You are on page 1of 69

Accepted Manuscript

Title: Adsorption removal of Zinc (II) from Aqueous Phase by


Raw and Base Modified Eucalyptus Sheathiana Bark:
kinetics, mechanism and equilibrium study

Author: Sharmeen Afroze Tushar Kanti Sen Ha Ming Ang

PII: S0957-5820(16)30026-X
DOI: http://dx.doi.org/doi:10.1016/j.psep.2016.04.009
Reference: PSEP 741

To appear in: Process Safety and Environment Protection

Received date: 15-12-2015


Revised date: 4-4-2016
Accepted date: 7-4-2016

Please cite this article as: Afroze, S., Sen, T.K., Ang, H.M.,Adsorption removal of
Zinc (II) from Aqueous Phase by Raw and Base Modified Eucalyptus Sheathiana Bark:
kinetics, mechanism and equilibrium study, Process Safety and Environment Protection
(2016), http://dx.doi.org/10.1016/j.psep.2016.04.009

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Research Highlights

Highlights
 Raw and chemically treated Eucalyptus sheathiana bark were used in Zn2+ removal
 The factors that affect the adsorption process were investigated in detail
 Both Freundlich and Langmuir isotherm models are applicable
 The Eucalyptus bark show great potential to treat metal ion bearing wastewaters

t
ip
cr
us
an
M
ed
pt
ce
Ac

Page 1 of 68
*Manuscript
Click here to view linked References

1 Adsorption removal of Zinc (II) from Aqueous Phase by Raw and Base
2 Modified Eucalyptus Sheathiana Bark: kinetics, mechanism and
3 equilibrium study
4
5 Sharmeen Afrozea, Tushar Kanti Sen*a, Ha Ming Anga
6
a
7 Department of Chemical Engineering, Curtin University,

8 GPO Box U1987, Perth, Western Australia 6845, Australia.

t
9

ip
10

cr
11

us
12

13

14
an
15
M
16

17
ed

18

19
pt

20
ce

21

22
Ac

23

24

25

26

27

*
28 Corresponding author email address: T.Sen@curtin.edu.au

Page 2 of 68
2

1 ABSTRACT

2 In this study, potential application of abundantly available agricultural by-product Eucalyptus

3 sheathiana bark in its raw and sodium hydroxide (NaOH) modified form to remove Zn2+

4 from its aqueous solutions was investigated by considering parameter identification and

5 optimization, reusability, equilibrium, kinetic and thermodynamic studies. The adsorbent was

t
6 characterised by SEM-EDX, FTIR, XRD, BET surface area and bulk density and point of

ip
7 zero charge were also determined. The process was strongly pH dependent and the adsorption

cr
8 percentage of Zn2+ was increased with an increase in solution pH from 2.5 to 5.1. Conversely,

us
9 the adsorption percentage of Zn2+ decreased with the increase in adsorbent dosage, initial

10 metal concentration, temperature and ionic strength. Kinetic measurements showed that the

11
an
process was multistep, rapid and diffusion controlled. It was found to follow the pseudo-

12 second-order rate equation. Equilibrium adsorption studies showed that both Freundlich and
M
13 Langmuir models are applicable for both raw and base modified eucalyptus bark. MPSD

14 error function was used to treat the equilibrium data using non-linear optimization technique
ed

15 for evaluating the fit of the isotherm equations. The maximum sorption capacity of modified

16 eucalyptus bark was 250.00 mg g−1 at 300C which was comparative to other adsorbents.
pt

17 Various thermodynamic parameters indicate that the process was spontaneous and physical in
ce

18 nature. Desorption studies were also performed to determine possible recovery potential of

19 Zn2+ and the re-usability of the biomass and to identify the mechanism of adsorption.
Ac

20 Keywords: Eucalyptus Bark; Chemically treated biomass; Zinc adsorption; Kinetics model;

21 Isotherm.

22

Page 3 of 68
3

1 1. Introduction

2 The main sources of zinc enters to the environment are from the wastewater discharging

3 through many industrial sectors such as the manufacturing of brass and bronze alloys,

4 purifying zinc, lead and cadmium ores, steel production and coal burning and burning of

5 wastes (Bhattacharya, Mandal, & Das, 2006). Zinc is also present in high concentration in

t
ip
6 wastewater of pharmaceuticals, galvanizing, paints, pigments, insecticides, cosmetics, etc.

7 that causes serious problem to the environment (Bhattacharya et al., 2006). Zinc is an

cr
8 essential element for life and acts as a micronutrient when present in trace amounts (Sen &

us
9 Gomez, 2011). The maximum contamination levels (MCLs) of heavy metals for surface or

10 groundwater to be used in the drinking supply set up by The United States Environmental

11
an
Protection Agency (USEPA) where the maximum acceptable concentration of zinc is 0.8

12 mg/L (Nguyen et al., 2013). Beyond the permissible limits, Zn2+ is toxic and the symptoms of
M
13 zinc toxicity include irritability, muscular stiffness, loss of appetite and nausea (Safe

Drinking Water Committee & Council, 1977). Various advanced chemical or physical
ed

14

15 treatment processes such as electro coagulation, ion exchange, irradiation, ozonation,


pt

16 membrane separation, advanced oxidation, adsorption on activated carbon are applied for the

17 removal of heavy metal ions including Zn2+ ions from aqueous solutions. Most of the
ce

18 treatment processes have their own advantages and disadvantages. Among all these

conventional technologies, adsorption has been recognized as one of the versatile effective
Ac

19

20 technique practiced widely for heavy metal infested wastewater treatment because of its

21 operational simplicity, sludge free operation and reuse potential of adsorbents during long-

22 term applications (Acharya, Sahu, Sahoo, Mohanty, & Meikap, 2009; Bharathi & Ramesh,

23 2013). In recent years, a number of non-conventional cost effective biomass based adsorbents

24 alternative to costly activated carbon such as pine chips (McLaughlan, Hossain, & Al-

25 Mashaqbeh, 2015), watermelon rind (Liu, Ngo, & Guo, 2012), cedrus deodara sawdust

Page 4 of 68
4

1 (Mishra, Balomajumder, & Agarwal, 2012), cashew nut shell (Senthil Kumar et al., 2012),

2 potato peels (Taha, Arifien, & El-Nahas, 2011), orange waste (Pérez Marín et al., 2010),

3 eucalyptus leaf biomass (Mishra, Balomajumder, & Agarwal, 2010), citrus reticulata

4 (Kinnow) waste (Boota, Bhatti, & Hanif, 2009), carrot residue (Nasernejad, Zadeh, Pour,

5 Bygi, & Zamani, 2005), almond shell, olive and peach stones (Ferro-Garcia, Rivera-Utrilla,

t
6 Rodriguez-Gordillo, & Bautista-Toledo, 1988) have been used for Zn2+ adsorption from its

ip
7 aqueous solution. Readers are encouraged to go through review article by Yagub et al

cr
8 (Yagub, Sen, Afroze, & Ang, 2014) and Nguyen et al (Nguyen et al., 2013) in this research

us
9 area. The application of agricultural solid wastes as effective adsorbents offer several

10 advantages; their easy availability in large quantities, requirement of less processing time,

11
an
renewable in nature, low cost, eco-friendly and good adsorption potential due to their unique

12 chemical composition makes the adsorption process more attractive for heavy metal and dye
M
13 remediation (Choi, Chung, Hong, Kim, & Lee, 2012; B. Singha & Das, 2013; Sud, Mahajan,

14 & Kaur, 2008). In most of the situations direct use of these raw biomass adsorbents is
ed

15 restricted due to their low adsorption capacity. Therefore pre-treatment of adsorbent is also

16 commonly used to increase the adsorption capacity of these materials (Bhattacharya et al.,
pt

17 2006). Pretreatment with dilute sodium hydroxide solution has been the most popular method
ce

18 of improving surface properties and removing soluble organic components of plant wastes

19 applied for adsorption (A. Ofomaja, Naidoo, & Modise, 2009). Dilute sodium hydroxide
Ac

20 solution treatment onto lingo cellulosic agricultural waste adsorbents is known to solubilize a

21 small fraction of the lignin and soluble organics in the plant waste (Wartelle & Marshall,

22 2000); improving the penetration of modifying agents into the biosorbent matrix and thereby

23 increasing porosity and changing surface properties and hence metal sorption capacity (A.

24 Ofomaja et al., 2009). Several investigators reported the effectiveness of chemically modified

25 agricultural biomass in the removal of metals and dyes and authors are encouraged to through

Page 5 of 68
5

1 few recently published review articles such as Afroze et al (S. Afroze, T. Sen, & M. Ang,

2 2015), Gautam et al (Gautam, Mudhoo, Lofrano, & Chattopadhyaya, 2014), and Hubbe et al

3 (Hubbe, Hasan, & Ducoste, 2011).

4 Eucalyptus trees are evergreen, fast growing and abundantly available worldwide. Due to the

5 high number of eucalyptus trees, massive amounts of barks are disposed each year in

t
6 Australia as waste material. Cost consideration of an adsorbent is an important parameter for

ip
7 selection and design of an adsorption process. In general, a sorbent can be assumed to be

cr
8 “low cost” if it requires little processing and is abundant in nature, or waste material either

us
9 from industry or agricultural by-product with almost no economic value. Eucalyptus bark is

10 freely available, non-hazardous in nature and with throughway price which fulfils the

11
an
precursor to become a low cost adsorbent alternative to costly commercial adsorbents. Hence

12 the utilization of this agricultural solid waste as adsorbent is of great significance in liquid-
M
13 phase contaminant separation and also in efficient solid waste management. Previously

14 eucalyptus barks were used in the adsorptive removal of Cu(II), Cr(III), Cd(II), and Ni(II)
ed

15 (Saliba, Gauthier, Gauthier, & Petit-Ramel, 2002), Hg(II) (Ghodbane & Hamdaoui, 2008)

16 from aqueous phase. It was observed that eucalyptus bark is an efficient sorbent for the
pt

17 removal of these metal ions. Therefore aim of this present research work were to (a) evaluate
ce

18 the effectiveness of raw and chemically NaOH treated eucalyptus bark biomass in the

19 removal of Zn2+ metal ions from its aqueous solution; (b) to study the equilibrium, kinetics,
Ac

20 mechanism and thermodynamics of Zn2+ adsorbate onto raw and chemically treated

21 adsorbent under various physico-chemical process conditions; (c) identify the main surface

22 and morphological and functional groups characteristics of treated adsorbents and their

23 mechanistic contribution towards capacity of adsorbents; (d) assess the reusability and

24 applicability of treated eucalyptus bark in the Zn2+ adsorption and (e) comparison of treated

Page 6 of 68
6

1 eucalyptus bark adsorbents capacity with other published adsorbents including commercial

2 activated carbon.

3 2. Materials and Methods


4 2.1 Materials

t
5 2.1.1 Raw Adsorbent

ip
6

7 Eucalyptus barks were collected from Eucalyptus sheathiana trees at the campus of Curtin

cr
8 University, Perth, Western Australia between February and March 2013. The barks were

us
9 washed repeatedly with distilled water to remove impurities such as sand and leaves, and then

10 dried at 105˚C for 24 h in an oven. The dried biomass was ground using a mechanical grinder

11
an
and the resultant powder was sieved and the particles below 106 µm were collected, stored in

12 an airtight plastic container and used for conducting adsorption experiments.


M
13 2.1.2 Chemical Modification of Eucalyptus Bark Adsorbent and Characterisation
ed

14

15 A weighed amount (50 g) of raw eucalyptus bark powder was contacted with 500 mL of 0.1
pt

16 M sodium hydroxide (NaOH) solution to prepare base modified eucalyptus bark powder. The

17 whole slurry was stirred overnight with a magnetic stirrer and then the powder was filtered
ce

18 and rinsed with distilled water. This procedure was repeated twice to ensure removal of

excess NaOH from the powder. The residues were then dried overnight at 90˚C and used for
Ac

19

20 adsorption experiments.

