Download as pdf or txt
Download as pdf or txt
You are on page 1of 42

Accepted Manuscript

Title: Adsorption of Cr(VI) from aqueous phase by high


surface area activated carbon prepared by chemical activation
with ZnCl2

Authors: Arvind Kumar, Hara Mohan Jena

PII: S0957-5820(17)30098-8
DOI: http://dx.doi.org/doi:10.1016/j.psep.2017.03.032
Reference: PSEP 1023

To appear in: Process Safety and Environment Protection

Received date: 4-8-2016


Revised date: 18-3-2017
Accepted date: 26-3-2017

Please cite this article as: Kumar, Arvind, Jena, Hara Mohan, Adsorption of
Cr(VI) from aqueous phase by high surface area activated carbon prepared
by chemical activation with ZnCl2.Process Safety and Environment Protection
http://dx.doi.org/10.1016/j.psep.2017.03.032

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 phase by high surface area activated carbon prepared

by chemical activation with ZnCl2

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)

Graphical abstract

1
Highlights

 Activated carbon (FNAC) was prepared from Fox nutshell by ZnCl2 activation.

 BET surface areaand total pore volume of FNAC were 2869 m2/g and 1.96 cm3/g.

 The maximum adsorption capacity of FNAC to Cr(VI) was 43.45 mg/g.

 Thermodynamic study revealed spontaneous and endothermic Cr(VI) adsorption.

Abstract

Prepared Activated carbon from Fox nutshell by chemical activation with ZnCl 2 in the N2

atmosphere was used for Cr(VI) removal from aqueous solution. Activated carbon was produced

at 600 oC activation temperature with a 2.0 impregnation ratio for 1h of activation time has been

found as 2869 m2/g, 1.96 and 1.68 cm3/g of highest BET surface area, total pore volume and

micropore volume, respectively. Batch mode adsorption studies were carried out by varying

agitation speed, pH, temperature, agitation time and initial Cr(VI) concentration. The adsorption

of Cr(VI) was pH dependent and showed maximum removal efficiency of Cr(VI) at pH 2.0.

Adsorption studies of 10–25 mg/L initial Cr(VI) concentration were conducted at pH of 2.0 and

temperature of 30 oC. The equilibrium, kinetics, and thermodynamics of Cr(VI) adsorption were

studied. The adsorption capacity of Cr(VI) from the synthetic wastewater was 43.45 mg/g. The

experimental adsorption equilibrium data was fitted to Langmuir adsorption model with an

adequate adsorption capacity of 46.21 mg/g.

Keywords: Activated carbons; Fox nutshell; Chemical activation; Cr(VI);Adsorption

1. Introduction

2
Chromium is a primary metal pollutant introduced into the water bodies from a variety of

industrial processes such as plastic, leather tanning, electroplating, metal processing, paint

manufacturing, steel fabrication and fertilizer, and also used in explosives, ceramics, and

photography (Selvi et al., 2001; Zhou et al., 2016). 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 (CrO42−) and dichromate (Cr2O72−). Cr(VI) possesses significantly higher

levels of toxicity than the other valency states which can cause severe diseases such as

dermatitis, kidney circulation, lung cancer and even death (Selvi et al., 2001; Sharma and

Forster, 1995; Zhou et al., 2016). The tolerance limit of Cr(VI) for discharge into inland surface

waters is 0.1 mg/L and in potable water is 0.05 mg/ L (Kobya, 2004). So, the removal of Cr(VI)

from contaminated water is important to protect the environment.

Various methods such as chemical precipitation (Zhou et al., 1993), ion exchange (Tiravanti et

al., 1997), reduction (Seaman et al., 1999), adsorption (Cronje et al., 2011; Yang et al., 2015),

solvent extraction (Pagilla and Canter, 1999), membrane separation (Chakravarti et al., 1995),

and biological method (Sahinkaya et al., 2012) have been developed for chromium removal from

wastewater. Among these methods, the adsorption technique by using activated carbon is the

most suitable method due to its efficiency; high adsorption capacity and low operational cost

(Acharya et al., 2009; Kumar and Jena, 2016b).

Adsorption process has proved to be the most efficient for the removal of heavy metals from

wastewaters (Acharya et al., 2009; Cronje et al., 2011). The high cost of adsorbents is the main

barrier to the application by the industries. The cost of adsorption technology application can be

reduced if the adsorbent is inexpensive. So there is a need to develop low-cost and easily

available adsorbents for the removal of heavy metal ions from the wastewater (Cronje et al.,

3
2011; Kumar and Jena, 2016a). Agricultural waste, biomass, and various solid waste substances

are used to prepare activated carbons for reducing the cost. In recent years, activated carbon are

being produced from agricultural products like paulownia wood (Yorgun and Yıldız, 2015),

marigold straw (Qin et al., 2014), candlenut shell (Prahas et al., 2008), corncob (Sych et al.,

2012), coconut shells (Cazetta et al., 2011), reedy grass leaves (Xu et al., 2014), lotus stalks (Liu

et al., 2013), and buriti shells (Pezoti et al., 2014).

