equilibrium, kinetics and thermodynamics study of phenols

16
Equilibrium, kinetics and thermodynamics study of phenols adsorption onto activated carbon obtained from lignocellulosic material (Eucalyptus Globulus labill seed) Nelson Giovanny Rinco ´n-Silva 1 Juan C. Moreno-Piraja ´n 1 Liliana Giraldo 2 Received: 30 August 2015 / Revised: 11 November 2015 / Accepted: 20 November 2015 Ó Springer Science+Business Media New York 2015 Abstract Activated carbon was prepared from lignocel- lulosic material (Eucalyptus Globulus labill seed) by chemical activation with ZnCl 2 at two different concen- trations (10 and 25 % m/v) named ACS25 and ACS10. The textural characteristics of the activated carbons (ACs) were determined by N 2 adsorption isotherms; these exhibit B.E.T. surface areas of 250 and 300 m 2 g -1 for ACS25 and ACS10, respectively, with micropore volume contents of 0.140 and 0.125 cm 3 g -1 in the same order. In addition, the FTIR and Boehm methods were conducted for the chemi- cal characterisation of ACs, where many groups with basic character were found, which favours the adsorption of phenols. The prepared carbonaceous adsorbents were used in the adsorption of wide pollutants monosubstituted phe- nol derivatives: phenol, 4-nitrophenol and 4-chlorophenol. The effect of temperature on the thermodynamics, kinetic and equilibrium of phenols adsorption on ACs was thor- oughly examined. The adsorption kinetics adjusted prop- erly for a pseudo-second-order kinetic model. However, the Elovich model (chemisorption) confirms that phenols adsorption did not occur via the sharing of electrons between the phenolic ring and basal plane of ACs because is not properly adjusted, so the process is given by physisorption. The thermodynamic parameters [i.e. Gibbs free energy change (DG°), enthalpy change (DH°) and entropy change (DS°)] were also evaluated. The overall adsorption process was exothermic and spontaneous in nature. The values found in the thermodynamic study, confirm that the adsorption process corresponds to a clearly physical process. Keywords Activated carbon Kinetic Thermodynamics 4-Chlorophneol Elovich equation Abbreviations Q t Amount of adsorbate adsorbed at time t (mg g -1 ) K f The pseudo-first-order rate constant (h -1 ), t Time K s The pseudo-first-order rate constant (g gm -1 min -1 ) a Desorption constant in Elovich Equation b Initial adsorption rate in Elovich Equation k id The intraparticle diffusion rate constant (mg g -1 h -1/2 ) I Thickness of the boundary layer in intraparticle model of Weber and Morris k Rate constant of adsorption of Dumwald–Warner model (min -1 ) F Fractional attainment of equilibrium in the Dumwald-Warner model k fd Film diffusion rate coefficient in the Dumwald– Wagner mode Q max Maximum phenol uptake in Langmuir model k L Constant in Langmuir model that denoted the energy of adsorption and affinity of the binding sites (L mg -1 ) k f Constant in Freundlich model giving an indication of adsorption capacity [mg g -1 (L mg -1 ) n ] & Liliana Giraldo [email protected] 1 Departamento de Quı ´mica, Facultad de Ciencias, Grupo de Investigacio ´ n de So ´ lidos Porosos y Calorimetrı ´a, Universidad de los Andes, Bogota ´, Colombia 2 Departamento de Quı ´mica, Facultad de Ciencias, Universidad Nacional de Colombia, Bogota ´, Colombia 123 Adsorption DOI 10.1007/s10450-015-9724-2

Upload: nelson-giovanny-rincon-silva

Post on 26-Jan-2017

414 views

Category:

Engineering


6 download

TRANSCRIPT

Page 1: Equilibrium, kinetics and thermodynamics study of phenols

Equilibrium, kinetics and thermodynamics study of phenolsadsorption onto activated carbon obtained from lignocellulosicmaterial (Eucalyptus Globulus labill seed)

Nelson Giovanny Rincon-Silva1 • Juan C. Moreno-Pirajan1 • Liliana Giraldo2

Received: 30 August 2015 / Revised: 11 November 2015 / Accepted: 20 November 2015

� Springer Science+Business Media New York 2015

Abstract Activated carbon was prepared from lignocel-

lulosic material (Eucalyptus Globulus labill seed) by

chemical activation with ZnCl2 at two different concen-

trations (10 and 25 % m/v) named ACS25 and ACS10. The

textural characteristics of the activated carbons (ACs) were

determined by N2 adsorption isotherms; these exhibit

B.E.T. surface areas of 250 and 300 m2 g-1 for ACS25 and

ACS10, respectively, with micropore volume contents of

0.140 and 0.125 cm3 g-1 in the same order. In addition, the

FTIR and Boehm methods were conducted for the chemi-

cal characterisation of ACs, where many groups with basic

character were found, which favours the adsorption of

phenols. The prepared carbonaceous adsorbents were used

in the adsorption of wide pollutants monosubstituted phe-

nol derivatives: phenol, 4-nitrophenol and 4-chlorophenol.

The effect of temperature on the thermodynamics, kinetic

and equilibrium of phenols adsorption on ACs was thor-

oughly examined. The adsorption kinetics adjusted prop-

erly for a pseudo-second-order kinetic model. However, the

Elovich model (chemisorption) confirms that phenols

adsorption did not occur via the sharing of electrons

between the phenolic ring and basal plane of ACs because

is not properly adjusted, so the process is given by

physisorption. The thermodynamic parameters [i.e. Gibbs

free energy change (DG�), enthalpy change (DH�) and

entropy change (DS�)] were also evaluated. The overall

adsorption process was exothermic and spontaneous in

nature. The values found in the thermodynamic study,

confirm that the adsorption process corresponds to a clearly

physical process.

Keywords Activated carbon � Kinetic �Thermodynamics � 4-Chlorophneol � Elovich equation

Abbreviations

Qt Amount of adsorbate adsorbed at time t (mg g-1)

Kf The pseudo-first-order rate constant (h-1),

t Time

Ks The pseudo-first-order rate constant

(g gm-1 min-1)

a Desorption constant in Elovich Equation

b Initial adsorption rate in Elovich Equation

kid The intraparticle diffusion rate constant

(mg g-1 h-1/2)

I Thickness of the boundary layer in intraparticle

model of Weber and Morris

k Rate constant of adsorption of Dumwald–Warner

model (min-1)

F Fractional attainment of equilibrium in the

Dumwald-Warner model

kfd Film diffusion rate coefficient in the Dumwald–

Wagner mode

Qmax Maximum phenol uptake in Langmuir model

kL Constant in Langmuir model that denoted the

energy of adsorption and affinity of the binding

sites (L mg-1)

kf Constant in Freundlich model giving an

indication of adsorption capacity [mg g-1 (L

mg-1) n]

& Liliana Giraldo

[email protected]

1 Departamento de Quımica, Facultad de Ciencias, Grupo de

Investigacion de Solidos Porosos y Calorimetrıa, Universidad

de los Andes, Bogota, Colombia

2 Departamento de Quımica, Facultad de Ciencias, Universidad

Nacional de Colombia, Bogota, Colombia

123

Adsorption

DOI 10.1007/s10450-015-9724-2

Page 2: Equilibrium, kinetics and thermodynamics study of phenols

n Constant in Freundlich model giving an

indication of adsorption intensity

QmDRK Amount adsorbed of solute on the monolayer in

Dubinin–Radusckevisch–Kanager model

Cs Concentrations of equilibrium saturation in

Dubinin–Radusckevisch–Kanager model

ES Is related to the energy characteristic of the

process in Dubinin–Radusckevisch–Kanager

model

DG0 Gibbs free energy change of adsorption process

R R is the gas constant (8.314 J mol-1 K-1)

Ko K0 is the apparent equilibrium constant, in this

study the Langmuir constant was used

DS0 Entropy change of adsorption process

DH0 Enthalpy change of adsorption process

T Temperature in Kelvin

A Arrhenius factor

Ea Arrhenius activation energy of adsorption

1 Introduction

Water pollution is one of the most undesirable environ-

mental problems in the world and requires solutions (Rain-

bown 2002). This is caused by the introduction of substances

which may be non-toxic but affect biological cycles at high

concentrations, in the case of substances such as nitrates,

phosphates and some organic compounds, and secondly the

introduction of toxic substances, such as heavy metals,

hydrocarbons, pesticides, phenols, etc. (Blanchard et al.

1984; Dabrowski et al. 2005; Qing-Song et al. 2010). The

latter are considered priority pollutants due to their toxicity

in living organisms, even at low concentrations (Kumar et al.

2007). Phenolic wastewater is generated from chemicals,

pharmaceuticals, papermaking, rubber, wood, dye, and

pesticide industries and are highly toxic and harmful. In

addition, phenols are considered priority pollutants by the

EPA since they are not only carcinogenic but also cause an

unpleasant taste and odour, even at low concentrations

(Ahmaruzzaman and Sharma 2005).

Various technologies have been developed to treat

phenolic wastewater. Among them, adsorption is an

effective technology for the removal of phenols from

wastewater (Qing-Song et al. 2010; Rincon-Silva et al.

2015). Activated carbon (AC) is an extremely versatile

carbonaceous material that is widely used as an adsorbent

and catalyst support in industries (Rodrıguez-Reinoso and

Linares-Solano 1989, Rodriguez-Reinoso 2007). More-

over, its large surface area and micropore volume content,

favourable pore size distribution, surface chemistry

including the oxygen functional groups, the degree of

polarity and the active surface area lend AC as an appro-

priate adsorbent for a variety of environmental applica-

tions, i.e. the removal of organic materials, and the

purification and storage of gases and organic compounds

from aqueous solution. The adsorption efficiency of AC

relies strongly on its special surface and structural char-

acteristics (Rodrıguez-Reinoso and Linares-Solano 1989;

Dabrowski et al. 2005; Rodriguez-Reinoso 2007).

