You are on page 1of 49

Journal Pre-proof

Competitive adsorption between Cu2+ and Ni2+ on corn cob activated


carbon and the difference of thermal effects on mono and bicomponent
systems

Natália F. Campos (Conceptualization) (Methodology) (Formal


analysis) (Investigation)<ce:contributor-role>Data Curation) (Writing
- original draft) (Visualization) (Project administration), Giovanna A.
J.C. Guedes (Methodology) (Investigation), Leticia P.S. Oliveira
(Conceptualization) (Methodology) (Formal analysis)
(Visualization), Brı́gida M.V. Gama (Methodology) (Visualization),
Deivson C.S. Sales (Conceptualization) (Methodology) (Software)
(Validation) (Formal analysis) (Writing - review and editing), Joan M.
Rodrı́guez-Dı́az (Validation) (Writing - review and editing), Celmy
M.B.M. Barbosa (Resources) (Funding acquisition), Marta M.M.B.
Duarte (Conceptualization) (Validation) (Resources) (Writing -
review and editing) (Supervision) (Project administration) (Funding
acquisition)

PII: S2213-3437(20)30581-9
DOI: https://doi.org/10.1016/j.jece.2020.104232
Reference: JECE 104232

To appear in: Journal of Environmental Chemical Engineering

Received Date: 8 April 2020


Revised Date: 31 May 2020
Accepted Date: 27 June 2020

Please cite this article as: Campos NF, Guedes GAJC, Oliveira LPS, Gama BMV, Sales DCS,
Rodrı́guez-Dı́az JM, Barbosa CMBM, Duarte MMMB, Competitive adsorption between Cu2+
and Ni2+ on corn cob activated carbon and the difference of thermal effects on mono and
bicomponent systems, Journal of Environmental Chemical Engineering (2020),
doi: https://doi.org/10.1016/j.jece.2020.104232

This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. 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.

© 2020 Published by Elsevier.


Competitive adsorption between Cu2+ and Ni2+ on corn cob activated carbon and

the difference of thermal effects on mono and bicomponent systems

Natália F. Camposa*, Giovanna A. J. C. Guedesa, Leticia P. S. Oliveiraa, Brígida M. V.

Gamaa, Deivson C. S. Salesb, Joan M. Rodríguez-Díazc,d,e, Celmy M. B. M. Barbosaa,

Marta M. M. B. Duartea*

a
Chemical Engineering Department, Federal University of Pernambuco. Avenida Artur

of
de Sá, s/n, 50740-521 Recife, Brazil.
b
Polytechnic School of Pernambuco, University of Pernambuco, Rua Benfica, nº 455,

ro
Madalena, 50720-001, Recife, Brazil.
c -p
Laboratorio de Análisis Químicos y Biotecnológicos. Instituto de Investigación.

Universidad Técnica de Manabí, Av. Urbina y Che Guevara, Portoviejo, Manabí,


re
Ecuador
d
lP

Departamento de Procesos Químicos. Facultad de Ciencias Matemáticas, Físicas y

Químicas. Universidad Técnica de Manabí, Ecuador


e
Programa de Pós-graduação em Engenharia Química. Universidade Federal da Paraíba,
na

58051-900. João Pessoa, Brazil.


ur

*Corresponding Author:
Email address: mmmbduarte@gmail.com (Marta M. M. B. Duarte)
Jo

nataliaferreiracamp@hotmail.com. (Natália F. Campos)

1
Abstract
Generally, a single heavy metal is not discarded into natural waters and wastewaters,
but a mixture of at least two components. When adsorption is the treatment used, the
competition for active sites on surface of adsorbents determines the removal efficiency.
In this work, the adsorption of Cu2+ and Ni2+ considering mono and bicomponent
systems was evaluated using corn cob activated carbon as an adsorbent. Effects of pH,
initial concentration and temperature were investigated. Kinetics and equilibrium of
adsorption studies were performed. The equilibrium was established after 240 min and
100 min for Cu2+ and Ni2+, respectively. The adsorption capacities were 0.39 mmol·g-1
and 0.28 mmol·g-1 for Cu2+ and Ni2+, respectively. High values of pH and initial
concentration favors the adsorption, although Cu2+ adsorption was not affected by Ni2+

of
presence. Thermal effects on the amount adsorbed were not statistically significant on
monocomponent system. For bicomponent system, the temperature was statistically

ro
significant only on Cu2+ adsorption. The work clearly shows that under the investigated
conditions, for monocomponent systems, only aspects related to concentration in
-p
solution and characteristics of the adsorbent influence the mass transfer and removal of
metal. In this case, the adsorption process must be designed as isothermal. On the other
re
hand, for multicomponent systems, the temperature also affects the process, being one
more parameter that need be optimized. This behavior has not been previously reported
lP

in the literature about the adsorption of heavy metals.

Keywords: Adsorption, agro-industrial waste, corn cob, heavy metals, temperature


na

1. Introduction

Several industries such as mining, foundry, surface finishing, electrolysis,


ur

electrical devices, circuits and boards, agricultural, paper and cellulose are sources of
Jo

pollution due to disposal of heavy metals in the environment [1,2]. The industry of

paper and cellulose deserves to be highlighted, since approximately 150 to 200 m³ of

sludge with high content iron (Fe), zinc (Zn), lead (Pb), chromium (Cr), cadmium (Cd),

manganese (Mn), copper (Cu) and nickel (Ni) per ton of paper are produced [3].

2
The simultaneous presence of copper and nickel ions can also be observed in

the effluents of industries of electroplating, mining, agricultural, livestock (manure and

sewage sludge), specialized steel alloys and landfill leachate [4]. The environmental

impact caused by the presence of these metals is serious since they are not

biodegradable accumulating in organisms and being harmful to health [5,6].

Some techniques are being used for treatment of effluent containing heavy

metals such as: ion exchange [7], chemical precipitation [8]; membrane separation [9]

and adsorption [10]. In this context, adsorption is interesting due to simplicity,

of
efficiency, easy operation, possibility of regeneration of adsorbent, minimization of

secondary waste and great availability of adsorbents (the choice of the most suitable

ro
adsorbent is necessary) [2,11].
-p
Several adsorbents such as Saccharomyces cerevisiae/alginate composites beads

[12], rice straw biochar [13], and bentonite-alginate composite [14] have been studied,
re
however, the activated carbon is the adsorbent commonly used for heavy metals
lP

adsorption. Different precursors for activated carbon can be used, however, agro-

industrial wastes are interesting due to low cost and great potential for reuse [15].

Among the precursors are peanut shell [16], grape marc [17] and corn cob [18]. In this
na

context, corn cob is viewed as a waste material with no commercial value, discarded or

burned in open spaces without energetic utility [19]. On the other hand, corn cob is a
ur

good alternative to produce activated carbon since it is renewable, available, and low-
Jo

cost precursor [20]. In addition, the corn cob contains 40 – 45% cellulose, 30 – 35%

hemicellulose, and 10 – 20% lignin [19], with a high carbon content and low ash

content compared to other biomasses [21].

Liu et al. [22] used an activated carbon from corn cob, chemically activated with

KOH, for Hg2+ removal from aqueous solutions obtaining maximum adsorption

3
capacity of 2.39 mg·g-1 for an initial concentration of 100 μg·L-1. Wang et al. [18]

prepared a magnetic activated carbon from corn cob using hydrochloric acid from the

surface treatment of iron and steel for methyl orange dye removal obtaining maximum

adsorption capacity of 595.24 mg·g-1 for an initial concentration of 500 mg·L-1. Sych et

al. [23] evaluated the adsorption capacity of an activated carbon from corn cob,

chemically activated with H3PO4, for copper and methylene blue in monocomponent

systems, obtaining 30% of copper removal and adsorption capacity of 215 mg·g-1 of

methylene blue.

of
Generally, the presence of a single metallic ion (monocomponente) is not

observed in natural waters and wastewaters, but a mixture containing at least two ions

ro
(bicomponent). In this case, the metal ions in the mixture can compete for available
-p
active sites of the adsorbents, leading to synergism, antagonism, or non-interaction [24].

One of the main problems related to multicomponent adsorption is that its behavior
re
depends on the components being treated. In the case of adsorption of heavy metals,
lP

both synergism and antagonism cannot be predicted from specific chemical properties

such as, for example, their solubility in water [14].

In order to further investigate the bicomponent adsorption of metallic ions, this


na

work aims at the competitive adsorption of Cu2+ and Ni2+ on activated carbon from

corn cob precursor. Kinetics and equilibrium of adsorption were investigated and the
ur

effect of temperature on adsorption for mono and bicomponent systems was evaluated
Jo

for equilibrium. Mathematical models were investigated in the prediction of

experimental data for both systems.

2. Materials and Methods

2.1. Preparation of Solutions

4
Stock solutions of concentration 10 mmol·L-1 of copper (Cu2+) and nickel (Ni2+)

were prepared from their respective salts: Cu(NO3)2·3H2O and Ni(NO3)2·6H2O (brand:

VETEC; purity: 99% and 97%). The working solutions were obtained by diluting the

stock solutions. The levels of metals were quantified by flame atomic absorption

spectrometry (brand: VARIAN; model: AA 240 FS – Fast Sequential Atomic

Absorption Spectrometer; λCu = 218.2 nm; λNi = 341.5 nm). The analytical curves

with linear range of 0.05 to 1.25 mmol·L-1 for both ions were built (Limit of Detection

– LODCu = 0.049 mg·L-1; Limit of Quantitation – LOQCu = 0.062 mg·L-1; Coefficient

of
of Determination – R2Cu = 0.9999; Coefficient of Variation – COVCu = 3.33%; LODNi =

0.046 mg·L-1; LOQNi = 0.065 mg·L-1; R2Ni = 0.9974; COVNi = 4.61%). According to

ro
Montgomery [25], R2 > 0.99 and COV < 5% indicates good fit of linear model.

