You are on page 1of 9

Journal of Chemical Technology and Biotechnology

J Chem Technol Biotechnol 80:189197 (2005) DOI: 10.1002/jctb.1178

Extraction and pertraction of phenol through bulk liquid membranes


Wojciech Cichy,1 Stefan Schlosser2 and Jan Szymanowski1
1 Institute 2 Department

of Chemical Technology and Engineering, Poznan University of Technology, Pl Sklodowskiej-Curie 2, 60-965 Poznan, Poland of Chemical and Biochemical Engineering, Slovak University of Technology, Radlinskeho 9, 81237 Bratislava, Slovakia

Abstract: The extraction and pertraction of phenol through a bulk liquid membrane (BLM) with Cyanex 923, Amberlite LA-2 and trioctylamine (TOA) as carriers were studied. Cyanex 923 was selected as the best carrier for pertraction. The distribution coefcient of phenol for solvents with carrier and pure n-alkanes, the individual mass-transfer coefcient at the extraction interface and the initial ux of phenol through the extraction interface (JFo ) decreased in the order: Cyanex 923 > Amberlite LA-2 > TOA >> pure n-alkanes. The opposite order was observed for the value of the mass-transfer coefcient in BLM and the maximum ux of phenol through the stripping interface (JRmax ). At constant driving forces the maximum uxes through the extraction and stripping interfaces were similar when amine carriers were used. However, JRmax was lower than JFo for Cyanex 923. Although the kinetics of stripping was the rate-determining step, the ux of phenol was signicantly higher than in pertraction with amine carriers. The adsorption of the carrier at aqueous phase/membrane interfaces was probably responsible for the rapid and slow transfer of phenol through the extraction and stripping interface, respectively. 2004 Society of Chemical Industry

Keywords: phenol; extraction; pertraction; bulk liquid membranes; Cyanex 923; trioctylamine; Amberlite LA-2

NOTATION A Interface surface area (m2 ) c Concentration (g dm3 ) D Diffusion coefcient (m2 s1 ) D Distribution ratio Da Distribution ratio in the absence of carrier J Flux density (mol m2 s1 ) k Individual mass-transfer coefcient (m s1 ) Ka Dissociation constant (mol dm3 ) Kd Dimerization constant (dm3 mol1 ) Kex Extraction constant (dm3 mol1 ) n Degree of association t Time (h) V Phase volume (m3 ) S Extractant (carrier) Kinematic viscosity (m2 s1 )

Superscripts T *

Overall concentration of component Equilibrium value

With overbar Organic phase Without overbar Aqueous phase INTRODUCTION Phenol and its derivatives are toxic pollutants, frequently found in surface and tap waters, and in aqueous efuents from various manufacturing processes.1 As a consequence, they are listed in the US Environmental Protection Agency priority list of pollutants and in the 76/464/EEC Directive of the European Union, related to dangerous substances discarded into the aquatic environment. Solvent extraction is the most often used technique to recover phenol (pKa = 9.98) in its neutral form.2 The use of simple organic solvents, such as benzene, heptane, toluene, methylisobutyl ketone, diisopropyl ether, isopropyl ether, methyltert-butyl ether, isopropyl acetate, etc, is now limited due to the high hydrophilicity and solubility of these solvents in aqueous streams and/or their toxicity. The use of various basic and solvating reagents, including

Subscripts F Feed phase max Maximum ux of phenol M Membrane phase o Initial value R Stripping solution MF Interface on the feed side MR Interface on the strip side

Correspondence to: Stefan Schlosser, Department of Chemical and Biochemical Engineering, Slovak University of Technology, Radlinskeho 9, 81237 Bratislava, Slovakia E-mail: schlosser@cvt.stuba.sk Contract/grant sponsor: Polish Scientic Committee; contract/grant number: DS 32/044/2004 Contract/grant sponsor: VEGA; contract/grant number: 9136/02 (Received 15 December 2003; revised version received 23 June 2004; accepted 13 August 2004) Published online 8 December 2004

2004 Society of Chemical Industry. J Chem Technol Biotechnol 02682575/2004/$30.00

189

W Cichy, S Schlosser, J Szymanowski

different alkylamines (eg Amberlite LA-2, Alamine 336),3 tributylphosphate, trialkylphosphine oxides4 7 and trialkylphosphine sulde8,9 is now preferred. Classical extraction is sometimes connected with the formation of a third phase and/or emulsication. As a result, phase disengagement problems may occur and can cause signicant contamination of aqueous streams with chemicals and even stop the continuous process. An emulsion is often formed in the stripping of phenol from loaded organic phases with alkaline aqueous solutions as extractants, including commercial trialkylphosphine oxides. Such reagents contain some acidic substances which after neutralization form strongly surfaceactive salts. These problems can be eliminated using pertraction through liquid membranes or membrane-based solvent extraction in hollow ber contactors. Although these processes also have some limitations, including lower mass-transfer coefcients and problems with stability of hollow ber modules in contact with solvent10 they seem to be natural successors to traditional extraction. The transport of phenol through liquid membranes can be studied in contactors with a bulk liquid membrane (BLM). The advantage of these contactors is their relative simplicity, the possibility of experimental determination of concentrations in all three phases, and the possibility of visual observation of interfaces and phases. For a BLM with a lower density than that of the aqueous phase, the use of H-type contactors11,12 is very convenient because all three phases can be stirred in such a way that identical hydrodynamic conditions at both interfaces can be reached. The aim of this work was to compare different extractants/carriers, such as Cyanex 923, trioctylamine (TOA) and Amberlite LA-2, in the extraction and the transport of phenol through a BLM. The additional aim was to study the effect of the purication of Cyanex 923 from acidic substances upon the extraction and pertraction performance and the mass transfer. In previous work13 it was found that the reagent exhibited atypical adsorption behavior at the membrane/alkaline strip phase, and the stripping step was signicantly slower than extraction.