Activated carbon is mainly produced by thermal and

chemical activation (Rodrıguez-Reinoso and Linares-

Solano 1989; Sun and Jian 2010). Thermal or physical

activation involves the primary carbonisation of a car-

bonaceous precursor (below 700 �C) followed by activa-

tion of the obtained char with oxidising gases such as air,

CO2 or steam at high temperature in the range of

700–1000 �C. Chemical activation consists of the

impregnation of raw material with chemical agents such as

ZnCl2, H3PO4 or KOH followed by carbonisation at tem-

peratures between 400 and 800 �C under a N2 atmosphere

(Rodrıguez-Reinoso and Linares-Solano 1989). An enor-

mous range of lignocellulosic materials including rice

husk, corn cobs, fruit stone, almond shell, coconut shell,

sugar cane bagasse, palm shell, pistachio-nut shell and

cotton stalk have been used as activated carbon precursors

(Li et al. 2008; Li et al. 2010; Chandra et al. 2009; Nor

et al. 2013). Eucalyptus Globulus labil is other kind of

lignocellulosic material which has a reasonably high con-

tent of carbon, utilized also as raw material for AC (Tan-

credi et al. 2004; Mojica-Sanchez et al. 2012; Rincon Silva

et al. 2014).

The objective of this work is study the capacity of

adsorption of derivatives phenolics monosubstituted: phe-

nol, 4-clorophenol and 4-nithrophenol (phenol, 4-CP and

4-NP, respectively) on activated carbons prepared from

Eucalyptus shell by chemical activation. Additionally, an

extensive kinetic study was conducted to evaluate the

efficiency of the adsorption process. The pseudo-first-order

and pseudo-second-order models were used to correlate the

adsorption kinetics data of phenols onto AC, the kinetic as

well as the diffusion parameters were evaluated. Thermo-

dynamics studies have also been performed to understand

the process of removal of the selected phenols on ACs.

2 Materials and methods

2.1 Preparation of activated carbon

In this research, the lignocellulosic materials were treated

by impregnation with zinc chloride at low concentrations in

order to obtain microporous activated carbons at low costs.

This type of activation leads to ACs with high porosity, and

although the distribution of pore size is largely determined

Adsorption

123

Page 3: Equilibrium, kinetics and thermodynamics study of phenols

by the precursor, the amount of zinc chloride used also

influences the porosity of the final product (Khalilia et al.

2000; Azevedo et al. 2007).

The suggested mechanism for activation using zinc

chloride can be summarised as follows: during impregna-

tion, the chemical reagent is introduced into the interior of

the precursor particles and causes some hydrolysis reactions

which are seen in a weight loss, in the output volatile mate-

rial, in the weakening of the structure and the increase in

elasticity, the chemical also causes the swelling of particles.

The two processes are more evident with increasing con-

centrations of zinc chloride. During heat treatment, zinc

chloride prevents the formation of volatiles thus increasing

process yield. During the impregnation and carbonisation at

low ratios impregnation occurs minimal weight loss, since

the amount of zinc chloride may be distributed uniformly

throughout the precursor with a large dispersion in the

interior of the particles resulting of activated carbons after

extensive washing with uniform microporosity and low

macroporosity. At higher impregnation ratios, hydrolysis

and swelling are accentuated, the zinc chloride may not be

distributed uniformly within the particles, and although the

total pore volume increases, the pore size distribution is more

heterogeneous, with meso- and macroporosity becoming

more important (Khalilia et al. 2000; Azevedo et al. 2007).

Thus, activated carbon was prepared by mixing of the

eucalyptus shell (ranging in size from 2.5 to 3.5 mm) at

two different concentrations of ZnCl2 (10 and 25 m/v),

which will be called ACS10 and ACS25. The impregnation

ratio of ZnCl2 and eucalyptus shell was 2.0/1.0; this pro-

cess was performed for 12 h. After mixing both together,

the salt impregnated material was dried overnight at 90 �C,

then a weighed sample was carbonised in a quartz tube

furnace at temperature of 550 �C, holding time of 2 h and

heating rate of 5 �C min-1. After carbonisation, the sample

was cooled down to room temperature in a flow of nitrogen

and then removed from the reactor. In order to remove

impurities in the synthesised ACs, it was dispersed in

distilled water at 80 �C. After that, the sample was washed

sequentially with hot and cold distilled water until the wash

water reached a pH of 6-7 (Qing-Song et al. 2010; Rincon-

Silva et al. 2015).

2.2 Characterisation of the ACs samples

The textural characterisation of the prepared ACs included

the surface area, the extent of micro- and mesoporosity was

conducted using N2 adsorption/desorption at 77 K using a

computer system Autosorb 3B, Quantachrome Co. The

specific surface areas were calculated from the N2

adsorption isotherm with Brunauer, Emmet and Teller

equation (B.E.T.) at the relative pressure in the range of

0.001–0.3 bar. The total pore volume was determined from

the amount of N2 adsorbed at P/P0 0.99. The volume of

micropores was estimated using the Dubinin–Radushke-

vich (D.R.) method (Qing-Song et al. 2010; Mojica-San-

chez et al. 2012; Anisuzzaman et al. 2014)

The selective method for determining of the total acid

and basic sites on the carbon surface was employed

(Boehm 1994, 2002; Lopez-Ramon et al. 1999). The pH at

point of zero charge (pHPZC) was determined using the

titration method of mass using a CG840B Schott pH meter

(Babic et al. 1999; Mojica-Sanchez et al. 2012).

The surface functional groups of the ACs samples were

also detected by Fourier Transform Infrared (FTIR) spec-

troscope (FTIR—Nicolet Impact 410) using a potassium

bromide (KBr) pellet prepared by mixing 0.033 % of dried

AC sample in KBr. The spectra were recorded between

4000 and 400 cm-1 (Pakulaa et al. 2005; Saka 2012).

2.3 Adsorption kinetics

Adsorption kinetic experiments were carried out using a

shaker water bath. Kinetic experiments were carried out by

agitating 50 mL of solution of phenols 100 mg of ACs at a

constant agitation speed, 20 �C and natural pH, because

measuring the pH of the solution showed that this was

below the pKa of phenols. Therefore, the molecular form of

each phenol of the ionic form predominates, favouring the

adsorption process due to electrostatic interactions (Mor-

eno-Castilla et al. 1995; Moreno-Castilla 2004; Qing-Song

et al. 2010; Rincon-Silva et al. 2015). Agitation was per-

formed for 120 min, which is more than sufficient time to

reach equilibrium at a constant stirring speed of 120 rpm.

Preliminary experiments had shown that the effect of

separation time on the adsorbed amount of phenol was

negligible. Two millilitres of samples was drawn at suit-

able time intervals. The samples were then centrifuged at

500 rpm and the omitted concentration in the supernatant

solution was analysed using UV–VIS spectrophotometry

with Milton Roy Co, Spectronic Genesys, equipment by

monitoring the absorbance changes at a wavelength of

maximum absorbance: 269, 280 and 319 nm for phenol,

4-CP and 4-NP, respectively (Tseng et al. 2010; Qing-Song

et al. 2010). Each experiment continued until equilibrium

conditions were reached when no further decrease in the

phenol concentration was measured; then, we proceeded to

plot the adsorption capacity of phenols on activated car-

bons versus time (Kumar et al. 2007; Kunwar et al. 2008;

Tseng et al. 2010; Qing-Song et al. 2010).

2.4 Adsorption isotherms of phenol, 4-nitrophenol

and 4-chlorophenol

The equilibrium isotherm of phenols adsorption on ACs

was determined by performing an adsorption test in

Adsorption

123

Page 4: Equilibrium, kinetics and thermodynamics study of phenols

100 mL flasks, where 50 mL of phenols solutions with

different initial concentrations (50–1500 mg L-1) were

placed in each flask. Then, 100 mg of each of the prepared

activated carbon was added to each flask and kept in shaker

of 120 rpm at 20 �C for 72 h to reach equilibrium. The

equilibrium concentration of phenolic compounds was

determined with respect to calibration curves, at wave-

lengths (kmax) (Dabrowski et al. 2005; Qing-Song et al.

2010; Rincon-Silva et al. 2015).

Data amount Qe adsorbed at equilibrium (mg g-1) were

calculated from the following equation:

Qe ¼ V � C0 � Ce

mð1Þ

where Co and Ce are the initial and equilibrium phenol

concentration in mg L-1, V is the volume of solution

(L) and m is the mass of ACs (g) (Qing-Song et al. 2010;

Rincon-Silva et al. 2015).

2.5 Thermodynamic analysis

A thermodynamic study of adsorption process of phenols

on ACs to estimate the feasibility of the adsorption process

was performed. Therefore, the enthalpy of adsorption was

determined by the isosteric method which is known as the

standard differential enthalpy of adsorption, where exper-

imental adsorption isotherms were used at different tem-

peratures (20, 30 and 40 �C) with the procedure described

in Sect. 2.4 (Humpola et al. 2013; Rincon-Silva et al. 2015;

Ashraf et al. 2014).

This method is useful for estimating the enthalpy dif-

ferential adsorption, under the constraint of constant frac-

tion of coating of the surface adsorption. Initially, for a

system in thermodynamic equilibrium, the following fun-

damental equation is satisfied (Stoeckli et al. 1995;

Douillard 1996; Arias et al. 2009):

DG�ads ¼ �RTLnkeq ¼ DH�

ads � TDS�ads ð2Þ

where Keq is the equilibrium constant dynamic, chemical

equilibrium type during the adsorption process. Thus, the

constant kL of the Langmuir isotherm (see Eq. (11) in

Sect. 3.4), can be directly related to the equilibrium con-

stant of Eq. 2 in the vicinity of the boundary of Henry law

(Douillard 1996; Arias et al. 2009). Incorporating the

Langmuir isotherm in this equation, we obtain:

DG�ads ¼ �RTIn

hð1 � hÞ

1

p

� �¼ DH�

ads � TDS�ads ð3Þ

As a result, it is possible to obtain the equation of Vant

Hoff, which can be used to calculate the values of entropy

and enthalpy. Additionally, the equilibrium conditions

could be used to obtain the free energy of the adsorption

process (Stoeckli et al. 1995; Douillard 1996; Arias et al.

2009; Ashraf et al. 2014).

