Heavy Metal Lead Removal by Biosorption – A Review

DOI : 10.17577/IJERTV5IS110351

Download Full-Text PDF Cite this Publication

Text Only Version

Heavy Metal Lead Removal by Biosorption – A Review

Chris Sheba M. & Monica Nandini G. K.

Department of Civil Engineering,

Sri Ramakrishna Institute of Technology, Pachapalayam, Perur Chettipalayam, Coimbatore 641 010, India.

Abstract Biosorption technique is increasingly researched and used in the biosorption of various heavy metals from industrial effluents over the past decades. Factors such as temperature, pH, initial metal ion concentration, biomass concentration and the metal affinity to the biosorbent, affect the efficiency of the biosorption capacity. Biosorbents such as large scale industrial fermentation by-products, agricultural waste, seaweeds, etc are used for lead removal. This article reviews the various research on biosorbents reported from 2013 2016, for the biosorption capacity & the kinetics and equilibrium isotherm models that the sorbents follow for optimum process. The thermodynamics of the process of certain biosorbents have also been selectively overviewed.

Keywords Lead; Biosorption; Biosorbent;, Isotherm; Adsorption capaciy.

  1. INTRODUCTION

    Heavy metal pollution is one of the major environmental problems prevalent today. Lead exists as Pb (II) and is toxic to living organisms in its concentrated forms. It is a metabolic poison which accumulates in the blood [1], mineralizing tissues (bones and teeth) [2] and soft tissues of human body (liver, kidneys, lungs, brain, spleen, muscles, and heart) [3]. It is an enzyme inhibitor which forms complexes with Oxo- groups in enzymes to affect virtually all steps in the synthesis of haemoglobin and porphyria metabolism [4]. It causes various health problems such as anaemia, Alzheimers disease, nervous system deterioration, failure of kidneys, bone problems and nervous system deterioration [5]. It affects infants and small children as well [6, 7]. The World Health Organization considers lead to be in the top ten chemical of public concern. [8] The permissible limit of lead in drinking water and surface water intended for drinking, as set by EU, US-EPA and WHO are 0.010, 0.015 and 0.010mgL-1

    respectively [9, 10 and 11].

    Industrial effluents containing high concentrations of lead are introduced in the water bodies from mining, manufacturing of batteries, paint, pigments, burning of coal, etc. Thus the removal and recovery of heavy metals from effluent streams are essential to the protection of the environment.

  2. BIOSORPTION PROCESS

    Biosorption is a property of certain types of inactive, dead, microbial biomass to bind and concentrate heavy metals from even dilute aqueous solutions. It is a metabolically passive process, meaning it does not require energy, and the amount of contaminants that a sorbent can remove is dependent on kinetic equilibrium and the composition of the sorbents cellular surface [12].

    The biosorption process involves a solid phase (sorbent or biosorbents; biological material) and a liquid phase (solvent, normally water) containing the dissolved metal ions (adsorbate, heavy metal) to be sorbent. Due to the higher affinity of the adsorbent for the adsorbate species, the latter is attracted and bound by different mechanisms such as ion exchange, complexation, chelation, microprecipitation, etc. The process continues till equilibrium is established between the amount of solid-bound adsorbate species and its portion remaining in the solution. The degree of adsorbent affinity for the adsorbate determines its distribution between the solid and liquid phases [13].

    1. Advantages of Biosorption Process

      Compared with the conventional heavy metal removal methods, the potential advantages of biosorption process includes [14]: High selectivity for recovery and removal of specific heavy metals High affinity thereby removing residual metals to < 1ppb Ability to treat large volumes of mixed wastes and wastewater with multiple heavy metals Use of naturally renewable biomaterial that are cheap and abundant, reducing the need for expensive reagents Wide range of physicochemical operational conditions including pH, temperature and also presence of other ions Relatively low investment and operational cost but with enhanced recovery of bound heavy metals ions The hazardous waste produces is greatly reduced in volume.

    2. Factors Affecting Biosorption of Metals

      Biosorption depends on many factors that are related to the biomass and metal; while some other factors are related to environmental conditions.

      1. Temperature: For most biomass, the biosorption activity occurs at the room temperature. However in certain cases, the temperature plays a major role in determining the optimum conditions for the process. In the biosorption study using peanut shells, its capacity to adsorb Pb(II) decreased as the temperature increased. It was also observed that the process was exothermic and spontaneous [15].

      2. pH: pH dependent experiments were advised not to be conducted at pH > 5 in order to avoid precipitation of Pb ions as hydroxides [16]. Since protons can be adsorbed or released in the biosorption process, the pH affects the metal ion solubility and the total charge of the biosorbent [17]. This behavior also depends on the functional groups present on the biomass cell wall.

      3. Initial metal ion and biosorbent concentrations: The concentration of exchangeable sites on the cell wall of the biosorbents affects its removal capacity. At minute initial concentrations of metal ions, 1g (say) of biomass would be enough for complete removal by biosorption. But when the metal ion concentration increases, the exchangeable sites in the 1g of biomass will be not be sufficient for biosorption, resulting in obvious reduced removal capacity. The study by Hasan et al, 2016 [18], the gradual increase in equilibrium uptake of the biosorbent with the increase in initial metal ion concentration was reported. This was probably due to the higher interaction between the exchangeable sites of the biosorbent and the metal ions. Hence the metal ion to adsorbent ratio plays a critical role in the biosorption process. In contrast, the presence of NaCl reduced the sorption efficiency due to the competition for binding sites between the Na+ and divalent cations [19]. Hence the biosorption capacity could be determined by the temperature, pH, initial metal ion concentration, biomass concentration and the metal affinity to the biosorbent.