21

22 The modified eucalyptus bark biomass was characterized in terms of surface area and bulk

23 density. Bulk density of NaOH modified eucalyptus bark were measured as per Eq. (1)

24 (Afroze, Sen, Ang, & Nishioka, 2015) :

Page 7 of 68
7

1 (1)

2 Scanning electron microscope (SEM) named Zeiss NEON 40EsB, USA was used to examine

3 the morphology of NaOH treated EB biomass before and after adsorption and to observe the

4 components of the adsorbent, the electron diffraction spectrum (EDS) was taken for NaOH

5 treated EB biomass before and after adsorption. To determine the crystal phase of raw EB

t
ip
6 and NaOH treated EB, the samples were analysed using powder diffractometer D8 Advance

(Bruker AXS, Germany) X-ray Diffraction (XRD) machine. Further, NaOH treated

cr
7

8 eucalyptus bark (EB) powder was analysed by PerkinElmer Spectrum 100 FT-IR

us
9 Spectrometer to determine the functional groups. Brunauer–Emmett–Teller (BET) method

10 was used to determine BET surface area of modified EB biomass using Tristar II 3020,

11 Micromeritics Instrument Corporation.


an
M
12 2.1.3 Adsorbate (Zn2+) and Other Chemicals
13

14 All chemicals used were of analytical grade. The stock standard solution of Zn2+ of 1000
ed

15 mg/L was prepared by dissolving an accurately weighted amount of Zn(NO3)2∙6H2O salt in

ultra-pure water. The working solutions were then prepared by diluting the stock solution
pt

16

17 with ultra-pure water to give the specific concentration of the working solutions. Similarly,
ce

18 salt solutions of 100 mg/L, 200 mg/L and 300 mg/L were prepared by dissolving the

19 appropriate amount of laboratory grade NaCl, CaCl2 and FeCl3 separately in a litre of ultra-
Ac

20 pure water to perform experiments with salt effects. The pH of the solutions was adjusted

21 using reagent grade HCl and NaOH respectively. All plastic sample bottles and glassware

22 were cleaned, then rinsed with deionised water and dried at 60°C in a temperature-controlled

23 oven. All measurements were conducted at room temperature of 30°C. The concentration of

24 Zn2+ was measured using Shimadzu AA-7000 Atomic Absorption Spectrophotometer (AAS).

25 The pH was measured by WP-81 pH-Cond-Salinity meter.

Page 8 of 68
8

1 2.2 Methods
2

3 2.2.1 Adsorption Kinetic Experiments


4

5 Batch adsorption experiments were conducted by using a known amount of the adsorbent

6 sample with 50 mL of aqueous Zn2+ solutions in a series of 100 mL plastic bottles. The

t
7 mixture was shaken in the Thermo Line Scientific Orbital Shaker Incubator at 120 rpm at a

ip
8 constant temperature of 30◦C for 100 min to attain equilibrium time. The suspensions were

cr
9 taken out at predetermined time intervals and then filtered using micro filter of pore size 0.47

us
10 µm and the filtrates were analysed using atomic absorption spectrophotometer with air–

11 acetylene flame. Several experiments were carried out by varying solution pH, initial Zn2+

12 an
solution concentration, mass of adsorbent, temperature, and with the addition of salts. The

13 amount of metal adsorbed onto raw and NaOH treated eucalyptus bark powder at time t, qt
M
14 (mg/g) and metal removal efficiency, i.e. % adsorption are calculated using Eqs. (2) and (3)

15 respectively:
ed

16 qt = (2)
pt

17 % adsorption (3)

where C0 is the initial Zn2+ concentration (mg/L), Ct is the concentration of Zn2+ at any time t
ce

18

19 (min), V is the volume of Zn2+ solution (L) and m is the mass of eucalyptus bark powder
Ac

20 (grams). Each experiment was repeated in twice to check reproducibility. Measurements are,

21 in general, reproducible within ±10 % accuracy.

22 2.2.2 Isotherm Experiments


23 Equilibrium isotherm adsorption studies were conducted by contacting 50 mL of Zn2+

24 solutions of different initial concentration of 20, 30, 40, 50, 60 and 70 ppm with 20 mg of raw

25 and NaOH treated eucalyptus bark powder separately in a series of 100 mL plastic bottles at a

26 fixed pH 5.1 for a period of 2 h. From the adsorption kinetic experiments, it was found that

Page 9 of 68
9

1 almost steady state adsorption reaction was reached within 100 min reaction time and

2 therefore reaction time for equilibrium was assumed as 100 min for convenience of this

3 isotherm study and therefore a period of 2 hr was sufficient for equilibrium study.

4 The bottles were then shaken at a constant speed of 120 rpm in the Thermo Line Scientific

5 Orbital Shaker Incubator at 30◦C and the optimum pH of all the samples were to be fixed at

t
ip
6 5.1. After 2 hr shaking, the samples were separated by filtration and the filtrate was analysed

7 for the remaining Zn2+ concentration by Shimadzu AA-7000 atomic absorption

cr
8 spectrophotometer (AAS).

us
9 2.2.3 Desorption Experiments

10 Desorption studies were carried out to assess recovery of Zn2+ and the nature of interaction of
an
11 Zn2+ with biomass adsorbents respectively. The method of Marshall and Johns (W. E.
M
12 Marshall & Johns, 1996) was modified and adopted. Zinc-loaded adsorbent biomass was

13 separated and gently washed with distilled deionized water to remove any unadsorbed Zn 2+.
ed

14 The biomass was then agitated in 50 mL of 0.1 M HCl for 100 min and the amount of

15 desorbed Zn2+ estimated.


pt

16
ce

17 2.3 Theory
Ac

18 2.3.1 Adsorption Kinetics and Mechanism


19

20 For process design and to control the adsorption phenomenon, it is significant to understand

21 the adsorption kinetics and mechanism of a system. The sorption kinetics depends on the

22 properties of the sorbent, sorbate and experimental conditions. To analyse kinetics of the

23 sorption process, three different reaction kinetic models (pseudo-first order equation, pseudo-

Page 10 of 68
10

1 second order equation and intra-particle diffusion model) were applied to determine the

2 reaction order and rate constant which are given below:

3 Pseudo-First-Order Kinetic Model The linearized integral equation for solid-liquid sorption

4 system of the Pseudo-first-order model developed by Lagergren (Lagergren, 1898) is written

5 as:

t
ip
K1
6 Log (qe  qt )  Log (qe )  t (4)
2.303

cr
7 where qe represents the amount of Zn2+ adsorbed (mg/g) at equilibrium which is a constant

us
8 steady value at equilibrium for the system and qt represents the amount of Zn2+ adsorbed

9 (mg/g) at any time which varies with time, t (min). K1 is the pseudo-first-order adsorption

10
an
rate constant (min-1). A plot of Log (qe – qt) against t should give a straight line if the pseudo-

11 first-order adsorption kinetics is applicable and the adsorption rate constant K1 and qe can be
M
12 calculated from the slope and intercept of the linear plot of Log (qe – qt) against (t)

13 respectively.
ed

14
15 Pseudo-Second-Order Kinetic Model The linearized form of the pseudo-second-order kinetic
pt

16 model (Sen & Sarzali, 2008) is shown in Eq. (5):


ce

t 1 1
17  2
 t (5)
qt K 2 qe qe

18 where K2 is the pseudo-second-order adsorption rate constant [g/(mg min)]. A plot between
Ac

19 t/qt against t should give a straight line and yield the values of rate constant K2 (g/mg min),

20 and the equilibrium amount of adsorption, qe (mg/g) from the slope and intercept of plot t/qt

21 against t respectively.

22
23 The constant K2 is used to calculate the initial sorption rate h, at t→0, as follows (Afroze,

24 Sen, Ang, et al., 2015):

Page 11 of 68
11

1 (6)

2 Intra Particles Diffusion Model To identify the adsorption mechanism, intraparticle mass

3 transfer diffusion model is important to consider for any adsorption system. According to

4 Weber and Morris (Weber & Morris, 1963), for most adsorption processes the adsorption

5 varies almost proportionately with t1/2 rather than with the contact time (t) as shown in Eq.

t
(7):

ip
6

7 qt  K id t 0.5 + C (7)

cr
8 where Kid is the intra-particle diffusion rate constant (mg/(g.min0.5)) and C is the intercept.

us
9 The rate constant of intraparticle transport, Kid can be estimated from the slope of the linear

10 portion of the plot of amount sorbed, qt (mg/g) against square root of time, t0.5. When intra-

11
an
particle diffusion plays a significant role in controlling the kinetics of the adsorption

12 processes, the plots of qt versus t0.5 yield straight lines passing through the origin and the
M
13 slope gives the rate constant Kid. When pore diffusion limits the adsorption process, the

relationship between the initial solute concentration and the rate of adsorption will not be
ed

14

15 linear (Knocke & Hemphill, 1981).


pt

16 2.3.2 Adsorption Equilibrium Isotherm and Error Analysis


17
ce

18 In order to optimize the design of a sorption system and to find out maximum adsorption

capacity of adsorbents, two well-known and widely applied used models, Freundlich
Ac

19

20 Isotherm Model and Langmuir Isotherm Model, were selected to explain the interaction

21 between metal ions (Zn2+) with raw eucalyptus and modified eucalyptus bark adsorbent.

22 Freundlich Isotherm: The linearized Freundlich isotherm (Freundlich, 1906), is based on

23 adsorption to be taken place on heterogeneous surfaces, can be represented as:

24 lnqe = ln Kf + bf lnCe (8)

Page 12 of 68
12

1 where qe (mg/g) is the equilibrium amount of Zn+2 adsorbed by the adsorbent, Ce (mg/L) is

2 the equilibrium concentration of adsorbate in solution, Kf and bf are the Frendlich isotherm

3 constants. The slope and intercept of plot between lnqe versus lnCe using Eq. (8) gives the

4 values of Kf and bf respectively .

5 Langmuir Isotherm: The Langmuir Isotherm model (Langmuir, 1918) deals with adsorption

t
ip
6 on homogeneous surface with a finite number of identical sites. The most common linearized

7 forms of Langmuir I and Langmuir II can be written as per Eqs. (9) and (10) respectively.

cr
8

us
Ce 1 C
  9e (9)
qe K a q m q m

1  1
10

 1 111
an (10)
  
qe  K a qm  Ce q m
M
12

13 where qm is the maximum adsorption capacity (mg/g) and Ka is the Langmuir equilibrium
ed

14 constant related to the energy of adsorption (L/mg) which can be calculated from the slope

15 and intercept of the linearized forms of the plots where (Ce/qe versus Ce) for Langmuir I and
pt

16 (1/qe versus 1/Ce) for Langmuir II respectively.


ce

17 The essential characteristics of the Langmuir isotherm can be expressed in terms of a

18 separation factor (RL) which is dimensionless and is given by Eq. (10) (Knocke & Hemphill,
Ac

19 1981):

20 RL = 1/(1+ KaCo) (11)

21 where Ka is the Langmuir constant and Co is the initial concentration of zinc (mg/L). The

22 parameter RL indicates the shape of isotherm as follows:

23 RL > 1 unfavorable; RL = 1 linear; 0 < RL < 1 favorable; RL = 0 irreversible (Ghodbane, Nouri,

24 Hamdaoui, & Chiha, 2008).

Page 13 of 68
13

1 Error Analysis:

2 The use of R2 is limited to solve linear forms of isotherm equation, which measures the

3 difference between experimental and theoretical data in linearized plots only, but not the

4 errors in non-linear form of isotherm curves. For that reason, the Marquardt’s percent

5 standard deviation (MPSD) error function was used to test the adequacy of the isotherm

t
ip
6 equations to represent the experimental data. The error function of Marquardt’s Percent

7 Standard Deviation (MPSD) (Marquardt, 1963) is given as:

cr
us
8 (12)

9
an
The MPSD account for the number of degrees of freedom of the system where n is the

10 number of data point and p is the number of parameters of the isotherm and is similar in some
M
11 respects to a geometric mean error distribution.
ed

12 2.4 Thermodynamic Study


13

The thermodynamic parameters, such as change in Gibb’s free energy of activation (∆Go),
pt

14

15 change in enthalpy of activation (∆Ho) and change in entropy of activation (∆So) for the
ce

16 adsorption of Zn+2 ions on raw and NaOH treated eucalyptus bark adsorbent have been

17 determined by using the following equations (Sen & Gomez, 2011):


Ac

18 ∆Go = ∆Ho-T∆So (13)

q  S o  H19o (14)
log  e   
 Ce  2.303R 2.303RT
20

21 where qe is the solid-phase concentration at equilibrium (mg/L), Ce is equilibrium

22 concentration in solution (mg/L), T is temperature in K and R is the gas constant (8.314 J/mol

Page 14 of 68
14

1 K). The entropy change (∆So) and enthalpy change (∆Ho) can be calculated from the intercept

2 and slope of the linear Van’t Hoff plot log (qe/Ce) versus 1/T. From those values, change in

3 Gibb’s free energy of activation (∆Go) can be calculated using the Eq. (13).

4 3. Results and Discussion


5

t
ip
6 3.1 Surface and Porosity Characteristics
7

cr
8 Surface and porosity characteristics of untreated raw and NaOH treated eucalyptus bark is

us
9 presented in Table 1.

10 Table 1 – Surface characteristic parameters of raw and NaOH treated eucalyptus bark (EB)
11
12
adsorbent an
13
M
14 Table 1 clearly shows that NaOH treated bark adsorbent has better surface properties in terms

15 of surface area and porosity compared to raw eucalyptus bark materials which gives good
ed

16 indication to become more effective adsorbent.

17 The information of bulk density is significant for the designing of adsorption tower in real
pt

18 industrial situation. It affects the overall performance of the adsorption process and it is
ce

19 inversely proportional to the particle size (Dawood, Sen, & Phan, 2014). As per American

20 Water Work association, bulk density should be greater than 0.25 g/cm3 for practical
Ac

21 purposes. Plant components such as sugars, tannins and pigments occupy void spaces and

22 contribute largely to their bulk densities and therefore extraction of these plant components

23 with dilute base solution reduces their bulk densities (Wartelle & Marshall, 2000). Hence,

24 NaOH treated eucalyptus bark showed a reduction in bulk density over the raw eucalyptus

25 bark (Table 1). Reduction in bulk density of some agricultural solid wastes after washing

Page 15 of 68
15

1 with 0.1 M of NaOH has also been observed for pine cone (A. Ofomaja & Naidoo, 2011) and

2 sugarcane bagasse (Wartelle & Marshall, 2000).

