Anda di halaman 1dari 35

Accepted Manuscript

Title: Adsorption of Cr(VI) from aqueous solution by


prepared high surface area activated carbon from Fox nutshell
by chemical activation with H3 PO4

Authors: Arvind Kumar, Hara Mohan Jena

PII: S2213-3437(17)30127-6
DOI: http://dx.doi.org/doi:10.1016/j.jece.2017.03.035
Reference: JECE 1541

To appear in:

Received date: 21-10-2016


Revised date: 23-3-2017
Accepted date: 26-3-2017

Please cite this article as: Arvind Kumar, Hara Mohan Jena, Adsorption of Cr(VI)
from aqueous solution by prepared high surface area activated carbon from Fox
nutshell by chemical activation with H3PO4, Journal of Environmental Chemical
Engineeringhttp://dx.doi.org/10.1016/j.jece.2017.03.035

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.
Adsorption of Cr(VI) from aqueous solution by prepared high surface area activated

carbon from Fox nutshell by chemical activation with H3PO4

Arvind Kumar, Hara Mohan Jena *

Department of Chemical Engineering, National Institute of Technology (NIT), Rourkela

769008, Orissa, India

*Corresponding author. Tel.: 0661-2462264

E-mail addresses: arvindkr202@gmail.com (A. Kumar), hmjena@nitrkl.ac.in (H.M. Jena)

Highlights

 Cr(VI) adsorption study onto prepared activated carbon with H3PO4 activator.
 The maximum adsorption capacity of Cr(VI) is 74.95 mg/g.
 The adsorption process is endothermic and spontaneous one.
 A maximum removal of 71.86% is achieved in the column study.

Abstract

Adsorption studies of Cr(VI) onto activated carbon, FNAC-700-1.5 (BET surface area of 2636

m2/g and total pore volumeof 1.53 cm3/g), prepared from Fox nutshell by chemical activation

with H3PO4 at an impregnation ratio of 1.5 and activation temperature of 700 oC under N2

atmosphere were carried out in the present work. In batch adsorption experiments the effect of

agitation speed, pH, temperature, adsorbent dosage, initial concentration of Cr(VI), and contact

time were studied. Also adsorption kinetics, equilibrium, and thermodynamics of Cr(VI)

adsorption were studied. The adsorption was found to follow the pseudo second order model.

A maximum adsorption of 74.95 mg/g was obtained at the operating condition of 35 mg/L of

initial concentration of Cr(VI) at pH of 2.0, temperature 45 oC and contact time of 3 h.

Thermodynamic parameters such as ΔGo, ΔHo and ΔSo were calculated by using Van´t hoff

equation. The adsorption process was endothermic and spontaneous one. In column

experiments, the effects of the bed height of the packing material and flow rate of Cr(VI)

1
solution on adsorption characteristics were studied. A maximum removal of 71.86% was

achieved in the column study for 10 mg/L initial Cr(VI) concentration at 4 cm bed height, and

5 mL/min flow rate under ambient conditions at a temperature around 30 oC.

Keywords: Activated carbon; Fox nutshell; Adsorption equilibrium; Thermodynamics;

Column study.

1. Introduction

Primary sources of wastes containing heavy metals are the industrial activities (e.g.

mining, painting, car manufacturing, metal plating, and tanneries) and agricultural activities

(when fertilizers and fungicidal sprays are intensively used) [1]. Heavy metals are considered

one of the most hazardous contaminants in wastewater. Cadmium, chromium, copper, lead,

mercury, and nickel are most toxic metals according to the WHO, 2004 and WHO, 2006 [1].

Chromium is an essential nutrient for plant and animal metabolism (glucose

metabolism, amino, and nucleic acid synthesis). When chromium accumulated at high levels,

generate serious problems (nausea, skin ulcerations, lung cancer) and, when the concentration

reaches 0.1 mg/g body weight, it becomes lethal [2]. Chromium is considered as an important

toxic material as it does not undergo biodegradation [3]. Chromium as pollutant is introduced

into natural waters by a variety of industrial wastewaters coming out from industries such as;

textile, steel fabrication, agricultural runoff, paint manufacturing, leather tanning,

electroplating, and metal finishing ones [3-5]. Chromium occurs as both trivalent [Cr(III)] and

hexavalent [Cr(VI)] states in the aquatic environment. Cr(VI) is primarily present in the form

of chromate (CrO4−) and dichromate (Cr2O7−). Cr(VI) possesses significantly higher levels of

toxicity than the other valency states [4-5]. Cr(VI) tolerance limit for discharge into inland

2
surface waters is 0.1 mg/L and in potable water is 0.05 mg/L [6]. Thus removal of Cr(VI) from

contaminated water is of prime importance to protect the environment.

Among the various removal techniques, the adsorption by using activated carbon is

effective than other adsorbents (clay material, natural and synthetic polymers and their

composites) due to its high specific surface area, adequate pore size distribution, variable

characteristics of surface chemistry [7-9]. The activated carbons are most widely used in many

applications such as wastewater treatment, removal of harmful gasses, solvent recovery, colour

removal, and as electrode materials in electrochemical devices and processes [9-10]. In the

adsorption process, the high cost of adsorbents is the main barrier in industrial use. The

adsorbent cost can be reduced if prepared from some waste materials such as; agricultural

waste, biomass, and various solid waste substances [11] and if appropriate regeneration is done

[12]. So there is a need to develop low-cost and easily available adsorbent for removal of

Cr(VI) from contaminated wastewater.

In the present work adsorptive removal of Cr(VI) from aqueous solution was carried

out by previously prepared high surface area activated carbon from Fox nutshell (FNAC-700-

1.5) by activation with H3PO4 [13]. Among the various dehydrating agents used for chemical

activation, phosphoric acid (H3PO4) is preferred recently due to environmental and economic

concerns. Phosphoric acid allows the development of both micropores and mesopores in the

resulting activated carbon [14]. Polyphosphoric acids are formed by condensation of two or

more H3PO4 molecules with the elimination of water as a complex mixture of linear molecules

of various chain lengths. At high temperatures, the reactions with carbon matrix increases and

develop more pores and release P4, CO2, H2O and other volatile materials [15]. To the best of

our knowledge, no study was reported on the removal of hexavalent chromium from aqueous

solution by FNAC-700-1.5. Fox nutshell, a low ash (5%) content and high volatile matter

(70.1%), is a good precursor for the production of an effective activated carbon due to its

3
abundance and availability. The adsorption kinetics, equilibrium, and thermodynamics of

Cr(VI) adsorption on the prepared activated carbon were studied. In the batch adsorption study

the influence of operating parameters, such as agitation speed, pH, temperature, contact time

and initial concentration of Cr(VI) on adsorption characteristics were investigated. Linear least

square method and non-linear method of three widely used isotherms, Langmuir, Freundlich,

and Tempkin, was examined to experiment Cr(VI) adsorption onto FNAC-700-1.5. A trial and

error procedure was used for non-linear method by using the solver add-in with Microsoft’s

spreadsheet, Microsoft Excel. In the column study effect of bed height of carbon and flow rate

on breakthrough and saturation times, adsorption capacity and percentage removal were

studied.

