution is highly dependent on this factor. According to Joseph et al. [32], there is a consensus that low pH values (<4) hinder the adsorption of heavy metals, while pH values between 5 and 7 are the most effective. In this context, we have studied the effect of pH on the Pb(II) absorption capacity qe, at different C0 (initial concentration of Pb ions) values. The pH range considered was 2.5 to only 5.5 because Pb precipitates at higher values, forming basically Pb(OH)2 (See Figure 5, speciation of Pb). We can observe that the trend of each qe vs. pH (Figure 5) curve is similar for all considered C0 values, that is, qe increases with increasing pH until reaching a plateau between pH 4.5 and 5.5. It is interesting to note that, even at low pHs, the qe values for NCM are significant. This result would be due to NCM having a pHPZC = 2.5, which indicates the basic nature of its surface, with numerous active sites, such as –OH, –COOH, –COH and –NH functional groups capable of favoring the adsorption of Pb(II). The increments of biosorption until pH 5.5 could be associated also with the deprotonation of carboxyl, hydroxyl and other negatively charged groups, leading to the electrostatic attraction of positively charged Pb(II) [4,33].

2.4. Influence of NCM Dose and Initial Concentration of Pb(II) Ions, C0

Figure 6 shows the influence of the biomass dose of NCM on the adsorption capacity qe and on the %R removal percentage of Pb(II). For each initial concentration C0, the same trend is observed in both qe vs. dose (descending) and in %R vs. dose (ascending) plots, that is: i/significant increase in %R up to an NCM dose of 0.5 g L−1, from which this increase is less pronounced, particularly for the lowest C0 (127.3 mg L−1) where %R reaches almost 97%. On the contrary, for the highest C0 (311.1 mg·L−1), %R reaches almost 64%, which would be related to the decrease and saturation of active adsorbent sites due to the agglomeration of biomass particles [34]. ii/For a given C0, qe decreases with increasing NCM dose. This behavior was also reported by Mangwandi et al. [35], in the removal of Cr(IV) with date pits and olive stones.
Taking in account these results, we consider 0.5 g·L−1 the optimal NCM dose for Pb(II) removal.

2.5. Kinetic of Biosorption

We can note (Table 2) a better correlation (R2 = 1) with pseudo-second order than with the first-order adjustment models, although this last one, according to Vijayaraghavan et al. [36], can be applied during the initial periods of the biosorption process. Accordingly, we can affirm that i/the adsorption of Pb(II) is a chemisorption process wherein the sorption capacity is related to the number of occupied active sites [37]; ii/the calculated biosorption capacity at time t, qt, is very close to the corresponding experimental value; iii/the calculated biosorption capacity at equilibrium, qe,cal = 384.6 mg g−1, is the value that we consider the maximum qe value (qe,max) for the Pb(II) removal capacity of NCM.
The pseudo-second order kinetic for NCM is the same model that governs the Pb(II) adsorption kinetics of several biosorbents, such as taro [38], olive pit [4], Geldium amansii [25] or shells of peanut [39].
There is an acceptable correlation (R2 = 0.8) for the Elovich kinetic model, which assumes that the active sites of the sorbent are heterogeneous and therefore exhibit different sorption energies. The α rate constant value is comparable to that obtained by Elwakeel et al. [40], which can be attributed to the negative surface-charge nature of the sorbent that allows the rapid sorption of Pb(II).
The qt vs. t0.5 plot (Figure 7) was fitted to the intra-particle diffusion Weber–Morris model. We can distinguish three parts: the first part, showing a rapid growth of qt at the time t (0 < t1/2 < 3) which would indicate the rapid absorption of Pb(II) ions on the outer surface of the biosorbent; the second part, a slower growth of qt with t (3 < t1/2 < 5.5), which would be related to a gradual adsorption process, wherein Pb ions would enter and fill the biosorbent pores and the intraparticle diffusion would be the rate-limiting step [41,42]; finally, the third part (t1/2 > 5.5), wherein qt is practically constant (plateau region), would indicate the equilibrium uptake, wherein the intra particle diffusion is not the only rate-controlling step [34].

2.6. Adsorption Isotherms

Data obtained from Pb(II) biosorption experiments were fitted to Langmuir, Freundlich and Dubinin–Radushkevich (D–R) equilibrium models. The isotherms were determined for optimal conditions at pH 4.5 and 5.5, room temperature, NMC dose 0.5 g L−1, contact time t = 60 min and a range of C0 concentrations between 127 and 442 mg L−1. Figure 8 shows adsorption isotherms fitted to Langmuir and Freundlich models.
The determined parameters are consigned in Table 3. We can see that the experimental data of the adsorption isotherms fit well to the Langmuir model, particularly at pH 4.5, with parameters qmax (maximum adsorption capacity) and KL practically similar to those obtained at pH 5.5. It is interesting to mention that the value of qmax = 384.6 mg g−1 is the same as the qe,max calculated by the pseudo-second order kinetic.
Nevertheless, a best fit, at pH 4.5, is obtained with the Freundlich model, given the lowest χ2 value with R2 close to 1. The results indicate that the biosorption of Pb(II) on NCM is a chemisorption process (nF >1) with a high affinity between Pb ions and the heterogeneous surfaces of NCM [43].
The fits of the biosorption data to the Dubinin–Radushkevich (D–R) model are acceptable. The values of qmax obtained are close to those determined with the Langmuir model (deviation between 15 and 20%), and the values of the parameter E (>500 kJ mol−1) confirm that the biosorption of Pb(II) on NCM is a chemisorption process [41].
The fits to the Temkin model give low χ2 values with R2 > 0.9. The parameter B (>75 J mol−1) values, related to the heat of biosorption [44,45], reinforce the chemical–biosorption mechanisms already indicated by the Freundlich and Dubinin–Radushkevich models.
The maximum Pb(II) biosorption capacity qe,max values of biomasses and agro-industrial wastes similar to that studied in this work are shown in the Table 4. It is interesting to note that the qe,max of NCM is among the highest.