3 BET surface area: Table 1 compares the values of BET surface area (m2/g) for the raw and

4 modified eucalyptus bark as well as the total pore volume measurements for the individual

5 adsorbent samples obtained by BET adsorption model. The results in Table 1 reveal that BET

t
ip
6 surface area of raw eucalyptus bark (6.55 m2/g) is quite less than the surface area obtained for

7 NaOH treated eucalyptus bark biomass (20.13 m2/g) indicating the treatment of eucalyptus

cr
8 bark with NaOH produced an increase in surface area. Increase in surface area with base

us
9 treatment has been attributed to extraction of plant components such as sugars, cementing

10 materials and lignin which usually block pores on adsorbent materials (W. Marshall,

11
an
Wartelle, Boler, Johns, & Toles, 1999; Wartelle & Marshall, 2000) leading to further

12 exposure of the adsorbent surface of plant based materials. Studies have shown that some
M
13 plant based materials applied as adsorbent are porous in nature (Elizalde-González, Mattusch,

Peláez-Cid, & Wennrich, 2007; Zvinowanda, Okonkwo, Agyei, & Shabalala, 2009). Table 1
ed

14

15 shows the values of total pore volume for raw and NaOH treated EB. As explained earlier,
pt

16 the increase in total pore volume can be related to extraction of plant components which may

17 have blocked existing pores and also the creation of new pores due to decrease in
ce

18 crystallinity, separation of structural linkages between lignin and carbohydrates and

disruption of the lignin structure (A. Ofomaja et al., 2009). This observation suggests that
Ac

19

20 base treatment leads to creation of new pores and that mesopore formation predominates with

21 the base treatment in the order NaOH treated EB > raw EB.

Page 16 of 68
16

3 3.1.1 Functional groups identification through Fourier Transform Infrared


4 (FTIR)
5

6 The presence of functional groups on the surface of adsorbents was confirmed by using FTIR

t
analysis. The FTIR spectra patterns for raw eucalyptus bark and NaOH modified eucalyptus

ip
7

8 bark powder are shown in Fig. 1. The complex nature of the FTIR spectrum shown in Fig. 1

cr
9 suggests that eucalyptus bark is composed of mixture of functional groups. Based on Fig.

us
10 1(a), –OH functional group can be easily recognized by the bands at 3314.9 cm −1 and the

11 peak representing C═O stretch appears at 1617.4 cm−1 while the nitro group N═O is

12
an
indicated by 1317.8 cm−1 (University of Colorado, 1985). The peaks between 1033.8 and

13 784.5 cm−1 may be assigned to the C–O and C–Cl stretching respectively (University of
M
14 Colorado, 1985).
ed

15 The spectra’s of NaOH treated eucalyptus bark sample in Fig. 1(b) showed the formation of

16 new adsorption bands with some slight changes in the intensity of the peak at 1617.4 cm -1,
pt

17 1317.8 cm-1, 1033.8 cm-1 and 784.5 cm-1. All these peaks tend to decrease slightly after the

treatment with NaOH. The appearance of new shoulder peaks at 2922.5 cm-1 represents the
ce

18

19 aliphatic group C–H while two sharp peaks at 1423.1 cm-1 and 1369.9 cm-1 are indicating
Ac

20 nitro group N═O and the observed peak at 1157.2 cm-1 is attributed for C–O stretching

21 (University of Colorado, 1985).

22 From FTIR study, the formation of new absorption bands, the change in absorption intensity

23 and the shift in wavenumber of functional groups could be due to the changed pore structure

24 and increased surface area of eucalyptus bark activated by NaOH treatment.

25 Fig. 1 – FTIR spectrum of (a) raw and (b) NaOH treated eucalyptus bark
26 powder

Page 17 of 68
17

1 3.1.2 SEM Analysis


2

3 The surface morphology of the adsorbents before and after metal ion adsorption was

4 observed using SEM analysis. SEM (BSE) images clearly reveal the surface texture and

5 morphology of the adsorbents before and after adsorption (Fig. 2). There are significant

6 changes to the surface morphology of the adsorbents, as well as the formation of discrete

t
ip
7 aggregates on their surfaces following metal ion adsorption. The surface morphology of

cr
8 NaOH treated EB is different from that of raw EB. After treatment with sodium hydroxide

9 (NaOH), eucalyptus bark (EB) has more irregular and porous structure than that of raw EB,

us
10 and therefore more specific surface area. This surface characteristic would result in the higher

11 adsorption capacity which was later confirmed by BET analysis.


an
12 As presented in Fig. 2, raw EB displayed a dense and porous surface texture. Interaction of
M
13 raw EB with Zn2+ has changed the surface texture of raw EB to fractured and fragmented

14 walls of the pores on its surface [Fig. 2(b)]. Similar SEM replication with more porous
ed

15 structure has been observed for modified EB biomass before adsorption. The secondary SEM

16 image of Fig. 2(c) indicates higher porous and fibrous texture of the modified EB adsorbent
pt

17 with high heterogeneity compared with the raw eucalyptus bark [Fig 2(a)] which may
ce

18 contribute to the higher adsorption of Zn2+ ions. The greater the number of pores, the greater

19 will be the adsorption of metal Zn2+ ions onto the adsorbent surface. On the other hand, the
Ac

20 porous texture of modified EB disappeared and lump-like deposits were formed after contact

21 with Zn2+ [Fig. 2(d)].

22 Fig. 2 – SEM (BSE) images of (a) raw EB before adsorption (b) raw EB after
2+
23 Zn adsorption (c) NaOH treated EB before adsorption (d) NaOH treated EB after
24 Zn2+ adsorption
25

Page 18 of 68
18

2 3.1.3 X-ray Diffraction (XRD)


3

4 Chemical analysis of raw and NaOH treated eucalyptus bark was determined by using XRD

5 technique which is presented in Fig. 3. The phases present were analysed using powder X-ray

6 diffraction in the 2θ region of 10-900 at 250C by using Cu Kα radiation (λ=1.54 A0). Major

t
ip
7 peaks obtained during XRD were identified using EVA software. Fig. 3 illustrates the

cr
8 presence of a significant amount of amorphous material due to lignin and tannin in the

9 samples and structure of eucalyptus bark remained unaltered even after treatment with

us
10 NaOH. The results in Fig. 3 showed the XRD pattern of raw and NaOH treated eucalyptus

11 bark where the characteristic main peaks obtained at 2θ values of 14.90, 24.360, 30.070,
an
12 38.260, 43.40, 52.50and 57.550 of the two samples were exactly the same. These peaks are
M
13 indicative of highly organized crystalline cellulose.

14 XRD analysis in Fig.3 showed that there was no obvious difference in the crystal structure of
ed

15 the raw and base treated eucalyptus bark. In general, there should not be any significant

16 difference to be observed for raw and modified samples for the same material. Powder x-ray
pt

17 diffraction (XRD) is used to measure the crystalline content of adsorbent materials and to
ce

18 identify the crystalline phases present within the material containing important structural

19 information about the crystal. The XRD for both the samples was implemented to observe the
Ac

20 crystalline patterns which are responsible for adsorption and to reconfirm the identical

21 crystalline cellulose of both the samples.

22 Fig. 3 – X-ray diffraction spectra’s of (a) raw and (b) NaOH treated eucalyptus bark
23 (EB) powder
24

25

Page 19 of 68
19

1 3.2 Adsorption Kinetic Experiments


2

3 3.2.1 Effect of Initial Solution pH on Zn2+ Adsorption


4 The magnitude of electrostatic charges imparted by the ionization of the adsorptive metal

5 ions and the surface properties of adsorbent are primarily controlled by pH of the medium (A.

6 Ofomaja et al., 2009). Therefore, pH of the aqueous solution plays a significant role on the

t
efficiency of adsorption. The effect of initial solution pH on Zn2+ adsorption from aqueous

ip
7

8 solution by raw eucalyptus bark (EB) and base (NaOH) treated eucalyptus bark was

cr
9 investigated over an initial solution pH range of 2.5–5.6 and the amounts of Zn2+ adsorbed

us
10 with initial solution pH relationship is presented in Fig. 4. It was shown that the amount of

11 Zn2+ adsorption increased with the increased initial solution pH range 2.5–5.1 for both the

12
an
systems. The increase in Zn2+ sorption in the above mentioned pH range is due to the

13 electrostatic interaction of cationic Zn2+ metal ions with the negatively charged surface of the
M
14 eucalyptus bark.

15 Fig. 4 - Effect of initial solution pH on the adsorption of Zn2+ by EB and NaOH treated
ed

16 EB powder: (Conditions: Mass of Adsorbent = 20 mg, Volume of Zn 2+ Solution = 50 ml,


17 Initial Zn2+ Solution Concentration = 20 ppm, Temp = 30°C, Shaker Speed = 120 rpm
18 and Time of Adsorption = 100 min).
pt

19

20 According to Surface Complexation Model (SCM), cations with the hydroxyl group on the
ce

21 surface of adsorbent can create a surface complex. Adsorption of metal ions may result in the

22 formation of more than one type of surface complex which attributes the following reactions
Ac

23 for Zn2+ with an oxide surface site where S in the symbol of adsorbent hydroxyl surface

24 (Davis & Leckie, 1978), i.e.,

25 ≡ SOH + Zn2+ ↔ SO−−Zn2+ + H+ (15)

26 ≡ SOH + Zn2+ + H2O ↔ [SO−−ZnOH+] + 2H+ (16)

27 As pH influences the sorbent surface charge and with the increase in pH, less competition

28 from protons to reaction sites increase the concentration of Zn(OH)+ species, and therefore

Page 20 of 68
20

1 higher amount of Zn2+ adsorption was observed up to solution pH 5.1. The total adsorption

2 density of Zn2+ is a sum of the surface concentrations of [SO−−Zn2+] and [SO−−ZnOH+].

3 [SO−−ZnOH+] is the predominant Zn2+ surface species for pH ≥ 5. This was also supported by

4 point of zero surface charge (pHzpc) of eucalyptus bark 5.0 (S. Afroze, T. K. Sen, & H. Ang,

5 2015) which influences adsorbent surface charge. Therefore, this can be attributed to the

t
6 increase in negative charge on the adsorbents surface with increasing solution pH, as solution

ip
7 pH rises towards and above pHPZC and the reduction of H+ ions in solution which competes

cr
8 with Zn2+ for sorption sites. Significantly, the adsorption of monohydroxo complexes is

us
9 consistent with the frequent occurrence of adsorption edges near the pH at which hydrolysis

10 of the free metal ion begins (James, Stiglich, & Healy, 1975) and as pH increases, the

11
an
formation of dissociated surface sites, SO-, is favoured by decreasing [H+].

12
For this study, after initial solution pH 5.1, the sorption capacity for all treated and untreated
M
13

14 samples was found to decrease at pH 5.6 (Fig. 4), due to the decreasing concentration of Zn2+
ed

15 and Zn(OH)+ at higher pH and the appearance of Zn(OH)2 species in solution. Higher initial

16 solution pH were not examined further to avoid Zn2+ precipitation as Zn(OH)2 at pH above
pt

17 5.5 which was experienced here and also confirmed by literature data (N. 1996, Balintova

and Petrilakova 2011). Ofamaja and Naidoo (A. Ofomaja & Naidoo, 2011) also reported
ce

18

19 similar results in their previous work for Cu(II) adsorption on chemically activated pine cone.
Ac

20 Furthermore, higher percentage of Zn2+ removal occurred at higher pH due to presence of

21 less H+ competing for sorption sites on the treated biomass. The removal percentage of Zn2+

22 ions by NaOH treated eucalyptus bark at solution pH 2.5 was 38.80% while at pH 5.1,

23 69.38% removal of Zn2+ ions was observed which was much higher than the removal

24 percentage obtained using raw eucalyptus bark for this identical pH range (for pH 2.5 to 5.1,

25 removal efficiency of Zn2+ ions was only from 19.85% to 40.12%) for which plots are not

26 presented here. From BET analysis, it was observed that the results of total pore volume for

Page 21 of 68
21

1 the treated samples are higher than the raw sample. Therefore, the internal surface of the

2 treated samples is larger than that of the raw samples which leads to increased adsorption of

3 Zn2+ for the treated samples over the raw sample. Similar trend was reported for adsorption of

4 Zn(II) ions by coconut tree sawdust (Putra et al., 2014), neem bark (Bhattacharya et al., 2006)

5 and by carrot residues (Nasernejad et al., 2005).

t
ip
6 3.2.2 Effect of Adsorbent Dosage on Zn2+ Adsorption
7

cr
8 Adsorbent dosage is one of the important parameters in adsorption processes because it

determines the capacity of an adsorbent for a given initial concentration of the adsorbate

us
9

10 under a given set of operating conditions. Adsorbent loading is also significant for the design

11 an
of an adsorber for the optimisation of the process. The adsorption of Zn2+ by raw and NaOH

12 treated eucalyptus bark were conducted with different adsorbent doses (0.01 - 0.03 g) in the
M
13 solution while maintaining the initial metal ions concentration (20 mg/L), temperature (30◦C),

14 stirring speed (120 rpm), contact time (100 min), and pH (5.1) constant. The effect of sorbent
ed

15 dosage on the amount of Zn2+ sorbed at equilibrium is shown in Fig. 5. At equilibrium, Fig. 5

16 showed that increasing the dose of adsorbent from 0.01 to 0.03 g, decreased the amount of
pt

17 Zn2+ adsorbed per unit mass of eucalyptus bark from 44.94 mg/g to 10.74 mg/g for raw EB
ce

18 and for NaOH treated EB, Zn2+ adsorbed amount decreased from 72.52 mg/g to 17.57 mg/g

19 respectively.
Ac

20 Fig. 5 - Effect of adsorbent dosages on the adsorption of Zn2+ onto raw and treated EB
21 powder: (Conditions: Volume of Zn2+ Solution = 50 ml, Solution pH = 5.1, Initial Zn2+
22 Concentration = 20 ppm, Temp = 30°C, Shaker Speed = 120 rpm and Time of
23 Adsorption = 100 min).
24
25
26 The number of adsorption sites per unit mass of an adsorbent should remain constant,

27 independent of the total adsorbent mass; increasing the adsorbent amount in a fixed volume

28 reduces the number of available sites as the effective surface area is likely to decrease (Sen &