2. Experimental

2.1. Materials

The chemicals used in the present study, such as; ZnCl2, K2Cr2O7, NaOH and HCl were

of AR grade of brand Merck, India. Stock solution of 1000 mg/L of Cr(VI) was prepared by

dissolving 2.8269 g K2Cr2O7 in 1000 mL distilled water. Standard working solutions of Cr(VI)

ions of various concentrations ranging from 10 to 35 mg/L were obtained by dilution of the

stock solution.

2.2. Effect of pH and determination of PZC

The effect of pH on the amount of Cr(VI) removal was analyzed over the pH range

from 2 to 7 by using 0.1 N NaOH and 0.1 N HCl solutions for the pH adjusting. In this study,

100 mL of Cr(VI) solution of 10 mg/L was agitated with 0.02 g of FNAC-700-1.5 using orbital

shaker at 30 °C temperature. Agitation was made for 3 h that is more than sufficient to reach

equilibrium at a constant agitation speed of 150 rpm. The samples were then centrifuged, and

the left out concentrations in the supernatant solution were analyzed by a spectrophotometer.

4
The PZC was measured as follows: 20 mL of NaCl solution (0.1 N) was placed in an

Erlenmeyer flask including 0.1 g of AC. The initial pH was adjusted between 2 and 12 by the

addition of 0.1 N NaOH and 0.1 N HCl. After a contact time of 24 h under magnetic agitation,

the final pH was determined.

2.3. Batch adsorption and kinetic experiment

The batch adsorption experiments were conducted in a set of 250 ml of Erlenmeyer

flasks containing 100 mL of Cr(VI) (10, 15, 20, 25, 30 and 35 mg/L) solution with

predetermined amount of adsorbent (in the range of 0.01g 0.07 g). The flasks were agitated in

an isothermal orbital shaker at 150 rpm and temperature range of 25-45 °C for 3 h. The pH was

maintained by the addition of 0.1 N NaOH and 0.1 N HCl as required and was recorded by

using pH meter (361 model, Systronic). The following relations were used to determine the

equilibrium adsorption capacity, qe (mg/g) and percentage removal, R (%):

(𝐶0 − 𝐶𝑒 )𝑉
𝑞𝑒 = , (1)
𝑚𝑠

𝐶0 − 𝐶𝑒
𝑅(%) = 100 , (2)
𝐶0

Kinetic studies were also performed according to the method as described above.

Kinetic study provides the information of uptake rate and controls the equilibrium time [16].

As per literature the most appropriate pseudo first and pseudo second order model were used

to describe Cr(VI) adsorption onto FNAC-700-1.5. The adsorption capacity qt (mg/g) at

different contact time t (min) was determined using the following equation:

(𝐶0 − 𝐶𝑡 )𝑉
𝑞𝑡 = , (3)
𝑚𝑠

where C0, Ce, and Ctwere the initial, equilibrium, and at time t (min) of Cr(VI) concentrations

(mg/L) respectively, V the volume of solution (L) and ms was the dry weight of the adsorbent

(g) added.

5
2.4. Fixed-bed column adsorption studies
A Perspex glass cylindrical tube having 2.5 cm internal diameter and 20 cm height was

used to construct the adsorption column. The schematic of the experimental arrangement and

photographic view of the adsorption column is shown in Fig.1. The column was packed with

calculated amount of FNAC-700-1.5 to obtain a desired bed height of 4 cm. For this amount

of activated carbon used was 6.0 g. The bed was held in place between two plugs of cotton.

The glass beads supported the cotton in the lower and upper part of the column. The volumetric

flow rate of Cr(VI) solution was adjusted to 5, 10 and 15 mL/min in the experiments at various

flow rates in the range 5-15 mL/min. The adsorbate solutions were fed continuously to the

column in an upward flow mode using the peristaltic pump (Rivtek, India) at room temperature

of 30 °C. To study the effect of bed height, sample collection ports were provided at each 1 cm

interval. Effluent samples were collected from the samplecollecting sites of the column at

different time intervals, and the concentration of Cr(VI) was determined colorimetrically.

The loading behavior of Cr(VI) in its dynamic adsorption from the solution by FNAC-

700-1.5 was shown in the form of breakthrough (BT) curves. The BT curve is usually expressed

in terms of normalized concentration, as the ratio of outlet adsorbate concentration to the inlet

adsorbate concentration (Ct/C0) for a given bed height. The total adsorbed quantity of

adsorbate, qtot (mg) was estimated as area under the BT curve which was obtained by

integrating the plot as per the Eq. (4) [17-18] shown below.
𝑡=𝑡𝑡𝑜𝑡
𝑄
𝑞𝑡𝑜𝑡 = ∫ 𝐶𝑎𝑑 𝑑𝑡, (4)
1000 𝑡=0

where Q (mL/min) is the volumetric flow rate, Cad (mg/L) is the difference between the initial

and final concentrations of adsorbates at the end of the total flow time till exhaustion, ttot (min).