2.7. Biosorption Thermodynamics

For the thermodynamic study, the Van’t Hoff equation (1) was used, which is derived from the equation Δ G o = Δ H o T Δ S o that relates the Gibbs energy (ΔG0), enthalpy (ΔH0) and entropy (ΔS0) of the adsorption process. ΔG0, at different temperatures, were evaluated from the expression Δ G 0 = R T ln K c where Kc is the equilibrium constant, and K c = C es C e ; Ces and Ce are the equilibrium Pb(II) concentrations, respectively, in the biosorbent and in the solution. R is the universal gas constant, and T is the temperature of the solution.
ln   K c = Δ S 0 R Δ H 0 R T
The parameters obtained are consigned in Table 5. The negative values of ΔG0 indicate that the Pb(II) adsorption process on NCM is spontaneous and favorable for all the considered temperatures (see Table 5). The negative value of the entropy change ΔS0 proves the decrease in randomness at the solid/solution interface during biosorption. On the other hand, the negative value of ΔH0 indicates that the process studied is exothermic, and its magnitude (65.1 kJ mol−1), according to Cardoso et al. [49], confirms the chemical nature of the Pb(II) adsorption process (chemisorption). Similar results were found in the removal of Pb(II) with Cystoseira stricta marine algae [19] and also with raw/treated maize stover [46].

2.8. NCM for the Pb(II) Removal in Real-Wastewater

For the possible application of NCM in the treatment of contaminated water, we consider real wastewater that contains, in addition to Pb, other metals such as Ca, K and Na, whose concentrations are shown in Table 6. For the removal of Pb(II), it has been considered optimal biosorption conditions, which were determined in this work: dose of NCM = 0.5 g L−1 and pH 5.5.
We can appreciate significant percentages of removal (%R) of Pb(II) (~98%) and Ca (~64%), to a lesser extent of K (10%) and to a practically null amount of Na (Table 6).
The results obtained show that NCM is a cheap, abundant and very effective biomass to remove Pb(II) and is also effective to remove Ca from real wastewater with high concentrations of these metals.

2.9. Regeneration of NCM Biosorbent

The adsorption/desorption studies are related to the economic and environmental viability of the sorbent material, since good stability (effective regeneration) and multiple re-uses of it are desirable [50]. We can see in Figure 9, that after four adsorption/desorption cycles, the regeneration efficiency, %D, of NCM is greater than 92%. This result shows that NCM is a stable, reusable and efficient biosorbent for the aqueous removal of Pb(II). In this regard, Bangaraiah et al. [50] removed Pb(II) with chemically modified green algae, reusing it up to three times with %D = 87%.

3. Materials and Methods

3.1. Preparation of Nostoc Commune Biosorbent

  • Untreated biomass (NC)
Untreated biomass was obtained from gelatinous colonies of Nostoc commune (Figure 10) which were collected at Pampa Cruz in the Junín Region of Peru. The material was washed, then dried in an oven at 333 K for 48 h and finally ground.
  • Treated biomass (NCM)
NC was placed in contact with 0.1 M NaOH solution in a ratio of 1 g:10 mL for 24 h with constant stirring at 300 rpm. Afterwards, the mixture was filtered and washed with abundant deionized water until reaching a colorless solution and neutral pH. Finally, the sample was dried at 333 K for 48 h, then it was ground and sieved with a 0.3 mm mesh.

3.2. Biosorbent Characterization

  • The point zero of charge pH values (pHPZC) were determined according to the procedures described by do Nascimento et al. [20]. It has been prepared as a mixture of 0.05 g of biosorbent with 50 mL of aqueous solutions under different initial pHs (pH0) ranging from 1 to 8. The acid solutions were prepared from 1 M HCl, while the basic solutions were prepared from 1 M NaOH. After 24 h of equilibrium, the final pHs (pHf) were determined.
  • A Fourier transform infrared spectrophotometer (FTIR, SHIMADZU- 8700) was used to identify the functional groups present on the surface of biosorbents. The wavelength was set to 4000 to 400 cm−1.
  • Morphological and elemental analysis of the biosorbent surface were performed by scanning electron microscopy (SEM) coupled with EDX (energy-dispersive X-ray spectroscopy) (Hitachi SU8230 model).
  • The concentration of acid and basic sites was determined by the Boehm method following the procedures described by do Nascimento et al. [51].
All measurements were performed in triplicate.