The Gibbs free energy change (DG�) values can discern

whether a process is spontaneous or not; negative values of

DG� imply a spontaneous process. The enthalpy change

(DH�) provides information about the exothermic or

endothermic nature of the process and differentiates

between physical and chemical adsorption process. The

entropy change (DS�) predicts the magnitude of the chan-

ges on the adsorbent surface, allowing the randomness of

the adsorbate-adsorbent interface to be evaluated (Hum-

pola et al. 2013; Rincon-Silva et al. 2014; Ashraf et al.

2014).

3 Results and discussion

3.1 Characterisation of activated carbons

3.1.1 Textural and chemical properties

The surface area of samples was calculated by B.E.T.

equation. Figure 1 shows the nitrogen adsorption isotherms

for the samples ACS25 and ACS10, demonstrating that

activated carbons obtained fit to Langmuir isotherm or

Type I related with the IUPAC classification (Martinez

1988; Rodrıguez-Reinoso and Linares-Solano 1989;

Lovera 2003). A greater volume was adsorbed at low rel-

ative pressures, characteristic of microporous adsorbents,

and the next part of the isotherm is not completely linear,

indicating the presence of a larger pore size generated by

the activation type (Martinez 1988; Lovera 2003). In the

adsorption isotherms for the carbon ACS10, it can be

observed that the curve displayed a small hysteresis loop,

indicating the presence of mesopore volume (Rodrıguez-

Reinoso and Linares-Solano 1989; Lovera 2003). Finally, it

0.0 0.2 0.4 0.6 0.8 1.00

20

40

60

80

100

120

Relative Pressure (P/P0)

V a

dsor

bed

(cm

3 g-1

STP

)

ACS10ACS25

Fig. 1 Nitrogen adsorption isotherms at 77 K for activated carbons

Adsorption

123

Page 5: Equilibrium, kinetics and thermodynamics study of phenols

is evident that the ACS25 sample adsorbs greater volumes

of nitrogen, exceeding 100 cm3 g-1; this happens because

the increase in concentration of ZnCl2 favours the devel-

opment of porosity and apparent surface area, because zinc

chloride is a dehydrating agent, which produces the greater

removal of water molecules from the lignocellulosic

matrix, which increases porosity development. Also, due to

the increase in zinc chloride atoms on ACS25 sample,

more pores will be developed when the temperature of

carbonisation and washing the material remove particles of

this activating agent.

Table 1 shows the apparent surface area calculated by

the B.E.T. method, the micropore volume content, meso-

pore and total pore volume which were calculated by the

D.R. method, where the characteristic energy also was

determined. The results show changes in textural charac-

teristics of carbonaceous materials. It was observed that for

sample ACS25, the apparent surface area was 300 m2 g-1

and for sample ACS10, it was 250 m2 g-1, demonstrating

similar values. Results of micropore volume obtained by

the equation D.R. showed a similar tendency to the car-

bonaceous samples, in which the micropore volume was

0.140 and 0.125 cm3 g-1 for ACS25 and ACS10, respec-

tively. On the other hand, the volume of mesopores pre-

sented low values at both samples. However, an increased

content of mesopores for ACS10 carbon is observed.

Additionally, the average pore diameter was determined

and reported in Table 1, which shows the larger diameter

pore (1,820 nm) for the sample with the higher surface area

and greater micropore volume content, i.e. the activated

carbon ACS25.

In general, the textural properties presented in Table 1

are best developed for ACS25 sample, which happens

because with increasing concentration of zinc chloride

the high development of porosity is evidenced and a

higher surface area is apparent. As reported in other

studies of lignocellulosic materials activated by zinc

chloride (Khalilia et al. 2000; Azevedo et al. 2007).

Likewise, as shown in Fig. 1 and Table 1, the microp-

orosity can be attributed to concentrations of activated

agent in samples, which favours the development of

microporosity (Rodrıguez-Reinoso and Linares-Solano

1989; Lovera 2003).

The results of total surface groups are also shown in

Table 2. The carbons immersed in HCl, determine the total

amount of basic sites, indicating that sample ACS25 had

the largest concentration with 0.238 meq g-1, and sample

ACS10 presented the lowest concentration of basic sites

with a value of 0.178 meq g-1. Moreover, immersion in

NaOH carbons can determine the concentration of total

acid sites, finding that ACS25 sample contained the highest

concentration of total acid sites with 0.092 meq g-1, while

the other sample had a value of 0.057 meq g-1. The

obtained carbons had a higher content of basic groups and

pH values in the relatively neutral zero point of charge,

with values between 6.98 and 6.91, which favours

adsorption of phenols derivatives. This is because if the

samples have a greater amount of acidic groups, they are

located on the edges of graphene layers, withdrawing the

electron density of electrons p, leading to a weaker inter-

action between the p electrons of the aromatic ring of the

phenol and graphene layers, which reduces the adsorptive

capacity (Moreno-Castilla et al. 1995; Moreno-Castilla

2004).

In order to detect the functionality present in ACs,

adsorption in the infrared (IR) region takes place

(4000–400 cm-1) due to the rotational and vibrational

movement of the molecular groups and chemical bond of a

molecule. The FT-IR spectra were obtained to evaluate

qualitatively the chemical structures of ACs.

Figure 2 Shows the FT-IR spectrum of ACs, which

indicated various surface functional groups. The broad

band at around 3500 cm-1 is typically attributed to the

hydroxyl group of phenol, alcohol, and carboxylic acid.

The relatively intense band at about 1200 cm-1 observed

in the samples is attributed to C–O–C stretching in ethers.

In the FT-IR spectra, the peaks observed at 1577 and

1586 cm-1 can be attributed to C = O stretching in

ketones. Moreover, the region of the spectrum of

2220 cm-1 is attributed to alkyne group (C:C) (Saka

2012). The bands observed from 700 to 750 correspond to

stretching by the presence of the C–Cl group, due to the

activating agent used. Finally, to appreciate the bands

corresponding to specific surface chemical groups, no

specific differences were evidenced by variation in the

concentration of the activating agent (Pakułaa et al. 2005).

Table 1 Textural and chemical properties of the activated carbons

Sample AB.E.T.

(m2 g-1)

Model D.R. Chemistry properties

Vlpore

(cm3 g-1)

Vmesopore

(cm3 g-1)

Vtotal

(cm3 g-1)

E0

(KJ mol-1)

Average pore

diameter (nm)

Acidity

(meq g-1)

Basicity

(meq g-1)

pHPCC

ACS25 300 0.140 0.007 0.146 10.946 1.510 0.092 0.238 6.98

ACS10 250 0.125 0.015 0.140 8.354 1.820 0.057 0.178 6.91

Adsorption

123

Page 6: Equilibrium, kinetics and thermodynamics study of phenols

Table

2P

seu

do

-firs

to

rder

,p

seu

do

-sec

on

do

rder

and

chem

iso

rpti

on

mo

del

con

stan

tsan

dco

rrel

atio

nco

effi

cien

tsfo

rp

hen

ols

adso

rpti

on

on

toA

Cs

Ph

eno

lsC

arb

on

C0

(mg

L-

1)

Qeexp

(mg

g-

1)

Pse

ud

o-fi

rst

ord

erP

seu

do

-sec

on

do

rder

Elo

vic

hk

inet

icm

od

el

Qe

(mg

g-

1)

Kf*

10-

2(h

-1)

R2

DQ

(%)

Qe

(mg

g-

1)

b(g

-1

h-

1)

R2

DQ

(%)

a(m

gg-

1

h-

1)

b(g

h-

1)

R2

DQ

(%)