Equilibrium adsorbates uptake in the column or maximum capacity of the column (bed

capacity), qbed (mg/g) is defined by Eq. (5) [17-18]:

6
𝑞𝑡𝑜𝑡
𝑞𝑏𝑒𝑑 = , (5)
𝑊

where W is the amount of prepared activated carbon FNAC-700-1.5. The quantity of Cr(VI)

sent to the column (mtot)is given in Eq. (6) [19-20]:

𝐶0 𝑄𝑡𝑡𝑜𝑡
𝑚𝑡𝑜𝑡 = , (6)
1000

The % removal of Cr(VI) are given by Eq. (7) [20]:

𝑞𝑡𝑜𝑡
%𝑟𝑒𝑚 = 100, (7)
𝑚𝑡𝑜𝑡

2.5. Analytical methods

Aliquot of adsorbate was withdrawn from the suspension at pre-set time intervals and

centrifuged (2 mL) at 10,000 rpm for 5 min to remove the adsorbent particles before analytical

measurements were made. The concentrations of Cr(VI) after adsorption was carried out

colorimetrically with the 1,5-diphenyl carbazide method [12] using a UV–Visible

spectrophotometer (Jasco, Model V-530, Japan) at 540 nm.

3. Results and discussion

3.1. Cr(VI) adsorption onto prepared activated carbon FNAC-700-1.5

3.1.1. Effect of agitation speed

Figure 2 shows the effect of agitation speed on % removal of Cr(VI) adsorption onto

FNAC-700-1.5. From Fig.2, the results show that the removal efficiency of Cr(VI) increases

with increase in agitation speed and maximum removal by FNAC-700-1.5 are achieved at 150

rpm, and above 150 rpm, the % removal is almost constant. Thus, the agitation speed was fixed

at 150 rpm for Cr(VI) adsorption for the further experiments. These results indicate that an

effective transport of Cr(VI) ions towards the adsorbent surfaces occurred, due to less

resistance to diffusion at higher agitation speed. Some studies also report similar variation in

Cr(VI) adsorption and speed of 150 rpm as sufficient to break all diffusion resistances [21-24].

3.1.2. Effect of pH

7
The solution pH is one of the important parameter having considerable influence on the

heavy metal ions adsorption from aqueous solution, because the pH affects surface charge

density of the adsorbent and the charge of the metallic species present [25-26]. To study the

effect of pH on adsorption, pH was varied from 2-7 at 10 mg/L of initial Cr(VI) concentration

with carbon dose of 0.02 mg/100 mL adsorbate solution and 30 °C temperature. pH below 2

was not experimented as per the published reports, which says “Cr(VI) could be reduced to

Cr(III) in the presence of activated carbon under highly acidic conditions (pH < 2)” [12]. Figure

3 shows the influence of solution pH on Cr(VI) adsorption by FNAC-700-1.5. As seen from

Fig. 3, the removal of Cr(VI) from aqueous solution is greatly dependent on the pH values, and

the maximum adsorption occurs when the pH value is 2.0.

Hexavalent chromium adsorption with pH vary could be explained as follows.

Hexavalent chromium exists in different forms in varying pH of aqueous solution and the

stability of these forms, such as HCrO4−,CrO42−, or Cr2O72− [12]. In the pH ranging from 2.0 to

6.0, HCrO4− and Cr2O72− ions mainly exist in equilibrium; HCrO4− is the dominant form of

Cr(VI) at pH 2 and the predominant form shifts to chromate ion (CrO42−) as pH increases. A

chromate ion (CrO42−) needs two active sites due to its two negative charges, whereas an

HCrO4− ion only needs one active site. It is well known that an increase in the adsorption

capacity of Cr(VI) is observed due to more HCrO4− ions formed at the expense of CrO42−ions

with decreasing the pH value. So that HCrO4− is adsorbed preferentially on carbon [24]. The

behavior for better adsorption at low pH by activated carbon may be attributed to the large

number of H+ ions present which in turn neutralize the negatively charged adsorbent surface,

thereby reducing hindrance to the diffusion of chromate ions [12]. At higher pH, decreasing

the adsorption capacity may be due to the competitive adsorption of chromate and hydroxyl

ions [24].

8
The pH at point zero charge (PZC) was found to be 2.7. At pH < pHPZC, the carbon

surface takes up more H+, increasing Cr(VI) ions bind on the adsorbent surface. At pH > pHPZC,

the adsorbent surface is negatively charged, the increasing electrostatic repulsion between

negative adsorbate species and adsorbent particles would lead to decrease adsorption capacity

of Cr(VI) ions. From Fig. 3, the adsorption was high as 46.32 mg/g at pH value of 2 that was

below PZC of the FNAC-700-1.5. However, the adsorption was decreasing at pH of solution

more than the PZC. This is in agreement with our experimental observations showing

adsorption decreases at pH > 2.

3.1.3. Effect of temperature

To access the effect of temperature on Cr(VI) adsorption, the temperature of the system

was varied from 20-45 °C at pH of 2, initial concentration of 10 mg/L, carbon dose of 0.02

g/100 mL and shaken in the orbital shaker for 3 h at agitation speed of 150 rpm. Figure 4 shows

the effect of temperature on % removal of Cr(VI). The figure indicates an increase in removal

of the adsorbate with temperature that indicates the adsorption process is endothermic in nature.

At this particular condition, % removal increased from 55.88 to 60.02% for an increase in

temperature from 20-45 °C.

3.1.4. Effect of adsorbent dose

The experiments were conducted at an initial solution concentration of Cr(VI) of 10

mg/L at 45 °C temperature and pH 2 while the amount of adsorbent added was varied to

evaluate the effect of carbon dose on removal of Cr(VI) and the result is presented in Fig. 5.

Activated carbon dosage were ranged from 0.01 to 0.07 g/100 mL of chromium solution and

equilibrated for 3 h. The percentage removal of Cr(VI) is found to increases with increase in

adsorbent dose up to 0.04 g significantly and after that remained unchanged. Thus, to get the

better Cr(VI) removal, 0.04 g was chosen as an optimal mass of the adsorbent in further

9
experiments. With increasing adsorbent dose, there is increase in adsorbent surface area and

lead to the adsorption of more adsorbate from the solution.

3.1.5. Effect of contact time and initial concentrations

To see the effect of contact time and initial concentration of Cr(VI) percentage removal

from the solution and the uptake of FNAC-700-1.5, experiments were conducted at different

initial concentrations (10, 15, 20, 25, 30 and 35 mg/L) by varying the time till 3 h., at 45 ºC

and pH 2. It can be readily observed from Fig. 10(a) that the uptake capacity of FNAC-700-1.5

to Cr(VI) drastically increased during the initial stage because of high available surface area

and vacant site of AC and then marginal increase with raising the time and reach equilibrium.

