Download as pdf or txt
Download as pdf or txt
You are on page 1of 122

Chapter – III Literature Review

LITERATURE REVIEW

The major pollution for the present society is Environmental pollution.

The review of the literature indicates various methods for removal of heavy metals

from aqueous solutions using different biosorbents. The author has obtained

literature from various sources and the literature obtained on various methods for

removal of chromium and lead is summarized below along with response surface

methodology.

3.1 CHROMIUM (Cr):

Agarwal et al [27] have studied on biosorption of aqueous chromium (VI) by

Tamarindus indica seeds. In this study the effectiveness of low cost agro-based

materials namely, Tamarindus indica seed (TS), crushed coconut shell (CS),

almond shell (AS), ground nut shell (GS) and walnut shell (WS) were evaluated

for Cr (VI) removal. Batch test indicated that hexavalent chromium sorption

capacity (qe) followed the sequence qe (TS) > qe (WS) > qe (AS) > qe (GS) > qe

(CS). Due to high sorptive capacity, tamarind seed was selected for detailed

sorption studies. Sorption kinetic data followed first order reversible kinetic fit

model for all the sorbents. The equilibrium conditions were achieved within 150

min under the mixing conditions employed. Sorption equilibria exhibited better

fit to Freundlich isotherms (R > 0.92) than Langmuir isotherm (R < 0.87).

Hexavalent chromium sorption by TS decreased with increase in pH, and slightly

reduced with increase in ionic strength. Cr (VI) removal by TS seems to be

mainly by chemisorption.

87
Chapter – III Literature Review

Ahmet Sarı et al [28] have studied on biosorption of total chromium from

aqueous solution by red algae (Ceramium virgatum): Equilibrium, kinetic and

thermodynamic studies. This study focused on the biosorption of total chromium

onto red algae (Ceramium virgatum) biomass from aqueous solution.

Experimental parameters affecting biosorption process such as pH, contact time,

biomass dosage and temperature were studied. The effect of biomass dosage on

the biosorption of total chromium was studied using different biomass dosage in

the range, 0.4–40 g/L and optimum is attrained at 10 g/L. The effect of pH on the

removal efficiency of total chromium onto C. virgatum was studied by changing

pH values in the range of 1–8 and the results were presented and the maximum

value attained at pH 1.5. The biosorption capacity of C. virgatum biomass for

total chromium was found to be 26.5 mg/g at and 10 g/L biomass dosage, pH 1.5,

90 min equilibrium time and 20 ◦C. From the D–R isotherm model, the mean free

energy was calculated as 9.7 kJ/mol, indicating that the biosorption of total

chromium was taken place by chemisorptions.

Annadural et al [29] examined the liquid-phase adsorption removal of Cu2+,

C02+, Ni2+, Zn2+, and Pb2+ in the concentration range of 5-25 mg/L using low-cost

banana and orange peel wastes at 30 °C. Under comparable conditions, the

amount of ldsorption decreased in the order Pb2+ > Ni2+ > Zn2+ > Cu2+ > C02+ for

both adsorbents. The adsorption isotherms were better described by the

Freundlich equation. The amount of adsorption increased with increasing pH and

reached a plateau at pH > 7, which was confirmed by the variations of zeta

potentials. The application potential of such cellulose-based wastes for metal

88
Chapter – III Literature Review

removal (up to 7.97 mg Pb2+ per gram of banana peel at pH 5.5) at trace levels

appeared to be promising.

Aoyama [30] investigated the effect of initial metal concentration, contact time

and solution temperature on the removal of Cr (VI) from solution by waste

London plane leaves, generated by the pruning of street trees, in batch mode

conditions. The removal of Cr (VI) was highly concentration dependent and was

shown to be mainly governed by physico-chemical adsorption under the weak

acidic conditions. The equilibrium data fitted well in the Langmuir isotherm

model. The Langmuir constants were calculated at different temperatures and

both the adsorption capacity and adsorption intensity increqsed with rising

temperature. The endothermic nature of the Cr (VI) adsorption was confirmed by

the thermodynamic parameters. The study showed that the waste leaves could be

used as an effective adsorbent for removal of Cr (VI) from wastewater.

Aoyama and Tsuda [31] investigated the ability of larch (Larix leptolepis Gold)

bark to remove Cr (VI) from dilute aqueous solutions. The research parameters

included the solution pH, contact time, temperature and initial concentration of Cr

(VI) in solution. Of the parameters studied, the solution pH was found to be the

most crucial. The Cr (VI) removal decreased steadily throughout the pH range

studied (pH 2-6), while the Cr adsorption peaked at pH 3. Because the chemical

reduction of Cr (VI) to trivalent state occurred to lesser extent even in strong

acidic media, the Cr (VI) removal was mainly governed by physico-chemical

adsorption. The positive value of the heat adsorption (HO) indicated that the

endothermic nature of the Cr (VI) adsorption. The relatively slow rate and

89
Chapter – III Literature Review

irreversible nature of the adsorption as well as the order of the magnitude of the

heat adsorption value suggested that the adsorption was of a chemical type. The

adsorption data obtained from equilibrium experiments were well fitted to both

Langmuir find Freundlich isotherms.

Atac Uzel et al [32] have studied on metal biosorption capacity of the organic

solvent tolerant Pseudomonas fluorescens TEM08. The effects of pH, contact

time and metal ion concentration on the biosorption of Cr (VI) and Ni (II) were

studied. The time intervals of metal sorption were 2, 5, 15, 30, 60, 90 and 120

min. The effects of the initial concentrations of Cr (VI) and Ni (II) ions were

studied at pH 2.0 and 5.0. All pH adjustment were made using reagent grade HCl

and NaOH. The samples were centrifuged at 10,000 g for 20 min at room

temperature. According to the Q values, the biosorption capacity of the type I

(40.8 mg/g) and type II (40.7 mg/g) cells was nearly same for Cr (VI). However,

a 9.67% decrease was observed between type I (12.4 mg/g) and type II (11.2

mg/g) cells for Ni (II) biosorption.

Awan ct al [33] studied the removal of heavy metals i.e., Pb, Cr, Cu, and Zn from

their aqueous solutions using ordinary sand as an adsorbent. The experiments

were carried out at 20 °C. The amount of metal adsorbed to form monolayer on

sand (am), which was obtained from Langmuir isotherm, exhibited the preference

of metals for sand in the order Pb>Cr>Cu>Zn. The heavy metal-sand adsorption

phenomenon was illustrated on the basis of the interaction between surface

functional group of silicates (sand) and the metal ions. It was deduced that sand

90
Chapter – III Literature Review

can be used as a low-cost adsorbent: or the removal of heavy metal from waster

(containing low concentrations of metals), especially in the developing countries.

Bala Kiran et al [34] have studied on biosorption of Cr (VI) by native isolate of

Lyngbya putealis (HH-15) in the presence of salts. The present study reports on

biosorption of Cr (VI) by native isolate of Lyngbya putealis HH-15 in batch

system under varying range of pH (2.0–10.0), initial metal ion concentration (10–

100 mg/L) and salt concentration (0–0.2 %). Maximum metal removal (94.8 %)

took place at pH 3.0 with initial Cr concentration of 50 mg/L, which got reduced

(90.1 %) in the presence of 0.2 % salts. Adsorption equilibrium and kinetic

behavior of Cr (VI) in solution was also examined. Both Langmuir and

Freundlich models fitted well to explain the adsorption data (R2 = 0.90 and 0.87,

respectively) at 0.2 % salt concentration. Pseudo-second order kinetic model also

fitted well to both the systems, viz. Cr (VI) and Cr (VI) salt.

Balibek et al [35] studied the Chromium (VI) biosorption by dried Rhizopus

arrhizus. In this study the biosorption of chromium (VI) from saline solutions on

dried Rhizopus arrhizus was studied as a function of pH, initial chromium (VI)

and salt (NaCl) concentrations in a batch system. In the absence of salt, when the

initial chromium (VI) concentration increased from 25 to 250 mg/L

approximately, the loading capacity increased from 23.2 to 108.9 mg/g due to the

increase in the number of ions competing for the available binding sites on the

biomass surface. The uptake of chromium (VI) reached a plateau at 250 mg/L

showing the saturation of binding sites at higher concentration levels. For the

optimum pH experiments were performed at different initial pH values (1.0–6.0)

91
Chapter – III Literature Review

and at different initial NaCl concentrations (0–50 g /L). The maximum sorption

obtained at pH 2.0 Equilibrium uptakes of chromium (VI) increased with

increasing chromium (VI) concentration and decreased considerably by the

presence of increasing concentrations of salt. And as the concentration of salt

increased from 0 to 50 g/L, the amount of chromium (VI) adsorbed by dried

R.arrhizus and chromium (VI) removal yield diminished from 23.2 to 18.4 mg/g

and from 92.1 to 72.2 %.

Bayat [36] compared the removals of Cr (VI) and Cd (II) from an aqueous

solution using two different Turkish fly ashes Afsin-Elbistan and Seyitomer as

adsorbents. The influence of four parameters (contact time, solution pH, initial

metal concentration in solution and ash quality) on the removal at 20 ± 2 °C was

studied. Fly ashes were found to have a higher adsorption capacity for the

adsorption of Cd (II) as compared to Cr (VI) and both Cr (VI) and Cd (II) required

an equilibrium time of 2h. The adsorption of Cr (VI) was higher at pH 4.0 for

Afsin-Elbistan fly ash (25.46 %) and pH 3.0 for Seyitomer fly ash (30.91 %)

while Cd (II) was adsorbed to a greater extent (98.43 % for Afsin-Elbistan and

65.24 % for Afsin-Elbistan fly ash) at pH 7. The adsorption of Cd (II) increased

with an increase in the concentrations of these metals in solution while Cr (VI)

adsorption decreased by both fly ashes. The lime (crystalline CaO) content in fly

ash seemed to be a significant factor in influencing Cr (VI) and Cd (II) ions

removal. The linear forms of the Langmuir and Freundlich equations were

utilized for experiments with metal concentrations- of 55 ± 2 mg/I for Cr (VI) and

6 ± 0.2 mg/I for Cd (II) as functions of solution pH (3.0-8.0). The adsorption of

Cr (VI) on both fly ashes was not described by both the Langmuir and Freundlich

92
Chapter – III Literature Review

isotherms while Cd (II) adsorption on both fly ashes satisfied only the Langmuir

isotherm model. The adsorption capacities of both fly ashes were nearly three

times less than that of activated carbon for the removal of Cr (VI) while Afsin-

Elbistan Oy ash with high-calcium content was as effective as activated carbon for

the removal of Cd (II). Therefore, there were possibilities for use the adsorption

of Cd (II) ions onto fly ash with high-calcium content in practical applications in

Turkey.

Blázquez et al [37] have studied on the effect of pH on the biosorption of Cr (III)

and Cr (VI) with olive stone. In this work, the biosorption of trivalent and

hexavalent chromium has been studied using olive stone, a low-cost natural

sorbent, analysing the pH effect in the biosorption process. Removal of Cr (III)

and Cr (VI) with olive stone is very sensitive to solution pH. The percentage of

Cr (III) removed reaches a maximum at range pH between 4 and 6, reaching

values close to 90 %. However, at pH 3 the retention of Cr (III) is less than 50 %.

The percentage of Cr (VI) removed is higher than 80 % when the pH is equal or

lower than 2, whereas as the pH increases this percentage drops to values than

15%. Also, the results obtained for the final concentration of total Cr, Cr (VI) and

Cr (III) indicate that a combined effect of biosorption and reduction is involved in

the Cr (VI) removal specially when the pH value is lower than 3.

Calero et al [38] have made a study of Cr (III) biosorption in a fixed-bed column.

The effects of flow rate and inlet concentration on Cr (III) biosorption by olive

stone were studied by varying the flow rate from 2 to 6 mL/min and inlet Cr (III)

concentration from 10 to 50 mg/L, while the bed height was held constant at

93
Chapter – III Literature Review

8.9cm (10 g of olive stone). The olive stone sorption capacity, qe, increases as the

inlet Cr (III) concentration increases until a value close to 0.800 mg/g is reached,

then qe remains constant. Column data obtained at different conditionswere

described using the Adams–Bohart, Thomas, Yoon and Nelson, and Dose–

Response models. The breakthrough curve prediction by the Adams–Bohart and

Dose–Response models were found to be very satisfactory. In particular, the

Adams–Bohart model can be used to represent the initial region of breakthrough

curve, whereas the Dose–Response model is the one which best reproduces the

whole curve for all the inlet Cr (III) concentrations used.

Carlo Solisio et al [39] have studied on the effect of acid pre-treatment on the

biosorption of chromium (III) by sphaerotilus natans from industrial wastewater.

Batch experiments carried out at starting acid conditions (pH 3.0 ± 3.5) show that

the pH progressively increases but the removing activity starts only when

conditions closed to neutrality are reached. The equilibrium concentration is

achieved more rapidly (only 25 h) during the tests carried out at Cmo 113 ± 16

mg/L and at an intermediate initial biomass concentration (0.40 g/L), while a

longer time is necessary using both lower (0.27 g/L) and higher (0.85 g/L) initial

biomass levels.

Chergui et al [40] have studied on simultaneous biosorption of Cu2+, Zn2+ and

Cr6+ from aqueous solution by Streptomyces rimosus biomass. The Cu2+, Zn2+

and Cr6+ biosorption capacity of the Streptomyces rimosus biomass pretreated

with NaOH was studied in the batch mode. Five grams of native biomass were

maintained for 30 min in 500 ml NaOH (0.1 M) with a stirring speed of 250 rpm

94
Chapter – III Literature Review

and at ambient temperature (around 25 °C). Under optimal experimental

conditions, a biosorption capacity of 30 mg/g Cu2+ biomass, 27.4 mg/g Zn2+

biomass and 26.7 mg/g Cr6+ biomass was obtained. The equilibrium data poorly

fitted the Langmuir and Freundlich model isotherms over the whole range of

initial Cu2+, Zn2+ and Cr6+ concentrations tested (0–300 mg/L).

Deepa Prabhu Mungasavalli et al [41] have studied on biosorption of chromium

from aqueous solutions by pretreated Aspergillus niger: Batch and column

studies. An initial factorial design of experiments showed that factors such as pH

of the solution, temperature and biomass mass were important. The kinetics of

biosorption of chromium was found to follow Ho pseudo-second order reaction.

Isotherm studies conducted at 5 ± 2, 15 ± 2, 22 ± 2 and 30 ± 2 ◦C provided

maximum biosorption capacities of 14.5, 15.2, 10.6, and 11.6 mg/g, respectively.

Reusability of biomass was examined by the desorption studies, in which NaOH

eluted 90 % chromium.

Donghee Park et al [42] have studied on XAS and XPS studies on chromium-

binding groups of biomaterial during Cr (VI) biosorption. In this study, abiotic Cr

(VI) reduction by the biomass was performed with various contact times, pHs and

initial Cr (VI) concentrations, and surface and bulk characteristics of the Cr-laden

biomass was then investigated using X-ray photoelectron spectroscopy (XPS) and

X-ray absorption spectroscopy (XAS). In conclusion, it was obvious that oxygen

containing groups, such as carboxyl and phenolic groups, play a major role in the

binding of the Cr (III) resulting from the abiotic reduction of Cr (VI) by the

biomass.

95
Chapter – III Literature Review

Donghee Park et al [43] have worked on how to study Cr (VI) biosorption: Use

of fermentation waste for detoxifying Cr (VI) in aqueous solution. To determine

the effects of Cr (VI) removal of pH variation, pH values of 1–4 were used; of

initial Cr (VI) concentration variation, concentrations of 25, 50, 75 and 100 mg/L

were used; and of biomass concentration variation, biomass concentrations of 5,

10, 15 and 20 g/L were used. The flasks were agitated on a shaker at 200 rpm.

For the 25 mg/L of initial Cr (VI) concentration, Cr (VI) was completely removed

in 2 h, while the complete removal of 100 mg/L of Cr (VI) required 10 h of

contact time.

Donghee Park et al [44] have worked on studies of hexavalent chromium

biosorption by chemically-treated biomass of Ecklonia sp. The aims of the

present investigation were to enhance the Cr (VI) reducing capacity of the

biomass using various chemical treatments and to elucidate the mechanisms

governing Cr (VI) reduction. Among the various chemical treatments, acid-

treatment showed the best performance with regards the improvement of Cr (VI)

removal from the aqueous phase, while organic solvent-treatment significantly

improved the removal efficiency of total Cr in the equilibrium state. These

findings indicated that the amino and carboxyl groups take part in the Cr (VI)

removal from the aqueous phase.

Elangovan et al [45] have studied on biosorption of hexavalent and trivalent

chromium by palm flower (Borassus aethiopum). Batch kinetic and equilibrium

experiments were conducted to determine the adsorption kinetic rate constants and

maximum adsorption capacities. Both Cr (III) and Cr (VI) adsorption followed

96
Chapter – III Literature Review

the second-order kinetics. For Cr (III), maximum adsorption capacity was 6.24

mg/g by raw adsorbent and 1.41 mg/g by acid treated adsorbent. In case of Cr

(VI), raw adsorbent exhibited a maximum adsorption capacity of 4.9 mg/g,

whereas the maximum adsorption capacity for acid treated adsorbent was 7.13

mg/g. There was a significant difference in the concentrations of Cr (VI) and total

chromium removed by palm flower. In case of Cr (VI) adsorption, first it was

reduced to Cr (III) with the help of tannin and phenolic compounds and

subsequently adsorbed by the biosorbent. Acid treatment significantly increased

Cr (VI) adsorption capacity of the biosorbent whereas, alkali treatment reduced

the adsorption capacities for Cr (VI). However, in case of Cr (III), acid treatment

significantly reduced the adsorption capacity whereas the adsorption capacity of

alkali treated biosorbent was slightly less than that of raw adsorbent. FT-IR

spectrum showed the changes in functional groups during acid treatment and

biosorption of Cr (VI) and Cr (III).

Elangovan et al [46] have studied on biosorption of chromium species by aquatic

Weeds. Batch kinetic and equilibrium experiments were conducted to determine

the adsorption kinetic rate constants and maximum adsorption capacities of

selected biosorbents. In most of the cases, adsorption followed a second-order

kinetics. For Cr (III), maximum adsorption capacity was exhibited by reed mat

(7.18 mg/g). In case of Cr (VI), mangrove leaves showed maximum

removal/reduction capacity (8.87 mg/g) followed by water lily (8.44 mg/g). There

was a significant difference in the concentrations of Cr (VI) and total chromium

removed by the biosorbents. In case of Cr (VI) removal, first it was reduced to Cr

(III) with the help of tannin, phenolic compounds and other functional groups on

97
Chapter – III Literature Review

the biosorbent and subsequently adsorbed. Acid treatment significantly increased

Cr (VI) removal capacity of the biosorbents where as, alkali treatment reduced the

Cr (VI) removal capacities of the biosorbents. FTIR spectrum showed the

changes in functional groups during acid treatment and biosorption of Cr (VI) and

Cr (III).

Erol Pehlivan et al [47] have studied on biosorption of chromium (VI) ion from

aqueous solutions using walnut, hazelnut and almond shell. The potential to

remove Cr (VI) ion from aqueous solutions through biosorption was investigated

in batch experiments. Kinetic experiments revealed that the dilute chromium

solutions reached equilibrium within 100 min. The effect of pH on adsorption of

Cr (VI) was studied at room temperature by varying the pH of metal solution–

shell suspension from 2.0 to 9.0. The biosorptive capacity of the shells was

dependent on the pH of the chromium solution, with pH 3.5 being optimal.

Adsorption of Cr (VI) ion uptake is in all cases pH dependent showing a

maximum at equilibrium pH values between 2.0 and 3.5, depending on the

biomaterial,that correspond to equilibrium pH values of 3.5 for (WNS), 3.5 for

(HNS) and 3.2 for (AS). The adsorption data fit well with the Langmuir isotherm

model. The sorption process conformed to the Langmuir isotherm with maximum

Cr (VI) ion sorption capacities of 8.01, 8.28, and 3.40 mg/g for WNS, HNS and

AS, respectively. Percentage removal by WNS, HNS and AS was 85.32, 88.46

and 55.00 %, respectively at a concentration of 0.5 mM.

Ertugay et al [48] have studied on biosorption of Cr (VI) from aqueous solutions

by biomass of Agaricus bisporus. In this study, biosorption of Cr (VI) ion was

98
Chapter – III Literature Review

investigated by using biomass of Agaricus bisporus (a species of mushroom) in a

temperature and shaking speed controlled shaker.the batch adsorption studies at

different pH values were carried out in the range of 1.0, 2.0, 3.0, 5.0 and 7.0. The

highest uptake yield of Cr (VI) at pH 1.0 was found to be 92.4 %. For biosorption

experiments, Cr (VI) solution having 25–125 mg/L was prepared and used.

Optimum biosorption conditions were found to be pH 1, C0 = 50 mg/L, m= 10

g/L, shaking speed = 150 rpm, T = 20 ◦C Cr (VI), respectively.

Etinkaya Donmez et al [49] have done a comparative study on eavy metal

biosorption characteristics of some algae. The experiments were conducted with

pH range 1-6 and varying concentrations from 25 to 250 mg/L. The flasks were

agitated on a shaker for 24 h. Samples were taken at pre-determined time

intervals (0, 5, 15, 30, 60, 120, 180, 240, 300, 960 and 1440 min) for the residual

metal ion concentrations in the solution. Optimum adsorption pH values of

copper (II), nickel (II) and chromium (VI) were determined as 5.0, 4.5 and

2.0.respectively, for all three algae. At the optimal conditions, metal ion uptake

increased with initial metal ion concentration up to 250 mg/L.

Fadali ct al [50] studied the removal of chromium from tannery effluents by

adsorption onto cement kiln dust. Tannery effluent was characterized not only by

heavy loads but also with toxic heavy metals especially chromium ions.

Chromium was considered as an important source of contamination due to large

volume of exhaust liquid discharged and solid sludge produced. Details on

adsorption studies were carried out using synthetic chromium salts (chromium

chloride) as adsorbate, and cement kiln dust (a waste from white cement industry)

as adsorbent. Equilibrium isotherms have been determined for the adsorption of

99
Chapter – III Literature Review

chromium ions on cement kiln dust. Kinetic study provided that the adsorption

process is diffusion controlled. The experimental results have been fitted using

Freundlich. Langmuir isotherms. The maximum adsorption capacity of cement

kiln dust was found to be 33 mg/g. Industrial tannery effluent (22 mg/L

chromium and COD 952 mg/L) was also treated by cement dust. The treated

effluent (using 20 g cement dust per 1 L) contained only 0.6 mg/L chromium and

COD 200 mg/L.

Gholamreza Moussavi et al [51] have studied on biosorption of chromium (VI)

from industrial wastewater onto pistachio hull waste biomass. In the present

work, pistachio hull powder (PHP) was investigated for the removal of hexavalent

chromium (Cr (VI)) from wastewater. The effects of pH (2–8), PHP concentration

(0.5–8 g/L), Cr (VI) concentration (50–200 mg/L), temperature (5–50 ◦C) and

contact time (1–60 min) were studied on the removal of Cr (VI) from aqueous

solution. The results revealed that PHP adsorbs over 99% of chromium from

solutions containing 50–200 mg/L of Cr (VI) at a pH of 2 and an adsorbent

concentration of 5 g/L after 60 min of equilibration. The percent chromium

adsorbed from solution increased with an increase in temperature from 5 to 40 ◦C.

Kinetic and isotherm modeling studies demonstrated that the experimental data

best fit a pseudo-second order and Langmuir model, respectively. The maximum

Langmuir adsorption capacity was 116.3 mg/g. In the second part of the study,

the efficacy of PHP was examined by analyzing the removal of Cr (VI) from

industrial wastewater. Results revealed that 2 g/L of PHP decreased the Cr (VI)

concentration from 25 mg/L to less than 0.05 mg/L after 30 min of equilibration.

100
Chapter – III Literature Review

Gokhale et al [52] have studied on kinetic and equilibrium modeling of

chromium (VI) biosorption on fresh and spent Spirulina platensis/Chlorella

vulgaris biomass. In the current study, pH was varied from 1.5 to 5.5. The

optimum pH was found to be 1.5. Algal cell concentration varied from 0.2 g/L to

2.4 g/L. The initial chromium (VI) ion concentration was kept fixed at 100 mg/L.

For both the species it was found that at an algal concentration of 2.4 g/L, almost

complete removal of chromium (VI) ions from the solution was achieved. Fresh

algal biomass of S. platensis gave maximum of 73.6 % biosorption of chromium

(VI) in 100 ppm solution at 1 g/L cell loading. The maximum biosorption by

spent biomass was increased to 86.2 %. Thus, this two step process not only

showed improved efficiency in biosorption (17% increase) but also gave valuable

byproduct, namely b-carotene.

Gokhale et al [53] have studied on modeling of chromium (VI) biosorption by

immobilized Spirulina platensis in packed column. This study describes

biosorption of chromium (VI) by immobilized Spirulina platensis, in calcium

alginate beads. Three aspects viz. optimization of bead parameters, equilibrium

conditions and packed column operation were studied and subsequently modeled.

In these experiments, 10 g of beads were contacted to 100 ml chromium (VI) ion

solution of concentration 100 mg /L. Samples were taken at 0, 1, 2, 3, 4, 5, 7.5,

10, 12.5, 15, 17.5, 20, 25, 30, 40, 50, 60, 75, 90, 105, 120 min and analyzed for

chromium (VI) ion concentration. Optimized beads were used for these

equilibrium studies. In these experiments, beads were contacted to chromium

(VI) ion solution in varying quantities (4, 8, and 12 g per 100 ml). Chromium

(VI) ion concentration in the solutionwas varied as 100, 150, 200 mg/L. The

101
Chapter – III Literature Review

flaskswere rotated on a shaker at 180 rpm for 4 h and the samples were analyzed

for residual chromium (VI) ions in solution. Under optimized bead diameter (2.6

mm), calcium alginate concentration (2 %, w/v) and biomass loading (2.6 %, w/v)

maximum biosorption was achieved. 140 g/L loading of optimized beads resulted

in 99 % adsorption of chromium (VI) ions from an aqueous solution containing

100mg/L of chromium (VI).