In the present study, the prepared Fox nutshell activated carbon by chemical activation method

with ZnCl2 is used as an adsorbent for Cr(VI) removal. To the best of our knowledge, no study is

reported on the removal of hexavalent chromium from aqueous solution by the prepared Fox

nutshell activated carbon.

2. Experimental

2.1. Material

Preparation of activated carbon from Fox nutshell was carried out by chemical activation process

with a ZnCl2 activating agent. The details of the preparation of high surface area activated carbon

was published in the separate study (Kumar and Jena, 2015). The resulting Fox nutshell activated

carbon (FNAC) was kept in sealed container and used as an adsorbent for the adsorption of

Cr(VI) from synthetic wastewater. The characteristics of the prepared activated carbon are

summarized in Table 1.

2.2. Adsorbate and analytical method

A stock solution (1000 mg/L) of Cr(VI) was prepared by dissolving 2.829 g analytical grade

K2Cr2O7 in 1000 mL distilled water. Standard solutions of Cr(VI) ions concentration range from

4
10 to 25 mg/L were obtained by dilution of the stock solution. 0.1N HCl and 0.1N NaOH

solutions were used to adjust the solution pH. Some experimental variables such as agitation

speed, pH (2–7), contact time (0–3h) and concentration (10–25 mg/L) were studied to investigate

the Cr(VI) removal process. Samples at different time intervals were taken and centrifuged (2

mL) at 10,000 rpm for 5 min. The concentrations of Cr(VI) before and after adsorption was

analyzed by spectrophotometer (Jasco, Model V-530, Japan) using 1,5-diphenyl carbazide

method at 540 nm.

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 and 25 mg/L) solution with a fixed adsorbent dosage of

0.05 g. The flasks were agitated in a temperature-controlled shaker at optimized 150 rpm and 30

°C temperature for 3h study. The equilibrium adsorption capacity, qe (mg/g) and percentage

removal, R (%) were calculated by using the equations given below.

(𝐶0 − 𝐶𝑒 )𝑉
𝑞𝑒 = , (1)
𝑚𝑠

𝐶0 − 𝐶𝑒
𝑅(%) = × 100 , (2)
𝐶0

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 Ct are the initial, equilibrium, and at time t (min) of Cr(VI) concentrations

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

(g).

3. Results and discussion

5
3.1. Surface morphologies of the prepared activated carbon

Field Emission Scanning Electron Microscopy (FESEM) image of the FNAC is shown in Fig. 1.

The prepared activated carbon has well-developed pores which are of different sizes and

different shapes. Due to well-developed pores, the activated carbon possessed high BET surface

area (2869 m2/g). The pores development on the surfaces of the activated carbon resulted from

the evaporation of the activating agent of ZnCl2 during carbonization, leaving space previously

occupied by the activating agent (Deng et al., 2010; Prahas et al., 2008).

Energy dispersive X-ray spectrometer with the field emission scanning electronic microscope

(EDX) was used to determine the chemical composition of FNAC, and it is shown in Fig. 2. It

shows clearly the presence of C and O peaks as component elements of the FNAC. The result

shows that the carbon has the maximum value (89.15%). From Fig. 2, Zn is not present in the

prepared activated carbon means that 0.5N HCl washing was important step for complete

removal of Zn and released closed pores to open pores. Also, the EDX result confirms that ZnCl2

is a suitable dehydrating agent that promotes the decomposition of carbonaceous material during

the pyrolysis process for obtaining high carbon content carbon.

Transmission Electron Microscope (TEM) visualized the microporous network of the FNAC.

The TEM image is shown in Fig. 3. It clearly shows the uniform pore distribution of micropores

like honeycomb due to which the prepared activated carbon attained high micropore volume.

From Fig. 3, the average pore diameter of the micropore present on the FNAC surface is 0.64

nm.

3.2. Cr(VI) adsorption onto FNAC

3.2.1. Effect of agitation speed

6
Figure 4 shows the effect of agitation speed on Cr(VI) removal was examined in the range of 90–

170 rpm. The results reveal that the Cr(VI) removal efficiency did not change above the agitation

speed of 150 rpm. This result indicates that an effective transport of Cr(VI) ions towards the

adsorbent surfaces occurred due to a decrease in boundary layer thickness around the adsorbent

particles by increasing agitation speed (Gupta et al., 2011). Hence, for the further experiments,

the agitation speed of 150 rpm was selected as the optimal mixing speed.