Page 22 of 68
22

1 Gomez, 2011). This is because a fixed mass of eucalyptus bark can only adsorb a fixed

2 amount of metal ions. Therefore the higher the adsorbent dosage, the larger the volume of

3 effluent that a fixed mass of eucalyptus bark biomass can purify. The decrease in amount of

4 Zn2+ adsorbed, qe (mg/g) with increasing adsorbent mass, is due to the split in the flux or the

5 concentration gradient between solute concentration in the solution and the solute

t
6 concentration in the surface of the adsorbent (Afroze, Sen, Ang, et al., 2015). Thus, with

ip
7 increasing adsorbent mass, the amount of Zn2+ adsorbed onto unit weight of adsorbent gets

cr
8 reduced and hence causing a decrease in qe value with increasing adsorbent mass

us
9 concentration. Various researchers also explained this matter by the effect of particle

10 concentration which is caused by particle-particle interactions. For the system with higher

11
an
solid content, these interactions may physically block some adsorption sites from the

12 adsorbing solutes and thus, causing decreased adsorption, or the electrical surface charges on
M
13 the closely packed particles reduce attractions between the adsorbing solutes and surfaces of

14 individual adsorbent particles by creating electrostatic interferences (Sen & Gomez, 2011). A
ed

15 similar behaviour was also reported for Zn (II) removal on natural bentonite (Sen & Gomez,

16 2011) and by horseradish tree biomass (Bhatti, Mumtaz, Hanif, & Nadeem, 2007). Another
pt

17 observation from the pattern in the results of Fig. 6 is that even though the trend of adsorbed
ce

18 amount of Zn2+ was decreased with increasing adsorbent concentration for both the systems

19 whereas Zn2+ adsorption capacity was lower for raw EB and higher for treated EB which
Ac

20 indicating the increase in Zn2+ capacity with NaOH treated EB. The interpretation to this is

21 also that the NaOH treated EB has a higher surface sites for adsorbate ions compared to

22 untreated raw EB. Similar observation was also observed for the removal of Cu(II) by raw

23 and base modified pine cone (A. Ofomaja et al., 2009).

24

Page 23 of 68
23

1 3.2.3 Effect of Contact Time and Initial metal ions Concentration on Zn2+ Adsorption
2 Kinetics
3

4 The adsorptive metal removal mechanism is of great significance for developing sorbent-

5 based water technology and is particularly dependent on the initial heavy metal ions

6 concentration. In order to establish equilibration time for maximum adsorption and to know

t
7 the kinetics of adsorption process, the effects of initial metal concentration with time on the

ip
8 adsorption of Zn2+ ions were investigated onto raw and NaOH treated eucalyptus bark, and

cr
9 the results for the treated eucalyptus bark which are presented in Fig. 6(a) and Fig. 6(b)

us
10 respectively.

11

percentage (%) of
an
Fig. 6(a) – Effect of initial metal concentrations and (b) the removal
Zn2+ adsorption by NaOH treated EB powder: 13
12

(Conditions: Mass of Adsorbent = 20 mg, Volume of Zn2+ Solution = 50 ml,


14
Solution pH =5.1, Temp = 30°C and Shaker Speed = 120 rpm).
M
15
16
ed

17 From Fig. 6, it was found that metal sorption occurred rapidly and an increase in initial Zn2+

18 concentration leads to an increase in the sorption capacity of Zn2+ by both raw and treated
pt

19 eucalyptus barks. The equilibrium uptake increased with higher initial metal ions

concentration at the range of experimental initial metal concentrations and equilibrium is


ce

20

21 attained within 100 min. From Fig. 6(a), the amount of Zn2+ adsorption, qt (mg/g) onto NaOH
Ac

22 treated eucalyptus bark, was observed to increase from 34.69 mg/g to 59.82 mg/g with an

23 increase in initial Zn2+ concentration from 20 ppm (mg/L) to 70 ppm (mg/L) which is higher

24 than the amount of Zn2+ adsorbed qt (mg/g) by untreated raw bark adsorbent (20.06 mg/g to

25 59.06 mg/g with the same initial Zn2+ concentration range from 20 to 70 ppm, respectively)

26 for which plot are not been presented here. Further, it was also witnessed from Fig. 6(b) that

27 the percentage removal of Zn2+ increased from 34.18% to 69.38% with decreasing initial

28 concentration of Zn2+ from 70 to 20 ppm (mg/L). In general, it can be concluded that the

Page 24 of 68
24

1 amount of adsorption increases and removal percentage (%) decreases with increasing initial

2 metal ion concentration [Figs. 6(a) and 6(b)] for both the systems.

3 As discussed by Wang et al. (Wang, Chen, Yang, & Ma, 2015), high initial metal

4 concentration accelerates the driving force and reduces the mass transfer resistance. Hence,

5 the increase in initial metal concentration also enhances the interaction between adsorbent

t
6 and metal ions in aqueous solution. It is also found from Fig. 6(b) that the removal percentage

ip
7 of Zn2+ ions by adsorption on NaOH treated EB is very fast at the initial period of contact but

cr
8 slowed down with time. This kinetic experiment clearly indicated that adsorption of Zn2+

us
9 metal ions on NaOH treated eucalyptus bark is a more or less two-step process: a very rapid

10 adsorption of Zn2+ metal ions to the external surface followed by possible slow intraparticle

11
an
diffusion in the interior of the adsorbent. The rapid kinetics has significant practical

12 importance, as it facilitates smaller reactor volumes, ensuring high efficiency and economy
M
13 (Arias & Sen, 2009; Sen, Afroze, & Ang, 2011). These findings are in agreement with the

14 observation of other researchers for the removal of Zn(II) by eggshell (Putra et al., 2014), by
ed

15 neem leaf (Naiya, Chowdhury, Bhattacharya, & Das, 2009) and adsorption of lead(II) onto

16 raw and NaOH modified pine cone powder (A. Ofomaja & Naidoo, 2010).
pt

17
ce

18 3.2.4 Effect of Temperature on Zn2+ Adsorption


19
Ac

20 To observe the effect of temperature on the adsorption capacity, experiments were carried out

21 at three different temperatures of 30, 45 and 60°C for a fixed initial metal ion concentration

22 of 20 ppm (mg/L). Fig. 7(a-b) illustrates the effect of temperature dependence on the

23 adsorption of Zn2+ onto raw treated eucalyptus bark (EB) where the results showed that the

24 amount of adsorption, qt (mg/g) for NaOH modified adsorbent is higher compared with

25 untreated raw adsorbent. But with an increase in temperature from 30oC to 60oC equilibrium

26 adsorption capacity, qe, is decreased from 20.06 mg/g to 17.47 mg/g for raw eucalyptus bark

Page 25 of 68
25

1 [Fig. 7(a)] and 34.69 mg/g to 20.84 mg/g for NaOH treated eucalyptus bark [Fig. 7(b)]. It was

2 also found that the increase in temperature from 30oC to 60oC resulted in decrease in the

3 removal percentage (%) of Zn2+ adsorption from 69.38% to 41.68% for NaOH treated

4 samples and 40.12% to 34.94% for untreated samples for which plots are not presented here.

5 Fig. 7 - Effect of temperature on Zn2+ adsorption (a) onto raw EB powder and (b)
6 onto NaOH modified EB powder: (Conditions: Mass of Adsorbent = 20 mg, Volume of

t
7 Zn2+ Solution = 50 ml, Solution pH = 5.1, Initial Zn2+ Concentration = 20 ppm and

ip
8 Shaker Speed = 120 rpm).
9

cr
10

The fact that the adsorption of Zn2+ metal ions was not in favor of temperature indicates that

us
11

12 the attractive forces and mobility of the adsorbate Zn2+ metal ions towards eucalyptus bark

13 an
adsorbent decreased with increase in temperature which resulted in decreased sorption. This

14 reduction in adsorption capacity with increasing temperature suggests that the process of Zn2+
M
15 removal by both raw and NaOH treated eucalyptus bark biomass is exothermic in nature. It is

16 also indicating that the adsorption process is sustainable and energy efficient as it is favorable
ed

17 at room temperature. Similar types of results are also obtained by earlier investigators for

18 different adsorbent systems (Gupta & Bhattacharyya, 2006; Sen & Gomez, 2011).
pt

19
ce

20 3.2.5 Effect of Presence of Salts on Zn2+ Adsorption Kinetics


21

Industrial wastes and natural water often contain dissolved salts in a wide range of
Ac

22

23 concentrations depending on the source and the quality of water. Presence of electrolytes in

24 solution strongly influences interfacial potential, compete with heavy metal ions for binding

25 sites on the sorbent which affect the sorption of metal ions and thus influences adsorption

26 process also. Hence, there is a need to ascertain the influence of this factor on eucalyptus bark

27 material. The effect of salt concentration (ionic strength) on the amount of Zn2+ sorbed by

28 eucalyptus bark (raw and treated) was analysed over a series of experiments by adding NaCl,

Page 26 of 68
26

1 CaCl2 and FeCl3 salts of initial concentrations ranging from 100 – 300 mg/L (ppm) to the

2 Zn2+ solution. The obtained results for NaOH treated bark biomass are reported in Fig. 8(a-b).

3 The results presented in Fig. 8(a) and 8(b) indicate that due to the presence of mono (Na+), di

4 (Ca2+) and trivalent (Fe3+) salt ions in metal Zn2+ solution the removal efficiency of Zn2+ by

5 chemically treated eucalyptus bark biomass was reduced and by increasing the salt

t
6 concentrations the adsorption capacity of eucalyptus bark further decreased more

ip
7 significantly. Similar trends were observed for the samples examined with raw eucalyptus

cr
8 bark for which results are not presented here. Further Figs. 8(a-b) indicate that the adsorption

us
9 of positively charged Zn2+ on negatively charged eucalyptus bark biomass (at this pH)

10 enhanced with the nature of salt in the order of Na+>Ca2+>Fe3+.

11
an
Fig. 8(a) - Effect of presence of salts on the amount of Zn2+ adsorbed onto NaOH treated
12 EB powder: (Conditions: Mass of Adsorbent = 20 mg, Total Reaction Volume = 50 ml
13 (48 ml of Zn2+ Solution + 2 ml salt solution), Initial Zn2+ Concentration = 20 ppm,
M
14 Solution pH = 5.1, Temp = 30°C, Shaker Speed = 120 rpm and Time of Adsorption =
15 100 min).
16
ed

17
pt

18 Fig. 8(b) - Effect of presence of salts on removal percentage (%) of Zn2+ onto NaOH
19 treated EB powder: (Conditions: Mass of Adsorbent = 20 mg, Total Reaction Volume =
50 ml (48 ml of Zn2+ Solution + 2 ml salt solution), Initial Zn2+ Concentration = 20 ppm,
ce

20
21 Solution pH = 5.1, Temp = 30°C, Shaker Speed = 120 rpm and Time of Adsorption =
22 100 min)
23
Ac

24 The trend of Fig. 8(a) indicated that adsorbed amount of Zn2+ by treated eucalyptus bark was

25 decreased due to the presence of ionic strength which could be attributed to the competitive

26 effect between Zn2+ ions and cations from the salt (Na+/Ca2+/Fe3+) for the sites available for

27 the sorption process and it was more significantly decreased with increasing salt

28 concentration. However, in every test series we experienced that the higher the Na+ ionic

29 concentration, the smaller the percentage amount that adsorbed onto the adsorbent. The

Page 27 of 68
27

1 reason behind this may be the increase in salt concentration (Na+) changes the equilibrium

2 constant between the interface and the bulk of the liquid ultimately influencing adsorption.

3 Similar trend was observed for divalent (CaCl2) and trivalent (FeCl3) salts but comparatively

4 to a lesser extent.

5 The molecular weight, size, atomic radii and charge density may also play a major role of the

t
ip
6 adsorption efficiency and either limit or increase the possibility of the adsorption onto

7 adsorbent (Arivoli, Hema, Parthasarathy, & Manju, 2010). From Fig. 8(a), it was also

cr
8 observed that removal efficiency of Zn2+ adsorption significantly decreased with

us
9 divalent/trivalent salt cations compared to monovalent cations. Hence, divalent calcium may

10 apparently reduce more effectively than monovalent ions binding sites of treated EB for co-

11
an
adsorbing ions (Karppi, Åkerman, Åkerman, Sundell, & Penttilä, 2010). Again, transition

12 metals (Fe) have a much greater affinity for the surface as compared to alkaline earth metals
M
13 (Ca) (Trivedi, Axe, & Dyer, 2001). Therefore, due to relatively small atomic radius of

transition metal element (Fe) and having more electrons that can participate in the chemical
ed

14

15 bonding (Sahin & Peeters, 2013), stronger binding of Fe3+ to the eucalyptus bark is expected
pt

16 which further reduces adsorption efficiency of Zn2+. From Fig. 8(a), the adsorbed amounts

17 suggest that adsorption affinity of studied cations onto the treated EB biomass followed the
ce

18 order Na+ < Ca2+ <Fe3+. The decrease in percentage removal capacity of Zn2+ by treated EB

[Fig. 8(b)] with increasing ionic strength is primarily attributed to the reduction in difference
Ac

19

20 in the ionic osmotic pressure between the sorbent and the external solution (Adeyinka, Liang,

21 & Tina, 2007; Ansari & Pornahad, 2010). The difference in osmotic pressure between the

22 sorbent and the external solution decreases as the ionic strength of external solution increases.

23 Therefore, the sorption of metal ion decreases when the ionic strength of external solution

24 increases.