Gupta et al [54] have studied on biosorption of hexavalent chromium by raw and

acid-treated green alga Oedogonium hatei from aqueous solutions. Batch

experiments were conducted to determine the biosorption properties of the

biomass. It is observed that the equilibrium chromium sorption was favored by

acidic pH range of 1–2.0 and maximum adsorption by the algal biomass was

observed at pH 2.0. The effect of temperature on the adsorption of Cr (VI) on the

biomass is investigated at three different temperatures (298, 308 and 318

K).optimum temperature is 318 K. To assess the effect of biosorbent dose,

different amounts (0.1–1.0 g/L) of biosorbent the optimum algal biomass dose

selected was 0.8 g/L. It has been observed that the metal adsorption rate is high at

the beginning for both the algal biomasses and then decreases slowly till

saturation levels were completely reached at equilibration point (110 min). Under

the optimal conditions, the biosorption capacities of the raw and acid-treated algae

were 31 and 35.2 mg Cr (VI) per gm of dry adsorbent, respectively.

Gupta et al [55] have studied on biosorption of chromium (VI) from aqueous

solutions by green algae spirogyra species. Batch experiments were conducted to

determine the adsorption properties of the biomass and it was observed that the

102
Chapter – III Literature Review

adsorption capacity of the biomass strongly depends on equilibrium pH.

Equilibrium isotherms were also obtained and the percent adsorption of Cr (VI)

increases with increase in pH from pH 1.0 to 2.0 and thereafter decreases with

further increase in pH, the maximum removal of Cr (VI) was around 14.7 x 10 (3)

mg metal/kg of dry weight biomass at a pH of 2.0 in 120 min with 5 mg/L of

initial concentration.

Guven Ozdemir et al [56] have studied on biosorption of chromium (VI),

cadmium (II) and copper (II) by Pantoea sp. TEM18. In this work, among

microorganisms isolated from wastewater treatment of a petrochemical industry, a

gram-negative bacterium Pantoea sp. TEM18 exhibited the greatest copper

tolerance. It was able to survive in the medium containing copper at

concentrations as high as 180 mg/L. The biosorption properties of bacterial

biomass for cadmium and the effects of environmental factors (i.e. pH, metal

concentration contact time) on the chromium, cadmium and copper biosorption

were explored. Optimum adsorption pH values ranging from 1.0 to 8.0 of

chromium (VI), cadmium (II) and copper (II) were determined as 3.0, 6.0 and 5.0,

respectively. Experimental results also showed the influence of initial metal

concentration on the metal uptake for dried biomass.

Guven Ozdemir et al [57] have studied on heavy metal biosorption by biomass

of Ochrobactrum anthropi producing exopolysaccharide in activated sludge. In

these tests, initial pH values of 100 mg/L aqueous metal solutions were adjusted in

the range 1.0–8.0, Optimum adsorption pH values of chromium (VI), cadmium

(II) and copper (II) were 2.0, 8.0 and 3.0 respectively. Maximum equilibrium

uptakes of copper (II) ions were also determined as 26.0 mg/g at 91.6 mg/L of

103
Chapter – III Literature Review

initial copper (II) ion concentrations. Experimental results also showed the

influence of initial metal concentration on the metal uptake for dried biomass.

Handan Ucun et al [58] have studied on biosorption of chromium (VI) from

aqueous solution by cone biomass of Pinus sylvestris. The study was done with

variation in the parameters of pH, initial metal ion concentration and agitation

speed. The biosorption of Cr (VI) was carried out at different initial Cr (VI) ion

concentrations ranging from 50 to 300 mg/L. The biosorption of Cr (VI) was

increased when pH of the solution was decreased from 7.0 to 1.0. The maximum

chromium biosorption occurred at 150 rpm agitation. An increase in

chromium/biomass ratio caused a decrease in the biosorption efficiency.

Hua Yin et al [59] have studied on improvement of chromium biosorption by

UV–HNO2 cooperative mutagenesis in Candida utilis. Batch sorption

experiments were performed at various aqueous phase pH values (1.0, 3.0, 5.0,

7.0, 9.0, 11.0 and 13.0) by keeping all other experimental conditions constant

(temperature, 30 ◦C; agitation, 120 rpm; biosorbent, 2.5 g; chromium

concentration, 20 mg/L). pH 1–5 and pH 1–3 were optimal ranges for CRC2811-

1 and CR-001, respectively. The experiments were carried out using Cr (VI)

solutions (pH 2.0) ranging from 5–200 mg/L at 30 ◦C. 3.2.4. Effect of absorbent

dosage on biosorption Different amounts of biomass ranging from 0–30 g/L were

added to the final concentration 20 mg/L Cr (VI) solutions at pH 2, after shaking

for 2 h. The removal efficiency of these mutants CRU132-26, CRC7-2,

CRC2811-1, CRC2811-2, CRC2814-8 and CRY182-1 for 20 mg/L Cr (VI)

solutions were 85.6 %, 95.2 %, 87.0 %, 82.5 %, 94.7 % and 82.7 %, respectively.

104
Chapter – III Literature Review

Hugo Figueiredo et al [60] have made a study on the ffect of the supporting

zeolite structure on Cr biosorption: Performance of a single-step reactor and of a

sequential batch reactor. This work presents a study on the applicability of a

zeolite-biomass system to the entrapment of metallic ions, starting from Cr (VI)

solutions up to 100 mg Cr/L, in batch processes. The effect of the zeolitic support

on the overall system performance was evaluated comparing two large pore

zeolitic structures which differ in chemical composition and ion-exchange

capacity. The systems were tested in single-step and in sequential processes. In

single-step studies, HY zeolite was found to be the most efficient support when

applied to low Cr concentrations (overall Cr removal of 93.4 %), whereas for the

higher initial Cr concentration, the higher ion-exchange capacity of NaY zeolite

was determinant to achieve the highest overall Cr removal of 77.6 %.The

evolution of Cr (VI) entrapment was strongly dependant on the zeolitic support

used in the system. In sequential batch processes, HY zeolite was found to be the

most efficient support with a 98.2 % overall Cr removal. The reduction of Cr (VI)

promoted by the biomass is more suited to the dynamics of the sequential process.

NaY zeolite behaved similarly to HMOR and NaMOR zeolites, as these systems

removed between 87.3 and 93.4 % of the initial Cr.

Huidong Li et al [61] have studied a novel technology for biosorption and

recovery of hexavalent chromium in wastewater by bio-functional magnetic

beads. The goal of this study was to develop an applied technique for the removal

and recovery of heavy metal in wastewater. The results indicated that the beads

were kept in good operating conditions under different pH values ranging from 1

to 12 at agitation speed 200 rpm for 96 h. At the same time, the performance of

105
Chapter – III Literature Review

these beads was not affected by temperature between 0 ◦C and 50 ◦C. The

parameters effecting Cr (VI) removal were obtained: the optimum pH 1.0 and

optimum temperature 28 ◦C. It was observed that the sorption capacity increased

with lowering pH and the maximum sorption capacity was found to be 6.73 mg/G

at initial concentration 40 mg/L and pH 1.0.

Jang et al [62] conducted a series of adsorption experiments in order to assess the

ability of three mulches to remove several of the heavy metal ions typically

encountered in urban runoff. Three types of mulch, cypress bark (C), hardwood

bark (H), and pine bark nugget (P), were selected as potential adsorbents to

capture heavy metals in urban runoff. The hardwood bark (H) mulch had the best

physicochemical properties for adsorption of heavy metal ions. In addition,

because of its fast removal rate and acceptably high capacity for all the heavy

metal ions, it was concluded that the H mulch is the best of the three adsorbents

for treatment of urban runoff containing trace amounts of heavy metals. In order

to investigate the sorption isotherm, two equilibrium models, the Freundlich and

Langmuir isotherms, were analyzed. The sorption of these metals on H mulch

conformed to the linear form of the Langmuir adsorption equation. At pH 5 and

6, the Langmuir constants (S(m)) for each metal were found to be 0.324 and 0.359

mmol/g (Cu); 0.306 and 0.350 mmol/g (Pb); and 0.185 and 0.187 mmol/g (Zn) at

25 °C.

Jinshao Ye et al [63] have studied on biosorption of chromium from aqueous

solution and electroplating wastewater using mixture of Candida lipolytica and

dewatered sewage sludge. In this study, the objective was to investigate Cr

106
Chapter – III Literature Review

removal from aqueous solutions, as well as Cr, Cu, Ni and Zn from electroplating

wastewaters by the mixture of Candida lipolytica and sewage sludge. The effect

of biosorbent dosage on the biosorption of 30 mg/ L Cr (VI) solution at initial pH

3.0 was studied using different dosage ranging from 0.1 to 5.0 g/L. 1.0 g/L of C.

lipolytica and 3.0 g/L of sewage sludge was selected for further experiments. The

biosorption studies were carried out with different initial pH values ranging from

1.0 to 12.0. The range of optimal pH for the mixture (1–5) was wider than C.

lipolytica (1–4) and sewage sludge (2–4), respectively. The effect of Cu and Zn in

combination was significant on the removal of total Cr and the bioreduction of

Cr (VI).

Jorge L. Brasil et al [64] have studied on statistical design of experiments as a

tool for optimizing the batch conditions to Cr (VI) biosorption on Araucaria

angustifolia wastes. Based on this, preliminary experiments about the

optimization of the biosorption of Cr (VI) on pinh˜ao wastes were carried out

using 5.00–100.0 mg/L Cr (VI), the acidity was adjusted at pH 2.0, biosorbent

concentration of 5 g/L, speed of agitation of 120 strikes/min and contact time of

5–360 min. It was observed that for solutions 5.00 mg/L of Cr (VI) after shaking

for 5 min, the removal of Cr (VI) from solution was 100.0 %. For a 100.0 mg/L of

Cr (VI) solution a similar behavior was observed; however, at 5 min the removal

of the element was 90.2 % and attained 100.0 % after 30 min of contact with the

biosorbent.

Kadirvelu et al [65] predicted the efficiency of activated carbons prepared from

various agricultural wastes for the removal of dyes and metal ions from aqueous

107
Chapter – III Literature Review

solutions. Activated carbons were prepared from agricultural solid wastes, silk

cotton hull, coconut tree sawdust, sago waste, maize cob and banana pith and used

to eliminate heavy metals and dyes from aqueous solution. Adsorption of all dyes

and metal ions required a very short time and gave quantitative removal.

Experimental results showed that all carbons were effective for the removal of

pollutants from water. Since all agricultural solid wastes used in this investigation

were freely, abundantly and locally available, the resulting carbons were expected

to be economically viable for wastewater treatment. Based on the above

information, it is found that the use of neem leaf powder as adsorbent is meager.

So in his present investigation equilibrium and kinetic studies are carried out for

removal of chromium using neem leaf powder as adsorbent.

Kondapalli Srividya et al [66] have studied on biosorption of hexavalent

chromium from aqueous solutions by Catla catla Scales. The percentage removal

of Cr (VI) increased from 35.06 to 60.89 when the dose was changed from 0.05 to

0.4 g as the doseage range is from 0-045 g. the effect of contact time for five

different initial concentrations of Cr (VI) (10–0 mg/L) with biosorbent dose (0.05

g) at natural pH of the solution. The figure shows that percentage removal of Cr

(VI) adsorption increases with time from 0 to 180 min ormore and then becomes

almost constant up to the end of the experiment. The highest metal uptake was

reported to be 27.18 mg/g obtained at pH 1.0 and the overall metal uptake of Cr

(VI) decreased to 2.44 mg/g as pH increased up to 6.

Lee ct al [67] studied the feasibility of using green sands as a low cost reactive

medium in permeable reactive barriers (PRBs). Waste green sands were industrial

108
Chapter – III Literature Review

byproducts of the gray iron foundry industry. These green sands were composed

of fine silica sand, clay binder, organic carbon, and residual iron particles.

Because of their potential sorptive and reactive properties, tests were performed to

determine the feasibility of using green sands as a low cost reactive medium in

permeable reactive barriers (PRBs). Serial batch kinetic tests and conventional

batch sorption tests were conducted to determine the removal characteristics for

zinc in aqueous solutions. Removal characteristics for zinc in the presence of

green sands are comparable to those of Peerless iron, a common reactive medium

used in PRBs. High removal capacities for zinc of green sands were attributed to

clay, organic carbon, and residual run particles, which were known sorptive media

for heavy metals. Furthermore, high pH values in the presence of clay and

residual iron particles enhanced sorption and precipitation of zinc.

Lei Yang et al [68] have studied on biosorption of hexavalent chromium onto raw

and chemically modified Sargassum sp. Hexavalent chromium biosorption by raw

algae is always accompanied with significantly high organic leaching. In this

study, hydrochloric acid, sodium hydroxide, calcium chloride, formaldehyde, and

glutaraldehyde were used for modification of raw Sargassum sp. Seaweed. At pH

ranging from 1 to 14, there are four soluble Cr (VI) species, which are CrO2+4,

Cr2O2+7, HCrO+4 and H2CrO4. From pH 1 to pH 6, the percentage of HCrO+4

changes from 86.1 % to 93.8 %. From pH 6 to pH 10 or even higher, the

percentage of H2CrO4 increases sharply from 2.3 % to 100 %. It is shown that pH

2.0 is the optimal pH for the Cr (VI) uptake for both biosorbents. At the optimal

pH of 2.0, the maximum biosorption capacities of MSW and RSW are 1.123 and

0.601 mmol g/L, respectively. The surface treatment improves the reduction

109
Chapter – III Literature Review

capacity of the biosorbents. The complete uptake of hexavalent chromium is

achieved in 20 h.

Levankumar Lakshmanraj et al [69] have studid on the biosorption of

hexavalent chromium from aqueous solutions by using boiled mucilaginous seeds

of Ocimum americanum. Batch experiments were conducted to study the

biosorption kinetics of chromium removal for the concentrations 10 mg/L, 20

mg/L and 40 mg/L of chromium (VI) solutions. The chromium reduction was

82% and 78 % for the pH values 1 and 1.5. However it decreased to 20 % with

increase in pH of chromium (VI) solution to 7. The optimum chromium reduction

takes place at the pH range 1–1.5. The continuous column study was also carried

out at the flow rate of 27 mL/h for the initial concentration 25 mg/L of chromium

(VI) feed solution using a packed bed column filled with boiled mucilaginous

seeds. The maximum reduction of chromium (VI) to chromium (III) in the packed

bed was 80 %. The percentage removal of reduced chromium from the aqueous

solution was 56.25 %. This valuewas maintained constant until 0.52 L of

chromium (VI) solutionwas pumped through the packed bed column.

Liping Deng et al [70] have studied on biosorption of Cr (VI) from aqueous

solutions by nonliving green algae Cladophora albida. The experiments were

conducted in 250 mL flasks containing 50 mL Cr (VI) solution and 0.1 g biomass

with varying pH from 0.5 to 12. Effect of initial Cr (VI) concentration from 20 to

154 mg/L was studied. The mixtures were shaken on a rotary shaker (agitation

rate, 200 rpm) for 24 h. the optimum pH was chosen at a range of 1.0–3.0 algal

dosage was varied from 0.2 to 10 g/ in 50 mL metal solution at 25 oC. The

110
Chapter – III Literature Review

optimum algal dosage 2 g/L was used in the experiment. The removal rate of Cr

(VI) was relatively rapid in the first 60 min, but then the rate decreased gradually.

Margarita Enid R. Carmona et al [71] have studied on biosorption of chromium

using factorial experimental design. The three factors considered were pH,

temperature, and metal concentration at two markedly different levels: Cr3+, pH

(2.0 and 6.0), T (29 and 55 oC), and metal concentration (10 and 1200 mg/L);

Cr6+, pH (1.0 and 3.0), T (29 and 55 oC), and metal concentration (10 and 1200

mg/L). Experiments were carried out in a batch type reactor system with 0.2 g of

biosorbent (Sargassum sp.), and 50mL of Cr3+ or Cr6+ solutions. The efficiency of

chromium removal during an exposition time of 6 h was then evaluated. The most

significant effect regarding Cr3+ uptake was ascribed to interaction between metal

concentration and pH. For Cr6+, the most significant effect was ascribed to metal

concentration.

Maria X. Loukidou et al [72] have studied on equilibrium and kinetic modeling

of chromium (VI) biosorption by Aeromonas caviae. Biosorption of hexavalent

chromium, from aqueous solutions, on Aeromonas caviae particles was

investigated in a well-stirred batch reactor. Equilibrium and kinetic experiments

were performed at various initial bulk concentrations, biomass loads, temperatures

and ionic background. a detailed analysis has been conducted testing several

chemical reaction kinetic models in order to identify a suitable kinetic equation,

assuming that biosorption is chemical sorption controlled.

111
Chapter – III Literature Review

Martin-Dupont et al [73] investigated equilibria and mechanisms involved in the

adsorption process of metal ions (Cr3+, Cu2+, Ni2+, Pb2+ and Zn2+) from aqueous

solutions using coniferous barks as biosorbent substrate. Crude barks were used

in this study since previous experiments showed a decreasing uptake for

chemically treated barks in the considered granulometry. In the experimental

conditions, the maximum binding capacity of barks followed, the decreasing order

Cr3+ > Cu2+ > Pb2+ > Ni2+ > Zn2+ whereas their general binding affinity decreased

as: Pb2+ > Cr3+ > Ni2+ > Zn2+ > Cu2+. Adsorption isotherms at the optimal

physicochemical conditions were established and the adsorption phenomenon was

described by the non-competitive Langmuir adsorption model which fitted well

the experimental data. An evaluation of adsorption capability was carried out

using model parameters which were graphically determined. Models for removal

of cations have been discussed; they represent efficient tools for predicting the

behaviour of the biosorbents in metal ion adsorption systems.

Murphy et al [74] have studied on comparative study of chromium iosorption by

red, green and brown seaweed biomass. pH experiments in this study were

conducted in the pH range 1.5–6. Metal sorption was highly pH dependent with

maximum Cr (III) and Cr (VI) sorption occurring at pH 4.5 and pH 2,

respectively. Extended equilibrium times were required for Cr (VI) binding over

Cr (III) binding (180 and 120 min, respectively) thus indicating possible

disparities in binding mechanism between chromium oxidation states. However,

at high initial metal concentrations F. vesiculosus had the greatest removal

efficiency for Cr (III) and performed almost identically to P. lanosa in terms of Cr

(VI) removal. The Langmuir Isotherm mathematically described chromium

112
Chapter – III Literature Review

binding to the seaweeds where F. vesiculosus had the largest qmax for Cr (III)

sorption (1.21 mmol/g) and P. lanosa had the largest Cr (VI) uptake (0.88

mmol/g). P. palmata had the highest affinity for both Cr (III) and Cr (VI) binding

with b values of 4.94 m/M and 8.64 m/M, respectively.

Murugan and Subramanian [75] investigated Cupressus Female Cone (CFC),

an inexpensive plant material, as an adsorbent for the removal of Cr (VI) from

synthetic solution as well as from industrial wastewater so as to help its safe

disposal. Batch sorption studies were conducted, involving various process

parameters such as pH, agitation time, initial Cr (VI) concentration, particle size

and adsorbent dosage. Maximum sorption occurred at the acidic pH range of 0.2-

0.5 but it decreased on increasing the pH. Increase in initial concentration of Cr

(VI) and adsorbent particle size was found to reduce the amount of sorption,

demonstrating the role of surface effects on sorption. When the temperature was

raised the amount of sorption increased, exhibiting an endothermic nature. The

sorption capacity was remarkably higher, amounting 119.4 mg/g (91 per cent) at

30 ◦C and at a pH 1 to 5; it further increased to 124.5 mg/g (99.6 per cent) at 50◦C.

Sorption process followed first order kinetics and the data fitted to the Langmuir

adsorption isotherm.

Nagesh and Krishnaiah [76] developed a method for removal of chromium from

water effluents by adsorption process on different carbon materials. In order to

understand the sorption behavior of chromium and also to evaluate the extent of

reduction from hexavalent chromium to trivalent chromium, a number of batch

experiments were conducted on each adsorbent-adsorbate system at various pH

113
Chapter – III Literature Review

values (pH 1-6). The results illustrated that the reduction was maximum at pH 1

and adsorption was maximum when the values of pH between 2 and 3 depending

upon the adsorbent. The results were interpreted on the basis of interaction

between adsorbent and adsorbate.

Narsi R. Bishnoi et al [77] have studied on biosorption of Cr (III) from aqueous

solution using algal biomass Spirogyra spp. In the present investigation, a fresh

water green algae Spirogyra spp. was used as an inexpensive and efficient

biosorbent for Cr (III) removal from aqueous solution. The effects of various

physico-chemical parameters were studied, e.g. pH 3.0–6.0, initial metal ions

concentration 20–150 mg/L, algal dose 1.0–3.0 g/ L, and contact time 15–180

min, respectively. Biosorption of Cr (III) is highly pH dependent. Maximum

81.02 % adsorption of Cr (III) was observed with 0.2 M CaCl2 treated biomass at

pH 5.0. Removal of Cr (III) was more than 70 % in 45 min of contact time with

different treated and untreated algal biomass at concentration 30 mg/L. Maximum

metal uptake (Qmax) was observed as 30.21 mg/g with 0.2 M CaCl2 treated algal

biomass indicate good biosorbents than other treated and untreated biomass.

Neetu Tewari et al [78] have made a study on biosorption of Cr (VI) by Mucor

hiemalis. The effect of initial Cr (VI) concentration was studied in the range 10–

600 mg/L. Different amounts of sorbent were added to vary the sorbent to solute

ratio in the range of 20–100 g/g. The pH of the solution was varied from 1.0 to

8.0. The highest Cr (VI) uptake of 53.5 mg/g at an initial pH of 2.0. The

activation energy of the biosorption (Ea) was estimated as 4.0 kJ/mol using

Arrhenius equation.

114
Chapter – III Literature Review

Ozgur Dogan Uluozlu et al [79] have studied on biosorption of Pb (II) and Cr

(III) from aqueous solution by lichen (Parmelina tiliaceae) biomass. Throughout

the study, the contact time was varied from 5 to 120 min, the pH from 2 to 8, the

initial metal concentration from 25 to 400 mg/L, and the biosorbent dosage from

0.8 to 20 g/L. The monolayer biosorption capacity of P. tiliaceae biomass for Pb

(II) and Cr (III) ions was found to be 75.8 mg/g and 52.1 mg/g, respectively.

From the D–R isotherm model, the mean free energy was calculated as 12.7kJ/mol

for Pb (II) biosorption and 10.5 kJ/mol for Cr (III) biosorption, indicating that the

biosorption of both metal ions was taken place by chemical ion-exchange.

Pérez Marín et al [80] have studed on biosorption of chromium (III) by orange

(Citrus cinensis) waste. Batch kinetic and isotherm studies were carried out on a

laboratory scale to evaluate the adsorption capacity of orange waste. The effects

of particle size, adsorbent dose and solution pH on Cr (III) removal were also

studied. The results showed that the higher the adsorbent dosage and the pH, the

higher the percentage of metal removal. Equilibrium assays displayed a

maximum sorption capacity ranging from 0.57 mmol/g to 1.44 mmol/g when the

pH increased from 3 to 5, according to the Sips model, which along with the

Redlich–Peterson equation, is very suitable for correlating equilibrium data.

Prakasham et al [81] have studied on biosorption of chromium VI by free and

immobilized Rhizopus arrhizus. Biosorption of chromium (VI) was studied by

using non-living free and immobilized biomass of Rhizopus arrhizus at pH 2. A

biphasic chromium adsorption pattern was observed in all experimental

115
Chapter – III Literature Review

conditions. Chromium removal rate was slightly more in free biomass conditions

over immobilized state. Stirred tank reactor studies indicated maximum

chromium biosorption at 100 rpm and at 1:10 biomass ± liquid ratio. Fluidized

bed reactor is more efficient in chromium removal over stirred tank reactor.

Immobilization of biomaterial has a little effect on chromium biosorption by this

R. arrhizus biomass.

Prasad and Freitas [82] investigated the potential of Quercus ilex phytomass

from stem, leaf and root as an adsorbent of chromium (Cr), nickel (Ni), copper

(Cu), cadmium (Cd) and lead (Pb) at ambient temperature. The metal uptake

capacity of the root for different metals was found to be in the order:

Ni>Cd>Pb>Cu>Cr; stem Ni>Pb>Cu>Cd>Cr; and leaf Ni>Cd>Cu>Pb>Cr. The

highest amount adsorbed was Ni (root>leaf>stem). Data from this paper

demonstrated that Ni was sequestered mostly in the roots, where concentrations

could be as high as 428.4 ng/g dry wt., when I-year-old seedlings were treated

with Ni (2000 mg/L) in pot culture experiments, compared to 7.6] mg/g dry wt.,

control (garden and greenhouse soil) topsoil where Ni was present in trace

amounts. This proved that the root biomass of Q. ilex has the capacity for

complexing Ni. Cr exhibited the least adsorption values for all the three types of

phytoma$S compared to other metals. The trend of adsorption of the phytomass

was similar for Ni and Cd, i.e. root>lcaf>stem. Desorption with 10 mM Na (4)

EDTA was effective (55-90 %) and hence, there exists the possibility of recycling

the phytomass. The biosorption results of recycled phytomass suggested that the

selected adsorbents were re-usable. The advantages and potential of the Q. ilex

116
Chapter – III Literature Review

phytomass as a biofilter of toxic trace metals. The scope and need for enhancing

the efficiency of the Q. lilex phytomass as an adsorbent of metals were presented.