2.2. Preparation of the Activated Carbon


-p
The corn cob precursor was split to facilitate drying, dried at 378 K in an oven
re
(brand: Splabor; model: SP-100A), crushed, washed with distilled water to remove
lP

impurities and dried again in the oven for 1 h at 333 K. The material was chemically

activated using phosphoric acid (H3PO4; brand: VETEC; purity: 85%) in the ratio m: V
na

(mass of precursor:volume of acid) of 5:3 at 386 K for 16 h. After, the carbonization

was performed under the atmosphere generated within the muffle furnace (brand:

Quimis®; model: Q318M21; heating rate: 10 K·min-1) at 773 K for 1 h, by adapting the
ur

methodology proposed by Patnukao and Pavasant [26] and Brito et al. [27].
Jo

The prepared activated carbon was washed with 1% (w/w) sodium bicarbonate

solution (NaHCO3; brand: Fmaia; purity: 99.7%) and distilled water to remove residual

acid until the filtrate reaches pH 6, dried in the oven for 1 h at 333 K and classified in a

series of Tyler sieves with particle sizes of 0.2 – 1.0 mm.

2.3. Characterization of the Materials

5
Thermogravimetric (TG) and differential thermogravimetric (DTG) analyses of

the corn cob were performed using a thermobalance (brand: NETZSCH; model: STA

449 F3 Jupiter; carrier gas: nitrogen; flow rate: 100 mL·min-1; alumina crucibles;

heating rate: 20 K·min-1; mass of material: 6.0 ± 0.5 mg; temperature range: 310 to

1073 K).

Absorption spectra of the corn cob precursor and prepared activated carbon were

obtained by infrared spectrometry (brand: Bruker; model: VERTEX 70v – FTIR-ATR;

ν = 4000 to 500 cm-1; Δν = 2 cm-1).

of
The acidic and basic functional groups of corn cob precursor and activated

carbon were identified by Boehm method as described by Li et al. [28]. Samples of 0.5

ro
g of each material were placed in 50 mL of solutions 0.1 mol·L-1: HCl, NaOH, Na2CO3
-p
and NaHCO3 under 200 rpm for 24 h (in duplicate). The mixture was filtered (brand:

Unifil; diameter: 2 μm) and the solution was titrated (in triplicate) with solutions of 0.1
re
mol·L-1 NaOH and 0.1 mol·L-1 HCl to quantify the concentrations of the acid and the
lP

base, respectively.

The pH of point of zero charge (pHPZC ) was determined for corn cob precursor

and prepared activated carbon using 1.0 g of each material placed in 25 mL of distilled
na

water, varying the pH between 2 and 10 under 300 rpm at 24 h. The pH of the solutions

was adjusted with HNO3 and NaOH (both 0.1 mol·L-1), measured by a pH meter (brand:
ur

Quimis; model Q400AS). The ΔpH (pHinitial − pHfinal ) versus pHinitial graph was built
Jo

and pHPZC was determined according to Pezoti et al. [29].

2.4. Effect of Initial pH of Solution

In the evaluation, 0.1 g of activated carbon was placed in 50 mL of solutions

1.00 mmol·L-1: Cu2+, Ni2+ and Cu2+/Ni2+ under 300 rpm at 303 K for 3 h and pH in

range of 2 to 7. This pH range was based on the chemical speciation diagrams for each

6
ion according to Huang et al. [30] and Razavian et al. [31]. The pH of the solutions was

adjusted with HNO3 and NaOH (both 0.1 mol·L-1), measured by a pH meter (brand:

Quimis; model Q400AS).

2.5. Kinetic of Adsorption for Mono and Bicomponent Systems and the Effect of Initial

Concentration of Solution

Adsorption kinetics tests were performed using 0.1 g of activated carbon with 50

mL of solution of Cu2+ or Ni2+ at the concentrations of 0.50, 1.00 and 1.25 mmol·L-1 for

3 to 480 min time interval under 300 rpm at 303 K. Initially for bicomponent system,

of
the concentration of Cu2+ was fixed at 1.00 mmol·L-1 while Ni2+ varied for 0.50, 1.00

and 1.25 mmol·L-1. After, Ni2+ was fixed at 1.00 mmol·L-1 while Cu2+ varied for 0.50,

ro
1.00 and 1.25 mmol·L-1. The experiments were performed in duplicate.

2.5.1. Modelling of Monocomponent Adsorption Kinetic -p


In the modelling of monocomponent adsorption kinetic were evaluated: pseudo-
re
first order model (PFO, Eq. 1) and pseudo-second order model (PSO, Eq. 2) according
lP

to Dotto et al. [32], and intraparticle diffusion model (Eq. 3) according to Goswami and

Phukan [33].

dqt
na

= k1 (qe − qt ) (1)
dt

where q t [mmol ∙ g −1 ] is the adsorbed amount of ion, t [min] is the time, k1 [min−1] is
ur

the PFO adsorption rate constant and q e [mmol ∙ g −1] is the adsorbed amount of ion at

equilibrium.
Jo

dqt
= k 2 (qe − qt )2 (2)
dt

where k 2 [min−1] is the PSO adsorption rate constant.

1⁄
q t = k id t 2 +c (3)

where k id is the intraparticle diffusion rate constant [mg ∙ g-1 ∙ min-1/2] and c is the
constant related to diffusion resistance [mg ∙ g-1].

7
2.5.2. Modelling of Bicomponent Adsorption Kinetic

Initially, in the modelling of bicomponent adsorption kinetic, PFO (Eq. 1) and

PSO (Eq. 2) models were tested to predict the behavior of adsorbed amount. In these

cases, although in the bicomponent system there is mutual influence between the ions,

the models do not consider this interaction and the fit to experimental data uses only

independent data from each.

In order to predict bicomponent adsorption kinetic considering interaction

between ions, a model was proposed (PM) as presented in Eq. 4.

of
dq t,i
= ra,i − rd,i (4)
dt

ro
where ra,i [mmol ∙ g −1 ∙ min−1] is the adsorption rate and rd,i [mmol ∙ g −1 ∙ min−1] is

the desorption rate. The adsorption and desorption rates are determined by Eq. 5 and 6,

respectively.
-p
re
ni
2

ra,i = k a,i Ci (1 − ∑ θj ) (5)


j=1
lP

n
rd,i = k d,i θi i (6)
na

where k a,i [L ∙ g −1 ∙ min−1] is the adsorption kinetic constant for ion i (i = 1 for Cu2+

and i = 2 for Ni2+), k d,i [mmol ∙ g −1 ∙ min−1] is the desorption kinetic constant for ion
ur

i, Ci [mmol ∙ L−1] is the concentration in solution for ion i, θi [dimensionless] is


Jo

fractional surface coverage for ion i, θj [dimensionless] is fractional surface coverage

for ion j (j = 1 for Cu2+; j = 2 for Ni2+) and ni is the heterogeneity parameter of ion i.

The fractional surface coverages are determined by θi = q t,i /q m,i and θj = q t,j /q m,j,

where q m [mmol ∙ g −1 ] is the maximum adsorbed amount. The ion concentration in

solution is related to the amount adsorbed by Eq. 7.

8
m
Ci = C0,i − q (7)
V i

where C0,i [mmol ∙ L−1] is the initial concentration in solution for ion i, m [g] is the

mass of adsorbent and V [L] is the volume of solution. Thus, Eq. 3 can be rewritten,

resulting in Eq. 8.
ni
2 ni
dq t,i m qj qj
= k a,i [C0,i − q i ] (1 − ∑ ) − k d,i ( ) (8)
dt V q m,j q m,i
j=1

The unknown parameters (k1 , k 2 , qe , k id , c, qm, k a , k d and n) were obtained by

of
fit of the experimental data to the models using the MATLAB R2019a (MathWorks

Inc., Natick, Massachusetts, USA) by minimizing the objective function (fobj ): fobj =

ro
2
∑N
j=1(q exp,j − q mod,j ) where j [dimensionless] is an index; N [dimensionless] is the

-p
number of data points, q exp,j [mmol·g-1] is the experimental adsorbed amount for j and

q mod,j [mmol·g-1] is the adsorbed amount obtained from model for j. The 4th-order
re
Runge-Kutta method was used to solve the equations. The R2 and reduced chi-square
lP

(χ2ν ) were determined.

2.6. Equilibrium of Adsorption for Mono and Bicomponent Systems


na

The equilibrium of adsorption for mono (Cu2+ or Ni2+) and bicomponent (Cu2+

and Ni2+) systems was evaluated using 0.1 g of activated carbon with 50 mL of solution
ur

for an initial concentration in the range of 0.10 to 2.00 mmol·L-1 by adapting the

methodology proposed by Demiral and Güngör [17]. For bicomponent system were
Jo

used the same initial concentrations (ratio 1:1). The samples were filtered after 16 h. All

experiments were performed at 293 K, 303 K, 313 K and 323 K in duplicate.

2.6.1. Modelling of Monocomponent Adsorption Equilibrium

9
In the modelling of the monocomponent adsorption equilibrium isotherms of

Cu2+ and Ni2+ ions, the models: Langmuir (Eq. 9) [34]; Freundlich (Eq. 10) [5]; and

Sips (Langmuir-Freundlich, Eq. 11) [35]; were evaluated.

q mL K L Ce
qe = (9)
1 + K L Ce

where q e [mmol ∙ g −1] is the adsorbed amount of ion at equilibrium, Ce [mmol ∙ L−1] is

the ion concentration in solution at equilibrium, q mL [mmol ∙ g −1 ] is the maximum

adsorbed amount of ion at equilibrium for Langmuir model and K L [L ∙ mmol−1] is the

of
Langmuir equilibrium constant.
1
q e = K F (Ce )nF (10)

ro
where K F [mmol1−1/nF ∙ g −1 ∙ L1/nF ] is the Freundlich equilibrium constant and nF

[dimensionless] is the Freundlich intensity parameter.


1
-p
re
q mS (K S Ce )nS
qe = 1
(11)
1 + (K S Ce )nS
lP

where q mS [mmol ∙ g −1 ] is the maximum adsorbed amount of ion at equilibrium for

Sips model, K S [L ∙ mmol−1 ] is the Sips equilibrium constant and nS [dimensionless] is


na

the Sips heterogeneity parameter.

The unknown parameters (q mL, K L , K F , nF , q mS , K S and nS ) were obtained by fit


ur

of the experimental data to the models by nonlinear curve fitting using the Origin® 2018

(OriginLab Cooperation, Northhampton, Massachusetts, USA). The R2 and reduced chi-


Jo

square (χ2ν ) were determined.