25

Page 28 of 68
28

1 In the presence of salts, similar results were also observed by Osemeahon et al. 2013 on the

2 removal of Zinc ions by Immobilized Bombax costatum calyx (Osemeahon, Barminas, &

3 HammaAdama, 2013) and by Ghodbane et al. 2008 on the adsorption of Cadmium ions by

4 eucalyptus bark (Ghodbane et al., 2008). However, despite decrease of adsorption efficiency

5 with the presence of salts, sorbent potential is still relatively high and it may still effectively

t
6 remove Zn2+ metal ions from heavy metal bearing solutions.

ip
7

cr
8 3.3 Desorption Studies
9

us
10 Results of desorption study is presented in Table 2. Desorption studies on the raw and base

11 modified eucalyptus bark adsorbents indicated that Zn2+ was poorly desorbed from the

12
an
biomass. Percentage desorption was defined as released Zn2+ concentration over initially

sorbed concentration of Zn2+ in percentage which was less than 15% for the concentration
M
13

14 range studied and lesser amounts of Zn2+ were desorbed from biomass with higher Zn2+
ed

15 concentration. Only about 14% and 7% of the bound Zn2+ was recovered from raw and

16 NaOH treated bark biomass respectively. This suggests that strong binding affinity exists
pt

17 between the biomass and Zn2+ implying possible presence of strong complexes and ligands

between adsorbent and Zn2+. Further eucalyptus bark is a solid waste with no value and hence
ce

18

19 regeneration of adsorbent is insignificant


Ac

20 Table 2 – Percentage desorption of Zn2+ from raw and NaOH treated eucalyptus bark
21 (EB) biomass

22

23 3.4 Adsorption Kinetics and Mechanism of Adsorption


24

25 3.4.1 Application of Pseudo-First-Order and Pseudo-Second-Order Model


26

Page 29 of 68
29

1 In this present study, to investigate the adsorption kinetics of Zn2+ metal ions onto raw and

2 treated eucalyptus bark and to evaluate the applicability of the pseudo-first- order, pseudo-

3 second-order and intraparticle diffusion kinetic models were applied to experimental data at

4 various physico-chemical conditions such as for initial dye concentration, initial solution pH,

5 adsorbent dosages, temperature and addition of salts. Kinetic experimental data were fitted

t
6 into plots of log (qe – qt) against t using Eq.(4) according to the pseudo first order model for

ip
7 which figures are not presented here. From those plots, the pseudo-first-order rate constant

cr
8 (K2) and linear regression coefficient (R2) have been calculated which gives poor value for

us
9 both the systems. In addition, pseudo-first-order kinetic model predicts a much lower value of

10 the equilibrium adsorption capacity than the experimental value and hence gives the

11
an
inapplicability of this model. Conversely, pseudo-second-order model was fitted into

12 experimental data (for which plots are not presented here) where the calculated values of qe
M
13 were very close to the experimental qe values. For both systems, the initial sorption rate h,

14 pseudo-second-order rate constant K2, amount of Zn2+ adsorbed at equilibrium qe and the
ed

15 higher regression coefficient R2 values obtained from the slope of the linearized form of the

16 plot t/qt verses time (t) [Eq. (5)] which are presented in Table 3. Chi-square test (χ2) was
pt

17 performed as per Eq. (17) where qcalc presents the calculated adsorption capacity (mg/g) and
ce

18 qe is the equilibrium capacity (mg/g) from the experimental data. χ2 of less than 0.3 indicated

19 that Zn2+ adsorption onto raw and NaOH treated eucalyptus bark biomass followed pseudo-
Ac

20 second order model as shown in Table 3. Therefore, it can be concluded that the adsorption

21 kinetics was followed by pseudo-second-order model which relies on the assumption that

22 adsorption may be the rate-limiting step involving valence forces through sharing or

23 exchange of electrons between adsorbent and sorbate as covalent forces (Saeed, Akhter, &

24 Iqbal, 2005). From Table 3, it can be observed that for both raw and NaOH treated eucalyptus

25 bark, the adsorption capacity increases with increase in initial metal ion concentration, initial

Page 30 of 68
30

1 solution pH but decreases with adsorbent amount and temperature respectively. Moreover,

2 the equilibrium adsorption capacity, qe values are much higher for NaOH treated samples

3 than raw ones which indicate stronger bond formation between Zn2+ and eucalyptus bark

4 treated with NaOH and hence the surface of the eucalyptus bark samples increased with

5 NaOH treatment. It was also noted that the values of the pseudo-second order capacities, qe,

t
6 were pretty close to the experimentally determined capacities signifying the applicability of

ip
7 this model. The high linear correlation coefficient R2 values also support the applicability of

cr
8 pseudo-second order model for both the systems.

us
9 (17)
10

11
12 NaOH treated eucalyptus bark (EB)
an
Table 3 - Pseudo Second Order Model Parameters for adsorption of Zn2+ on raw and

13
M
14 3.4.2 Application of Intra Particle Diffusion Model and Mechanism of Zn2+ Adsorption

15 To investigate the mechanism of Zn2+ adsorption on raw and pre-treated eucalyptus bark by
ed

16 NaOH, the experimental data were fitted in the intraparticle diffusion plot of the amount of

17 Zn2+ sorbed per unit mass of sorbent against the square root of contact time for both the
pt

18 systems. From the plot of amount sorbed, qt (mg/g) versus t0.5 (min0.5), the obtained results
ce

19 for treated samples depicted in Fig. 9 shows that the sorption process is not linear over the

20 whole period of time and tends to be split into two to three regions which confirm the multi-
Ac

21 stages of adsorption. The initial linear portion of the plot indicates a boundary layer effect,

22 while the second portion is due to intraparticle diffusion and the third to the chemical reaction

23 (A. E. Ofomaja, 2008).

24

Fig. 9 - Intra-particle diffusion model of Zn2+ uptake by NaOH treated 25


eucalyptus bark (Conditions: Mass of Adsorbent = 20 mg, Volume of Zn 2+
Solution = 50 ml, Solution pH = 5.1, Temp = 30°C, Adsorption Time = 100 26
min
and Shaker Speed = 120 rpm).

Page 31 of 68
31

1 The experimental data in Fig. 9 indicates that the metal ions were transported to the external

2 surface of the treated eucalyptus bark particle through film diffusion and its rate was fast.

3 After that, Zn2+ metal ions entered into the adsorbent particles by intraparticle diffusion

4 through pores. This indicates that the intra-particle diffusion is involved in the adsorption

5 process of Zn2+ onto the treated biomass but not the only rate controlling step. Using Eq. (7),

t
6 the slope of the second portion of the plot is defined as the intraparticle diffusion parameter,

ip
7 Kid (mg/(g min0.5)) and the intercept of the plot, C reflects the boundary layer effect. Table 4

cr
8 shows the intraparticle diffusion parameters, Kid and C obtained from the plot of amount Zn2+

us
9 adsorbed per unit mass versus the square root of time, t0.5 for the eucalyptus bark and NaOH

10 treated eucalyptus bark samples. The order in the magnitude of Kid and C for raw and treated

11
an
NaOH adsorbent increased from 20 to 70 ppm (mg/L). Increase in Kid and C with initial

12 concentration can be attributed to increase in concentration gradient as initial concentration of


M
13 Zn2+ in solution is increased, forcing more Zn2+ ions to migrate to the adsorbent surface (A.

14 Ofomaja & Naidoo, 2011). Consequently, the larger the intercept, the greater the contribution
ed

15 of the surface sorption in the rate limiting step (A. Ofomaja et al., 2009).

16 Table 4 – Parameters of Intraparticle Diffusion Model for Zn2+ adsorption on raw and
pt

17 NaOH treated Eucalyptus Bark (EB) at different concentrations.


18
ce

19 3.5 Adsorption Equilibrium Isotherm


20
Ac

21 Equilibrium studies determine the capacity of the adsorbent and describe the adsorption

22 mechanism. For this purpose, the experimental data were fitted to the most popular isotherm

23 models of Langmuir and Freundlich within the metal ion concentration range of 20 to 70 ppm

24 (mg/L) and isotherm parameters were calculated. Linear regression analysis was then used to

25 determine the best fitted isotherm.

26

Page 32 of 68
32

1 3.5.1 Freundlich Adsorption Isotherm


2

3 To study the Freundlich Isotherm, adsorption equilibrium data were fitted with experimental

4 data for both raw and NaOH treated system and the obtained results for both samples are

5 presented in Fig. 10. The high value of linear correlation coefficient, R2 of the Freundlich

6 isotherm is 0.9331 and 0.9273 for raw eucalyptus bark and NaOH modified eucalyptus bark

t
ip
7 adsorbent respectively which indicates the applicability of this model (Fig. 10). The

calculated values of Kf and bf from the slope and intercept of the plot (Fig. 10) of lnqe versus

cr
8

9 lnCe using Eq. (8) are tabulated on Table 5. The Freundlich isotherm constant, Kf was found

us
10 to increase with NaOH pre-treatment. The Freundlich constant bf is a measure of the

11 deviation from linearity of the adsorption or the adsorption affinity of the adsorbent for a

12
an
metal ion in solution and in this study, the numerical values of 1/bf for the adsorbents are

always greater than unity, indicating that Zn2+ are favourably adsorbed by raw eucalyptus and
M
13

14 NaOH treated eucalyptus bark (Demiral, Demiral, Tümsek, & Karabacakoğlu, 2008).
ed

15

16 Fig. 10 - Freundlich plot for raw and NaOH treated EB: Amount of adsorbent
10 mg; Initial Zn2+ Concentration = 20, 30, 40, 50, 60 and 70 ppm, Soluiton pH
pt

17
≈ 5.1, Temperature = 30◦C; Shaker Speed = 120 rpm and Time of Adsorption
18 = 2 hr.
ce

19

20 3.5.2 Langmuir Adsorption Isotherm


Ac

21 According to Langmuir model, adsorption occurs uniformly on the active sites of the

22 adsorbent containing finite numbers of identical sites with no transmigration of adsorbate in

23 the plane of surface. Langmuir isotherms are also found to be fitted well for Zn2+ adsorption

24 by both raw and treated eucalyptus bark adsorbents and examined results for raw and NaOH

25 treated EB samples are presented in Figs. 11(a) and 11(b) which make this study suitable to

26 optimize the operating conditions for effective adsorption. The maximum adsorption

Page 33 of 68
33

1 capacity, qm (mg/g) and Ka values for Langmuir constants can be obtained from the linear

2 equations of different plots [Figs. 11(a-b)]. The Langmuir isotherm fits the experimental data

3 very well and Langmuir I showed higher adsorption capacity (mg/g) compared to Langmuir

4 II model for both raw and treated bark adsorbents. The calculated values for Langmuir I show

5 that the monolayer capacities of the different adsorbents were: raw eucalyptus bark (131.58

t
6 mg/g) and NaOH treated eucalyptus bark (222.22 mg/g). These results indicate that

ip
7 monolayer capacity increases with NaOH pre-treatment. Moreover, the Langmuir II isotherm

cr
8 shows that the trends in the parameters, qm and Ka are similar to those of the Langmuir-I

us
9 isotherm.

10

11
12
an
Fig. 11(a) - Equilibrium Adsorption Isotherm fitted to the Langmuir-2 Model : Amount
of Adsorbent added = 10 mg; initial Zn+2 Concentration = 20, 30, 40, 50, 60, 70 ppm;
13 Soluiton pH ≈ 5.1; Temperature = 30°C; Shaker Speed = 120 rpm and Time of
M
14 Adsorption = 2 hr min.
15
16
ed

17

Fig. 11(b) - Equilibrium Adsorption Isotherm fitted to the Langmuir-I Model 18 :


pt

+2
Amount of Adsorbent added = 10 mg; Initial Zn Concentration = 20, 30, 40,
50, 60, 70 ppm; Soluiton pH ≈ 5.1 ; Temperature = 30°C; Shaker Speed = 19 120
rpm and Time of Adsorption = 2 hr.
ce

20
.
21 The correlation coefficient, R2, for the equilibrium isotherm models applied in this study

showed that the Langmuir I isotherm had the highest R2 values followed by the Langmuir II
Ac

22

23 and Freundlich isotherms. The Langmuir I isotherm is therefore the best fitting of the three

24 models; indicating monolayer coverage of the adsorbate on the adsorbent and homogeneity of

25 the actives sites on the adsorbent surface. The values of Freundlich & Langmuir model

26 parameters for the adsorption of Zn2+ ions onto raw and treated eucalyptus bark adsorbents

27 are summarised and presented in Table 5 below

Page 34 of 68
34

1 Table 5 – A summary of Freundlich and Langmuir calculated values


2

3 Further, the dimensionless constant separation factor (RL) was used to investigate the

4 adsorption system feasibility at different initial Zn2+ concentrations and obtained RL values,

5 calculated from Langmuir plots using Eq. (11) were were observed to decrease with the

6 increase of initial Zn2+ solution concentration for both raw eucalyputs and NaOH pretreated

t
ip
7 eucalyptus bark and as 0<RL<1, this confirms the favourable uptake of Zn2+ metal ions onto

raw and NaOH treated eucalyptus bark biomass.

cr
8

9
For each parameter set, the values of MPSD error function for adsorption of Zn2+ on raw and

us
10

11 NaOH treated eucalyptus bark are calculated and presented in Table 4. By comparing the

12 an
results of the values for the MPSD error function, it can be concluded that the Freundlich

13 isotherms result in the lowest values for the error function and therefore fits the data better
M
14 than Langmuir isotherm models.