The dissociation will be further important in the BLM experiments at the stripping interface, as the complex is decomposed and phenol is effectively stripped. When the physical extraction and association of phenol to a dimer is considered, then the following mass balance equations can be formulated: [PhOH]T = [PhOH S] + Da [PhOH]w + nKd (Da )n [PhOH]n w [S] = [S] + [PhOH S]
T

(2) (3)

The results of the BLM experiments are well described by the model used by Boyadzhiev and Atanassova14 for L-lysine recovery and by Schlosser and Sabolov 11 for the pertraction of butyric acid. a The following assumptions were taken into account in the development of a mass-transfer model in the two-compartment cell with a bulk liquid membrane: (i) fast chemical reactions of the formation and decomposition of the complex phenolcarrier at the interfaces, which means that diffusional masstransfer in boundary layers is important; (ii) all three phases are well mixed and hydrodynamic conditions at both interfaces in the membrane are identical, so kMF = kMR = kM ; (iii) equilibrium at both interfaces; (iv) volumes of the feed, membrane, and the stripping phases do not change; (v) concentration of undissociated phenol in the stripping solution at the interface R/M (stripping solution/membrane phase) is practically zero, which means that the concentration of the complex in the membrane at the strip interface is close to zero, as well. The resulting model equations are as follows: kF kM AF (DF cF cM ) dcF = dt VF (kF + kM DF ) dcM kM AR = dt VM kF AF (DF cF cM ) cM AR (kF + kM DF ) (4) (5) (6)

kM AR dcR = cM dt VR

THEORY At pH values above 4 for TOA and Cyanex 923 and above 5.5 for Amberlite LA-2, the studied extractants (carriers) form 1:1 complexes with phenol (PhOH):7 PhOHw + S (PhOH)S (1)

where for the concentration of the permeate in the membrane phase the following equation can be written from the material balance: cM = VF (cFo cF ) VR cR VM (7)

the bar denoting the organic phase, hereafter called the membrane phase. Simultaneously the physical extraction of phenol into the membrane phase and then the association of phenol, at least to the dimer, have to be considered. At pH > 8, dissociation of phenol also occurs.
190

The initial ux of phenol through the extraction interface and the maximum ux of phenol through the stripping interface were estimated from the following equations: JFo = JMmax VF dcF AF dt VR dcR = AR dt (8) (9)

J Chem Technol Biotechnol 80:189197 (2005)

Extraction and pertraction of phenol

The slope of the linear part of the time dependence of cR , calculated from experimental data, is taken as dcR /dt.

EXPERIMENTAL METHODS The following carriers were used as delivered without any purication: Cyanex 923 (Cytec, Canada), trioctylamine (TOA) (Fluka, Germany) and Amberlite LA-2 (Fluka, Germany). Cyanex 923 contained trialkyl (C6 , C8 ) phosphine oxides as the active components and the high purity of this industrial reagent was recently demonstrated.15 The reagent contains, however, small amounts of dialkylphosphinic acids that were observed on a gas chromatogram, but only after sample derivatization. This impurity causes an atypical change of the interfacial tension and formation of emulsions when the aqueous phase contains sodium hydroxide. Dialkylphosphinic acids were removed by gentle shaking of Cyanex 923 solution several times with fresh portions of 2 kmol m3 aqueous NaOH solution. The purication was stopped when the formation of a haze during shaking was not observed. TOA had a purity better than 99%. Amberlite LA2 contained N-dodecyl-N-(1,1,3,3,5,5-hexamethylhexylamine) as the active substance. n-Alkanes (dodecane fraction, Slovnaft, Slovakia), puried by zeolite adsorption, were used as diluents. Their composition was as follows: aliphatic hydrocarbons99.8% (n-C10 : 7.18%, n-C11 : 32.39%, n-C12 : 33.11%, n-C13 : 26.84% and n-C14 and higher homologes: 0.28%) and aromatics: 0.03%. The composition of the membrane phases and their characteristics are shown in Table 1. Viscosities of organic phases were determined by an Ubbelohde capillary viscometer. Extraction experiments were carried out at room temperature. Phases of various ratios were mechanically shaken for 3 h and the content of phenol was determined in the aqueous phases. The concentration of phenol in the aqueous feed was changed from 1 to
Table 1. Physical and transport properties of membrane phases (solvents): distribution coefcient of phenol D (at equilibrium phenol concentration in the feed 0.8 g dm3 , and pHF = 4), kinematic viscosity of the membrane M , and diffusion coefcient of the carrierphenol complex in the membrane phase D at 25 C calculated by WilkeChang equation16

