You are on page 1of 36

Accepted Manuscript

Separation of 2,4,6-Trinitrophenol from Aqueous Solution by Liquid-Liquid


Extraction Method : Equilibrium, Kinetics, Thermodynamics and Molecular
Dynamic Simulation

Hasan Uslu, Dipaloy Datta, Dheiver Santos, Hisham S. Bamufleh, Cuma Bayat

PII: S1385-8947(16)30530-7
DOI: http://dx.doi.org/10.1016/j.cej.2016.04.080
Reference: CEJ 15086

To appear in: Chemical Engineering Journal

Received Date: 9 February 2016


Revised Date: 14 April 2016
Accepted Date: 16 April 2016

Please cite this article as: H. Uslu, D. Datta, D. Santos, H.S. Bamufleh, C. Bayat, Separation of 2,4,6-Trinitrophenol
from Aqueous Solution by Liquid-Liquid Extraction Method : Equilibrium, Kinetics, Thermodynamics and
Molecular Dynamic Simulation, Chemical Engineering Journal (2016), doi: http://dx.doi.org/10.1016/j.cej.
2016.04.080

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.
Separation of 2,4,6-Trinitrophenol from Aqueous Solution by Liquid-Liquid Extraction
Method : Equilibrium, Kinetics, Thermodynamics and Molecular Dynamic Simulation
Hasan Uslua,d,e,*, Dipaloy Dattab, Dheiver Santosc, Hisham S. Bamuflehc, Cuma Bayate

a
Beykent University, Engineering and Architecture Faculty, Chemical Engineering Department,
Ayazağa, İstanbul, Turkey. E-mail: hasanuslu@gmail.com.
b
Malaviya National Institute of Technology (MNIT), Department of Chemical Engineering,
Jaipur, Rajasthan, India.
c
Chemical and Materials Engineering Department, Faculty of Engineering, King
Abdulaziz University, Jeddah, Saudi Arabia.
d
Department of Chemical Engineering, Polytechnic School, Federal University of Bahia
(UFBA), Salvador, Brazil.
e
İstanbul Esenyurt University, Engineering and Architecture Faculty, Industrial Engineering
Department, Esenyurt, İstanbul, Turkey.
ABSTRACT
In this paper, the equilibrium and kinetic studies on the extraction of 2,4,6-Trinitrophenol (picric
acid) (0.021 - 0.061 kmol·m-3) using Amberlite LA2, a secondary amine (ALA2: 0.118 - 0.588
kmol·m-3) dissolved in a polar active solvent, methyl-iso-butyl ketone (MIBK) are performed.
Also, the temperature effect (293.2 ± 1 K, 303.2 ± 1 K and 313.2 ± 1 K) on the extraction
mechanism and efficiency is evaluated. Thermodynamic parameters like the change in entropy
and enthalpy are determined. From the values of loading ratio (Z < 0.5), it is inferred that the
amine molecule form 1:1 complex with the acid molecule in the organic phase. The mass transfer
coefficient (kL = 3.1×10-5 m·s-1) of picric acid in MIBK is determined. The Hatta number is
calculated, and observed to vary in the range of 0.0032 to 0.0054, indicating that there is a very
slow chemical reaction taking place between the acid and the amine molecule in the bulk of the
organic phase. The reaction order is estimated to be 0.9 w.r.t picric acid, and 0.6 w.r.t ALA2
with rate constants of 14.95 × 10-6 (kmol·m-3)-0.5s-1, and 8.94 × 10-7 (kmol·m-3)-0.5s-1, for forward
and backward reaction, respectively. Kinetic and potential energies of components during
reactive extraction have been determined by molecular dynamic modeling.

Keywords: Picric acid; extraction; equilibrium; kinetics; thermodynamics; molecular modeling.

1
1. INTRODUCTION
Large amount of wastewater containing pollutants is produced by the processing industries.
Among these pollutants, the presence of phenols and/or nitrophenols [1,2] in the wastewater
streams causes tremendous environmental problems. The prime sources of phenols are
wastewater from different industries like paint, pesticide, coal conversion, polymeric resin,
gasoline, rubber proofing, steel, petroleum and petrochemicals [3-5].

Nitrophenols are essential ingredients used to produce many important chemicals like pesticides,
herbicides, explosives, dyes and plasticizers [1,2]. These nitrophenols are non-biodegradable
particularly at high concentrations [6], and can toxicate the lives of growing plants,
microorganisms and humans [1,2]. As per the guidelines provided by US Environmental
Protection Agency (EPA) and European Union (EU), the amount of phenols in wastewater
should not exceed 1 mg·L-1, and 5 mg·L-1 in Japan [7]. Poly-nitro-aromatic, 2,4,6-trinitrophenol
or picric acid [8] is one of the nitrophenols exists in the wastewater coming out of the industry
and causes harmful effects on the environment [1,2]. Serine et al. [9] investigated the extraction
of picric acid from the aqueous chloride solutions into a solvent phase consisting of
trioctylphosphine oxide (TOPO) or trioctylamine (TOA) diluted by cyclohexane. The association
of acid molecule took place with one or two TOPO molecule, and with only one TOA molecule
as confirmed from the results obtained from UV-visible spectrum of the organic phase. The
TOPO spectrum showed dependency on the free extractant amount whereas TOA spectrum did
not. Kusakabeet al. [10] performed the extraction of phenol and nitrophenols using TOPO +
cyclohexane as the extract phase at 298 K. They also analyzed the association of these phenols
with TOPO in the organic phase. It was observed that a linear relationship existed between
logarithm of partition coefficient of phenols and logarithm of their association constants with
TOPO. Equilibrium extraction of picric acid from water or solutions containing (NH4)2SO4,
Na2SO4 and NH4NO3 was studied by 4-methyl-2-pentanone, 2-methylpropyl ethanoate, and
diisopropyl ether at 303.15 K and at a pH of 3 [11]. The extraction power of solvents was found
to be in the order of 4-methyl-2-pentanone > 2-methylpropyl ethanoate > diisopropyl ether. The
presence of ammonium nitrate in the solution enhanced the extraction yield but the presence of
sulfates showed irregular behavior and limited the extraction process. Recovery of picric acid
from aqueous solutions was studied at 298 K by TOA dissolved in isoamyl alcohol, 1-octanol,

2
and 1-decanol [12]. Isoamyl alcohol with TOA showed highest extraction efficiency with a
distribution coefficient of 19.33 (extraction efficiency = 95.08 %). The author also proposed
Linear solvation energy relationship (LSER) models to predict the distribution coefficient. The
experimental study on the equilibrium and kinetics for the adsorption of picric acid on the
calcined and uncalcined mesoporous silicate (MCM-41) adsorbent was studied [13]. MCM-41
showed highest adsorption capacity of picric acid in the pH range of 1.4 to 4. Equilibrium,
kinetic, and thermodynamic adsorption study of picric acid [14] were performed by a weak basic
adsorbent (Amberlite IRA-67) at three different temperatures (298 K, 308 K, and 318 K).
Optimum amount of Amberlite IRA-67 was found to be 1 g. Langmuir isotherm and pseudo-
second-order model best described the experimental data of equilibrium and kinetic study,
respectively. Thermodynamic parameters like ∆Hads0 (= -21.204 kJ· mol-1), ∆Sads0 (= -200.043
J· mol−1·K−1), and ∆Gads0 were also estimated.