Preetha et al [83] have studied on batch and continuous biosorption of chromium

(VI) by Rhizopus arrhizus. This was studied by evaluating the physicochemical

parameters of the solution such as initial chromium ion concentration in both

batch and packed bed reactor. Batch experiments demonstrate that the sorption

process corresponds to the second-order kinetic model and the kinetics of sorption

indicates the process to be diffusion controlled. The diffusivity of chromium in

Rhizopus-alginate gel beads was calculated using the shrinking core model, and

the diffusivity values were in the ranges of 0.049 × 10−5 to 0.521 × 10−5 cm2/s.

Quintelas et al [84] have studied on biosorption of Cr (VI) by three different

bacterial species supported on granular activated carbon. The results showed

metal uptake values of 5.82, 5.35 and 4.12 mg/g bios, respectively, for S.

equisimilis, B. coagulans and E. coli, for an initial metal concentration of 100

mg/L. In the same order and for the initial concentration of 50 mg/L, metal

uptake values were 2.33, 1.98 and 3.60 mg/g bios. Finally, for the initial metal

concentration of 10 mg/L, those values were, respectively, 0.66, 1.51 and 1.12

mg/g bios. Studies made with an industrial effluent, with the aim of testing these

biofilms in a real situation, showed values of Cr uptake of 0.083, 0.090 and 0.110

mg/g bios, respectively, for S. equisimilis, B. coagulans and E. coli, for an initial

concentration of 4.2 mg/L of total Cr. The quantification of polysaccharides,

playing a key role in the whole process, was made and it was concluded that the

117
Chapter – III Literature Review

production of polysaccharides is higher for B. coagulans followed by S.

equisimilis and E. coli (9.19, 7.24 and 4.77 mg/g bios.).

Quintelas et al [85] have studied on biosorption of Cr (VI) by a Bacillus

coagulans biofilm supported on granular activated carbon (GAC). The ability of a

biofilm of Bacillus coagulans supported on granular activated carbon (GAC) to

biosorb Cr (VI) was investigated in batch and column studies so it may be applied

to low metal concentration wastewater treatment. The quantification of

polysaccharides and polymeric net revealed a value of 9.19 mg/g biosorbent for

the polysaccharides and 75 mg/g biosorbent, for the polymeric net. The results

obtained with open systems showed uptake values of 1.50, 1.98 and 5.34

mg/gbiosorbent, respectively, for initial concentrations of 10, 50 and 100 mg/L of

Cr (VI). Column studies performed with an industrial effluent showed values of

Cr uptake of 0.090 mg/gbiosorbent, for an initial concentration of 4.2 mg/L. The

presence of functional groups on the cell wall surface of the biomass that may

interact with the metal ion, was confirmed by FTIR. The equilibrium studies in

batch systems were described by Freundlich, Langmuir, Reddlich–Peterson,

Dubinin–Radushkevich, Sips and Toth model isotherms. Best fit was obtained

with Toth model isotherm.

Rabbani et al [86] have studied on biosorption of chromium (III) by new

bacterial strain (NRC-BT-2). The effect of initial concentration of the metal

solution in the range of 10–250 ppm pH from 1-4 and contact time 5-150 min has

been studied, and the maximum removal occurs in 10 ppm, in pH 4 and 1 h of

118
Chapter – III Literature Review

contact time (100 %). Sodium acid, high temperature and g-ray did not affect the

biosorption.

RadojkaRazmovski et al [87] have studied on biosorption of Cr (VI) and Cu (II)

by waste tea fungal biomass. Batch experiments were carried out with live and

dried biosorbent samples (both 0.5 g/L of dry matter) at 25 oC in Erlenmeyer

flasks on a rotary shaker at 200 rpm in the pH range 2.0–8.0 at various sorbent

dosage, ranging from 0.15 to 4.0 g/L of DM. the optimum pH values for Cr (VI)

and Cu (II) were 2.0 and 4.0, respectively. Itwas observed that the percentage of

metal ions removal increased with increase in biosorbent dosage.

Rajender Kumar et al [88] have stdied on biosorption of chromium (VI) from

aqueous solution and electroplating wastewater using fungal biomass. The

influence of pH of the solution, biosorbents dose, concentration of ions and

contact time on biosorption capacity of Cr (VI) ions was studied. The optimum

pH for biosorption of Cr (VI) ions was found to be 2.0. The removal of Cr (VI)

was 91.03% with A. niger at biosorbent dose 0.6 g/50 mL, whereas, 87.95 % and

86.61 % with A. sydoni and P. janthinellum at biosorbent dose 0.8 g/50mL but

uptake capacity (mg/g) of Cr (VI) ions decreased with increased biosorbent dose.

Initially percent removal of Cr VI) ions from solution was increased with increase

in concentration from 10 to 30 mg/L and maximum percent removal was observed

at concentration 30 mg/L after that percent removal decreased. Whereas, uptake

capacity was increased with increase in concentration of Cr (VI) ions from 10 to

60 mg/L. Uptake rate of Cr (VI) increased from 1.72 to 2.39 mg/g with A. niger,

1.22 to 1.76 mg/g with A. sydoni and 1.18 to 1.77 mg/g with P. janthinellum with

119
Chapter – III Literature Review

increases time from 15 to 120 min. Removal of Cr (VI) from electroplating

wastewater was observed less than from synthetic solution.

Ringqvist ct al [89] determined metal adsorption and surface charge of well

characterized Sphagnum and Carex peat samples. The aim of this investigation

was to determine metal adsorption from complex wastewaters onto these peat

samples and compared it to the adsorption onto peat granules, clinoptilolite,

glauconite and a flue dust from steel production. A sulphide mine leachate, a

landfill leachate and a laundry wastewater were chosen and was given a variation

in pH, ionic strength, total organic carbon and concentrations of metals. Metal

adsorption was determined in batch and column experiments. The wastewater

composition was of great importance for metal removal efficiency, mainly due to

the difference in dominating metal species. In the sulphide mine leachate,

containing free metal ions, a high metal adsorption was observed onto both peat

and inorganic adsorbents. In the landfill leachate the metals formed carbonate and

organic complexes and a low metal removal was achieved. Contrary to the

leachates, the laundry wastewater contained suspended particles. The high

amount of metals removed, 80 % of the Cu and 30-60 % of the Zn concentration,

was probably withdrawn bound to the particle fraction. The highest removal of

metal ions was obtained in the sulphide mine leachate with Carex peat, removing

97-99 % of the Zn and 85-100 % of the Cu content. The Sphagnum peat sample

removed 37-77 % of the Zn and 80-100 % of the Cu content. The differences

found between Sphagnum and Carex peat were attributed to the original chemistry

of the plant material and the habitat conditions at. the time of peat formation.

120
Chapter – III Literature Review

Generally, the combination of glauconite or clinoptilolite with the peat samples in

column experiments gave a minor improvement in metal removal.

Rio ct at [90] tested the fly ashes obtained from two fluidized bed power plants to

Pb2+ ,Cu2+,Cr3+, Ni2+, Zn2+ and Cr6+ from aqueous solutions. Experimental design

methodology was used to study the removal and the leaching as a function of (i)

the water pollutant content, (ii) the metal concentration in water, (iii) the pH of the

solution and (iv) the addition of lime to fly ashes. The results showed that the

percentage of adsorbed ions was greater when they were in contact with silica-

aluminous fly ashes than silico-aluminous fly ashes, except in the case of the ion

Ni2+. The removal of metallic ions increased with increasing pH. The authors

compared the adsorption capacity of fly ashes with activated carbon or low cost

adsorbents by carrying out similar tests and concluded that adsorption onto fly ash

was an interesting alternative.

Sandra Lameiras et al [91] have studied on biosorption of Cr (VI) using a

bacterial biofilm supported on granular activated carbon and on zeolite. In this

study Two mini-columns partially filled with granular activated carbon (GAC)

and/or a natural zeolite, covered by a bacterial biofilm of Arthrobacter viscosus,

were used in a continuous flow system to remove Cr (VI) from solutions with

initial concentration of 70 mg/L and a working pH ranging between 4.5 and 5.5.

Comparatively, the biosorption system supported on GAC reaches similar removal

values, 19 %, as the one supported on the zeolite, 18 %, but when these two beds

are used in combination better performances are reached, i.e. 42% removal. The

121
Chapter – III Literature Review

maximum uptake values ranged from 0.57 mg/g Cr Support to 3.58 mgCr/g

support.

Saravananc et al [92] presented a comparative study of removal efficiency of

heavy metals [Chromium—Cr (VI), Copper—Cu (II), Manganese—Mn (II),

lron—Fe (Il), Nickel—Ni (Il), Lead—Pb (II) and Zinc—Zn (II)] from aqueous

solution by adsorption on non-conventional materials and on chemically activated

non-conventional materials. It was found that the adsorption potential varied as a

function of contact time, concentration, particle size, pH and flow rate. Of all the

low cost adsorbents used in the study, saw dust was found to possess greater

adsorption efficiency for all metals, than rice husk under identical experimental

conditions. Chemically activated saw dust could remove 98.28% of Cu (II); 100%

of Mn (II); 96.72 % of Fe (II); 96.72 % of Cd (II); 75 % of Cr (VI); 80 % of Ni

(Il); 95 % of Pb (II) and 93 % of Zn (II), from the corresponding metal solution

taken one at a time.

Schmuhl et al [93] studied the ability of chit os an as an adsorbent for Cu (II) and

Cr (VI) ions in aqueous solution. The experiments were done as batch processes.

Equilibrium studies were done on both cross-linked and non-cross-linked chitosan

for both metals. Cr (VI) adsorption behavior could be described using the

Langmuir isotherm over the whole concentration range of 10 to 1000 mg/L Cr.

The maximum adsorption capacity for both types of chitosan was found to be 78

mg/g for the non-cross-linked chitosan and 50 mg/g for the cross-linked chitosan

for the Cr (VI) removal. For the Cu (II) removal the Freundlich isotherm

described the experimental data over the whole concentration range of 10 to 1000

mg/L Cu (II). The maximum adsorption capacity for both types of chitosan can

122
Chapter – III Literature Review

be estimated to be greater than 80 mg/g. Cr (VI) removal was the highest at pH 5

but pH did not have a large influence on Cu (II). From these results it is clear that

the adsorption of heavy metals is possible with chitosan, but that with this method,

end concentrations of below 1mg/L can hardly be obtained.

Semra Ilhon et al [94] investigated the selective biosorption of chromium, lead

and copper ions from industrial wastewaters by microorganisms. Microorganisms

were isolated from soil and in this research a bacterium that is identified as

staphylococcus saprophyticus was used. The effects of pH, temperature and

initial concentration of metal ions on the hiosorption capacity were investigated.

The optimum pH values for chromium, lead and copper biosorption were found to

be 2.0, 4.5 and 3.5 respectively. The maximum adsorption was observed for Cr6+,

Pb2+ and Cu2+ at the initial concentrations of 193.66 mg/L Cr6+; 100 mg/L Pb6+

and 105 mg/L Cu6+ and under these conditions the biosorption values were found

to be 88.66 mg/L Cr6+, 100 mg Pb6+ / L and 44.94 mg Cu6+ / L, respectively. The

results indicated that saprophytic us was suitable for biosorption of lead and

chromium from wastewaters.

Shaik Basha et al [95] have studied on biosorption of hexavalent chromium by

chemically modified seaweed, Cystoseira indica. The sorption of hexavalent

chromium by marine brown algae Cystoseira indica, which was chemically-

modified by cross-linking with epichlorohydrin (CB1, CB2), or oxidized by

potassium permanganate (CB3), or only washed by distilled water (RB) was

studied with variation in the parameters of contact time, pH, initial metal ion

concentration and solid/liquid ratio. The adsorption characteristics of Cr (VI) with

123
Chapter – III Literature Review

various pH in the range of 1.0–5.0 are studied at initial concentration of 30 mg/L.

The highest uptake capacity of Cr (VI) is 22.7 and 24.2 mg/g for CB1 and CB2,

respectively, at pH 3.0. The biosorption of Cr (VI) was carried out at different

initial Cr (VI) ion concentrations ranging from 10 to 100 mg/L at pH 3.0. The

uptake of Cr (VI) by all the four sorbents gave a plateau at 30–100 mg/L showing

the saturation of binding sites at higher concentration levels. The maximum

amounts of Cr (VI) adsorbed are 22.7, 24.2, 20.1 and 17.8 mg/g, respectively for

CB1, CB2, CB3 and RB.

Shailendra Mishra et al [96] have studied on Novel chromium tolerant

microorganisms: Isolation, characterization and their biosorption capacity. A total

of 29 chromium tolerant isolates from solid waste and 39 isolates from liquid

effluent were obtained. At a maximum chromium concentration of 700 mg/L,

three organisms from solid waste and six from liquid effluent were isolated. It is

seen from the kinetics that chromium concentration is lowered to less than 3 % of

200 mg/L concentration in about 20 h. The initial rate of this biological removal

of chromium with most active organism isolated from solid waste (solid 1) and

liquid effluent (liquid 4) are 90.4 and 89.9 mg/L/h, respectively. The removal of

90 % of 200 mg/L of chromium in 5 h and more than 95 % in about 20 h has been

observed in the current study.

Shrivastava and Thakur [97] reported the bio absorption potential of

Acinetobacter sp. Strain 1ST 103 of bacterial consortium for removal of

chromium from tannery effluent. A mixed microbial community obtained from

pulp and paper mill and tannery effluent was enriched in a chemostat in the

124
Chapter – III Literature Review

presence of potassium chromate as sole source of carbon and energy. The

microbial community of pulp and paper mill indicated significant reductien of

chromium (70 per cent) in comparison to tannery community (60 %) in batch

culture in the presence of potassium chromate and sodium acetate as additional

carbon source. The community comprised three bacterial strains, identified as CP

1, CP2 and CP3, were further applied for reduction of chromium indicated higher

removal of chromium (68 %) by strain CP3 followed by CP2 and CPl. The strain

CP3 was identified as Acinetobacter sp. by biochemical test. The bacterial strain,

CP3, applied for the removal of chromium in tannery effluent in a sequential

bioreactor indicated the reduction of chromium (80 %).

Sibel Tunali et al [98] have studied on chromium (VI) biosorption characteristics

of Neurospora crassa fungal biomass. The effect of pH on the biosorption rate of

Cr (VI) on fungal biomass was investigated by equilibrating the sorption mixture

containing dried biomass (0.1 g) and Cr (VI) solutions (50 ml of 100 mg/L) in the

pH range 1.0–6.0. The greatest capacity of heat inactivated N. crassa biomass

was obtained at pH 1.0 using Cr (VI) concentration of 100 mg/L at 25 ± 0.1 ◦C.

biosorption capacity of acetic acid pretreated biomass was found to be 15.85 

0.94 mg/g biomass under optimum conditions. The adsorption constants were

found from the Freundlich isotherm model at 25 ◦C. The biosorbent was

regenerated using 10 mM NaOH solution with up to 95 % recovery and reused

five times in biosorption–desorption cycles successivelly.

Srinath et al [99] have studied on chromium (VI) biosorption and accumulation

by chromate resistant bacteria. In this study, strains that are capable of

125
Chapter – III Literature Review

bioaccumulating Cr (VI) were isolated from treated tannery effluent of a common

effluent treatment plant. The growth response curve of strains H and A at various

concentrations of Cr (VI) (0, 25, 50, 100 and 150 mg/L) showed that the lag phase

as well as the optical density attained by the strains depended greatly on the

concentration of Cr (VI) in the medium. At the concentration of 100 mg/L Cr

(VI) and above, the lag phase was extended and the maximal cell density reduced

below 90 % of the control. Two strains, identified as Bacillus circulans and

Bacillus megaterium were able to bioaccumulate 34.5 and 32.0 mg/g Cr dry

weight, respectively and brought the residual concentration of Cr (VI) to the

permissible limit in 24 h when the initial concentration was 50 mg/L Cr (VI).

Sudha Bai et al [100] have made Studies on enhancement of Cr (VI) biosorption

by chemically modified biomass of Rhizopus nigricans. For biosorption

experiments, Cr solution having 50–500 mg/L was prepared and used. Treatment

of the biosorbent with mild alkalies (0.01 N NaOH and ammonia solution) and

formaldehyde (10 %, w/v) deteriorated the biosorption efficiency. However,

extraction of the biomass powder in acids (0.1N HCl and H2SO4), alcohols (50 %

v/v, CH3OH and C2H5OH) and acetone (50 %, v/v) improved the Cr uptake

capacity. Biomass modification experiments conducted using Cetyl Trimethyl

Ammonium Bromide (CTAB), Polyethylenimine (PEI), and Amino Propyl

Trimethoxy Silane (APTS) improved the biosorption efficiency to exceptionally

high levels. The FTIR spectroscopic analysis of the native, Cr bound and the

other types of chemically modified biomass indicated the involvement of amino

groups of Rhizopus cell wall in Cr binding.

126
Chapter – III Literature Review

Sujoy K. Das et al [101] have studied on biosorption of hexavalent chromium by

Termitomyces clypeatus biomass: Kinetics and transmission electron microscopic

study. Biosorption of Cr+6 by Termitomyces clypeatus has been investigated

involving kinetics, transmission electron microscopy (TEM) and Fourier

transform infrared spectroscopic (FTIR) studies. Kinetics experiments reveal that

the uptake of chromium by live cell involves initial rapid surface binding followed

by relatively slow intracellular accumulation. Of the different chromate analogues

tested, only sulfate ion reduces the uptake of chromium to the extent of 30 %

indicating chromate ions accumulation into the cytoplasm using sulfate transport

system. Metabolic inhibitors, e.g. N,N-dicyclohexylcarbodiimide, 2,4-ditrophenol

and sodium azide inhibit chromate accumulation by 30 % in live cell. This

indicates that accumulation of chromium into the cytoplasm occurs through the

active transport system.

Sujoy K. Das et al [102] have studied on biosorption of chromium by

Ermitomyces clypeatus. The influence of hydrogen ion concentration on the

sorption of chromium was monitored at pH values 2.0–7.0 by suspending 0.2 g of

dry live TC biomass in 25 mL 50 mM acetate buffer containing either 100 mg/L

Cr6+ or Cr3+, varying Cr6+ concentration from 10 to 1000 mg/L. The sorption of

hexavalent chromium by live TCB depends on the pH of the solution, the

optimum pH value being 3.0. The process follows Langmuir isotherm (regression

coefficient 0.998, χ2-square 5.03) model with uniform distribution over the

surface which gets strong support from the X-ray elemental mapping of chromium

adsorbed biomass. Desorption and FTIR studies also exhibited that Cr6+ is

reduced to trivalent chromium on binding to the cell surface. The level of

127
Chapter – III Literature Review

chromium concentration present in the effluent of tannery industries’ is reduced to

a permissible limit using TCB as adsorbent.

Vieira et al [103] have studied on biosorption of chromium (VI) using a

Sargassum sp. packed-bed column. The objective of the study was to examine

chromium (VI) removal from an aqueous solution using a packed-bed column

with Sargassum sp. algae as a biosorbent. The kinetic and equilibrium results for

a biomass of 4 g and a flow rate of 0.015 mL/s for several initial chromium

concentrations from 200 to 1240 mg/L and different tme intervals from 0 to 210

min. The capacity of removal obtained at optimum conditions was 19.06 mg of

metal/g biosorbent.

Vinoda et al [104] have studied on biosorption of nickel and total chromium from

aqueous solution by gum kondagogu (Cochlospermum gossypium) The optimum

conditions of biosorption were determined by investigating pH, contact time, and

initial metal ion and biosorbent concentrations. The maximum biosorption

capacity of gum kondagogu as calculated by Langmuir model were found to be

50.5 mg/g for nickel at pH 5.0 ± 0.1 when the range is maintained from (2.0–6.0)

and 129.8 mg/g for total chromium at pH 2.0 ± 0.1 when the range is maintained

from (2.0–6.0), respectively. It was observed the adsorption capacity of nickel

(Ni2+) and chromium by gum kondagogu was increased as the contact time

increased and the equilibrium was reached after 120 min. The result shows that

with the increase in the gum kondagogu concentration, the metal removal

efficiency also increased. The % metal biosorption increased from 55 ± 1.4 to

82.2 ± 2.8 % for Ni2+ and 58.9 ± 1.54 to 87.5 ± 3.05 % for total chromium.

128
Chapter – III Literature Review

However, the adsorption capacity decreased from 55 ± 1.4 to 16.4 ± 0.45 mg/g for

Ni2+ and 58.9 ± 1.54 to 7.5 ± 0.49 mg/g for total chromium, when gum kondagogu

concentration was increased from 1000 to 5000 mg/L.

Xu Han et al [105] have studied on biosorption and bioreduction of Cr (VI) by a

microalgal isolate, Chlorella miniata. Cr (VI) removal was determined under

different initial pH (from 1.0 to 4.0) and biomass concentrations (from 1.0 to 5.0

g/L) at an initial Cr (VI) concentration of 100 mg/L. The biosorption capacity of

total Cr [Cr (III) and Cr (VI)] reached the maximum at initial pH of 3.0. The

spectrum of Fourier Transform Infrared Spectrometer analysis (FTIR) further

confirmed that amino group on the algal biomass was the main adsorption site for

Cr (VI) biosorption in acidic pH while the reduced Cr (III) was mainly

sequestered by carboxylate group.

Yasemin S ahin et al [106] have studied on biosorption of chromium (VI) ions

from aqueous solution by the bacterium Bacillus thuringiensis. Batch

experiments were done as a function of pH, initial metal ion concentration and

temperature. When the initial chromium (VI) ions concentration varied from 25.0

to 250.0 mg/L, the loading capacity of B. thuringiensis’vegetative cell has

increased from 14.3 to 54.7 mg/L while its spore–crystal mixture’s uptake has

increased from 20.6 to 61.5 mg/L. The optimum pH value for the biosorption of

chromium (VI) ions was determined as 2.0 for both of the spore– crystal mixture

and the vegetative cell. In this work, the maximum initial adsorption percentage

yields were found to be 38.3 % for the vegetative cells and 59.3 % for the spore–

crystal mixture at 25 oC Chromium (VI) ions uptake of B. thuringiensis’ spore–

129
Chapter – III Literature Review

crystal mixture at 250 mg/L was 24.1 %, whereas its vegetative cell metal uptake

was 18.0 %. Chromium (VI) biosorption experiments were carried out at three

different temperatures, 15, 25 and 35 0C. The best temperature for biosorption

was 25 0C.

Yasmin Khambhaty et al [107] have studied on kinetics, equilibrium and

thermodynamic studies on biosorption of hexavalent chromium by dead fungal

biomass of marine Aspergillus niger. The present study was carried out in a batch

system using dead biomass of marine Aspergillus niger for the sorption of Cr (VI).

The removal rate of Cr (VI) was increased with a decrease in pH and an increase

in Cr (VI) and biomass concentration. A. niger exhibited the highest Cr (VI)

uptake of 117.33 mg/g of biomass at pH 1.0 in the presence of 400 mg/L Cr at

50oC.

Yetis et al [108] determined the heavy metal biosorption potentials of two white-

rot fungi, Poloporous versicolor and Phanarochaete chrysosporium, which were

commonly used in wastewater treatment. Biosorption studies were performed for

Cu (II), Cr (III), Cd (II), Ni (II) and Pb (II) at the same operational conditions and

the effectiveness of both fungi at removing these heavy metals was compared. It

was found that both P. versicolor and P. chrysosporium were the most effective in

removing Pb (II) from aqueous solutions with biosorption capacities of 57.5 and

110 mg Pb (II)/g dry biomass, respectively. With P.versicolor, the adsorptive

capacity order was determined to be Pb (II) > Ni (II) > Cr (II) > Cd (II) > Cu (II)

whereas the order was Pb (II) > Cr (III) > Cu (II) = Cd (II) > Ni (II) with P.

chrysosporium. It was observed that, as a general trend, the metal removal

130
Chapter – III Literature Review

efficiency with these fungi decreased as the initial metal ion concentration

increased.

Ziagova et al [109] have worked on Comparative study of Cd (II) and Cr (VI)

biosorption on Staphylococcus xylosus and Pseudomonas sp. in single and binary

mixtures. In tis study Langmuir and Freundlich models were applied to describe

metal biosorption and the influence of pH, biomass concentration and contact time

was determined. Biosorption experiments were conducted at initial Cd (II)

concentrations from 10 to 1000 mg/L. The effect of pH was investigated in the

range of 3.0–7.0. Maximum uptake capacity of cadmium was estimated to 250

and 278 mg/g, whereas that of chromium to 143 and 95 mg/g for S. xylosus and

Pseudomonas sp., respectively. In binary mixtures with Cd (II) ions as the

dominant species, there is a profound selectivity for cadmium biosorption,

reaching 96 % and 89 % for Pseudomonas sp. and S. xylosus, respectively, at 10

mg/L Cd (II) and 5 mg/L Cr (VI). Interesting, when chromium (VI) ions are the

dominant species, there is selectivity towards chromium around 92 % with S.

xylosus only.

Zumriye Aksu et al [110] have studied on biosorption of chromium (VI) ions by

Mowital®B30H resin immobilized activated sludge in a packed bed: comparison

with granular activated carbon. The effect of operating parameters such as flow

rate and inlet metal ion concentration on the sorption characteristics of each

sorbent was investigated in a continuous packed bed column. From the batch

system studies the working sorption pH value (1-7) was determined as 1.0 for both

sorbents and packed bed sorption studies were performed at this pH value.

131
Chapter – III Literature Review

Saturation was reached within 210-/1800 min at different flow rates from 0.8 to

3.2 mL/min and prolonged exposure time did not increase removal of chromium

(VI). The sorption breakthrough curves obtained by changing inlet chromium

(VI) concentration from 50 to 500 mg/L at 0.8 ml/min flow rate. At 50 mg/L of

inlet chromium (VI) concentration and at the lowest flow rate of 0.8 ml/min, the

sorption, although continuous with time, was very efficient in the initial steps of

the process.