The growing trend stopped when a state of equilibrium was reached. FNAC-700-1.5 removed

a larger amount of Cr(VI) in the first 20 min of contact time, and the equilibrium was

established in 60 min for all different adsorbate concentration studied (Fig 6(a)). A large

number of vacant sites with active functional groups were available on FNAC-700-1.5 at an

early stage of adsorption for the Cr(VI). The equilibrium adsorption increased from 24.92 mg/g

to 74.95 mg/g (Fig 6(a)), which may be due to the availability of more number of Cr(VI) ions

in solution for adsorption [27]. Furthermore, higher initial adsorbate concentration provided

higher driving force to overcome all mass transfer resistances, resulting in high adsorption of

Cr(VI) [27-29]. The percentage removal of Cr(VI) was observed in reverse behaviour, as

decreased from 99.67% to 86.85% with initial concentration increased from 10 mg/L to 35

mg/L (Fig. 6(b)).

3.1.6. Adsorption kinetic studies

The study of adsorption kinetics of Cr(VI) is significant as it provides valuable insights

into the reaction pathways and the mechanism of the reactions. Any adsorption process is

normally controlled by the three diffusion steps: (i) transport of the solute from bulk solution

to the film surrounding the adsorbent, (ii) from the film to the adsorbent surface, (iii) from the

10
surface to the internal sites followed by binding of the metal ions to the active sites. The slowest

steps determine the overall rate of the adsorption process, and usually, it is thought that the step

(ii) leads to surface adsorption, and the step (iii) leads to intra-particle adsorption [12]. The

pseudo first and pseudo second order models were used to fit the experimental data of Cr(VI)

adsorption by using nonlinear regression method. To evaluate the goodness of fitting and

suitability of the model, the Coefficient of Determination (R2) and normalized standard

deviation ∆q (%) were used in the kinetic model study. A higher value of R2 and lower value

of ∆q denoted better model fitting. The standard deviation ∆q (%) was calculated as follows:

[(𝑞𝑒𝑥𝑝 − 𝑞𝑐𝑎𝑙 )/𝑞𝑒𝑥𝑝 ]2


∆q (%) = √ 100, (8)
𝑁−1

where, qexp and qcal (mg/g) are the experimental adsorption capacity and calculated adsorption

capacity, respectively, and N is the number of experimental data points.

Nonlinear and linear forms of pseudo first order equations are given in Eqs. (9) and

(10), respectively.

−𝑘1 𝑡
𝑞𝑡 = 𝑞𝑒 (1 − 𝑒𝑥𝑝𝑒 ), (9)

ln(𝑞𝑒 − 𝑞𝑡 ) = 𝑙𝑛𝑞𝑒 − 𝑘1 𝑡 , (10)

Nonlinear and linear forms of pseudo second order are given in Eqs. (11) and (12),

respectively.

𝑘2 𝑞𝑒2 𝑡
𝑞= , (11)
1 + 𝑘2 𝑞𝑒 𝑡

𝑡 1 1
= 2
+ ( )𝑡 , (12)
𝑞𝑡 𝑘2 𝑞𝑒 𝑞𝑒

where qe and qt (mg/g) are the adsorbed Cr(VI) amounts onto FNAC-700-1.5 at the equilibrium

and any time t (min), respectively and k1 (min−1) and k2 (g/min/mg) are the rate constant of the

pseudo first and pseudo second order model.

11
The study of adsorption kinetics describes the solute uptake rate, and evidently, this

rate controls the residence time of adsorbate at the solid–solution interface. The adsorption

kinetics of pseudo first order and pseudo second order for Cr(VI) onto FNAC-700-1.5 is shown

in Figs. 7 (a) and (b). The derived kinetic parameters of pseudo first and pseudo second order

model was estimated by nonlinear regression method and tabulated in Table 1. As observed,

the experimental kinetic data are better fitted by the pseudo second order model (R2 = 1 for all

Cr(VI) concentrations). Also, the calculated value (qcal) that was derived from the pseudo

second order equation is quite similar to those obtained experimentally, which indicates that

the pseudo second order model is suitable for the observed kinetics than pseudo first order

model. Moreover, all the calculated normalized standard deviation Δq (%) of pseudo first and

pseudo second order model are shown in Table 1. The resulted Δq (%) values are relatively

lower for the pseudo second order than the pseudo first order model.

3.1.7. Adsorption isotherms

At equilibrium state, the adsorption isotherm is very useful to describe how the

adsorbed molecules distribute between the liquid phase and the solid phase. The Langmuir,

Freundlich, and Temkin isotherm models were used for the adsorption isotherm. The results of

the fitting done for used models of Cr(VI) adsorption are listed in Table 2.

The Langmuir isotherm is valid for monolayer and homogeneous sites within the

adsorbent surface with a uniform distribution of energy level. The model assumes uniform

adsorption and no transmigration in the plane of the adsorbent surface [30]. The nonlinear and

linear form of the Langmuir equation is represented as follows:

𝑞𝑚 𝑘𝐿 𝐶𝑒
𝑞𝑒 = , (13)
1 + 𝑘𝐿 𝐶𝑒

𝐶𝑒 1 𝐶𝑒
= + , (14)
𝑞𝑒 𝑘𝐿 𝑞𝑚 𝑞𝑚

12
where qm (mg/g) represents the maximum adsorption capacity of the solid phase loading and

kL (L/mg) is the Langmuir constant. Figure 8(a) shows a linear relationship of C e/qe vs. Ce

using experimental data obtained for Cr(VI) adsorption. The qm and kL values are obtained

from slope and intercept of the plot and are tabulated in Table 2. Figure 8 shows experimental

data and the predicted equilibrium curve using non-linear method for the three-equilibrium

isotherms Freundlich, Langmuir and Temkin at 45 oC. The obtained isotherm parameters are

listed in Table 2.

The separation factor (RL) is dimensionless quantity and it is an essential characteristic

of the Langmuir isotherm [31] and is defined as:

1
𝑅𝐿 = , (15)
(1 + 𝑘𝐿 𝐶𝑜 )

where kL is the Langmuir constant and Co is the highest Cr(VI) concentration (mg/L). The RL

value indicates the isotherm typeto be either unfavourable (RL> 1), linear (RL = 1), favourable

(0 < RL<1) or irreversible (RL = 0). The value of RL was found to be 0.0069 for Cr(VI)

adsorption, and it confirms that the Langmuir isotherm is favorable for Cr(VI) adsorption on

the FNAC-700-1.5 under the conditions used in the present study.