The separation methods such as the advanced oxidation [15,16], adsorption [17] and biological
treatment [18-20] were tried for the treatment of wastewater to remove different harmful
compounds. These technologies have drawbacks of high costs and the formation of other toxic
side products [2] and limitation for application in industrial scale [8]. Solvent extraction is one of
the most important techniques used in the pretreatment of wastewater when the aqueous solution
is highly concentrated with phenol [21, 22]. Using this method, most phenols in the wastewater
can be recovered as a by-product with economic benefit, and also COD value of the effluent
wastewater stream can be significantly reduced which makes the biochemical treatment easier
[23]. Phenol or phenolic wastewater generally can be treated by using methods such as
biodegradation, thermal decomposition and adsorption. Under aerobic or anaerobic conditions,
some aerobic bacteria and fungi utilize phenol as a source of carbon and energy, and can degrade
it [24] but survival of microorganisms at high concentrations of phenol limits the process.
Moreover, longer time is required for the degradation of phenol. Again, thermal decomposition
could be used to treat phenolic wastewater but needs high energy consumption. In the wet air
oxidation process, the temperature varies from 200 to 330°C, and pressure lies between 2 and
20 MPa [25, 26]. Therefore, thermal treatment is not energy efficient way of treating phenol by
degradation. Adsorption is proposed to be an effective method of treating dilute phenolic
wastewater [27-30] but relatively high cost of activated carbon or other adsorbents make this

3
process uneconomical to treat high concentrated phenolic wastewater. Due to the different
solubility of phenolic compounds in solvent/extractant and in water, solvent extraction is more
preferred than any other wastewater treatment method.

Recently, a three phase liquid extraction (TPLE) system, a new technique was developed for the
extraction of multi-component mixtures using multiphase separation in one extraction stage. The
process is advantageous as there exists differences in physicochemical properties of the three
coexisted phases [31]. It is also efficient and easy to be operated in a continuous industrial scale
[31]. In TPLE process, an ionizable compound is extracted from the donor phase to the organic
phase, and then back extracted to the acceptor phase [32]. Liquid-liquid-micro-extraction was
involved in the organic membrane phase [32, 33]. Liquid membranes were employed for the
extraction of phenol from wastewater [34-37] by using unsupported liquid membrane [38, 39],
liquid membrane supported with a hollow fiber [40] and hollow fiber membrane [20]. These
techniques were applied for the treatment of wastewater containing o-nitrophenol [41], p-
nitrophenol [42-45] and 2, 4 nitrophenol [46].

In designing an extraction process, there is a need to gather information and data on the
equilibrium and kinetics of extraction, and mass-transfer parameters for the acid-extractant-
solvent system to be considered. Sufficient work on the equilibrium of several acid-amine
systems [12, 47-54] is available in the literature, but the knowledge of interaction of picric acid
with new extractant-solvent system is inadequate. Therefore, the objective of this work is to
study the liquid-liquid extraction of picric acid from wastewater by a secondary amine,
Amberlite LA2 in methyl-iso-butyl ketone (MIBK) as the acceptor phase. Equilibrium, kinetics,
and thermodynamic data were generated experimentally, and analyzed in the present work to be
used for scaling-up the extraction process in the industrial application.

2. EXPERIMENTAL

2.1. Materials

4
Picric acid (2,4,6-trinitrophenol) (Sigma-Aldrich, purity >99 % in mass), Amberlite LA-2
(Sigma-Aldrich, purity > 99 % in mass), MIBK (Sigma-Aldrich, purity > 99 % in mass), and
NaOH (Merck, purity > 99 % in mass) were used as supplied.

2.2. Equilibrium
Three initial concentrations of picric acid (0.021 kmol·m-3, 0.041 kmol·m-3, and 0.061 kmol·m-3)
were prepared by using distilled water. 0.061 kmol·m-3 concentration was the highest solubility
of acid in water. For the organic solvent phase, five different initial concentrations of Amberlite-
LA2 in MIBK were considered (0.118 kmol·m-3, 0.235 kmol·m-3, 0.353 kmol·m-3, 0.470 kmol·m-
3
, 0.588 kmol·m-3). MIBK was selected as a diluent for making the organic phase. Equal volumes
(20 mL) of the organic and aqueous phases were mixed in an Erlenmeyer flask by shaking for 2
h at 50 rpm and at constant temperature in a temperature controlled shaker. After equilibrium,
the phases were left for 2 h to have a distinct separation of phases. After phase separation, the
picric acid concentration in the aqueous phase was measured by base titration (0.01 N NaOH)
with phenolphthalein as the indicator. Picric acid concentration in the organic phase after
equilibrium was determined by mass balance. Equilibrium experiment and chemical analysis
were performed in duplicate, and the average values were used for calculation. Equilibrium
experiments were done at three different equilibrium temperatures (293.2 ± 1 K, 303.2 ± 1 K,
and 313.2 ± 1 K).

2.3. Kinetics
Kinetic experiments were done in a custom-made glass stirred cell with inside diameter of 0.036
m, and height of 0.113 m as schematically shown in Figure 1. Two agitators (made of stainless
steel) were placed at the center of each phase. The cell has water inlet for keeping temperature
constant. 50 mL of each phase was taken to achieve experimental data on the kinetics of reactive
extraction. The aqueous phase was poured to the cell first, and then the organic phase was poured
slowly into the agitated cell without disturbing the aqueous-organic interface. The stirring speed
was selected as 20 rpm, 40 rpm, 60 rpm and 80 rpm so that the interface should not get disturbed
and to maintain the interfacial area almost equal to the geometric area of the cell. Samples were
collected from the bottom side of the cell at 20 min interval of time until equilibrium is reached.
Aqueous and organic phase acid concentrations at different time were determined as mentioned

5
in the equilibrium studies. Kinetic experiments and chemical analysis were performed in
duplicate, and the average values were used for the analysis and calculation of kinetic
parameters.

6
Agitator

Sealing
Water In
36
Stirrers
Interface

113

Sample Port

Water Out

Figure 1. Experimetal set-up of the stirred cell used in this study (all dimensions are in mm).

3. RESULTS AND DISCUSSION

3.1. Equilibrium
The extraction of picric acid at equilibrium by a secondary amine (Amberlite LA2: ALA2)
,which is not soluble in water therefore it is not as a contaminant, dissolved in an active polar
solvent (MIBK) can be explained by the theory of Mass Action [47]. It is the undissociated part
of the acid molecule which forms complex with amine molecule. At the interface of the aqueous
and organic phase, the following reaction takes place.

mHPic + ALA2 ↔ (HPic) m (ALA2) (1)

[(HPic) m (ALA2)]
Equilibrium complexation constant, K E = (2)
[HPic]m [ALA2]

7
To analyze the performance of the extraction process, the experimental results are generally
presented in terms of distribution coefficient [KD, a ratio between total number of acid molecules
in the organic phase (Corg) and aqueous phase (Caq) at equilibrium] and loading ratio [Z, a ratio
between acid concentration in the organic phase at equilibrium and the initial extractant
concentration ( [ALA2 ]in ) in the organic phase] as given by Eqs. 3 and 4, respectively.