2.6.2. Modelling of Bicomponent Adsorption Equilibrium

Considering the modelling of the bicomponent adsorption equilibrium isotherms,

the models: non-modified Langmuir (Eq. 12) [36]; extended Langmuir (Eq. 13) [37];

10
modified Langmuir (Eq. 14) [29]; non-modified Sips (Eq. 15) [36]; and extended Sips

(Eq. 16) [39]; were evaluated.

q mL,i K L,i Ce,i


q e,i = (12)
1 + ∑Nj=1 K L,j Ce,j

where q e,i [mmol ∙ g −1 ] is the adsorbed amount of ion i (i = 1 for Ni2+ and i = 2 for

Cu2+) at equilibrium, Ce,i [mmol ∙ L−1] is the ion i concentration in solution at

equilibrium, q mL,i [mmol ∙ g −1 ] is the maximum adsorbed amount of ion i at

equilibrium for Langmuir model, K L,i [L ∙ mmol−1] is the Langmuir equilibrium

of
constant of ion i (obtained from monocomponent system), K L,j is the monocomponent

Langmuir equilibrium constant of ion j (j = 1 for Ni2+; j = 2 for Cu2+; obtained from

ro
monocomponent system) and N [dimensionless] is the number of ions (N = 2).

q e,i =
q mEL K EL,i Ce,i
1 + ∑Nj=1 K EL,j Ce,j
-p (13)
re
where q mEL [mmol ∙ g −1 ] is the maximum adsorbed amount for Langmuir model (both

ions together) and K EL,i [L ∙ mmol−1 ] is the Langmuir equilibrium constant of ion i
lP

(obtained from simultaneous fit of the equations to bicomponent system) and K EL,j [L ∙

mmol−1] is the Langmuir equilibrium constant of ion j (obtained from simultaneous fit
na

of the equations to bicomponent system).

q mL,i K L,i (Ce,i /ηL,i )


ur

q e,i = (14)
1 + ∑Nj=1 K L,j (Ce,j /ηL,j )

where ηL,i [dimensionless] is the interaction coefficient of ion i and ηL,j [dimensionless]
Jo

is the interaction coefficient of ion j.


1
nS,i
q mS,i K S,i Ce,i
q e,i = 1 (15)
n S,j
1+ ∑N
j=1 K S,j Ce,j

11
where q mS,i [mmol ∙ g −1 ] is the maximum adsorbed amount of ion i at equilibrium for

Sips model, K S,i [L ∙ mmol−1 ] is the Sips equilibrium constant of ion i (obtained from

monocomponent system), K S,j [L ∙ mmol−1 ] is the Sips equilibrium constant of ion j

(obtained from monocomponent system), nS,i [dimensionless] is the Sips heterogeneity

parameter of ion i (obtained from monocomponent system) and nS,j [dimensionless] is

the Sips heterogeneity parameter of ion 𝑗 (obtained from monocomponent system).


nES −1
q mES K ES,i Ce,i (∑ K ES,j Ce,j )
q e,i = nES (16)
1 + (∑ K ES,j Ce,j )

of
where q mES [mmol ∙ g −1 ] is the maximum adsorbed amount for extended Sips model

ro
(both ions together), K ES,i [L ∙ mmol−1 ] is the extended Sips equilibrium constant of ion

i (obtained from bicomponent system), K ES,i [L ∙ mmol−1 ] is the extended Sips


-p
equilibrium constant of ion i (obtained from monocomponent system), K ES,j [L ∙
re
mmol−1] is the extended Sips equilibrium constant of ion j (obtained from

monocomponent system), nES [dimensionless] is the Sips heterogeneity parameter.


lP

The unknown parameters (q mL,i, K L,j , q mEL,i, K EL,j , ηL,j , q mS,i, K S,j, nS,j, q mES ,

K ES,j and nES ) were obtained by fit of the experimental data to the models by nonlinear
na

curve fitting using the Origin® 2018 (OriginLab Cooperation, Northhampton,

Massachusetts, USA). The R2 and reduced chi-square (χ2ν ) was determined.


ur

3. Results and discussions

3.1. Characterization of the Materials


Jo

In Fig. 1 are shown the results of TG and DTG analyses for corn cob precursor.

12
110
15
100
90 10

Derivative mass (%/K)


80 5
70
Mass (%)

0
60
-5
50
-10
40
-15
30
20 -20

10 -25

of
273 373 473 573 673 773 873 973 1073 1173
T (K)

ro
Fig. 1. TG and DTG analyses for corn cob precursor.

According to Fig. 1, the corn cob presented three stages of thermal


-p
decomposition. The first mass loss (below 523 K) is related to desorption of physically

adsorbed water [40]. The second mass loss (between 523 K and 673 K) is related to
re
degradation of cellulose, hemicellulose, and some lignin [41]. The double peak
lP

observed around 600 K (from DTG analysis) is attributed to higher temperature of

cellulose degradation compared to that of hemicellulose [42]. The third loss of mass is
na

due to degradation of more stable compounds such as lignin with aromatic structures

[20]. In this context, the corn cob is indicated as a good precursor for production of

activated carbons, because 83% of its mass is composed of cellulose, hemicellulose, and
ur

lignin [21]. Due to low mass loss after 673 K, Zhu et al. [43] established that
Jo

temperatures above this value are indicated for the carbonization of corn cob.

The spectra of FTIR-ATR for activated carbon and corn cob precursor are

shown in Fig. 2.

13
1.00

0.95

Transmittance
0.90

0.85

Corn cob
0.80
Activated carbon
4000 3500 3000 2500 2000 1500 1000 500
-1
 (cm )

Fig. 2. Spectra of FTIR-ATR for activated carbon and corn cob precursor.

of
For corn cob precursor was observed a band at 3333 cm−1 related to presence of

bond -OH of alcohols and phenols groups [35], while bands at 2924 cm−1 and 2855

ro
cm−1 are assigned to bond C-H of alkanes groups [16]. Bands at 1738 cm−1 and 1633
-p
cm−1 are assigned to stretching of bond C=O of carbonyl groups, and the bands at 1242

cm−1 and 1032 cm−1 to bond C-H of heterocyclic rings [44].


re
After activation, a band was observed at 2362 cm−1 related to stretching of bond
lP

C≡C [16]. The band at 1567 cm−1 is assigned to bond C=O of carboxyl groups [44] and

at 1163 cm−1 is assigned to stretching of bond C-O alcohols groups [35]. The presence
na

of these functional groups favors the adsorption of metallic ions.

The activation of the corn cob resulted in a decrease in phenolic groups and an
ur

increase in carboxylic groups on surface. In addition, for activated carbon the number of

carboxylic groups is greater than the phenolic and lactonic groups. When carboxylic
Jo

groups are present in high concentrations on surface, in comparison with the other

functional groups, there is an improvement in the adsorption process, especially if the

pH of solution is acidic, which favors the protonation of these groups as observed by

Liu et al. [45]. Yin et al. [46] produced activated carbon from Trapa natans husk by

chemical activation using H3PO4 observing an increase in the number of carboxylic

14
groups on surface when compared to other functional groups, favoring the adsorption

process.

The values of pHPZC were 4.7 and 3.0 for corn cob and activated carbon,

respectively, indicating a decrease due to activation using H3PO4. This effect can be

related to increase in the number of functional groups such as carboxylics on surface

observed by FTIR-ATR and Boehm method. The decrease in pHPZC allows an increase

in the pH range of solution for which more negative charges are distributed on the

surface of activated carbon favoring the interaction with cations (Cu2+ and Ni2+).

of
3.2. Effect of Initial pH of Solution

The influence of pH (range 2 to 7) in solution on adsorption of Cu2+ and Ni2+ for

ro
mono and bicomponent systems is shown in Fig. 3. Two conditions were evaluated: first
-p
with solutions containing only ions (Cu2+ and Ni2+); after with mixtures containing the

ions and the adsorbent (ads + ion).


re
lP
na
ur
Jo

15
0.30
(a) Cu
2+

0.25

0.20

q (mmol.g )
-1
0.15

0.10

0.05

0.00
2 3 4 5 6 7
pH
0.30

of
(b) Ni
2+

0.25

ro
0.20
q (mmol.g )
-1

0.15

0.10
-p
0.05
re
0.00
2 3 4 5 6 7
lP

pH
0.30
2+
(c) Cu
2+
0.25 Ni
na

0.20
q (mmol.g )
-1

0.15
ur

0.10

0.05
Jo

0.00
2 3 4 5 6 7
pH
Fig. 3. Evaluation of pH of solution for adsorption of Cu2+ and Ni2+ on corn cob

activated carbon: (a) Cu2+, (b) Ni2+ and (c) Cu2+/Ni2+. Conditions: mass of adsorbent =

0.1 g; particle size = 0.2 – 1.0 mm; volume of solution = 50 mL; initial concentration:

1.00 mmol·L-1; stirring speed = 300 rpm; temperature = 303 K; time = 3 h.

16
According to Fig. 3a and 3b, in monocomponent systems no significant change of

adsorptive capacity associated to varying of pH was observed for both ions until pH 7,

except for pH 2. However, precipitation occurred after pH 6 for the bicomponent

system, resulting in a significant decrease in adsorption capacity (Fig. 3c). The

variations in adsorption capacity observed for low pH (between 2 and 3) were possibly

due to protonation of excess of H+ in the solution. These ions compete with Cu2+ and

Ni2+ for available active sites on surface of adsorbent. Due to the increase in pH, the

amount of H+ in solution is reduced, decreasing competition for active sites, and

of
providing increase of access for metallic ions, promoting an increase in the removal of

Cu2+ and Ni2+ ions. Moreover, the highest adsorption capacities were observed at pH 4

ro
for monocomponent and pH 3 for bicomponent, therefore being chosen for use in the

other evaluations. -p
3.3. Kinetic of Adsorption for Mono and Bicomponent Systems and Effect of Initial
re
Concentration of Solution

In Fig. 4 is shown the kinetic of adsorption for Cu2+ and Ni2+ considering initial
lP

concentrations of 0.50, 1.00 and 1.25 mmol ∙ g −1 .