Zumriye Aksu et al [111] have studied on single and binary chromium (VI) and

Remazol Black B biosorption properties of Phormidium sp. In this study

biosorption of chromium (VI) and Remazol Black B reactive dye by dried

Phormidium sp., a thermophilic cyanobacterium, was studied as a function of

initial chromium (VI) concentration and temperature in no dye and 100mg/L dye

containing media at an initial pH value of 2.0 at which the biomass exhibited the

maximum chromium (VI) and dye uptakes. Experiments were performed at five

different initial pH values ranging from 1.0 to 6.0 in single 100 mg/L chromium

(VI) and in single 100 mg/L dye-containing media. The highest uptake values

were found at pH 2.0 for both situations tested. The adsorption of chromium (VI)

and dye has been studied over a range of 25–45 oC. At 25 oC, 22.8 mg/g

chromium (VI) and 91.3 mg/g dye were sorbed by the biomass in binary 100mg/L

chromium (VI) and 100 mg/L dye-containing medium. At 25 oC, on changing the

initial chromium (VI) concentration from 10 to 100 mg/L, the amount of

biosorbed chromium (VI) increased from 3.5 to 15.2 mg/g and from 5.1 to 22.8

mg/g in the absence and in the presence of 100 mg/L dye, respectively.

132
Chapter – III Literature Review

3.2 LEAD (Pb):

Adil Hammaini et al [112] investigated the effect of the presence of Pb on the

biosorption efficiency of Cu, Cd and Zn by activated sludge was investigated. The

interference of Pb with the Zn uptake was very pronounced. The amount of Zn

adsorbed [Cf [Zn] = 1mM] when Pb was present at 0.2 mM was 87 % less than

when Cf [Pb] = 0.05 mM. At Cf [Pb] = 0.2 mM and when 0.5 and 2 mM of Zn

were present in the system (final equilibrium Zn concentration), the Pb uptake was

0.5 and 0.48 mmol/g, respectively. When the residuals of Pb and Zn were the

same,about 99 % of the total metal uptake was due to Pb uptake. The equilibrium

concentration of Zn would have to be 70 times greater than that of Pb to obtain the

same proportion of uptake for each metal. When the residuals of Pb and Cd were

the same, about 88 % of the total metal uptake was due to Pb.

Ahmet Cabuk et al [113] studied the Pb (II) biosorption properties of

immobilized Bacills sp.ATS-2 cells were investigated in the fixed bed column and

batch system. The maximum Pb (II) biosorption capacity of the immobilized

biosorbent was obtained at the initial pH of 4.0.The effect of initial pH on the

biosorption of Pb (II) onto silica gel-immobilized Bacillus sp. ATS-2 was

investigated at a constant flow rate of 180 ml/h, fixed bed column with the bed

length of 10 cm and the initial Pb (II) concentration of 100 mg/L. The lower

biosorption yield under the initial pH values of 4.0 has been attributed to the

competition of the metal ions with the protons for the available binding sites on

the immobilized biosorbent. The FTIR spectra of the immobilized biosorbent

before and after Pb (II) sorption in the range of 400–4000 cm−1 were taken and

133
Chapter – III Literature Review

compared with each other to find out which functional groups are responsible for

the Pb (II) biosorption

Ahmet Cabuk et al [114] suggested that the S. cerevisiae immobilized on cone

biomass of P. nigra appears as a low cost possible biosorbent and to be used for

treatment of Pb (II) bearing solutions. The maximum biosorption capacity was

determined to be 30.04 mg/g at pH 5.0, 2.0 g/L biosorbent dosage and 30 min.

The experimental data were evaluated by Langmuir, Freundlich and Dubinin–

Radushkevich isotherms and fitted well to all of the isotherm models with good

regression coefficients. The Pb (II) biosorption capacity of biosorbent increased

with increasing pH from 1.0 to 5.0. It is well known that, at low pH values, cell

wall ligands were closely associated with the hydronium ions and restricted the

approach of positively charged metal ions as a result of the repulsive force. The

mechanism of the process was evaluated by FT-IR and EDAX analysis.

Alaa H. Hawari et al [115] investigated the feasibility of anaerobic granules as a

novel type of biosorbent, for lead, copper, cadmium, and nickel removal from

aqueous solutions. Over the pH range 4–5.5, pH-related effects were not

significant. Meanwhile, at the pH values of 3.5 and 3, the q values started to

decrease. At low pH, protons would compete for active binding sites with metal

ions. The protonation of active sites thus tends to decrease the metal sorption. At

a low pH, of almost 2.0, all the binding sites may be protonated, thereby desorbing

all originally bound metals from the biomass. Water was the most abundant

single compound in the cell and it makes almost 70 % of the total weight of the

cell. Inorganic salts and mineral elements were found to be 25 % of the total dry

134
Chapter – III Literature Review

weight (TFS). Organic matter (TVS), on the other hand, constitute 75 % of the

solid fraction of the biomass.

Ankit Balaria et al [116] investigated the study on biosorption of lead (Pb) from

aqueous solution using citrus peels can provide an efficient and cost-effective

solution for lead removal from industrial wastewaters. These peels contain the

biopolymer pectin that has a strong affinity formetal ions. A plateau can be

observed for pH 6–9 at surface charge values of 0.8 meq/g for LM pectin and 0.4

meq/g for HM pectin. LM pectin contains fewer methoxylated carboxylic groups

and more free carboxylic groups than HM pectin, which explains the higher

negative charge for LM pectin.For the LM pectin, 91 % of the carboxyl groups

should be free (based on a methoxyl content of 9 %) in contrast to 36 % of the

HM pectin. If only free carboxyl groups contribute to the charge, one would

expect the LM pectin to have a charge 2.5 times higher than the HM pectin, which

is similar to the observed factor of two. The FTIR spectra comparisons of pure

versus Pb–laden pectins reveal that the absorption peak corresponding to the

carboxylate (–COO−) groups shifted significantly from 1627 cm−1 to 1636 cm−1 in

the case of the HM pectin after treatment with Pb. A similar shift in the same

peak from 1624 cm−1 to 1635 cm−1 can be observed for LM pectins. This shift in

wave number corresponds to a change in bonding energy corresponding to the

functional group, which further reflects that the bonding pattern of carboxylate

groups changes after biosorption. This result confirmed the involvement of

carboxylic acid groups in binding of Pb in the case of pectin.

135
Chapter – III Literature Review

Anushree Malik et al [117] observed a set of studies that demonstrated the

feasibility of employing rotating biological contactor (RBC), supporting the

immobilized growing bio films for removal of Cu, Zn and Cd from synthetic

wastewater. The inoculum consisted of the enrichment cultures out of the sewage

activated sludge and the nutrient broth concentration had been judiciously chosen

to avoid metal complexation. The disc rotation speed (10 rev/min) and flow rate

(6.9 ml/min) had to be previously optimized for maximum biofilm growth and

metal removal.Candida spp. isolated from sewage samples could accumulate

significant amount of Ni (57–71 %) and Cu (52–68 %), but the process was found

to be affected by initial metal concentration and pH (optimum 3–5) of the

medium. This method of uptake is independent of the biological metabolic cycle

and is known as ‘‘biosorption’’ or ‘‘passive uptake’’. The three strains of bacteria

isolated from industrial effluents (Enterobacter cloacae and Klebsiella spp.) were

resistant to high concentrations of Cd, Pb and Cr in the growth media and could

remove approximately 85 % Cd during growth.

Arzu Y. Dursun et al [118] investigated the kinetics and thermodynamics of

copper (II) and lead (II) biosorption onto Aspergillus niger pretreated with NaOH

were studied with respect to pH, temperature and initial metal ion concentration.

The optimum pH values were determined as 5.0 and 4.0 for copper (II) and lead

(II) at 25 oC, respectively. Maximum biosorption capacities were obtained at pH

5 and 4 for Cu (II) and Pb (II), respectively. An increase or decrease in the pH

from these optimum pH resulted in a reduction in the biosorption of metal ions.

Little biosorption took place for pH less than 3.0.Solution pH influences both cell

surface metal binding sites and metal chemistry in water. It was shown that the

136
Chapter – III Literature Review

removal of both metal ions increased with increasing temperature up to 35 ◦C.

22.5 mg/g Cu (II) and 27.1 mg/g Pb (II) of dried microorganism were adsorbed at

equilibrium at 35 ◦C.

Aysegul Seker et al [119] studies has demonstrated that S. platensis is an

effective biosorbent for the adsorption of Pb2+, Cd2+, and Ni2+ ions from aqueous

solution. The uptake of metal ions by S.platensis seemed to be quite rapid and the

experimental data obeyed well the pseudo-second-order model. Faster adsorption

kinetics was observed for Pb (II) ions in comparison to Cd2+ ions. Aqueous Pb2+,

Cd2+, and Ni2+ ions were prepared from analytical reagent grade lead nitrate,

Pb(NO)3 (Riedel, 99 %, product code: 31137, CAS no. [10099-74-8]) cadmium

chloride, CdCl2 ( Fluka, > 99, product code: 20899, CAS no. [10108-64-2]) and

nickel chloride, NiCl2·6H2O (Carlo Erba, 99 %, product code: 464645, CASno.

[7791-20-0]) respectively. According to the thermodynamic parameters such as

G◦, H◦and S◦ calculated, the sorption process was endothermic and largely

driven towards the products. Sorption activities in a three metal ion system were

studied which indicated that there is a relative selectivity of the biosorbent

towards Pb2+ ions.

Bahadir et al [120] investigated the effects of the media conditions (pH,

temperature, biomass concentration) on the biosorption of Pb (II) ions to Rhizopus

arrhizus in a batch reactor. In order to determine the effect of biomass

concentration on Pb (II) removal, stock solutions of 10.0 g/L were prepared for

different microorganism particle size ranges, such as 0.045–0.063, 0.063–0.090

and 0.090–0.125 mm. These stock solutions were diluted to 0.1–1.0 g/L standard

137
Chapter – III Literature Review

solutions in order to find the effect of microorganism concentration on

biosorption. To determine the effect of pH of the biosorption medium on the

initial biosorption rates of Pb (II) ions, the pH was varied between 2.0 and 5.0.

The optimum initial pH of the biosorption medium was found to be 4.5. Results

of removal of Pb (II) ions from storage battery industry wastewater experiments

carried out at different temperatures ranging from 20 to 45 C. The optimum

biomass concentration was obtained at 1.0 g/L dose. It may be possible to

consider the biomass dose as 0.5 g/L but it may take time to reach at equilibrium

conditions at this dose. After the treatment of wastewater, the Pb (II)

concentration was measured as 0.157 mg/L which was far from the maximum

discharge limit of legacy.

In this study, biosorption of lead (Pb) from aqueous solution using citrus peels can

provide an efficient and cost-effective solution for lead removal by Balaria and

Silke [116] from industrial wastewaters. These peels contain the biopolymer

pectin that has a strong affinity for metal ions. A plateau can be observed for pH

6–9 at surface charge values of 0.8 meq/g for LM pectin and 0.4 meq/g for HM

pectin. LM pectin contains fewer methoxylated carboxylic groups and more free

carboxylic groups than HM pectin, which explains the higher negative charge for

LM pectin. For the LM pectin, 91 % of the carboxyl groups should be free (based

on a methoxyl content of 9 %) in contrast to 36 % of the HM pectin. The FTIR

spectra comparisons of pure versus Pb–laden pectins reveal that the absorption

peak corresponding to the carboxylate (–COO−) groups shifted significantly

from1627 cm−1 to 1636 cm−1 in the case of the HM pectin after treatment with Pb.

A similar shift in the same peak from 1624 cm−1 to 1635 cm−1 can be observed for

138
Chapter – III Literature Review

LM pectins. This shift in wave number corresponds to a change in bonding

energy corresponding to the functional group, which further reflects that the

bonding pattern of carboxylate groups changes after biosorption. This result

confirmed the involvement of carboxylic acid groups in binding of Pb in the case

of pectin.

Chang et al. [121] studied the biosorption kinetics of lead (Pb), copper (Cu) and

cadmium (Cd) ions on the biomass of pseudomonas aeruginosa PU21 (Rip64)

was investigated. Effects of environmental factors and growth conditions on the

biosorption were studied. The biomass originated from different growth phases

exhibited different adsorption capacities for Pb and Cd whereas the effect of

growth phase on biosorption of Cu was negligible. The saturation uptake capacity

and metal-cell affinity tended to increase as pH increased, until metals precipitated

as metal hydroxides when the pH exceeded some threshold values. Adjusting the

pH value to about 2.0 resulted in 98, 98 and 82 % recovery of Pb, Cu and Cd

respectively. The biomass resulted from desorption processes was able to retain

approximately 80 % of original adsorption capacity for Pb and Cu with four

repeated adsorption and recovery runs. The maximal adsorption capacity of

resting cells was nearly 110 mg Pb/g dry cell, whereas for the inactivated cells its

around 70 mg Pb/g dry cell.

Changzhou Yan et al [122] stated that the experimental results showed the

biosorption of lead and was adversely affected by low acidic pH values and the

presence of high concentrations of other metal ions. The removal efficiency of

lead ions was 37 % at a solution pH of 2.0, and it increased when solution pH

139
Chapter – III Literature Review

increased from 2.0 to 5.0. At pH 5.0, the removal rate was 76 %. It showed that

the sorption amount of lead rose markedly with increasing solution pH and that

the sorption process was pH dependent. The rise in lead biosorption value with

increasing pH can be explained by the presence of a strong relation between

biosorption and the number of negatively charged sites, which is highly dependent

on the dissociation of functional groups. In order to characterize the interference

by the Langmuir competitive model, another experiment was carried out at room

temperature (25 ± 1 ◦C) with initial pH of 5.0 for three metal solutions: Pb (II)

with varies concentration of Cu(II), 10, 30 and 50 mg/L, respectively. In all cases,

the Pb (II) concentration was varied from 10 to 250 mg/L and the samples were

placed on the shaker for a period of 120 min. Overall, this study suggested that

biosorption of Pb (II) ions by M. spicatum can be an inexpensive and effective

way of metal-ion treatment.

Comte et al [123] presented the influence of pH on the metal biosorption of

extracellular polymeric substances (EPS) extracted from two different activated

sludges called A and B. All samples were treated in exactly the same way in

order to obtain the polarographic titration curve. The polarographic measurements

were performed at 25 ± 1 oC under a nitrogen atmosphere and with a 20 mL total

volume of solution containing EPS. The pH of the different samples in the

analysis cell was continuously measured using a CRISON pH meter. The pH was

adjusted to 4.0, 6.0, 7.0 and 8.0 ± 0.1 by the addition of micro quantities of a

dilute solution of sodium hydroxide or nitric acid. Two pKa values were found

for each EPS. pKa1 are 6.6 and 5.7 and pKa2 are 8.7 and 9.4 for EPS A and B,

respectively.

140
Chapter – III Literature Review

Cynthia Urdaneta et al [124] studied, after a preliminary physical and chemical

characterization of the vermicompost, the optimal parameters for the heavy metal

adsorption. A synthetic multielemental solution of Pb, Cr and Ni and a solution of

NH4VO3 for vanadium were evaluated. The effect of pH on the adsorption

efficiency was evaluated in a range of 3.5 to 7.0 using aqueous solution of NaOH

or HNO3 of 0.10 mol/L. The effects of the agitation time were evaluated using 5,

10, 15, 30, 45 and 60 min. The solution volume was fixed to 75 mL, the pH to

5.00, the vermin compost mass was1g and the particle size ranging between 841

and 75 μm. Removing levels for Pb, Cr andNi are in the order of 95 % showing

that vermicompost is adequate for this removing process. Metal adsorption for Ni,

Pb and Cr in real wastewater from Barquisimeto city was 100, 65 and 40 %,

respectively.

Durali Mendil et al [125] experimented Bacillus thuringiensis var. israelensis

immobilized on Chromosorb 101 that is a new solid phase extractor has been

presented in his work for the preconcentration and separation of cadmium(II),

lead(II), manganese(II), chromium(III), nickel(II) and cobalt(II) in environmental

samples. The influences of the some metal ions as concomitant were investigated.

Under the optimized conditions, the detection limits by 3 sigma for analyte ions

were in the range of 0.37–2.85 g/L. The presented biosorption procedure is an

easy, safe, rapid, and inexpensive for the preconcentration and separation of trace

metals in aqueous solutions. The limits of detection (LOD) of the proposed

method for the determination of investigated elements were studied by passing

250 mL of blank solutions from the column under the optimal experimental

141
Chapter – III Literature Review

conditions. The LOD, defined as the concentration equivalent to three times the

standard deviation (N = 11) of the reagent blank were found as: Co(II): 2.11 g/L,

Cd(II): 0.37 g/L, Pb(II): 2.85 g/L, Mn(II): 0.71 g/L, Cr(III): 1.15 g/L and Ni(II):

1.67 g/L. The method was also applied successfully to the determination of

analytes in microwave digested red wine, rice and canned fish samples and sea

water, spring water and urine samples.

El-Naas et al [126] studied and the experimental results showed that the uptake

increased with increasing pH from 3.0 to an optimum value of 5.0. The

biosorption of Pb (II)was found to be adversely affected by the presence of Cu (II)

ions, while Zn (II) ions seemed to have negligible effect on the process. The

results also revealed that biosorption mechanism does not depend only on the

presence of active sites, and that it may involve a combination of complexation

and nucleation at cellar wall of the biosorbent. Modeling of the controlling

mechanism indicated that both intrinsic kinetics and mass transfer played major

roles in controlling the process. The results indicated that the uptake of lead ions

was significantly affected by solution pH and that the biosorption capacity of C.

vulgaris increased with increasing pH as shown in. At pH values of less than 3,

there seemed to be negligible uptake of lead; however, as the pH is increased to 4.

Fang Luo et al [127] investigated in his paper that marine brown algae Laminaria

japonica was chemically-modified by crosslinking with epichlorohydrin (EC1,

EC2), or oxidizing by potassium permanganate (PC), or only washed by distilled

water (DW). They were used for equilibrium sorption uptake studies with lead.

The biosorption is solution pH dependent and the highest removal of lead is

142
Chapter – III Literature Review

occurred at pH 4.0–5.3 for EC1 and EC2. With regard to PC and DW, the

maximum Pb2+ uptakes are obtained in pH range of 3.0–5.3. L–F isotherm model

is in good agreement with all the experimental data compared with the results of

Langmuir and Freundlich isotherm models. The adsorption characteristics of lead

with various pH in the range of 1.4–5.3 are examined using four biomasses. The

initial concentration of Pb2+ solution is 1.0 mmol/L and the investigation of pH

values above 5.5 was not possible since lead precipitation appeared. The

maximum lead uptakes were 1.67 mmol g/L, 1.62 mmol g/L, 1.54 mmol/g and

1.21 mmol/g, respectively for EC1, EC2, PC and DW. The order of maximum

lead uptakes for different pretreated and raw alga was EC1 > EC2 > PC > DW

Flavio A. Pavan et al [128] experimented and from the study, it can be concluded

that natural ponkan peel is an effective and efficient biosorbent for removing Pb

(II) ions from aqueous solution. The effects of initial pH on biosorption capacity

of Pb (II) ions using ponkan peel were evaluated within the pH range of 2.0–8.0

Electron micrographs and EDX spectra of the ponkan peel before and after

biosorption with 1000 mg/L of solution contained Pb (II) ions. The broad and

intense absorption with a maximum at 3400 cm−1 was assigned to O–H stretching

group, originally bonded to organic polymeric structure. In order to evaluate the

influence of pH parameter on biosorption, experiments were carried out at

different pH values between 2.0 and 8.0. For this experiments the initial Pb (II)

concentration used was 15.0 mmol/L, pH of the solution was 5.0 and contact time

120 min.

143
Chapter – III Literature Review

Francesca Beolchini et al [129] investigated that the experimental data of single

and binary metal biosorption in membrane reactor and is reported in this work

denoted by the effect of metal competition for the active sites on the biosorbent.

Biosorption tests using single and binary metallic solutions (Cu, Pb and Cu–Pb)

denoted the biomass affinity (PbNCu), the competition among metals

simultaneously present in the system, the filtrate flux decline and the change of

metal retention coefficient on the membrane for pore plugging by cell fragments.

Dynamic modelling is developed considering the unsteady mass balances of the

metal in the system and the equilibrium parameters obtained by biosorption batch

tests using Langmuir models.

Francisco W. Sousa et al [130] investigated that green coconut shells can be used

as adsorbent for the removal of toxic metal ions from aqueous effluents using

column adsorption. Using HNO3 1.0 mol/L, the adsorbent can be reused, but

removal efficiencies decrease from the first cycle. Investigation of a real sample,

containing Ni (2.787 mg/L); Zn (98.02 mg/L) and Cu (320.4 mg/L), showed that

the fixed-bed technology can be applied, provided that the pH of the sample is

adjusted to 5.0. It can be observed that the reuse of the adsorbent decreases the

removal efficiency for all metal ions from the first cycle. Removal capacities

were reduced more than 50 % for Pb, 70 % for Ni; 67 % for Cd; 76 % for Zn and

75 % for Cu.

Gulsad Uslu et al [131] studies include the biosorption of lead (II) and copper (II)

ions, in single and binary systems using P. putida in the batch system. The lead

(II) and copper (II) ions binding capacity of microorganism was shown as a

144
Chapter – III Literature Review

function of initial pH, temperature and single–binary-pollutant concentrations.

The single and binary component sorption phenomena were expressed by the

single component (noncompetitive) and multicomponent (competitive) Langmuir

and Freundlich adsorption models and adsorption isotherms were developed for

both single and binary component systems at different temperature. For binary

lead (II) and copper (II) ions mixture studies, desired combinations of lead (II) and

copper (II) ions were obtained by diluting 1.0 gd/m3 of stock solutions of

component and mixing them in the test medium before mixing with the bacterial

suspension. Adsorption data were well described by the Langmuir model,

although they could be modeled by the Freundlich equation. The thermodynamics

constants of the adsorption process: H◦, S◦ and G◦ were evaluated.

Gabr et al [132] investigated the optimum conditions for biosorption and

bioaccumulation of lead and nickel using a tolerant bacterial strain isolated from

El-Malah canal, Assiut, Egypt, and identified as Pseudomonas aeruginosa ASU

6a. The optimal pH values for Ni (II) and Pb (II) adsorption were, respectively,

7.0 and 6.0. The maximum adsorption uptake (qmax) of nickel and lead

calculated from Langmuir equation for biosorption by heat-dried cells and

lyophilized cells or accumulated by living cells is 113.6, 77.5, 70 and 123, 93, 79

mg/g for both metals, respectively. The appropriate equilibrium time for

measurements was taken at 30 min. Here it is worth mentioning that the

correlation coefficients for all the lead and nickel Pseudomonas systems were

found to be close to unity.

145
Chapter – III Literature Review

García-Rosales et al [133] investigated the removal of Pb (II) from aqueous

solutions by a maize (Zea mays) stalk sponge. The sorption process was found to

be influenced by the pH of the solution; the percentage of sorbed lead increased

with the pH value, with maximum sorption percentage of 70 % at pH 7. Specific

metal uptake was found to decrease with an increase in pH (after pH > 7). The

adsorption capacities increased with increasing pH, but pH 7.0 was too high for

Pb because it precipitates at pH > 5. The results obtained showed that Zea may

stalk sponge was a useful biomaterial for Pb (II) sorption and that pH has an

important effect on metal biosorption capacity.

Groudeva et al [134] found that the Tulenovo oil deposit, Northeastern Bulgaria,

waters contaminated with crude oil and toxic heavy metals Cd, Cu, Pb, Mn, Fe

and were treated by a passive system of the type of the constructed wetlands. The

oil content in the fluid recovered from the different wells varies in the range of

about 0.1–0.5 %. The oil is heavy, rich in asphalthene-resinous substances and

with a high viscosity. The total ion concentration in the brine is about 2–4 g/L

and the pH is in the range of 6.8–7.7. The oil content of the waters after treatment

was decreased to less than 0.2 mg/L, and the concentrations of heavy metals were

decreased below the relevant permissible levels. The waters had a pH of about

5.3–6.8 and contained about 2–10 mg/L oil.

Gupta et al [135] conducted the batch studies to provide significant information

regarding biosorption of lead on green algae Spirogyra species in terms of

optimum pH and biomass dose for maximum removal of Pb (II) from the aqueous

solution. It has been observed that maximum adsorption took place within first

146
Chapter – III Literature Review

100 min. As the pH of the lead solution (100 and 200 mg/L) increased from 2.99

to 7.04, the adsorption capacity of lead was changed, i.e. it first increased from

2.99 pH to pH 5.0 and then dramatically decreased up to pH 7.04. The extent of

lead biosorption was 31.2 % for 0.05 g/L of algal biomass, while it was greatly

increased to 80 % for 10 g/L of adsorbent. For an increase in temperature from

298 to 318 K, an increase in the adsorption of lead was observed. The maximum

amount adsorbed increased from 96.4 to 104 mg/g at 150 mg/L, initial lead

concentration.

The biosorption of different metals (Cu2+, Cd2+, Zn2+, Ni2+ and Pb2+) was

investigated using activated sludge by Hammaini et al [136]. The optimum pH

was 4 for Cd, Cu and Pb sorption and 5 for Ni and Zn. Biomass metal uptake

clearly competed with protons present in the aqueous medium, making pH an

important variable in the process. Protons consumed by biomass in control tests

versus protons exchange in biosorption tests confirmed a maximum exchange

between metal cations and protons at pH 2. pH had a clear influence on the

sorption capacity of activated sludge for Cu, Cd, Ni, Zn and Pb. The competition

between metal and protons sorption was greater at pH 2. At very low pH values

(pH 1–2) metal uptake was negligible. Metal uptake increased with pH up to 4–5,

beyond these pH values no improvement in the sorption capacity was observed.