10 g dm3 , and the pH of the feed was above 2 and varied up to 11. In membrane experiments the aqueous feeds contained phenol (pure, POCh, Poland) (0.52 g dm3 ), Na2 SO4 (0.1 kmol m3 ) with pH values in the range 1.26.0 (adjusted with H2 SO4 ). The aqueous receiving phase contained NaOH (0.052 kmol m3 ). The concentration of the carrier in the membrane phase was 0.4 kmol m3 . BLM experiments were carried out at 25 C in a twocompartment contactor designed by Schlosser11,12 and also described in previous work.13 The membrane phase was mixed with Teon disc mixers (29 mm in diameter, distance between the center of disc and interface 12 mm) rotating in the same direction with an adjustable frequency set at 120 min1 in the current experiments. Interfaces were stable up to a frequency of the membrane phase mixers of about 135 min1 . Aqueous feed and stripping phases were mixed by magnetic Teon rod mixers (diameter 5 mm and length 25 mm) at a frequency of 95 min1 . As found for the transport of citric acid and zinc,17 the permeate ux reached a steady state value between frequencies from about 90 to 130 min1 . Above this value interfaces became unstable and the transport rate increased due to the increase in the interfacial area. Volumes of phases were equal to: feed VF = 310 cm3 , membrane VM = 90 cm3 , stripping solution VR = 130 cm3 . Interfacial areas were equal to: AF = 21.53 cm2 , AR = 20.19 cm2 . The content of phenol in the aqueous phases was determined spectrophotometrically using a (spectrophotometer UNICAM 8625, UK). Samples were diluted with NaOH (0.1 mol dm3 ) to adjust the pH to well above the pKa , and the content of phenolate was determined at 270 nm. The concentration of phenol in the organic (membrane) phase was determined after stripping of phenol into alkali solution where it was analyzed spectrophotometrically.

Membrane phase n-Alkanes 0.4 mol dm3 TOA in n-alkanes 0.4 mol dm3 LA-2 in n-alkanes 0.4 mol dm3 Cyanex 923 in n-alkanes

M 106 (m2 s1 ) 1.60 2.03

D D

1010 (m2 s1 ) 14.2 3.8

DD 1010 (m2 s1 ) 1.7 5.1

0.124 1.35

2.22

2.20

3.8

8.4

2.30

42

4.7

197.4

DISCUSSION AND RESULTS Liquid/liquid equilibria and extraction of phenol Initial screening experiments showed that each of the extractants considered could be used to recover phenol from aqueous acidic and neutral solutions. Phenol could be efciently stripped from the loaded organic phase with NaOH (0.2 mol dm3 ). However, stable emulsions were formed in the stripping step in the system containing Cyanex 923. The formation of emulsions was also observed in extraction when the concentration of phenol in the aqueous feed was above 2 g dm3 . Centrifuging was necessary to achieve separation of phases. This negative effect was not observed in systems containing amines and when the puried Cyanex 923 was used. Thus, the commercial Cyanex 923 had to be pre-washed with an alkaline solution before use. However, it is necessary to point out that the purication did not change the extraction abilities of the reagent.
191

J Chem Technol Biotechnol 80:189197 (2005)

W Cichy, S Schlosser, J Szymanowski

Cyanex 923 is a neutral reagent, while amines could be protonated to give mixtures of sulfates and hydrosulfates. The pKa values for TOA and Amberlite LA-2 are 3.5 and 5.3, respectively. Thus, in the region of aqueous phase acidity in which the protonation of amines could be neglected (>4 for TOA and >6 for Amberlite LA-2) the reagents could be considered as solvating, S, and the extraction of phenol could be described by eqn (1). The extraction isotherms (Figs 1 and 2) for TOA and Amberlite LA-2 indicate that the distribution coefcient of phenol did not change with the equilibrium concentration of phenol in the aqueous phase (over the considered range of concentrations). The logarithm of distribution coefcient was a linear function of the amine concentration in the organic phase. log D = log Kex + log S (10) Linear relationships for log D vs log S were observed but the slopes were equal to 0.77 for TOA and 0.84 for Amberlite LA-2. These equations are statistically valid with correlation coefcients of 0.997 and 1.000. This means that some other phenomena had also to be considered.
5 4
n-alkanes 0.05 mol dm-3 TOA 0.2 mol dm-3 TOA 0.4 mol dm-3 TOA 0.6 mol dm-3 TOA 0.8 mol dm-3 TOA