0.45
na

(a)
0.40

0.35

0.30
ur q (mmol.g )
-1

0.25

0.20
Jo

0.15

0.10 2+ -1
Cu (0.50 mmol.L )
2+ -1
0.05 Cu (1.00 mmol.L )
2+ -1
Cu (1.25 mmol.L )
0.00
0 100 200 300 400 500
t (min)

17
0.45
(b)
0.40

0.35

0.30

q (mmol.g )
-1
0.25

0.20

0.15

0.10 2+ -1
Ni (0.50 mmol.L )
2+ -1
0.05 Ni (1.00 mmol.L )
2+ -1
Ni (1.25 mmol.L )
0.00
0 100 200 300 400 500
t (min)

of
0.45
(c)
0.40 2+ 2+
Cu : Ni
2+
0.35 Cu (1.00:0.50)

ro
2+
Cu (1.00:1.00)
0.30 2+
Cu (1.00:1.25)
q (mmol.g )
-1

2+
0.25 Cu (0.50:1.00)
0.20

0.15
-p 2+
Cu (0.50:1.25)
2+
Ni (1.00:0.50)
2+
Ni (1.00:1.00)
2+
re
0.10 Ni (1.00:1.25)
2+
Ni (0.50:1.00)
0.05 2+
Ni (0.50:1.25)
0.00
lP

0 100 200 300 400 500 600 700 800 900


t (min)
Fig. 4. Kinetic of adsorption for Cu and Ni2+: (a) Cu2+, (b) Ni2+ and (c) Cu2+/Ni2+.
2+

Conditions: mass of adsorbent = 0.1 g; particle size = 0.2 – 1.0 mm; volume of solution
na

= 50 mL; initial concentration: 0.50 – 1.25 mmol·L-1; stirring speed = 300 rpm;

temperature = 303 K; time = 0 – 480 min.


ur

According to Figs. 4a and 4b, most of removal of Cu2+ and Ni2+ considering
Jo

monocomponent system occurred for t < 50 min. The equilibrium was established after

240 min and 100 min for Cu2+ and Ni2+, respectively. Similar behavior was observed by

Wang et al. [47] evaluating the adsorption of Cu2+ and Ni2+ on willow wood biochar.

For bicomponent system, the equilibrium time was longer compared to

monocomponent, over 350 min. The interaction between ions had effect on adsorption

18
each one. Initially, both ions presented rapid adsorption, due to the high availability of

active sites. After 100 min, the adsorbed amount of Ni2+ decreased while for Cu2+

increased due to a greater affinity of the adsorbent for Cu2+ compared to Ni2+. This

behavior indicates the Ni2+ adsorbed desorbs and Cu2+ occupies the active site.

According to Mahdi et al. [48], this behavior is due to bond strength between Cu2+ and

functional groups on the surface of the adsorbent, such as carboxylics. In addition, Ni2+

ions with smaller ionic radius (0.69 Å) compared to Cu2+ (0.73 Å) presents higher

hydration energy forming aquocomplexes with large residence time of water molecules

of
in the first hydration shell interfering the interaction with the adsorbent.

In Table 1 are shown the parameters obtained from fit of PFO, PSO and

ro
intraparticle diffusion monocomponent models to experimental data for Cu2+ and Ni2+

kinetics of adsorption on the activated carbon. -p


According to Table 1, PSO presented best fit to experimental data (R2 ≥ 0.618
re
and χ2ν ≤ 2 ∙ 10−3) compared to PFO (R2 ≥ 0.585 and χ2ν ≤ 2 ∙ 10−3 ). The values of q e
lP

obtained by the PSO model were closer to the experimental data (q exp ) for both Cu2+

and Ni2+. Feng and Aldrich [49] attribute this behavior to the fact that both Cu2+ and

Ni2+ interacts with two different active sites on surface: two carboxylic groups, two
na

phenolic groups or one carboxylic and one phenolic, because the ions are divalent as

predicted by PSO model.


ur

For bicomponent system (Table 2), PM presented best fit to experimental data
Jo

(R2 ≥ 0.567 and χ2ν ≤ 9·10-5) compared to PFO (R2 ≥ 0.585 and χ2ν ≤ 1·10-2) and PSO

(R2 ≥ 0.536 and χ2ν ≤ 6·10-4) for both Cu2+ and Ni2+. The PSO model (R2 ≥ 0.892);

χ2ν ≤: 1 ∙ 10−4 ) presented best fit compared to PFO (R2 ≥ 0.801); χ2ν ≤ 1 ∙ 10−3) for

Cu2+ obtaining values of q e close to q exp , similar to observed for monocomponent

system, indicating the low influence of Ni2+ on Cu2+ adsorption. The opposite of

19
previous behavior was observed for Ni2+ adsorption in which Cu2+ had influence. A bad

fit of PFO and PSO models to experimental data of Ni2+ to fact that these models nor

consider competition and desorption due to affinity difference between ions. On the

other hand, according to PM model the increase of Cu2+ in solution resulted in

increasing qm values for Cu2+ and establishment for Ni2+ (Cu2+ has not affected by Ni2+

in solution). In addition, k a,Cu2+ > k a,Ni2+ and k d,Cu2+ > k d,Ni2+ indicating that the

intensity of adsorption and desorption for Cu2+ is greater than for Ni2+. Values of

K Cu2+ ≈ K Ni2+ indicated similar behaviors towards equilibrium and n ≠ 1 established

of
heterogeneity on adsorption of both ions.

ro
-p
re
lP
na
ur
Jo

20
In Fig. 5 are shown the fit of experimental data to intraparticle diffusion model
for monocomponent systems.

0.45
(a)
0.40

0.35

0.30
q (mmol.g )
-1

0.25

0.20

0.15
2+ -1
0.10 Cu (0.50 mmol.L )

of
2+ -1
Cu (1.00 mmol.L )
0.05 2+ -1
Cu (1.25 mmol.L )
0.00

ro
0 5 10 15 20 25
1/2 1/2
t (min )
0.45

0.40
(b)
-p
re
0.35

0.30
q (mmol.g )

lP
-1

0.25

0.20

0.15
na

2+ -1
0.10 Ni (0.50 mmol.L )
2+ -1
Ni (1.00 mmol.L )
0.05 2+ -1
Ni (1.25 mmol.L )
0.00
ur

0 5 10 15 20 25
1/2 1/2
t (min )
Fig. 5. Kinetics of adsorption for Cu2+ and Ni2+. Intraparticle diffusion model for
Jo

monocomponent systems: (a) Cu2+ and (b) Ni2+. Conditions: mass of adsorbent = 0.1 g,
particle size = 0.2 – 1.0 mm, volume of solution = 50 mL, stirring speed = 300 rpm,
temperature = 303 K.

21
According to Fig. 5, good fits of experimental data to intraparticle model were

observed for first step (R21 ≥ 0.898 and χ2ν,1 ≤ 14.95) and moderate were observed for

second step (R22 ≥ 0.550 and χ2ν,2 ≤ 8.61). In case of Ni2+ (1.00 mmol·L-1), due to high

data dispersion was not possible accurately establish what happened in these steps,

however, observing the behavior in the other concentrations (0.50 and 1.25 mmol·L-1),

the same mechanisms can be considered. Since none of the lines went through the

origin, the intra-particle diffusion was not only the rate-controlling step during the

adsorption process Neris et al. [35]. Different linear fit indicated the existence of

of
different diffusion mechanisms. The first step (from 1.7 min1/2 to 9.5 min1/2 for Cu2+ and

ro
4.5 min1/2 for Ni2+) was related to the internal transport through the diffusion of the

molecules to the most internal adsorption sites of the adsorbent. After 90 min for Cu 2+
-p
and 20 min for Ni2+, the adsorption rate decreased toward equilibrium, in which the

intraparticle diffusion began to decline (k d1 > k d2 ) due to lower availability of sites for
re
adsorption and decreased concentration of adsorbate in the solution according to Kumar
lP

and Porkodi [50] and Goswami and Phukan [33].

3.4. Evaluation of Adsorption Equilibrium for Mono and Bicomponent Systems

In Fig. 6 is shown the thermal effect on equilibrium of adsorption by Cu2+ and


na

Ni2+ isotherms. According to Giles et al. [51], the isotherms are type H subtype 2,
ur

indicating high affinity to activated carbon for both metals in the evaluated

concentration range. The adsorption capacities were 0.39 mmol·g-1 and 0.28 mmol·g-1
Jo

for Cu2+ (Fig. 6a) and Ni2+ (Fig. 6b), respectively.

22
0.45
(a)
0.40

0.35

0.30

q (mmol.g )
-1 0.25

0.20

0.15 2+
Cu 293 K
2+
Cu 303 K
0.10 2+
Cu 313 K
2+
0.05 Cu 323 K
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
-1
C (mmol.L )

of
0.45
(b)
0.40

ro
0.35

0.30
q (mmol.g )
-1

0.25

0.20
-p
0.15 2+
re
Ni 293 K
2+
0.10 Ni 303 K
2+
Ni 313 K
0.05 2+
Ni 323 K
lP

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8
-1
C (mmol.L )
Fig. 6. Adsorption isotherms for monocomponent system: (a) Cu2+ and (b) Ni2+.
na

Conditions: mass of adsorbent = 0.1 g; particle size = 0.2 – 1.0 mm; volume of solution

= 50 mL; initial concentration: 0.10 – 2.00 mmol·L-1; stirring speed = 300 rpm;
ur

temperature = 293 – 323 K; time = 16 h.

Similar results were reported by Lee, Park and Chung [52], evaluating the Cu2+
Jo

adsorption on peanut shell biochar, whose adsorption capacity was 0.164 mmol·g-1 for

concentration of 7.87 mmol·L-1 at 298 K. Neris et al. [35] using water hyacinth fibers

modified with NaOH to remove Ni2+ reached an adsorption capacity of 0.21 mmol·g-1

for concentration of 1.7 mmol·L-1 at 298 K.

23
The results also show low variation among the adsorbed amount related to

increase of temperature, mainly at lower concentrations. This behavior is according to

Wei et al. [53], who evaluated the influence of temperature on adsorption microalgae

biomass of on cationic starch.

In order to evaluate the significance of this variation of temperature, a one-way

ANOVA test was performed, and the result is shown in Table 3.

The high p-value (close to 1.0) indicates that the difference among the means of

the adsorbed amounts at each temperature are not significant considering 95% of

of
confidence level. The p-value (Ni2+) > p-value (Cu2+) indicates greater proximity among

the means of Ni2+, compared to the means of Cu2+. In Fig. 7 is shown the box plot of

ro
Cu2+ and Ni2+ adsorption at 293, 303, 313 and 323 K. The medians and variability are

similar, comparing all temperatures. -p


re
lP
na

(a)
ur
Jo

(b)

24
Fig. 7. Box plot for monocomponent adsorption at 293, 303, 313 and 323 K: (a) Cu2+

and (b) Ni2+. Conditions: mass of adsorbent = 0.1 g; particle size = 0.2 – 1.0 mm;

volume of solution = 50 mL; initial concentration: 0.10 – 2.00 mmol·L-1; stirring speed

= 300 rpm; time = 16 h.