3.2.2. Effect of pH

The pH of the system determines the adsorption capacity due to its influence on the surface

properties of the FNAC and different ionic forms of the chromium solutions. Hexavalent

chromium, Cr(VI) exists in different ionic forms in aqueous solution and the stability of these

forms, such as HCrO4−, CrO42− or Cr2O72−, depend on the pH of the aqueous solution and

chromate concentration. The following are the equilibrium reactions and corresponding pKa

values reported in few literature for different Cr(VI) species.

𝐻2 𝐶𝑟𝑂4 ↔ 𝐻 + + 𝐻𝐶𝑟𝑂4− (𝑖)

pKa = 0.8 (Sengupta and Clifford, 1986), 0.75 (Kotaś and Stasicka, 2000), -0.26 (Cox, 2000)

𝐻𝐶𝑟𝑂4− ↔ 𝐻 + + 𝐶𝑟𝑂42− (𝑖𝑖)

pKa = 6.5 (Sengupta and Clifford, 1986), 6.45 (Kotaś and Stasicka, 2000), 5.9 (Ajouyed et al.,

2010; Cox, 2000)

2𝐻𝐶𝑟𝑂4− ↔ 𝐶𝑟2 𝑂72− + 𝐻2 𝑂 (𝑖𝑖𝑖)

pKa = -1.52 (Sengupta and Clifford, 1986), -2.2 (Ajouyed et al., 2010; Cox, 2000; Kotaś and

Stasicka, 2000)

𝐻𝐶𝑟2 𝑂7− = 𝐶𝑟2 𝑂7− + 𝐻 + (𝑖𝑣)

pKa = -0.07 (Sengupta and Clifford, 1986), -0.85 (Cox, 2000)

7
The reaction (iii) does not contain any H+ terms, i.e., in a certain pH range (2-5) this reaction is

independent of pH and depends only on total Cr(VI) concentration. This may be regarded as a

dimerization reaction for HCrO4− at acidic pH. In the pH ranging from 2.0 to 6.0, HCrO4− ions

mainly exist in equilibrium and the predominant form of HCrO4− shifts to chromate ion (CrO42−)

as pH increases. Above pH 7 only CrO42− ions exist in solution throughout the concentration

range and in the pH between 1.0 and 6.0, HCrO4− is the predominant form up to the Cr(VI)

concentration 10−2 M then it starts to condense yielding the orange-red dichromate ion. HCrO4−

ion only needs one active site whereas chromate ion (CrO42−) needs two active sites due to its

two minus charges (Yang et al., 2014). The behaviour 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 hydroxyl group (−OH) on the adsorbent surface, thereby

reducing hindrance to the diffusion of chromate ions (Demiral et al., 2008). At higher pH values,

the reduction in adsorption may be possible due to the abundance of OH− ions causing increased

hindrance to diffusion of dichromate ions.

Figure 5 shows the influence of solution pH on Cr(VI) adsorption by FNAC in the pH range of

2.0 to 7.0. From Fig. 5, the removal of Cr(VI) from aqueous solution is greatly influenced by the

pH of the solution, and the maximum removal of 85.35 % occurs at pH 2.0. The Cr(VI)

adsorption decreases from 2.0 to 7.0. This result indicates that the adsorption of the adsorbent is

clearly pH dependent. Other investigators have also reported similar observations (El-Sikaily et

al., 2007; Hamadi et al., 2001; Karthikeyan et al., 2005; Sharma and Forster, 1994; Yang et al.,

2015).

3.2.3. Effect of adsorbent dose

8
The effect of the AC dosage on the adsorption of Cr(VI) was studied. The percentage removal of

Cr(VI) varied linearly with the amount of the adsorbent as shown in Fig. 6. With the increase of

adsorbent dosage, the time needed to reach equilibrium is reduced due to the increase of efficient

adsorption sites at higher dosages. FNAC dosage was ranged from 0.01 to 0.07 g/100 mL of

Cr(VI) solution and equilibrated for 3h. From Fig. 6, the percentage uptake capacity of Cr(VI) is

increased with the increase in adsorbent dose up to 0.05 g significantly and after that remains

unchanged. Thus, to get the better Cr(VI) removal, 0.05 g was chosen as an optimal mass of the

adsorbent for the further experiments.

3.2.4. Effect of contact time and initial Cr(VI) concentrations

Figure 7 displays the impact of contact time and initial concentrations of the solution. It can be

readily observed that the adsorption capacity of Cr(VI) on FNAC drastically increased during the

initial stage and then at a slow speed. The growing trend stopped when a state of equilibrium was

reached. FNAC is removed a larger amount of Cr(VI) in the first 20 min of contact time, and the

equilibrium is established in 60 min for all different absorbent concentration studied [Fig 7(a)]. It

is clear that the adsorption of Cr(VI) ion is rather quick because it reaches its maximum removal

in 60 min and after that, no more adsorption occurs. A large number of vacant sites with active

functional groups were available on FNAC at an early stage of adsorption for the Cr(VI). The

equilibrium adsorption increases from 19.18 mg/g to 43.45 mg/g while % removal of Cr(VI)

observed in reverse behaviour, as decreases from 99.08 to 86.89% when initial concentrations

are increased from 10 mg/L to 25 mg/L [Fig. 7(b)].