3 2 1 0

Phenol can also be transferred to the organic phase by physical extraction (Da = 0.124). Moreover, phenol can associate in the hydrocarbon phase, giving various linear self-associates.18,19 This association of phenol is often limited to the formation of dimers with the dimerization constant Kd within the range 2060 dm3 mol1 .20 Thus, the extraction could be described by eqns (2) and (3) in which n = 2 when the formation of dimers is considered. Actually, when n equaled 2 the dimerization constant was evaluated as 44.6 dm3 mol1 for Amberlite LA-2 and 10.8 dm3 mol1 for TOA. However, when Cyanex 923 was used physical extraction could be neglected and phenol dimerization was not considered. Cyanex 923 extracted signicantly stronger phenol than TOA and Amberlite LA-2 (Fig 3). As a result, the extraction isotherms were linear only for low concentrations of phenol in the aqueous phase (<0.5 g dm3 ) and achieved the asymptote for the concentrations of phenol above 1.5 g dm3 . Cyanex 923 (0.4 mol dm3 ) could be loaded with phenol up to 40 g dm3 . Similar values of the distribution coefcients were obtained in the pH region 25. High values of D (40100) were obtained for the equilibrium concentration of phenol in the aqueous phase below 0.3 g dm3 decreasing to about 20 for [PhOH]w = 1.5 g dm3 . This meant that the distribution coefcients for Cyanex 923 were one or two orders higher than those determined for TOA and Amberlite LA-2. Physical extraction of phenol could be neglected for the system with Cyanex 923 because it is relatively low compared with reactive extraction. The extraction of phenol with the considered extractants depended on the acidity of the aqueous phase (Fig 4). The sharp decrease of extraction at pH > 8 was caused by the dissociation of phenol as only
100 80 60 40 Cyanex 923

co*, g dm-3

cw*, g dm-3
Figure 1. The isotherms of phenol extraction with trioctylamine in n-alkanes for different extractant concentrations (the concentration 0.00 mol dm3 means that pure hydrocarbons were used for extraction, pH 6). =

10 9 8 7 6 5 4 3 2 1 0

20 2.5 Amberlite LA-2

co*, g dm-3

0.00 mol dm-3 LA-2 0.05 mol dm-3 LA-2 0.2 mol dm-3 LA-2 0.4 mol dm-3 LA-2 0.6 mol dm-3 LA-2

2.0 1.5 1.0 0.5 0.0 0.0 0.4 0.8 TOA n-alkanes 1.2 1.6

cw*, g dm-3
0 1 2 3 4

cw*, g dm-3
Figure 2. The isotherms of phenol extraction with Amberlite LA-2 in n-alkanes for different extractant concentrations (the concentration 0.00 mol dm3 means that pure hydrocarbons were used for extraction, pH 6). =

, pH* 2, Cyanex 923 , pH* 4, Cyanex 923

, pH* 3, Cyanex 923 , pH* 5, Cyanex 923

Figure 3. Distribution coefcient of phenol vs equilibrium aqueous concentration of phenol. Extractant concentration: 0.4 kmol m3 , pH 6 for systems with TOA, Amberlite LA-2 and pure n-alkanes; pH = 2, 3, 4 and 5 for Cyanex 923.

192

J Chem Technol Biotechnol 80:189197 (2005)

Extraction and pertraction of phenol


100 Cyanex 923

10

Amberlite LA-2

TOA

0.1

10

12

pH*F
Figure 4. Effect of aqueous phase equilibrium pH on the distribution ratio of phenol (concentrations of extractants in n-alkanes, 0.4 mol dm3 ; equilibrium concentration of phenol in the aqueous phaseimportant only for Cyanex 923, approx 0.8 g dm3 ).

neutral species are extracted. As a result, stripping could be easily achieved with an alkaline aqueous solution, eg 0.2 mol dm3 NaOH. An atypical increase of the distribution ratio of phenol was observed at pH values below 6 and 4 for the systems containing Amberlite LA-2 and TOA, respectively. In these acidic media the amines were protonated and the extraction of phenol could be described by the following set of equations:21 2(TOA) + 2H+ F + SO2 4F = (TOA)2 H2 SO4 (11)

with the feed (Fig 5), the enrichment factor depending on the volumes of the feed and stripping phase and for the system under consideration could be equal to 2.4. Decreasing the stripping phase volume will increase the enrichment factor. Time proles of the phenol concentration in the three phases of the two-chamber contactor are presented in Figs 68. From these gures, where the solid lines represent correlation of experimental data according to the model, follows good agreement of the tested model with the experimental data. Only in one case does the model not t the experimental data so very well, ie when Cyanex 923 was used as carrier and the stripping solution contained only 0.05 mol dm3 NaOH. This can be explained by a low capacity of the stripping solution when, considering assumption (v), the developed model is not appropriate. Therefore, the concentration of NaOH in the stripping phase was increased to 0.2 mol dm3 for Cyanex 923. The purication of the carrier signicantly affected the concentration proles in the aqueous feed and in the membrane, but affected the aqueous

c DF = cMF
FM

cRM cMF cR

cF

cMR

c DR = cMR
RM

PhOHF + 0.61(TOA)2 H2 SO4 = [0.61(TOA)2 H2 SO4 ] PhOH (12)