According to the statistical evaluation, the temperature had no significant effect

on the amount adsorbed for monocomponent system. Thus, the isotherm at room

temperature (303 K) was used for modelling. In Table 4 are shown the values of

parameters obtained from fit of models to monocomponent adsorption equilibrium

of
isotherm at 303 K.

Based on Table 4, Sips model presented best fit to experimental data (greater R2

ro
and lower χ2ν ) compared to other models, indicating that the active sites of adsorbent
-p
have different energies. The values of nS far from 1.0 indicate that the adsorption

occurred on a heterogeneous surface. This behavior is observed also by good fit of


re
Freundlich model. Similar order of magnitude (nS = 0.185) was observed by Deniz and
lP

Ersanli [54] investigating Cu2+ adsorption on natural macroalgae. Wang et al. [47] also

observed similar order of magnitude (nS = 0.7) for adsorption of Ni2+ on willow wood.
na

On the other hand, the consideration of the Langmuir model to represent the process,

even if disregarding the heterogeneity of the surface, would not result in an inadequate

prediction of the equilibrium behavior.


ur

The activated carbons produced from several biomasses reported in the literature
Jo

for Cu2+ and Ni2+ removal by using a batch adsorber, as well as those used in this work,

are shown in Table 5.

According to Table 5, the activated carbon from corn cob chemically activated

with H3PO4 produced in this work presented better performance in adsorption of Cu2+

and Ni2+, when compared to studies reported in the literature. Although the initial

25
concentration used in this work is lower in relation to the works developed by Bohli et

al [55], Gao et al. [57] and Gupta et al. [59], q m values were close to or higher than

those observed by these authors. Despite the adsorption capacity obtained by Hasar [60]

being 1.44 times greater than that obtained in this work, this adsorption capacity is

related to an initial concentration 2.13 times greater.

0.45
2+
0.40 Cu 293 K
2+
Cu 303 K
0.35 2+
Cu 313 K
2+
0.30 Cu 323 K

of
2+
Ni 293 K
q (mmol.g )
-1

0.25 2+
Ni 303 K
2+
0.20 Ni 313 K
2+
N1 323 K

ro
0.15

0.10

0.05

0.00
-p
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4
-1
C (mmol.L )
re
Fig. 8. Adsorption isotherms for bicomponent system (Cu2+/Ni2+). Conditions: mass of

adsorbent = 0.1 g; particle size = 0.2 – 1.0 mm; volume of solution = 50 mL; initial
lP

concentration: 0.10 – 2.00 mmol·L-1; stirring speed = 300 rpm; temperature = 293 – 323

K; time = 16 h.
na

Considering the bicomponent system (Cu2+/Ni2+), the thermal effect on

equilibrium of adsorption is show in Fig. 8. The increase of Cu2+ in solution decrease


ur

the adsorption of Ni2+. According to Neris et al. [36], the ratio between the adsorption

capacity of Cu2+ in bicomponent (q bi ) and monocomponent (q mono ) system is q bi /


Jo

q mono = 0.9, however, for Ni2+ is q bi /q mono = 0.1, which highlights the antagonistic

effect between ions. According to Silva Correia et al. [61], in cases where this ratio is

close to 1.0, the reduction in adsorption is not significant as observed for Cu2+,

indicating that the presence of Ni2+ has no effect on the process.

26
In the bicomponent system, as previously in the monocomponent, a low variation

of adsorbed amount due to temperature increase was observed (Fig. 8). A one-way

ANOVA test for each ion individually was performed and the result is shown in Table

6.

The low p-value for Cu2+ indicated significant difference among means

considering 95% of confidence level, the opposite of what was observed for Ni2+. In

Fig. 9 is shown the box plot of Cu2+/Ni2+ system.

of
ro
-p
re
(a)
lP
na
ur

(b)
Jo

Fig. 9. Box plot for bicomponent system at 293, 303, 313 and 323 K: (a) Cu2+ and (b)

Ni2+. Conditions: mass of adsorbent = 0.1 g; particle size = 0.2 – 1.0 mm; volume of

solution = 50 mL; initial concentration: 0.10 – 2.00 mmol·L-1; stirring speed = 300 rpm;

time = 16 h.

27
The medians and variability are not similar for Ni2+, comparing all temperatures.

The temperature presented effect on bicomponent adsorption and cannot be neglected.

The variation of temperature affects the Cu2+ adsorption, due to the presence of Ni2+,

since in monocomponent system (only Cu2+) the temperature effect was not significant.

In this case, considering that the effects of the two ions are combined in the adsorption

of each one, the modeling of the adsorption isotherms was performed for both ions and

all temperatures. In Table 7 are shown the values of parameters obtained from fit of

models to bicomponent adsorption equilibrium isotherms at 293, 303, 313 and 323 K.

of
According to Table 7, non-modified models do not present good fit to

experimental data ( R2 ≤ 0.877 and χ2ν ≥ 0.418) due to the fact of not considering

ro
interaction between the ions. Modified Langmuir model presented good fit ( R2 ≥ 0.942
-p
and χ2ν ≤ 9 ∙10-4) by considering these interactions, furthermore ηL,Cu < ηL,Ni indicates

high affinity of Cu2+ to adsorbent compared to Ni2+ [36]. In case of Langmuir and Sips
re
extended models ( R2 ≥ 0.954 and χ2ν ≥ 7 ∙10-4), best fits were obtained and K Cu2+ >
lP

K Ni2+ also indicates high affinity of Cu2+ to adsorbent. Extended Langmuir and Sips

models presented best fit to experimental data because they use parameters that allow
na

evaluate the competition of ions for active sites.

4. Conclusion
ur

The effects of initial concentration, pH, and temperature on Cu2+ and Ni2+

removal using a fixed bed adsorber were evaluated for monocomponent and
Jo

bicomponent systems. The results indicated best removal for higher initial

concentrations, and appropriate pH of 4 and 3 for mono and bicomponent systems,

respectively. The variation of temperature does not affect the adsorption for

monocomponent system; thus, the adsorption process must be designed as isothermal.

Opposite behavior was observed in the effect of temperature on the bicomponent

28
system. The adsorption of Cu2+ was not affected by Ni2+ and antagonistic effect for Ni2+

in the presence of Cu2+ was verified. PSO model predicted better the kinetic evolution

for monocomponent system because it was developed to bivalent metals adsorption.

Sips model predicted better the adsorption equilibrium for monocomponent system

because of heterogeneity and saturation of the adsorbent. The kinetics model proposed

in present work (considers interaction between ions) and extended equilibrium models

(considers competition between ions) presented best prediction for bicomponent system.

This work presented a way of using an abundant residue with no commercial value for

of
the preparation of adsorbent used in metal removal.

ro
Credit Author Statement
-p
Natália F. Campos: Conceptualization, Methodology, Formal analysis, Investigation,

Data Curation, Writing - Original Draft, Visualization, Project administration.


re
Giovanna A. J. C. Guedes: Methodology, Investigation. Leticia P. S. Oliveira:
lP

Conceptualization, Methodology, Formal analysis, Visualization. Brígida M. V. Gama:

Methodology, Visualization. Deivson C. S. Sales: Conceptualization, Methodology,

Software, Validation, Formal analysis, Writing - Review & Editing. Joan M.


na

Rodríguez-Díaz: Validation, Writing - Review & Editing. Celmy M. B. M. Barbosa:

Resources, Funding acquisition. Marta M. M. B. Duarte: Conceptualization, Validation,


ur

Resources, Writing - Review & Editing, Supervision, Project administration, Funding


Jo

acquisition.

Declaration of interests
The authors declare that they have no known competing financial interests or personal relationships that
could have appeared to influence the work reported in this paper.

Acknowledgements

29
This study was partially funded by CAPES – Brazil – Financing Code 001. The

authors are grateful to NUQAAPE/FACEPE, FADE-UFPE and PIBIC/UFPE/CNPq for

their support in the several phases of the research project.

References

[1] J. Wang, C. Chen, Biosorbents for heavy metals removal and their future,

Biotechnol. Adv. 27 (2009) 195-226. https://doi.org/10.1016/j.biotechadv.2008.11.002.

[2] J.K. Ratan, M. Kaur, B. Adiraju, Synthesis of activated carbon from agricultural

of
waste using a simple method: characterization, parametric and isotherms study, Mater.

Today-Proc. 5 (2018) 3334-3345. https://doi.org/10.1016/j.matpr.2017.11.576.

ro
[3] R. Chandra, S. Yadav, S. Yadav, Phytoextraction potential of heavy metals by native
-p
wetland plants growing on chlorolignin containing sludge of pulp and paper industry,

Ecol. Eng. 98 (2017) 134-145. https://doi.org/10.1016/j.ecoleng.2016.10.017.


re
[4] R.K. Gautam, A. Mudhoo, G. Lofrano, M.C. Chattopadhyaya. Biomass-derived
lP

biosorbents for metal ions sequestration: Adsorbent modification and activation

methods and adsorbent regeneration. J. Environ. Chem. Eng. 2 (2014) 239-259.

[5] J. Dong, Y. Du, R. Duyu, Y. Shang, S. Zhang, R. Han, Adsorption of copper ion
na

from solution by polyethylenimine modified wheat straw, Bioresour. Technol. Rep. 6

(2019) 96-102. https://doi.org/10.1016/j.biteb.2019.02.011.


ur

[6] S.A. Maddodi, H.A. Alalwan, A.H. Alminshid, M.N. Abbas, Isotherm and
Jo

computational fluid dynamics analysis of nickel ion adsorption from aqueous solution

using activated carbon. S. Afr. J. Chem. Eng. 32 (2020) 5-12.

https://doi.org/10.1016/j.sajce.2020.01.002.

30
[7] C.-W. Wong, J.P. Barford, G. Chen, G. Mckay, Kinetics and equilibrium studies for

the removal of cadmium ions by ion exchange resin, J. Environ. Chem. Eng. 2 (2014)

698-707. https://doi.org/10.1016/j.jece.2013.11.010.