  3. BIOSORBENTS

    The first major challenge for the biosorption field was to choose the most promising types of biomass/ biosorbent from enormously available and inexpensive biomaterials. Even though several materials of biological origin bind heavy metals, biomaterials with sufficiently high metal-binding capacity and selectivity for heavy metals are appropriate for full-scale biosorption process [20].

    A. Types of Biomass

    The types of biomass used for biosorption process includes large scale industrial by-products of fermentation (e.g. mold Rhizopus sp. [21, 22] and bacterium Bacillus sp. [23]), agricultural waste (like husks of rice [24], coffee [25], etc), seaweeds (e.g. algae) [26], etc. These biomass types can accumulate in excess 25% of their dry weight in heavy metals. Moreover chemically modified biomass shows improved biosorption capacity. The functionalized Saccharomyces cerevisiae cell with biogenic intracellular CaCO3 mineral scaffold was efficient for removing Pb(II) ions from aqueous solutions at the maximum removal capacity of 116.69 mg g1[27]. HNO3 modified Phytolacca americana was found to have higher uptake capacty than the original P. americana [28]. The different lead removal capacities of several biosorbents under optimum conditions are summarized in table I.

    For the biomass to be effective biosorbents, the vital evaluation elements for an adsorption process are the mechanism of biosorption and the reaction rate. Biosorbate uptake rate determines the time required for completing the adsorption reaction and can be described from the kinetic analysis.

    define the metal interactions (physisorption or chemisorptions). The intraparticle diffusion model defines the internal diffusion that establishes the adsorption rate in most of the liquid systems [32]. The linear equations of these kinetic models are presented in table II

    Lagergrens pseudo first order model describes the rate of adsorption reaction to be dependent on the available concentration of adsorbent. In previous studies, this model was not able to correlate with the experimental values, probably due to the time lag causing the presence of boundary layer that controls at the beginning of the biosorption process [37]. Pseudo second order model indicates chemisorptions to be the rate limiting step [42]. Biosorption of lead using corn husk [37], peanut shells [15], C. interruptus [38], Marula seed husk (S. birrea) [39], P. americana [28], maize stover [40], nanostructured cedar leaf ash [35], cashew nut shell (A. occidentale L.) [41], Casuarina Leaf Powder [16], S. melongena leaf powder [42], and Castor leaf (R. communis)

    [43] have been reported to follow pseudo second order model for adsorption.

    TABLE I. LEAD REMOVAL CAPACITIES OF VARIOUS BIOSORBENTS

    Biosorbent

    Operation Conditions

    Biosorption Capacity, q

    Ref.

    pH

    Biomass Concentration

    Peanut shells

    5.5

    2 g/L

    33 mg/g

    [15]

    Coffee Husk

    5

    2 g

    50 mg/g

    [25]

    P. americana L.

    6

    20 g/L

    10.83 mg/g

    [28]

    HNO3 modified P. americana L.

    12.66 mg/g

    Marula seed husk (Sclerocarya birrea)

    5

    0.8 g

    20mg/g

    [39]

    Castor Leaf Powder

    5

    0.02 g

    0.327 mmol/g

    [44]

    Ulva lactuca

    5

    0.8 g/L

    68.9 mg/g

    [50]

    Sargassum ilicifolium

    3.7

    0.2 g/L

    195 ± 3.3 mg/g

    [51]

    (live) Spirulina (Arthospira) maxima,

    7

    0.24 (g/L)

    27.25 mg/gdw

    [52]

    (live) Spirulina (Arthospira) indica,

    28.2 mg/gdw

    (live) Spirulina (Arthospira) platensis

    28.25 mg/gdw

    Curtobacterium sp.

    FM01

    6

    1 mg

    180.6 mg/g

    [53]

    Streptomyces fradiae

    biomass

    5

    1g dm-3

    138.88 mg/g

    [54]

    Sophora japonica

    pods

    6.0 – 7.0

    0.5 g/L

    25.13 mg/g

    [55]

    Otostegia persica

    5.5

    1.06 g/L

    17.3 mg/g

    [56]

    TABLE II. KINETIC MODEL LINEAR EQUATION [33]

    Kinetic Model

    Linear Equation

    Eqn. No.

    Plot representation

    Pseudo – first order

    log ( )

    = log ( ) 1

    2.303

    (1)

    log ( )

    vs

    Pseudo – second order

    = 2 +

    22

    (2)

    vs

    Intra – particle diffusion

    = 0.5 +

    (3)

    vs 0.5

    Kinetic Model

    Linear Equation

    Eqn. No.