F

cFM M R

Actually, a mixture of amine sulfate and amine hydrosulfate was formed, eg: (R3 NH)2 SO4 + 2H+ + SO2 = 2R3 NH HSO4 F 4F (13) with the equilibrium of reaction (1) shifted to the right with increasing concentration of sulfuric acid.22, 23 The extraction constants for reaction (1) amounted to 1230, 5.74 and 3.54 dm3 mole1 for Cyanex 923, Amberlite LA-2 and TOA, respectively. Pertraction of phenol through a bulk liquid membrane The driving force for the transport of phenol in a BLM is a concentration difference of undissociated phenol molecules between the feed and the stripping phases. When sufciently concentrated NaOH solution (pH >> pKa ) was used as the stripping phase then the concentration of undissociated phenol in the strip phase is zero and phenol could be totally removed from the feed. The phenol concentration in the stripping solution could be signicantly higher in comparison
J Chem Technol Biotechnol 80:189197 (2005)

Figure 5. Concentration prole diagram for phenol extraction through BLM.

1.2 R

0.8

c, g dm-3
0.4 F 0.0 0 , , 8 t, h ; TOA , , ; Amberlite LA-2 , , , Cyanex 923 4 12
Figure 6. Time course of phenol concentration in three phases of pertractor for phenol pertraction with various carriers. F: cFo = 1 g dm3 , pHFo = 4; M: 0.4 kmol m3 TOA and Amberlite LA-2, and 0.2 kmol m3 Cyanex 923; R: 0.05 kmol m3 NaOH.

193

W Cichy, S Schlosser, J Szymanowski

1.2

c, g dm-3

0.8

0.4

0.0 0 4 8

F 12

t, h
, , ; reagent as delivered , , ; purified reagent

Figure 7. Time course of phenol concentration in three phases of pertractor for phenol pertraction with trialkyl phosphine oxides. F: cFo = 1 g dm3 , pHFo = 4; M: 0.2 kmol m3 Cyanex 923; R: 0.2 kmol m3 NaOH.

1.2

F R M model

0.8

0.4

0.0

12

t, h
Figure 8. Time course of phenol concentration in three phases of pertractor for phenol pertraction with TOA. F: cFo = 1 g dm3 , pHFo = 2.5; M: 0.4 kmol m3 TOA; R: 0.05 kmol m3 NaOH.

receiving phase only slightly (Fig 7). Thus, the purication caused only a signicant increase in the extraction rate. As a result, a sharp decrease of phenol content in the aqueous feed and a high phenol accumulation in the membrane phase was observed. Purication caused a large increase of the initial ux of phenol through the extraction interface, ie from 5.7 mol m2 s1 for unpuried carrier to 13.0 mol m2 s1 for puried Cyanex 923. Simultaneously, the maximum concentration in the membrane phase increased from 0.80 to 1.17 g dm3 . However, the uxes of phenol through the stripping interface had similar values for the commercial and puried carriers. Actually, the purication even caused a small decrease of JMmax . The following values of maximum uxes were determined: 2.4 and 2.1 mol m2 s1 for commercial and puried carriers, respectively. This difference could be in the range of
194

experimental error but it could be also the result of different maximum concentrations and times needed to achieve these concentrations in the membrane. The overall rate (ux) of a process consisting of consecutive steps is always limited by the slowest step. Thus, the purication of the carrier did not improve the limiting step and, as a result, the overall process rate. This meant that the slow stripping observed in several different systems with a stable interface, eg in the pertraction of carboxylic acids,11,24 could be caused by another phenomenon. A possible explanation could be the adsorption of the carrier at the aqueous phase/membrane interfaces. Such adsorption is a well established phenomenon.20,25 28 The adsorption of the carrier at the aqueous feed/membrane interface facilitates extraction due to a dense population of carrier molecules in the adsorption layer and their orientation with the hydrophilic reactive groups penetrating the aqueous layers near the interface. However, adsorption at the membrane/receiving aqueous phase retards the stripping as the adsorbed carrier molecules block the interface and decrease the access of the complex to the interface. Another explanation could be saturation of the adsorbed layer with molecules of the complex at its relatively low concentration in the bulk membrane phase. Further increase in the concentration of the complex in the membrane phase, eg by increased carrier concentration or by its purication, does not change its interfacial concentration and consequently does not change the stripping rate. All this means that the high potential of Cyanex 923 to transfer phenol rapidly could not be exploited and could not be improved by purication of reagent due to its adsorption ability. A sharp increase of the distribution coefcient (Fig 4) and phenol transfer with TOA (Fig 8), both to the membrane and receiving phase, were observed when pHFo of the aqueous feed decreased from 4.0 to 2.5. These effects were caused by a change of the complexation reaction due to protonation of TOA (eqns (11)(13)). The effect of pH was also observed for Amberlite LA-2 at pH values below 5. Amberlite LA-2 is a more basic reagent (pKa = 5.3) than trioctylamine (pKa = 3.5) and protonation occurred at a higher pH. Cyanex 923 was not protonated under such mild conditions and the effect of feed acidity was not observed. A further decrease of pH below 2 was impossible for the system containing TOA because a new organic phase was formed at the feed/membrane interface, which collected phenol and retarded the further transport. It was impossible to destroy this phase by an increase of the frequency of the mixer because oscillation of the interface was observed at 130 min1 and further increase of the rate of mixing caused the formation of a dispersion. The concentration of sodium hydroxide in the stripping solution did not affect the uxes of
J Chem Technol Biotechnol 80:189197 (2005)