The Freundlich isotherm equation is based on adsorption onto a heterogeneous surface

and given as [32]:


1⁄
𝑞𝑒 = 𝑘𝐹 + 𝐶𝑒 𝑛, (16)

where kF ((mg/g)(L/mg)1/n) is the Freundlich constant related to adsorption capacity and 1/n is

dimensionless heterogeneity factor. The linear form of Eq. (16) is

1
ln 𝑞𝑒 = ln 𝑘𝐹 + ln 𝐶𝑒 , (17)
𝑛

The plot of ln qe vs. ln Ce confirms the validity of the Freundlich model and is shown in Fig.

8(b). The value of 1/n (0.213 by linear regression and 0.305 by nonlinear regression) < 1

represents a favorable condition.

13
Temkin and Pyzhev have suggested that the heat of adsorption should decrease linearly

with the surface coverage because of the existence of adsorbate-adsorbate interactions [33].

The nonlinear and linear form of the Temkin and Pyzhev equation is represented as follows:

𝑞𝑒 = (𝑅𝑇⁄𝐵 )𝑙𝑛(𝐴𝐶𝑒 ), (18)

𝑞𝑒 = 𝐵 𝑙𝑛(𝐴) + 𝐵𝑙𝑛𝐶𝑒 , (19)

where B=RT/b is related to the heat of adsorption (L/g), and A is the dimensionless Temkin

isotherm constant. The constant A and B values are listed in Table 2.

The Coefficient of Determination (R2) values of these isotherm models are shown in Table 2.

For linear regression analysis, the Langmuir model is best fitted for Cr(VI) adsorption onto

FNAC-700-1.5. Thus, from Table 2, the comparison of tested models for the description of

Cr(VI) adsorption equilibrium isotherms on FNAC-700-1.5 is as follows: Langmuir >

Freundlich > Temkin. From Table 2, it is observed that, the isotherm parameters estimated for

different isotherms by linear and non-linear methods are entirely different. R2 values suggest

that the Freundlich isotherm was best-fitted model with the larger value of coefficient of

determination (Table 2). For nonlinear regression analysis, the comparison of tested models for

the description of Cr(VI) adsorption equilibrium isotherms on FNAC-700-1.5 is as follows:

Freundlich > Langmuir > Temkin (Table 2).

3.1.8. Comparison of the results of present study with literature ones

To justify the viability of the prepared activated carbon as effective adsorbents for

Cr(VI) removal, the adsorption capacity of FNAC on Cr(VI) were compared to the efficiency

of other low-cost adsorbents found in literature with similar batch studies. Table 3 shows a

summary of Cr(VI) removal capacity (mg/g), at optimum pH, and maximum concentration of

Cr(VI) used (mg/L) for various adsorbents in the present study and previously studied. The

adsorption capacity of Olive bagasse activated carbon (136.63 mg/g) [12] is larger than FNAC-

14
700-1.5 activated carbon (74.95 mg/g). Hence, Fox nutshell activated carbon can be considered

to be viable adsorbent for the removal of Cr(VI) from dilute solutions.

3.1.9. Thermodynamic studies

Evaluation of thermodynamic parameters is required to have the knowledge of change

in the reaction. The thermodynamic parameters such as; Gibbs free energy (∆Go) for

adsorption, change in enthalpy (∆Ho) i.e. heat of adsorption of Cr(VI) onto FNAC-700-1.5 and

change in entropy (∆So) were determined by using the following equations [16,44-45]:

∆𝐺 𝑜 = −𝑅𝑇 𝑙𝑛 𝐾𝐶 , (20)

∆S 𝑜 −∆H 𝑜
ln 𝐾𝐶 = + , (21)
𝑅 𝑅𝑇

𝐾𝐶 = (1000 × 55.5)𝐾𝐿 , (22)

where R (8.314 J/mol.K) is the universal gas constant, T (K) is the solution temperature, and

KL (L/g) is the adsorption affinity (Langmuir equilibrium constant) and is calculated as the ratio

between adsorption capacity (qe, mg/g) and equilibrium concentration (Ce, mg/L) [42-43]. KC

is the thermodynamic equilibrium constant (dimensionless) [44-45]. The values of ΔHo and

ΔSo were determined from the slope and intercept of the Van´t hoff plot of ln KC vs. 1/T (Fig.

9). The calculated thermodynamic parameter values are tabulated in Table 4. Generally, a value

of ΔGo in between 0 and -20 kJ/mol indicates physical adsorption i.e., electrostatic interaction

between adsorption sites and the adsorbing ion while a more negative ΔGo value ranging from

-80 to -400 kJ/mol indicates that the adsorption involves charge sharing or transferring from

the adsorbent surface to the adsorbing ion to form a coordinate bond (chemisorption) [46-47].

From Table 4, the negative Gibbs free energy of the experimental value indicates the adsorption

process to be spontaneous. The positive value of the enthalpy change (ΔHo = 10.29 kJ/mol)

shows that the Cr(VI) adsorption process onto FNAC-700-1.5 is endothermic in nature. The

positive value of ΔSo indicates the increased randomness at the solid-solution interface during

the fixation of the Cr(VI) on the active sites of the adsorbent.


15
3.2. Column study for Cr(VI) removal

The operating variables such as volumetric flow rate (Q), and length of the bed (L) have

a great influence on breakthrough and saturation times, and on the dynamics of the column. In

this study, the effect of these variables on the breakthrough curves of Cr(VI) onto FNAC-700-

1.5 adsorbent was investigated.

3.2.1. Effect of activated carbon bed height

Fig. 10(a) shows the Breakthrough (BT) curves obtained for Cr(VI) adsorption onto

prepared activated carbon at bed heights of 2, 3 and 4 cm, and a constant flow rate of 5 mL/min

and initial adsorbate concentration of 10 mg/L. The figure shows that both the BT and

exhaustion times increase as bed height is increased. The increase in the breakthrough time

could be ascribed to the longer distance it takes the mass transfer zone to move from the

entrance of the bed to the exit when the bed height is increased. Furthermore, higher uptake of

Cr(VI) by FNAC-700-1.5 was observed at higher bed height, which could be attributed to rising

in the surface area of the adsorbent, which provide more fixation binding sites for adsorbate to

adsorb. From Table 5, it can be noticed that the breakthrough time and exhaustion time

increased, uptake capacity decreased from 17.16-29.10 mg/g and percentage removal of Cr(VI)

from the solution increased from 58.18-71.86% as the bed height increased from 2-4 cm.