Ph

eno

lA

CS

25

10

4.6

74

3.6

00

2.3

03

0.9

15

0.0

95

4.7

50

1.9

61

0.9

54

0.0

54

0.1

09

6.5

79

0.8

37

1.4

52

40

19

.02

51

7.8

59

1.6

12

0.8

56

0.3

42

20

.85

90

.09

70

.96

80

.09

50

.64

72

.94

10

.80

00

.89

5

70

36

.52

63

5.4

73

1.3

82

0.9

27

0.1

62

40

.47

30

.60

90

.97

90

.00

90

.79

40

.92

30

.86

60

.96

3

10

04

5.0

55

39

.46

64

.60

60

.96

80

.09

34

6.4

66

0.0

72

0.9

48

0.0

28

1.4

08

0.5

51

0.8

81

2.6

24

AC

S1

01

03

.60

14

.41

74

.93

00

.90

80

.25

62

.41

71

4.0

13

0.9

88

0.0

31

0.1

16

5.6

50

0.9

38

2.6

58

40

20

.20

81

5.5

97

1.3

82

0.9

07

0.4

27

16

.59

71

3.1

53

0.9

79

0.2

56

0.3

95

1.7

57

0.8

83

0.8

96

70

31

.40

03

0.8

97

1.1

52

0.8

56

0.2

27

32

.89

70

.06

80

.98

80

.05

60

.51

21

.50

60

.83

42

.36

5

10

04

1.9

58

36

.93

21

.15

20

.92

80

.12

84

6.9

32

2.9

86

0.9

79

0.0

25

1.0

56

0.8

91

0.8

70

3.3

65

4-N

PA

CS

25

10

6.3

44

3.0

76

3.6

85

0.9

95

0.1

74

6.0

76

1.9

55

0.9

88

0.0

43

1.5

26

3.3

67

0.9

88

3.5

41

40

22

.93

52

0.5

40

8.9

82

0.9

94

0.3

51

24

.54

02

.70

60

.99

80

.17

41

.25

01

.12

20

.98

12

.65

8

70

37

.56

33

9.2

35

9.9

82

0.9

27

0.1

30

38

.65

23

.60

90

.92

70

.35

11

.23

52

.94

10

.88

00

.98

5

10

05

6.6

24

51

.26

31

1.9

82

0.9

38

0.0

86

68

.21

44

.60

90

.98

90

.00

31

.21

00

.92

30

.86

01

.02

3

AC

S1

01

05

.53

72

.05

43

.22

40

.99

50

.78

14

.05

44

.03

40

.98

80

.00

40

.78

84

.44

40

.93

32

.36

5

40

29

.64

71

8.2

96

2.7

64

0.9

96

0.3

88

22

.29

66

.23

00

.96

90

.78

11

.12

81

.15

30

.97

71

.36

5

70

38

.52

63

6.1

23

3.7

36

0.9

25

0.1

44

33

.24

19

.23

00

.93

90

.38

81

.15

92

.05

40

.87

92

.98

4

10

06

8.2

35

47

.14

24

.76

49

.40

40

.08

26

0.6

35

14

.03

40

.98

80

.00

52

.14

82

.65

40

.87

84

.36

5

4-C

PA

CS

25

10

7.7

90

3.1

68

2.7

64

0.9

87

0.4

21

7.1

68

2.3

40

0.9

89

0.0

05

0.1

00

5.2

08

0.8

45

3.6

54

40

23

.09

41

9.8

77

2.3

03

0.9

75

0.1

82

28

.87

71

.61

20

.96

90

.42

10

.51

70

.94

20

.91

71

.36

5

70

43

.30

23

5.6

70

1.3

82

0.9

26

0.0

55

48

.67

00

.84

20

.98

90

.00

90

.74

50

.65

40

.90

40

.89

5

10

07

1.9

58

57

.25

71

.15

20

.96

50

.04

56

9.2

57

0.6

36

0.9

78

0.0

05

1.2

29

0.4

00

0.9

15

1.3

54

AC

S1

01

06

.46

54

.53

82

.53

30

.90

40

.12

26

.53

81

0.5

69

0.9

89

0.0

20

0.1

07

6.3

29

0.9

21

3.5

64

40

21

.70

92

2.3

30

1.3

82

0.9

24

0.1

88

25

.33

01

.77

80

.95

80

.12

20

.32

41

.55

50

.92

22

.69

8

70

49

.61

73

9.1

77

1.3

82

0.9

95

0.0

66

41

.17

70

.65

20

.97

80

.10

91

6.7

90

0.0

60

0.9

11

4.6

58

10

07

8.2

68

59

.03

01

.15

20

.96

50

.04

76

9.0

30

0.4

93

0.9

89

0.0

14

0.9

90

0.4

69

0.9

02

5.6

54

Adsorption

123

Page 7: Equilibrium, kinetics and thermodynamics study of phenols

3.2 Adsorption kinetics

The study of adsorption kinetics is important because the

rate of adsorption (which is one of the criteria for efficiency

of adsorbent) and also the mechanism of adsorption can be

concluded from kinetic studies (Ho and McKay 2000; Ho

2006b; Wu et al. 2009; Tseng et al. 2010; Boparai et al.

2011; Aljeboree et al. 2014).

The relationship between contact time and phenol

adsorption onto ACs at different initial phenol concentra-

tions (10.0 to 100 mg L-1) is shown in Fig. 3. This graph is

obtained by plotting the absorption capacity versus time t

in the time required to reach equilibrium system in hours,

proving the variation of the amount of adsorbed (Q) as a

function of time. The rate of adsorption for phenols is high

at initial times of adsorption (Azizian 2008; Plazinski and

Plazinska 2012). For phenols, most of the adsorption takes

place within 300 min which indicate that the rate of phe-

nols adsorption by ACs is high (Figure to 4-CP and 4-NP

not was shown here, but had a similar behaviour). Figure 3

indicates that while the adsorption of phenol was quite

rapid initially, the rate of adsorption became slower with

the time and reached a constant value (equilibrium time).

Additionally, it showed that the speed increased at lower

concentrations. The initial faster rate may be due to the

availability of the uncovered surface area of the adsorbents

(Ho and McKay 1999; Ho 2006a; Manasi and Rajesh

2014).

In order to analyse the adsorption kinetics of mono-

substituted phenols by ACs the pseudo-first-order, pseudo-

second order, Elovich equation, Dumwald–Wagner equa-

tion and intra-particle diffusion model were tested (Leyva-

Ramos and Geankoplis 1994; Qiu et al. 2009; Feng-Chin

et al. 2009; Siminiceanu et al. 2010; Gihan and El-Khaiary

2010; Plazinski et al. 2013) The result of fitting is listed in

Tables 2 and 3.

A simple kinetic analysis of adsorption (pseudo-first-

order equation or Lagergren equation) is in the form:

Qt ¼ Qe 1 � expðkf tÞ� �

ð4Þ

Where Qt is the amount of adsorbate adsorbed at time t

(mg g-1), Qe is the adsorption capacity in the equilibrium

(mg g-1), kf is the pseudo-first-order rate constant (h-1),

and t is the contact time (Rudzinski and Plazinski 2006;

Qiu et al. 2009).

A pseudo-second-order equation based on adsorption

equilibrium capacity may be expressed in the equation:

Qt ¼ksQ

2e t

1 þ ksQetð5Þ

where ks is the pseudo-first-order rate constant (g gm-1 -

min-1) and t is the contact time (Qiu et al. 2009).

The applicability of the kinetic model to describe the

adsorption process was further validated by the normalised

standard deviation DQ (%), which is defined as (Qiu et al.

2009; Valderrama et al. 2010):

DQ ¼ 100

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiPðQexp � Qcal=QexpÞ� �2

N � 1

s8<:

9=; ð6Þ

The Elovich equation has been applied satisfactorily to

some chemisorption processes and has been found to cover

a wide range of slow adsorption rates. Moreover, this

describes the adsorption of adsorbate by solid adsorbents in

aqueous medium. The same equation is often valid for

3600 3000 2400 1800 1200 600

60

70

80

90

100

C=CC=O

C-O-COH-

1586

700

1200

1577

2220

3500Tran

smitt

ance

% T

Wavenumber cm-1

ACS10ACS25

Fig. 2 FT-IR spectra for the ACs derivate from Eucalyptus shell

activated with ZnCl2

0 1 2 3 4 5 6 7 8 9 100

10

20

30

40

50

ACS25 10 mg L-1 ACS10 10 mg L-1

ACS25 40 mg L-1 ACS10 40 mg L-1

ACS25 70 mg L-1 ACS10 70 mg L-1

ACS25 100 mg L-1 ACS10 100 mg L-1

Qt (

mg

g-1)

Time (h)

Fig. 3 Kinetics of phenol adsorption at different concentrations onto

ACs

Adsorption

123

Page 8: Equilibrium, kinetics and thermodynamics study of phenols

systems in which the adsorbing surface is heterogeneous,

and is formulated as (Qiu et al. 2009; Feng-Chin et al.

2009):

Qt ¼1

b

� �LnðabÞ þ 1

b

� �LnðtÞ ð7Þ

The intra-particle diffusion model based on the theory

proposed by Weber and Morris was used to identify the

diffusion mechanism (Gihan and El-Khaiary 2010).

According to this theory, the adsorbate uptake Qt varies

almost proportionally with the square root of the contact

time, t� rather than t, Eq. (7) (Qiu et al. 2009; Gihan and

El-Khaiary 2010):

Qt ¼ kidffiffit

pþ I ð8Þ

where I is the intercept and kid (mg g-1 h-1/2) is the

intraparticle diffusion rate constant.

The Dumwald–Warner model is another intraparticle

diffusion model, which is written as (Acharya et al. 2009;

Qiu et al. 2009; Siminiceanu et al. 2010; Gihan and El-

Khaiary 2010; Qing-Song et al. 2010; Theydana and

Ahmed 2012):

F ¼ Qt

Qe

¼ 1 � 6

p2

X1n¼1

1

n2expð�n2ktÞ ð9Þ

where k (min-1) is the rate constant of adsorption. Equa-

tion (9) can be simplified as:

Inð1 � FÞ ¼ �kfdt ð10Þ

Where F is the fractional attainment of equilibrium

(F = Qt/Qe) and kfd (min-1) is the film diffusion rate

coefficient. The Dumwald–Wagner model proved to be a

reasonable model for different kinds of adsorption systems

(Qiu et al. 2009; Siminiceanu et al. 2010).

The results of the kinetic data of adsorptions of phenol,

4-NP and 4-CP at different initial concentrations are given

in Tables 2 and 3. In the pseudo-first order model kf and

Qe, were calculated using the slope and intercept of plots of

Log (Qe - Qt) versus t (Fig. 4; Table 2). It shows that the

Table 3 Intraparticle diffusion and liquid film diffusion model constants and correlation coefficients for phenols adsorption onto ACs

Phenols Carbon C0 (mg L-1) Intraparticle diffusion Liquid film diffusion

kid (mg g-1 h0.5) I R2 DQ (%) Qe

(mg g-1)

kfd (h-1) R2 DQ (%)

Phenol ACS25 10 0.094 0.042 0.979 0.854 3.546 0.009 0.993 1.526

40 0.487 0.365 0.978 1.562 18.5621 0.009 0.998 2.651

70 0.733 0.552 0.989 3.541 40.621 0.010 0.990 2.865

100 1.221 0.806 0.998 2.654 40.896 0.404 0.982 3.365

ACS10 10 0.099 0.045 0.998 2.456 3.456 0.010 0.978 2.365

40 0.380 0.304 0.998 1.365 12.365 0.010 0.985 0.954

70 0.456 0.332 0.989 2.365 25.658 0.008 0.993 1.984

100 0.768 0.338 0.998 2.321 40.998 0.010 0.971 2.654

4-NP ACS25 10 0.196 0.428 0.899 2.654 5.654 0.019 0.863 3.256

40 0.262 1.419 0.959 2.365 20.652 0.066 0.988 3.651

70 0.362 2.019 0.979 2.123 36.654 0.009 0.928 2.365

100 0.982 2.119 0.979 2.415 62.365 0.046 0.938 2.654

ACS10 10 0.140 0.34 0.819 1.025 3.658 0.023 0.801 1.254

40 0.502 0.514 0.899 2.125 23.654 0.043 0.883 2.365

70 0.620 0.564 0.979 2.451 35.568 0.078 0.873 2.854

100 0.730 0.614 0.979 3.562 65.651 0.033 0.889 4.654

4-CP ACS25 10 0.216 0.101 0.989 3.214 6.854 0.011 0.998 3.254

40 0.700 0.362 0.999 3.245 25.658 0.014 0.985 3.652

70 1.034 0.584 0.998 3.354 46.365 0.012 0.990 4.562

100 2.122 0.988 0.988 3.654 60.652 0.016 0.965 4.658

ACS10 10 0.295 0.042 0.997 2.124 4.658 0.011 0.984 2.654

40 0.425 0.197 0.997 2.365 20.654 0.011 0.974 2.854

70 0.629 0.497 0.918 2.854 40.654 0.011 0.989 3.654

100 1.303 0.969 0.978 2.654 62.687 0.013 0.972 3.854

Adsorption

123

Page 9: Equilibrium, kinetics and thermodynamics study of phenols

correlation coefficients (R2) for the pseudo-first order

kinetic model fit are far from 1.00 and the standard devi-

ation values are high compared with the pseudo-second-

order. Moreover, a low correlation was also observed

between Qe exp and Qe calculated of model. Thereby, the

pseudo-first-order model isn’t a suitable equation to

describe the adsorption kinetics of phenols on the ACs (Qiu

et al. 2009; Siminiceanu et al. 2010).