C org
KD = (3)
C aq

Corg
Z= (4)
[ALA2]in

A mathematical model based on the loading ratio, Z for the formation of different types of picric
acid-ALA2 solvates (1:1, 2:1 etc.) can be presented, and the stoichiometry of the extraction
process can be estimated from the plots of Z versus the acid concentration in the aqueous phase
at equilibrium. At lower loading of the organic phase with acid (Z < 0.5), it is assumed that
mostly 1:1 types of solvates are formed, and plots of Z/(1-Z) against [HPic] leads to a straight
line passing through the origin, and the slope gives the value of complexation constant (K11).

Z
= K11[ HPic ] (5)
1− Z

The equilibrium extraction of picric acid from aqueous solution with MIBK (physical) and with
ALA2 + MIBK (chemical) were carried out. Extraction with pure MIBK gave a distribution
KD
coefficient of 0.17 (extraction efficiency, %E = = 14.8%) at 0.061 kmol· m-3 of initial
1+ KD
picric acid concentration showing a very low distribution of acid in MIBK, and hence
necessitates the use of an extractant. Therefore, a secondary amine with five different
concentrations in MIBK (ALA2: 0.118 to 0.588 kmol· m-3) was used to carry out further
experiments at three different equilibrium temperatures (293.2 ± 1 K, 303.2 ± 1 K and 313.2 ± 1

8
K). The equilibrium results are shown in Figure 2. The addition of ALA2 showed a considerable
increase in the extraction efficiency. Also, by increasing the concentration of ALA2 in MIBK
from 0.118 to 0.588 kmol· m-3 improved the uptake of picric acid, and facilitates easy transfer of
acid molecules.

100

80
Degree of Exraction (%)

60

40

20

0.1 0.2 0.3 0.4 0.5 0.6


-3
[ALA2], kmol.m

(a)

100

80
Degree of Exraction (%)

60

40

20

0.1 0.2 0.3 0.4 0.5 0.6


-3
[ALA2], kmol.m

9
(b)

100

Degree of Exraction (%) 80

60

40

20

0.1 0.2 0.3 0.4 0.5 0.6


-3
[ALA2], kmol.m

(c)

Figure. 2. Degree of extraction at different picric acid and Amberlite LA2 concentrations. *,
0.021 kmol· m-3; ■, 0.041 kmol· m-3; ∆, 0.061 kmol· m-3 (a) 293.2 ± 1 K (b) 303.2 ± 1 K (c) 313.2
± 1 K.

The loading of acid molecules (Z) in the organic phase may be used to determine the
stoichiometry of the overall extraction process. In the present study, the values of Z were plotted
against equilibrium concentrations of acid in the aqueous phase for five different ALA2
concentrations in MIBK, and at three different temperatures as shown in Figure 3. The observed
values of Z (0.036 to 0.221 at 298.2 ± 1 K; 0.036 to 0.204 at 303.2 ± 1 K; 0.036 to 0.187 at 313.2
± 1 K) showed an increasing trend with the equilibrium concentration of acid in the aqueous
phase, and which is irrespective of the initial ALA2 concentration in MIBK. These interpretation
were in good agreement with the results published by Kings and Kertes [52], and also validated
the formation of same type of acid-amine solvates in the organic phase at different ALA2
concentrations. In this study, the equilibrium constants (K11) were estimated by plotting Z/(1-Z)
versus [HPic], and fitting it by a straight line for low values of loading ratio. From the slope of

10
this line, the complexation constants (K11) were determined and shown in Table 1. An average
value of K11 [= 16.71 (kmol·m-3)-1] was calculated , and used to determine the intrinsic kinetics of
extraction process at 293.2 ± 1 K.

0.25

0.20

0.15
Z

0.10

0.05

0.00
0.00 0.01 0.02 0.03 0.04 0.05
-3
Caq, kmol.m

(a)

0.25

0.20

0.15
Z

0.10

0.05

0.00
0.00 0.01 0.02 0.03 0.04 0.05
-3
Caq, kmol.m

(b)

11
0.25

0.20

0.15
Z

0.10

0.05

0.00
0.00 0.01 0.02 0.03 0.04 0.05
-3
Caq, kmol.m

(c)

Figure 3. Loading of picric acid on Amberlite LA2 with MIBK. *, 0.021 kmol· m-3 ; ■, 0.041
kmol· m-3; ∆, 0.061 kmol· m-3; (a) 293.2 ± 1 K, (b) 303.2 ± 1 K, and (c) 313.2 ± 1 K.
The results obtained in the current study were compared with the previous work of the Uslu et al.
[12, 14]. In the first study [12], picric acid (0.061 kmol·m-3) was extracted using TOA (0.25 to
3.75 kmol·m-3) dissolved in isoamyl alcohol, octan-1-ol, decan-1-ol at 298 K. The diluent
showed solvation ability of the acid molecule with TOA (at an intermediate concentration of
1.75 kmol·m-3) in the order of isoamyl alcohol (KD = 19.333), octan-1-ol (KD = 5.1), and
decan-1-ol (KD = 3.066). LSER modeling was done to predict distribution coefficient. In the
second study [14], adsorption of picric acid (4 to 14 g·L-1) was done by using Amberlite IRA-67
(0.25 to 1.25 g) at three different temperatures (298 K, 308 K, and 318 K). Adsorption of picric
acid was investigated in terms of equilibrium, kinetics, and thermodynamic conditions. The
Langmuir isotherm and pseudo-second-order model best described the equilibrium and kinetic
data, respectively. The use of IRA67 (1 g, optimum amount) gave removal of 96 % picric acid
(4 g·L-1) from aqueous solution at 298 K. FTIR results showed that ion exchange during the
adsorption of acid by Amberlite IRA-67 was through hydrogen atoms. In the current study,
equilibrium and kinetic experiments on the extraction of picric acid (0.021 to 0.061 kmol·m-3)
were performed by using Amberlite LA2 (0.118 to 0.588 kmol·m-3) dissolved in methyl-iso-

12
butyl ketone at three different temperatures (293.2 ± 1 K, 303.2 ± 1 K and 313.2 ± 1 K). In the
equilibrium experiment, a maximum of 99.52 % (KD = 209) recovery of acid (0.021 kmol·m-3)
was achieved using ALA2 (0.470 kmol·m-3) + MIBK system. From the kinetic experiments, the
reaction order and the rate constants were estimated. Also, molecular dynamic modeling of the
reacting system was performed to determine kinetic and potential energies of molecules
interacting during the extraction process.