3.3. Biosorption Assays

Experimental parameters such as pH, biosorbent dosage, initial and final or equilibrium Pb(II) concentrations and those related to kinetic and isothermal properties of Pb(II) biosorption on NC and NCM samples have been determined. Between 0.0125 and 0.5 g of each biomass was added to 50 mL of Pb(NO3)2 solution with a Pb(II) initial concentration between 127.3 to 440.4 mg L−1. These solutions had been adjusted to a pH in the range of 2.5 to 5.5 by adding 0.1 M HNO3 or 0.1 M NaOH. The suspension obtained was stirred at 300 rpm for a time period of 60 min. The temperature was kept at 293 K (room temperature). The Pb(II) adsorption capacity, qe (in mg·g−1), onto biosorbents was determined by Equation (2) measuring the concentration of Pb before and after the biosorption process by means an atomic absorption spectrophotometer (SHIMADZU-AAS 6800). The Pb(II) removal efficiency (%R) was calculated by using Equation (3)
q e = C 0 C e m × V
% R = C 0 C e C 0 × 100
where C0 and Ce (in mg L−1) are the initial and equilibrium or final Pb(II) concentrations, respectively; V(in L) is the volume of solution and m (in g) is the biosorbent mass. The adsorption experiments were repeated three times, and the average values were reported.
The experimental data of isotherms and biosorption kinetics on NC and NCM were evaluated and correlated with different models, described in Table 7. The quality of the corresponding adjustments is reflected by chi-squared χ2 and/or correlation coefficient R2 values [52].

3.4. Desorption Experiments

A total of 50 mg of NCM biosorbent previously loaded with Pb(II), from the mixture with 100 mL of Pb solution ([Pb(II)] = 127.3 mg L−1), was then then filtered, washed and dried, then was subjected to the desorption process by adding 50 mL of 0.1 M HNO3 eluent and then stirring the result at 200 rpm for 60 min. After that, the biosorbent was washed with distilled water, dried and re-used again. The adsorption/desorption operation was repeated up to four times. The [Pb(II)] concentration adsorbed and desorbed was analyzed by an atomic absorption spectrophotometer (before being described).
The desorption or regeneration efficiency (%D) of the NCM biomass was calculated using the following expression [53]:
% D = P b I I d e s o r b e d   f r o m   b i o m a s s P b I I a d s o r b e d   o n   b i o m a s s   × 100

4. Conclusions

  • Nostoc commune cyanobacteria was chemically modified, with a 0.1 M NaOH. The Pb(II) adsorption capacity of the treated biomass (NCM) was almost 1.6 times higher than for untreated biomass (NC).
  • Point zero of charge, pHPZC, of NCM (= 2.5) was greater than for NC (=1.3). It is consistent with the concentration of basic sites, which is almost six times higher for treated than untreated biosorbents. The basic sites would be associated with OH, C=O, COH, COO and NH functional groups (identified by FTIR).
  • SEM/EDX analyses of NCM showed a more porous and cracked surface than NC, but once charged with Pb(II), the morphology changed to a less rough and porous surface.
  • For a given initial Pb ions concentration, C0, the biosorption capacity qe of NCM reached a maximum plateau at a pH between pH 4.5 and 5.5.
  • We consider 0.5 g L−1 the optimal NCM dose for Pb(II) removal, given that the corresponding efficiency, %R, can reach almost 97% for low C0 values.
  • The adsorption kinetic data were well fitted with the pseudo-second order model, indicating that the Pb(II) biosorption on NCM was a chemisorption process, with a removal capacity of qe = 384.6 mg g−1. The Elovich kinetic model indicated a rapid sorption of Pb(II). It is consistent for the first stage described by the intra-particle diffusion Weber–Morris model, since, in the following stages, Pb(II) adsorption was very slow.
  • The adsorption isotherms’ data were well fitted with the Freundlich model (heterogeneous adsorption) at pH 4.5 and 5.5. Freundlich, Dubinin–Radushkevich and Temkin models confirmed that Pb(II) adsorption on NCM is a chemisorption process, which was thermodynamically characterized as exothermic (ΔH0 < 0), feasible and spontaneous (ΔG0 < 0) and with a decreasing randomness (ΔS0 < 0) at the solid/liquid interface.
  • The maximum Pb(II) biosorption capacity of NCM, qe,max= 384.6 mg g−1, is higher than for other similar treated biosorbents reported in the literature.
  • Desorption–regeneration experiments showed that NCM can be recovered efficiently (%D > 92%) up to four times.
  • NCM was tested as an inexpensive and efficient biosorbent to remove Pb and Ca, from real wastewater, with an efficiency %R of almost 98% and 64%, respectively.