    Plot representation

    Pseudo – first order

    log ( )

    = log ( ) 1

    2.303

    (1)

    log ( )

    vs

    Pseudo – second order

    = 2 +

    22

    (2)

    vs

    Intra – particle diffusion

    = 0.5 +

    (3)

    vs 0.5

  4. KINETICS OF ADSORPTION

    To explain the solid/liquid biosorption processes, Lagergren first described the process and presented the first order rate equation. In order to differentiate the kinetic equations based on adsorption capacity from solution concentration, the Lagergrens first order rate equation is also known as the pseudo first order kinetic equation [29, 30].The pseudo second order kinetic equation [31] and intra- particle diffusion model is also used to predict the sorption kinetics. The pseudo first order and pseudo second order

    Where the parameters are qe (mg/g) and qt (mg/g) are the adsorption capacity at equilibrium and time t (min), respectively; k1 (1/min) and k2 (g/ (mg min)) are the pseudo- first and pseudo-second order rate constants, respectively; Kid (mg/g min0.5) is the intraparticle diffusion rate constant and C is the intercept.

    The overall rate of adsorption can be described by the following three steps: (1) fluid transport, (2) film diffusion,

    (3) surface diffusion. If the nature of diffusion could not be identified by the first and second pseudo order model the intra-particle diffusion model is investigated. If the intraparticle diffusion is the sole rate-limiting step, intercept of the line, in the plot representation, should pass through the origin, i.e. C=0 [34]. The study on rapeseed [33] suggests that it underwent pseudo second order reaction and intra-particle diffusion.

  5. EQUILIBRIUM MODELS OF ADSORPTION

    In order to understand the relationship between the concentrations of lead adsorbed to its equilibrium concentration in the biosorbent solution, kinetic isotherms are applied. Isotherms define the adsorption efficiency. The different isotherm models along with its linear equation are given in table III.

    TABLE III. BIOSORPTION ISOTHERM MODEL – LINEAR EQUATION [33]

    Isotherm Model

    Linear Equation

    Eqn. No.

    Plot representation

    Freundlich [33]

    log

    = log

    + 1log

    (4)

    log log

    Langmuir [33]

    = + 1

    (5)

    Temkin [58]

    = ln

    + ln

    (6)

    ln

    Dubinin

    Radushkevich (D R [59]

    ln

    = ln 2

    = ln( 1

    + 1 )

    = 12

    (7)

    (8)

    (9)

    ln ln

    Sips (Langmuir Freundlich) [35]

    ln )

    (

    = ln

    + ln

    (10)

    ln ) ln

    (

    RedlichPeterson [35]

    ln ( 1)

    = ln + ln

    (11)

    ln ( 1) vs ln

    Where the qe is the sorption capacity at equilibrium (mg/g); qm is the maximum sorption capacity (mg/g); Ce (mg/L) is the equilibrium concentration of biosorbent; KL is the Langmuir constant (L/mg); KF is the Freundlich constants of the adsorption capacity and exponent n is the adsorption intensity which varies with heterogeneity of the material. Ks is the affinity constant for adsorption in Sips model (L/g); KR (L /g) and b (L/mg) are the RedlichPeterson constants of adsorption. KT is the Temkins equilibrium binding constant

    the Temkin isotherm constant (Jmol-1); R is the gas constant (8.314JK-1mol-1); T is the temperature in Kelvin; is the Polanyi potential for adsorption; Ea is the change in free energy when 1 mol of ion is transported to the surface of the solid in solution. It is calculated from , which is the D-R constant. If the magnitude of Ea is between 8 and 16 kJ/mol, then chemisorption process takes place and if the values of Ea is less than 8kJ/mol, then physisorption process occurs.

    The Langmuir isotherm model assumes that there is finite number of active sites distributed homogeneously over the surface of the adsorbent. The affinity of these active sites for adsorption is same towards a mono molecular layer and the interaction between adsorbed molecules is nil [37]. The biosorbents that undergo the respective isotherm models are summarized in table IV.

    TABLE IV. ISOTHERM PARAMETERS OF THE VARIOUS BIOSORBENTS

    Biosorbent

    Isotherm Model

    Remarks

    Ref.

    Peanut shells

    Langmuir

    qm = 39 mg g-1

    [15]

    Rice Husk Ash

    Langmuir

    qm = 0.0561mmol/g

    [24]

    P. americana

    Freudlich

    Kf = 0.6006 (mg g-1) (L

    mg-1)1/n

    1/n = 0.6742

    [28]

    Langmuir

    qm = 10.83 mg g-1

    HNO3 modified P. americana

    Freudlich

    Kf = 1.244 (mg g-1) (L mg-

    1)1/n

    1/n = 0.5906

    Langmuir

    qm = 12.66 mg g-1

    Rapeseed biomass

    Langmuir

    qm = 18.35mg/L at 4OC

    [33]

    Langmuir

    qm = 21.29mg/L at 20OC

    Langmuir

    qm = 22.7mg/L at 25OC

    Nanostructured cedar leaf

    Langmuir

    qm = 7.23 mg g-1

    [35]

    Sips

    qm = 8.045 mg g-1

    Corn husk

    Langmuir

    qm = 3.034 mg g-1

    [37]

    Cyclosorus interruptus

    Langmuir

    qm = 46.25 mg g-1

    [38]

    Raw maize stover

    Langmuir

    qm = 19.65 mg g-1

    [40]

    Treated maize stover

    Langmuir

    qm = 27.1 mg g-1

    Cashew nut

    Langmuir

    qm = 11.23 mg g-1

    [41]