In the pseudo-second order adsorption parameters Qe

and ks in Eq. (3) were determined by plotting t/Qt and

t (Fig. 5; Table 2). The values of Qe calculated of model

are close from the experimental value, the R2 values

derived from the second-order kinetic model were rela-

tively high in comparison with the Pseudo-first order

model. Therefore this model fit the adsorption process of

phenol onto ACs.

As shown in Table 2, phenol and 4-CP demonstrate

similar adsorption kinetics, while other phenols exhibit

slower initial adsorption rates. For phenol, it should be

noted that its adsorption driving force is weaker due to a

relatively lower ultimate uptake; in fact, its adsorption

kinetics is remarkable, as suggested by the second-order

rate index. For 4-NP slower adsorption rates should be

ascribed to the steric effects, i.e., the adsorbate molecules

have difficulties in moving within pores with size not large

enough. Adsorption kinetics is more sensitive to the steric

effects, which demonstrates their influences during the

adsorption process. The similar adsorption kinetics of

phenol and 4-CP indicate that their adsorption process is

not hindered to an appreciable extent, suggesting that steric

effects are negligible if the molecular dimensions are

below some limits (Qiu et al. 2009; Siminiceanu et al.

2010).

In the Elovich model, a and b were calculated from the

slope and intercept of the plot of Qt vs In t and the results

are present in Table 2 (figure not shown here). The values

show that the correlation coefficients were not satisfactory

for most of the cases, which indicated that the Elovich

model is not appropriate for the description of phenol, 4-CP

and 4-NP adsorption by ACs. Therefore, phenols adsorp-

tion on the ACs does not follow chemisorption models.

Thus, suggested phenols adsorption does not occur via the

sharing of electrons between the phenolic ring and basal

plane of ACs (Qiu et al. 2009; Wu et al. 2009; Ahmaruz-

zaman and Laxmi 2010; Al-Khateeb et al. 2014).

3.3 Diffusion kinetic model

A detailed understanding of adsorption mechanisms facil-

itates the determination of the rate-limiting step. This

information can then be used to optimise the design of

adsorbents and adsorption conditions. The overall rate of

adsorption can be described by the following three steps:

(i) film or surface diffusion where the sorbate is transported

from the bulk solution to the external surface of sorbent (ii)

intraparticle or pore diffusion, where sorbate molecules

move into the interior of sorbent particles, and (iii)

adsorption on the interior sites of the sorbent (Theydana

and Ahmed 2012). Since the adsorption step is very rapid,

it is assumed that it does not influence the overall kinetics.

The overall rate of adsorption process, therefore, will be

0 1 2 3 4 5 6 7 8 9 100.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Log

( Qe-Q

t )

ACS25 10 mg L-1 ACS10 10 mg L-1 ACS25 40 mg L-1

ACS10 40 mg L-1 ACS25 70 mg L-1 ACS10 70 mg L-1

ACS25 100 mg L-1 ACS10 100 mg L-1

Time (h)

Fig. 4 Pseudo-first-order kinetic model fit for phenol adsorption at

different concentrations onto ACs

0 1 2 3 4 5 6 7 8 9 100

15

30

45

60

Time (h)

t/Qt (

h g /

mg)

ACS25 10 mg L-1 ACS10 10 mg L-1 ACS25 40 mg L-1

ACS10 40 mg L-1 ACS25 70 mg L-1 ACS10 70 mg L-1

ACS25 100 mg L-1 ACS10 100 mg L-1

Fig. 5 Pseudo-second-order kinetic model fit for phenol adsorption at

different concentrations onto ACs

Adsorption

123

Page 10: Equilibrium, kinetics and thermodynamics study of phenols

controlled by either surface diffusion or intraparticle dif-

fusion (Gihan and El-Khaiary 2010; Theydana and Ahmed

2012).

The Weber–Morris intraparticle diffusion model has

often been used to determine whether intraparticle diffu-

sion is the rate-limiting step. According to this model, a

plot of Qt versus t0.5 should be linear if intraparticle dif-

fusion is involved in the adsorption process and if the plot

passes through the origin then intraparticle diffusion is the

sole rate-limiting step. It has also been suggested that in

instances when Qt versus t0.5 is multilinear two or more

steps govern the adsorption process, the multilinearity of

this plot for adsorption on activated carbon suggests that

adsorption occurred in three phases. The initial steeper

section represents surface or film diffusion, the second

linear section represents a gradual adsorption stage where

intraparticle or pore diffusion is rate-limiting and the third

section is final equilibrium stage. In Fig. 6, the adjustment

graph of the intraparticle diffusion model for phenol is

shown (Figures for 4-CP and 4-NP are not shown here), as

the plot passes through the origin, intraparticle diffusion

could be rate-limiting step for most phenol concentrations

studied; however, for some concentrations of phenol

(100 mg L-1), in both carbons, there were three processes

controlling the adsorption rate but only one was rate lim-

iting in any particular time range. The intraparticle diffu-

sion rate constant kid was calculated from the slope linear

section (Fig. 6; Table 3). The value of the intercept I in the

second section provides information related to the thick-

ness of the boundary layer. Larger intercepts suggest that

surface diffusion has a larger role as the rate-limiting step

(Gihan and El-Khaiary 2010; Theydana and Ahmed 2012;

Ocampo-Perez and Leyva-Ramos 2013).

In the liquid film diffusion models, a linear plot of

In(1 - F) versus t with zero intercept suggests that the

kinetics of the adsorption process is controlled by diffusion

through the liquid film (this Figure is not shown here).

Application of the liquid film diffusion model to the

adsorption of phenol, 4-CP and 4-NP by ACs did not

converge, and the regression coefficient values were very

low; however, the standard deviation was higher, as shown

in Table 3. This indicates that the liquid film diffusion was

not the rate-determining step (Acharya et Al 2009; Gihan

and El-Khaiary 2010; Qing-Song et al. 2010; Theydana and

Ahmed 2012; Ocampo-Perez and Leyva-Ramos 2013).

Based on the results presented in Table 3, it is clear that

the mechanism of interaction between the phenolic com-

pounds and the ACs is somewhat complex. The application

of the intra-particle diffusion model, and liquid film dif-

fusion model, on the experimental data yielded different

straight lines but only in intraparticle model the line

passing through the origin, which indicates some degree of

boundary layer control. This further show that the intra-

particle diffusion and liquid film diffusion are not the only

rate-controlling step, but other processes may also control

the rate and mechanism of adsorption (Acharya et Al 2009;

Gihan and El-Khaiary 2010; Theydana and Ahmed 2012;

Ocampo-Perez and Leyva-Ramos 2013).

3.4 Adsorption isotherms

The adsorption isotherm can describe the distribution of

phenol between solid phase and the solution at a certain

temperature when the equilibrium was reached.

Figure 7 shows the experimental adsorption isotherms

for phenol, 4-CP and 4-NP, for the two samples, which

depicts the phenol adsorption capacity (Qe expressed in mg

per gram of activated carbon retained). It shows that phenol

adsorption behaviour follows the Freundlich isotherm,

because the mass of adsorbed phenol in a wide range of

concentrations as considered in this work does not became

asymptotic at high concentrations; the same behaviour

applied for the adsorption of 4-NP by ACS10. Although in

some isotherms, e.g. phenol adsorption on ACS10, it

behaves according to the Langmuir model, since the mass

of phenol remains constant when the phenol concentration

at equilibrium is greater than 400 mg L-1. Compared with

phenol, substituted phenols showed greater intensity of

adsorption and adsorption capacity given in the following

order: 4-CP[ 4-NP[ phenol. As phenol has smaller

molecular size than substituted phenols, these results imply

that only a small part of the micropores is filled in phenol,

adsorption and the micropore filling phenomenon is more

evident for the substituted phenols (Moreno-Castilla et al.

1 2 3 40

10

20

30

40

Qt

(mg

g-1)

Time 0.5 (h0.5)

ACS25 10 mg L-1 ACS10 10 mg L-1 ACS25 40 mg L-1

ACS10 40 mg L-1 ACS25 70 mg L-1 ACS10 70 mg L-1

ACS25 100 mg L-1 ACS10 100 mg L-1

Fig. 6 Intraparticle diffusion plots for phenols adsorption at different

concentrations onto ACs

Adsorption

123

Page 11: Equilibrium, kinetics and thermodynamics study of phenols

1995; Dabrowski et al. 2005; Derylo-Marczewska et al.

2010; Qing-Song et al. 2010; Rincon-Silva et al. 2015).

3.5 Equilibrium adsorption of phenols

The experimental data of the adsorption isotherms were

fitted to models Langmuir, Freundlich and Dubinin–

Raduskevich–Kaganer (D.R.K.), which are presented in

Table 3. Each isotherm model was expressed by relative

certain constants which characterised the surface properties

and indicated adsorption capacity of this material (Kumar

et al. 2007; Derylo-Marczewska et al. 2010; Qing-Song

et al. 2010; Rincon-Silva et al. 2015).

The Langmuir model proposes that monolayer sorption

occurs on the solid surface with identical homogeneous

sites. It also suggests that no further adsorption takes place

once the active sites are covered with phenols molecules.

The saturated monolayer isotherm is presented by the fol-

lowing equation:

Qe ¼QmaxkLCe

1 þ kLCe

ð11Þ

where Ce is the concentration of phenol at equilibrium in

solution (mg L-1); Qe is unit equilibrium adsorption

capacity; Qmax is the maximum phenol uptake, giving the

information about adsorption capacity for a complete

monolayer (mg g-1); and KL is a constant denoted the

energy of adsorption and affinity of the binding sites

(L mg-1) (Qing-Song et al. 2010; Okelo and Odebunmi

2010; Rincon-Silva et al. 2015).