Author Contributions

C.L.-M. and L.D.l.C.-C.: experiments, data analysis, writing—original draft., J.A.-S. and C.L.-P.: experimental characterization, paper review. J.Z.D.-P.: data analysis, writing final version. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Lee, J.W.; Choi, H.; Hwang, U.K.; Kang, J.C.; Kang, Y.J.; Kim, K., II; Kim, J.H. Toxic effects of lead exposure on bioaccumulation, oxidative stress, neurotoxicity, and immune responses in fish: A review. Environ. Toxicol. Pharmacol. 2019, 68, 101–108. [Google Scholar] [CrossRef] [PubMed]
  2. Lentini, P.; Zanoli, L.; de Cal, M.; Granata, A.; Dell’Aquila, R. Lead and Heavy Metals and the Kidney. In Critical Care Nephrology, 3rd ed.; Elsevier: Rome, Italy, 2019; Chapter 222; pp. 1324–1330. [Google Scholar]
  3. Morosanu, I.; Teodosiu, C.; Paduraru, C.; Ibanescu, D.; Tofan, L. Biosorption of lead ions from aqueous effluents by rapeseed biomass. New Biotechnol. 2017, 39, 110–112. [Google Scholar] [CrossRef] [PubMed]
  4. Blázquez, G.; Calero, M.; Ronda, A.; Tenorio, G.; Martín-Lara, M.A. Study of kinetics in the biosorption of lead onto native and chemically treated olive stone. J. Ind. Eng. Chem. 2014, 20, 2754–2760. [Google Scholar] [CrossRef]
  5. Gupta, P.; Diwan, B. Bacterial Exopolysaccharide mediated heavy metal removal: A Review on biosynthesis, mechanism and remediation strategies. Biotechnol. Rep. 2017, 13, 58–71. [Google Scholar] [CrossRef] [PubMed]
  6. Al-Marghany, A.; Badjah Hadj Ahmed, A.Y.; AlOthman, Z.A.; Sheikh, M.; Ghfar, A.A.; Habila, M. Fabrication of Schiff’s base-functionalized porous carbon materials for the effective removal of toxic metals from wastewater. Arab. J. Geosci. 2021, 14, 1–13. [Google Scholar] [CrossRef]
  7. Alothman, Z.A.; Habila, M.A.; Moshab, M.S.; Al-Qahtani, K.M.; AlMasoud, N.; Al-Senani, G.M.; Al-Kadhi, N.S. Fabrication of renewable palm-pruning leaves based nano-composite for remediation of heavy metals pollution. Arab. J. Chem. 2020, 13, 4936–4944. [Google Scholar] [CrossRef]
  8. Gupta, V.K.; Rastogi, A. Sorption and desorption studies of chromium(VI) from nonviable cyanobacterium Nostoc muscorum biomass. J. Hazard. Mater. 2008, 154, 347–354. [Google Scholar] [CrossRef]
  9. Gupta, V.K.; Rastogi, A. Biosorption of lead (II) from aqueous solutions by non-living algal biomass Oedogonium sp. and Nostoc sp.−A comparative study. Colloids Surf. B Biointerfaces 2008, 64, 170–178. [Google Scholar] [CrossRef] [PubMed]
  10. Fomina, M.; Gadd, G.M. Biosorption: Current perspectives on concept, definition and application. Bioresour. Technol. 2014, 160, 3–14. [Google Scholar] [CrossRef]
  11. Šoštarić, T.D.; Petrović, M.S.; Pastor, F.T.; Lončarević, D.R.; Petrović, J.T.; Milojković, J.V.; Stojanović, M.D. Study of heavy metals biosorption on native and alkali-treated apricot shells and its application in wastewater treatment. J. Mol. Liq. 2018, 259, 340–349. [Google Scholar] [CrossRef]
  12. Patel, A.; Matsakas, L.; Rova, U.; Christakopoulos, P. Bioresource Technology A perspective on biotechnological applications of thermophilic microalgae and cyanobacteria. Bioresour. Technol. 2019, 278, 424–434. [Google Scholar] [CrossRef]
  13. Kulal, D.K.; Loni, P.C.; Dcosta, C.; Some, S.; Kalambate, P.K. Cyanobacteria: As a promising candidate for heavy-metals removal. In Advances in Cyanobacterial Biology; Academic Press: Cambridge, MA, USA, 2020; Chapter 19; pp. 291–300. [Google Scholar]
  14. Bhunia, B.; Shankar, U.; Uday, P.; Oinam, G.; Mondal, A. Characterization, genetic regulation and production of cyanobacterial exopolysaccharides and its applicability for heavy metal removal. Carbohydr. Polym. 2018, 179, 228–243. [Google Scholar] [CrossRef] [PubMed]
  15. Tran, H.T.; Vu, N.D.; Matsukawa, M.; Okajima, M.; Kaneko, T.; Ohki, K.; Yoshikawa, S. Heavy metal biosorption from aqueous solutions by algae inhabiting rice paddies in Vietnam. J. Environ. Chem. Eng. 2016, 2, 2529–2535. [Google Scholar] [CrossRef]
  16. Liang, Y.; Shu, X.; Wang, W. Biochemical composition, heavy metal content and their geographic variations of the form species Nostoc commune across China. Food Sci. Technol. 2022, 42, 1–8. [Google Scholar] [CrossRef]
  17. Dodds, W.K.; Gudder, D.A.; Mollenhauer, D. The ecology of Nostoc. J. Phycol. 1995, 31, 2–18. [Google Scholar] [CrossRef]