    Olive stone

    Langmuir

    qm = 6.33 mg g-1

    [44]

    HNO3 modified Olive Stone

    Langmuir

    qm = 49.13 mg g-1

    H2SO4 modified Olive Stone

    Langmuir

    qm = 14.83 mg g-1

    NaOH modified Olive Stone

    Langmuir

    qm = 38.93 mg g-1

    Olive Tree Pruning

    Langmuir

    qm = 26.72 mg g-1

    HNO3 modified Olive Tree Pruning

    Langmuir

    qm = 86.4 mg g-1

    H2SO4 modified Olive Tree Pruning

    Langmuir

    qm = 72.78 mg g-1

    NaOH modified Olive Tree Pruning

    Langmuir

    qm = 123.8 mg g-1

    S. obliquus

    Langmuir

    qm = 112 mg g-1

    [45]

    Freudlich

    Kf = 22.35 (mg g-1) (L mg-

    1)1/n

    1/n = 0.768

    D R

    = – 0.64 x 10-8 mol2J-2

    qm = 31.6 mmol/g Ea = 8.84 kJ/mol

    Modified S. obliquus

    Langmuir

    qm = 207.2 mg g-1

    Freudlich

    Kf = 78.10 (mg g-1) (L mg-

    1)1/n

    1/n = 0.6019

    D R

    = – 0.46 x 10-8 mol2J-2

    qm = 24.9 mmol/g Ea = 10.44 kJ/mol

    Lentil husk

    Langmuir

    qm = 81.43 mg g-1

    [46]

    Biosorbent

    Isotherm Model

    Remarks

    Ref.

    Peanut shells

    Langmuir

    qm = 39 mg g-1

    [15]

    Rice Husk Ash

    Langmuir

    qm = 0.0561mmol/g

    [24]

    P. americana

    Freudlich

    Kf = 0.6006 (mg g-1) (L

    mg-1)1/n

    1/n = 0.6742

    [28]

    Langmuir

    qm = 10.83 mg g-1

    HNO3 modified P. americana

    Freudlich

    Kf = 1.244 (mg g-1) (L mg-

    1)1/n

    1/n = 0.5906

    Langmuir

    qm = 12.66 mg g-1

    Rapeseed biomass

    Langmuir

    qm = 18.35mg/L at 4OC

    [33]

    Langmuir

    qm = 21.29mg/L at 20OC

    Langmuir

    qm = 22.7mg/L at 25OC

    Nanostructured cedar leaf

    Langmuir

    qm = 7.23 mg g-1

    [35]

    Sips

    qm = 8.045 mg g-1

    Corn husk

    Langmuir

    qm = 3.034 mg g-1

    [37]

    Cyclosorus interruptus

    Langmuir

    qm = 46.25 mg g-1

    [38]

    Raw maize stover

    Langmuir

    qm = 19.65 mg g-1

    [40]

    Treated maize stover

    Langmuir

    qm = 27.1 mg g-1

    Cashew nut

    Langmuir

    qm = 11.23 mg g-1

    [41]

    Olive stone

    Langmuir

    qm = 6.33 mg g-1

    [44]

    HNO3 modified Olive Stone

    Langmuir

    qm = 49.13 mg g-1

    H2SO4 modified Olive Stone

    Langmuir

    qm = 14.83 mg g-1

    NaOH modified Olive Stone

    Langmuir

    qm = 38.93 mg g-1

    Olive Tree Pruning

    Langmuir

    qm = 26.72 mg g-1

    HNO3 modified Olive Tree Pruning

    Langmuir

    qm = 86.4 mg g-1

    H2SO4 modified Olive Tree Pruning

    Langmuir

    qm = 72.78 mg g-1

    NaOH modified Olive Tree Pruning

    Langmuir

    qm = 123.8 mg g-1

    S. obliquus

    Langmuir

    qm = 112mg g-1

    [45]

    Freudlich

    Kf = 22.35 (mg g-1) (L mg-

    1)1/n

    1/n = 0.768

    D R

    = – 0.64 x 10-8 mol2J-2

    qm = 31.6 mmol/g Ea = 8.84 kJ/mol

    Modified S. obliquus

    Langmuir

    qm = 207.2 mg g-1

    Freudlich

    Kf = 78.10 (mg g-1) (L mg-

    1)1/n

    1/n = 0.6019

    D R

    = – 0.46 x 10-8 mol2J-2

    qm = 24.9 mmol/g Ea = 10.44 kJ/mol

    Lentil husk

    Langmuir

    qm = 81.43 mg g-1

    [46]

    corresponding to the maximum binding energy (L/mg); bt is

    Anabaena sphaerica

    Langmuir

    qm = 121.95 mg g-1

    [47]

    Freundlich

    Kf = 28.28 (mg g-1) (L mg-

    1)1/n

    1/n = 0.2631

    D R

    = – 0.2433 x 10-8 mol2J-2

    qm = 1.104 mmol g-1 Ea = 14.3 kJ/mol

    Salvinia natans (at temperature = 15oC)

    Langmuir

    qm = 0.614 mmol g-1

    [48]

    Freundlich

    Kf = 0.439 mmol g-1 n = 2.964

    D R

    = – 0.022 x 10-5 mol2J-2

    qm = 0.508 mmol/g Ea = 4.814 kJ/mol

    Salvinia natans (at temperature = 30oC)