3.6 Thermodynamic Studies


ed

15
16

17 To understand the changes in the reaction during the process, the thermodynamic parameters
pt

18 have been calculated. The values for change in Gibbs free energy of activation (ΔGo), were
ce

19 calculated by knowing the values of enthalpy change (ΔHo) and change of entropy (ΔSo)

20 using Eq. (13) and Eq. (14) which were obtained from the slope and intercept of the plot log
Ac

21 (qe/Ce) verses 1/T shown in Fig.12 for treated and untreated eucalyptus bark. Further, the

22 corresponding values for change in free energy of activation (ΔGo), enthalpy change (ΔHo)

23 and entropy changes (ΔSo) for raw and NaOH treated eucalyptus bark are presented in Table

24 6

25 Fig. 12 – Van’t Hoff plot for adsorption of Zn2+ onto raw and NaOH treated EB
26

Page 35 of 68
35

1 From Table 6, the change in Gibbs free energy (ΔGo) and enthalpy (ΔHo) were found to be

2 negative at all temperatures suggesting exothermic and spontaneous nature of the process.

3 The negative value of entropy (ΔSo) indicates decreased randomness during adsorption.

4 Similar results for thermodynamic parameters were also reported by earlier researchers for

5 the adsorption of Zn2+ from aqueous solution on various adsorbents (Arias & Sen, 2009;

t
6 Bhattacharya et al., 2006; Sen & Gomez, 2011). Moreover, it was observed that the entropy

ip
7 change (ΔSo), for all raw eucalyptus bark samples were found to be smaller than zero, and the

cr
8 values tend to move towards zero with NaOH treatment. The change in enthalpy (ΔHo)

us
9 obtained is lower than TΔSo for both systems, demonstrating that the reorientation step is

10 mostly entropy controlled at the activation state. The contribution of the reorientation step to

11
an
activation tends to be higher for NaOH treated eucalyptus bark powder biosorption than raw

12 eucalyptus bark as seen from the higher value of enthalpy (ΔHo) for NaOH treated samples.
M
13 The free energy of activation (ΔGo) was found to increase with NaOH treatment indicating

14 that activation becomes more spontaneous with NaOH treatment.


ed

15 Table 6 - Thermodynamic parameters for adsorption of Zn2+ at different temperatures


16 onto eucalyptus bark and NaOH modified eucalyptus bark
pt

17 3.7 Comparison of Zn(II) removal with various low cost adsorbents


reported in literature
ce

18
19

20 The adsorption capacities, qm (mg/g) of the adsorbents obtained in the present study for the
Ac

21 removal of Zn(II) from its aqueous solution have been compared with those of other

22 adsorbents reported in literature and the values of adsorption capacities are presented in Table

23 7. The experimental data of the present investigations are comparable with the reported

24 values for many other agricultural solid wastes and activated carbon adsorbents.

25 Table 7 – Comparison of the adsorption capacity (qm in mg/g) of different sorbents for
26 the removal of Zn(II)
27

Page 36 of 68
36

1 4 Conclusion

3 The removal of Zn2+ from aqueous solution using eucalyptus bark and NaOH modified

4 eucalyptus bark was investigated under different experimental conditions in batch process.

5 The adsorptive effectiveness of raw eucalyptus bark was increased by activation with NaOH.

t
6 This has also been confirmed by BET surface area, bulk density, FTIR analysis and XRD

ip
7 analysis of raw and treated adsorbents. The sorption kinetics was found to be dependent upon

cr
8 contact time, initial metal ion concentration, sorbent dose, solution initial pH, ionic strength

and temperature. The amount of metal ion adsorption on both eucalyptus bark and NaOH

us
9

10 modified eucalyptus bark increases with initial metal ion concentration, the contact time and

11 an
the solution pH. Modelling of sorption kinetics showed good agreement of experimental data

12 with the pseudo-second-order kinetic equation for different initial metal concentrations. The
M
13 mechanism of adsorption was analysed with intraparticle diffusion kinetic model. Both the

14 Freundlich and Langmuir isotherm models were fitted well in the selected range of
ed

15 concentrations. The highest monolayer adsorption capacity was obtained 250.00 mg/g for

16 NaOH pre-treated eucalyptus bark and lowest for raw eucalyptus bark 128.21 mg/g at
pt

17 optimum pH 5.1. The RL values showed that eucalyptus bark was favourable for the sorption
ce

18 of zinc. Desorption experiments clearly indicated that adsorption of Zn2+ metal followed ion-

19 exchange and strong physical-chemical adsorption. It may be possible that high


Ac

20 concentrations of the acid may have positive effect in desorption of Zn2+. Further studies

21 should however be carried out to explore other solvents which may be more effective for

22 Zn2+ desorption from eucalyptus bark biomass. The negative values of free energy change

23 and enthalpy change indicated that the Zn2+ sorption process is spontaneous and exothermic

24 in nature. The present study also demonstrated that the raw and modified eucalyptus bark

Page 37 of 68
37

1 may be an effective adsorbents alternative to many expensive adsorbents used in the removal

2 of Zn2+ metal ions from wastewater.

3 Nomenclature

4 bf Freundlich constant

Ce equilibrium Zn2+ concentration, ppm (mg/L)

t
5

ip
6 C0 initial Zn2+ concentration, ppm (mg/L)

cr
7 Ct Zn2+ concentration at time t, ppm (mg/L)

8 ΔG0 Gibbs free energy change, (kJ/mol)

us
9 ΔH0 enthalpy change, (kJ/mol)

10 h initial adsorption rate (mg/g-min) an


11 Ka Langmuir constant
M
12 K1 pseudo-first-order rate constant (min-1)

13 K2 pseudo-second-order rate constant, (mg/g min)


ed

14 Kf Freundlich adsorption constant, (L/g)

15 Kid intra-particle rate constant ((mg/g) min0.5)


pt

16 MPSD Marquardt’s percent standard deviation

mass of adsorbent per unit volume (g L-1)


ce

17 M

18 m amount of adsorbent added (g)


Ac

19 n number of data points

20 p number of isotherm parameters

21 q amount of adsorbate per g of adsorbent (mg/g)

22 qe amount of adsorbate per g of adsorbent at equilibrium, (mg/g)

23 qt amount of adsorbate per g of adsorbent at any time, t

24 qm equilibrium adsorption capacity

25 qmax maximum adsorption capacity (mg/g)

Page 38 of 68
38

1 qe,calc Calculated metal equilibrium solid phase concentration (mg g-1)


2
3 qe,meas Measured metal equilibrium solid phase concentration (mg g-1)
4
5 R2 linear regression coefficient

6 RL separation factor

7 ΔS0 entropy change, (J/k mol)

t
8 t time (min)

ip
9 T temperature (K)

cr
10 V volume of Zn2+ solution (L)

us
11

12 Acknowledgement

13
an
The authors would like to thank the Chemical Engineering Department of Curtin University-

14 Perth for financial support and Chemical Engineering Laboratory technicians for their
M
15 support during experiments.
ed

16

17
pt

18
ce

19

20
Ac

21

22

23

24

25

26

27 References

Page 39 of 68
39

1 Acharya, J., Sahu, J., Sahoo, B., Mohanty, C. and Meikap, B., 2009, Removal of chromium (VI) from
2 wastewater by activated carbon developed from Tamarind wood activated with zinc
3 chloride, Chemical Engineering Journal, 150(1): 25-39.
4 Adeyinka, A., Liang, H. and Tina, G., 2007, Removal of metal ion form waste water with natural
5 waste, School of Engineering and Technology, 33: 1-8.
6 Afroze, S., Sen, T. and Ang, M., 2015, Agricultural solid wastes in aqueous phase dye adsorption: A
7 Review, In Foster, C. N. (Ed.), Agricultural wastes characteristics, types and management,
8 New York: Nova.
9 Afroze, S., Sen, T. K. and Ang, H., 2015, Adsorption performance of continuous fixed bed column for
10 the removal of methylene blue (MB) dye using Eucalyptus sheathiana bark biomass,

t
11 Research on Chemical Intermediates, 1-22.

ip
12 Afroze, S., Sen, T. K., Ang, M. and Nishioka, H., 2015, Adsorption of methylene blue dye from
13 aqueous solution by novel biomass Eucalyptus sheathiana bark: equilibrium, kinetics,
14 thermodynamics and mechanism, Desalination and Water Treatment, 1-21.

cr
15 Ansari, R. and Pornahad, A., 2010, Removal of cerium (IV) ion from aqueous solutions using sawdust
16 as a very low coast bioadsorbent, Journal of Applied Science and Environmental
17 Sanitation,239-247.

us
18 Arias, F. and Sen, T. K., 2009, Removal of zinc metal ion (Zn 2+) from its aqueous solution by kaolin
19 clay mineral: a kinetic and equilibrium study, Colloids and Surfaces A: Physicochemical and
20 Engineering Aspects, 348(1): 100-108.
21
22
23
an
Arivoli, S., Hema, M., Parthasarathy, S. and Manju, N., 2010, Adsorption dynamics of methylene blue
by acid activated carbon, Journal of Chemical and Pharmaceutical Research, 2(5): 626-641.
Bharathi, K. and Ramesh, S., 2013, Removal of dyes using agricultural waste as low-cost adsorbents:
24 a review, Applied Water Science, 3(4): 773-790.
M
25 Bhattacharya, A., Mandal, S. and Das, S., 2006, Adsorption of Zn (II) from aqueous solution by using
26 different adsorbents, Chemical Engineering Journal, 123(1): 43-51.
27 Bhatti, H. N., Mumtaz, B., Hanif, M. A. and Nadeem, R., 2007, Removal of Zn (II) ions from aqueous
28 solution using Moringa oleifera Lam.(horseradish tree) biomass, Process Biochemistry, 42(4):
ed

29 547-553.
30 Boota, R., Bhatti, H. N. and Hanif, M. A., 2009, Removal of Cu (II) and Zn (II) using lignocellulosic fiber
31 derived from Citrus reticulata (Kinnow) waste biomass, Separation Science and Technology,
32 44(16): 4000-4022.
pt

33 Choi, J.-W., Chung, S.-G., Hong, S.-W., Kim, D.-J. and Lee, S.-H., 2012, Development of an
34 environmentally friendly adsorbent for the removal of toxic heavy metals from aqueous
ce

35 solution, Water, Air, & Soil Pollution, 223(4): 1837-1846.


36 Davis, J. A. and Leckie, J. O., 1978, Surface ionization and complexation at the oxide/water interface
37 II. Surface properties of amorphous iron oxyhydroxide and adsorption of metal ions, Journal
38 of Colloid and Interface Science, 67(1): 90-107.
Ac

39 Dawood, S., Sen, T. K. and Phan, C., 2014, Synthesis and characterisation of novel-activated carbon
40 from waste biomass pine cone and its application in the removal of congo red dye from
41 aqueous solution by adsorption, Water, Air, & Soil Pollution, 225(1): 1-16.
42 Demiral, H., Demiral, İ., Tümsek, F. and Karabacakoğlu, B., 2008, Adsorption of chromium (VI) from
43 aqueous solution by activated carbon derived from olive bagasse and applicability of
44 different adsorption models, Chemical Engineering Journal, 144(2): 188-196.
45 Elizalde-González, M. P., Mattusch, J., Peláez-Cid, A. A. and Wennrich, R., 2007, Characterization of
46 adsorbent materials prepared from avocado kernel seeds: Natural, activated and carbonized
47 forms, Journal of Analytical and Applied Pyrolysis, 78(1): 185-193.
48 FENG, N.-c. and GUO, X.-y., 2012, Characterization of adsorptive capacity and mechanisms on
49 adsorption of copper, lead and zinc by modified orange peel, Transactions of Nonferrous
50 Metals Society of China, 22(5): 1224-1231.

Page 40 of 68
40

1 Ferro-Garcia, M., Rivera-Utrilla, J., Rodriguez-Gordillo, J. and Bautista-Toledo, I., 1988, Adsorption of
2 zinc, cadmium, and copper on activated carbons obtained from agricultural by-products,
3 Carbon, 26(3): 363-373.
4 Freundlich, H., 1906, Over the adsorption in solution, Journal of Physical Chemistry, 57(385): e470.
5 Gautam, R. K., Mudhoo, A., Lofrano, G. and Chattopadhyaya, M. C., 2014, Biomass-derived
6 biosorbents for metal ions sequestration: Adsorbent modification and activation methods
7 and adsorbent regeneration, Journal of Environmental Chemical Engineering, 2(1): 239-259.
8 Ghodbane, I. and Hamdaoui, O., 2008, Removal of mercury (II) from aqueous media using eucalyptus
9 bark: kinetic and equilibrium studies, Journal of Hazardous Materials, 160(2): 301-309.
10 Ghodbane, I., Nouri, L., Hamdaoui, O. and Chiha, M., 2008, Kinetic and equilibrium study for the

t
11 sorption of cadmium (II) ions from aqueous phase by eucalyptus bark. Journal of Hazardous

ip
12 Materials, 152(1): 148-158.
13 Gupta, S. S. and Bhattacharyya, K. G., 2006, Adsorption of Ni (II) on clays, Journal of Colloid and
14 Interface Science, 295(1): 21-32.

cr
15 Hubbe, M. A., Hasan, S. H. and Ducoste, J. J., 2011, Cellulosic substrates for removal of pollutants
16 from aqueous systems: A review. 1. Metals, BioResources, 6(2): 2161-2287.
17 James, R. O., Stiglich, P. and Healy, T., 1975, Analysis of models of adsorption of metal ions at

us
18 oxide/water interfaces, Faraday Discussions of the Chemical Society, 59: 142-156.
19 Karppi, J., Åkerman, S., Åkerman, K., Sundell, A. and & Penttilä, I., 2010, Adsorption of metal cations
20 from aqueous solutions onto the pH responsive poly (vinylidene fluoride grafted poly (acrylic
21
22
23
an
acid)(PVDF-PAA) membrane, Journal of Polymer Research, 17(1): 71-76.
Knocke, W. and Hemphill, L., 1981, Mercury (II) sorption by waste rubber, Water Research, 15(2):
275-282.
24 Lagergren, S., 1898, About the theory of so-called adsorption of soluble substances, Kungliga
M
25 Svenska Vetenskapsakademiens, Handlingar, 24(4): 1-39.