  18. Atoku, D.I.; Ojekunle, O.Z.; Taiwo, A.M.; Shittu, O.B. Evaluating the efficiency of Nostoc commune, Oscillatoria limosa and Chlorella vulgaris in a phycoremediation of heavy metals contaminated industrial wastewater. Sci. Afr. 2021, 12, e00817. [Google Scholar] [CrossRef]
  19. Iddou, A.; Hadj Youcef, M.; Aziz, A.; Ouali, M.S. Biosorptive removal of lead (II) ions from aqueous solutions using Cystoseira stricta biomass: Study of the surface modification effect. J. Saudi Chem. Soc. 2011, 15, 83–88. [Google Scholar] [CrossRef] [Green Version]
  20. Ronda, A.; Martín-Lara, M.A.; Almendros, A.I.; Pérez, A.; Blázquez, G. Comparison of two models for the biosorption of Pb(II) using untreated and chemically treated olive stone: Experimental design methodology and adaptive neural fuzzy inference system (ANFIS). J. Taiwan Inst. Chem. Eng. 2015, 54, 45–56. [Google Scholar] [CrossRef]
  21. Santos, S.C.R.; Ungureanu, G.; Volf, I.; Boaventura, R.A.R.; Botelho, C.M.S. Macroalgae Biomass as Sorbent for Metal Ions. In Biomass as Renewable Raw Material to Obtain Bioproducts of High-Tech Value; Elsevier: Amsterdam, The Netherlands, 2018; pp. 69–112. [Google Scholar]
  22. Bulgariu, L.; Bulgariu, D.; Macoveanu, M. Adsorptive Performances of Alkaline Treated Peat for Heavy Metal Removal. Sep. Sci. Technol. 2011, 46, 1023–1033. [Google Scholar] [CrossRef]
  23. Bulgariu, L.; Bulgariu, D. Enhancing Biosorption Characteristics of Marine Green Algae (Ulva lactuca) for Heavy Metals Removal by Alkaline Treatment. J. Bioprocess. Biotech. 2014, 4, 1–8. [Google Scholar] [CrossRef]
  24. Zafar, M.N.; Aslam, I.; Nadeem, R.; Munir, S.; Rana, U.A.; Khan, S.U.D. Characterization of chemically modified biosorbents from rice bran for biosorption of Ni(II). J. Taiwan Inst. Chem. Eng. 2015, 46, 82–88. [Google Scholar] [CrossRef]
  25. El-Naggar, N.E.A.; Hamouda, R.A.; Mousa, I.E.; Abdel-Hamid, M.S.; Rabei, N.H. Biosorption optimization, characterization, immobilization and application of Gelidium amansii biomass for complete Pb2+ removal from aqueous solutions. Sci. Rep. 2018, 8, 13456. [Google Scholar] [CrossRef] [Green Version]
  26. Petrović, M.; Šoštarić, T.; Stojanović, M.; Petrović, J.; Mihajlović, M.; Ćosović, A.; Stanković, S. Mechanism of adsorption of Cu2+ and Zn2+ on the corn silk (Zea mays L.). Ecol. Eng. 2017, 99, 83–90. [Google Scholar] [CrossRef]
  27. Calero, M.; Pérez, A.; Blázquez, G.; Ronda, A.; Martín-lara, M.A. Characterization of chemically modified biosorbents from olive tree pruning for the biosorption of lead. Ecol. Eng. 2013, 58, 344–354. [Google Scholar] [CrossRef]
  28. Morsy, F.M.; Hassan, S.H.A.; Koutb, M. Biosorption of Cd (II) and Zn (II) by Nostoc commune: Isotherm and Kinetics Studies. Clean Soil Air Water 2011, 39, 680–687. [Google Scholar] [CrossRef]
  29. Ohki, K.; Le, N.Q.T.; Yoshikawa, S.; Kanesaki, Y.; Okajima, M.; Kaneko, T.; Thi, T.H. Exopolysaccharide production by a unicellular freshwater cyanobacterium Cyanothece sp. isolated from a rice field in Vietnam. J. Appl. Phycol. 2014, 26, 265–272. [Google Scholar] [CrossRef]
  30. Raj, K.; Sardar, U.R.; Bhargavi, E.; Devi, I.; Bhunia, B.; Tiwari, O.N. Advances in exopolysaccharides based bioremediation of heavy metals in soil and water: A critical review. Carbohydr. Polym. 2018, 199, 353–364. [Google Scholar]
  31. Warjri, S.M.; Syiem, M.B. Analysis of Biosorption Parameters, Equilibrium Isotherms and Thermodynamic Studies of Chromium (VI) Uptake by a Nostoc sp. Isolated from a Coal Mining Site in Meghalaya, India. Mine Water Environ. 2018, 37, 713–723. [Google Scholar] [CrossRef]
  32. Joseph, L.; Jun, B.M.; Flora, J.R.V.; Park, C.M.; Yoon, Y. Removal of heavy metals from water sources in the developing world using low-cost materials: A review. Chemosphere 2019, 229, 142–159. [Google Scholar] [CrossRef]
  33. Ibrahim, W.M. Biosorption of heavy metal ions from aqueous solution by red macroalgae. J Hazard Mater. 2011, 192, 1827–1835. [Google Scholar] [CrossRef]
  34. Reddy, D.H.K.; Harinath, Y.; Seshaiah, K.; Reddy, A.V.R. Biosorption of Pb(II) from aqueous solutions using chemically modified Moringa oleifera tree leaves. Chem. Eng. J. 2010, 162, 626–634. [Google Scholar] [CrossRef]