    Langmuir

    qm = 0.295 mmol g-1

    Freundlich

    Kf = 0.245 mmol g-1 n = 4.204

    D R

    = – 0.015 x 10-5 mol2J-2

    qm = 0.301 mmol/g Ea = 5.814 kJ/mol

    Mimusops elengi

    leaves

    Langmuir

    qm = 15.408 mg g-1

    [49]

    Freundlich

    Kf = 1.392 mg g-1 (mg g-1) (L mg-1)1/n

    1/n =

    Redlich Peterson

    A = 5.805 L/g; B = 3.261 L/mg

    Temkin

    AT = 2.03 L/mg; bT = 865.71

    D R

    = 0.0099 mol2kJ-2 qm = 29.18 mg/g

    Ea = 7.106 kJ/mol

    Jatropha curcas L.

    seed husk ash

    Langmuir

    qm = 263.1 mg g-1

    [57]

    Anabaena sphaerica

    Langmuir

    qm = 121.95 mg g-1

    [47]

    Freundlich

    Kf = 28.28 (mg g-1) (L mg-

    1)1/n

    1/n = 0.2631

    D R

    = – 0.2433 x 10-8 mol2J-2

    qm = 1.104 mmol g-1 Ea = 14.3 kJ/mol

    Salvinia natans (at temperature = 15oC)

    Langmuir

    qm = 0.614 mmol g-1

    [48]

    Freundlich

    Kf = 0.439 mmol g-1 n = 2.964

    D R

    = – 0.022 x 10-5 mol2J-2

    qm = 0.508 mmol/g Ea = 4.814 kJ/mol

    Salvinia natans (at temperature = 30oC)

    Langmuir

    qm = 0.295 mmol g-1

    Freundlich

    Kf = 0.245 mmol g-1 n = 4.204

    D R

    = – 0.015 x 10-5 mol2J-2

    qm = 0.301 mmol/g Ea = 5.814 kJ/mol

    Mimusops elengi

    leaves

    Langmuir

    qm = 15.408 mg g-1

    [49]

    Freundlich

    Kf = 1.392 mg g-1 (mg g-1) (L mg-1)1/n

    1/n =

    Redlich Peterson

    A = 5.805 L/g; B = 3.261 L/mg

    Temkin

    AT = 2.03 L/mg; bT = 865.71

    D R

    = 0.0099 mol2kJ-2 qm = 29.18 mg/g

    Ea = 7.106 kJ/mol

    Jatropha curcas L.

    seed husk ash

    Langmuir

    qm = 263.1 mg g-1

    [57]

  6. THERMODYNAMICS OF ADSORPTION

In order to assert the thermodynamic nature of lead biosorption, several basic parameters including change in free energy (G0), enthalpy (H0), entrophy (S0) were found by researchers by the following formula.

G0 = RT ln K (12)

CONCLUSION

The removal of lead by various biosorbents has been reviewed. The major demerits of conventional heavy metal removal techniques are high cost, low removal efficiency, increased chemical and biological effluent sludge, additional chemical requirements for regeneration of adsorption medium and difficulty in biosorbent recovery. Thus by using biosorbents, the problems caused by the conventional treatments can be minimized and the process can be performed in an environmental friendly way.

REFERENCES

  1. Agency for Toxic Substances and Disease Registry (ATSDR), Toxicological profile for Lead, Atlanta, GA:

    U.S. Department of Health and Human Services, Public Health Service, 2007.

  2. E. K. Silbergeld, J. Sauk, M. Somerman, A. Todd, F. McNeill, B. Fowler, et al., Lead in bone: storage site, exposure source, and target organ, Neurotoxicology, vol. 14 (2-3), pp. 225-36, Summer-Fall, 1993.

  3. P. S. L. Barry, and D. B. Mossman, Lead concentrations in human tissues, Brit. J. industr. Med., vol. 27 (4), pp. 339 351, Oct. 1970.

  4. C. M. A. Ademorati, Pollution by Heavy metals, Environmental Chemistry and Toxicology, Foludex press Ibadan, pp. 171-172, 1996.

  5. A. Navas-Acien, E. Guallar, K. E. Silbergeld, J. S. Rothenberg, Lead exposure and cardiovascular disease a systematic review, Environ Health Perspect, vol. 115, pp. 472482, 2007.

  6. K. Schumann, The toxicological estimation of the heavy metal content (Cd, Hg, Pb) in food for infants and small children, Z. Ernahrungswiss, vol. 29, pp. 54-73, 1990.

  7. Theodore I. Lidsky, Jay S. Schneider, Lead neurotoxicity in children: basic mechanisms and clinical correlates, Brain,

ln K = S0 H0

(13)

vol. 126 (1), pp. 5-19, 2003.

R RT

Where T is the temperature (Kelvin, K); R is the ideal gas

constant (8.314 Jmol-1K-1) and K is the thermodynamic equilibrium constant. According to the equation 12 and 13, H0 and S0 can be calculated from the slope and intercept of the plot of ln(K) versus 1/T respectively. The thermodynamic equilibrium constant is the ratio of the amount of adsorbate (mg) adsorbed per litre to the equilibrium concentration of the adsorbate in mg/L.