Freundlich isotherm is an empirical model assuming that

the distribution of the heat on the adsorbent surface is non-

uniform, namely a heterogeneous adsorption. The equation

is stated as follows (Kumar et al. 2007; Qing-Song et al.

2010; Okelo and Odebunmi 2010; Rincon-Silva et al.

2014):

Qe ¼ kf ðCeÞ1=n ð12Þ

where n and Kf [mg g-1 (L mg-1)n] are both the Fre-

undlich constants giving an indication of adsorption

intensity and capacity, respectively. The degree of non-

linearity between solution concentration and adsorption is

n dependent as follows: if the value of n is equal to unity,

the adsorption is linear; if the value is below to unity, this

implies that adsorption process is chemical; if the value is

above unity adsorption is a favourable physical process

(Mourao et al. 2006; Kumar et al. 2007; Qing-Song et al.

2010; Rincon-Silva et al. 2014).

The Dubinin–Radusckevisch–Kanager model is repre-

sented in Eq. (13)

Qe ¼ QmDRK exp �RTIn Cs=Ce

� �ES

nð13Þ

where QmDRK represents the amount adsorbed of solute on

the monolayer, Ce and Cs are the concentrations of equi-

librium saturation and adsorbate, respectively. ES is related

to the energy characteristic of the process, and n relates to

the heterogeneity variations of the microporous adsorbents.

The parameters n and Es are in principle responsible of

surface heterogeneity for adsorbate-adsorbent system

(Mourao et al. 2006; Rincon-Silva et al. 2015).

The coefficients of determination (R2) and isotherm

parameters from nonlinear regressive method were listed in

Table 4.

When the Langmuir model is applied for experimental

adsorption isotherms, it is observed that the value of Qmax

was greater for ACS25 with values of 55.566, 137.005

and 200.004 mg g-1 for phenol 4-nitrophenol and

4-chlorophenol respectively, the sample ACS10 also had

the same order in the adsorption capacity. The adsorption

was greater to 4-CP followed by 4-NP and finally phenol.

In general, ACS25 carbon present the best able to

adsorption of phenols compounds and this succeed because

this is the activated carbon with the highest apparent sur-

face area (300 m2 g-1) and the largest volume of microp-

ores (0.140 cm3 g-1), favouring the ability to adsorption of

phenolic compounds on this adsorbent.

In Freundlich analysis, the kf was higher for the sample

ACS25, in adsorption of 4-chlorophenol with a magnitude

of 29.818 mg1-1/nL1/ng-1, which is associated with

monolayer adsorption capacity and textural properties of

this sample. The value of 1/n is a measure of heterogeneity

of the surface, in which value close to 0 indicates a

heterogeneous surface. When the value of 1/n is less than 1

it is said that the adsorption process is favourable, as

0 400 800 12000

40

80

120

160

200

ACS10 phenol CS25 phenol ACS10 4-NP ACS25 4-NPACS10 4-CP ACS25 4-CP

Qe (

mg

g-1)

Ce (mg L-1)

Fig. 7 Experimental adsorption isotherms of phenols compounds at

20 �C

Adsorption

123

Page 12: Equilibrium, kinetics and thermodynamics study of phenols

happened in this study. However, there was a greater value

in adsorption favouring 4-chlorophenol.

In the D.R.K. model, the maximum quantity of QmDRK

was on ACS25 for all adsorbates in the order

4-chlorophenol[ 4-nitrophenol[ phenol with values

155.544, 122.836 and 48.525 mg g-1 respectively; it was

noted that there was a correlation of values of adsorption in

monolayer between the two models. Value Es is the char-

acteristic energy of adsorption process, and it can be

observed that value ES was higher for the sample ACS25,

an aspect that is also related with the amount adsorbed on

the monolayer. Additionally, its increase was directly

proportional to values of QmDRK and Qmax. On the other

hand, it was observed that ES values were higher for the

phenol compared with 4-nitrophenol, because this value is

directly related to the solubility of phenol, and this com-

pound had the highest solubility with a value of 93 g L-1,

compared with the value 1.7 g L-1 of 4-nitrophenol. The

values of Es in the adsorption process of phenol were

similar of the adsorption process of 4-chlorophenol, this is

also explained by solubility, since 4-chlorophenol is also

significantly soluble in water, and finally, this is the one

with greater adsorption. Therefore, high values are also

related to this parameter. Again, the activated carbon

ACS25 is the solid that has greater adsorption capacity in

D.R.K model, which happens for the same reason as the

Langmuir model, that is, the textural properties favoured in

this solid compared with the other carbon ACS10, in

relation to the highest amount of used concentration of zinc

chloride.

According to the results presented in Table 4, the acti-

vated carbon with zinc chloride to 25 % m/V is more

efficient in the adsorption of phenolic compounds, relative

to the sample activated to 10 % m/V, since it is the higher

adsorption capacity presented in adjusting models of

Langmuir and D.R.K. This happens because, as explained

above, this sample is the one with higher development of

textural properties. Likewise, ACS25 is the one with higher

content of basic groups which favours adsorption, because

these do not withdraw electron density, which does not

weaken the interaction between the electrons of the aro-

matic ring and of the graphene layers of coal (Moreno-

Castilla et al. 1995). Similarly it is clear in Table 4 that the

best fit of the data is for the Freundlich model because its

correlation coefficient is very close to unity, Furthermore,

this is confirmed because it has the lowest percentage of

deviation (0.025–0.365 %), so it is assumed that the

adsorption occur on adsorbents with the energetically

heterogeneous surfaces (Kumar et al. 2007; Okelo and

Odebunmi 2010; Qing-Song et al. 2010).

Finally, another important aspect to consider in the

adsorption process is the pH of the solution, because this

also interacts with the solid charge, and indicates the

concentration of phenolic compound in solution. From the

diagram of phenol speciation it is known that a high pro-

portion of protonated species at low pH values and pres-

ence of species deprotonated at high pH values. The pH

experimental values of phenols isotherms in this study

were between 6.56 and 7.86. With these data and the dis-

tribution of species, it can be demonstrated that the process

was conducted for phenol protonated species, i.e. they are

not dissociated, and the process favours the dispersion

forces, because if solutions were in basic medium, they

would have decreased the adsorption due to electrostatic

repulsion between the negatively charged surface and the

anions phenolates and between each other (Kumar et al.

2007; Qing-Song et al. 2010; Rincon-Silva et al. 2015).

This behaviour is evidenced by the pKa value of phenol

(9.89) which is greater than the data obtained from

experimental solution pH, which proves that the adsorbed

species were in their protonated form (Kumar et al. 2007;

Qing-Song et al. 2010; Rincon-Silva et al. 2015).

3.6 Thermodynamic study

The thermodynamic behaviours for adsorption of phenol,

4-CP and 4-NP on ACs were investigated. The thermody-

namic parameters such as Gibbs free energy change (DG�),enthalpy change (DH�) and entropy change (DS�) were

calculated using the following equations:

Table 4 Parameters of the models, Freundlich, Langmuir and D.R.K. in the adsorption of phenols compounds on carbon samples obtained

Phenol Carbon Langmuir Freundlich Dubinin–Raduskevich–Kaganer

Qmax

(mg g-1)

KL (L

mg-1)

R2 %

Dev

kf (mg1-1/nL1/

n*g-1)

1/n R2 %

Dev

QmDRK

(mg g-1)

ES

(kJ mol-1)

R2 %

Dev

Phenol ACS10 27.035 0.010 0.964 1.524 3.735 0.285 0.997 0.025 28.504 18.804 0.994 2.365

ACS25 55.566 0.010 0.985 0.958 2.852 0.444 0.985 0.124 48.525 19.448 0.985 1.895

4-NP ACS10 54.954 0.010 0.978 0.745 5.596 0.315 0.997 0.365 45.384 9.376 0.986 3.541

ACS25 137.005 0.030 0.996 1.214 21.984 0.296 0.996 0.018 122.836 12.578 0.855 4.654

4-CP ACS10 125.008 0.030 0.975 1.523 28.447 0.221 0.997 0.087 101.606 22.364 0.935 2.365

ACS25 200.004 0.030 0.984 2.587 29.818 0.282 0.998 0.124 155.544 20.205 0.934 1.365

Adsorption

123

Page 13: Equilibrium, kinetics and thermodynamics study of phenols

DG0 ¼ �RTInK0 ð14Þ

where K0 is the apparent equilibrium constant [In this case,

is the equilibrium constant of Langmuir model kL, as has

already been explained in Sect. 2.5), R is the gas constant

(8.314 J mol-1 K-1)], and T is absolute temperatures

(K) (Qing-Song et al. 2010; Al-Khateeb et al. 2014; Rin-

con-Silva et al. 2015).

The enthalpy change (DH�) of adsorption and entropy

change (DS�) of adsorption were calculated from adsorp-

tion data at different temperatures using the Van’t Hoff

Eq. (9) as follows (Kumar et al. 2007; Rodriguez et al.

2009; Qing-Song et al. 2010; Al-Khateeb et al. 2014):

InK0 ¼ DS0

R� DH0

RTð15Þ

Other thermodynamic quantity is Ea (Arrhenius activa-

tion energy), which can be calculated from the relationship

from the pseudo-second order rate constant of phenols

adsorption, which is expressed as a function of temperature

(Ashraf et al. 2014):

InkS ¼ LnAEa

RTð16Þ

where Ea is the Arrhenius activation energy of adsorption,

A is the Arrhenius factor, R is the gas constant is equal to

8.314 J mol-1 K-1 and T is the operating temperature. A

linear plot of Ln ks vs 1/T for the adsorption of phenols

onto ACs was constructed to generate the activation energy

from the slope (-Ea/R). The magnitude of activation

energy gives an idea about the type of adsorption, which is

mainly physical or chemical. Low activation energies

(5–40 kJ mol-1) are characteristics for physisorption,

while higher activation energies (40–800 kJ mol-1) sug-

gest chemisorptions (Stoeckli et al. 1995; Kumar et al.