26 Langmuir, I. , 1918, The adsorption of gases on plane surfaces of glass, mica and platinum, Journal of
27 the American Chemical Society, 40(9): 1361-1403.
ed

28 Liu, C., Ngo, H. H. and Guo, W., 2012, Watermelon rind: agro-waste or superior biosorbent? Applied
29 Biochemistry and Biotechnology, 167(6): 1699-1715.
30 Maheshwari, U., Mathesan, B. and Gupta, S., 2015, Efficient adsorbent for simultaneous removal of
pt

31 Cu (II), Zn (II) and Cr (VI): Kinetic, thermodynamics and mass transfer mechanism, Process
32 Safety and Environmental Protection, 98: 198-210.
33 Marquardt, D. W., 1963, An algorithm for least-squares estimation of nonlinear parameters, Journal
ce

34 of the Society for Industrial & Applied Mathematics, 11(2): 431-441.


35 Marshall, W., Wartelle, L., Boler, D., Johns, M. and Toles, C., 1999, Enhanced metal adsorption by
36 soybean hulls modified with citric acid, Bioresource Technology, 69(3): 263-268.
37 Marshall, W. E. and Johns, M. M., 1996, Agricultural by‐products as metal adsorbents: sorption
Ac

38 properties and resistance to mechanical abrasion, Journal of Chemical Technology and


39 Biotechnology, 66(2): 192-198.
40 McLaughlan, R., Hossain, S. and Al-Mashaqbeh, O. A., 2015, Zinc sorption by permanganate treated
41 pine chips, Journal of Environmental Chemical Engineering, 3(3): 1539-1545.
42 Mishra, V., Balomajumder, C. and Agarwal, V. K., 2010, Zn (II) ion biosorption onto surface of
43 eucalyptus leaf biomass: Isotherm, kinetic, and mechanistic modeling, CLEAN–Soil, Air,
44 Water, 38(11): 1062-1073.
45 Mishra, V., Balomajumder, C. and Agarwal, V. K., 2012, Kinetics, mechanistic and thermodynamics of
46 Zn (II) ion sorption: a modeling approach, CLEAN–Soil, Air, Water, 40(7): 718-727.
47 Naiya, T. K., Chowdhury, P., Bhattacharya, A. K. and Das, S. K., 2009, Saw dust and neem bark as low-
48 cost natural biosorbent for adsorptive removal of Zn (II) and Cd (II) ions from aqueous
49 solutions, Chemical Engineering Journal, 148(1): 68-79.

Page 41 of 68
41

1 Nasernejad, B., Zadeh, T. E., Pour, B. B., Bygi, M. E. and Zamani, A., 2005, Camparison for biosorption
2 modeling of heavy metals (Cr (III), Cu (II), Zn (II)) adsorption from wastewater by carrot
3 residues, Process Biochemistry, 40(3): 1319-1322.
4 Nguyen, T., Ngo, H., Guo, W., Zhang, J., Liang, S., Yue, Q. and Nguyen, T., 2013, Applicability of
5 agricultural waste and by-products for adsorptive removal of heavy metals from
6 wastewater, Bioresource Technology, 148: 574-585.
7 Ofomaja, A. and Naidoo, E., 2010, Biosorption of lead (II) onto pine cone powder: studies on
8 biosorption performance and process design to minimize biosorbent mass, Carbohydrate
9 Polymers, 82(4): 1031-1042.
10 Ofomaja, A. and Naidoo, E., 2011, Biosorption of copper from aqueous solution by chemically

t
11 activated pine cone: a kinetic study, Chemical Engineering Journal, 175: 260-270.

ip
12 Ofomaja, A., Naidoo, E. and Modise, S., 2009, Removal of copper (II) from aqueous solution by pine
13 and base modified pine cone powder as biosorbent, Journal of Hazardous Materials, 168(2):
14 909-917.

cr
15 Ofomaja, A. E., 2008, Sorptive removal of Methylene blue from aqueous solution using palm kernel
16 fibre: Effect of fibre dose, Biochemical Engineering Journal, 40(1): 8-18.
17 Osemeahon, S., Barminas, J. and HammaAdama, M., 2013, Studies on the removal of metal ions

us
18 from aqueous solution using Immobilized Bombax costatum calyx, IOSR Journal Of
19 Environmental Science, Toxicology And Food Technology (IOSR-JESTFT), 3(6): 6-13.
20 Paduraru, C., Tofan, L., Teodosiu, C., Bunia, I., Tudorachi, N. and Toma, O., 2015, Biosorption of zinc
21
22
23
an
(II) on rapeseed waste: Equilibrium studies and thermogravimetric investigations, Process
Safety and Environmental Protection, 94: 18-28.
Pérez Marín, A. B., Aguilar, M. I., Ortuño, J. F., Meseguer, V. F., Sáez, J. and Lloréns, M., 2010,
24 Biosorption of Zn (II) by orange waste in batch and packed‐bed systems, Journal of Chemical
M
25 Technology and Biotechnology, 85(10): 1310-1318.
26 Putra, W. P., Kamari, A., Yusoff, S. N. M., Ishak, C. F., Mohamed, A., Hashim, N. and Isa, I. M., 2014,
27 Biosorption of Cu (II), Pb (II) and Zn (II) ions from aqueous solutions using selected waste
28 materials: Adsorption and characterisation studies, Journal of Encapsulation and Adsorption
ed

29 Sciences, 4(1).
30 Saeed, A., Akhter, M. W. and Iqbal, M., 2005, Removal and recovery of heavy metals from aqueous
31 solution using papaya wood as a new biosorbent, Separation and Purification Technology,
32 45(1): 25-31.
pt

33 Safe Drinking Water Committee, N. R. and Council, 1977, Drinking Water and Health, Volume 1: 948.
34 Sahin, H. and Peeters, F. M., 2013, Adsorption of alkali, alkaline-earth, and 3 d transition metal
ce

35 atoms on silicene. Physical Review B, 87(8): 085423.


36 Saliba, Gauthier, Gauthier and Petit-Ramel,. 2002, The use of eucalyptus barks for the adsorption of
37 heavy metal ions and dyes, Adsorption Science & Technology, 20(2): 119-129.
38 Sen, T. K., Afroze, S. and Ang, H., 2011, Equilibrium, kinetics and mechanism of removal of methylene
Ac

39 blue from aqueous solution by adsorption onto pine cone biomass of Pinus radiata, Water,
40 Air, & Soil Pollution, 218(1-4): 499-515.
41 Sen, T. K. and Gomez, D., 2011, Adsorption of zinc (Zn 2+) from aqueous solution on natural
42 bentonite, Desalination, 267(2): 286-294.
43 Sen, T. K. and Sarzali, M. V., 2008, Removal of cadmium metal ion (Cd 2+) from its aqueous solution
44 by aluminium oxide (Al2O3): a kinetic and equilibrium study, Chemical Engineering Journal,
45 142(3): 256-262.
46 Senthil Kumar, P., Ramalingam, S., Abhinaya, R. V., Kirupha, S. D., Murugesan, A. and Sivanesan, S.,
47 2012, Adsorption of metal ions onto the chemically modified agricultural waste, CLEAN–Soil,
48 Air, Water, 40(2): 188-197.
49 Singha, A. S. and Guleria, A., 2015, Utility of chemically modified agricultural waste okra biomass for
50 removal of toxic heavy metal ions from aqueous solution, Engineering in Agriculture,
51 Environment and Food, 8(1): 52-60.

Page 42 of 68
42

1 Singha, B. and Das, S. K., 2013, Adsorptive removal of Cu (II) from aqueous solution and industrial
2 effluent using natural/agricultural wastes, Colloids and Surfaces B: Biointerfaces, 107: 97-
3 106.
4 Sud, D., Mahajan, G. and Kaur, M., 2008, Agricultural waste material as potential adsorbent for
5 sequestering heavy metal ions from aqueous solutions–A review. Bioresource Technology,
6 99(14): 6017-6027.
7 Taha, G., Arifien, A. and El-Nahas, S., 2011, Removal efficiency of potato peels as a new biosorbent
8 material for uptake of Pb (II) Cd (II) and Zn (II) from their aqueous solutions, The Journal of
9 Solid Waste Technology and Management, 37(2): 128-140.
10 Trivedi, P., Axe, L. and Dyer, J., 2001, Adsorption of metal ions onto goethite: single-adsorbate and

t
11 competitive systems, Colloids and Surfaces A: Physicochemical and Engineering Aspects,

ip
12 191(1): 107-121.
13 University of Colorado, 1985, Table of Characteristic IR Absorptions, Retrieved December 27, 2013,
14 from orgchem.colorado.edu/Spectroscopy/specttutor/irchart.pdf

cr
15 Wang, L., Chen, Z., Yang, J. and Ma, F., 2015, Pb (II) biosorption by compound bioflocculant:
16 performance and mechanism, Desalination and Water Treatment, 53(2): 421-429.

us
17 Wartelle, L. and Marshall, W., 2000, Citric acid modified agricultural by-products as copper ion
18 adsorbents, Advances in Environmental Research, 4(1): 1-7.
19 Weber, W. J. and Morris, J. C., 1963, Kinetics of adsorption on carbon from solution, Journal of the
20
21
22
Sanitary Engineering Division, 89(2): 31-60. an
Yagub, M. T., Sen, T. K., Afroze, S. and Ang, H., 2014, Dye and its Removal from aqueous solution by
Adsorption: A review, Advances in Colloid and Interface Science, 204: 172-184.
23 Zvinowanda, C. M., Okonkwo, J. O., Agyei, N. M. and Shabalala, P. N., 2009, Physicochemical
M
24 characterization of maize tassel as an adsorbent. I. Surface texture, microstructure, and
25 thermal stability, Journal of applied polymer science, 111(4): 1923-1930.

26
ed

27

28
pt

29
ce

30

31
Ac

32

33

34

35

36

37

38 List of Figure Caption

Page 43 of 68
43

1 Fig. 1 – FTIR spectrum of (a) raw and (b) NaOH treated eucalyptus bark powder

2 Fig. 2 – SEM (BSE) images of (a) raw EB before adsorption (b) raw EB after Zn 2+
3 adsorption (c) NaOH treated EB before adsorption (d) NaOH treated EB after Zn 2+
4 adsorption
5
6 Fig. 3 – X-ray diffraction spectra’s of (a) raw and (b) NaOH treated eucalyptus bark
7 (EB) powder
8 Fig. 4 - Effect of initial solution pH on the adsorption of Zn2+ by EB and NaOH treated
9 EB powder: (Conditions: Mass of Adsorbent = 20 mg, Volume of Zn 2+ Solution = 50 ml,

t
10 Initial Zn2+ Solution Concentration = 20 ppm, Temp = 30°C, Shaker Speed = 120 rpm

ip
11 and Time of Adsorption = 100 min).
12
Fig. 5 - Effect of adsorbent dosages on the adsorption of Zn 2+ onto raw and treated EB

cr
13
14 powder: (Conditions: Volume of Zn2+ Solution = 50 ml, Solution pH = 5.1, Initial Zn2+
15 Concentration = 20 ppm, Temp = 30°C, Shaker Speed = 120 rpm and Time of

us
16 Adsorption = 100 min).
17
18 Fig. 6(a) – Effect of initial metal concentrations and (b) the removal percentage (%) of
19 Zn2+ adsorption by NaOH treated EB powder: (Conditions: Mass of Adsorbent = 20
20
21 = 120 rpm).
an
mg, Volume of Zn2+ Solution = 50 ml, Solution pH =5.1, Temp = 30°C and Shaker Speed

22
Fig. 7 - Effect of temperature on Zn2+ adsorption (a) onto raw EB powder and (b) onto
M
23
24 NaOH modified EB powder: (Conditions: Mass of Adsorbent = 20 mg, Volume of Zn2+
25 Solution = 50 ml, Solution pH = 5.1, Initial Zn2+ Concentration = 20 ppm and Shaker
26 Speed = 120 rpm).
ed

27
28 Fig. 8(a) - Effect of presence of salts on the amount of Zn2+ adsorbed onto NaOH treated
29 EB powder: (Conditions: Mass of Adsorbent = 20 mg, Total Reaction Volume = 50 ml
30 (48 ml of Zn2+ Solution + 2 ml salt solution), Initial Zn2+ Concentration = 20 ppm,
pt

31 Solution pH = 5.1, Temp = 30°C, Shaker Speed = 120 rpm and Time of Adsorption =
32 100 min).
33
ce

34 Fig. 8(b) - Effect of presence of salts on removal percentage (%) of Zn 2+ onto NaOH
35 treated EB powder: (Conditions: Mass of Adsorbent = 20 mg, Total Reaction Volume =
36 50 ml (48 ml of Zn2+ Solution + 2 ml salt solution), Initial Zn2+ Concentration = 20 ppm,
Ac