The most adequate sorption pH was 4 for Cd, Cu and Pb, 5 for Ni and 6 for Zn. In

the latter case, the pH was also set at 5 to avoid its precipitation, especially in

those experiments carried out to draw the sorption isotherms in which the initial

metal concentration was higher.

147
Chapter – III Literature Review

The effect of the presence of Pb on the biosorption efficiency of Cu, Cd and Zn by

activated sludge was investigated by Hammaini et al. [137]. The interference of

Pb with the Zn uptake was very pronounced. The amount of Zn adsorbed [Cf [Zn]

= 1mM] when Pb was present at 0.2 mM was 87 % less than when Cf [Pb] = 0.05

mM. At Cf [Pb] = 0.2 mM and when 0.5 and 2 mM of Zn were present in the

system (final equilibrium Zn concentration), the Pb uptake was 0.5 and 0.48

mmol/g respectively. When the residuals of Pb and Zn were the same, about 99 %

of the total metal uptake was due to Pb uptake. The equilibrium concentration of

Zn would have to be 70 times greater than that of Pb to obtain the same proportion

of uptake for each metal. When the residuals of Pb and Cd were the same, about

88 % of the total metal uptake was due to BP.

This work confines to biosorption of copper and lead ions on waste beer yeast, a

byproduct of brewing industry by Han et al. [138]. Batch biosorption tests were

done at different contact time at the initial concentration of 0.315 mmol/L for Cu

(II) and 0.393 mmol/L for Pb (II) respectively, and waste beer yeast dose

concentration is 8 g/L in 10 mL solution. The temperature was controlled with a

water bath at the temperature of 293 K. Experiments could not be performed at

higher pH values due to hydrolysis and precipitation of the lead ions. Flasks were

agitated on a shaker for 60 min to ensure that equilibrium was reached. There was

an increase in biosorption capacity of biomass with increasing pH from 2.0 to 6.0

for both metal ions. The lowest metal uptake values were determined at pH < 2.0

for both metal ions. The adsorption quantity of Pb (II) decreased with the

increasing of Cu (II) concentration. The biosorption quantity of Pb (II) decreased

148
Chapter – III Literature Review

from 0.0218 to 0.0140 mmol/g when the concentration of Cu (II) ranged from 0 to

0.472 mmol/L.

Biosorption of lead (II) onto a cone biomass of Pinus syivestris was studied by

Handan Ucun et al [139] with variation in the parameters of pH, initial metal ion

concentration and impeller speeds. The biosorption medium was stirred at

constant speed for 1 h at 25 oC.The role of contact time on biosorption of lead by

cone biomass was studied under shake flask conditions at pH 4.0, 150 rpm, using

50 mg/L lead ion concentration. The pH of the biosorption medium affects the

biosorption rate on biomass. The pH solution ranging from 2.0 to 5.5 was studied.

Biosorption studies of lead with cone biomass was carried out in a shaker working

at pH = 4.0 using a lead solution of 50 mg/L lead solution. By varying the

impeller speed from 100 to 240 rpm in different sets, keeping biomass constant,

the effect of contact time on the biosorption capacity.

The results of the study given by Harikishore Kumar Reddy et al [140] revealed

that MOL wastes could be converted into an innovative low-cost biosorbent with

good biosorption capacity by modifying it with NaOH and citric acid. The

biosorption of Pb (II) by CAMOL from aqueous solution was found to be greatly

dependent on solution pH. Moringa oleifera leaves (MOL); an agro-waste

material has been used as a precursor to prepare a new biosorbent. The leaves

were washed with base and citric acid, and obtained new chemically modified

MOL biosorbent (CAMOL) for sequestration of Pb (II) from aqueous solution.

The biosorbent was characterized by SEM, FTIR spectral and elemental

analyses.The equilibrium data were analyzed using Langmuir, Freundlich,

149
Chapter – III Literature Review

Dubinin–Radushkevick and Temkin isotherm models. Langmuir model provided

the best correlation with biosorption capacity of 209.54 mg/g at 313 K. The

thermodynamic properties showed that biosorption of Pb (II) onto CAMOL was

spontaneous, endothermic and feasible in the temperature range of 293–313 K.

Desorption experiments showed feasibility of regeneration of the biosorbent for

further use after treating with dilute HCl.

Ilhami Tuzun et al [141] studied the biosorption of lead with cone biomass was

carried out in a shaker working at pH 4.0 using a lead solution of 50 mg/L lead

solution. By varying the impeller speed from 100 to 240 rpm in different sets,

keeping biomass constant, the effect of contact time on the biosorption capacity.

The biosorption of Hg (II), Cd (II) and Pb (II) ions by microalgae biomass

increased as the initial concentration of Hg (II), Cd (II) and Pb (II) ions increased

in the biosorption medium. The maximum biosorption capacities of microalgae

for Hg (II), Cd (II) and Pb (II) ions were 72.2 G 0.67, 42.6 G 0.54 and 96.3 G 0.86

mg/g dry biomass, respectively. The biosorption of Hg (II), Cd (II) and Pb (II)

ions by microalgae biomass increased as the initial concentration of Hg (II), Cd

(II) and Pb (II) ions increased in the biosorption medium. The maximum

biosorption capacities of microalgae for Hg (II), Cd (II) and Pb (II) ions were 72.2

G 0.67, 42.6 G 0.54 and 96.3 G 0.86 mg/g dry biomass, respectively.

Immobilized Microcystis aeruginosa in a flow-through sorption column was

evaluated for the potential to remove Pb2+, Cd2+ and Hg2+ from aqueous solutions

by Jianhua Pan et al [142]. M. aeruginosa showed high affinity for the three

heavy metals with removal efficiency of 90 % for Cd2+ and Hg2+, and 80 % for

150
Chapter – III Literature Review

Pb2+ at saturation conditions. The removal efficiency reached 90 % for Cd2+ and

Hg2+, and 80 % for Pb2+ when the biosorption attained a plateau. Competitive

uptake experiments carried out in binary and ternary systems revealed a high

selectivity for Pb2+ over Cd2+ and Hg2+. Pb2+ and Hg2+ interfered slightly with the

uptake of Cd2+. As such, lower interference of Cd2+ on Hg2+ was also observed in

this study. In contrast, the influence of Pb2+ on Hg2+ adsorption was significant at

certain time. Cooperation of Pb2+ and Hg2+ significantly interfered with the

uptake of Cd2+ in the first 10 min, and thereafter the cadmium removal efficiency

remained lower than that for Cd2+ added individually and paired with Pb2+ and

Hg2+ during the experimental timeframe except for 70 min.

To elucidate the potential mechanisms involved in the biosorption of metal ions,

atomic force microscopy (AFM) and Fourier transform infrared (FTIR)

spectroscopy were used to characterize the interaction between Pb2+ and Bacillus

cereus by Jianhua et al. [4]. The scanning speed and the loop gain factors were

varied during the imaging process. Changes in the spectra are attributed to the

interaction of Pb2+ with the carboxyl, hydroxyl and amino groups present on the

surface of the biosorbent. The shift of peak at 1455.6 cm−1 transposing to lower

frequencies up to disappearance is due to the complexation of amino and hydroxyl

groups with Pb2+. Kinetic studies demonstrate that the biosorption of Pb2+ by B.

cereus biomass is a rapid process and follows the pseudo-second-order rate law.

The AFM imaging of B. cereus biomass surfaces after Pb2+ biosorption indicates a

major morphological difference on mica, the assembly structures changing from

rod-like to spindle-like. Among the rest, the Redlich–Peterson model yielded the

best fit of experimental data.

151
Chapter – III Literature Review

Ji Bing et al. [143] investigated the potential of Sedum alfredii hance for the

biosorption of some metals from synthetic wastewater. S. alfredii hance was

grown in synthetic wastewater in 2.5 L capacity containers contaminated with

(mg/L) 19.20 Zn, 11.24 Cd, 3.27 Cu and 0.53 Pb respectively. The initial Pb

concentration of in the control solution (without plants) was 0.53 mg/L that is

decreased to 0.23 ± 0.04 mg/L after the experiment with a removal rate of 56.32 ±

8.28 %, likely resulting from precipitation. The removal efficiencies of Zn, Cd,

Cu and Pb were 94.07 %, 82.03 %, 96.03 % and 69.61 % respectively. Both root

surface adsorption and plant absorption contributed to the removal of metals from

contaminated water. The amounts of heavy metals removed by root surface

adsorption accounted for 67.74 %, 24.03 %, 66.53 % and 70.77 % of the total

plant removal for Zn, Cd, Cu and Pb respectively. The supplied metal

concentrations were comparable with those in plating industrial wastewater. The

results indicated that heavy metal removal was mostly contributed by adsorption

on root surface in case of Zn, Cu and Pb but less for Cd than root absorption.

The study experimented by Jose T. Matheickal et al [144] indicated that

abundantly and cheaply available brown marine algae from the Australian marine

environment can be used for the development of efficient biosorbent materials for

heavy metal recovery from wastewater. A two stage modification process

substantially improved the leaching characteristics of the biomass. Batch

equilibrium experiments showed that the maximum adsorption capacities of

DP95Ca for lead and copper were 1.6 and 1.3 mmol/g, respectively. The

corresponding values for ER95Ca were 1.3 and 1.1 mmol/g. These capacities are

152
Chapter – III Literature Review

comparable with those of commercial ion exchange resins and are much higher

than those of natural zeolites and powdered activated carbon. The heavy metal

uptake process was found to be rapid with 90 % of the adsorption completed in

about 10 min in batch conditions. Heavy metal adsorption was observed at pH

values as low as 2.0 and maximum adsorption was obtained approximately at a pH

of 4.5.

Jo-Shu Chang et al [145] investigated the biosorption kinetics of lead (Pb),

copper (Cu) and cadmium (Cd) ions on the biomass Pseudomonas aeruginosa

PU21 (Rip64) was investigated. Effects of environmental factors and growth

conditions on the biosorption were studied. The biomass originated from different

growth phases exhibited different adsorption capacities for Pb and Cd, whereas

the effect of growth phase on biosorption of Cu was negligible. The saturation

uptake capacity and metal-cell affinity tended to increase as pH increased, until

metals precipitated as metal hydroxides when the pH exceeded some threshold

values. Adjusting the pH value to about 2.0 resulted in 98, 98 and 82 % recovery

of Pb, Cu and Cd, respectively. The biomass resulted from desorption processes

was able to retain approximately 80% of original adsorption capacity for Pb and

Cu with four repeated adsorption and recovery runs. Regeneration of biomass

appears to enhance the uptake capacity of Cd by nearly 35% after four

adsorption/desorption cycles.

Junxia Yu et al [146] studied a simple method was used to prepare modified

biomass to improve its adsorption capacity for Cd2+ and Pb2+. This study showed

that grafting effective groups on the biomass could result in a biosorbent with high

adsorption capacity for metal ions.the dependence of Cd2+ and Pb2+ uptake with

153
Chapter – III Literature Review

the solution pH. It could be seen a fast increase in lead and cadmium uptake with

increasing pH from 2.5 to 3.5, while around pH 4.5, the Cd2+ and Pb2+ adsorption

capacity leveled off at a maximum value reaching a plateau. In these experiments,

the Cd2+ and Pb2+ concentration were 150 mg/L and 250 mg/L, respectively. After

0.0500 g of the biomass was used to adsorb Cd2+ at 80 mg/L in 50 mL solution

over 0.5 h, the adsorbent was regenerated, and then rinsed with distilled water and

used in subsequent adsorption experiments. The adsorption capacity and

desorption efficiency of the modified biomass for Cd2+ over six successive

adsorption–desorption cycles. It could be seen that little loss of uptake capacity of

the biomass was observed after using for six times, and the desorption efficiency

was above 90 %.

Biosorption of lead (II) and cadmium (II) from aqueous solutions by protonated

Sargassum glaucescens biomass was studied in a continuous packed bed column

by Kazem Naddafi et al [147]. The selective biosorption capacities of Pb2+ and

Cd2+ at complete exhaustion point were determined to be 1.18 and 0.22 mmol/g,

respectively; therefore, the biosorbent exhibited much higher relative affinity for

Pb2+ than for Cd2+. The efficiency of biosorbent regeneration by 0.1MHCl was

achieved about 60 %, so that the maximum uptake capacity of Pb2+ by the

regenerated biomass was determined to be 0.75 mmol/g while the same value for

the original biomass was 1.24 mmol/g.the effluent pH decreased to about 2.3 by

the most replacement of hydrogen ions in the binding sites with Pb2+ and then by

reduction of protonated binding sites as well as decrease of Pb2+ biosorption rate,

the effluent pH increased gradually until the effluent pH became equal to the

154
Chapter – III Literature Review

influent pH at complete exhaustion of the biosorption bed, so the rate of metal ion

biosorption was proportionate to release of hydrogen ions.

Kılıc et al [148] studied adsorption of Pb (II) and Hg (II) on nonviable activated

sludge biomass at both batch and continuous-flow operational conditions.

Equilibrium concentration, equilibrium period, optimum pH and temperature for

adsorption were found out to be 0.387 mmol/g, 3.5 and 30 oC for Pb (II) and 0.097

mmol/g, 105 min, 5.8 and 20 oC for Hg (II), respectively. Optimum pH’s for both

metal ions were investigated using solutions containing 0.05 g dried raw biomass

and 1mg metal ion within 50 ml of deionized water. The equilibrium was reached

around 90 min though 62 % of the biosorption capacity for Pb (II) was completed

in first 15 min. In the tests, the highest biosorption capacity was observed at 30oC

as 0.091 mmol Pb (II) per g biomass. Optimum temperature for Pb (II) and Hg

(II) biosorption were found out to be 30 oC and 20 oC, respectively.

Lalhruaitluanga et al [149] demonstrated that the charcoal biomass could be

used as adsorbents for the treatment of Pb (II) from aqueous solution. The effect

of chemical pretreatment on the adsorption of Pb (II) ions by MBRC and MBAC

showed that 60 % KOH pretreatment shows highest percentage of adsorption.

The maximum biosorption of Pb (II) was found at pH 5. In order to determine the

functional groups involved in metal adsorption, the unloaded biomass, MBRC-Pb

(II) loaded and MBACPb (II) loaded with 50 mg/L Pb (II) ions were analyzed

using Fourier Transform Infrared Spectroscopy (JASCO FTIR-5300) in a range of

400–4000 cm−1. Chemical activation was carried out using phosphoric acid

(Qualigens) and potassium hydroxide (Himedia) at various concentrations (20, 40,

155
Chapter – III Literature Review

60 and 80 %, w/w). In the activation system, the charcoal powder with a mass

concentration of 40 g/L was in contacted with the above various concentrations of

chemical reagents, respectively, for 24 h by shaking on an orbital shaker with a

speed of 150 rpm.

Lawal et al [150] showed that the uptake of the metal ions increased with increase

in initial metal ion concentration. The lead ion was optimally adsorbed at pH 4,

and its uptake was rapid in the first 10 min of adsorption after which adsorption

decreased. Pseudo-secondorder kinetics explained the biosorption of the metal ion

better than the pseudo-first-order. The pH of the Pb (II) solution was first adjusted

to a desired value, from 1.0 to 6.0 with HNO3 or NaOH. It was observed that the

biosorption process reached equilibrium after 75 min when the biosorption of the

Pb (II) ions on C. inophyllum seed husk was 95 %.It is reasonable that increase in

metal ion concentration facilitated enhanced displacement of hydrogen ions on the

surface of the biomass at pH 4 because of high concentration of the metal ions.

The stretching vibrations of OH group shifted from 3391 to 3385 cm−1 after Pb

(II) ions adsorption.

Lenka Svecova et al [151] studied the evaluation of the impact of pH on the

sorption of selected metals and continued with the sorption isotherms

determination. The maximum uptake capacity is close to 270 mg/g mercury (i.e.

1.34 mmol/g) and the affinity coefficient is quite low (about 0.07 L/mg, or 14.04

L m/mol). The optimum pH for sorption of mercury is obtained at initial pH close

to pH 5; however, during the sorption process the pH rises to 6.5 and the sorption

is not increased when controlling the pH at the initial value; the fluctuation of pH

156
Chapter – III Literature Review

reveals favorable to metal recovery.Results obtained at pH lower than 3 are not

presented since the material was partially dissolved in acidic pH region. For

biosorbent P3, the sorption efficiency remained almost constant between pH 4 and

6. After the treatment with ethanol (P2), Penicillium biomass lost a significant

percentage of C (from 48 % to 43 %) while the percentage of nitrogen slightly

increased (from 4.4 % to 5 %). Hydrogen content slightly decreased (from 7.6 %

to 7.1 %).

Biosorption potential of Penicillium simplicissimum (Penicillium sp.) immobilized

within loofa sponge (PSILS) for lead and copper from aqueous media was

explored by Li et al. [152]. The effect of pH on the biosorption capacity was

investigated at initial pH values range of 2.0–6.0 and the desired pH of the

suspensions was maintained by adding HCl or NaOH at the beginning of the

experiment and not controlled afterwards. Results showed that the biosorption of

Pb (II) and Cu (II) on PSILS increased significantly as the pH increased from 2.0

to 5.0. The low Pb (II) and Cu (II) biosorption capacity at pH values below 3.0

may be attributed to hydrogen ions that compete with metal ions on the sorption

sites. The contact time of approximately 60 min was required to reach the

equilibrium. The effect of initial single metal ion concentration was investigated

in the range of 10–500 mg/L. sites. According to Langmuir isotherm, the

monolayer saturation capacity of PSILS is 144.9 mg/g for Pb (II) and 106.4 mg/g

Cu (II).

Luo et al. [153] studied about marine brown algae Laminaria japonica which was

chemically-modified by crosslinking with epichlorohydrin (EC1, EC2) or

157
Chapter – III Literature Review

oxidizing by potassium permanganate (PC) or only washed by distilled water

(DW) and are used for equilibrium sorption uptake studies with lead. The

biosorption is solution pH dependent and the highest removal of lead is occurred

at pH 4.0–5.3 for EC1 and EC2. With regard to PC and DW, the maximum Pb2+

uptakes are obtained in pH range of 3.0–5.3. Langmuir and Freundlich isotherm

models are in good agreement with all the experimental data. The adsorption

characteristics of lead with various pH in the range of 1.4–5.3 are examined using

four biomasses. The initial concentration of Pb2+ solution is 1.0 mmol/L and the

investigation of pH values above 5.5 was not possible since lead precipitation

appeared. The maximum lead uptakes were 1.67 mmol/g, 1.62 mmol/g, 1.54

mmol/g and 1.21 mmol/g, respectively for EC1, EC2, PC and DW. The order of

maximum lead uptakes for different pretreated and raw alga was EC1 > EC2 > PC

> DW.

Response surface methodology was applied to optimize the removal of lead ion by

Aspergillus niger in an aqueous solution by Malihe Amini et al [154].

Optimization of biosorption from aqueous solution by response surface

methodology resulted in 96.21 % lead removal than that of the pre-optimized

condition. The level of the three variable, initial solution pH, 4.27; initial ion

concentration, 30 mg/L; A. niger dosage, 2.17 g/l, were found to be optimum for

maximum lead ion removal.A minimum level of biomass dosage (1.6 g/l),

maximum levels of initial lead ion concentration (30 mg/L) and lead removal

(96.21 %) and level of initial solution pH within range of 2.8–7.2 were set for

maximum desirability. The importance of each goal was changed in relation to the

other goals.

158
Chapter – III Literature Review

Gymnogongrus torulosus adsorption efficiency for cadmium (II), copper (II), lead

(II) and zinc (II) were studied in batch mode in different acidic conditions by

María Mar Areco et al [155]. Lead biosorption exhibits a slight dependence on

pH. Both zinc (II) and copper (II) have a similar profile with a higher dependence

on pH, while cadmium shows moderate pH dependence. Initial solution pH plays

a significant role on metal biosorption, with maximum uptake at pH values above

5. Temperature can influence the sorption process rate. An increase in temperature

from 7 to 25 oC at pH 5.5 increases the pseudo-second-order kinetic constant by a

factor of 2.1, 17.3, 13.6 and 10.5 for lead (II), cadmium (II), zinc (II) and copper

(II), respectively. Increasing the temperature from 25 to 55 oC decreases the

pseudo-second-order kinetic constant by a factor of 0.1, 0.2 and 0.6 for lead (II),

cadmium (II) and copper (II), respectively. The sea weed was washed thoroughly

with deionized water, then dried overnight at 60 oC and finally stored in

desiccators before being used. Afterwards, the dried seaweed was blended in a

homogenizer into finer particles. A stainless steel standard sieve was used to

obtain fine particles (0.5–2 mm) of sea weed. Biosorption capacities were affected

by solution parameters. The maximum metal uptake (qmax) increased with

increasing pH.

Mata et al [156] determined the effect of immobilized brown alga Fucus

vesiculosus in the biosorption of heavy metals with alginate xerogels. Calcium in

the xerogels was displaced by heavy metals from solution according to the “egg-

box” model, creating a more uniform and organized structure. The Langmuir

maximum biosorption capacity increased two fold for cadmium, 10 times for lead,

159
Chapter – III Literature Review

and decreased by half for copper. According to this model, the affinity of the

metals for the biomass was as follows: Cu > Pb > Cd without alga and Pb > Cu >

Cd with alga. Tests were run for 24 h, time enough to reach the equilibrium, at the

initial pH values previously mentioned, with 0.5 g/l initial biomass concentration

and 50 ml of different initial metal concentration solutions: 10, 25, 50, 100 and

150 mg/L.

Meral Yurtsever I. Ayhan Sengil et al [157] investigated the effect of

temperature, pH and initial metal concentration on Pb (II) biosorption on modified

quebracho tannin resin (QTR) was investigated. The specific BET surface area of

QTR was found to be 0.820 m2/g.The adsorption of Pb (II) increases with time

from 0 to 10 min and then becomes almost constant up to the end of

experiment.The analysis is carried out through an adsorption of N2 gas (gas

effluent temperature = 75 oC versus bath temperature = 77.35 oC). The measured

BET specific surface area for QTR was 0.820 m2/g.The kinetic data was tested

using pseudo-first-order, pseudo-second-order, Elovich and intraparticle diffusion

model. The results suggested that the pseudo-second-order model (R2 > 0.999)

was the best choice among all the kinetic models to describe the adsorption

behavior of Pb (II) onto QTR. Langmuir, Freundlich and Tempkin adsorption

models were used to represent the equilibrium data.

Mohammad Mehdi Montazer-Rahmati et al [158] studied the biosorption of Cd

(II), Ni (II) and Pb (II) from aqueous solutions using brown algae, C. indica, S.

glaucescens, N.zanardini and P. australis in batch systems indicates that the

adsorption of metal ions is significantly dependent on the pH of solution. The

160
Chapter – III Literature Review

results are best fitted by the Freundlich model among two-parameter models and

the both, Khan and Radke-Prausnitz models among three-parameter isotherm

models for Cd (II), Ni (II) and Pb (II), respectively. The kinetic data were fitted

by models including pseudo-first-order and pseudo-second-order. The adsorption

characteristic of metal biosorption at various pH values in the range of 2.5–7.0

was examined. Experiments were not conducted beyond a pH of 7.0 to avoid

metal precipitation. The adsorption rate tests were performed on an equilibrium

batch basis. 0.06 g/L of the biomass was contacted with a solution bearing a metal

concentration of 0.25, 1 and 0.5 mmol/L for Cd (II), Ni (II) and Pb (II),

respectively at 25 oC and the optimum pH. The pseudo second-order kinetic

model fits the experimental data well (R2 : 0.99 for all metals).

Mustafa Tuzen et al [159] investigated the biosorption of copper (II), lead (II),

iron (III) and cobalt(II) on Bacillus sphaericus-loaded Diaion SP-850 resin for pre

concentration–separation of them have been investigated. Biosorption on B.

sphaericus for the atomic absorption spectrometric determination of copper (II),

lead (II), iron (III) and cobalt (II) ions at trace level is described. In order to study

the adsorptive capacity of B. sphaericus loaded Diaion SP-850, batch method was

used. To 0.1 g resin was added 50 mL of solution containing 1.0 mg of metal ion

at pH 6.0. After shaking, the mixture was filtered. To estimate the accuracy of the

procedure, different amounts of the investigated metal ions were spiked in natural

waters and urine sample. The detection limits by 3 sigma for analyte ions were in

the range of 0.20–0.75 µg/L for aqueous samples and 2.5–9.4 mg/g for solid

samples.

161
Chapter – III Literature Review

Nabanita Chakraborty et al [160] studied the lead accumulation by free and

immobilized cyanobacteria, Lyngbya majuscula and Spirulina subsalsa. The

column packed with glass wool only retained 18–20 % Pb from 5 mg/L solution,

but 95.8 % Pb was adsorbed by the column packed with growing cyanobacteria

filaments and only 4.2 % Pb was found in Elute I. After the first wash with 0.1 M

EDTA, 92.5 % of accumulated Pb was recovered in Elute II. A second exposure

to 5 mg/L Pb solution showed 59–60 % removal of Pb (40 % remaining in Elute

III) indicating its ability to be used in repeated Pb removal process. In L.

majuscula accumulation of Pb was 3.5, 3.0 and 2.89 mg/g at pH 6, 7 and 8

respectively. The corresponding values were 1.07, 1.05 and 0.9 mg/g for S.

subsalsa. The absorption was concentration dependant. L. majuscule accumulated

15 times more Pb when exposed to 20 mg/L Pb solution for 3 h than that to 1

mg/L while a 2.5 to 3-fold increase in accumulation of Pb was observed in S.

subsalsa under the same experimental conditions. A 92.5 % recovery of

accumulated Pb from the immobilized biomass suggests that repeated absorption–

desorption is possible.