2007; Qing-Song et al. 2010; Rincon-Silva et al. 2015).

The thermodynamic parameters of the system: Gibbs

free energy change (DG�), enthalpy change (DH�) and

entropy change (DS�) are presented in Table 5. The free

energy values at temperatures 20, 30 and 40 �C in the

adsorption of phenol; 4-NP and 4-CP were negative in all

cases, showing the spontaneous nature in the process. That

is to say, for three phenols, it was shown that the adsorption

process on ACs was a spontaneous process, and the

decrease of DG� values with the increase of temperature

indicated that the adsorption became less favourable at

higher temperatures. It is also shown that free energy

values increases with the addition of nitro or chlorine

groups to phenol. The maximum value of DG� for

adsorption at 20 �C was on ACS25 for the three phenols,

data directly related to the adsorption capacity of this

carbon. In this study the free energy values indicated that

the adsorption process occurs by physisorption (Stoeckli

and Centeno 1997; Kumar et al. 2007; Qing-Song et al.

2010; Boparai et al. 2011; Rincon-Silva et al. 2015).

DH� values are negative for all process on ACs, showing

the exothermic nature of the adsorption process. Enthalpy

change values are between -6.032 and -30.352 kJ mol-1;

therefore, it follows a behaviour of physisorption in the

process. DS� values were also reported; positive values

indicated the entropy of the system increased during the

adsorption, as in most cases. However, it should also be

noted that the entropy of the universe (including the system

and the surroundings) might increase because the adsorp-

tion reaction was not an isolated process. Additionally,

negative cases of DS� are also given, suggesting a

decreased randomness at the solid–liquid interface during

the adsorption process, as was observed in the adsorption

of 4-NP and 4-CP with values of -10.231 and

-5.583 J mol-1, respectively, on ACS25. Finally, the

values of Arrhenius activation energy are presented in

Table 4 also, the values indicated that the adsorption pro-

cess has a low potential barrier and also confirms the

physisorption process, since these are in the range

6.852–14.564 kJ mol-1 (Huang and Gao 2007; Kumar

et al. 2007; Qing-Song et al. 2010; Rincon-Silva et al.

2015).

4 Conclusions

Eucalyptus shell, a waste solid from trees, was successfully

utilised to synthesise activated carbon, a low cost alterna-

tive adsorbent for the removal of phenols (phenol,

4-chlorophenol and 4-nithrophenol). The samples obtained

present apparent surface area values of 250 and

300 m2 g-1 and micropore volume between 0.007 and

0.015 cm3. The micropore and mesopore distribution is

appropriate for carrying out efficient adsorption processes,

especially in aqueous phenols. Also, the adsorption iso-

therms of phenols on ACs were studied and modelled using

three isotherm models. Freundlich model gives the best

fitting for the adsorption isotherms in most cases, while

Langmuir model is reasonably applicable in all cases.

The values for the adsorption capacity were between

27.035 and 200.004 mg g-1. In kinetic studies, ACs show

high adsorption rate. The pseudo-second-order model gives

satisfactory fitting, and the intraparticle diffusion model

describes the adsorption process well. Steric effects on

adsorption kinetics were found for 4-NP, due to the limi-

tation of the pore structure and the retardation of the

adsorbed molecules. It is proposed that the steric effects are

notable on the adsorption kinetics. The classification of the

kinetic models according to the adsorption study is:

pseudo-first-order[ pseudo-second-order[ intra-particle-

diffuse[ chemisorption.

Adsorption

123

Page 14: Equilibrium, kinetics and thermodynamics study of phenols

The thermodynamic study demonstrates the spontaneous

and exothermic nature of the adsorption process due to

negative values of both free energy change and enthalpy

change, the increased entropy change with the increased

substitution degree was observed; based on thermodynamic

parameters, the adsorption is primarily physical in nature.

Finally, the uptakes were observed at pH\ pKa which

indicates that the adsorbed phenols species were in their

protonated form which improves the adsorption process on

ACs.

Acknowledgments The authors wish to thank the Master Agree-

ment established between the ‘Universidad de los Andes’ and the

‘Universidad Nacional de Colombia’ and the Memorandum of

Understanding entered into by the Departments of Chemistry of both

Universities. Additionally, the authors are also thankful for financial

support to the Additionally, the authors are also thankful for financial

support to the convocation ‘Proyecto Semilla’ of the faculty of sci-

ences of Universidad de los Andes in the category Master Student

2015.

References

Acharya, J., Sahu, J.N., Sahoo, B.K., Mohanty, C.R., Meikap, B.C.:

Removal of Chromium (VI) from wastewater by activated

carbon developed from Tamarind wood activated with Zinc

Chloride. Chem. Eng. J. 150, 25–39 (2009)

Ahmaruzzaman, M., Sharma, D.: Adsorption of phenols from

wastewater. Coll. Interface Sci. 287, 14–24 (2005)

Ahmaruzzaman, M., Laxmi, S.: Batch adsorption of 4-nitrophenol by

acid activated jute stick char: equilibrium, kinetic and thermo-

dynamic studies. Chem. Eng. J. 158, 173–180 (2010)

Anisuzzaman, S.M., Bono, A., Krishnaiah D., Tan, Y.: A study on

dynamic simulation of phenol adsorption in activated carbon

packed bed column. J. King Saud Uni., 1–9 (2014)

Aljeboree, A., Alshirifi, A., Alkaim, A.: Kinetics and equilibrium

study for the adsorption of textile dyes on coconut shell activated

carbon. Arab. J. Chem. 30, 1–13 (2014)

Al-Khateeb, L.A., Abdualah, Y., Asiri, A., Salam, M.: Adsorption

behavior of estrogenic compounds on carbon nanotubes from

aqueous solutions: kinetic and thermodynamic studies. J. Ind.

Eng. Chem. 20, 916–924 (2014)

Arias, J.M., Paternina, E., Barragan, D.: Adsorcion fısica sobre

solidos: aspectos termodinamicos. Quim. Nova 32(5),

1350–1355 (2009)

Ashraf, A., El-Bindary, A., Mostafa Hussien, A., Ahmed, M.E.:

Adsorption of Acid Yellow 99 by polyacrylonitrile/activated

carbon composite: kinetics, thermodynamics and isotherm

studies. J. Mol. Liq. 197, 236–242 (2014)

Azevedo, D.C.S., Araujo, J.S., Bastos-Neto, M., Torres, E.B.,

Jaguaribe, E.F., Cavalcante, C.L.: Microporous activated carbon

prepared from coconut shells using chemical activation with zinc

chloride. Microporous Mesoporous Mater. 100, 361–364 (2007)

Azizian, S.: Comments on ‘‘Biosorption isotherms, kinetics and

thermodynamics’’ review. Sep. Purif. Technol. 63(2), 249–250

(2008)

Babic, B.M., Milonjic, S.K., Polovina, M.J., Kaludierovic, B.V.: Point

of zero charge and intrinsic equilibrium constants of activated

carbon. Carbon 37, 477–481 (1999)

Blanchard, G., Maunaye, M., Martin, G.: Removal of heavy metals

from waters by means of natural zeolites. Water Res. 18(12),

1501–1507 (1984)

Boparai, H.K., Joseph, M., O’carroll, D.M.: Kinetics and thermody-

namics of cadmium ion removal by adsorption onto nano-

zerovalent iron particles. J. Hazard. Mater. 186, 458–465 (2011)

Boehm, H.P.: Some aspects of the surface chemistry of carbon blacks

and other carbons. Carbon 5, 759–769 (1994)

Boehm, H.P.: Surface oxides on carbon and their analysis: a critical

assessment. Carbon 40, 145–149 (2002)

Chandra, T.C., Mirna, M.M., Sunarso, J., Sudaryanto, Y., Ismadji, S.:

Activated carbon from durian Shell. Preparation ad characteri-

sation. J. Taiwan Inst. Chem. Eng. 40(4), 457–462 (2009)

Dabrowski, A., Podkoscielny, P., Hubicki, Z., Barczak, M.: Adsorp-

tion of phenolic compounds by activated carbon—a critical

review. Chemosphere 58, 1049–1070 (2005)

Derylo-Marczewska, A., Marczewski, A.W., Winter, S.Z., Sternik,

D.: Studies of adsorption equilibria and kinetics in the systems:

aqueous solution of dyes–mesoporous carbons. Appl. Surf. Sci.

256(17), 5164–5170 (2010)

Douillard, J.M.: What can really be deduced from enthalpy of

immersional wetting experiments? J. Coll. Interface Sci.

182(0468), 308–311 (1996)

Feng-Chin, W., Ru-Ling, T., Ruey-Shin, J.: Characteristics of Elovich

equation used for the analysis of adsorption kinetics in dye-

chitosan systems. Chem. Eng. J. 150, 366–373 (2009)

Gihan, F.M., El-Khaiary, M.I.: Piecewise linear regression: a

statistical method for the analysis of experimental adsorptiondata by the intraparticle-diffusion models. Chem. Eng. J. 163,

256–263 (2010)

Ho, Y.S.: Review of second-order models for adsorption systems.

J. Hazard. Mater. 136(3), 681–689 (2006a)

Ho, Y.S.: Second-order kinetic model for the sorption of cadmium

onto tree fern: a comparison of linear and non-linear methods.

Water Res. 40(1), 119–125 (2006b)

Table 5 Thermodynamic parameters for the adsorption of phenols on ACs

Phenols ACs -DG� 20 �C(kJ mol-1)

-DG� 30 �C(kJ mol-1)

-DG� 40 �C(kJ mol-1)

-DH�(kJ mol-1)

DS�(J mol-1)

EA

(kJ mol-1)

Phenol ACS10 4.413 4.031 4.009 6.032 2.833 12.655

ACS25 5.643 5.181 4.719 6.352 17.44 14.564

4-NP ACS10 9.213 6.871 6.689 7.032 40.532 7.254

ACS25 12.443 7.181 6.756 11.371 -10.231 8.455

4-CP ACS10 18.013 10.831 9.653 19.038 11.912 6.852

ACS25 22.143 16.181 13.719 30.352 -5.583 8.427

Adsorption

123

Page 15: Equilibrium, kinetics and thermodynamics study of phenols

Ho, Y.S., McKay, G.: Pseudo-second-order model for sorption

processes. Process Biochem. 34(5), 451–465 (1999)

Ho, Y.S., McKay, G.: The kinetics of sorption of divalent metal ions

onto sphagnum moss peat. Water Res. 34(3), 735–742 (2000)

Huang, N.Y., Gao, Q.L.: Qiao.: thermodynamics and kinetics of

cadmium adsorption onto oxidised granular activated carbon.