Negative G0 indicates thermodynamic feasibility and spontaneity of the biosorption process. The magnitude of H0 indicates whether the biosorption is physical or chemical. The heat of adsorption for a physical reaction is around 2.1 20.9 kJmol-1 and the energy of activation for chemical reaction (Ea) is in the range of 20.9 418.4 kJmol-1 [37]. S0 indicates the randomness during biosorption.

C. interruptus [38] could biosorb lead chemically while peanut shells [15] and M. elengi leaves [49] manifested physical biosorption. Rice husk ash [24] and P. americana

[28] were found to exhibit both physico-sorption and chemisorption process, exothermally and spontaneously, while the biosorption of lead by maize stover [40], S. melongena leaf powder [42], corn husk [37] and S. ilicifolium [51], was found to be spontaneous and endothermic.

http://dx.doi.org/10.1093/brain/awg014

  1. World Health Organization, Ten chemicals of major public health concern, 2016, Retrieved from http://www.who.int/ipcs/assessment/public_health/chemical s_phc/en/ Assessed on 24th October 2016.

  2. European Union (Drinking Water) Regulations, Government Publications, Dublin, 2014. ISBN 978-1-4468-1806-0.

  3. https://nepis.epa.gov/Exe/ZyPDF.cgi?Dockey=60001N8P.tx t Assessed on 24th October 2016.

  4. World Health Organization, Lead in Drinking-water, Background document for development of WHO Guidelines for Drinking-water Quality, incorporating 1st and 2nd addenda, Vol.1, Recommendations. 3rd ed, WHO, Geneva, Switzerland, 2011.

  5. L. Velásquez, and J. Dussan, "Biosorption and bioaccumulation of heavy metals on dead and living biomass of Bacillus sphaericus". J. Hazard. Mater. vol. 167 (1-3), pp. 7136, 2009.

  6. N. Ahalya, T. V. Ramachandra, and R. D. Kanamadi, Biosorption of Heavy Metals, Research Journal of Chemistry and Environment, vol. 7, pp. 71-79, 2003.

  7. Z. Aksu, Y. Sag, and T. Kutsal, The biosorption of copper by C. vulgaris and Z. ramigera, Environ. Technol., vol. 13, pp. 579-586, 1992.

  8. . Taar, F. Kaya, and A. Özer, Biosorption of lead (II) ions from aqueous solution by peanut shells: Equilibrium, thermodynamic and kinetic studies, Journal of Environmental Chemical Engineering, vol. 2 (2), pp. 1018 1026, June 2014. http://dx.doi.org/10.1016/j.jece.2014.03.015

  9. S. J. Rao, K. C. Rao, and G. Prabhakar, Optimization of Biosorption Performance of Casuarina Leaf Powder for the Removal of Lead Using Central Composite Design, J. Environ. Anal. Toxicol., vol. 3, no. 2, pp. 166, 2013. http://dx.doi:10.4172/2161- 0525.1000166

  10. Romera, E., F. González, A. Ballester, M. L. Blázquez, and

    J. A. Munoz. "Comparative study of biosorption of heavy metals using different types of algae." Bioresource Technology, vol. 98, no. 17, pp. 3344-3353, 2007.

  11. Hasan, H. A., Abdullah, S. R. S., Kofli, N. T., and Yeoh, S. J., Interaction of environmental factors on simultaneous biosorption of lead and manganese ions by locally isolated Bacillus cereus. Journal of Industrial and Engineering Chemistry, vol. 37, pp. 295-305, 2016.

  12. Lutts, S., Qin, P., and Han, R. M., Salinity influences biosorption of heavy metals by the roots of the halophyte plant species Kosteletzkya pentacarpos, Ecological Engineering, vol. 95, pp. 682-689.

  13. Norton L., Baskaran K., and Mckenzie T., Biosorption of zinc from aqueous solutions using biosolids, Adv. Environ. Res., vol. 8, pp. 62935, 2004.

  14. Uslu G., Dursun A. Y., Ekiz H. I., Aksu Z., The effect of Cd(II), Pb(II) and Cu(II) ions on the growth and bioaccumulation properties of Rhizopus arrhizus, Process Biochem., vol. 39, pp. 10510, 2003.

  15. Bahadir T., Bakan G., Altas L., and Buyukgungor H., The investigation of lead removal by biosorption: an application at storage battery industry wastewaters, Enzyme and Microbial Technology, vol. 41, no. 1, pp. 98-102, July 2007.

  16. Tunali, S., Cabuk, A. and Akar, T., Removal of lead and copper ions from aqueous solutions by bacterial strain isolated from soil, Chemical Engineering Journal, vol. 115, no. 3, pp. 203-211, 2006.

  17. Vieira, M. G. A., A. F. de Almeida Neto, M. G. C. Da Silva,

    C. N. Carneiro, and A. A. Melo Filho, "Adsorption of lead and copper ions from aqueous effluents on rice husk ash in a dynamic system." Brazilian Journal of Chemical Engineering 31, no. 2, pp. 519-529, 2014.

  18. Berhe, S., Ayele, D., Tadesse, A. and Mulu, A., Adsorption Efficiency of Coffee Husk for Removal of Lead (II) from Industrial Effluents: Equilibrium and Kinetic Study, International Journal of Scientific and Research Publications, vol. 5, no. 9, September 2015.