  35. Mangwandi, C.; Kurniawan, T.A.; Albadarin, A.B. Comparative biosorption of chromium (VI) using chemically modified date pits (CM-DP) and olive stone (CM-OS): Kinetics, isotherms and influence of co-existing ions. Chem. Eng. Res. Des. 2020, 56, 251–262. [Google Scholar] [CrossRef]
  36. Vijayaraghavan, J.; Pushpa, T.B.; Basha, S.J.S.; Jegan, J.; Pushpa, T.B.; Basha, S.J.S. Isotherm, kinetics and mechanistic studies of methylene blue biosorption onto red seaweed Gracilaria corticata. Desalination Water Treat. 2015, 57, 13540–13548. [Google Scholar] [CrossRef]
  37. Albadarin, A.B.; Solomon, S.; Daher, M.A.; Walker, G. Efficient removal of anionic and cationic dyes from aqueous systems using spent Yerba Mate “Ilex paraguariensis”. J. Taiwan Inst. 2018, 82, 144–155. [Google Scholar] [CrossRef]
  38. Saha, G.C.; Hoque, M.I.U.; Miah, M.A.M.; Holze, R.; Chowdhury, D.A.; Khandaker, S.; Chowdhury, S. Biosorptive removal of lead from aqueous solutions onto Taro (Colocasiaesculenta (L.) Schott) as a low cost bioadsorbent: Characterization, equilibria, kinetics and biosorption-mechanism studies. J. Environ. Chem. Eng. 2017, 5, 2151–2162. [Google Scholar] [CrossRef]
  39. Taşar, Ş.; Kaya, F.; Özer, A. Biosorption of lead(II) ions from aqueous solution by peanut shells: Equilibrium, thermodynamic and kinetic studies. J. Environ. Chem. Eng. 2014, 2, 1018–1026. [Google Scholar] [CrossRef]
  40. Elwakeel, K.Z.; Elgarahy, A.M.; Mohammad, S.H. Journal of Environmental Chemical Engineering Use of beach bivalve shells located at Port Said coast (Egypt) as a green approach for methylene blue removal. J. Environ. Chem. Eng. 2017, 5, 578–587. [Google Scholar] [CrossRef]
  41. Koyuncu, H.; Kul, A.R. Biosorption study for removal of methylene blue dye from aqueous solution using a novel activated carbon obtained from nonliving lichen (Pseudevernia furfuracea (L.) Zopf.). Surf. Interfaces 2020, 19, 100527. [Google Scholar] [CrossRef]
  42. Srivastava, S.; Agrawal, S.B.; Mondal, M.K. Biosorption isotherms and kinetics on removal of Cr(VI) using native and chemically modified Lagerstroemia speciosa bark. Ecol. Eng. 2015, 85, 56–66. [Google Scholar] [CrossRef]
  43. Elkhaleefa, A.; Ali, I.H.; Brima, E.I.; Shigidi, I.; Elhag, A.B.; Karama, B. Evaluation of the adsorption efficiency on the removal of lead(II) ions from aqueous solutions using Azadirachta indica leaves as an adsorbent. Processes 2021, 9, 559. [Google Scholar] [CrossRef]
  44. Lebron, Y.A.R.; Moreira, V.R.; Santos, L.V.S.; Jacob, R.S. Remediation of methylene blue from aqueous solution by Chlorella pyrenoidosa and Spirulina maxima biosorption: Equilibrium, kinetics, thermodynamics and optimization studies. J. Environ. Chem. Eng. 2018, 6, 6680–6690. [Google Scholar] [CrossRef]
  45. Araújo, C.S.T.; Almeida, I.L.S.; Rezende, H.C.; Marcionilio, S.M.L.O.; Léon, J.J.L.; de Matos, T.N. Elucidation of mechanism involved in adsorption of Pb(II) onto lobeira fruit (Solanum lycocarpum) using Langmuir, Freundlich and Temkin isotherms. Microchem. J. 2018, 137, 348–354. [Google Scholar] [CrossRef]
  46. Guyo, U.; Mhonyera, J.; Moyo, M. Pb(II) adsorption from aqueous solutions by raw and treated biomass of maize stover—A comparative study. Process. Saf. Environ. Prot. 2015, 93, 192–200. [Google Scholar] [CrossRef]
  47. Ye, H.; Yu, Z. Adsorption of Pb(II) onto Modified Rice Bran. Nat. Resour. 2010, 01, 104–109. [Google Scholar] [CrossRef] [Green Version]
  48. Moyo, M.; Pakade, V.E.; Modise, S.J. Biosorption of lead(II) by chemically modified Mangifera indica seed shells: Adsorbent preparation, characterization and performance assessment. Process. Saf. Environ. Prot. 2017, 111, 40–51. [Google Scholar] [CrossRef]
  49. Cardoso, N.F.; Lima, E.C.; Royer, B.; Bach, M.V.; Dotto, G.L.; Pinto, L.A.A.; Calvete, T. Comparison of Spirulina platensis microalgae and commercial activated carbon as adsorbents for the removal of Reactive Red 120 dye from aqueous effluents. J. Hazard. Mater. 2012, 241–242, 146–153. [Google Scholar] [CrossRef]
  50. Bangaraiah, P.; Peele, K.A.; Venkateswarulu, T.C. Removal of lead from aqueous solution using chemically modified green algae as biosorbent: Optimization and kinetics study. Int. J. Environ. Sci. Technol. 2021, 18, 317–326. [Google Scholar] [CrossRef]