Biosorption of lead (II) and cadmium (II) from aqueous solutions by protonated

sargassum glaucescens biomass was studied by Naddafi et al. [161] in a

continuous packed bed column. The selective biosorption capacities of Pb2+ and

Cd2+ at complete exhaustion point were determined to be 1.18 and 0.22 mmol/g

respectively. Therefore, the biosorbent exhibited much higher relative affinity for

Pb2+ than for Cd2+. The efficiency of biosorbent regeneration by 0.1 M HCl was

achieved about 60 %, so that the maximum uptake capacity of Pb2+ by the

regenerated biomass was determined to be 0.75 mmol/g while the same value for

162
Chapter – III Literature Review

the original biomass was 1.24 mmol/g. The effluent pH decreased to about 2.3 by

the most replacement of hydrogen ions in the binding sites with Pb2+ and then by

reduction of protonated binding sites as well as decrease of Pb2+ biosorption rate,

the effluent pH increased gradually until the effluent pH became equal to the

influent pH at complete exhaustion of the biosorption bed, so the rate of metal ion

biosorption was proportionate to release of hydrogen ions.

Several number of biosorption experiments have been done considering dead

biomass, weeds as biosorbents until now but, in this Nadeem et al. [162] stepped

away from conventional methods and took another form, a locally available fish

(Labeo rohita) scales for Pb (II) removal from aqueous solutions under different

experimental conditions. Sorption experiments beyond pH 5 were not conducted

since insoluble precipitates appeared beyond this pH. Stock Pb (II) solution (1000

mg/L) was prepared by dissolving 1.598 g of Pb(NO3)2 in 100 mL deionized water

shaking it for 15 min on a magnetic stirrer to obtain complete dissolution and then

diluting quantitatively to 1000 mL using deionized water. Sorption capacities

(mg/g) of fish scales pretreated using bases were in the following order: Ca(OH)2

(200.76) > nontreated (196.8) > Al(OH)3 (192.76) > NaOH (149.14). The two

factors are responsible for increase or decrease in sorption capacity of a particular

biomass after basic pretreatment. The highest removal capacities 196.80 mg/g of

Pb (II) ions by fish scales were obtained at pH 3.5 and the overall removal

capacity of Pb (II) decreased to 86.96 mg/g as pH increased up to 5.

Nurbas Nourbakhsh et al [163] investigated the biosorption of various metal

ions onto Bacillus sp. (OGUB 001). At 27 oC, pH values of 2.0, 4.5 and 4.5 were

163
Chapter – III Literature Review

determined for Cr6+, Pb2+ and Cu2+ ions, respectively. At pH 4.0, 71.8 % of the

capacity was used by lead, while 16.3 and 11.9 % were used by copper and

chromium ions, respectively ; and, at pH 7.0, 75.42 % of the capacity was used by

lead while 0.9 and 23.68 % by copper and by chromium ions, respectively. The

multi-metal ions, containing different concentrations of Cr6+, Pb2+ and Cu2+ ions

together were prepared and partial competitive adsorptions of these mixtures were

investigated at 27 oC and for the pH values of 4.0 and 7.0 that are the most

frequently seen conditions in industrial waste waters.

Ofomaja et al [164] investigated the kinetic modeling is frequently performed on

both raw and modified biosorbents and changes in model parameters. Kinetic

experiments were carried out by agitating 100 cm3 of lead (II) solution of

concentration ranging from 60 to 120 mgd/m3 with 0.40 g of pine cone and base

modified pine cone powder in a 250 cm3 beaker at 291K at an optimum pH of 5.0

and at a constant agitation speed of 160 rpm for 15 min. Samples (3.0 cm3) were

pipette out at different time intervals, centrifuged and the concentration of lead

(II) analyzed using a Perkin-Elmer model 2100 atomic absorption spectrometer

(AAS). To a series of 100 cm3 conical flasks, 45 cm3 0.01 mol/dm3 of KNO3

solution of known concentration was transferred. The pH values of the solution

were roughly adjusted from pH 2 to 12 by adding ether 0.10 mol/dm3 HCl or

NaOH on a pH meter. It was observed that for two stages a total of 3.82, 2.30 and

1.57 kg of PCP Raw, PCP 0.01 and PCP 0.05 is required to remove 99 % of 120

mg/dm3 lead (II) from 2.5 m3 solution.

164
Chapter – III Literature Review

Ofomaja et al [165] studied the biosorption of lead and copper onto Mansonia

wood sawdust. Biosorbents usually contains organic functional groups such as

alcohol, aldehydes, ketones, carboxylic, phenolic and ether groups on their

surface. These groups have been shown to participate in cation binding due to

their ability to ionize in aqueous solution. The temperature dependence of copper

and lead ion biosorption onto Mansonia wood sawdust was studied at a

concentration range of 60–140 mg/dm3 at 200 rpm with Mansonia sawdust dose of

2.0 g/dm3 at temperatures ranging from 299 to 329 K. Accurately weighed

amount (0.25 g) of Mansonia sawdust was added to five 250 ml beakers

containing 100 ml of 120 mg/dm3 of lead and copper ion solutions, each adjusted

to pH of 2.0, 3.0, 4.0, 5.0, and 6.0 using either 0.1 MHCl or NaOH solutions. The

solutions were stirred at 200 rpm for 4 h at 299 K. The optimization procedure

requires an error function to be defined in order to evaluate the fit of the equation

to the experimental data. In this study, linear coefficient of determination and

nonlinear chi-square were examined.

Ozgur Dogan Uluozlu et al [166] focused on the biosorption of Pb (II) and Cr

(III) ions onto P. tiliaceae biomass from aqueous solution. The effect of pH on the

biosorption of Pb (II) and Cr (III) ions onto P. tiliaceae biomass was studied by

changing pH values in range of 2–8. Negligible biosorption could be found when

the pH values were lower than 2.0 and also at pH < 5 the biosorption yield was in

range of 30–60 % for both Pb (II) and Cr (III) ions. The maximum biosorption

was found to be 100 % for Pb (II) and 95 % for Cr (III) ions at pH 5. Therefore,

all the biosorption experiments were carried out at pH 5.It can be seen that the

biosorption yield of Pb (II) and Cr (III) increases with rise in contact time up to 90

165
Chapter – III Literature Review

min at 20–50 oC. After this time there was no considerable increase. For instance,

at 20 oC and during 90 min, when the biosorption efficiency for Pb (II) and Cr

(III) was 99 % and 96 %, respectively. Experimental data were also tested in

terms of biosorption kinetics using pseudo-first-order and pseudo-second-order

kinetic models. The results showed that the biosorption processes of both metal

ions followed well pseudo-second-order kinetics.

Pagnanelli et al [167] studied the experimental data of lead, copper, zinc and

cadmium biosorption onto Sphaerotilus natans at different equilibrium pH. The

amount of metal adsorbed per gram of biomass weight at the equilibrium, Ceq is

the residual (equilibrium) metal concentration left in solution after binding, qmax is

the maximum possible amount of metallic ion adsorbed per unit of weight of

adsorbent and b is the equilibrium constant related to the affinity of the binding

sites for the metals. In his research a Sphaerotilus natans biomass was used as a

potential adsorbent for heavy metal removal from aqueous solutions. This

bacterium was isolated from a biological depuration plant where it is usually

present as a characteristic component of activated sludge. An original empirical

model was proposed to represent the effect of pH on heavy metal biosorption

inserting qmax vs pH empirical functions into the classical Langmuir isotherm.

Pan Jian-hua et al [168] studied the surface chemical functional groups of

Bacillus cereus biomass were identified by Fourier transform infrared (FTIR)

analytical technique. The Langmuir isotherm can yield the best fit to absorption

experimental data for two metals on the biomass. A certain amount of 3 g/L

biomass stock suspension was added to a 100 ml flask, and 1 mol/L NaNO3 was

166
Chapter – III Literature Review

used to stabilize the system at a fixed ionic strength. Bacteria cells were cultured

in 600 ml of nutrient broth for 72 h with growth condition: pH 7.2–7.5,

temperature 30 °C, and stirring speed of 160 r/min. The bacteria were then

harvested from the growth media by centrifugation at 7000 r/min for 30 min,

rinsed 4–6 times in DDI water. The mixture was then agitated on a shaker at 25°C

for 24 h, which was more than ample time for adsorption equilibrium, based on

the results of the previous kinetics experiments. After 24 h of equilibration, the

bacterial suspensions were separated by centrifugation and the supernatant was

analyzed for dissolved metal content via flame atomic absorption

spectrophotometer.

The biosorption of lead (II) by fruiting bodies of Pleurotus ostreatus immobilized

in calcium alginate was studied by Pan Xiangliang et al [4]. The kinetic

biosorption experiment was carried out as follows: 10 g of beads (wd) were added

to flasks containing 100 ml of lead-bearing solution with pH adjusted to a value of

6.5 with 0.01 M HCl and 0.01 M NaOH. Flasks were shaken at 150 rpm at 25 0C.

Aliquot amounts (2 ml) of solution were collected, periodically. Samples were

filtered and lead concentrations in filtrate were determined by AAS. The

biosorption tests were replicated twice and the standard error values ranged from

1.5 to 4 %. The maximum adsorption capacity (qmax) onto P. ostreatus

immobilized in calcium alginate based on Langmuir isotherm model was up to

121.21 mg/g for Pb (II).

Puranik et al [169] evaluated the infuence of co-cations (cadmium, copper,

cobalt and nickel) on lead and zinc biosorption by Streptoverticillium

167
Chapter – III Literature Review

cinnamoneum and Penicillium chrysogenum in binary and multimetal systems.

On the basis of the covalent index we may prepare a `covalent series' of the

cations used in the present investigation as: Pb2+ > Cu2+ > Ni2+ > Co2+ > Cd2+ >

Zn2+. The results of metal biosorption by S. cinnamoneum and P. chrysogenum

from the solutions containing mixture of cations at pH 4.0. The results of the

present studies also indicated that there was preferential uptake of lead in addition

to the direct competition effect for the binary metal systems used; Pb2+- Cd2+, Pb2+

- Zn2+, Pb2+ - Cu2+, Pb2+ - Ni2+ and Pb2+ - Co2+. The metal cations used in the

studies can be classified into three classes on the basis of the extent of metal

biosorption from multi metal solutions as; high affinity (Pb2+), intermediate affnity

(Cd2+, Cu2+ and Zn2+), and low affinity (Ni2+ and Co2+). The order of metal

biosorption in multimetal systems could be predicted well for P. chrysogenum on

the basis of Langmuir parameters evaluated in binary metal systems.

Puranik et al [170] studied the linear transformation of the adsorption data using

Freundlich and Langmuirian models that allowed the computation of the metal

adsorption capacities (r2= 0.97 for both the models). The results indicate that

0.1M sodium carbonate, 0.1 M sodium bicarbonate and 0.1 M potassium chloride

were poor desorbing agents for both lead and zinc. Among the eluents tried for

lead desorption, 0.1 M hydrochloric acid, 0.1 M nitric acid and 0.1 M EDTA were

found to be most efficient with 90 % desorption efficiency in the first desorption

cycle. In the case of zinc, in addition to these three eluents, 0.05 M sulphuric acid

could also desorb 90 % of the metal. The results obtained indicated that the

optimum pH range for the biosorption of lead and zinc by S. cinnamoneum

biomass was between 3.5–4.5 and 5.0–6.0, respectively. It has been reported

168
Chapter – III Literature Review

previously that the biosorption requires a few min to reach > 90 % of the total

metal uptake.

Rajesh Singh et al [171] studied that biosorption has been found to be

spontaneous and relatively fast process. The pure colony was preserved on the

slants at 4◦C and identified from Microbial Type Culture Collection on the basis of

spore’s morphology. The fungal biomass was prepared in the nutrient broth by

inoculating the spore suspension in the 250 ml flasks containing 100 ml of

nutrient broth on the rotary shaker cum incubator at 35 oC and 125 rpm. The

results are obtained with the experimental design that was aimed at identifying the

best levels of the selected variables, i.e.temperature (20–40 oC), adsorbent loading

(0.5–2.5 g/L), pH (2–6) and initial metal ions concentration (20–100 ppm). The

effects of temperature, adsorbent loading, pH and initial metal ions on the

biosorption were studied.

Raziya Nadeem et al [172] evaluated the biosorption of Pb (II) using dried

untreated and pretreated distillery sludge biomass. The sorption of Pb (II) ions

rapidly occurred within first 30 min followed by slow diffusion controlled step till

the equilibrium was reached. Maximum biosorption capacity of untreated and

pretreated distillery sludge was observed at pH 5. The maximum Pb (II)

adsorption was achieved with a biomass concentration of 0.5 g/L and sorption

capacity declined smoothly with increase in biomass level. The parameters such

as contact time (24 h), volume of solution (100 mL), particle size (0.250 mm) and

biomass loading (0.5 g/L) were kept constant. A significant increase in Pb (II)

uptake per weight of biomass was obtained as the pH increased from 2 to 5. The

169
Chapter – III Literature Review

spent wash is acidic in character having very high BOD (40,000–50,000 mg/L)

and COD (10,000–125,000 mg/L). The most of solid waste generated by sugar

mills are used for land fills.

Raziya Nadeem et al [173] used the locally available fish (Labeo rohita) scales

for Pb (II) removal from aqueous solutions under different experimental

conditions. Sorption experiments beyond pH 5 were not conducted since

insoluble precipitates appeared beyond this pH. The highest removal capacities

196.80 mg/g of Pb (II) ions by fish scales were obtained at pH 3.5 and the overall

removal capacity of Pb (II) decreased to 86.96 mg/g as pH increased up to 5.

Stock Pb (II) solution (1000 mg/L) was prepared by dissolving, 1.598 g of

Pb(NO3)2 in 100mL deionized water shaking it for 15 min on a magnetic stirrer to

obtain complete dissolution and then diluting quantitatively to 1000mL using

deionized water. For chemical pretreatment, 10 g of powdered sample was soaked

in 150 mL of 0.1 M HCl, H2SO4, H3PO4, NaOH, Ca(OH)2 and Al(OH)3 for 2 h at

30 oC and 100 rpm at an orbital shaker while for physical pretreatment 10 g

powdered sample was heated at 100 oC for 24 h, boiled and autoclaved at 121 oC

and 15 Psi for 30 min in 150 mL of deionized water.

Renmin Gong et al [174] studied the biosorption of lead by using intact biomass

and pretreated biomass of S. maxima. The external pH significantly influenced

lead biosorption and maximum adsorption was observed approximately at pH 5.5.

The biosorption rate of lead was rapid and equilibrium was reached at 60 min.

The removal percentage of lead was increased along with increase in sorbent dose.

At a pH below 2, the amount of lead biosorption was less. As pH increases, the

170
Chapter – III Literature Review

amount of biosorption observably increased at the range of pH 2–5.5 and the

maximum adsorption of lead ions was observed at pH 5.5. Experiments were not

conducted beyond pH 5.5 to avoid heavy metal precipitation. Among all four

eluent solutions studied, more then 90 % of lead adsorbed were desorbed from

biosorbents with nitric acid and EDTA. The lead adsorbed could be desorbed

effectively by 0.1 M nitric acid, EDTA and hydrochloric acid.

P. chrysosporium has been successfully used as the sorbing agent for removal of

metals ions from artificial wastewaters by Rõdvan Say et al [175]. The

maximum biosorption of heavy metal species on the biomass was observed at

around pH 6.0. The amounts of adsorbed heavy metal ions (Cd (II), Pb (II) and

Cu (II)) on the dry fungal biomass at pH 6.0 were found to be 13.24, 45.25 and

10.72 mg/g dry biomass, respectively. There was an increase in metal ions

adsorption per unit weight of fungal biomass with increasing pH from 2.0 to 6.0,

but it seemed to level at pH greater than 6.0. Each heavy metal ion (100 mg/L

was prepared in 150 mM NaCl solution (50 ml) and dry fungal biomass (0.2 g)

was transferred to this medium and agitated magnetically at 100 rpm. Once

inoculated, asks were incubated on an orbital shaker at 150 rpm for 7 days at

30°C. After incubation, the biomass was harvested from the medium and washed

with distilled water; it was then dried at 90 °C in an oven for 24 h.

Rong Pan et al [176] investigated the effects of single and multiple heavy metals

on the growth and uptake of consortium of two types of fungal strains, Penicillium

sp. A1 and Fusarium sp. A19. The biomasses of A1 + A19 cultivated in PDB

absorbed more metals than A1 or A19 under the different treatments of heavy

171
Chapter – III Literature Review

metals. Nevertheless, the metal concentrations absorbed by the biomass of A1 +

A19 cultivated in CDM were not significantly higher than those by A1 or A19.

However, in PDB, the fungal biomasses of A1, A19, and A1 + A19 reached the

highest values in the presence of Cu2+ compared with other treatments. The fungal

biomass cultivated in PDB in the presence of single Pb2+ was lower than those

cultivated in the media containing multimetals. Then 1 g (fresh weight) of the

washed biomass was added into each of 25 ml heavy metal solution containing

100 mg/L of Cd2+, Cu2+, Zn2+, and Pb2+, respectively, in a conical flask. The

flasks were shaken on a rotatory shaker (160 rpm) for 45 min at 25 oC. The

biomass was separated by the filteration, dried at 80 oC, and weighted. The

filtrate was digested in 67 % HNO3. The HNO3 solution was evaporated and the

solids were redissolved in 0.1 M HCl.

Ruhan Altun Anayurt et al [177] studied the use of L. scrobiculatus as a

biosorbent for removing of Pb (II) and Cd (II) ions from aqueous solution. The

biosorption efficiency was increased from 60 % to 96 % for Pb (II) and from 52 %

to 95 % for Cd (II) ion as pH was increased from 2 to 5. The maximum

biosorption was found to be 98 % for Pb (II) and 96 % for Cd (II) ions at pH 5.5.

Therefore, the remaining all biosorption experiments were carried out at this pH

value. The biosorption yield steeply increases with concentration as the biomass

concentration was increased from 0.1 to 4 g/L. The biosorption yield of Pb (II)

and Cd (II) increased considerably until the contact time reached 60 min. The

biosorption decreased from 98 % to 90 % for Pb (II) and from 95 % to 88 % for

Cd (II) as temperature was increased from 20 to 50 oC during the equilibrium

time, 60 min.

172
Chapter – III Literature Review

Runping Han et al [178] worked on the biosorption of copper and lead ions on

waste beer yeast. Batch biosorption tests were done at different contact time at the

initial concentration of 0.315 mmol/L for Cu (II) and 0.393 mmol/L for Pb (II),

respectively, and waste beer yeast dose concentration is 8 g/L in 10 mL solution.

The temperature was controlled with a water bath at the temperature of 293 K.

Experiments could not be performed at higher pH values due to hydrolysis and

precipitation of the lead ions. Flasks were agitated on a shaker for 60 min to

ensure that equilibrium was reached. There was an increase in biosorption

capacity of biomass with increasing pH from 2.0 to 6.0 for both metal ions. The

lowest metal uptake values were determined at pH < 2.0 for both metal ions. The

adsorption quantity of Pb (II) decreased with the increasing of Cu (II)

concentration. The biosorption quantity of Pb (II) decreased from 0.0218 to

0.0140 mmol/L when the concentration of Cu (II) ranged from 0 to 0.472 mmol/L.

Runping Han et al [179] studied the adsorption potential of MOCS for the

removal of Cu (II) and Pb (II) from single (noncompetitive) and binary

(competitive) aqueous systems. A volume of 20 ml of Cu (II) and Pb (II) solution

with initial concentrations ranged from 0.077 to 1.27 mmol/L and 0.097 to 1.54

mmol/L were placed in 125 ml conical flasks, respectively. The initial pH was

4.0. An accurately weighed MOCS sample (20 g/L) with particle size range from

0.99 to 0.67 mm was then added to the solution. In another binary system, the

initial concentration of Pb (II) was constant in 0.386 mmol/L, and the

concentration of Cu (II) were varied from 0 to 1.57 mmol/L. The adsorption

capacities qe obtained from the experiment results for the binary component

173
Chapter – III Literature Review

system at described conditions were ranging from 0.162 to 2.94 mol/g and 2.82 to

7.14 mol/g for Cu (II) and Pb (II), respectively

Safa Özcan et al [180] investigated the biosorption of lead (II) ions onto

Phaseolus vulgaris L. waste with the variation in the parameters like pH, contact

time, biosorbent and lead (II) concentrations and temperatures. The nature of the

possible biosorbent and metal ion interactions was examined by the FTIR

technique. The lead (II) biosorption equilibrium was attained within 20 min. This

study has shown that an agricultural byproduct, P. vulgaris L., can be used to

remove lead (II) ions from aqueous solution as a function of pH, contact time and

temperature. The maximum biosorption capacity of biosorbent for the removal of

lead (II) was obtained at pH 5. Biosorption of lead (II) ions onto P. vulgaris L.

waste followed by the Langmuir and Dubinin–Radushkevich isotherm models.

Sarabjeet Singh Ahluwalia et al [181] has desired to develop biosorbents

(microbial and plant derived biomass) with a wide range of metal affinities that

can remove a variety of metal cations. These will be particularly useful for

industrial effluents, which carry more than one type of metals. Copper

biosorption by non-living wood rotting fungus Ganoderma lucidum was studied

and it was found that protein interaction with metals did not play a significant role

in copper (II) uptake. At pH 4–5, the saturated uptake capacity for lead sorption

was higher than that of activated charcoal and that reported for some other

organisms. Dead cells of Saccharomyces cerevisiae removed 40 % more uranium

or zinc than live cultures and biosorption rapidly reached 60 % of the final uptake

value within 15 min of contact.Dried Caulerpa lentillifera was shown to have

174
Chapter – III Literature Review

adsorption potential for Cu, Cd, Pb and Zn. The adsorption of metals from binary

mixture of heavy metal was competitive and the adsorption capacity of any single

metal decreased by 10– 40 % in the presence of the other metallic species and the

overall adsorption capacity of the algae decreased by 30–50 %. Further,

adsorption capacities for these metals were found to in the order of Cr (VI) > Cd

(II) > Co (II) > Cr (III) > Ni (II) > Cu (II) > Pb (II), respectively.

Selatnia et al [182] examined that this method of eliminating Pb2+ ions is very

promising and confirm the technical and economic interest compared to the

conventional processes such as the ion exchange on resins. The quantity of

adsorbed Pb2+ per unit mass of biosorbent increased when the initial Pb2+ ions

concentration increased. In our study, initial Pb2+ concentration varies from 10 to

800 mg/L. After pretreatment of biomass at the ambient temperature, optimum

conditions of biosorption were found to be: a biomass particle size between 50

and 160 µm, an average contact time of 3 h, a biomass concentration of 3 g/L and

a stirring speed of 250 rpm. The equilibrium data could be fitted by Langmuir

isotherm equation. Under these optimal conditions, 135 mg Pb2+/g biomass was

obtained.

Senthilkumar et al [183] utilized T. conoides as a suitable biosorbent to be

utilized for lead removal in batch and column mode of operations. In column

experiments, 35.1 g of T. conoides was packed in the column to yield 25 cm bed

height with a packing density of 447.1 g/L. Sorption isotherms, obtained at

different pH (4–5) and temperature (25–35 oC) conditions were fitted using

Langmuir and Sips models. According to the Langmuir model, the maximum lead

175
Chapter – III Literature Review

uptake of 439.4 mg/g was obtained at optimum pH (4.5) and temperature (30 oC).

The influence of co-ions (Na+, K+, Mg2+ and Ca2+) on lead uptake was well

pronounced in the case of divalent ions compared to monovalent ions. The

solution of 0.1 M HCl performed well in eluting lead from the biomass, with

elution efficiency of 99.7 %.

Shihabudheen M. Maliyekkal et al [184] reported that physisorption plays a

dominant role in Pb (II) adsorption by both NMO and C–NMOC in his work with

the use of a novel cellulose-nano scale manganese oxide composite (C–NMOC)

for removing Pb (II) ions from contaminated water. The performance of C–

NMOC was compared with unsupported nanoscale-manganese oxide (NMO)

prepared through a similar route. The studies report: (i) the synthesis and detailed

characterization of NMO and C–NMOC, (ii) application of these materials for Pb

(II) ions removal from aqueous medium, and (iii) spectroscopic and microscopic

studies to probe the mechanism of Pb (II) removal by NMO and C–NMOC. Pb

(II) adsorption by C–NMOC is less sensitive to pH variations compared to NMO.

Optimum removal of Pb (II) by NMO was achieved at pH 5 whereas C–NMOC

removed more than 90 % of the Pb (II) even at an initial pH of 2.5. It is apparent

that adsorption of Pb (II) onto NMO and C–NMOC is a fast process and most of

the adsorption happened in a short period (< 10 min) and the system achieved

pseudo equilibrium at 60 min of contact time from the commencement of the

adsorption process.

Sibel Tunali et al [185] examined the biosorption of lead (II) onto

Cephalosporium aphidicola at constant temperatures of 20, 30 and 40 oC for the

176
Chapter – III Literature Review

pseudo-first-order and pseudo-second-order kinetic models and for the biosorption

isotherms at 20 oC. With an increase in biosorbent concentration, from 0.4 to 2.0

g/dm3 the percentage of lead (II) removal increased from 14.58 to 62.48 %, as the

number of binding sites would be increased, but the amount of lead (II) biosorbed

decreased from 35.80 to 31.24 mg/g. It was observed that there was almost no

change in adsorption capacity with increasing pH of the metal solution during the

biosorption process up to pH’s 4–6 and the maximum lead (II) removal was

observed at pH 5.