3.2.2. Effect of solution flows rate

The effect of flow rate on Cr(VI) removal by prepared FNAC-700-1.5 was studied by

varying the flow rate in the range of 5, 10 and 15 mL/min, while the bed height of 4 cm and

initial Cr(VI) concentration of 10 mg/L held constant. The breakthrough curves of both

activated carbons obtained by plotting effluent Cr(VI) concentration versus time at different

flow rates are represented in Fig. 10(b). As the flow rate is increased, the breakthrough curves

became steeper, and the slope of the breakthrough curve increased. The residence time of the

Cr(VI) in the column decreases due to flow rate increases, which leads to the early exit of the

16
Cr(VI) solution from the column. From Table 5, it can be noticed that the breakthrough time,

exhaustion time, % removal decreased and uptake capacity is increased as the flow rate is

increased. From Table 5, it can be noticed that the breakthrough time and exhaustion time

decreased. The uptake capacity increased from 29.10-36.17 mg/g and percentage removal of

Cr(VI) from the solution decreased from 71.86-58.82% as the flow rate increased from 5-15

mL/min.

4. Conclusions

In the present study, activated carbon with a well-developed pore structure of high

surface area was prepared from Fox nutshell by chemical activation with phosphoric acid and

studied for Cr(VI) adsorption from the synthetic waste water. Under the prevailing

conditions,the maximum Cr(VI) removal efficiency was found to be 99.67%. Different

thermodynamic parameters as ∆Go, ∆So and ∆Ho conforms that the adsorption was feasible,

spontaneous and endothermic in nature. The results indicate that the FNAC-700-1.5 could be

used to adsorb Cr(VI) effectively from aqueous solutions. The highest bed capacity of 36.17

mg/g was obtained using 10 mg/L initial Cr(VI) concentration at 4 cm bed height, and 15

mL/min flow rate. FNAC-700-1.5 activated carbon proved to be an effective adsorbent for

removing Cr(VI) from aqueous solutions.

References

[1] J.M. Dias, M.C. Alvim-Ferraz, M.F. Almeida, J. Rivera-Utrilla, M. Sánchez-Polo, Waste

materials for activated carbon preparation and its use in aqueous-phase treatment: a review,

Journal of Environmental Management 85 (2007) 833-846.

[2] F.C. Richard, A.C. Bourg, Aqueous geochemistry of chromium: a review, Water Research

25 (1991) 807-816.

17
[3] R. Schneider, C. Cavalin, M. Barros, C. Tavares, Adsorption of chromium ions in activated

carbon, Chemical Engineering Journal 132 (2007) 355-362.

[4] K. Selvi, S. Pattabhi, K. Kadirvelu, Removal of Cr(VI) from aqueous solution by adsorption

onto activated carbon, Bioresource technology 80 (2001) 87-89.

[5] D. Sharma, C. Forster, Column studies into the adsorption of chromium(VI) using

sphagnum moss peat, Bioresource Technology 52 (1995) 261-267.

[6] M. Kobya, Removal of Cr(VI) from aqueous solutions by adsorption onto hazelnut shell

activated carbon: kinetic and equilibrium studies, Bioresource technology 91 (2004) 317-321.

[7] A. Kumar, H.M. Jena, High surface area microporous activated carbons prepared from Fox

nut (Euryale ferox) shell by zinc chloride activation, Applied Surface Science 356 (2015) 753-

761.

[8] M. Kumar, R. Tamilarasan, Modeling studies for the removal of methylene blue from

aqueous solution using Acacia fumosa seed shell activated carbon, Journal of Environmental

Chemical Engineering 1 (2013) 1108-1116.

[9] I. Ozdemir, M. Şahin, R. Orhan, M. Erdem, Preparation and characterization of activated

carbon from grape stalk by zinc chloride activation, Fuel Processing Technology 125 (2014)

200-206.

[10] J.i. Hayashi, A. Kazehaya, K. Muroyama, A.P. Watkinson, Preparation of activated carbon

from lignin by chemical activation, Carbon 38 (2000) 1873-1878.

[11] K. Cronje, K. Chetty, M. Carsky, J. Sahu, B. Meikap, Optimization of chromium(VI)

sorption potential using developed activated carbon from sugarcane bagasse with chemical

activation by zinc chloride, Desalination 275 (2011) 276-284.

[12] H. Demiral, I. Demiral, F. Tümsek, B. Karabacakoğlu, Adsorption of chromium(VI) from

aqueous solution by activated carbon derived from olive bagasse and applicability of different

adsorption models, Chemical Engineering Journal 144 (2008) 188-196.

18
[13] A. Kumar, H.M. Jena, Preparation and characterization of high surface area activated

carbon from Fox nut (Euryale ferox) shell by chemical activation with H3PO4, Results in

Physics 6 (2016) 651-658.

[14] S. Yorgun, D. Yıldız, Preparation and characterization of activated carbons from

Paulownia wood by chemical activation with H3PO4, Journal of the Taiwan Institute of

Chemical Engineers 53 (2015) 122-131.

[15] M. Olivares-Marín, C. Fernández-González, A. Macías-García, V. Gómez-Serrano,

Thermal behaviour of lignocellulosic material in the presence of phosphoric acid. Influence of

the acid content in the initial solution, Carbon 44 (2006) 2347-2350.

[16] K. Malwade, D. Lataye, V. Mhaisalkar, S. Kurwadkar, D. Ramirez, Adsorption of

hexavalent chromium onto activated carbon derived from Leucaena leucocephala waste

sawdust: kinetics, equilibrium and thermodynamics, International Journal of Environmental

Science and Technology 13 (2016) 2107-2116.

[17] Z. Aksu, F. Gönen, Biosorption of phenol by immobilized activated sludge in a continuous

packed bed: prediction of breakthrough curves, Process biochemistry 39 (2004) 599-613.

[18] J. Salman, V. Njoku, B. Hameed, Batch and fixed-bed adsorption of 2,4-

dichlorophenoxyacetic acid onto oil palm frond activated carbon, Chemical engineering

journal 174 (2011) 33-40.