  51. Do Nascimento, J.M.; de Oliveira, J.D.; Leite, S.G.F. Chemical characterization of biomass flour of the babassu coconut mesocarp (Orbignya speciosa) during biosorption process of copper ions. Environ. Technol. Innov. 2019, 16, 100440. [Google Scholar] [CrossRef]
  52. Tan, K.L.; Hameed, B.H. Insight into the adsorption kinetics models for the removal of contaminants from aqueous solutions. J. Taiwan Inst. Chem. Eng. 2017, 74, 25–48. [Google Scholar] [CrossRef]
  53. Nayak, A.K.; Pal, A. Green and ef fi cient biosorptive removal of methylene blue by Abelmoschus esculentus seed: Process optimization and multi-variate modeling. J. Environ. Manag. 2017, 200, 145–159. [Google Scholar] [CrossRef]
Figure 1. Determination of pHPZC values for NC and NCM biosorbents.
Figure 1. Determination of pHPZC values for NC and NCM biosorbents.
Molecules 28 00268 g001
Figure 2. Pb(II) adsorption capacities, qe, of treated (with NaOH) NCM and untreated NC biomass. pH 4.5, t = 60 min, biomass dose = 0.5 g L−1, T = 293 K.
Figure 2. Pb(II) adsorption capacities, qe, of treated (with NaOH) NCM and untreated NC biomass. pH 4.5, t = 60 min, biomass dose = 0.5 g L−1, T = 293 K.
Molecules 28 00268 g002
Figure 3. SEM images (left) and EDX spectra (rigth) of NC (a), NCM (b) and NCM-Pb (c).
Figure 3. SEM images (left) and EDX spectra (rigth) of NC (a), NCM (b) and NCM-Pb (c).
Molecules 28 00268 g003
Figure 4. FTIR spectra of the NCM before (left) and after Pb(II) biosorption (right).
Figure 4. FTIR spectra of the NCM before (left) and after Pb(II) biosorption (right).
Molecules 28 00268 g004
Figure 5. qe and Pb(II) species distribution vs. pH plots at different C0 initial concentrations of Pb ions. T = 293 K, t = 60 min, biosorbent dose = 0.5 g L−1.
Figure 5. qe and Pb(II) species distribution vs. pH plots at different C0 initial concentrations of Pb ions. T = 293 K, t = 60 min, biosorbent dose = 0.5 g L−1.
Molecules 28 00268 g005
Figure 6. qe (dotted) and %R (continuous) vs. biomass dose plots at different C0 concentrations. pH = 4.5, t = 60 min.
Figure 6. qe (dotted) and %R (continuous) vs. biomass dose plots at different C0 concentrations. pH = 4.5, t = 60 min.
Molecules 28 00268 g006
Figure 7. Weber–Morris plots of Pb(II) adsorption on NCM biosorbent.
Figure 7. Weber–Morris plots of Pb(II) adsorption on NCM biosorbent.
Molecules 28 00268 g007
Figure 8. NCM adsorption isotherms fitted to Langmuir (continuous lines) and Freundlich model (dotted lines). For pH 4.5 and pH 5; biosorbent dose = 0.5 g L−1, t = 60 min, room temperature.
Figure 8. NCM adsorption isotherms fitted to Langmuir (continuous lines) and Freundlich model (dotted lines). For pH 4.5 and pH 5; biosorbent dose = 0.5 g L−1, t = 60 min, room temperature.
Molecules 28 00268 g008
Figure 9. Regeneration efficiency %D of NCM vs. number of Pb(II) adsorption/desorption cycles. Biosorbent dose= 0.5 g L−1.
Figure 9. Regeneration efficiency %D of NCM vs. number of Pb(II) adsorption/desorption cycles. Biosorbent dose= 0.5 g L−1.
Molecules 28 00268 g009
Figure 10. Photograph of gelatinous colony of Nostoc commune.
Figure 10. Photograph of gelatinous colony of Nostoc commune.
Molecules 28 00268 g010
Table 1. Concentration of acidic and basic active sites on surface of NC and NCM evaluated by Boehm method.
Table 1. Concentration of acidic and basic active sites on surface of NC and NCM evaluated by Boehm method.
BiomassAcidic Sites (mmol g−1)Basic Sites (mmol g−1)
Untreated, NC0.520.02
Treated, NCM0.380.11
Table 2. Kinetic parameters for the adjustment of experimental data using kinetic models. T = 293 K, C0 = 442.42 mg·L−1.
Table 2. Kinetic parameters for the adjustment of experimental data using kinetic models. T = 293 K, C0 = 442.42 mg·L−1.
ModelParametersNCM Biosorbent
Pseudo first-orderqe,cal (mg g−1) (a)5.34
k1 (min−1)0.041
R20.6
Pseudo second-orderqe,cal (mg g−1) (a)384.6
k2 (g mg−1⋅min−1)0.0042
h (mg g−1⋅min−1)624.99
R21
Elovich β ´   (g mg−1) 0.16
α ×106 (mg g−1·min−1)4.27
R20.8
Intra-particle diffusionkid, I (mg g−1·min−1/2)115.9
BI7.25
R20.99
kid, II (mg g−1·min−1/2)7.4
BII332.84
R21
kid, III (mg g−1·min−1/2)0.3
BIII372.23
R20.81
(a) qe,cal: calculated adsorption capacity.