Sibel Tunali Akar et al [186] reported the results of lead (II) binding by the

natural and low cost biosorbent Symphoricarpus albus. A stock solution of lead

(II) was prepared by dissolving appropriate amount of anhydrous Pb(NO3)2 in 1 L

of deionized water, and the concentrations of lead (II) used in this study (50–500

mg/L) were obtained by dilution of the stock solution. The pH of the solutions

was adjusted to the desired value by adding a small quantity of 0.1 mol/L HCl, or

0.1 mol/L NaOH. The batch equilibrium studies were carried out with different

initial pH values ranging from 1.0 to 5.5 in order to study the effect of initial pH

on the biosorption capacity of S. albus biomass. The batch equilibrium studies

were carried out with different initial pH values ranging from 1.0 to 5.5 in order to

study the effect of initial pH on the biosorption capacity of S. albus biomas. The

biosorption yield for lead (II) onto S. albus biomass increased from 15.34 to

88.55% when the biomass dosage was increased from 0.6 to 4.0 g/L. In order to

determine the biosorption equilibrium time for lead (II) ions, the contact time was

varied from 10 to 70min at different temperatures

177
Chapter – III Literature Review

Sıtkı Baytak et al [187] employed the analysis of environmental water and

vegetable samples, P. digitatum immobilized on pumice stone as solid phase

extraction material. The sample solutions were adjusted to the desired pH with

diluted hydrochloric acid and/or a diluted ammonia solution. The resulting

solution was passed through the preconditioned column at a flow rate about

1mL/min and the retained metal ions were eluted with 10 mL of 1 mol/L HCl

solution. The quantitative recovery (> 95 %) was obtained at pH 6 for Cu (II) and

6–8 for Zn (II) and Pb (II).The time required for preconcentration of 500 mL of

sample solution (125 min, at flow rate of 4 mL/min), elution (10 min,at flow rate

of 1mL/min) and conditioning the column (about 5 min) was approximately 2 h.

The mean recovery of the analytes with their standard deviation (N = 5) were

found as to be 97 ± 2, 98 ± 2 and 98 ± 2% at 95 % confidence level for Cu (II), Zn

(II) and Pb (II), respectively.

Sudhir Dahiya et al [188] explored the sorption potential of pretreated crab and

arca shell biomass for lead and copper from aqueous media. The removal was

40% within 60 min and thereafter removal was slightly slow and attained

equilibrium after 4 h. Sorption behaviour of biosorbent at different dosages from

0.2 g/L to 25 g/L have been studied in 100 g/mL of solution under optimised

condition of pH and contact time. The residual concentration of lead was

decreased from 92.4 g/mL at pH 1 to less than one microgram per liter at pH 5.5

with a 99.1 % removal. For the Sorption of lead and copper on arca shell biomass

the percentage removal of copper increased from 7.7 % at pH 1 to 88.2 % at pH

4.5. The treated crab and arca shell biomasses are waste material from the fishing

178
Chapter – III Literature Review

industry can be converted in a biosorbent, which exhibit good sorption for both

the metals lead and copper.

Ting Fan et al [189] reported a very high potential for P. simplicissimum to

remove Cd (II), Zn (II) and Pb (II) from aqueous solution in his study. Maximum

biosorption capacities were obtained at pH 4.0, 6.0 and 5.0 for Cd (II), Zn (II) and

Pb (II), respectively. At low pH (< 1.0) the biosorption capacity for all metal ions

is very low, because large quantity of hydrogen ions competes with metal ions at

sorption sites. The influences of biomass dose and initial metal ion concentration

on biosorption were studied at temperature of 28 oC and pH 5.0. Equilibrium was

reached in 3 h for Cd (II), 4 h for Zn (II) and Pb (II), respectively. After the

equilibrium time, no more Cd (II), Zn (II) and Pb (II) were adsorbed. The

maximum Cd (II), Zn (II) and Pb (II) uptake capacity were determined as 52.50,

65.60 and 76.90 mg/g, respectively.

Tolga Bahadir et al [190] studied the effects of the media conditions (pH,

temperature, biomass concentration) on the biosorption of Pb (II) ions to R.

arrhizus in a batch reactor. In order to determine the effect of biomass

concentration on Pb (II) removal, stock solutions of 10.0 g/L were prepared for

different microorganism particle size ranges, such as 0.045–0.063, 0.063–0.090

and 0.090–0.125 mm. These stock solutions were diluted to 0.1–1.0 g/L standard

solutions in order to find the effect of microorganism concentration on

biosorption. To determine the effect of pH of the biosorption medium on the

initial biosorption rates of Pb (II) ions, the pH was varied between 2.0 and 5.0.

The optimum initial pH of the biosorption medium was found to be 4.5. Results

179
Chapter – III Literature Review

of removal of Pb (II) ions from storage battery industry wastewater experiments

carried out at different temperatures ranging from 20 to 45 oC.

Urdaneta et al. [191] obtained the optimal parameters for the heavy metal

adsorption after a preliminary physical and chemical characterization of the

vermicompost. A synthetic multielemental solution of Pb, Cr and Ni and a

solution of NH4VO3 for vanadium were evaluated. The effect of pH on the

adsorption efficiency was evaluated in a range of 3.5 to 7.0 using aqueous solution

of NaOH or HNO3 of 0.10 mol/L. The effects of the agitation time were

evaluated using 5, 10, 15, 30, 45 and 60 min. The solution volume was fixed to

75 mL, the pH to 5.00, the vermicompost mass was 1.0 g and the particle size

ranging between 75 and 841 μm. Removing levels for Pb, Cr and Ni are in the

order of 95 % showing that vermicompost is adequate for this removing process.

Metal adsorption for Ni, Pb and Cr in real wastewater from Barquisimeto city was

100, 65 and 40 % respectively.

Vimala et al [192] studied the sorption capacity of oyster mushroom (Pleurotus

platypus), button mushroom (Agaricus bisporus) and milky mushroom (Calocybe

indica). When the pH values increased, biosorbent surfaces were more negatively

charged and the biosorption of metal ions (positive charge) increased and reached

equilibrium at pH 6.0 for cadmium and pH 5.0 for lead. Decrease in biosorption

at higher pH (pH > 6.0 and > 5.0) is due to the formation of soluble hydroxylated

complexes of the metal ions the removal efficiency reached equilibrium at 60 min

and that by A. bisporus and C. indica reached equilibrium at 240 min. The

respective removal efficiencies of lead by P. platypus, A. bisporus and C. indica

180
Chapter – III Literature Review

reached at 120, 240 and 180 min. Langmuir isotherm model exhibited better fit to

the sorption data of both cadmium (II) and lead (II) for all the three mushrooms

than the Freundlich isotherm model. Higher values of qmax (34.96 mg/g) and Kf

(4.9980) in case of P. platypus for cadmium sorption and qmax (33.78 mg/g) and Kf

(4.9522) for lead sorption of A. bisporus indicate their greater suitability for

removal for cadmium (II) and lead (II) ions, respectively from wastewater.

Vinod K. Gupta et al [193] reported the ability of two non-living (dried) fresh

water algae, Oedogonium sp. and Nostoc sp. to remove lead (II) from aqueous

solutions in batch system. It has been observed that maximum adsorption took

place within first 90 and 70 min for Oedogonium sp. and Nostoc sp., respectively.

The extent of lead removal was found to be 11.1 and 16.4 % for 0.1 g/L of algal

biomass Oedogonium sp. and Nostoc sp., respectively. It greatly increased to 63.5

and 70 % for 0.5 g/L of biosorbent, respectively. As the pH of the lead solution

(200 mg/L) increased from 2.99 to 7.04, the adsorption capacity of lead increased

upto pH 5.0 and then dramatically decreased at pH 7.04 for both the algae. For an

increase in temperature from 25 to 45 oC, an increase in the adsorption of lead was

observed (111.4 mg/g at 25 oC to 124.8 mg/g at 45 oC at 200 mg/L, initial lead

concentration

Wei-Bin Lu et al [194] studied the biosorption kinetics and equilibria of lead

(Pb), copper (Cu) and cadmium (Cd) ions using the biomass of Enterobacter sp.

J1 isolated from a local industry wastewater treatment plant. The biosorbent was

suspended in 50 ml of heavy metal solutions (100 mg/L) in a glass container to

reach a cell concentration of 1.5–2.5 g/L. Metal loaded and metal free (control)

181
Chapter – III Literature Review

biosorbents were treated with glutaraldehyde for 1 h and were dehydrated by

acetone (50–100 %) for 30 min. Both models had a good agreement with the data

for Cu and Cd biosorption, evidenced by the high r2 values (all greater than

0.978). However, prediction of Pb adsorption by Langmuir isotherm was less

precisely with a lower r2 value of 0.883, in contrast to better fits of the Pb

adsorption data by Freundlich model (r2 = 0.970).Desoprtion of Cd was nearly

complete for pH 3, while desorption efficiency of Cu and Pb reached 90 % when

pH was decreased to below 2.0.

The biosorption of lead (II) by fruiting bodies of pleurotus ostreatus immobilized

in calcium alginate was studied by Xiangliang et al. [4]. The kinetic biosorption

experiment was carried out as follows: 10 g of beads (wd) were added to flasks

containing 100 mL of lead-bearing solution with pH adjusted to a value of 6.5

with 0.01 M HCl and 0.01 M NaOH. Flasks were shaken at 150 rpm at 25 C.

Aliquot amounts (2 mL) of solution were collected, periodically. Samples were

filtered and lead concentrations in filtrate were determined by AAS. The

biosorption tests were replicated twice and the standard error values ranged from

1.5 to 4 %. The maximum adsorption capacity (qmax) onto P. ostreatus

immobilized in calcium alginate based on Langmuir isotherm model was up to

121.21 mg/g for Pb (II).

Xiao-ming Li et al [152] reported that PSILS is an efficient biosorbent for

removal of heavy metal ions from aqueous solution. The effect of pH on the

biosorption capacity was investigated at initial pH values range of 2.0–6.0 and the

desired pH of the suspensions was maintained by adding HCl or NaOH at the

182
Chapter – III Literature Review

beginning of the experiment and not controlled afterwards. The results showed

that the biosorption of Pb (II) and Cu (II) on PSILS increased significantly as the

pH increased from 2.0 to 5.0, however, the increase was slightly on NLS, and then

decreased at pH 5.0–6.0. The low Pb (II) and Cu (II) biosorption capacity at pH

values below 3.0 may be attributed to hydrogen ions that compete with metal ions

on the sorption sites. The contact time of approximately 60minwas required to

reach the equilibrium. The effect of initial single metal ion concentration was

investigated in the range of 10–500 mg/L. sites. According to Langmuir isotherm,

the monolayer saturation capacity of PSILS is 144.9 mg/g for Pb (II) and 106.4

mg/g Cu (II).

Yasemin Bakircioglu et al [195] studied the filamentous fungal biomass loaded

TiO2 nano particles were used for the biosorption of lead (II) ions. Maximum

retention was maintained within the pH of 4.0–10.0 which is quite a wide range.

Therefore, a precise pH adjustment above 4 was not necessary. Nevertheless, the

pH of the solutions was adjusted to 4.0–4.5 throughout this study. In this study,

various concentrations of HCl and HNO3 (0.5–2.0 mol/L) were used for

desorption of the lead ions from the biosorbent following the on-line

preconcentration procedure. The highest signal was obtained with 1.0 mol/L HCl

whereas relatively lower signals and incomplete elution of the lead from the

biosorbent column was observed with HNO3 solutions. Results showed that the

absorbance was higher at flow rate of 4.3 mL/min by using distilled-deionized

water as a carrier.

183
Chapter – III Literature Review

Yongwen Liu et al [196] studied the biosorption behavior of the solid waste

Chinese herb Pang Da Hai (seeds of Sterculia lychnophera Hance) as a sorbent for

trace lead and cadmium. The optimum pH ranges for metal adsorption were 5.0–

8.0 for Pb and 6.0–8.0 for Cd. So pH 6.5 was selected for the recommended

procedure. It was observed that temperature changes between 15 and 45 ◦C have

slightly influence on the rate of biosorption, but did not affect the biosorption

capacity. The reproducibility of the proposed procedure was also measured using

model solution. The relative standard deviations (n = 10) were 7.2 % for Pb and

4.6 % for Cd.

Yurtsever and Ahyan [197] investigated the effect of temperature, pH and initial

metal concentration on Pb (II) biosorption on modified quebracho tannin resin

(QTR). The specific BET surface area of QTR was found to be 0.820 m2/g. The

adsorption of Pb (II) increases with time from 0 to 10 min and then becomes

almost constant up to the end of experiment. The analysis is carried out through

an adsorption of N2 gas (gas effluent temperature = 75 C versus bath temperature

= 77.35 C). The kinetic data was tested using pseudo-first-order, pseudo-second-

order, Elovich and intraparticle diffusion model. It was found that the kinetics Pb

(II) ions adsorption onto QTR followed by the pseudo-second-order rate equation.

Langmuir, Freundlich and Temkin adsorption models were used to represent the

equilibrium data. Langmuir model best describes the lead adsorption process (R2

> 0.999). Using the Langmuir model equation, the monolayer adsorption capacity

of QTR was found to be 86.207 mg/g for Pb (II) ions.

184
Chapter – III Literature Review

Yus Azila et al [198] investigated about response surface methodology (RSM) for

obtaining optimum process condition for Pb (II) removal using immobilized cells

of P. sanguineus as a biosorbent. It was observed that more than 90 % removal

was achieved with the initial Pb (II) ions concentration increased from 50 to 350

mg/L at pH 4. The interaction shows that for a higher biomass loading more than

9 g/L, the percentage of Pb (II) ions removal was more than 97 % with initial Pb

(II) ions concentration in the range of 50–350 mg/L. At solution pH lower than 3,

less than 80 % removal was achieved due to competition between hydrogen and

metal ions on the binding sites of biosorbent. As the solution pH increased more

than pH 4.0, which resulted in precipitation of Pb (II) ions thus, lower the

biosorption efficiency.

Zhexian Xuan et al [199] studied about preparation of a new biosorbent using

chemically modified orange peel and its biosorption of lead ion. As the pH

increases, the amount of lead uptake increases and the sharpest increase is

observed between pH 4.5 and 6.0 for all the biosorbents. When the final lead

concentrations are increased from 0.5 to 10.0 mmol/L, the uptake of lead

increased. But it has no change when the final concentrations are more than 10.0

mmol/L. The maximum removal rate is more than 90 % when the s/l ratio is 3.0–

5.0 g/L. Thus the optimum s/l ratio is 3.0 g/L in terms of the cost effect.

3.3 RESPONSE SURFACE METHODOLOGY (RSM):

Aleboyeh et al [200] studied the decolorization of C.I. Acid Red 14 (AR14) azo

dye by electro coagulation (EC) process in a batch reactor. Response surface

185
Chapter – III Literature Review

methodology (RSM) was applied to evaluate the simple and combined effects of

the three main independent parameters, current density, and time of electrolysis

and initial pH of the dye solution on the color removal efficiency and optimising

the operating conditions of the treatment process. A 23 full factorial central

composite face centred (CCF) experimental design was employed. Analysis of

variance (ANOVA) showed a high coefficient of determination (R2 =0.928) and

satisfactory second-order regression model was predicted. Maximum color

removal efficiency was predicted and experimentally validated. The optimum

current density, time of electrolysis and initial pH of the dye solution were found

to be 102 Am-2, 4.47 min and 7.27, respectively. Under optimal value of process

parameters, removal of > 91 % was obtained for Acid Red 14. This study clearly

showed that response surface methodology was one of the suitable methods to

optimize the operating conditions and maximize the dye removal. Graphical

response surface and contour plots were used to locate the optimum point.

Antonio et al [201] employed a factorial design to evaluate the quantitative

removal of an anionic red dye from aqueous solutions to epichlorohydrin-cross-

linked chitosan. The experimental factors and their respective levels studied were

the initial dye concentration in solution (25 to 600 mg/L), the absence or the

presence of the anionic surfactant sodium dodecylbenzenesulfonate (DBS) and the

adsorption temperature (25 or 55 oC). The adsorption parameters were analyzed

statistically using modeling polynomial equations. The results indicated that

increasing the dye concentration from 25 to 600 mg/L increased the dye

adsorption whereas the presence of DBS increased it. The principal effect of

temperature did not show a high statistical significance. The factorial results also

186
Chapter – III Literature Review

demonstrated the existence of statistically significant binary interactions of the

experimental factors. The adsorption thermodynamic parameters, namely AadsH,

AadsG and AadsS, were determined for all the factorial design results.

Exothermic and endothermic values were found in relation to the AadsH. The

positive AadsS values indicated that entropy was a driving force for adsorption.

The AadsG values were significantly affected by an important synergistic effect of

the factors and not by the temperature changes alone.

Antonio et al [202] employed a 23 factorial design to evaluate the quantitative

removal of the indigo carmine (IC) dye from aqueous solutions on glutaraldehyde

cross-linked chitosan. The variables were chitosan masses of 100 and 300 mg, IC

concentrations of 2.0 and 5.0 x 10-5 mol/L and temperatures of 25 and 350 ºC.

The quantitative and energetic adsorption parameters were analyzed statistically

using modeling with bilinear equations. The results indicated that increasing the

chitosan mass from 100 to 300 mg decreased the IC adsorption/mass ratio (mol/g)

whereas a temperature increase of 25-35 ºC increased it. The principal effect of

the IC concentration did not show statistical significance. The factorial

experiments demonstrated the existence of a significant antagonistic interaction

effect between the chitosan mass and temperature. The adsorption

thermodynamic parameters, namely AadsH, AadsG, and AadsS, were determined

for all the factorial design results. Endothermic values were found in relation to

the AadsH. The positive AadsS values indicated that entropy was a driving force

for adsorption. The AadsG values were also significantly affected by important

antagonistic and synergistic effects involving all principal and interactive factors.

187
Chapter – III Literature Review

The thermodynamically spontaneity of the IC adsorption parameters were greatly

influenced by the interactive factors and not by the temperature changes alone.

Ayla Ozer et al [203] investigated the biosorption of copper (II) ions on

Enteromorpha prolifera, green seaweed, in a batch system. The effects of

operating parameters such as initial pH, temperature, initial metal ion

concentration and biosorbent concentration on the copper (II) biosorption were

analysed using response surface methodology (RSM). The proposed quadratic

model for central composite design (CCD) fitted very well to the experimental

data that it could be used to navigate the design space according to ANOVA

results. The optimum biosorption conditions were determined as initial pH 4.0,

temperature 25 oC, biosorbent concentration 1.2 g/L and initial copper (II) ion

concentration 200 mg/L. The Langmuir and Freundlich isotherm models were

applied to the equilibrium data at different temperatures and initial pH values.

The maximum monolayer coverage capacity of E. prolifera for copper (II) ions

was found to be 57.14 mg/g at 25 ºC and initial pH 4.0 indicating that the

optimum biosorption temperature and initial pH. The external and intraparticle

diffusion models were also applied to biosorption data of copper (II) ions with E.

prolifera and it was found that both the external diffusion as well as intraparticle

diffusion contributes to the actual biosorption process. The pseudo-second order

kinetic model described the copper (II) biosorption process with a good fitting.

The thermodynamics of copper (II) biosorption on E. prolifera indicated its

spontaneous and exothermic nature.

188
Chapter – III Literature Review

Ayla Ozer et al [204] investigated the biosorption of nickel (II) ions on

Enteromorpha prolifera, a green algae, in a batch system. The single and

combined effects of operating parameters such as initial pH, temperature, initial

metal ion concentration and biosorbent concentration on the biosorption of nickel

(II) ions on E. prolifera were analyzed using response surface methodology

(RSM). The optimum biosorption conditions were determined as initial pH 4.3,

temperature 27 ºC, biosorbent concentration 1.2 g/L and initial nickel (II) ion

concentration 100 mg/L. At optimum biosorption conditions, the biosorption

capacity of E. prolifera for nickel (II) ions was found to be 36.8 mg/g after 120

min biosorption. The Langmuir and Freundlich isotherm models were applied to

the equilibrium data and defined very well both isotherm models. The monolayer

coverage capacity of E. prolifera for nickel (II) ions was found as 65.7 mg/g. In

order to examine the rate limiting step of nickel (II) biosorption, such as the mass

transfer and chemical reaction kinetics, the intraparticle diffusion model, external

diffusion model and the pseudo second order kinetic model were tested with the

experimental data. It was found that for both contributes to the actual biosorption

process. The pseudo second order kinetic model described the nickel (II)

biosorption process with a good fitting.

Azharul Islam et al [205] dealt with the multiple response optimization for the

removal of organ phosphorus pesticide quinalphos [QP: O, O-diethyl O-2-

quinoxalinyl phosphorothioate] from the aqueous solution onto low-cost material

and tried to overcome the drawbacks of univariate optimization. Used tea leaves

were used as low-cost adsorbent and batch equilibration method was followed for

this study. A Box-Behnken design was used to develop response model and

189
Chapter – III Literature Review

desirability function was then used for simultaneous optimization of all affecting

parameters in order to achieve the highest removal % of quinalphos. The

optimum conditions of factors predicted for quinalphos removal % were found to

be: pH 8.83, concentration 7 mg/L and dose 0.40 g. Under these conditions,

maximum removal % of quinalphos was obtained 96.31 %. Considering the

above optimum conditions, the adsorption isotherms were developed and provided

adsorption capacity of 196.07 (mg/g) by using Langmuir equation, indicating that

used tea leaves may be applied as a low-cost material for pesticides removal from

aqueous matrices.

Bala Kiran et al [206] presented the chromium adsorptive potential of

polysaccharide produced by a novel cyanobacterium Lyngbya putealis HH-15.

Batch mode experiments were performed to determine the adsorption equilibrium

and the equilibrium data was applied to different two-parameter models

(Langmuir, Freundlich, Temkin, Dubinin-Radushkevich, Flory-Huggins, and

Brunauer, Emmer & Teller (BET) model). The highest coefficient of

determination (0.9925) for Langmuir and BET models indicates best fitness of

these models in explaining the sorption as a multilayer process. Effect of different

variables like initial metal ion concentration (10-100 mg/L), pH (2-6) and

temperature (25-45 ºC) on chromium adsorption of exopolysaccharides (EPS)

were also examined, using Box-Behnken design model. Very high value of

regression coefficient between the variables and the response (R2 = 0.9982)

indicates excellent evaluation of experimental data by second-order polynomial

regression model. The response surface method indicated that 30-40 mg/L initial

chromium concentration, pH = 2 and 45 ºC temperature were optimal for

190
Chapter – III Literature Review

biosorption of chromium by the cyanobacterial exopolysaccharides. Adsorption

capacity of EPS increased from 45 to 157 mg/g of EPS as initial Cr (VI)

concentration increased from 10 to 30 mg/L. Surface adsorption of the metal at

surface of EPS was confirmed through scanning electron microscopy.

Bala Kiran and Anubha Kaushik [207] investigated the potential use of alginate

immobilized algal beads for the removal of chromium from aqueous solution

under optimized conditions using a novel cyanobacterium, Lyngbya putealis

isolated from metal contaminated soil. Batch mode experiments were performed

to determine the adsorption equilibrium and kinetic behavior of chromium in

aqueous solution allowing the computation of kinetic parameters and maximum

metal adsorption capacity. Influences of other parameters like initial metal ion

concentration (10-100 mg/L), pH (2-6) and temperature (25-45 ºC) on chromium

adsorption were also examined, using Box-Behnken design. Very high regression

coefficient between the variables and the response (R2 = 0.9984) indicates

excellent evaluation of experimental data by second-order polynomial regression

model. The response surface method indicated that 50-60 mg/L initial chromium

concentration, 2-3 pH and a temperature of 45 ºC were optimal for biosorption of

chromium by immobilized L. putealis, when 82 % of the metal is removed from

the solution.

Baral et al [208] studied the Cr (VI) adsorption efficiency of the seaweed,

Hydrilla verticillata, in batches. The adsorbent was characterized using SEM,

BET surface area analyzer, Malvern particle size analyzer, EDAX and FT-IR. Cr

(VI) removal efficiency of the adsorbent was studied as a function of different

191
Chapter – III Literature Review

adsorption parameters such as contact time, stirring speed, pH, adsorbent dose,

particle size, adsorbate concentration, and temperature. Langmuir, Freundlich,

and Temkin adsorption isotherm equations were used in the equilibrium modeling.

The adsorption process followed pseudo second-order kinetics and intra-particle

diffusion was found to be the rate-controlling step. Experimental data follow

Langmuir adsorption isotherm. Thermo-dynamic parameters such as Gibbs free

energy and enthalpy of the adsorption process were evaluated to find out the

feasibility of the adsorption process. The negative values of Gibb's free energy

and positive enthalpy values show the feasibility and endothermic nature of the

process. The significance of different adsorption parameters along with their

combined effect on the adsorption process has been established through a full 24

factorial design. Among the different adsorption parameters, pH has the most

influential effect on the adsorption process followed by adsorbate concentration

and combined effects of all the four parameters were tested. The correlation

among different adsorption parameters were studied using multi-variate analysis.

Corneliu et al [209] investigated the ability of dried yeast Saccharomyces

biomass to remove Cu (II) ions from aqueous solutions by using of batch

techniques. The influence of different parameters on copper uptake by dried

yeast, such as initial Cu (II) concentration, initial pH of solution and temperature,

was studied. The Freundlich, Langmuir, Redlich-Peterson and Sips isotherms

were applied to the obtained experimental data. According to Langmuir isotherm

the maximum adsorption capacity of investigated non-living biomass was found to

be 2.59 mg/g. The thermodynamic parameters (e.g. free energy and enthalpy)

were calculated and discussed. The adsorption of Cu (II) onto the dried cells of

192
Chapter – III Literature Review

Saccharomyces cerevisiae is an endothermic process and become more favorable

with the increasing of temperature in pH range from 3 to 4. Optimization studies

by means of response surface methodology were carried out, which resulted in

improvement of the efficiency of sorption removal by using of biomass. The

removal efficiency of real wastewater originating from electroplating industry

which contains Sn (II) ions was determined and compared with synthetic

wastewater obtained in laboratory.