Table 1. The estimated values of the equilibrium constants (K11) with R2 for picric acid (0.021 to
0.061 kmol· m-3) reactive extraction at three different temperatures.

Temperature / K Caq / kmol·m-3 K11 / (kmol·m-3)-1 R2 / (-)


0.021 25.97 0.876
293.2 ± 1 0.041 17.12 0.966
0.061 7.04 0.953
0.021 19.58 0.939
303.2 ± 1 0.041 12.51 0.956
0.061 5.45 0.956
0.021 15.15 0.943
313.2 ± 1 0.041 9.89 0.958
0.061 4.68 0.968

Effect of Temperature
The effect of temperature (293.2 ± 1 K to 313.2 ± 1 K) on the performance of picric acid reactive
extraction process by using ALA2 dissolved in MIBK was also studied, and the equilibrium
results are shown in Figure 2. It can be seen from Figure 2 that the acid distribution into the
organic solvent phase decreases sharply by increasing the temperature. This is because at higher
temperature back-extraction of the acid from the organic solvent phase to the fresh aqueous
phase took place which decreases the extraction efficiency. The equilibrium (K11) constants at
three different temperatures were estimated (Table 1). With change in the temperature of the
extraction process the parameters like pKa of acid, the acid-amine interaction, the solubility of
the acid in both phases, the extractant basicity, and water co-extraction [53] get affected. The

13
increase in the temperature decreases the pKa value of acid which leads to the dissociation of
acid molecule in the aqueous phase, and decreases the availability of the undissociated acid
molecule to be extracted by the extractant-solvent phase. The solubility of organic acids in water
increases as there is an increase in temperature [55]. The temperature also has effect on the acid-
ALA2 complex formation. At a temperature of 293.2 ± 1 K, higher values of K11 were obtained
as compared to what were obtained at 313.2 ± 1 K. Therefore, at higher temperatures the
reaction-extraction system becomes more unstable facilitating back-extraction [54]. At the
equilibrium, due to the formation of the acid-amine complexes the randomness of the system
decreases which also decreases the net entropy of the system. For constant enthalpy and constant
entropy in the range of studied temperature, the equilibrium constant (K11) may be related to the
temperature (T) by Eq. 6 [56].

− ∆H ∆S
ln K 11 = + (6)
RT R

Plots of lnK11 vs 1/T are fitted linearly to determine the enthalpy (∆H) from the slope and the
entropy (∆S) from the intercept (Table 2). Larger change in the values of ∆H and ∆S showed
that the reaction is exothermic, and sensitive to the equilibrium. Also, the negative value of ∆H
dictates that the extraction accompanied with reaction favored at low temperature.

Table 2. Values of entropy and enthalpy for the picric acid reactive extraction using ALA2
dissolved in MIBK.

Caq / kmol·m-3 -∆H, kJ·mol-1 -∆S, J·mol-1·K-1 R2


0.021 49.61 141.45 0.950
0.041 48.64 140.60 0.964
0.061 44.69 129.35 0.970

Effect of pH

14
Amberlite LA2 is a weak anion exchanger which can be used to significantly alter the polarity of
the target molecules. This secondary amine can react with the neutral form of the phenolic
compound, picric acid to form a complex that is generally extracted into the organic phase. The
anionic form of picric acid would dominate at a pH greater than pKa, and therefore, it is
important to understand the effect of pH on the extraction mechanism. Since, the pH of aqueous
solution of phenolic compound affects the molar fraction of neutral phenol, the distribution of
the molecule between the phases greatly depends on the equilibrium pH. The effect of pH on the
distribution coefficient at 0.061 kmol· m-3 of picric acid concentration and at 0.470 kmol· m-3 of
Amberlite LA2 concentration is shown in Figure 4 at 293.2 ± 1 K temperature. The values of KD
increased up to a pH of 5, and then started decreasing. It occurred because the anionic form of
picric acid appears at higher pH which has little reactivity with ALA2, so KD value decreased
with an increase in the pH of aqueous solution.

5.0

4.5
Distribution coefficient (KD)

4.0

3.5

3.0

2.5

2.0

1.5
0 2 4 6 8 10
pH

Figure 4. The effect of pH on the distribution coefficient for the extraction of picric acid (0.061
kmol· m-3) with Amberlite LA2 (0.470 kmol· m-3) dissolved in MIBK at temperature T = 293.2 ±
1 K.

3.2. Kinetics

15
The mechanism of simultaneous extraction and chemical reaction in a stirred cell was proposed
by Doraiswamy and Sharma [57]. Based on the film and renewal theories, they established the
fact that the specific extraction rate gets effected by the chemical reaction, and to identify this
fact there is a need to study the effect of physico-chemical and hydrodynamic parameters of the
system on the extraction. According to their theory, they indentified four regimes in the reactive-
separation system: (i) very slow, (ii) slow, (iii) fast and (iv) instantaneous (Table 3). Therefore,
to identify the reaction regime, and to determine the intrinsic kinetics of the reactive extraction
process, the effect of agitation speed (N), phase volume ratio (Vorg/Vaq), concentration of picric
acid and ALA2 on the specific extraction rate is studied and analyzed thoroughly.

Table 3. Limiting regimes for simultaneous reaction (irreversible)and extraction in a stirred cell
[47].

Effect on the specific extraction rate (kmol· m-2·s-1)


Hatta Volume
Stirrer
Regime Description Number [ HPic org ] [ ALA2] ratio of
speed (N,
(Ha) kmol· m-3 kmol· m-3 phases (-
rpm)
)

1. Very slow α [HPicorg ]m α[ALA2]n None α Vorg

Increases
with
<< 1
2. Slow α [ HPic org ] None increase in None
the stirring
speed
m+1 n
3. Fast α [ HPic org ] 2
α [ ALA2 ] 2 None None

Increases
>> 1 with
4. Instantaneous None α [ ALA2 ] increase in None
the stirring
speed

16
3.2.1. Determination of Mass-Transfer Coefficient (kL) in MIBK
Mass transfer coefficient has to be found out to determine the reaction regime in reacting-
extracting system. Therefore, its value is from the kinetic experiments carried out using only
pure MIBK. The molecules of picric acid show almost insufficient distribution in water-MIBK
system. Assuming, the diffusion film resistance in the aqueous phase is negligible, and
considering only the existence of diffusion film resistance in the organic phase, an equation
representing the mass-transfer rate or molar flux can be written in form of Eq. 7.

Vorg d C org
A dt
( *
= k L C org − C org ) (7)

where, A is the interfacial area (m2), Vorg is the organic phase volume (m3) and Corg
*
represents the

acid concentration at the aqueous-organic interface.

By integrating Eq. 7, the acid concentration in the organic phase can be expressed in terms of
time (Eq. 8).

 Corg*
 kL A

ln * = t (8)
C −C  V
 org org  org

 Corg *

From the plot of ln *  versus time (t), and from linear fit (Figure 5), the value of mass-

 Corg − Corg 
transfer coefficient of picric acid (kL) in MIBK was determined at a stirrer speed of 60 rpm, at
0.061 kmol· m-3 of acid concentration, and with equal volumes of phases (Vorg/Vaq = 1). From the
slope of Figure 5, the value of kL was found to be 3.1× 10-4 m·s-1.