3.2.5. 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

9
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 (Langmuir, 1918). The linear form of the

Langmuir equation is represented as follows:

𝐶𝑒 1 𝐶𝑒
= + , (4)
𝑞𝑒 𝑘𝐿 𝑞𝑚 𝑞𝑚

where qm represents the maximum adsorption capacity of the solid phase loading, and kL (L/mg)

is the Langmuir constant. Fig. 8 shows a linear relationship of Ce/qe versus 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.

The Freundlich isotherm equation is based on sorption onto a heterogeneous surface and given as

(Freundlich, 1906):
1⁄
𝑞𝑒 = 𝑘𝐹 + 𝐶𝑒 𝑛, (6)

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. (6) is

1
ln 𝑞𝑒 = ln 𝑘𝐹 + ln 𝐶𝑒 , (7)
𝑛

A linear plot of ln qe versus ln Ce confirms the validity of the Freundlich model and is shown in

Fig. 9. The value of n (4.66) > 1 represents a favorable condition.

Tempkin and Pyzhev have suggested that the heat of adsorption should decrease linearly with the

surface coverage because of the existence of adsorbate-adsorbate interactions (Tempkin and

Pyzhev, 1940). The following equation can adjust the corresponding adsorption isotherm:

10
𝑞𝑒 = 𝐵 𝑙𝑛(𝐴) + 𝐵 𝑙𝑛𝐶𝑒 , (8)

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

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

The R2 values of these isotherm models are shown in Table 2. In this study, the Langmuir model

is best fitted for Cr(VI) adsorption onto Fox nutshell AC. Thus, from Table 2, the comparison of

tested models for the description of Cr(VI) adsorption equilibrium isotherms on the Fox nutshell

AC is as follows: Langmuir > Freundlich > Tempkin.

3.2.6. Adsorption kinetic studies

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

insights into the reaction pathways and the mechanism of the reactions. Three diffusion steps

usually control any adsorption process: (i) transport of the solute from the bulk solution to the

film surrounding the adsorbent, (ii) from the film to the adsorbent surface, (iii) from the 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 (Demiral et al.,

2008). In the present study, the most used kinetic models of pseudo-first-order (Budinova et al.,

2006) and pseudo-second-order (Ho and McKay, 1999) were used to fit the experimental data.

Linear forms of pseudo-first-order and pseudo-second-order kinetic equations are given in Eqs.

(9) and (10), respectively.

ln(𝑞𝑒 − 𝑞𝑡 ) = 𝑙𝑛𝑞𝑒 − 𝑘1 𝑡 , (9)

𝑡 1 1
= + ( )𝑡 , (10)
𝑞𝑡 𝑘2 𝑞𝑒2 𝑞𝑒

11
where qe and qt (mg/g) are the adsorbed Cr(VI) amounts onto FNAC at the equilibrium and at

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

pseudo-first-order and pseudo-second-order adsorption.

To evaluate the goodness of fitting and suitability of the model, the linear correlation coefficient

(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 from given equation (11).

[(𝑞𝑒𝑥𝑝 − 𝑞𝑐𝑎𝑙 )/𝑞𝑒𝑥𝑝 ]2


∆q (%) = √ × 100, (11)
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.

The adsorption kinetics of pseudo-first-order and pseudo-second-order for Cr(VI) onto FNAC is

shown in Fig. 10 and 11. The derived kinetic parameters of pseudo-first-order and pseudo-

second-order are listed in Table 3. As observed, the experimental kinetic data are better fitted by

the pseudo-second-order model (R2 = 0.999 for all Cr(VI) concentrations). Also, the calculated

value (qe, cal = 43.99 mg/g) that was derived from the second-order equation is quite similar to

those obtained experimentally (qe, exp = 43.45 mg/g), which indicates that the second-order model

is suitable for the observed kinetics. Moreover, all the determined normalized standard deviation

Δq (%) of pseudo-first-order and pseudo-second-order are shown in Table 3. The resulted Δq

(%) values are relatively lower for the pseudo-second-order kinetic model than the pseudo-first-

order kinetic model.

12
The diffusion mechanismcan not identifyeither bythe pseudo-first-order or the pseudo-second-

order kinetic model. The intraparticle diffusion varies with the square root of time is given by

(Weber and Morris, 1963) as:

𝑞𝑡 = 𝑘𝑖 𝑡 0.5 + 𝑐 , (12)

whereki is the intraparticle diffusion rate constant (mg g−1 min−1/2), and c is the intercept. The

constant (ki) for Cr(VI) adsorption was calculated and are tabulated in Table 3 and plot of qt vs

t0.5 is shown in Fig. 12. From Fig. 12, the plots are not linear over the whole time range, that

confirms more than one process affect the Cr(VI) adsorption (Sharma and Bhattacharyya, 2005).