[8] K.K. Bhatluri, M.S. Manna, A.K. Ghoshal, P. Saha, Supported liquid membrane

based removal of lead(II) and cadmium(II) from mixed feed: Conversion to solid waste

by precipitation, J. Hazard. Mater. 299 (2015) 504-512.

https://doi.org/10.1016/j.jhazmat.2015.07.030.

[9] M. Garmsiri, H.R. Mortaheb, Enhancing performance of hybrid liquid membrane

of
process supported by porous anionic exchange membranes for removal of cadmium

from wastewater, Chem. Eng. J. 264 (2015) 241-250.

ro
https://doi.org/10.1016/j.cej.2014.11.061.
-p
[10] W.M. Ibrahim, A.F. Hassan, Y.A. Azab, Biosorption of toxic heavy metals from

aqueous solution by Ulva lactuca activated carbon, Egypt. J. Basic Appl. Sci. 3 (2016)
re
241-249. https://doi.org/10.1016/j.ejbas.2016.07.005.
lP

[11] D. Bulgariu, L. Bulgariu, Potential use of alkaline treated algae waste biomass as

sustainable biosorbent for clean recovery of cadmium (II) from aqueous media: batch

and column studies, J. Clean. Prod. 112 (2016) 4525-4533.


na

https://doi.org/10.1016/j.jclepro.2015.05.124.

[12] A. De Rossi, C.V.T. Rigueto, A. Dettmer, L.M. Colla, J.S. Piccin, Synthesis,
ur

characterization, and application of Saccharomyces cerevisiae/alginate composites


Jo

beads for adsorption of heavy metals. J. Environ. Chem. Eng. (2020).

https://doi.org/10.1016/j.jece.2020.104009.

[13] Y. Deng, S. Huang, D.A. Laird, X. Wang, Z. Meng, Adsorption behaviour and

mechanisms of cadmium and nickel on rice straw biochars in single-and binary-metal

31
systems. Chemosphere. 218 (2019) 308-318.

https://doi.org/10.1016/j.chemosphere.2018.11.081.

[14] L. Sellaoui, F.E. Soetaredjo, S. Ismadji, Y. Benguerba, G.L. Dotto, A. Bonilla-

Petriciolet, A.E. Rodrigues, A.B. Lamine, A. Erto, Equilibrium study of single and

binary adsorption of lead and mercury on bentonite-alginate composite: Experiments

and application of two theoretical approaches. J. Mol. Liq. 253 (2018) 160-168.

https://doi.org/10.1016/j.molliq.2018.01.056.

[15] R.F. Nascimento, A.C.A. Lima, C.B. Vidal, D.Q. Melo, G.S.C. Raulino, Adsorção

of
– aspectos teóricos e aplicações ambientais, first ed., Imprensa Universitária, Fortaleza,

2014.

ro
[16] B.M.V. Gama, G.E. Nascimento, D.C.S. Sales, J.M. Rodríguez-Díaz, C.M.B.M.
-p
Barbosa, M.M.M.B. Duarte, Mono and binary component adsorption of phenol and

cadmium using adsorbent derived from peanut shells, J. Clean. Prod. 201 (2018) 219-
re
228. https://doi.org/10.1016/j.jclepro.2018.07.291.
lP

[17] H. Demiral, C. Güngör, Adsorption of copper (II) from aqueous solutions on

activated carbon prepared from grape bagasse, J. Clean. Prod. 124 (2016) 103-113.

https://doi.org/10.1016/j.jclepro.2016.02.084.
na

[18] F. Wang, Y.-q. Dang, X. Tian, S. Harrington, Y.-q. Ma, Fabrication of magnetic

activated carbons from corn cobs using pickle liquor from the surface treatment of iron
ur

and steel, New Carbon Mater. 33 (2018) 303-309. https://doi.org/10.1016/S1872-


Jo

5805(18)60340-3.

[19] J. Shim, P. Velmurugan, B.-T. Oh, Extraction and physical characterization of

amorphous silica made from corn cob ash at variable pH conditions via sol gel

processing, J. Ind. Eng. Chem. 28 (2015) 110-116.

https://doi.org/10.1016/j.jiec.2015.05.029.

32
[20] S. Rovani, A.G. Rodrigues, L.F. Medeiros, R. Cataluña, E.C. Lima, A.N.

Fernandes, Synthesis and characterisation of activated carbon from agroindustrial waste

– Preliminary study of 17β-estradiol removal from aqueous solution, J. Environ. Chem.

Eng. 4 (2016) 2128-2137. https://doi.org/10.1016/j.jece.2016.03.030.

[21] Z. An, J. Xue, H. Cao, C. Zhu, H. Wang, A facile synthesis of silicon carbide

nanoparticles with high specific surface area by using corn cob. Adv. Powder Technol.,

30 (2019) 164-169. https://doi.org/10.1016/j.apt.2018.10.019.

[22] Z. Liu, Y. Sun, X. Xu, X. Meng, J. Qu, Z. Wang, C. Liu, B. Qu, Preparation,

of
characterization and application of activated carbon from corn cob by KOH activation

for removal of Hg(II) from aqueous solution. Bioresour. Technol. 306 (2020) 123154.

ro
https://doi.org/10.1016/j.biortech.2020.123154.
-p
[23] N.V. Sych, S.I. Trofymenko, O.I. Poddubnaya, M.M. Tsyba, V.I. Sapsay, D.O.

Klymchuk, A.M. Puziy, Porous structure and surface chemistry of phosphoric acid
re
activated carbon from corncob. Appl. Surf. Sci. 15 (2012) 75-82.
lP

https://doi.org/10.1016/j.apsusc.2012.07.084.

[24] R. Laus, V.T. De Fávere, Competitive adsorption of Cu(II) and Cd(II) ions by

chitosan crosslinked with epichlorohydrin-triphosphate, Bioresource Technol. 120


na

(2011) 8769-8776. https://doi.org/10.1016/j.biortech.2011.07.057.

[25] D.C. Montgomery, Introdução ao controle estatístico de qualidade, fourth ed.,


ur

LTC, Rio de Janeiro, 2012.


Jo

[26] P. Patnukao, P. Pavasant, Activated carbon from Eucalyptus camaldulensis Dehn

bark using phosphoric acid activation, Bioresource Technol. 99 (2008) 8540-8543.

https://doi.org/10.1016/j.biortech.2006.10.049.

[27] M.J.P.B. Brito, C.M. Veloso, L.S. Santos, R.C.F. Bonomo, R.C.I. Fontan,

Adsorption of the textile dye Dianix® royal blue CC onto carbons obtained from yellow

33
mombin fruit stones and activated with KOH and H3PO4: kinetics, adsorption

equilibrium and thermodynamic studies, Powder Technol. 339 (2018) 334-343.

https://doi.org/10.1016/j.powtec.2018.08.017.

[28] B. Li, L. Yang, C. Wang, Q. Zhang, Q. Liu, Y. Li, R. Xiao, Adsorption of Cd(II)

from aqueous solutions by rape straw biochar derived from different modification

processes, Chemosphere. 175 (2017) 332-340.

https://doi.org/10.1016/j.chemosphere.2017.02.061.

[29] O. Pezoti, A.L. Cazetta, K.C. Bedin, L.S. Souza, R.P. Souza, S.R. Melo, V.C.

of
Almeida, Percolation as new method of preparation of modified biosorbents for

pollutants removal, Chem. Eng. J. 283 (2016) 1305-1314.

ro
https://doi.org/10.1016/j.cej.2015.08.074.
-p
[30] J. Huang, F. Yuan, G. Zeng, X. Li, Y. Gu, L. Shi, W. Liu, Y. Shi, Influence of pH

on heavy metal speciation and removal from waster using micellar-enhanced


re
ultrafiltration. Chemosphere 173 (2017) 199-206.
lP

https://dx.doi.org/10.1016/j.chemosphere.2016.12.137.

[31] M. Razavian, S. Fatemi, A.T. Najafabadi, Extraction of highly pure nickel

hydroxide from spent NiO/Al2O3 catalyst: Statistical study on leaching by sulfuric acid
na

lixiviant and selective. J. Environ. Chem. Eng. 8 (2020) 103660.

https://doi.org/10.1016/j.jece.2020.103660.
ur

[32] G.L. Dotto, N.P.G. Salau, J.S. Piccin, T.R.S. Cadaval, L.A.A. Pinto, Adsorption
Jo

Kinetics in Liquid Phase: Modeling for Discontinuous and Continuous Systems, In: A.

Bonilla-Petriciolet, D. Mendoza-Castillo, H. Reynel-Ávila (Eds.), Adsorption Processes

for Water Treatment and Purification, Springer, Cham, 2017, pp. 53-76.

34
[33] M. Goswami, P. Phukan, Enhanced adsorption of cationic dyes using sulfonic acid

modified activated carbon. J. Environ. Chem. Eng. 5 (2017) 3508-3517.

https://doi.org/10.1016/j.jece.2017.07.016.

[34] K. Attar, D. Bouazza, H. Miloudi, A. Tayeb, A. Boos, A.M. Sastre, H. Demey,

Cadmium removal by a low-cost magadiite-based material: Characterization and

sorption applications, J. Environ. Chem. Eng. 6 (2018) 5351-5360.

https://doi.org/10.1016/j.jece.2018.08.014.

[35] J.B. Neris, F.H.M. Luzardo, P.F. Santos, O.N. Almeida, F.G. Velasco, Evaluation

of
of single and tri-element adsorption of Pb2+, Ni2+ and Zn2+ ions in aqueous solution on

modified water hyacinth (Eichhornia crassipes) fibers, J. Environ. Chem. Eng. 7

ro
(2019) 102885. https://doi.org/10.1016/j.jece.2019.102885.
-p
[36] J.B. Neris, F.H.M. Luzardo, E.G.P. Silva, F.G. Velasco, Evaluation of adsorption

processes of metal ions in multi-element aqueous systems by lignocellulosic adsorbents


re
applying different isotherms: A critical review, Chem. Eng. J. 357 (2019) 404-420.
lP

https://doi.org/10.1016/j.cej.2018.09.125.

[37] P. Loganathan, W.G. Shim, D.P, Sounthararajah, M. Kalaruban, T. Nur, S.

Vigneswaran, Modelling equilibrium adsorption of single, bicomponent, and ternary


na

combinations of Cu, Pb, and Zn onto granular activated carbon, Environ. Sci. Pollut. R.