c, g dm-3

Extraction and pertraction of phenol


6

0 0.0

0.5

1.0

1.5

2.0

2.5

[NaOH], mol dm-3 , JFo, Cyanex 923 , JFo, TOA , JRmax, Cyanex 923 , JRmax, TOA

Figure 9. Initial ux of phenol through the extraction (JFo ) and maximum ux through the stripping interface (JRmax ) vs concentration of NaOH in the stripping solution. F: cFo = 1 g dm3 , pHFo = 4, M: 0.4 kmol m3 of the carrier.

phenol with TOA and Cyanex 923 for hydroxide concentrations above about 0.2 mol dm3 (Fig 9). However, a decrease of the ux through the stripping interface was observed when the concentration was 0.05 mol dm3 , which was connected with the lower stripping rate. The uxes changed approximately linearly with increasing concentration of phenol (Fig 10). The highest uxes were obtained for Cyanex 923 and the lowest for the pure n-alkanes. The uxes through
15

the extraction interface, JFo , and stripping interface, JRmax , changed in the following order: Cyanex 923 >> Amberlite LA-2 > TOA >> n-alkanes. This is the same order of membrane phases as that obtained for the values of the distribution coefcients or extraction constants of these solvents. This shows the important role of the distribution ratio on the rate of membrane transport. For all membranes with a carrier, JFo > JRmax and the difference is especially large at higher phenol concentrations. A more complex parameter, characterizing passage through the membrane by solution-diffusion transport, is permeability, which in the rst approach (no inuence of reaction kinetics) can be dened as the product of the distribution coefcient and the diffusion coefcient (DD) and for studied membranes is shown in Table 1. While the addition of TOA and Cyanex 923 to n-alkanes increased their nominal permeability by factors of 3.0 and 116.1 (Table 1) (4.9 for Amberlite LA-2), the maximal ux through the stripping interface increased by factors of 2.8 and 6.16, (Fig 10). It is a much smaller increase in the phenol ux for the solvent with Cyanex 923. Thus, these results support the previous conclusions showing the importance of adsorption phenomena, especially at the stripping interface. Values of the individual mass-transfer coefcients (kF and kM ) obtained by tting of eqns (4)(7) to experimental data are given in Fig 11 as the function of phenol concentration in the feed. The value of the individual mass-transfer coefcients in the feed boundary layer (for an phenol initial concentration in the feed of 1 g dm3 ) increased in the order: n-alkanes < TOA < Amberlite LA-2 < Cyanex 923; that
8

J.105, mol.m-2 .s-1

J.105, mol.m-2 .s-1

10

k. 106, m.s-1
0.5 1.0 1.5 , JRmax, 2.0 Cyanex 923 2.5

0 0.0

cFo, g.dm-3
, JFo, Cyanex 923 , JFo, Amberlite LA-2 , JFo, TOA , JFo, Pure N-alkanes , JRmax, Amberlite LA-2 , JRmax, TOA , JRmax, Pure N-alkanes

0 0.0

0.5

1.0

1.5

2.0

2.5

cFo, g dm-3
, kF, Cyanex 923 , kF, , kF, TOA Amberlite LA-2 , kM, Cyanex 923 , kM, Amberlite LA-2 , kM, TOA

Figure 10. Initial ux of phenol through the extraction interface (JFo ) and maximum ux through the stripping interface (JRmax ), as a function of the initial phenol concentration in the feed. F: pHFo = 4; R: 0.050.2 kmol m3 NaOH (0.2 kmol m3 for Cyanex 923); M: 0.4 kmol m3 of the carrier.

Figure 11. Individual mass-transfer coefcients of phenol vs initial phenol concentration in the feed. F: pHFo = 4; R: 0.050.2 kmol m3 NaOH (0.2 kmol m3 for Cyanex 923); M: 0.4 kmol m3 of the carrier. Empty and full symbols denote individual mass-transfer coefcient kF and kM , respectively.

J Chem Technol Biotechnol 80:189197 (2005)