The plot shows two portions: the first straight portion depicting macropore and mesopore

diffusion and the second representing micropore diffusion (Fierro et al., 2008). If the value of c

is zero, then the rate of adsorption is controlled by intraparticle diffusion for the entire adsorption

period but in the present study, c is not zero which confirms more than one process affect the

Cr(VI) adsorption process onto the FNAC.

3.2.7. Thermodynamic studies

The measurement of thermodynamic parameters such as Gibbs free energy (∆G o), change in

enthalpy (∆Ho) and change in entropy (∆So) has been vital in the adsorption studies. Its original

concept assumes that the energy cannot be gained or lost, where entropy change is the driving

force. The values of ΔHo, ΔGo and ΔSo were calculated according to the following equations:

𝑞𝑒(𝑊⁄𝑉)
𝐾𝐶 = , (13)
𝐶𝑒

∆𝐺 o = −𝑅𝑇𝑙𝑛𝐾𝐶 , (14)

∆𝑆 𝑜 ∆𝐻 𝑜
𝑙𝑛𝐾𝐶 = − , (15)
𝑅 𝑅𝑇

13
where R (8.314 J/mol.K) is the universal gas constant, T (K) is the absolute solution temperature,

and KC is the Langmuir isotherm constant. The values of ΔHo and ΔSo were determined from the

slope and intercept of the van't Hoff plot of ln KC versus 1/T (Fig. 13). Thermodynamic

parameters obtained from Fig. 13 are tabulated in Table 4. From Table 4, the negative Gibbs free

energy (ΔGo) of the experimental value indicates a typical physical process. The positive value

of the enthalpy change (ΔHo = 14.74 kJ/mol) shows that the Cr(VI) adsorption process onto

FNAC is endothermic in nature. The positive value of ΔSo indicates an increase in the degree of

freedom (or disorder) of the adsorbed Cr(VI) onto FNAC means adsorption is increased with

increasing temperature.

3.3. Comparison to some other adsorbents

A comparison of the maximum adsorption capacities of the FNAC with other reported values for

some adsorbents is listed in Table 5. From Table 5, the FNAC seems to be an alternative

precursor for the commercial AC preparation, in which its sorption capacity of Cr(VI) is higher

than others agricultural waste-based carbonaceous adsorbents. FNAC could be employed as an

efficient carbonaceous adsorbent to compare with the commercial ones for the removal Cr(VI)

from wastewater.

4. Conclusions

In the present study, prepared activated carbon of high surface area with a well-developed pore

structure was used for Cr(VI) adsorption from aqueous solution. The results illustrated that the

prepared adsorbent showed a high efficiency in adsorption of Cr(VI). The optimum pH was 2 at

which the Cr(VI) removal was maximum. The maximum Cr(VI) removal efficiency was found

to be 99.08% of 10 mg/L concentration at pH 2 and temperature of 30 °C for 3h study. Different

14
thermodynamic parameters as ∆Go, ∆So and ∆Ho were evaluated and concluded that the

adsorption was feasible, spontaneous and endothermic in nature. Due to favorable performance

of FNAC in the removal of Cr(VI) from the aqueous solutions, it can be used as an efficient

adsorbent in the treatment of water and wastewater with no need for further filtering and

centrifugation, etc., and also it could be utilized as an alternative to commercial activated

carbons.

15
References

Acar, F., Malkoc, E., 2004. The removal of chromium(VI) from aqueous solutions by Fagus

orientalis L. Bioresource Technology 94, 13-15.

Acharya, J., Sahu, J., Mohanty, C., Meikap, B., 2009. Removal of lead(II) from wastewater by

activated carbon developed from Tamarind wood by zinc chloride activation. Chemical

Engineering Journal 149, 249-262.

Ajouyed, O., Hurel, C., Ammari, M., Allal, L.B., Marmier, N., 2010. Sorption of Cr(VI) onto

natural iron and aluminum (oxy) hydroxides: effects of pH, ionic strength and initial

concentration. Journal of Hazardous Materials 174, 616-622.

Alaerts, G., Jitjaturunt, V., Kelderman, P., 1989. Use of coconut shell-based activated carbon for

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

Babel, S., Kurniawan, T.A., 2004. Cr(VI) removal from synthetic wastewater using coconut shell

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

Chemosphere 54, 951-967.

Budinova, T., Ekinci, E., Yardim, F., Grimm, A., Björnbom, E., Minkova, V., Goranova, M.,

2006. Characterization and application of activated carbon produced by H3PO4 and water vapor

activation. Fuel Processing Technology 87, 899-905.