  19. Sheng P. X, Ting Y. P, Chen J. P, and Hong L. Sorption of lead, copper, cadmium, zinc, and nickel by marine algal biomass: characterization of biosorptive capacity and investigation of mechanisms, J. Colloid Interface Sci., vol. 275, pp. 13141, 2004

  20. Ma, X., Cui, W., Yang, L., Yang, Y., Chen, H. and Wang, K., Efficient biosorption of lead (II) and cadmium (II) ions from aqueous solutions by functionalized cell with intracellular CaCO3 mineral scaffolds, Bioresource technology, vol. 185, pp.70-78, 2015. http://dx.doi.org/10.1016/j.biortech.2015.02.074

  21. Wang G., Zhang S., Yao P, Chen Y, Xu X, Li T, and Gong G., Removal of Pb(II) from aqueous solutions by Phytolacca americana L. biomass as a low cost biosorbent, Arabian Journal of Chemistry, June 2015. http://dx.doi.org/10.1016/j.arabjc.2015.06.011

  22. Lodeiro P., Barriada J. L., Herrero R., Sastre D. E. and Vicente M. E., The marine macroalga Cystoseira baccata as biosorbent for cadmium (II) and lead (II) removal: kinetic and equilibrium studies, Environ. Pollu., vol. 142, pp. 264- 273, 2006.

  23. Low K. S., Lee C. K. and Liew S. C., Sorption of cadmium and lead from aqueous solutions by spent grain, Process Biochem., vol. 36, pp. 59-64, 2000.

  24. McKay G. and Ho Y. S., Pseudo-second-order model for sorption process, Process Biochem., vol. 34, pp. 451-465, 1999.

  25. I. Tsibranska, and E. Hristova, Comparison of different kinetic models for adsorption of heavy metals onto activated carbon from apricot stones, Bulgarian Chemical Communications, vol. 43, no. 3, pp. 370 377, 2011.

  26. Morosanu, I., Teodosiu, C., Paduraru, C., Ibanescu, D. and Tofan, L., Biosorption of lead ions from aqueous effluents by rapeseed biomass, New Biotechnology, 2016. http://dx.doi.org/10.1016/j.nbt.2016.08.002

  27. A. Maleki, E. Pajootan, and B. Hayati, Ethyl acrylate grafted chitosan for heavy metal removal from wastewater: equilibrium, kinetic and thermodynamic studies, J. Taiwan Inst. Chem. Eng., vol. 51, pp. 127134, 2015.

  28. Hafshejani L. D., Nasab S. B., Gholami R. M., Moradzadeh M., Izadpanah Z., Hafshejani SB, et al., Removal of zinc and lead from aqueous solution by nanostructured cedar leaf ash as biosorbent, Journal of Molecular Liquids, vol. 211, pp. 44856, 2015. http://dx.doi.org/10.1016/j.molliq.2015.07.044.

  29. Langmuir, I., The adsorption of gases on plane surfaces of glass, mica and platinum, J. American Chem. Soc., vol. 40, pp. 1361-1403, 1918.

  30. D. Sarkar, G. Paul and Ramachandra Murthy, T. T. S., Studies on biochemical thermodynamics of lead biosorption from aqueous system using corn husk biomass as biosorbent agent, International Journal of Current Research, vol. 8, no. 09, pp. 37592-37598, 2016.

  31. Zhou K, Yang Z, Liu Y, and Kong X., Kinetics and equilibrium studies on biosorption of Pb(II) from aqueous solution by a novel biosorbent: Cyclosorus interruptus, J. Environ. Chem. Eng. vol. 3, pp. 221928, 2015. doi:http://dx.doi.org/10.1016/j.jece.2015.08.002

  32. Moyo M., Guyo U., Mawenyiyo G., Zinyama N. P, and Nyamunda B. C., Marula seed husk (Sclerocarya birrea) biomass as a low cost biosorbent for removal of Pb(II) and Cu(II) from aqueous solution, J. Ind Eng Chem, vol. 27, pp. 12632, 2015.

    doi:http://dx.doi.org/10.1016/j.jiec.2014.12.026

  33. Guyo U., Mhonyera J., and Moyo M., Pb(II) adsorption from aqueous solutions by raw and treated biomass of maize stover a comparative study, Process Saf Environ Prot., vol. 93, pp. 192200, 2015. http://dx.doi.org/10.1016/j.psep.2014.06.009

  34. Coelho G. F., Gonçalves Jr. A. C., Tarley C. R. T., Casarin J., Nacke H., Francziskowski M. A., Removal of metal ions Cd (II), Pb (II), and Cr (III) from water by the cashew nut shell Anacardium occidentale L., Ecol. Eng., vol. 73, pp. 51425, 2014.

    http://dx.doi.org/10.1016/j.ecoleng.2014.09.103

  35. G. Yuvaraja, N. Krishnaiah, M.V. Subbaiah, and A. Krishnaiah, Biosorption of Pb(II) from Aqueous Solution by Solanum melongena Leaf Powder as a Low-Cost Biosorbent Prepared from Agricultural Waste, Colloids and Surfaces B: Biointerfaces, 2013. http://dx.doi.org/10.1016/j.colsurfb.2013.09.039

  36. Martns, A. E., Pereira, M. S., Jorgetto, A. O., Martines, M. A., Silva, R. I., Saeki, M. J. and Castro, G. R., The reactive surface of Castor leaf [Ricinus communis L.] powder as a green adsorbent for the removal of heavy metals from natural river water, Applied Surface Science, vol. 276, pp. 24-30, 2013.