[19] A. Kumar, H.M. Jena, Removal of methylene blue and phenol onto prepared activated

carbon from Fox nutshell by chemical activation in batch and fixed-bed column, Journal of

Cleaner Production 137 (2016) 1246-1259.

[20] C. Girish, V.R. Murty, Adsorption of Phenol from Aqueous Solution Using Lantana

camara, Forest Waste: Packed Bed Studies and Prediction of Breakthrough Curves,

Environmental Processes 2 (2015) 773-796.

19
[21] B. Abussaud, H.A. Asmaly, T.A. Saleh, V.K. Gupta, M.A. Atieh, Sorption of phenol from

waters on activated carbon impregnated with iron oxide, aluminum oxide and titanium oxide,

Journal of Molecular Liquids 213 (2016) 351-359.

[22] M. Auta, B. Hameed, Chitosan–clay composite as highly effective and low-cost adsorbent

for batch and fixed-bed adsorption of methylene blue, Chemical Engineering Journal 237

(2014) 352-361.

[23] H. Deng, L. Yang, G. Tao, J. Dai, Preparation and characterization of activated carbon

from cotton stalk by microwave assisted chemical activation—application in methylene blue

adsorption from aqueous solution, Journal of Hazardous Materials 166 (2009) 1514-1521.

[24] J. Yang, M. Yu, W. Chen, Adsorption of hexavalent chromium from aqueous solution by

activated carbon prepared from longan seed: Kinetics, equilibrium and thermodynamics,

Journal of Industrial and Engineering Chemistry 21 (2015) 414-422.

[25] A. Sharma, K.G. Bhattacharyya, Adsorption of chromium(VI) on Azadirachta indica

(Neem) leaf powder, Adsorption 10 (2005) 327-338.

[26] Z.A. Al-Othman, R. Ali, M. Naushad, Hexavalent chromium removal from aqueous

medium by activated carbon prepared from peanut shell: adsorption kinetics, equilibrium and

thermodynamic studies, Chemical Engineering Journal 184 (2012) 238-247.

[27] S.S. Baral, S.N. Das, P. Rath, Hexavalent chromium removal from aqueous solution by

adsorption on treated sawdust, Biochemical Engineering Journal 31 (2006) 216-222.

[28] M.A. Islam, A. Benhouria, M. Asif, B. Hameed, Methylene blue adsorption on factory-

rejected tea activated carbon prepared by conjunction of hydrothermal carbonization and

sodium hydroxide activation processes, Journal of the Taiwan Institute of Chemical Engineers

52 (2015) 57-64.

[29] H. Cherifi, B. Fatiha, H. Salah, Kinetic studies on the adsorption of methylene blue onto

vegetal fiber activated carbons, Applied Surface Science 282 (2013) 52-59.

20
[30] I. Langmuir, The adsorption of gases on plane surfaces of glass, mica and platinum,

Journal of the American Chemical society 40 (1918) 1361-1403.

[31] K.R. Hall, L.C. Eagleton, A. Acrivos, T. Vermeulen, Pore-and solid-diffusion kinetics in

fixed-bed adsorption under constant-pattern conditions, Industrial & Engineering Chemistry

Fundamentals 5 (1966) 212-223.

[32] H. Freundlich, Over the adsorption in solution, J. Phys. Chem 57 (1906) 385-471.

[33] M. Temkin, V. Pyzhev, Recent modifications to Langmuir isotherms, (1940).

[34] N.K. Hamadi, X.D. Chen, M.M. Farid, M.G. Lu, Adsorption kinetics for the removal of

chromium(VI) from aqueous solution by adsorbents derived from used tyres and sawdust,

Chemical Engineering Journal 84 (2001) 95-105.

[35] T. Karthikeyan, S. Rajgopal, L.R. Miranda, Chromium(VI) adsorption from aqueous

solution by Hevea Brasilinesis sawdust activated carbon, Journal of hazardous materials 124

(2005) 192-199.

[36] G. Alaerts, V. Jitjaturunt, P. Kelderman, Use of coconut shell-based activated carbon for

chromium(VI) removal, Water science and technology 21 (1989) 1701-1704.

[37] G. Cimino, A. Passerini, G. Toscano, Removal of toxic cations and Cr(VI) from aqueous

solution by hazelnut shell, Water research 34 (2000) 2955-2962.

[38] F. Acar, E. Malkoc, The removal of chromium(VI) from aqueous solutions by Fagus

orientalis L, Bioresource Technology 94 (2004) 13-15.

[39] D. Sharma, C. Forster, A preliminary examination into the adsorption of hexavalent

chromium using low-cost adsorbents, Bioresource Technology 47 (1994) 257-264.

[40] S. Babel, T.A. Kurniawan, Cr(VI) removal from synthetic wastewater using coconut shell

charcoal and commercial activated carbon modified with oxidizing agents and/or chitosan,

Chemosphere 54 (2004) 951-967.

21
[41] V. Garg, R. Gupta, R. Kumar, R. Gupta, Adsorption of chromium from aqueous solution

on treated sawdust, Bioresource Technology 92 (2004) 79-81.

[42] J. Fu, Z. Chen, M. Wang, S. Liu, J. Zhang, J. Zhang, R. Han, Q. Xu, Adsorption of

methylene blue by a high-efficiency adsorbent (polydopamine microspheres): kinetics,

isotherm, thermodynamics and mechanism analysis, Chemical Engineering Journal 259 (2015)

53-61.

[43] M. Ghaedi, M.D. Ghazanfarkhani, S. Khodadoust, N. Sohrabi, M. Oftade, Acceleration of

methylene blue adsorption onto activated carbon prepared from dross licorice by ultrasonic:

Equilibrium, kinetic and thermodynamic studies, Journal of Industrial and Engineering

Chemistry 20 (2014) 2548-2560.

[44] H.N. Tran, S.J. You, H.P. Chao, Thermodynamic parameters of cadmium adsorption onto

orange peel calculated from various methods: A comparison study, Journal of Environmental

Chemical Engineering 4 (2016) 2671–2682.

[45] X. Zhou, X. Zhou, The unit problem in the thermodynamic calculation of adsorption using

the langmuir equation, Chemical Engineering Communications 201 (2014) 1459–1467.