Cazetta, A.L., Vargas, A.M., Nogami, E.M., Kunita, M.H., Guilherme, M.R., Martins, A.C.,

Silva, T.L., Moraes, J.C., Almeida, V.C., 2011. NaOH-activated carbon of high surface area

produced from coconut shell: Kinetics and equilibrium studies from the methylene blue

adsorption. Chemical Engineering Journal 174, 117-125.

16
Chakravarti, A., Chowdhury, S., Chakrabarty, S., Chakrabarty, T., Mukherjee, D., 1995. Liquid

membrane multiple emulsion process of chromium(VI) separation from waste waters. Colloids

and Surfaces A: Physicochemical and Engineering Aspects 103, 59-71.

Cimino, G., Passerini, A., Toscano, G., 2000. Removal of toxic cations and Cr(VI) from aqueous

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

Cox, P., 2000. Advanced Inorganic Chemistry, by FA Cotton, G. Wilkinson, CA Murillo and M.

Bochmann, Wiley, Chichester, 1999. xv+ 1355 pp., ISBN 0-471-19957-5.£ 58.50. Elsevier.

Cronje, K., Chetty, K., Carsky, M., Sahu, J., Meikap, B., 2011. Optimization of chromium(VI)

sorption potential using developed activated carbon from sugarcane bagasse with chemical

activation by zinc chloride. Desalination 275, 276-284.

Demiral, H., Demiral, I., Tümsek, F., Karabacakoğlu, B., 2008. Adsorption of chromium(VI)

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

different adsorption models. Chemical Engineering Journal 144, 188-196.

Deng, H., Zhang, G., Xu, X., Tao, G., Dai, J., 2010. Optimization of preparation of activated

carbon from cotton stalk by microwave assisted phosphoric acid-chemical activation. Journal of

Hazardous Materials 182, 217-224.

El-Sikaily, A., El Nemr, A., Khaled, A., Abdelwehab, O., 2007. Removal of toxic chromium

from wastewater using green alga Ulva lactuca and its activated carbon. Journal of Hazardous

Materials 148, 216-228.

Fierro, V., Torné-Fernández, V., Montané, D., Celzard, A., 2008. Adsorption of phenol onto

activated carbons having different textural and surface properties. Microporous and mesoporous

materials 111, 276-284.

Freundlich, H., 1906. Over the adsorption in solution. J. Phys. Chem 57, e470.

17
Garg, V., Gupta, R., Kumar, R., Gupta, R., 2004. Adsorption of chromium from aqueous

solution on treated sawdust. Bioresource Technology 92, 79-81.

Gottipati, R., Mishra, S., 2016. Preparation of microporous activated carbon from Aegle

marmelos fruit shell and its application in removal of chromium(VI) from aqueous phase.

Journal of Industrial and Engineering Chemistry 36, 355-363.

Gupta, V., Agarwal, S., Saleh, T.A., 2011. Chromium removal by combining the magnetic

properties of iron oxide with adsorption properties of carbon nanotubes. Water research 45,

2207-2212.

Hamadi, N.K., Chen, X.D., Farid, M.M., Lu, M.G., 2001. Adsorption kinetics for the removal of

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

Chemical Engineering Journal 84, 95-105.

Ho, Y.-S., McKay, G., 1999. Pseudo-second order model for sorption processes. Process

biochemistry 34, 451-465.

Karthikeyan, T., Rajgopal, S., Miranda, L.R., 2005. Chromium(VI) adsorption from aqueous

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

192-199.

Kobya, M., 2004. Removal of Cr(VI) from aqueous solutions by adsorption onto hazelnut shell

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

Kotaś, J., Stasicka, Z., 2000. Chromium occurrence in the environment and methods of its

speciation. Environmental pollution 107, 263-283.

Kumar, A., Jena, H.M., 2015. High surface area microporous activated carbons prepared from

Fox nut (Euryale ferox) shell by zinc chloride activation. Applied Surface Science 356, 753-761.

18
Kumar, A., Jena, H.M., 2016a. Preparation and characterization of high surface area activated

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

Physics.

Kumar, A., Jena, H.M., 2016b. 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, 1246-1259.

Langmuir, I., 1918. The adsorption of gases on plane surfaces of glass, mica and platinum.

Journal of the American Chemical society 40, 1361-1403.

Liu, H., Zhang, J., Zhang, C., Bao, N., Cheng, C., 2013. Activated carbons with well-developed

microporosity and high surface acidity prepared from lotus stalks by organophosphorus

compounds activations. Carbon 60, 289-291.

Pagilla, K.R., Canter, L.W., 1999. Laboratory studies on remediation of chromium-contaminated

soils. Journal of Environmental Engineering 125, 243-248.