37 Solution pH = 5.1, Temp = 30°C, Shaker Speed = 120 rpm and Time of Adsorption =
38 100 min)
39
40 Fig. 9 - Intra-particle diffusion model of Zn2+ uptake by NaOH treated eucalyptus bark
41 (Conditions: Mass of Adsorbent = 20 mg, Volume of Zn2+ Solution = 50 ml, Solution pH
42 = 5.1, Temp = 30°C, Adsorption Time = 100 min and Shaker Speed = 120 rpm).
43
44 Fig. 10 - Freundlich plot for raw and NaOH treated EB: Amount of adsorbent 10 mg;
45 Initial Zn2+ Concentration = 20, 30, 40, 50, 60 and 70 ppm, Soluiton pH ≈ 5.1,
46 Temperature = 30◦C; Shaker Speed = 120 rpm and Time of Adsorption = 2 hr.
47
48 Fig. 11(a) - Equilibrium Adsorption Isotherm fitted to the Langmuir-2 Model : Amount
49 of Adsorbent added = 10 mg; initial Zn+2 Concentration = 20, 30, 40, 50, 60, 70 ppm;

Page 44 of 68
44

1 Soluiton pH ≈ 5.1; Temperature = 30°C; Shaker Speed = 120 rpm and Time of
2 Adsorption = 2 hr min.
3

Fig. 11(b) - Equilibrium Adsorption Isotherm fitted to the Langmuir-I Model


4 :
+2
Amount of Adsorbent added = 10 mg; Initial Zn Concentration = 20, 30, 40,
50, 60, 70 ppm; Soluiton pH ≈ 5.1 ; Temperature = 30°C; Shaker Speed = 120
5
rpm and Time of Adsorption = 2 hr.
. 6

t
7 Fig. 12 – Van’t Hoff plot for adsorption of Zn2+ onto raw and NaOH treated EB

ip
8
9

cr
us
an
M
ed
pt
ce
Ac

Page 45 of 68
Figure-1
Click here to download high resolution image

i
cr
us
an
M
ed
pt
ce
Ac

Page 46 of 68
Figure-2
Click here to download high resolution image

i
cr
us
an
M
ed
pt
ce
Ac

Page 47 of 68
Figure-3
Click here to download high resolution image

i
cr
us
an
M
ed
pt
ce
Ac

Page 48 of 68
Figure-4
Click here to download high resolution image

i
cr
us
an
M
ed
pt
ce
Ac

Page 49 of 68
Figure-5
Click here to download high resolution image

i
cr
us
an
M
ed
pt
ce
Ac

Page 50 of 68
Figure-6
Click here to download high resolution image

i
cr
us
an
M
ed
pt
ce
Ac

Page 51 of 68
Figure-7
Click here to download high resolution image

i
cr
us
an
M
ed
pt
ce
Ac

Page 52 of 68
Figure-8 (a)
Click here to download high resolution image

i
cr
us
an
M
ed
pt
ce
Ac

Page 53 of 68
Figure-8 (b)
Click here to download high resolution image

i
cr
us
an
M
ed
pt
ce
Ac

Page 54 of 68
Figure-9
Click here to download high resolution image

i
cr
us
an
M
ed
pt
ce
Ac

Page 55 of 68
Figure-10
Click here to download high resolution image

i
cr
us
an
M
ed
pt
ce
Ac

Page 56 of 68
Figure-11 (a)
Click here to download high resolution image

i
cr
us
an
M
ed
pt
ce
Ac

Page 57 of 68
Figure-11 (b)
Click here to download high resolution image

i
cr
us
an
M
ed
pt
ce
Ac

Page 58 of 68
Figure-12
Click here to download high resolution image

i
cr
us
an
M
ed
pt
ce
Ac

Page 59 of 68
Table-1

Table 1 – Surface characteristic parameters of raw and NaOH treated eucalyptus bark (EB)
adsorbent

Parameters Values

Sample – Raw EB Sample – NaOH treated EB

BET Surface Area (m2/g) 6.55 20.13


Bulk Density (gm/cm3) 0.39 0.27
Pore volume (cm3/g) 0.003 0.011

t
ip
cr
us
an
M
ed
pt
ce
Ac

Page 60 of 68
Table-2

Table 2 – Percentage desorption of Zn2+ from raw and NaOH treated eucalyptus bark (EB) biomass

Raw eucalyptus bark NaOH treated eucalyptus bark


2+
Initial Zn 2+ 2+
Zn Zn
Concentration
Concentration Removal Desorption Concentration Removal Desorption
(mg/L)
adsorbed (%) (%) adsorbed (%) (%)
(mg/L) (mg/L)
02 20.8 40 .1 2 13 .8 9 8.31 8 69 .3 8 19.6

t
03 1 .7 7 39 .2 4 89.6 17 .2 1 57 .3 8 2 .3

ip
04 14 .4 7 36 .1 8 09.4 18 .1 0 45 .2 6 51.
05 17 .8 9 35 .7 9 45.3 7.91 1 39 .4 1 65.0

cr
06 2.0 8 3 .8 0 58.2 21 .2 0 35 .3 4 32.0
07 6.32 3 3 .7 5 34.1 9.32 3 34 .1 8 41.0

us
an
M
ed
pt
ce
Ac

Page 61 of 68
Table-3

Table 3 - Pseudo Second Order Model Parameters for adsorption of Zn2+ on raw and NaOH treated
eucalyptus bark (EB)
System qe (mg/g), K2 qe (mg/g), h
Parameters Experimental (g/mg-min) Calculated (mg/g-min)
R
2
χ2
Adsorbent Dosage (mg) for raw EB
01 4 .9 4 0. 8 45 .6 6 16 .1 6 0.9 9 7 0. 1 14
02 20 .0 6 0. 2 1 20 .0 0 84. 0.9 9 7 0. 0 02
03 10 .7 4 30. 4 10 .7 8 21.4 0.9 9 8 0. 0 01
Adsorbent Dosage (mg) for treated EB

t
01 72 .5 2 0. 3 74 .0 7 13 .9 9 9 .0 2 0. 3 26

ip
02 34 .6 9 0. 4 35 .2 1 97.4 0.9 6 9 0. 0 7
03 17 .5 7 0. 1 7 17 .7 9 84.5 9 .0 7 0. 0 28

cr
Initial Metal Ion Concentration (ppm) for raw EB
02 20 .0 6 0. 2 1 20 .0 0 84. 0.9 9 7 0. 0 02
03 29 .4 3 0. 0 8 29 .5 0 79.6 0.9 9 2 0. 0 02

us
04 36 .1 8 0. 2 8 36 .1 0 36 .5 0 0.9 9 9 0. 0 02

05 4 .7 4 0. 1 1 4 .0 5 21 .6 0 0.9 9 7 0. 1 07
06 50 .6 9 0. 9 50 .2 5 21 .6 5 0.9 9 7 0. 0 38
07 59 .0 6 0.
Initial Metal Ion Concentration (ppm) for treated EB
6
an58 .1 4 19 .0 1 9 .0 3 0. 1 46
M
02 34 .6 9 0. 1 3 35 .2 1 97.4 0.9 6 9 0. 0 7
03 43 .0 4 0. 0 3 43 .6 7 64. 0.9 7 9 0. 0 90
04 45 .2 6 0. 0 5 45 .0 5 10 .6 5 9 .0 3 0. 0 10
05 49 .2 6 0. 0 4 50 .2 5 1 .2 2 0.9 9 2 0. 1 96
ed

06 53 .0 1 0. 0 8 52 .6 3 2 .8 8 9 .0 7 0. 0 27
07 59 .8 2 0. 0 8 59 .5 2 28 .5 7 0.9 9 8 0. 0 15
pH for raw EB
5.2 39. 0. 1 9 10 .0 6 59.7 0.9 9 4 0. 0 17
pt

10.3 12 .3 7 0. 1 1 1 .6 0 6 .9 0.9 5 6 0. 5 10
0.4 17 .5 4 0. 3 5 17 .3 9 12 .5 2 0.9 9 8 0. 0 13
25.4 19 .2 5 0. 2 7 19 .4 9 1 .3 9 0.9 9 9 0. 0 30
ce

1.5 20 .0 6 0. 2 1 20 .0 0 7 .4 0.9 9 7 0. 0 02
pH for treated EB
5.2 19 .4 0 0. 0 2 21 .7 9 21. 0.9 2 4 0.2 6 14
Ac

10.3 2 .5 8 0. 0 2 24 .5 7 43.1 98.0 1 0.1 6 12


0.4 27 .8 9 0. 6 28 .4 9 34. 0.9 9 0 0. 1 26
25.4 31 .8 4 0. 6 32 .7 9 21.6 9 .0 1 0. 2 73
1.5 34 .6 9 0. 4 35 .2 1 97.4 0.9 6 9 0. 0 7
o
Temperature ( C) for raw EB
03 20 .0 6 0. 2 1 20 .0 0 84. 0.9 9 7 0. 0 02
54 18 .6 1 0. 1 6 18 .9 4 87.5 9 .0 6 0. 0 57
06 17 .4 7 10. 8 17 .5 1 83.5 0.9 9 6 0. 0 01
o
Temperature ( C) for treated EB
03 34 .6 9 0. 4 35 .2 1 97.4 0.9 6 9 0. 0 7
54 25 .6 5 0. 1 3 25 .9 1 09.8 9 .0 8 0. 0 25
06 20 .8 4 20. 4 20 .4 1 97. 0.9 9 6 0. 0 91

Page 62 of 68
Ac
ce
pt
ed
M
an
us
cr
ip
t

Page 63 of 68
Table-4

Table 4 – Parameters of Intraparticle Diffusion Model for Zn2+ adsorption on raw and NaOH
treated Eucalyptus Bark (EB) at different concentrations.

System Parameters 20 mg/L 30 mg/L 40 mg/L 50 mg/L 60 mg/L 70 mg/L

Raw K
id 74.1 82. 05.2 1.3 2 37. 3.4 3
eucalyptus
bark C 6.7 4 1.9 3 4.61 6 3.81 1 19 .1 1 9.02 3
NaOH K

t
treated id 8.2 7 06.3 6.3 2 30.4 7.3 3 4. 2

ip
eucalyptus
bark C 7. 5 95. 12 .9 8 14 .2 8 5.12 4 2 .9 5

cr
us
an
M
ed
pt
ce
Ac

Page 64 of 68
Table-5

Table 5 – A summary of Freundlich and Langmuir calculated values

Adsorbent Model Parameters Values


FREUNDLICH Kf ( L/g) 16 .3 1
bf 54.0
2
R 39.0
MP S D 19.
LANGMUIR (TYPE I) qm ( m g/ ) 13 1. 85
Raw eucalyptus Ka ( L/ mg ) 0. 5
2
bark R 9.0 7

t
MP S D 69 .0 1

ip
LANGMUIR (TYPE II) qm ( m g/ ) 12 8. 12
Ka ( L/ mg ) 0. 6
2
R 49.0

cr
MP S D 37 .0 2
FREUNDLICH Kf ( L/g) 38 .2 8
bf 4 .0

us
2
R 9.0 3
MP S D 31.2
LANGMUIR (TYPE I) qm ( m g/ ) 2 2. 2
NaOH treated Ka ( L/ mg ) 01.
eucalyptus bark

LANGMUIR (TYPE II)


R
MP
an
2

qm ( m
S D
g/ )
9 .0
74
25
.4 4
0. 0
Ka ( L/ mg ) 0. 8
M
2
R 79.0
MP S D 53 .3 5
ed
pt
ce
Ac

Page 65 of 68
Table-6

Table 6 - Thermodynamic parameters for adsorption of Zn2+ at different temperatures onto


eucalyptus bark and NaOH modified eucalyptus bark

Adsorbent Temp. (K) ∆G˚(KJ/mole) ∆H˚(KJ/mole) ∆S˚(KJ/mole.K)


30 3. 15 - 2.4 4
Ra w eu c a l y p tus 31 8. 15 - 8.2 5 - 32 .4 3 - 0. 9
ba rk 3 3. 15 - 4.1 5

t
30 3. 15 - 2.1 9

ip
NaO H m od i f i ed 31 8. 15 - 0.1 5 - 1.6 9 - 0. 2
eu c al y pt us ba rk 3 3. 15 - 8.0 1

cr
us
an
M
ed
pt
ce
Ac

Page 66 of 68
Table-7

Table 7 – Comparison of the adsorption capacity (qm in mg/g) of different sorbents for the removal
of Zn(II)

SI No. Adsorbent Adsorption Reference

capacity (mg/g)

1 Okra biomass (modified) 57.11 (A. S. Singha & Guleria,

t
2015)

ip
2 Activated neem bark 11.904 (Maheshwari, Mathesan, &

cr
Gupta, 2015)

us
3 Rapeseed waste 13.859 (Paduraru et al., 2015)

4 Sugarcane bagasse 40.00 (Putra et al., 2014)

5 Coconut tree sawdust 23.81


an (Putra et al., 2014)

6 Orange Peel (Modified by 56.18 (FENG & GUO, 2012)


M
sodium hydroxide and calcium

chloride)
ed

7 Orange Peel (raw) 21.25 (FENG & GUO, 2012)

8 Water melon rind 6.845 (Liu et al., 2012)


pt

10 Citrus reticulata waste 138.42 (Boota et al., 2009)


ce

biomass

11 Moringa oleifera biomass 45.76 (Bhatti et al., 2007)


Ac

(NaOH treated)

12 Neem bark 13.29 (Bhattacharya et al., 2006)

13 Papaya wood 13.45 (Saeed et al., 2005)

14 Raw eucalyptus bark 128.21 Present study

15 Modified eucalyptus bark 250.00 Present study

(NaOH treated)

Page 67 of 68
Ac
ce
pt
ed
M
an
us
cr
ip
t

Page 68 of 68

You might also like