[46] Z.A. AlOthman, M.A. Habila, R. Ali, A.A. Ghafar, M.S.E.-d. Hassouna, Valorization of

two waste streams into activated carbon and studying its adsorption kinetics, equilibrium

isotherms and thermodynamics for methylene blue removal, Arabian Journal of Chemistry 7

(2014) 1148-1158.

[47] D. Singh, Studies of the adsorption thermodynamics of oxamyl on fly ash, Adsorption

Science & Technology 18 (2000) 741-748.

Figure Caption

22
Figure captions:

Figure 1. Schematic and real photograph of a fixed bed adsorption column.

Figure 2. Effect of agitation speed on the Cr(VI) removal (C0 = 10 mg/L, temperature = 30 oC,

FNAC-700-1.5 weight = 0.02 g, contact time = 3 h).

Figure 3. Effect of pH on the Cr(VI) removal (C0 = 10 mg/L, temperature = 30 oC, FNAC-700-

1.5 weight = 0.02 g, contact time = 3 h).

Figure 4. Effect of temperature on the Cr(VI) removal (C0 = 10 mg/L, pH = 2, FNAC-700-1.5

weight = 0.02 g, contact time = 3 h).

Figure 5. Effect of FNAC-700-1.5 dosage on the Cr(VI) removal (C0 = 10 mg/L, pH = 2.0,

temperature = 45 oC, contact time = 3 h).

Figure 6. Effects of contact time on the adsorption capacity at different initial concentrations

(FNAC-700-1.5 weight = 0.04 g, pH = 2.0, temperature = 45 oC).

Figure 7. (a) Pseudo first and (b) Pseudo second order plot for the Cr(VI) adsorption by FNAC-

700-1.5 at different concentrations (FNAC-700-1.5 weight = 0.04 g, pH = 2.0; temperature =45


o
C).

Figure 8. (a) Langmuir and (b) Freundlich isotherms for the adsorption of Cr(VI) onto FNAC-

700-1.5 (FNAC-700-1.5 weight = 0.04 g, pH = 2.0; temperature =45 oC).

Figure 9. Thermodynamic study of Cr(VI) adsorption onto prepared FNAC-700-1.5 (C0 = 10

mg/L, pH = 2, FNAC-700-1.5 weight = 0.02 g, contact time = 3 h).

Figure 10. Breakthrough curves for Cr(VI) adsorption on FNAC-700-1.5 at different bed

heights and flow rates.

23
Fig.1

1. Peristaltic pump
2. Wastewater reservoir
3. Packed column (glass beads +cotton +activated carbon +cotton +glass beads)
4./5./6. Sample storing vial (at each 1 cm height of filled activated carbon)

Fig. 2

24
Fig. 3

Fig. 4

25
Fig. 5

26
Fig. 6

27
Fig. 7

28
Fig. 8

29
Fig. 9

30
Fig. 10

31
Table captions

Table 1 Kinetic constants obtained by nonlinear regression for the adsorption of Cr(VI) onto

FNAC-700-1.5

Parameters ACPA-700-1.5
Cr(VI), C0 (mg/L)
10 15 20 25 30 35
qe,exp (mg/g) 24.92 35.82 46.47 56.31 65.83 74.95

Pseudo first order model


qe,cal (mg/g) 6.74 6.50 14.30 22.71 45.19 53.30

h0 (mg/g/min) 1.03 0.54 1.9 2.575 2.28 1.925

K1 (min−1) 0.103 0.036 0.095 0.103 0.076 0.055

R2 0.987 0.994 0.992 0.989 0.993 0.992

Δq (%) 14.61 13.68 10.15 7.95 3.86 3.34


Pseudo second order model
qe,cal (mg/g) 24.94 35.46 46.73 57.47 68.49 79.05

k2 (g/(mg/min)) 0.0412 0.0318 0.0183 0.0104 0.00304 0.00205

h0 (mg/g/min) 25.63 39.98 39.96 39.47 14.26 12.81

R2 1 1 1 1 1 1

Δq (%) 0.025 0.001 0.0003 0.004 0.015 0.027

32
Table 2 Langmuir, Freundlich, and Temkin parameters for Cr(VI) adsorption onto the FNAC-

700-1.5

Freundlich Langmuir Temkin


Linear regression analysis
1/n n 2 qm (mg/g) kL (L/mg) 2 b A (L/g) 2
kF (mg/g(L/mg) ) R R R

46.99 4.7 0.897 79.5 1.51 0.949 287.19 233.26 0.775


Nonlinear regression analysis
10.85 3.28 0.999 47.62 10.5 0.994 377.53 4.4E-9 0.992

Table 3 Comparison of adsorption capacities of Cr(VI) with other adsorbents

Adsorbents Optimum C0 Maximum adsorption Reference


pH (mg/L) capacity, Qm ( mg/g)
FNAC 2 25 74.95 In this
study
Olive bagasse activated 2 500 136.63 [11]
carbon
Tyres activated carbon 2 60 58.50 [33]
F400 (CAC) 2 60 48.54 [33]
Hevea Brasilinesis - 44.05 [34]
(Rubber wood) sawdust
Leaf mould 2 1000 43.10 [5]
Coconut shell carbon 2 - 20.00 [35]
Hazelnut shell - - 17.70 [36]
Beech sawdust 1 200 16.10 [37]
Sugarcane bagasse 2 500 13.40 [38]
Coconut shell carbon 4 25 10.88 [39]
Treated sawdust of Indian 3 10 10.00 [40]
rosewood
Coconut tree sawdust 3 20 3.6 [4]

Table 4 Thermodynamic parameters for the adsorption of Cr(VI) onto FNAC-700-1.5

T(K) ΔGo (kJ/mol) ΔHo (kJ/mol) ΔSo (J/mol K)


298 -35.92 10.29 155.05
303 -36.75
308 -37.51
313 -38.30
318 -39.02

33
Table 5 Column data parameters obtained at different bed heights and flow rates
Adsorbents Bed Flow rate Breakthrough exhaustion Bed %
height (mL/min) time (min) time (min) capacity, qeq Removal
(cm) (mg/g)
FNAC-700- 2 5 420 3060 17.16 58.18
1.5 3 5 1680 3660 22.70 66.77
4 5 2520 4500 29.10 71.86
4 10 1380 3020 35.61 64.75
4 15 780 2220 36.17 58.82

34

Anda mungkin juga menyukai