25 (2018) 16664-16675. https://doi.org/10.1007/s11356-018-1793-9.


ur

[38] Z. Mahdi, Q.J. Yu, A.E. Hanandeh, Competitive adsorption of heavy metal ions
Jo

(Pb2+, Cu2+, and Ni2+) onto date seed biochar: batch and fixed bed experiments, Sep.

Sci. Technol. 54 (2019) 888-901. https://doi.org/10.1080/01496395.2018.1523192.

[39] D.D. Do, Adsorption Analysis: Equilibria and Kinetics, Series on Chemical

Engineering, volume 2, London, 1998.

35
[40] M.S. Monteiro, R.F. Farias, J.A.P. Chaves, S.A. Santana, H.A.S. Silva, C.W.B.

Bezerra, Wood (Bagassa guianensis Aubl) and Green coconut mesocarp (cocos

nucifera) residues as textile dye removers (Remazol Red and Remazol Brilliant Violet).

J. Environ. Manage. 204 (2017) 23-30. https://doi.org/10.1016/j.jenvman.2017.08.033.

[41] N. Maaloul, N.; P. Oulego, M. Rendueles, A. Ghorbala, M. Díaz, Novel

biosorbents from almond shells: Characterization and adsorption properties modeling

for Cu(II) ions from aqueous solutions. J. Environ. Chem. Eng. 5 (2017) 2944-2954.

https://doi.org/10.1016/j.jece.2017.05.037.

of
[42] N. Worasuwannarak, T. Sonobe, W. Tanthapanichakoon. Pyrolysis behaviors of

rice straw, rice husk, and corncob by TG-MS technique. J. Anal. Appl. Pyrol. 78 (2007)

ro
266-271. https://doi.org/10.1016/j.jaap.2006.08.002.
-p
[43] J. Zhu, Y. Li, L. Xu, Z. Liu, Removal of toluene from waste gas by adsorption-

desorption process using corncob-based activated carbons as adsorbents. Ecotox.


re
Environ. Safe. 165 (2018) 115-125. . https://doi.org/10.1016/j.ecoenv.2018.08.105.
lP

[44] M. Duarte, G. Nascimento, M. Santos, T. Silva, N. Campos, J. Santos, C. Barbosa,

Adsorption of phenol on adsorbents produced from coconut tree waste: kinetic and

equilibrium studies, Environ. Eng. Manage. J. 18 (2019) 693-705.


na

https://doi.org/10.30638/eemj.2019.063.

[45] S.X. Liu, X. Chen, Z.F. Liu, H.L. Wang, Activated carbon with excellent
ur

chromium(VI) adsorption performance prepared by acid–base surface modification, J.


Jo

Hazard. Mater. 141 (2007) 315-319. https://doi.org/10.1016/j.jhazmat.2006.07.006.

[46] W. Yin, C. Zhao, J. Xu, J. Zhang, Z. Guo, Y. Shao, Removal of Cd(II) and Ni(II)

from aqueous solutions using activated carbon developed from powder-hydrolyzed-

feathers and Trapa natans husks, Colloids Surf. A Physicochem. Eng. Asp. 560 (2019)

426-433. https://doi.org/10.1016/j.colsurfa.2018.10.031.

36
[47] S. Wang, J.-H. Kwak, M.S. Islam, M.A. Naeth, M.G. El-Din, S.X. Chang, Biochar

surface complexation and Ni(II), Cu(II), and Cd(II) adsorption in aqueous solutions

depend on feedstock type, Sci. Total Environ. 712 (2020) 136538.

https://doi.org/10.1016/j.scitotenv.2020.136538.

[48] Z. Mahdi, Q.J. Yu, A.E. Hanandeh, Investigation of the kinetics and mechanisms

of nickel and copper ions adsorption from aqueous solutions by date seed derived

biochar. J. Environ. Chem. Eng., 6 (2018) 1171-1181.

https://doi.org/10.1016/j.jece.2018.01.021.

of
[49] D. Feng, C. Aldrich, Adsorption of heavy metals by biomaterials derived from the

marine alga Ecklonia maxima, Hydrometallurgy. 73 (2004) 1-10.

ro
https://doi.org/10.1016/S0304-386X(03)00138-5.
-p
[50] K.V. Kumar, K. Porkodi, Mass transfer, kinetics and equilibrium studies for the

biosorption of methylene blue using Paspalum notatum. J. Hazar. Mater. 146 (2007)
re
214–226. https://doi.org/10.1016/j.jhazmat.2006.12.010.
lP

[51] C.H. Giles, T.H. Macewan, S.N. Nakhwa, D. Smith, Studies in adsorption, Part XI,

A system of classification of solution adsorption isotherms, and its use in diagnosis of

adsorption mechanisms and in measurement of specific surface areas of solids, J. Chem.


na

Soc. 14 (1960) 3973-3993. http://doi.org/10.1039/jr9600003973.

[52] M.-E. Lee, J.H. Park, J.W. Chung, Comparison of the lead and copper adsorption
ur

capacities of plant sourcematerials and their biochars, J. Environ. Manage. 236 (2019)
Jo

118-124. https://doi.org/10.1016/j.jenvman.2019.01.100.

[53] C. Wei, Y. Huang, Q. Liao, A. Xia, X. Zhu, X. Zhu, Adsorption thermodynamic

characteristics of Chlorella vulgaris with organic polymer adsorbent cationic starch:

Effect of temperature on adsorptive capacity and rate, Bioresource Technol. 293 (2019)

122056. https://doi.org/10.1016/j.biortech.2019.122056.

37
[54] F. Deniz, E.T. Ersanli, A natural macroalgae consortium for biosorption of copper

from aqueous solution: Optimization, modeling and design studies, Int. J.

Phytoremediat. 20 (2018) 362-368. https://doi.org/10.1080/15226514.2017.1393387.

[55] T. Bohli, A. Ouederni, N. Fiol, I. Villaescusa. Evalution of an activation carbon

from olive stones used as an adsorbent for heavy metal removal from aqueous phases.

C. R. Chimie 18 (2015) 88-99. https://doi.org/10.1016/j.crci.2014.05.009.

[56] L. Wu, W. Wan, Z. Shang, X. Gao, N. Kobayashi, G. Luo, Z. Li, Surface

modification of phosphoric acid activated carbon by using non-thermal plasma for

of
enhancement of Cu(II) adsorption from aqueous solutions. Sep. Purif. Technol. 197

(2018) 156-169. https://doi.org/10.1016/j.seppur.2018.01.007.

ro
[57] X. Gao, L. Wu, Q. Xu, W. Tian, Z. Li, N. Kobayashi, Adsorption kinetics and

H3PO4 activation. Environ. Sci.


-p
mechanisms of copper ions on activated carbons derived from pinewood sawdust by

fast Pollut. Res. 25 (2018) 7907-7915.


re
https://doi.org/10.1007/s11356-017-1079-7.
lP

[58] T. Van Thuan, B.T.P. Quynh, T.D. Nguyen, V.T.T. Ho, L.G. Bach. Response

surface methodology approach for optimization of Cu2+, Ni2+ and Pb2+ adsorption

using KOH-activated carbon from banana peel. Surf. Interfaces 6 (2017) 209-217.
na

https://doi.org/10.1016/j.surfin.2016.10.007.

[59] V.K. Gupta, D. Pathania, S. Sharma, Adsorptive remediation of Cu(II) and Ni(II)
ur

by microwave assisted H3PO4 activated carbon. Arab. J. Chem. 10 (2017) S2836-


Jo

S2844. https://doi.org/10.1016/j.arabjc.2013.11.006.

[60] H. Hasar, Adsorption of nickel(II) from aqueous solution onto activated carbon

prepared from almond husk. J. Hazar. Mater. 97 (2003) 49-57.

https://doi.org/10.1016/S0304-3894(02)00237-6.

38
[61] I.K. Silva Correia, P.F. Santos, C.S. Santana, J.B. Neris, F.H.M. Luzardo, F.G.

Velasco, Application of coconut shell, banana peel, spent coffee grounds, eucalyptus

bark, piassava (Attalea funifera) and water hyacinth (Eichornia crassipes) in the

adsorption of Pb2+ and Ni2+ ions in water, J. Environ. Chem. Eng. 6 (2018) 2319-2334.

https://doi.org/10.1016/j.jece.2018.03.033.

of
ro
-p
re
lP
na
ur
Jo

39
Table 1. Parameters from pseudo-first order (PFO), pseudo-second order (PSO) and

intraparticle diffusion monocomponent models for Cu2+ and Ni2+ kinetics of adsorption

on the corn cob activated carbon.

CCu2+ CNi2+
Model Parameter
0.50 1.00 1.25 0.50 1.00 1.25

– qexp 0.238 0.369 0.422 0.180 0.192 0.267

PFO qe 0.214 ± 0.008 0.33 ± 0.02 0.36 ± 0.01 0.170 ± 0.003 0.20 ± 0.01 0.248 ± 0.005

k1 0.20 ± 0.05 0.06 ± 0.02 0.21 ± 0.05 0.25 ± 0.04 0.5 ± 0.3 0.22 ± 0.03

χ2ν 7 ∙10-4 0.002 0.002 1·10-4 0.002 3·10-4

of
R2 0.814 0.783 0.818 0.932 0.585 0.936

PSO qe 0.225 ± 0.006 0.35 ± 0.01 0.38 ± 0.01 0.176 ± 0.002 0.20 ± 0.01 0.257 ± 0.004

ro
k2 1.2 ± 0.3 0.29 ± 0.08 0.7 ± 0.2 2.4 ± 0.3 4±3 1.5 ± 0.2

χ2ν 3·10-4 1·10-3 1·10-3 5·10-5 2·10-3 1·10-4

R2 0.917 0.898 0.914


-p 0.975 0.618 0.973

k id1 (106 ± 6) ·10-4 (194 ± 8) ·10-4 (202 ± 6) ·10-4 (151 ± 9) ·10-4 -0.031 ± 0.001 (284 ± 10) ·10-4
re
c1 0.117 ± 0.004 0.126 ± 0.005 0.188 ± 0.004 0.084 ± 0.003 0.272 ± 0.003 0.102 ± 0.006

R21 0.905 0.914 0.898 0.971 0.441 0.995


lP

Weber- χ2ν,1 4.54 6.95 14.95 2.24 223.36 0.76

Morris k id2 (17 ± 4) ·10-4 (48 ± 4) ·10-4 (49 ± 5) ·10-4 (14 ± 3) ·10-4 (-13 ± 3) ·10-4 (26 ± 3) ·10-4
na

c2 0.203 ± 0.007 0.270 ± 0.006 0.314 ± 0.008 0.152 ± 0.005 0.219 ± 0.004 0.206 ± 0.005

R22 0.977 0.948 0.760 0.567 0.148 0.550

χ2ν,2
ur

0.06 1.29 5.73 1.38 8.61 6.76


Jo

40
Table 2. Parameters from pseudo-first order (PFO), pseudo-second order (PSO) and proposed in this work

(PM) bicomponent models for Cu2+ and Ni2+ kinetics of adsorption on the corn cob activated carbon.