Pezoti, O., Cazetta, A.L., Souza, I.P., Bedin, K.C., Martins, A.C., Silva, T.L., Almeida, V.C.,

2014. Adsorption studies of methylene blue onto ZnCl2-activated carbon produced from buriti

shells (Mauritia flexuosa L.). Journal of Industrial and Engineering Chemistry 20, 4401-4407.

Prahas, D., Kartika, Y., Indraswati, N., Ismadji, S., 2008. Activated carbon from jackfruit peel

waste by H3PO4 chemical activation: pore structure and surface chemistry characterization.

Chemical Engineering Journal 140, 32-42.

Qin, C., Chen, Y., Gao, J.-m., 2014. Manufacture and characterization of activated carbon from

marigold straw (Tagetes erecta L) by H3PO4 chemical activation. Materials Letters 135, 123-

126.

19
Sahinkaya, E., Altun, M., Bektas, S., Komnitsas, K., 2012. Bioreduction of Cr(VI) from acidic

wastewaters in a sulfidogenic ABR. Minerals Engineering 32, 38-44.

Seaman, J.C., Bertsch, P.M., Schwallie, L., 1999. In situ Cr(VI) reduction within coarse-

textured, oxide-coated soil and aquifer systems using Fe(II) solutions. Environmental Science &

Technology 33, 938-944.

Selvi, K., Pattabhi, S., Kadirvelu, K., 2001. Removal of Cr(VI) from aqueous solution by

adsorption onto activated carbon. Bioresource technology 80, 87-89.

Sengupta, A.K., Clifford, D., 1986. Some unique characteristics of chromate ion exchange.

Reactive Polymers, Ion Exchangers, Sorbents 4, 113-130.

Sharma, A., Bhattacharyya, K.G., 2005. Adsorption of chromium(VI) on Azadirachta indica

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

Sharma, D., Forster, C., 1994. A preliminary examination into the adsorption of hexavalent

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

Sharma, D., Forster, C., 1995. Column studies into the adsorption of chromium(VI) using

sphagnum moss peat. Bioresource Technology 52, 261-267.

Sych, N., Trofymenko, S., Poddubnaya, O., Tsyba, M., Sapsay, V., Klymchuk, D., Puziy, A.,

2012. Porous structure and surface chemistry of phosphoric acid activated carbon from corncob.

Applied Surface Science 261, 75-82.

Temkin, M., Pyzhev, V., 1940. Recent modifications to Langmuir isotherms.

Tiravanti, G., Petruzzelli, D., Passino, R., 1997. Pretreatment of tannery wastewaters by an ion

exchange process for Cr(III) removal and recovery. Water Science and Technology 36, 197-207.

Weber, W.J., Morris, J.C., 1963. Kinetics of adsorption on carbon from solution. Journal of the

Sanitary Engineering Division 89, 31-60.

20
Xu, J., Chen, L., Qu, H., Jiao, Y., Xie, J., Xing, G., 2014. Preparation and characterization of

activated carbon from reedy grass leaves by chemical activation with H3PO4. Applied Surface

Science 320, 674-680.

Yang, J., Yu, M., Chen, W., 2015. 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, 414-422.

Yang, J., Yu, M., Qiu, T., 2014. Adsorption thermodynamics and kinetics of Cr(VI) on KIP210

resin. Journal of Industrial and Engineering Chemistry 20, 480-486.

Yorgun, S., Yıldız, D., 2015. Preparation and characterization of activated carbons from

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

Engineers 53, 122-131.

Zhou, J., Wang, Y., Wang, J., Qiao, W., Long, D., Ling, L., 2016. Effective removal of

hexavalent chromium from aqueous solutions by adsorption on mesoporous carbon

microspheres. Journal of colloid and interface science 462, 200-207.

Zhou, X., Korenaga, T., Takahashi, T., Moriwake, T., Shinoda, S., 1993. A process

monitoring/controlling system for the treatment of wastewater containing chromium(VI). Water

Research 27, 1049-1054.

21
Figure captions:

Figure 1. FESEM of FNAC.

Figure 2. EDX spectrum of FNAC.

Figure 3. TEM image of FNAC.

Figure 4. Effect of agitation speed on the removal of Cr(VI) onto FNAC (C0 = 10 mg/L, FNAC

weight = 0.02 g, pH= 2, temperature = 30 oC, and contact time (t) = 3h).

Figure 5. Effect of pH on the Cr(VI) removal onto FNAC (C0 = 10 mg/L, FNAC weight = 0.02

g, temperature = 30 oC, and contact time (t) = 3h).

Figure 6. Effect of adsorbent dosage of FNAC on Cr(VI) removal (C0 = 10 mg/L, pH = 2.0,

contact time (t) = 3h, temperature = 30oC).

Figure 7. Effects of contact time on the adsorption capacity at different initial concentrations

(FNAC weight = 0.05 g, C0= 10 mg/L, pH = 2.0, temperature = 30 oC).

Figure 8. Langmuir isotherm for the adsorption of Cr(VI) onto FNAC.