195

W Cichy, S Schlosser, J Szymanowski

is the same order as for the distribution coefcient D. This effect is justied by the presence of the distribution coefcient in the considered model, ie in eqns (4) and (5). The value of the individual masstransfer coefcient in the membrane boundary layers changed in the opposite direction. The formation of more stable carrier/phenol complexes in the solvent with higher distribution coefcients results in a slower decomposition rate at the stripping interface that hinders the overall transport rate. The hydrodynamic conditions at both interfaces are similar. The viscosity of the membrane phase and diffusivity of the phenol complex depend on the carrier but the differences are relatively small (Table 1). The individual mass-transfer coefcients do not depend on phenol concentration in the feed for the pure n-alkanes and the membrane with Amberlite LA-2 (Fig 11), and approximately similar values of both coefcients are observed. Quite different behavior was observed in the membranes with Cyanex 923 and TOA. The value of the individual masstransfer coefcient in the feed, kF , increased and the individual mass-transfer coefcient in the membrane, kM , decreased with increasing feed concentration (Fig 11). The evaluated mass-transfer coefcients were of the same order (106 m s1 ) as those determined for the transport in hollow bers.29 31 However, when physical transport was considered (hexanewaterphenol) the mass-transfer coefcient was of the order 107 m s1 .29 More justied, however, was the comparison of uxes at the constant driving force equal to cFo for JFo and to cMmax for JRmax (Fig 12). In such a case, JFo = JRmax for the same driving force when amine carriers were used. Only in the case of Cyanex 923, JFo >> JRmax . This gives further evidence that the stripping process was slower when Cyanex 923 was used, but still the ux was signicantly higher than in the BLM experiments carried out with amine carriers. When amines were used as a carrier, the concentration of the phenol complex in the membrane phase was low and equal to 0.2 g dm3 for TOA and 0.4 g dm3 for Amberlite LA-2. This concentration increased to 1.5 g dm3 when Cyanex 923 was used and the NaOH concentration was 0.05 mol dm3 . Thus, the concentration of sodium hydroxide was too low; NaOH was quickly consumed at the interface and the transfer of phenol through the stripping interface was limited by the diffusion of NaOH from the bulk to the interface. As a result, the phenol complex accumulated and the model used did not t the experimental data very well. Thus, an increase of NaOH to 0.2 mol dm3 was necessary.

16

12

J.105, mol.m-2.s-1

0 0.0

0.8
, JF, Cyanax 923 , JF, AmberliteLA-2

1.6 c, g dm-3 , JRmax , JRmax , JRmax

2.4

, JF, TOA

Figure 12. Flux of phenol through the extraction (JF ) and maximum ux through the stripping interface (JRmax ) vs driving force (cFo for JFo and to cMmax for JRmax ). F: cFo = 1 g dm3 , pHFo = 4; M: 0.4 kmol m3 of the carrier.

CONCLUSIONS Cyanex 923 was the best carrier for phenol and the formation of emulsions observed in classical liquidliquid extraction was eliminated by pertraction
196

through the BLM and in addition, the formation of a new phase was not observed. The distribution coefcients of phenol for solvents containing carriers and pure n-alkanes, the individual mass-transfer coefcients in the feed and the maximum uxes of phenol through the extraction (JFo ) and stripping (JRmax ) interfaces decreased in the order: Cyanex 923 > Amberlite LA-2 > TOA >> pure n-alkanes. The opposite order was observed for the masstransfer coefcients in the membrane phase. This is connected with the fact that the value of kM resulting from tting the model to the experimental data reects also the resistance connected with the reaction kinetics of the decomposition of the complex (PhOH)S. The acidity of the aqueous feed inuenced the transport of phenol with TOA and Amberlite LA-2 at pH values below 4 and 6, respectively. This effect was caused by the protonation of amines and formation of a different complex with phenol and co-transport of mineral acid, as discussed elsewhere.24,31 At constant driving forces the maximum uxes through the extraction and stripping interfaces were closer when amine carriers were used. However, JRmax was much lower than JFo for Cyanex 923. Despite this, the ux of phenol to the stripping solution was signicantly higher than through BLM with amine carriers. The rate of the complex decomposition was the ratedetermining step for the overall transport rate. The adsorption of the carrier at aqueous phase/membrane interfaces was probably responsible for the rapid and slow transfer of phenol through the extraction and stripping interfaces, respectively. The adsorption of the carrier at the aqueous feed/membrane interface facilitated the extraction due to a dense population of carrier molecules in the adsorption layer and
J Chem Technol Biotechnol 80:189197 (2005)

Extraction and pertraction of phenol

their orientation with the hydrophilic reactive groups penetrated the aqueous layers near the interface. However, adsorption at the membrane/receiving aqueous phase retarded the stripping as the adsorbed carrier molecules blocked the interface and decreased the access of the complex to the interface.

ACKNOWLEDGEMENTS Two authors (WC and JS) thank the Polish Scientic Committee (grant DS 32/044/2004). One author (SS) thanks the Slovak grant VEGA No 9136/02. WC thanks Dr S Schlosser for his general help and scientic discussion during several working visits at the Department of Chemical and Biochemical Engineering of the Slovak University of Technology in Bratislava.