Farshid Ghorbani et al [210] Optimized cadmium biosorption process was

performed by varying three independent parameters (initial pH, initial cadmium

ion concentration, Saccharomyces cerevisiae dosage) using a central composite

design (CCD) under response surface methodology (RSM). For the maximum

biosorption of cadmium ion in an aqueous solution by S. cerevisiae, a total of 20

experimental runs were set and the experimental data fitted to the empirical

second-order polynomial model of a suitable degree. The potential of S.

cerevisiae as a bioadsorbent was evaluated as a pretreated material with 700 g/L

of ethanol. Furthermore, the quantitative relationship between the heavy metal

uptake (q) and different levels of these factors was used to work out optimized

levels of these parameters by a full factorial design (23). The analysis of variance

(ANOVA) of the quadratic model demonstrates that the model was highly

significant. The best set required 5 as initial pH, 3.8 g/L S. cerevisiae and 19

mg/L cadmium ion concentration within 240 min of contact time. Three

dimensional plots demonstrate relationships between the cadmium ion uptake with

the paired factors (when other factor was kept at its optimal level), describing the

behavior of biosorption system in a batch process. The model showed that

193
Chapter – III Literature Review

cadmium uptake in aqueous solution was affected by all the three factors studied.

An optimum cadmium uptake of 6.71 mg/g biomass was achieved at initial

cadmium ion concentration of 26.46 mg/L and S. cerevisiae dosage of 2.13 g/L.

The process kinetics were also evaluated by isotherm, pseudo-second-order and

intra-particle diffusion models. It showed that both monolayer adsorption and

intra-particle diffusion mechanisms were effective in the cadmium biosorption

process. Therefore, it is apparent that the response surface methodology not only

gives valuable information on interactions between the factors but also leads to

identification of feasible optimum values of the studied factors.

Harapriya and Rani Gupta [211] reported biosorption of Zn (II), Cu (II) and Co

(II) onto O. angustissima biomass from single, binary and ternary metal solutions,

as a function of pH and metal concentrations via Central Composite Design

generated by statistical software package Design Expert 6.0. The experimental

design revealed that metal interactions could be best studied at lower pH range i.e.

4.0, 5.0, which facilitates adequate availability of all the metal ions. The sorption

capacities for single metal decreased in the order Zn (II) > Co (II) > Cu (II). In

absence of any interfering metals, at pH 4.0 and an initial metal concentration of

0.5 mM in the solution, the adsorption capacities were 0.33 mmol/g Zn (II), 0.26

mmol/g Co (II) and 0.12 mmol/g Cu (II). In a binary system, copper inhibited

both Zn (II) and Co (II) sorption but the extent of inhibition of former was greater

than the latter; sorption values being 0.14 mmol/g Zn (II) and 0.27 mmol/g Co (II)

at initial Zn (II) and Co (II) concentration of 1.5 each, pH 4.0 and 1 mM Cu (II) as

the interfering metal. Zn (II) and Co(II) were equally antagonistic to each other’s

sorption; Zn (II) and Co (II) sorption being 0.23 and 0.24 mmol/g, respectively, at

194
Chapter – III Literature Review

initial metal concentration of 1.5 mM each, pH 4.0 and 1 mM interfering metal

concentration. In contrast, Cu (II) sorption remained almost unaffected at lower

concentrations of the competing metals. Thus, in binary system inhibition

dominance observed was Cu (II) > Zn (II), Cu (II) > Co (II) and Zn (II) > Co (II),

due to this the biosorbent exhibited net preference/ affinity for Cu (II) sorption

over Zn (II) or Co (II). Hence, the affinity series showed a trend of Cu (II) > Co

(II) > Zn (II). In a ternary system, increasing Co (II) concentration exhibited

protection against the inhibitory effect of Cu (II) on Zn (II) sorption. On the other

hand, the inhibitory effect of Zn (II) and Cu (II) on Co (II) sorption was additive.

The model equation for metal interactions was found to be valid within the design

space.

Hasan et al [212] successfully utilized the biomass of Aeromonas hydrophila for

the removal of lead from aqueous solution. The effects of process variables such

as pH, initial Pb (II) concentration, biomass dose and temperature on the uptake of

lead were investigated using two level four factor (24) full factorial central

composite designs with the help of MINITAB version15 software. The predicted

results thus obtained were found to be in good agreement (R2 = 98.6 %) with the

results obtained by performing experiments. The multiple regression analysis and

analysis of variance (ANOVA) showed that the concentration has positive and

temperature and biomass dose have negative whereas pH has curved relationship

with the uptake of Pb (II). The maximum uptake of Pb (II) predicted by

optimization plots was 122.18 mg/g at 20 oC, initial Pb (II) concentration of 259

mg/L, pH 5.0, temperature 20 oC and biomass dose 1.0 g. Langmuir isotherm

model was applicable to sorption data and sorption capacity was found to be 163.3

195
Chapter – III Literature Review

mg/g at 30 ºC, pH 5.0 and Pb (II) concentration range 51.8-259 mg/L indicate that

the biosorbent was better in comparison of the biosorbent reported in the

literature. Dubinin-Radushkevich (D-R) isotherm model was also applied and it

was found that sorption was chemisorptions (E = 12.98 kJ/mol). FT-IR studies

indicate the involvement of various functional groups present on biomass surface

in the sorption of Pb (II).

Helen Kalavathy et al [213] investigated adsorption capacity of Cu2+ from

aqueous solution onto H3PO4 activated carbon using rubber wood sawdust

(RSAC) in a batch system. Kinetic and isotherm studies were carried out, the

thermodynamic parameters like standard Gibb's free energy (ΔG), enthalpy (ΔH)

and entropy (ΔS) were evaluated. The pseudo-second-order model was found to

explain the kinetics of Cu2+ adsorption most effectively. The process optimization

was performed through Central Composite Rotary Design using response surface

methodology (RSM) by Design Expert Version 5.0.7 (STAT-EASE Inc.,

Minneapolis, USA). An initial concentration of 35 mg/L, temperature of 260 ºC,

carbon loading of 0.45 g (100 mg/L), adsorption time 208 min and pH of 6.5 was

found to be the optimum conditions for the maximum uptake of copper ions of 5.6

mg/L in batch mode.

Javad Zolgharnein et al [214] in their study introduced Robinia tree leaves as a

novel and efficient biosorbent for removing Pb (II) from aqueous solutions. In

order to reduce the large number of experiments and find the highest removal

efficiency of Pb (II), a set of full 23 factorial design with two blocks were

performed in duplicate (16 experiments). In all experiments, the contact time was

196
Chapter – III Literature Review

fixed at 25 min. The main interaction effects of the three factors including sorbent

mass, pH and initial concentration of metal-ion were considered. By using

Student's t-test and analysis of variances (ANOVA), the main factors, which had

the highest effect on the removal process, were identified. Twenty-six

experiments were designed according to Doehlert response surface design to

obtain a mathematical model describing functional relationship between response

and main independent variables. The most suitable regression model, that fitted

the experimental data extremely well, was chosen according to the lack-of-fit-test

and adjusted R2 value. Finally, after checking for possible outliers, the optimum

conditions for maximum removal of Pb (II) from aqueous solution were obtained.

The best conditions were calculated to be as: initial concentration of Pb (II) = 40

mg/L, pH = 4.6 and concentration of sorbent equal to 27.3 g/L.

Junkal Landaburu et al [215] conducted removal of cadmium and zinc from

their respective water samples by micellar-enhanced ultra filtration (MEUF),

using sodium dodecyl sulfate (SDS) as the surfactant. Response surface

methodology (RSM) was used for modeling and optimizing the process, and to

gain a better understanding of the process performance. Face Centered Composite

(CCF) Design was used as the experimental design. The factors studied were

pressure (P), nominal molecular weight limit (NMWL), heavy metal feed

concentration (CZn, CCd) and SDS feed concentration (CSDS). Using RSM the

retention of heavy metals was maximized while olarizati the surfactant to metal

ratio (S/M). Response surface plots improved the understanding the effect of the

factors on permeate flux. Concentration olarization was negligible and

therefore, high NMWL membranes with high pressure provided high flux with

197
Chapter – III Literature Review

negligible effect on the retention of heavy metals. The optimal conditions of zinc

removal were CSDS = 13.9 mM, CZn = 0.5 mM, NMWL = 10 kDa and P = 3.0

bar, and for cadmium removal CSDS = 14.2 mM, CCd = 0.5 mM, NMWL = 10

kDa and P = 3.0 bar. The retentions achieved were 98.0 ± 0.4 % for zinc and 99.0

± 0.4 % for cadmium. To improve resource efficiency, the surfactant was

reclaimed after use; 84 % of the initial SDS was recovered by precipitation.

Kaan et al [216] employed a three factor, three-level Box-Behnken experimental

design combining with response surface modeling (RSM) and quadratic

programming (QP) for maximizing Pb (II) removal from aqueous solution by

Antep pistachio (Pistacia vera L.) shells based on 17 different experimental data

obtained in a lab-scale batch study. Three independent variables (initial pH of

solution) pH ranging from 2.0 to 5.5, initial concentration of Pb (II) ions (C0)

ranging from 5 to 50 ppm, and contact time (t) ranging from 5 to 120 min were

consecutively coded as x1, x2 and x3 at three levels (-1, 0 and 1), and a second-

order polynomial regression equation was then derived to predict responses. The

significance of independent variables and their interactions were tested by means

of the analysis of variance (ANOVA) with 95 % confidence limits (a = 0.05). The

standardized effects of the independent variables and their interactions on the

dependent variable were also investigated by preparing a Pereto chart. The

optimum values of the selected variables were obtained by solving the quadratic

regression model, as well as by analysing the response surface contour plots. The

optimum coded values of three test variables were computed as x1 = 0.125, x2 =

0.707, and x3 = 0.107 by using a LOQO/AMPL optimization algorithm. The

experimental conditions at this global point were determined to be pH = 3.97, Co

198
Chapter – III Literature Review

= 43.4 ppm, and t = 68.7 min, and the corresponding Pb (II) removal efficiency

was found to be about 100 %.

Khattar and Shailza [217] used Response surface methodology (RSM) to

optimize the critical parameters responsible for higher Cd2+ removal by a

unicellular cyanobacterium Synechocystis pevalekii. A three-level Box-Behnken

factorial design was used to optimize pH, biomass and metal concentration for

Cd2+ removal. A coefficient of determination (R2) value (0.99), model F-value

(86.40) and its low p-value (F < 0.0001) along with lower value of coefficient of

variation (5.61 %) indicated the fitness of response surface quadratic model during

the present study. At optimum pH (6.48), biomass concentration (0.25 mg protein

ml-1) and metal concentration (5 gm/L) the model predicted 4.29 gm/L Cd2+

removal and experimentally, 4.27 gm/L Cd2+ removal was obtained.

Malihe et al [218] investigated the effects of biosorbent Aspergillus niger dosage,

initial solution pH and initial Ni (II) concentration on the uptake of Ni (II) by

NaOH pretreated biomass of A. niger from aqueous solution. Batch experiments

were carried out in order to model and optimize the biosorption process. The

influence of three parameters on the uptake of Ni (II) was described using a

response surface methodology (RSM) as well as Langmuir and Freundlich

isotherm models. Optimum Ni (II) uptake of 4.82 mg/gbiomass (70.30 %) was

achieved at pH 6.25, biomass dosage of 2.98 g/L and initial Ni (II) concentration

of 30.00 mg/L Ni (II). Langmuir and Freundlich were able to describe the

biosorption isotherm fairly well. However, prediction of Ni (II) biosorption using

Langmuir and Freundlich isotherms was relatively poor in comparison with RSM

199
Chapter – III Literature Review

approaches. The biosorption mechanism was also investigated by using Fourier

transfer infrared (FT-IR) analysis of untreated, NaOH pretreated, and Ni (II)

loaded A. niger biomass.

Malihe et al [219] investigated, the process of cadmium biosorption on NaOH

pretreated Aspergillus niger biomass in the batch mode. The effect of three

independent variables, initial pH of solution (1.3-8.7), biomass dosage (0.1-7.5

g/l) and initial cadmium ion concentration (0.5-37.5 mg/L) on the biosorption

process was determined and the process was then optimized by means of response

surface methodology (RSM). The process was evaluated by cadmium removal

efficiency as the process response. Twenty experiments designed by central

composite design (CCD) were carried out and the process response was modeled

using a polynomial equation as function of the variables. The optimum values of

the variables were found to be 5.96, 30.0 mg/L and 1.6 g/L for initial pH, initial

cadmium ion concentration and biomass dosage, respectively, at contact time of

1440 min. At optimal conditions, a biosorption capacity of 10.14 mg/g biomass

was obtained corresponding to 82.2 % cadmium removal efficiency. Under this

condition, a desirability value of 0.903 was obtained, showing that the estimated

function may represent the experimental model and give the desired conditions.

According to these observations, biomass A niger fungus particles with clean

surface and high porosity may have application as biosorbent for heavy metal

removal from industrial wastewater effluents.

Mevra et al [220] applied response surface methodology to optimize the removal

efficiency of Ni (II) using Pinus sylvestris ovulate cones. A 23 full-factorial

200
Chapter – III Literature Review

central composite design was employed for experimental design and analysis of

the results. The initial Ni (II) concentration (10-30 mg/L), pH (2.5-6.5) and

biomass concentration (5-25 g/L) were the critical components of the removal

optimized. The optimum pH, m (biomass concentration) and Co (initial Ni (II)

concentration) were found to be 6.17, 18.8 g/L and 11.175 mg/L, resp ectively.

Under these conditions, removal efficiency of Ni (II) was 99.91 %.

Mohammad et al [221] investigated the biosorption of Cd (II) and Ni (II) from

aqueous solution onto Saccharomyces cerevisiae and Ralstonia eutropha non-

living biomass. Biomass inactivated by heat and pretreated by ethanol was used in

determination of optimum conditions. The important process parameters, such as

initial solution pH (2-8), initial Ni (II) concentration (11-42 mg/L), initial Cd (II)

concentration (11-42 mg/L), and biomass dosage (0.2-4.7 g/L) were optimized

using design of experiments (DOE). A central composite design (CCD) under

response surface methodology (RSM) was applied to evaluate and optimize the

efficiency of removing each adsorbent. Moreover, the two responses were

simultaneously studied by using a numerical optimization methodology. The

optimum removal efficiency of Cd (II) and Ni (II) onto S. cerevisiae was

determined as 43.4 and 65.5 % at 7.1 initial solution pH, 4.07 g/L biomass dosage,

16 mg/L initial Ni (II) concentration and 37 mg/L initial Cd (II) concentration.

The optimum removal efficiency of Cd (II) and Ni (II) onto R. eutropha was

ascertained as 52.7 and 50.1 % at 5.0 initial solution pH, 2.32 g/L biomass dosage,

28 mg/L initial Ni (II) concentration and 37 mg/L initial Cd (II) concentration.

The present analysis suggests that the predicted values were in good agreement

with experimental data. The characteristics of the possible interactions between

201
Chapter – III Literature Review

biosorbents and metal ions were also evaluated by scanning electron microscope

(SEM) and Fourier transform infrared (FT-IR) spectroscopy analysis.

Olga Freitas et al [222] was used A Box-Behnken factorial design coupled with

surface response methodology to evaluate the effects of temperature, pH and

initial concentration in the Cu (II) sorption process onto the marine macro-algae

Ascophyllum nodosum. The effect of the operating variables on metal uptake

capacity was studied in a batch system and a mathematical model showing the

influence of each variable and their interactions was obtained. Study ranges were

10-40 ºC for temperature, 3.0-5.0 for pH and 50-150 mg/L for initial Cu (II)

concentration. Within these ranges, the biosorption capacity is slightly dependent

on temperature but markedly increases with pH and initial concentration of Cu

(II). The uptake capacities predicted by the model were in good agreement with

the experimental values. Maximum biosorption capacity of Cu (II) by A.

nodosum is 70 mg/g and corresponds to the following values of those variables:

temperature = 40 ºC, pH = 5.0 and initial Cu (II) concentration = 150mg/L.

Preetha and Viruthagiri [223] used Response surface methodology to study the

cumulative effect of the various parameters namely, initial copper ion

concentration, pH, temperature, biomass loading and to optimize the process

conditions for the maximum removal of copper. For obtaining the mutual

interaction between the variables and optimizing these variables, a 24 full factorial

central composite design using response surface methodology was employed. The

analysis of variance (ANOVA) of the quadratic model demonstrates that the

model was highly significant. The model was statistically tested and verified by

202
Chapter – III Literature Review

experimentation. A maximum copper removal of 98 % was obtained using the

biosorption kinetics of copper under optimum conditions.

Rajender Kumar et al [224] applied the Box-Behnken design matrix and

response surface methodology (RSM) to design the experiments to evaluate the

interactive effects of three most important operating variables: pH (2.0-6.0),

temperature (25-40 ºC) and initial concentration of metal ions (10.0-60.0 mg/L) on

biosorption of Cr (VI), Ni (II) and Zn (II) ions with immobilized bacterial strain

Bacillus brevis. The total 17 experiments were conducted in the present study

towards the construction of a quadratic model. Independent variables have

significant value 0.0001 which indicates the importance of these variables in the

biosorption process. Values of "Prob > F' less than 0.0500 indicate that model

terms are significant for the biosorption of Cr (VI), Ni (II) and Zn (II) ions. The

regression equation coefficients were calculated and the data fitted to a second-

order polynomial equation for removal of Cr (VI), Ni (II) and Zn (II) ions with

immobilized bacterial strains B. brevis.

Rajesh Singh et al [225] optimized various environmental conditions for

biosorption of Pb (II), Cd (II) and Cu (II) by investigating as a function of the

initial metal ion concentration, temperature, biosorbent loading and pH using

Trichoderma viride as adsorbent. Biosorption of ions from aqueous solution was

optimized in a batch system using response surface methodology. The values of

R2 0.9716, 0.9699 and 0.9982 for Pb (II), Cd (II) and Cu (II) ions, respectively,

indicated the validity of the model. The thermodynamic properties ΔG, ΔH, ΔE

and ΔS by the metal ions for biosorption were analyzed using the equilibrium

203
Chapter – III Literature Review

constant value obtained from experimental data at different temperatures. The

results showed that biosorption of Pb (II) ions by T. viride adsorbent are more

endothermic and spontaneous. The study was attempted to offer a better

understating of representative biosorption isotherms and thermodynamics with

special focuses on binding mechanism for biosorption using the FTIR

spectroscopy.

Sahu et al [226] presented application of RSM for optimizing the removal of Cr

(VI) ions from aqua solutions using activated carbon as adsorbent. All

experiments were performed according to statistical designs in order to develop

the predictive regression models used for optimization. The optimization of

adsorption of chromium on activated carbon was carried out to ensure high

adsorption efficiency at low adsorbent dose and high initial concentration of Cr

(VI). In the adsorption experiments a laboratory developed Tamarind wood

activated carbon made of chemical activation (zinc chloride) was used. A 24 full

factorial central composite design experimental design was employed. Analysis

of variance (ANOVA) showed a high coefficient of determination value (R2 =

0.928) and satisfactory prediction second-order regression model was derived.

Maximum chromium removal efficiency was predicted and experimentally

validated. The optimum adsorbent dose, temperature, initial concentration of Cr

(VI) and initial pH of the Cr (VI) solution were found to be 4.3 g/L, 32 ºC, 20.15

mg/L and 5.41 respectively. Under optimal value of process parameters, high

removal (> 89 %) was obtained for Cr (VI).

204
Chapter – III Literature Review

Umesh et al [227] investigated the effect of adsorbent dose, pH and agitation

speed on nickel removal from aqueous medium using an agricultural waste

biomass, Sugarcane bagasse. Batch mode experiments were carried out to assess

the adsorption equilibrium. The influence of three parameters on the removal of

nickel was also examined using a response surface methodological approach. The

central composite face-centered experimental design in response surface

methodology (RSM) by Design Expert Version 6.0.10 (Stat Ease, USA) was used

for designing the experiments as well as for full response surface estimation. The

optimum conditions for maximum removal of nickel from an aqueous solution of

50 mg/L were as follows: adsorbent dose (1500 mg/L), pH (7.52) and stirring

speed (150 rpm). This was evidenced by the higher value of coefficient of

determination (R2 = 0.9873).

Umesh et al [228] investigated the effect of adsorbent dose, pH and agitation

speed on the removal of chromium from aqueous medium using an agricultural

waste biomass (sugarcane bagasse treated with succinic acid). Batch mode

experiments were carried out to assess the adsorption equilibrium. The influence

of three parameters on the removal of chromium was examined using a response

surface methodological approach. The Central Composite Face-Centered

Experimental Design in Response (CCFD) Surface Methodology (RSM) by

Design Expert Version 6.0.10 (Stat Ease, USA) was used for designing the

experiments as well as for full response surface estimation and 20 trials as per the

model were run. The Chromium optimum conditions for maximum removal of

chromium from an aqueous solution of 50 mg/L were as it follows: adsorbent dose

205
Chapter – III Literature Review

(20 g/L), pH (2.0) and stirring speed (250 rpm). This was evidenced by the higher

Adsorption value of coefficient of determination (R2 = 0.9873).

Veglio et al [229] presented a study on the biosorption of toxic metals (copper,

manganese, nickel and lead) by a species of Arthrobacter. The equilibrium of the

process was in all cases well described by the Langmuir isotherm. The highest

values of specific uptake were 406 mg Mn/g, 148 mg Cu/g, 13 mg Ni/g and 130

mg Pb/g. Some factorial experiments were performed in order to investigate the

effect of operating and cultural conditions on the equilibrium. Experimental

results showed the influence of the biomass concentration on the specific uptake

(mg of metal/g of biomass as dry weight). A comparison with results reported in

the literature was made.

Yasar and Nuran [230] studied the effects of process parameters pH, adsorbent

dose and initial Cr (VI) concentration in the ranges 1.5-9.5, 1.8-24.2 g/L and 15-

95 mg/L, respectively. A predictive quadratic model was constructed by variance

analysis of data obtained from a total of 20 experimental runs with three replicates

each. Maximum removal was attained from a solution as concentrated as 30 ppm

at pH 3 with an adsorbent dosage of 13 g/L. The adsorption capacity of chitosan

flakes was determined as 22.09 mg/g at these specified conditions. However, the

adsorption capacity was recorded as high as 102 mg/g for 100 mg/L initial

concentration. Out of Langmuir, Freundlich and Dubinin-Radushkevich isotherm

models, adsorption data was best described by Langmuir isotherm with 0.99

consistencies. The process kinetics was evaluated by pseudo-first, pseudo-second

order and intra-particle diffusion models. Pseudo-second order kinetic model

206
Chapter – III Literature Review

exhibited the highest correlation with data. The results showed that both

monolayer adsorption and intra-particle diffusion mechanisms limited the rate of

Cr (VI) adsorption. Thermodynamic parameters revealed the feasibility,

spontaneity and exothermic nature of adsorption.

Yi- Ling Lai et al [231] tested the performance of a new biosorbent system,

consisting of a fungal biomass immobilized within an orange peel cellulose

absorbent matrix, for the removal of Zn2+ heavy metal ions from an aqueous

solution. The amount of Zn (II) ion sorption by the beads was as follows; orange

peel cellulose with Phanerochaete chrysosporium immobilized Ca-alginate beads

(OPCFCA) (168.61 mg/g) > orange peel cellulose immobilized Ca-alginate beads

(OPCCA) (147.06 mg/g) > P. chrysosporium (F) (125.0 mg/g) > orange peel

cellulose (OPC) (108.70 mg/g) > plain Ca-alginate bead (PCA) (98.26 mg/g). The

Zn2+ concentration was 100 to 1000 mg/L. The widely used Langmuir and

Freundlich isotherm models were utilized to describe the biosorption equilibrium

process. The isotherm parameters were estimated using linear and non-linear

regression analysis. The Box-Behnken model was found to be in close agreement

with the experimental values, as indicated by the correlation coefficient (R2) value

of 0.9999.

Yus Azila et al [232] employed central composite design (CCD) to optimize the

biosorption of Pb (II) ions onto immobilized cells of Pycnoporus sanguineus. The

independent variables were initial Pb (II) concentration, pH and biomass loading.

The combined effects of these variables were analyzed by response surface

methodology (RSM) using quadratic model for predicting the optimum point.

207
Chapter – III Literature Review

Under these conditions the model predicted a maximum of 97.7 % of Pb (II) ions

removal at pH 4, 200 mg/L of initial Pb (II) concentration with 10 g/L of

biosorbent. The experimental values are in good agreement with predicted values

within +0.10 to +0.81 % error.

Zulkali et al [233] investigated the effects of initial concentration of lead,

temperature, biomass loading and pH for an optimized condition of lead uptake

from the aqueous solution. The optimization process was analyzed using Central

Composite Face-Centered Experimental Design in Response Surface

Methodology (RSM) by Design Expert Version 5.0.7 (StatEase, USA). The

design was employed to derive a statistical model for the effect of parameters

studied on the removal of lead ion from aqueous solution. The coefficient of

determination, R2 was found to be 92.36 %. The initial concentration of 50.0

mg/L, temperature of 60 oC, biomass loading of 0.2 g and pH of 5.0 had been

found to be the optimum conditions for the maximum uptake of lead ions in 98.11

% batch mode. Under the optimum conditions, the lead uptake was attained to be

circa 8.60 mg/g.

It is evident from the above literature review that no research has been carried out

on biosorption of chrmomium and lead using Ageratum Conyzoides leaf powder

and Anacardium occidentale testa powder in batch study. Hence the author has

intended to follow up the present work.

208

You might also like