  37. Blázquez, G., Ronda, A., Martín-Lara, M.A., Pérez, A. and Calero, M., Comparative study of isotherm parameters of lead biosorption by two wastes of olive-oil production, Water Science and Technology, vol. 72, no.5, pp.711-720, 2015.

  38. Abdel Ghafar, H.H., Abdel-Aty, A.M., Ammar, N.S. and Embaby, M.A., Lead biosorption from aqueous solution by raw and chemically modified green fresh water algae Scenedesmus obliquus, Desalination and Water Treatment, vol. 52, no. 40-42, pp.7906-7914, 2014.

  39. Basu, M., Guha, A.K. and Ray, L., Biosorptive removal of lead by lentil husk, Journal of Environmental Chemical Engineering, vol. 3, no. 2, pp.1088-1095, 2015. http://dx.doi.org/10.1016/j.jece.2015.04.024

  40. Abdel-Aty, A.M., Ammar, N.S., Ghafar, H.H.A. and Ali, R.K., Biosorption of cadmium and lead from aqueous solution by fresh water alga Anabaena sphaerica biomass, Journal of advanced research, vol. 4, no. 4, pp.367-374, 2013. http://dx.doi.org/10.1016/j.jare.2012.07.004

  41. Lima L. K. S., Silva J. F. L., Da Silva M. G. C., Vieira M.

    G. A., Lead biosorption by Salvinia natans biomass: equilibrium study, Chemical Engineering Transactions, vol. 38, pp. 97-102, 2014.

  42. Kalyani G., Prasanna Kumar Y. and King P., Optimization of Lead Biosorption by an Ecofriendly Biosorbent Mimusops elengi Using Central Composite Design, Research Journal of Pharmaceutical, Biological and Chemical Sciences, vol. 7, no. 3, pp. 1586 1603, 2016.

  43. Ibrahim, W. M., Hassan, A. F., and Azab, Y. A., Biosorption of toxic heavy metals from aqueous solution by Ulva lactuca activated carbon, Egyptian Journal of Basic and Applied Sciences, vol. 3, no. 3, pp. 241-249, 2016. http//dx.doi.org/10.1016/j.ejbas.2016.07.005

  44. Tabaraki, R., Nateghi, A., and Ahmady-Asbchin, S., Biosorption of lead (II) ions on Sargassum ilicifolium: Application of response surface methodology, International Biodeterioration & Biodegradation, vol. 93, pp. 145-152, 2014.

  45. Siva Kiran R. R., Madhu G. M., Satyanarayana S. V., Kalpana P., Bindiya P., Subba Rangaiah G., Equilibrium and kinetic studies of lead biosorption by three Spirulina (Arthrospira) species in open raceway ponds, J. Biochem Tech., vol. 6, no. 1, pp. 894-909, 2015.

  46. Masoumi, F., Khadivinia, E., Alidoust, L., Mansourinejad, Z., Shahryari, S., Safaei, M., et al., Nickel and lead biosorption by Curtobacterium sp. FM01, an indigenous bacterium isolated from farmland soils of northeast Iran, Journal of Environmental Chemical Engineering, vol. 4, no. 1, pp.950-957, 2016.

  47. Gergana, K., Zdravka, V., Margarita, S., Yana, H., Iliev, I. and Velizar, G., Biosorption of Pb (II) ions from aqueous solutions by waste biomass of Streptomyces fradiae pretreated with NaOH, Biotechnology and Biotechnological Equipment (Bulgaria). vol. 29, no. 4, pp. 689-695, 2015.

  48. Amer, M.W., Ahmad, R.A. and Awwad, A.M., Biosorption of Cu (II), Ni (II), Zn (II) and Pb (II) ions from aqueous solution by Sophora japonica pods powder, International Journal of Industrial Chemistry, 6(1), pp.67-75, 2015. DOI 10.1007/s40090-014-0030-8

  49. Alavi, S. A., Zilouei, H. and Asadinezhad, A., Otostegia persica biomass as a new biosorbent for the removal of lead from aqueous solutions, Int. J. Environ. Sci. Technol., vol. 12, pp. 489 – 498, 2015. doi:10.1007/s13762-014-0705-x

  50. Shi, B., Zuo, W., Zhang, J., Tong, H. and Zhao, J., Removal of Lead (II) Ions from Aqueous Solution Using Jatropha curcas L. Seed Husk Ash as a Biosorbent, Journal of environmental quality, vol. 45, no. 3, pp. 984-992, 2016.

  51. Aharoni A. and Ungarish M., Kinetics of activated chemisorptions Part 2. Theoretical models, J Chem Soc Faraday Trans., vol. 73, pp. 456464, 1977.

  52. Huston N. D. and Yang R. T., Theoretical basis for the DubininRadushkevich (DR) adsorption isotherm equation, Adsorption, vol. 3, pp. 189195, 1997.

Leave a Reply