REFERENCES
1 Mulligan TJ and Fox RD, Treatment of industrial waste waters, in Industrial Wastewater and Solid Waste Engineering, ed by Cavaseno V. McGraw-Hill Publications Co, New York, pp 173190 (1980). 2 King CJ and Senetar JJ, Solvent extraction of industrial organic substances, in Ion Exchange and Solvent Extraction, ed by Marinsky JA and Marcus Y. M Dekker, New York, Vol 10, pp 3561 (1988). 3 Cichy W, Schlosser S and Szymanowski J, Transport of phenol through bulk liquid membranes, in Solvent Extraction for the21st Century, ISEC 99, ed by Cox M, Hidalgo M and Valiente M. Soc Chem Ind, London, pp 10651070 (2001). 4 McGlashan JD, Bixby JL and King CJ, Separation of phenols from dilute aqueous solution by use of tri(n-octyl)phosphine oxide as extractant. Solvent Extr Ion Exch 3:125 (1985). 5 Watson EK, Rickelton WA, Robertson AJ and Brown TJ, A liquid phosphine oxide: solvent extraction of phenol, acetic acid and ethanol. Solvent Extr Ion Exch 6:207220 (1988). 6 Urtiaga AM and Ortiz I, Extraction of phenol using trialkylphosphine oxides (Cyanex 923) in kerosene. Sep Sci Technol 32:11571162 (1997). 7 Drozdova MK, Nikolaeva IV and Torgov VG, Complexation and solvatation on extraction of phenol by trioctylphosphine and trioctylamine oxides. Russian J Phys Chem 71:498501 (1997). 8 Schlosser S, Rothov I and Frianov H, Hollow-ber pertractor a a with bulk liquid membrane. J Membr Sci 80:99106 (1993). 9 Schlosser S and Rothov I, A new type of hollow-ber a pertractor. Sep Sci Technol 29:765780 (1994). 10 Reed BW, Semmens MJ and Cussler EL, Membrane contactors, in Membrane Separation Technology. Principles and Applications, ed by Noble RD and Stern SA. Elsevier Science, Amsterdam, pp 467498 (1995). 11 Schlosser S and Sabolov E, Transport of butyric acid through a layered bulk liquid membranes. Chem Papers 53:403411 (1999).

12 Vajda M, Schlosser S and Kova ova K, Pertraction of silver c through bulk liquid membranes. Chem Papers 54:423429 (2000). 13 Cichy W, Schlosser S and Szymanowski J, Recovery of phenol with CYANEX 923 in membrane extraction-stripping systems. Solvent Extr Ion Exch 19:905923 (2001). 14 Boyadzhiev L and Atanassova I, Recovery of L-lysine from dilute water solutions by liquid pertraction. Biotechnol Bioeng 38:10591064 (1991). 15 Dziwinski E and Szymanowski J, Composition of Cyanex 923, Cyanex 925, Cyanex 921 and TOPO. Solvent Extr Ion Exch 16:14651492 (1998). 16 Reid RC, Prausnitz JM and Sherwood TK, The Properties of Gases and Liquids. McGraw-Hill, New York (1977). 17 Schlosser S and Forgova-Hovan kova E, Pertraction of zinc c through a bulk liquid membrane, in Proc 6th Int Conf Separat Ionic Solutes, Piestany. pp 2628 (1995). 18 Barela R and Buchowski H, Vaporliquid equilibria of dilute solutions of phenols in cyclohexane. Self-association of meta- and para-substituted phenols. Fluid Phase Equilibria 59:99107 (1990). 19 Barela R and Buchowski H, Vaporliquid equilibria of dilute solutions of phenols in cyclohexane. Self-association of ortho-substituted phenols. Fluid Phase Equilibria 39:293306 (1988). 20 Szymanowski J, Physiochemical modications of extractants. Critical Rev Anal Chem 25:143194 (1995). 21 Wang ML and Hu KH, Extraction of phenol using sulfuric acids salts of trioctylamine in a supported liquid membrane. Ind Eng Chem Res 33:914921 (1994). 22 Kyuchoukov G and Mihailov I, A novel method for recovery of copper from hydrochloric acid solutions. Hydrometallurgy 27:361369 (1991). 23 Kyuchoukov G and Mishonov I, A new extractant mixture for recovery of copper from etching solution. Solvent Extr Ion Exch 11:555567 (1993). 24 Schlosser S, Sabolov E, Kert sz R and Kubi ov L, Factors a e s a inuencing transport through liquid membranes and membrane based solvent extraction. J Sep Sci 24:509518 (2001). 25 Szymanowski J, Hydroxyoximes and Copper Hydrometallurgy. CRC Press, Boca Raton, USA (1993). 26 Szymanowski J and Tondre C, Kinetics and interfacial phenomena in classical and micellar extraction systems. Solvent Extr Ion Exch 12:873905 (1994). 27 Szymanowski J, Kinetics and interfacial phenomena. Solvent Extr Ion Exch 18:729751 (2000). 28 Prochaska K, Interfacial activity of metal ion extractant. Adv Colloid Int Sci 95:5172 (2002). 29 Yun CH, Prasad R and Sirkar KK, Membrane solvent extraction removal of organic pollutants from aqueous waste streams. Ind Eng Chem Res 31:17091719 (1992). 30 Sengupta A, Basu R and Sirkar KK, Separation of solutes from aqueous solutions by contained liquid membranes. AICHE J 34:16981708 (1988). 31 Kubisova L, Martak J and Schlosser S, Transport of 5-methyl2-pyrazinecarboxylic acid through a layered bulk liquid membrane. Chemical Papers-Chemicke Zvesti 56:418425 (2002).

J Chem Technol Biotechnol 80:189197 (2005)

197

You might also like