17
3.0

2
R = 0.97
2.5 -4
Slope = 6.32x10

ln[C org/(C org - Corg)]


2.0

1.5
*

1.0
*

0.5

0.0

0 500 1000 1500 2000 2500 3000

t, sec

 Corg *

Figure 5. The plot of ln *  versus time (t) to determine kL.

 Corg − Corg 

3.2.2. Reaction Kinetics


Commonly, the simultaneous extraction and reaction between the solute (picric acid) and the
solvent (ALA2 + MIBK) molecule is a reversible process. Therefore, in determining the
extraction kinetics, the application of the initial rate method (governed only by the forward
reaction) could be a better approximation which eliminates the problem of reversibility [58].
Hence, initial specific rate, RHPic,0 (kmol· m-2·s-1) was calculated from the experimental data on
kinetics, and using Eq. 9.

Vorg  dCorg 
RHPic , 0 =   (9)
A  dt t =0

The kinetic equation and the Hatta number (Ha) is given by Eqs. 10 and 11, respectively.

*
R HPic,0 = k [C org ]α [ ALA2 in ] β (10)

18
2 * α −1
k[Corg ] [ALA2in ]β DHPic
Ha = α + 1 (11)
kL

where, α and β are the orders of the reaction with respect to picric acid and ALA2, respectively;
k is the reaction rate constant; and DHPic is the diffusion coefficient of acid into diluent (m2·s-1).

The initial specific reaction rates (RHPic,0) were calculated from the experimental values at
different picric acid and ALA2 concentration, agitation speed, and phase-volume ratio. The
effects of agitation speed (N), and phase-volume ratio (Vorg/Vaq) on RHPic,0 will confirm the
regime of reaction [57].

3.2.3. Effect of Stirrer Speed (N) on RHPic,0


The diffusion or the chemical reaction taking place in the reacting system governs the extraction
process with chemical reaction in a stirred cell. Generally, in the diffusion controlled regime, an
increase in the stirring speed will increase the extraction rate and reaches to a constant value
afterwards but in the reaction controlled regime, there is no effect of stirring speed on the
extraction rate [58]. In the latter case, the chemical reaction limits the overall rate of mass
transfer, and diffusion contributions are almost negligible or do not exist. The reason may be that
the interfacial boundary layer becomes thinnest minimizing the individual film resistance to mass
transfer. In the present study, to determine the kinetics of extraction the experiments were
performed at four different stirring speed (20, 40, 60, and 80 rpm), at 0.061 kmol· m-3 of picric
acid concentration, at 0.588 kmol· m-3 of ALA2 in MIBK, at 1:1 Vorg/Vaq ratio, and at a
temperature of 293.2 ± 1 K (Figure 6). This study will confirm the hydrodynamic effect on the
initial extraction rate. The highest stirring speed was chosen to be 80 rpm. It was found that
beyond this speed the interfacial area between the aqueous and organic phase got disturbed, and
hence a constant area of mass transfer could not be maintained. In this range of stirrer speed, the
RHPic,0 was found to be almost constant (0.374 × 10-6 to 0.382 × 10-6 kmol· m-2·s-1) showing no
affect on the specific extraction rate. As per Table 3, and from the effect of N, it may be said that
the simultaneous extraction and reaction will fall either in Regime 1 (very slow) or 3 (slow).

19
0.5

0.4

RHPic, 0 x 10 , kmol/m s
2

0.3
6

0.2

0.1
20 40 60 80 100

N, rpm

Figure 6. Effect of N on RHPic,0 for picric acid reactive extraction with Amberlite LA2 in MIBK
(T = 293.2 ± 1 K, Vorg/Vaq = 1, CHPic,in = 0.061 kmol· m-3, [ALA2]in = 0.588 kmol· m-3).

3.2.4. Effect of Volume Ratio of Phases (Vorg/Vaq) on RHPic,0


Further study on the kinetics was carried out at four volume-phase ratios (Vorg/Vaq = 0.5, 1, 1.5
and 2) using 0.061 kmol· m-3 of picric acid solution, and 0.588 kmol· m-3 of ALA2 in MIBK at 40
rpm and at 293.2 K to make a conclusion between Regime 1 and 3. The initial specific extraction
rate was calculated and plotted against Vorg/Vaq in Figure 7. This showed a linear variation of
RHPic,0 with Vorg/Vaq confirming that the reaction between picric acid and ALA2 molecule is
taking place in the bulk of the organic solvent phase. Therefore, based on the analysis done in
previous two paragraphs, and the guidelines provided by Doraiswamy and Sharma, it may be
concluded that the simultaneous extraction and chemical reaction of picric acid with ALA2 +
MIBK system falls in Regime 1 (a very slow chemical reaction).

20
0.40

0.35

RHPic, 0 x 10 , kmol/m s
2
0.30

0.25
6

0.20

0.15

0.10
0.5 1.0 1.5 2.0

Vorg/Vaq

Figure 7. Variation of RHPic,0with Vorg/Vaq for picric acid reactive extraction with Amberlite LA2
in MIBK (T = 293.2 ± 1 K, N = 40 rpm, CHPic,in = 0.061 kmol· m-3, [ALA2]in = 0.588 kmol· m-3).

3.2.5.Order of the Reaction Kinetics (α, β)


To determine the reaction order with respect to acid (α), kinetic experiments were performed by
changing initial picric acid concentration from 0.021 to 0.061 kmol· m-3 at 0.588 kmol· m-3 of
ALA2 + MIBK, at 40 rpm, and 293.2 ± 1 K. The kinetic data were used to calculate RHPic,0, and
*
plotted against Corg as shown in Figure 8. Calculated RHPic,0 showed a linear trend with picric acid

concentration at equilibrium. The regression analysis gave a value of α equal to 0.9. Similarly,
the reaction order with respect to ALA2 (β) was determined by varying ALA2 concentration
from 0.118 to 0.588 kmol· m-3 at 0.061 kmol· m-3 of acid concentration, at 40 rpm, and at 293.2 K
(Figure 9). The value of β was estimated to be 0.6. The forward reaction rate constant [k =
14.95× 10-6 (kmol·m-3)-0.5s-1] was obtained using Figure 10, and from the value of equilibrium
constant, the backward reaction rate constant [k-1 = 8.94 × 10-7 (kmol·m-3)-0.5s-1] was estimated.

21
1.0

0.9

0.8

-2 -1
RHPic, 0 x 10 , kmol m s
0.7

0.6

0.5
6

0.4

0.3

0.2

0.1
0.01 0.02 0.03 0.04 0.05 0.06
* -3
C org, kmol m

*
Figure 8. Variation of RHPc,0 with Corg for picric acid reactive extraction with Amberlite - LA2 in

MIBK (T = 293.2 ± 1 K, N = 40 rpm, Vorg/Vaq= 1, [ALA2]in = 0.588 kmol· m-3).