J. Environ. Sci. 19, 1287–1292 (2007)

Humpola, P., Odetti, H., Fertitta, A., Vicente, J.: Thermodynamic

analysis of adsorption models of phenol in liquid phase on

different activated carbons. J. Chil. Chem. Soc. 58, 1541–1544

(2013)

Kunwar, S., Amrita, M., Sarita, S., Priyanka, O.: Liquid-phase

adsorption of phenols using activated carbons derived from

agricultural waste material. J. Hazard. Mater. 150, 626–641

(2008)

Kumar, A., Kumar, S., Kumar, S., Gupta, D.: Adsorption of phenol

and 4-nitrofenol on granular activated carbon in basal salt

medium: equilibrium and kinetics. J. Hazard. Mater. 147,

155–166 (2007)

Leyva-Ramos, R., Geankoplis, C.J.: Diffusion in liquid-filled pores of

activated carbon. I. Pore volume diffusion. Can. J. Chem. Eng.

72, 262–271 (1994)

Li, W., Yang, K., Peng, J., Zhang, L., Guo, S., Xia, H.: Effects of

carbonisation temperatures on characteristics of porosity in

coconut shell chars and activated carbon derived from carbonised

coconut shell chars. Ind. Crops Prod. 28(2), 190–198 (2008)

Lopez-Ramon, M., Stoeckli, F., Moreno-Castilla, C., Carrasco-Marin,

F.: On the characterisation of acidic and basic surface sites on

carbons by various techniques. Carbon 6, 1215–1221 (1999)

Lovera, R.: Caracterizacion Textural de Adsorbentes. Rev. Chil. ing

Concepcion. 6, 24–28 (2003)

Khalilia, N.R., Campbella, M., Sandib, G., Golas, J.: Production of

micro- and mesoporous activated carbon from paper mill sludge.

I. Effect of zinc chloride activation. Carbon 38, 1905–1915

(2000)

Manasi, V., Rajesh, N.: Adsorption isotherms, kinetics and thermo-

dynamic studies towards understanding the interaction between a

microbe immobilised polysaccharide matrix and lead. Chem.

Eng. J. 248, 342–351 (2014)

Martinez, M.J.: Adsorcion fısica de gases y vapores por carbones,

Alicante. Espana: Universidad de Alicante (Publicaciones)

27–30 (1988)

Mojica-Sanchez, L.C., Ramirez-Gomez, W.M., Rincon-Silva, N.G.,

Blanco-Martinez, D.A., Giraldo, L., Moreno-Pirajan, J.C.: Syn-

thesis of activated carbon from Eucalyptus seed physical and

chemical activation. Afinidad LXIX. 559, 203–210 (2012)

Moreno-Castilla, C.: Adsorption of organic molecules from aqueous

solutions on carbon material. Carbon 42, 83–94 (2004)

Moreno-Castilla, C., Rivera-Utrilla, J., Lopez-Ramon, M.V., Car-

rasco-Marin, F.: Adsorption of some substituted phenols on

activated carbons from a bituminous coal. Carbon 33(6),

845–851 (1995)

Mourao, P.A.M., Carrott, P.J.M., Ribeiro, M.M.L.: Application of

different equations to adsorption isotherms of phenolic com-

pounds on activated carbons prepared from cork. Carbon 44,

2422–2429 (2006)

Norhusna, M.N., Lau, L.C., Lee, K.T.: Mohamed. A.R.: Synthesis of

activated carbon from lignocellulosic biomass and its applica-

tions in air pollution control-a review. J. Environ. Chem. Eng.

1(4), 658–666 (2013)

Ocampo-Perez, R., Leyva-Ramos, R.: Application of diffusional and

kinetic models. Modelling the adsorption of pyridine onto

granular activated carbon. Boletın del Grupo Espanol del

Carbon. 30, 6–9 (2013)

Okelo, F.O., Odebunmi, E.O.: Freundlich and Langmuir isotherms

parameters for adsorption o methylene blue by activated carbon

derived from agrowastes. Adv. Nat. Appl. Sci. 4(3), 281–288

(2010)

Pakułaa, M., Swiatkowski, A., Walczykc, M., Biniak, S.: Voltam-

metric and FT-IR studies of modified activated carbon systems

with phenol, 4-chlorophenol or 1,4-benzoquinone adsorbed from

aqueous electrolyte solutions. Colloids Surf. A 260, 145–155

(2005)

Plazinski, W., Plazinska, A.: Equilibrium and kinetic modelling of

adsorption at solid/solution interfaces. In: Bhatnagar, A. (ed.)

Application of Adsorbents for Water Pollution Control. Bentham

Science, Sharjah (2012)

Plazinski, W., Dziuba, J., Rudzinski, W.: Modelling of sorption

kinetics: the pseudo-second-order equation and the sorbate

intraparticle diffusivity. Adsorption. 19, 1055–1064 (2013)

Qing-Song, L., Tong, Z., Peng, W., Ji-Ping, J., Nan, L.: Adsorption

isotherm, kinetic and mechanism studies of some substituted

phenols on activated carbon fibres. Chem. Eng. J. 157, 348–356

(2010)

Qiu, H., Lu, L., Bing-Cai, P., Zhang, Q., Zhang, W.M., Zhang. Q.X.:

Critical review in adsorption kinetic models. J. Zhejiang Uni.

Sci. A. 10, 716–724 (2009)

Rainbown, P.S.: Trace metal concentrations in aqueous vertebrates:

why and so what? Env. Poll. 120, 497–507 (2002)

Rincon-Silva, N.G., Moreno-Pirajan, J.C., Giraldo, L.: Thermody-

namic study of adsorption of phenol, 4-chlorophenol and

4-nitrophenol on activated carbon obtained from eucalyptus

seed. J. Chem. 2015, 01–12 (2015)

Rincon-Silva, N.G., Ramirez-Gomez, W.M, Mojica-Sanchez, L.C.,

Blanco-Martınez, D.A, Giraldo, L., Moreno-Pirajan, J.C.:

Obtaining of activated carbon from seeds of eucalyptus by

chemical activation with H3PO4. Characterisation and evaluation

of adsorption capacity of phenol from aqueous solution.

Ingenierıa y Competitividad. 16 (8), 207–219 (2014)

Rodriguez, G.A., Giraldo, L., Moreno, J.C.: Calorimetric study of the

immersion enthalpies of activated carbon cloths in different

solvents and aqueous solutions. J. Therm. Anal. Calorim. 96,

547–552 (2009)

Rodrıguez-Reinoso, F., Linares-Solano, A.: Microporous structure of

activated carbons as revealed by adsorption methods. Chem.

Phys. Carbon 21, 140–146 (1989)

Rodriguez-Reinoso, F.: Solidos Porosos: preparacion, Caracterizacion

y Aplicaciones, Bogota, Colombia: Ediciones Uniandes. El

Carbon Activado como Adsorbente. I, 01–43 (2007)

Rudzinski, W., Plazinski, W.: Kinetics of solute adsorption at

solid/solution interfaces: a theoretical development of the empir-

ical pseudo-first and pseudo-second-order kinetic rate equations,

based on applying the statistical rate theory of interfacial transport.

J. Phys. Chem. B 110(33), 16514–16525 (2006)

Saka, C.: BET, TG–DTG, FT-IR, SEM, iodine number analysis and

preparation of activated carbon from acorn shell by chemical

activation with ZnCl2. J. Anal. Appl. Pyrol. 95, 21–24 (2012)

Siminiceanu, I., Marchitan, N., Duca, G., Mereuta, A.: Mathematical

models based on thermodynamic equilibrium and kinetics of an

ion exchange process. Rev. Chim. 61, 623–626 (2010)

Stoeckli, F., Centeno, T.A., Donnet, J.B., Pusset, N., Papirer, E.:

Characterisation of industrial activated carbons by adsorption

and immersion techniques and by STM. Fuel 74(11), 1582–1588

(1995)

Stoeckli, F., Centeno, T.A.: On the characterisation of microporous

carbons by immersion calorimetry alone. Carbon 35(8),

1097–1100 (1997)

Sun, K., Jian, J.C.: Preparation and characterisation of activated

carbon from rubber-seed shell by physical activation. Biomass

Bioenergy 34, 539–544 (2010)

Tancredi, N., Medero, N., Moller, F., Piritz, J., Plada, C., Cordero, T.:

Phenol adsorption onto powdered and granular activated carbon,

Adsorption

123

Page 16: Equilibrium, kinetics and thermodynamics study of phenols

prepared from Eucalyptus wood. J. Coll. Interface Sci. 279,

357–363 (2004)

Theydana, S.K., Ahmed, M.J.: Adsorption of methylene blue onto

biomass-based activated carbon by FeCl3 activation: equilib-

rium, kinetics, and thermodynamic studies. J. Anal. Appl.

Pyrolysis 97, 116–122 (2012)

Tseng, R.L., Wu, K.T., Wu, F.C., Juang, R.S.: Kinetic studies on the

adsorption of phenol, 4-chlorophenol and 2,4-dichlorophenol

from water using activated carbons. J. Environ. Manag. 91,

2208–2214 (2010)

Valderrama, C., Barios, J.I., Caetano, M., Farran, A., Cortina, J.L.: Kinetic

evaluation of phenol/aniline mixtures adsorption from aqueous

solutions onto activated carbon and hyper-cross-linked polymeric

resin (MN200). React. Funct. Polym. 70, 142–150 (2010)

Wu, F.C., Tseng, R.L., Juang, R.S.: Characteristics of Elovich

equation used for the analysis of adsorption kinetics in dye-

chitosan systems. Chem. Eng. J. 150, 366–373 (2009)

Adsorption

123