Figure 9. Freundlich isotherm for the adsorption of Cr(VI) onto FNAC.

Figure 10. Pseudo-first order kinetic plot for the Cr(VI) adsorption onto FNAC (FNAC weight =

0.05 g, pH 2.0, and temperature = 30oC).

Figure 11. Pseudo-second order kinetic plot for the Cr(VI) adsorption onto FNAC (FNAC weight

= 0.05 g, pH 2.0, and temperature = 30 oC).

Figure 12. Intraparticle diffusion kinetic plot for the Cr(VI) adsorption by FNAC (FNAC weight

= 0.05 g, pH 2.0, and temperature = 30 oC).

Figure 13. Thermodynamic study of Cr(VI) adsorption onto FNAC.

22
Fig. 1

Fig. 2

23
Fig. 3

24
Fig. 4

25
Fig. 5

26
Fig. 6

27
Fig. 7

28
29
Fig. 8

30
Fig. 9

31
Fig. 10

32
Fig. 11

33
Fig. 12

34
Fig. 13

35
36
Table captions

Table 1 Characteristics of FNAC.

Proximate Activated carbon Ultimate Activated carbon Pore structure


analysis (wt%) Analysis (wt%) characteristics

Moisture 2.21 C 89.77 SBET 2,869


(m2/g)
Volatile matter 19.05 H 2.33 VT (cm3/g) 1.96

Ash 1.02 N 0.43 Vmicro(cm3 1.68


/g)
Fixed carbona 77.72 S 0.06 Vmeso(cm3/ 0.28
g)
a
by difference Oa 7.41 Vmicro/VT 85.71
(%)
SBET: BET surface area, Smicro: micropore surface area, Smeso: mesopore Dp (nm) 2.75
surface area, VT: total pore volume, Vmicro: micropore volume, Vmeso:
mesopore volume, microporosity = (Vmicro/VT) × 100, Dp: average diameter.

37
Table 2 Langmuir, Freundlich, and Tempkin parameters of Cr(VI) adsorption onto FNAC.

Freundlich Langmuir Tempkin


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

32.10 4.66 0.960 46.21 3.02 0.971 394.64 201.14 0.890

38
Table 3 Kinetic constants obtained for the adsorption of Cr(VI) adsorption onto FNAC.

Parameters Cr(VI), C0 (mg/L)


10 15 20 25
qe,exp (mg/g) 19.18 28.38 36.39 43.45
Pseudo-first-order
qe,cal (mg/g) 1.01 4.98 7.18 11.46
h0 (mg/g/min) 0.431 0.705 1.39 1.18
K1(min−1) 0.0431 0.047 0.0693 0.0472
R2 0.570 0.898 0.978 0.967
Δq (%) 31.58 27.48 26.76 24.54
Pseudo-second-order
qe,cal (mg/g) 19.93 28.48 36.58 43.99
k2 (g/(mg min)) 0.0705 0.039 0.027 0.012
h0 (mg/g/min) 28.00 31.63 36.13 23.22
R2 0.9999 0.9999 1 0.9999
Δq (%) 1.30 0.12 0.17 0.41
Intraparticle diffusion
-1/2
Ki (mg/g min ) 0.372 0.436 0.571 1.355
2
R 0.424 0.229 0.493 0.610

C 16.24 23.88 30.68 29.83

39
Table 4 Thermodynamic parameters for the adsorption of Cr(VI) onto FNAC.

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


298 -7.59
303 -7.99
308 -8.28 14.74 0.075
313 -8.64
318 -9.14

40
Table 5 The adsorptive capacities of various adsorbents for Cr(VI).

Adsorbents Optimum Max. Cr. concentration Qm ( Reference


pH used ( mg/L) mg/g)
Tyres activated carbon 2 60 58.50 (Hamadi et al.,
2001)
F400 (CAC) 2 60 48.54 (Hamadi et al.,
2001)
Hevea Brasilinesis 2 200 44.05 (Karthikeyan et al.,
sawdust AC 2005)
Leaf mould 2 1000 43.10 (Sharma and
Forster, 1995)
Coconut shell carbon 2 - 20.00 (Alaerts et al., 1989)
Hazelnut shell 2 30 17.70 (Cimino et al.,
2000)
Beech sawdust 1 200 16.10 (Acar and Malkoc,
2004)
Sugarcane bagasse 2 500 13.40 (Sharma and
Forster, 1994)
Coconut shell carbon 4 25 10.88 (Babel and
Kurniawan, 2004)
Treated sawdust of 3 10 10.00 (Garg et al., 2004)
Indian Rosewood
Aegle Marmelos fruit 2 10 4.27 (Gottipati and
shell Mishra, 2016)
Coconut tree sawdust 3 20 3.6 (Selvi et al., 2001)
FNAC 2 25 43.45 In this study

41

You might also like