0.9

0.8
-2 -1
RHPic, 0 x 10 , kmol m s

0.7

0.6
6

0.5

0.4

0.3
0.10 0.15 0.20 0.25 0.30 0.35 0.40
-3
[ALA2]in, kmol m

Figure 9. Variation of RHPic,0with [ALA2]in for picric acid reactive extraction with Amberlite
LA2 in MIBK (T = 293.2 ± 1 K, N = 40 rpm, Vorg/Vaq= 1, CHPic,in = 0.061 kmol· m-3).

22
0.8

0.7

0.6

-3 -1
RHPic, 0 x 10 , kmol m s
6 0.5

0.4

0.3

0.2

0.1
0.02 0.03 0.04 0.05
* 0.9 0.6
[Corg] [ALA2]in

Figure 10. Estimation of reaction rate constant (k).

3.2.6. Validation of the Reaction Regime


For α = 0.9 and β = 0.6, the Hatta number (Ha) was calculated using the following expression,
Eq. 12.

*
1.05k[Corg ]−0.1[Torg, in ]0.6 DHPc
Ha = (12)
kL

The value of DHPic was determined from the equation proposed by Reddy-Doraiswamy [59] and
Wilke-Chang [60] as 1.61×10-9 m2·s-1 at 293.2 ± 1 K. The values of Ha were found in the range
of 0.0032 to 0.0054 at 293.2 ± 1 K which also validated the Regime 1. The values of RHPic,0 were
predicted by the kinetic equation ( RHPic,0 = 14.95 × 10 -6 [C org
*
]0.9 [ ALA2 ]0in.6 ) and plotted against

experimental values of RHPc,0 in Figure 11. Therefore, these equilibrium, kinetic and
thermodynamic data and findings on the reactive extraction of picric acid with ALA2 + MIBK
system would be helpful in designing a continuous extraction process.

23
0.7

0.6

-1
RHPic,0 (pred) x 10 , kmol m s
-2
0.5
6

0.4

0.3

0.2

0.1
0.1 0.2 0.3 0.4 0.5 0.6 0.7
6 -2 -1
RHPic,0 (exp) x 10 , kmol m s

Figure 11. Comparison of predicted and experimental values of specific reaction rates, RHPic,0
(kmol· m-2·s-1).

3.2.7. Molecular Dynamic Simulation


CHARMM simulations were run with the LAMMPS [61] simulation package, using a 100 fs
time-step and using PPPM long-range electrostatics and Nose-Hoover [62] temperature/pressure
coupling.Starting from the known structure, molecular simulations were run in the constant stress
ensemble at 293 K and 1 bar, allowing the simulation cell to change size and shape. Different
boxes were created using playmol package with 200 picric acid, 200 MIBK, and 200 secondary
amine molecules and used as initial configurations. The constant temperature simulations were
run for 5 x 106 fs at 293 K, and data collected over the last 1 ns. Figure 12 shows the energy
change during the intra-molecular interaction for the system of picric acid + secondary amine +
MIBK which was obtained from molecular dynamics model.

24
+ +

Secondary amine + MIBK + Picric acid

Figure 12. Simulation box for the mixture of (picric acid + secondary amine + MIBK) in the
organic phase after extraction.

The figures 13 and 14 showed the statistics displayed below the image are kinetic energy,
potential energy (from interactions between atoms and interactions with the neighbours).

25
Figure 13. Behaviour of kinetic energy of different pure components using the technique of
molecular dynamics.

Figure 14. Behavior of potential energy of different pure components using the technique of
molecular dynamics

26
It is important to remember that the kinetic energy associated with the translational motion can
be exactly separated from other molecular motions, and potential energy consisting of N atoms is
a function of three parameter i.e. r1, r2 (inter-atomic distance), angle and conformation
difference of the molecule in question. The values of kinetic energy per atom, for picric acid,
MIBK and secondary amine were ~-45000 kcal/mol, ~10000 kcal/mol, ~-10000 kcal/mol,
respectively. A range of potential energy for the secondary amine and picric acid molecule was
obtained in the range of ~-1000 to ~1000 kcal/mol. For MIBK, the potential energy value was
observed in the order of ~ -20000 kcal/mol. The resemblance of potential energy between
secondary amine and picric acid molecules may be associated with the similarity in the
packaging degree, molecular weight and free volume.

In figures 15 to 17, the results obtained from the examination of the radial distribution functions
(RDFs) were presented.

Figure 15. Radial distribution function between constituent atoms of picric acid molecule using
molecular dynamics technique.

27
Figure 16. Radial distribution function between constituent atoms of MIBK molecule using
molecular dynamics technique.

Figure 17. Radial distribution function between constituent atoms of secondary amine molecule
using molecular dynamics technique.

28
Figure 15 for picric acid showed a small shoulder at ~3 A for the CB-OR and CB-O2CM group,
and a high interaction at ~4 A for CB-OR and CB - NC=C group. From the RDFs in Figure 16
for MIBK, it was observed that a first solvation minimum was obtained at ~6.25 A (C=O – O=C
and C=O - HCMM), and a maximum of solvation between C=O-HCMM groups was obtained at
~3.5 A. For secondary amine (Figure 17), the similar values were obtained at ~2.5 A (NR-HNR
and NR-HCMM) and at ~4.25 A (between NR-CR groups), respectively. The inter-atomic
interaction between the compound groups followed the order: CB - NC=C > CB - O2CM > CB -
OR for picric acid; C=O-HCMM > C=O-CR > C=O-O=C for MIBK; and NR-CR > NR-HCMM
> NR-HNR for secondary amine. Much of these results are associated with the high level of
vibration of the energy of hydrogen atoms.

4. CONCLUSIONS
Equilibrium, kinetic and thermodynamic studies were performed for the reactive extraction of
picric acid with Amberlite-LA2 + MIBK in a stirred cell. From the equilibrium data and
calculated values of loading ratio (Z < 0.5), the equilibrium constant of 1:1complex formation in
the organic phase were determined at three different temperatures (293.2 ± 1 to 313.2 ± 1 K).
The kinetic experiments were done to investigate the affect of agitation speed and volume-phase
ratio on the specific extraction rate, and confirm the regime of extraction with reaction. It was
observed that the specific rate of extraction remained almost constant with agitation speed but
increases with increase in the volume-phase ratio. Based on the Hatta number, the criterion
proposed by Doraiswamy and Sharma, the reaction regime was found to be very slow occurring
at the bulk of the organic phase. This proposed extraction-reaction model could be successfully
utilized for designing an extraction process to recover picric acid using Amberlite-LA2 in MIBK
from wastewater streams. Kinetic and potential energies of components during reactive
extraction have been determined by molecular dynamic model.

5. REFERENCES
[1] J. Shen, R. He, L. Wang, J. Zhang, Y. Zuo, Y. Li, X. Sun, J. Li, W. Han, Biodegradation
kinetics of picric acid by Rhodococcus sp.NJUST16 in batch reactors, J. Hazard. Mater. 167
(2009) 193-198.