PFO PSO PM

q exp K
Met
C0 qe k1 χ2ν R2 qe k2 χ2ν R2 qm ka kd = ka n χ2ν R2
al
/k d

0.16 0.14 0.02 0.16 0.2


5 6
Cu2 0.5 5 8± 8± 3·1 0.86 2± 6± 2·1 0.93 (6 ± (3 ± 2.0 3·1 0.93
± ±
+
0 0.00 0.00 0-4 9 0.00 0.0 0-4 0 2)·102 1)·102 4 0-6 0
2 2
7 5 7 6

0.06 2.
0.06 0.06 0.06 1

of
5 1.7 (3.9 ± (1.6 ± 0
1.0 2± 0± 8·1 0.82 5± 8·1 0.82 2 2.5 6·1 0.94
Ni2+ ± 0.5)·1 0.5)·1 ±
0 0.00 0.02 0-5 3 0.00 0-5 9 ± 3 0-5 3
02 02

ro
0.6 0.
3 0 3 4
8

0.27 0.24 0.03 0.27 0.1

Cu2
+
1.0

0
2 7±

0.00

0.00
6·1

0 -4
0.91

7

0.00

0.0
2·1

0 -4
-p
0.96

6
8

3
(6 ±

1)·10 2
(3 ±

2)·10 2
2.0

1
2

1
6·1

0 -6
0.82

1
re
9 5 8 3

0.05 0.07 0.11 0.07


3.0 2 (1.6 ± 3
1.0 1 3± 0± 2·1 0.71 5± 2·1 0.64 (3 ± 2.0 9·1 0.79
lP

Ni2+ ± ± 0.7)·1 ±
0 0.00 0.04 0-4 5 0.00 0-4 9 2)·102 6 0-5 3
2.0 1 02 1
4 0 6

0.28 1.
0.02 0.1 1
na

6 0.26 0.28 1
Cu2 1.2 7± 1·1 0.84 5± 6·1 0.92 6 (10 ± (6 ± 1.4 4·1 0.87
± ± ±
+
5 0.00 0-3 9 0.0 0-4 5 ± 4)·103 3)·103 7 0-6 2
0.01 0.01 0.
6 4 7
ur

0.05 0.88 1.
0.06 0.06 1
8 0.13 0 4
Jo

1.0 3± 5·1 5± 4± 6·1 0.84 0 (7 ± (5 ± 1.2 8·1 0.94


Ni2+ ± ±
0 0.00 0-5 0.00 1 0-5 4 ± 2)·103 1)·103 8 0-5 1
0.02 0.
2 3 2
3

0.27 0.03 0.2 1 0.


0.24 0.26 (2.3 ± (1.9 ±
Cu2 1.0 3 7± 1·1 0.80 5± 6·1 0.89 8 1.2 9 6·1 0.85
± ± 0.4)·1 0.9)·1
+ -3 -4 -6
0 0.00 0 1 0.0 0 2 ± 3 ± 0 5
0.01 0.01 03 03
8 7 6 0.

41
6

0.02 0.56 1.
0.40 0.04 1
0 0.09 7 0
0.5 0± 1·1 1± 5± 1·1 0.53 6 (6 ± (5 ± 1.0 9·1 0.63
Ni2+ ± ±
0 0.00 0-4 0.00 4 0-4 6 ± 5)·103 4)·103 8 0-5 0
0.04 0.
3 4 9
9

0.25 0.
0.23 0.02 0.25 0.1 1
4 (2.8 ± 9
Cu2 1.0 5± 8± 6·1 0.90 6± 7± 3·1 0.94 8 (2 ± 1.2 3·1 0.93
0.3)·1 ±
+
0 0.00 0.00 0-4 9 0.00 0.0 0-4 8 ± 1)·103 3 0-6 8
03 0.
9 4 9 3 6
6

0.05 0.65 0.

of
0.07 0.07 1
2 0.22 2 (1.3 ± (1.2 ± 9
1.2 0± 2·1 2± 6± 2·1 0.65 4 1.1 9·1 0.72
Ni2+ ± 0.7)·1 0.9)·1 ±
5 0.00 0-4 0.00 4 0-4 8 ± 0 0-5 0

ro
0.08 03 03 0.
4 4 9
7

-p
re
lP
na
ur
Jo

42
Table 3. One-way ANOVA test applied to Cu2+ and Ni2+ monocomponent adsorption on

the corn cob activated carbon at 293, 303, 313 and 323 K.

Metal Source SS Df MS F p-value = P(F > 0.255)

Cu2+ Temperature 0.0013 3 0.0004 0.0255 0.9944

Error 0.4653 28 0.0166

Total 0.4666 31

Ni2+ Temperature 0.0021 3 0.0007 0.1138 0.9513

Error 0.1692 28 0.0060

Total 0.1713 31

of
Temperature: between-groups variation; Error: within-groups variation; SS: sum of squares;
df: degrees of freedom; MS = SS/df: mean squared error; F = MSColumns /MSError : F-statistic

ro
-p
re
lP
na
ur
Jo

43
Table 4. Parameters of modelling of monocomponent adsorption equilibrium isotherms

for Cu2+ and Ni2+ on the corn cob activated carbon at 303 K.

Model Parameter Cu2+ Ni2+

Langmuir qmL 0.37 ± 0.02 0.27 ± 0.03

KL 54 ± 18 20 ± 16

χ2ν 1 ·10-3 2 ·10-3

R2 0.946 0.817

Freundlich KF 0.42 ± 0.02 0.28 ± 0.02

nF 4.3 ± 0.6 7±4

of
χ2ν 1 ·10-3 2 ·10-3

R2 0.964 0.818

ro
Sips qmS 0.5 ± 0.1 0.36 ± 0.04

KS 19.6 ± 0.1 17 ± 9

nS
-p
0.54 ± 0.08 0.3 ± 0.2

χ2ν 7·10-4 2 ·10-3


re
R2 0.964 0.829
lP

qexp 0.39 0.28


na
ur
Jo

44
Table 5. Activated carbons produced from several biomasses reported in the literature

for Cu2+ and Ni2+ removal by using a batch adsorber.

Precursor Activator C0 [mmol·L-1] pH w/V [g·L-1] q m [mmol·g-1] T [K] Reference

Cu2+

Olive stones H3PO4 5.00 5.0 6.0 0.30 303 [55]

Walnut shell H3PO4 1.57 5.0 3.0 0.15 313 [56]

Pinewood sawdust H3PO4 3.15 6.0 2.0 0.39 298 [57]

Corn cob H3PO4 2.00 4.0 2.0 0.50 303 This work

Ni2+

Banana peel KOH 1.54 6.4 2.4 0.36 - [58]

Ficus carica fiber H3PO4 8.52 5.0 5.0 0.32 323 [59]

of
Almond husk CO2 4.26 5.0 5.0 0.52 293 [60]

Corn cob H3PO4 2.00 4.0 2.0 0.36 303 This work

ro
-p
re
lP
na
ur
Jo

45
Table 6. One-way ANOVA test applied to Cu2+/Ni2+ for bicomponent adsorption on the

corn cob activated carbon at 293, 303, 313 and 323 K.

Metal Source SS Df MS F p-value = P(F > 0.255)

Cu2+ Temperature 0.0059 3 0.0018 0.150 0.9285

Error 0.3284 28 0.0177

Total 0.3337 31

Ni2+ Temperature 0.0046 3 0.0015 4.080 0.0160

Error 0.0102 28 0.0004

Total 0.0147 31

of
ro
-p
re
lP
na
ur
Jo

46
Table 7. Parameters of modelling of bicomponent adsorption equilibrium isotherms for

Cu2+ and Ni2+ on the corn cob activated carbon at 293, 303, 313 e 323 K.

Model Parameter 293 K 303 K 313 K 323 K

Non-
χ2ν 0.418 0.596 5.730 2.122
modified
R2 0.877 0.804 0.005 0.403
Langmuir

Extended qmEL 0.39 ± 0.02 0.40 ± 0.02 0.31 ± 0.01 0.32 ± 0.01

Langmuir K EL,Cu 51 ± 17 69 ± 23 116 ± 43 264 ± 141

Isotherm K EL,Ni

of
6±2 8±3 6±2 10 ± 4

χ2ν 2∙10-4 6 ∙10-4 4 ∙10-4 7 ∙10-4

ro
R2 0.966 0.966 0.969 0.954

Modified ηL,Cu 0.7 ± 0.3 0.5 ± 0.2 1.2 ± 0.4 1.0 ± 0.5

Langmuir ηL,Ni

χ2ν
2±1

7 ∙10-4
-p
1.4 ± 0.7

8 ∙10-4
15 ± 6

5 ∙10-4
5±2

9 ∙10-4
re
R2 0.954 0.953 0.966 0.942

Non-
χ2ν 17.548 1.226 16.194 1.972
lP

modified
R2 0.110 0.573 0.162 0.556
Sips

qmES
na

Extended 0.6 ± 0.2 0.6 ± 0.1 0.42 ± 0.08 0.35 ± 0.04

Sips K ES,Cu 5±2 12 ± 2 18 ± 2 23 ± 2

K ES,Ni 0.5 ± 0.4 1.7 ± 0.9 0.8 ± 0.1 7±7


ur

nES 0.31 ± 0.07 0.34 ± 0.07 0.4 ± 0.1 0.5 ± 0.1

χ2ν 2∙10-4 2∙10-4 2∙10-4 6 ∙10-4


Jo

R2 0.990 0.990 0.986 0.968

47

You might also like