Table 3. Adjustment parameters of Pb (II) biosorption isotherms on NCM.
Table 3. Adjustment parameters of Pb (II) biosorption isotherms on NCM.
pH
ModelParameters4.55.5
Langmuirqmax (mg g−1) (a)384.6384.6
KL (L mg−1) 0.120.13
R2
χ2
0.99
1.6
0.99
7.6
Freundlichn7.79.1
KF (mg L1/n g−1 mg−1/n)178.5200.4
R2
χ2
0.99
0.6
0.98
1.5
Dubinin–Radushkevich
(D–R)
BDR (mol2 kJ−2) 2 × 10−87 × 10−8
qmax (mg g−1) (a)327.3309.3
E (kJ mol−1) 5002673
R2
χ2
0.67
34.8
0.67
36.3
Temkin B (J mol−1)
KT (Lg−1)
R2
χ2
75.3
256.16
0.96
2.27
89.6
217.41
0.93
4.51
(a)qmax: maximum biosorption capacity.
Table 4. Comparative table of the Pb(II) biosorption capacity qe of biomasses and agro-industrial wastes.
Table 4. Comparative table of the Pb(II) biosorption capacity qe of biomasses and agro-industrial wastes.
BiomassTreatmentqe,max (mg·g−1)Reference
Algae Nostoc sp
Algae Oedogonium sp
Non
Non
93.5
145.0
Gupta & Rastogi [9]
Algae Cystoseira strictaNaOH65Iddou et al. [19]
Olive StoneNaOH
NaOH
15
≤25.48
Blázquez et al. [4]
Ronda et al. [20]
Maize stoverHNO327.1Guyo et al. [46]
Rice branNaOH78.9Ye and Yu [47]
Mangifera indica seed shellsNaOH
Carboxyl functionalized
59.25
306.33
Moyo et al. [48]
Moringa oleifera tree leavesNaOH209.554Reddy et al. [34]
Nostoc communeNaOH384.6This work
Table 5. Thermodynamic parameters of Pb(II) biosorption on NCM.
Table 5. Thermodynamic parameters of Pb(II) biosorption on NCM.
Ho (kJ mol−1)So (J mol−1 K−1)Go (kJ mol−1)
293 K303K313K
−65.1−200.8−6.5−3.7−2.5
Table 6. Metal concentrations in industrial wastewater (real effluent), before (untreated) and after (treated) Pb(II) removal with NCM.
Table 6. Metal concentrations in industrial wastewater (real effluent), before (untreated) and after (treated) Pb(II) removal with NCM.
Metal Concentration (mg L−1)
PbCaKNa
Untreated wastewater5.85125.9250.3540.32
Treated wastewater0.1245.8445.2940.14
removal efficiency, % R97.963.610.00.4
Table 7. Models used for evaluation of Pb(II) biosorption onto NCM.
Table 7. Models used for evaluation of Pb(II) biosorption onto NCM.
ModelEquationParameters
Kinetic model
Pseudo-first order log ( q e q t ) = log ( q e ) k 1 2.303 t
qe: adsorption capacity at equilibrium
qt: amount of Pb(II) retained per unit mass of biosorbent at time t.
k1: the first-order kinetic constant
k2: rate constant adsorption
h: initial adsorption rate
Pseudo-second order t q t = 1 k 2 q e 2 + t q e
h = k 2 q e 2
Elovich q t = 1 β ´ ln . β ´ + 1 β ´ ln t     ∝: rate constant
β: constant related to the covered surface and the activation energy by chemisorption
Intra-particle diffusion q t = k id t 1 / 2 + c kid: intraparticle diffusion rate constant
c: constant
Isotherms
Langmuir C e q e = 1 q m a x K L + C e q m a x Ce: adsorbate concentration at equilibrium
qmax: Langmuir constant related to the maximum biosorption capacity
KL: Langmuir constant related to the affinity between sorbent and sorbate
Freundlich ln q e = ln K F + 1 n ln C e   KF: equilibrium constant
n: constant related to the affinity between sorbent and sorbate.
Dubinin–Radushkevich (D–R) ln q e = ln q max B DR   ε 2
ε = RT ln ( 1 + 1 C e )
ΒDR: constant related to adsorption energy
ε: Polanyi potential
Temkin q e = RT B ln ( K T ) + RT B ln C e R: universal gas constant
T: temperature
K T : Temkin’s equilibrium constant
B: adsorption energy variation
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lavado-Meza, C.; De la Cruz-Cerrón, L.; Lavado-Puente, C.; Angeles-Suazo, J.; Dávalos-Prado, J.Z. Efficient Lead Pb(II) Removal with Chemically Modified Nostoc commune Biomass. Molecules 2023, 28, 268. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules28010268

AMA Style

Lavado-Meza C, De la Cruz-Cerrón L, Lavado-Puente C, Angeles-Suazo J, Dávalos-Prado JZ. Efficient Lead Pb(II) Removal with Chemically Modified Nostoc commune Biomass. Molecules. 2023; 28(1):268. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules28010268

Chicago/Turabian Style

Lavado-Meza, Carmencita, Leonel De la Cruz-Cerrón, Carmen Lavado-Puente, Julio Angeles-Suazo, and Juan Z. Dávalos-Prado. 2023. "Efficient Lead Pb(II) Removal with Chemically Modified Nostoc commune Biomass" Molecules 28, no. 1: 268. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules28010268

Article Metrics

Back to TopTop