29
[2] J. Shen, R. He, H. Yu, L.Wang, J. Zhang, X. Sun, J. Li, W. Han, L. Xu, Biodegradation of
2,4,6-trinitrophenol (picric acid) in a biological aerated filter (BAF), Bioresour. Technol. 100
(2009) 1922–1930.
[3] H. Parham, S. Saeed, Ultrasound-assisted solid phase extraction of nitro- and chloro-
(phenols) using magnetic iron oxide nanoparticles and Aliquat 336 ionic liquid, J.
Chromatogr. A 1336 (2014) 34–42.
[4] M. Ahmaruzzaman, Adsorption of phenolic compounds on low-cost adsorbents, A review.
Adv Colloid Interface Sci. 143 (2008) 48-67.
[5] S. Lin, R. Juang, Adsorption of phenol and its derivatives from water using synthetic resins
and low-cost natural adsorbents, A review. J. Environ. Manage. 90 (2009) 1336-1349.
[6] N. Wan, J. Gua, Y. Yan, Degradation of p-nitrophenol by AchromobacterxylosoxidansNs
isolated from wetland sediment, Int. Biodeterior. Biodegrad. 59 (2007) 90-96.
[7] T. Hirooka, H. Nagase, K. Hirata, K. Miyamoto, Degradation of 2,4-dinitrophenol by a
mixed culture of photoautotrophic microorganisms, Biochem. Eng. J. 29 (2006) 157–162.
[8] G. Perchet, G Merlina, J.C. Revel, M. Hafidib, C. Richard, E. Pinelli, Evaluation of a TiO2
photocatalysis treatment on nitrophenols and nitramines contaminated plant wastewaters by
solid-phase extraction coupled with ESIHPLC–MS, J. Hazard. Mater. 166 (2009) 284-290.
[9] T. Serine, Y. Katayama, Y. Wakabayashi, H. Naganawa, Solvent extraction of picric acid
from acid aqueous solutions into cyclohexane with trioctylphosphine oxide and
trioctylamine, Solvent Extr. Ion Exch. 7 (1989) 73-86.
[10] S. Kusakabe, Y. Ogihara, T. Sekine, Solvent extraction and association of nitrophenols
with trioctylphosphine oxide in cyclohexane, Bull. Chem. Soc. Jpn. 65 (1992) 2534-2536.
[11] L.M. Ferreira, E. Lopes, Solvent extraction of picric acid from aqueous solutions, J.
Chem. Eng. Data. 41 (1996) 698-700.
[12] H. Uslu, Separation of picric acid with trioctyl amine (TOA) extractant in diluents, Sep.
Sci. Technol. 46 (2011) 1178-1183.
[13] H. Sepehrian, J. Fasihi, M.K. Mahani, Adsorption behavior studies of picric acid on
mesoporous MCM-41, Ind. Eng. Chem. Res. 48 (2009) 6772-6775.
[14] H. Uslu, G. Demir, Adsorption of picric acid from aqueous solution by the weakly basic
adsorbent amberlite IRA-67, J. Chem. Eng. Data. 55 (2010) 3290-3296.

30
[15] O. Gimeno, M. Carbajo, F.J. Beltran, F.J. Rivas, Phenol and substituted phenols AOPs
remediation, J. Hazard. Mater. B. 119 (2005) 99-108.
[16] F.R. Zaggout, N.A. Ghalwa, Removal of o-nitrophenol from water by electrochemical
degradation using a lead oxide/titanium modified electrode, J.Environ. Manage. 86 (2008)
291-296.
[17] D.Tang, Z. Zheng, K. Lin, J. Luan, J. Zhang, Adsorption of p-nitrophenol from aqueous
solutions onto activated carbon fiber, J. Hazard. Mater. 143 (2007) 49–56.
[18] J.L. Ramos, M.M. Gonzalez-Perez, A. Caballero, P. van Dillewijn, Bioremediation of
polynitrated aromatic compounds: plants and microbes put up a fight, Curr. Opin. Microbiol.
16 (2005) 275-281.
[19] S.K. Samanta, B. Bhushan, A. Chauhan, R.K. Jain, Chemotaxis of a Ralstoniasp. SJ98
toward different nitroaromaticcompounds and their degradation, Biochem. Biophys. Res.
Commun. 269 (2000) 117–123.
[20] M.C. Tomei, M.C. Annesini, R. Luberti, G. Cento, A. Senia, Kinetics of 4-nitrophenol
biodegradation in a sequencing batch reactor, Water Res. 37 (2003) 3803-3814.
[21] M. Medir, A. Arriola, D. Mackay, F. Giralt, Phenol recovery from water effluents with
mixed solvents, J. Chem. Eng. Data, 30 (1985) 157-159.
[22] L. Zhen, W. Minghong, J. Zheng, B. Borong, L. Senlin, Extraction of phenol from
wastewater by N-octonoylpyrrolidine, J. Hazard. Mater., B114 (2004) 111-114.
[23] Y. Chufen, Q. Yu, Z. Lijuan, F. Jianzhong, Solvent extraction process development and
on-site trial-plant for phenol removal from industrial coal-gasification wastewater, Chem.
Eng. J., 117 (2006) 179-185.
[24] M.M. Broholm, A. Erik, Biodegradation of phenols in a sandstone aquifer under aerobic
conditions and mixed nitrate and iron reducing conditions, J. Contamin. Hydro. 44 (2000)
239-273.
[25] J.L. Yu, E.S. Phillip, Phenol oxidation over CuO/Al2O3 in supercritical water, Appl.
Catal. Part B: Environ. 28 (2000) 275-288.
[26] J.R. Portela, E. Nebot, E.M. Ossa, Kinetic comparison between subcritical and
supercritical water oxidation of phenol, Chem. Eng. J. 81 (2001) 287-299.
[27] F.A. Banat, B. Al-Bailey, S. Al-Asheh, O. Hayajneh, Adsorption of phenol by bentonite,
Environ. Pollut. 107 (2000) 391-398.

31
[28] V.M. Hebatpuria, A.H. Hassan, H.S. Rho, Immobilization of phenol in cement-based
solidified/stabilized hazardous wastes using regenerated activated carbon: leaching studies, J.
Hazard. Mater. 70 (1999) 117-138.
[29] S. Rengaraj, S.H. Moon, R. Sivabalan, B. Arabindoo, V. Murugesan, Removal of phenol
from aqueous solution and resin manufacturing industry wastewater using an agricultural
waste: rubber seed coat, J. Hazard. Mater. 89 (2002) 185-196.
[30] T. Viraraghavan, F.D.M. Alfaro, Adsorption of phenol from wastewater by peat, fly ash
and bentonite, J. Hazard. Mater. 57 (1998) 59-70.
[31] H. Xiuqiong, H. Kun, Y. Pinhua, Z. Chao, X. Keng, L. Pengfei, W. Juan, A. Zhentao, L.
Huizhou, Liquid-liquid-liquid three phase extraction apparatus: operation strategy and
influences on mass transfer efficiency, Chin. J. Chem. Eng. 20 (2012) 27-35.
[32] L. Zhu, H. Lee, Liquid–liquid–liquid microextraction of nitrophenols with a hollow fiber
membrane prior to capillary liquid chromatography, J. Chromatogr. A 924 (2001) 407–414.
[33] R.E. Terry, W.S. Ho, N.N. Li, Extraction of phenolic-compounds and organic-acid by
liquid membranes, J. Membr. Sci. 10 (1982) 305-323.
[34] X.J. Zhang, Q.J. Fan, X.T. Zhang, Z.F. Liu, New surfactant LMS-2 used for industrial
application in liquid membrane separation in separation technology, in: N.N. Li, H.
Strathmann (Eds.), New York. Umbo Engineering Truslus, 1988, pp. 215-226.
[35] M.L. Wang, K.H. Hu, Extraction of phenol using sulfuric acids salts of trioctylamine in a
supported liquid membrane, Ind. Eng. Chem.Res. 33 (1994) 914-921.
[36] Y.H. Wan, X.D. Wang, X.J. Zhang, Treatment of high-concentration phenolic
wastewater by liquid membrane with N-503 as mobile carrier, J. Membr. Sci. 135 (1997)
263-270.
[37] T. Kataoka, T. Nishiki, S. Kimura, Phenol permeation through liquid surfactant
membrane-permeation model and effective diffusivity, J.Membr. Sci. 41 (1989) 197-209.
[38] M.H. Ma, F.F. Cantwell, Solvent microextraction with simultaneous back-extraction for
sample cleanup and preconcentration: quantitative extraction, Anal. Chemistry 70 (1998)
3912-3919.
[39] M. Ma, F.F. Cantwell, Chain unfolding in an ODS-bonded phase caused by the sorbed
tetra-n-butylammonium ion, Anal. Chem. 71 (1999) 1879-84.

32
[40] S. Pálmarsdóttir, E. Thordarson, L.E. Edholm, J.A. Jönsson, L. Mathiasson, Miniaturized
supported liquid membrane device for selective on-line enrichment of basic drugs in plasma
combined with capillary zone electrophoresis, Anal. Chem. 69 (1997) 1732-1737.
[41] Y. Zeng, Study on recovering of o-nitrophenol by liquid membrane process,
Shuichulijishu (Ch). 3 (1993) 124-135.
[42] S.V. Ho, P.W. Sheridan, E. Krupetsky, Supported polymeric liquid membranes for
removing organics from aqueous solutions I. Transport characteristics of polyglycol liquid
membranes, J. Membr. Sci. 112 (1996) 13-27.
[43] P. Harriott, S.V. Ho, Mass transfer analysis of extraction with a supported polymeric
liquid membrane, J. Membr. Sci. 135 (1997) 55-63.
[44] P.T. Gadekar, Recovery of nitrophenols from aqueous solutions by a liquid emulsion
membrane system, Sep. Sci. Technol. 4 (1992) 427-445.
[45] S.W. Peretti, C.J. Tompkins, J.L. Goodall, A.S. Michaels, Extraction of 4-nitrophenol
from 1-octanol into aqueous solutions in a hollow fiber liquid contactor, J. Membr. Sci. 195
(2001) 193-202.
[46] J. Luan, A. Plaisier, Study on treatment of wastewater containing nitrophenol compounds
by liquid membrane process, J. Membr. Sci. 229 (2004) 235-239.
[47] D. Datta, S. Kumar, Equilibrium and kinetic studies on reactive extraction of nicotinic
acid with tri-n-octylamine dissolved in MIBK, Ind. Eng. Chem. Res. 52 (2013) 14680-14686.
[48] D. Datta, Y. S. Aşçı, A. F. Tuyun, Extraction equilibria of glycolic acid using tertiary
amines: experimental data and theoretical predictions, J. Chem. Eng. Data, 60 (2015) 3262-
3267.
[49] H. Uslu, D. Datta, H. S. Bamufleh, Reactive extraction of phenol from aqueous solution
using tri-octylamine dissolved in alkanes and alcohols, J. Mol. Liq. 212 (2015) 430-435.
[50] D. Datta, S. Kumar, Reactive extraction of picolinic acid using tri-n-octylamine dissolved
in different diluents: effect of solvent polarity, J. Chem. Eng. Data, 60 (2015) 2709-2716.
[51] H. Uslu, D. Datta, S. Kumar, Investigations on the reactive extraction of glyoxylic acid
by amberlite-la2 dissolved in alcoholic diluents, Sep. Sci. Technol. 50 (2015) 2658-2667.
[52] A.S. Kertes, C. King, Extraction chemistry of fermentation product carboxylic acids,
Biotechnol. Bioeng. 28 (1986) 269-282.

33
[53] R. Canari, A.M. Eyal, Temperature effect on the extraction of carboxylic acids by amine-
based extractants, Ind. Eng. Chem. Res. 43 (2004) 7608-7617.
[54] D. Datta, S. Kumar, Intensification of recovery of formic acid from aqueous stream using
reactive extraction with N, N-dioctyloctan-1-amine: effect of diluent and temperature, Chem.
Eng. Commun. 200 (2013) 678-700.
[55] A. Apelblat, E. Manzurola, Solubility of oxalic, malonic, succinic, adipic, maleic, malic,
citric, and tartaric acids in water from 278.15 to 338.15 K, J. Chem. Thermodyn. 19 (1987)
317-320.
[56] J.A. Tamada, C.J. King, Extraction of carboxylic acids with amine extractants. 3. Effect
of temperature, water co-extraction, and process considerations, Ind. Eng. Chem. Res. 29
(1990) 1333-1338.
[57] L.K Doraiswamy, M.M. Sharma, Heterogeneous Reactions: Analysis, Examples, and
Reactor Design, New York: John Wiley & Sons Inc. 1984.
[58] G.J. Hanna, R.D. Noble, Measurement of liquid–liquid interfacial kinetics, Chem. Rev.
85 (1985) 583-598.
[59] K.A. Reddy, L.K. Doraiswamy, Estimating liquid diffusivity, Ind. Eng. Chem. Fundam.
6 (1967) 77-79.
[60] C.R.Wilke, P. Chang, Correlation of diffusion coefficient in dilute solutions, AIChE J. 1
(1955) 264-270.
[61] M. L. Parks, P. Seleson, S.J. Plimpton, R.B. Lehoucq, S. Silling, Peridynamics with
LAMMPS : A User Guide 2 (2010) 32.
[62] D.J. Evans, B.L. Holian, The Nose-Hoover thermostat, J. Chem. Phys. 83 (1985) 4069-
4074.

34
Highlights

• Picric acid extraction was studied by secondary amine.

• The mass transfer coefficient (kL = 3.1 × 10-5 m·s-1) of picric acid in MIBK is estimated

• the thermodynamic parameters (entropy and enthalpy) are determined.

• Molecular dynamic behavior was simulated for the system of picric acid+secondary amine +

MIBK in the organic phase.

35

You might also like