12.07.2015 Views

11. Interfacial Mechanism and Kinetics of Phase-Transfer Catalysis

11. Interfacial Mechanism and Kinetics of Phase-Transfer Catalysis

11. Interfacial Mechanism and Kinetics of Phase-Transfer Catalysis

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

11<strong>Interfacial</strong> <strong>Mechanism</strong> <strong>and</strong> <strong>Kinetics</strong> <strong>of</strong><strong>Phase</strong>-<strong>Transfer</strong> <strong>Catalysis</strong>HUNG-MING YANG National Chung Hsing University, Taichung, Taiwan,Republic <strong>of</strong> ChinaHO-SHING WU Yuan-Ze University, Taoyuan, Taiwan, Republic <strong>of</strong> ChinaI. INTRODUCTIONA. General ConsiderationsAs the chemical reactants reside in immiscible phases, phase transfer (PT) catalysts havethe ability to carry one <strong>of</strong> the reactants as a highly active species for penetrating theinterface, into the other phase where the reaction takes place, <strong>and</strong> to give a high conversion<strong>and</strong> selectivity for the desired product under mild reaction conditions. This type <strong>of</strong>reaction was termed ‘‘phase-transfer catalysis’’ (PTC) by Starks in 1971 [1]. Since then,numerous efforts have been devoted to the investigation <strong>of</strong> the applications, reactionmechanisms, <strong>and</strong> kinetics <strong>of</strong> PTC. Nowadays, PTC becomes an important choice inorganic synthesis <strong>and</strong> is widely applied in the manufacturing processes <strong>of</strong> specialty chemicals,such as pharmaceuticals, dyes, perfumes, additives for lubricants, pesticides, <strong>and</strong>monomers for polymer synthesis. The global usage <strong>of</strong> PT catalysts was estimated at overone million pounds in 1996, <strong>and</strong> PTC in industrial utilization is continuously growing atan annual rate <strong>of</strong> 10–20% [2]. PTC is a very effective tool in many types <strong>of</strong> reactions, e.g.,alkylation, oxidation, reduction, addition, hydrolysis, etherification, esterification, carbene,<strong>and</strong> chiral reactions [2,3].1. Reaction Cycle <strong>of</strong> PTCThe first reaction scheme addressed by Starks in 1971 was for the reaction <strong>of</strong> aqueoussodium cyanide <strong>and</strong> organic 1-chloro-octane. In contrast with the result <strong>of</strong> no apparentreaction occurring after more than 24 h in the absence <strong>of</strong> catalyst, the cyanide displacementreaction takes place rapidly with only 1% <strong>of</strong> the quaternary ammonium salt(C 6 H 13 Þ 4 N þ Cl added, <strong>and</strong> achieving near 100% yield <strong>of</strong> 1-cyano-octane product in2–3 h [1]. The reaction scheme for the PT catalyzed cyanide displacement reaction inaqueous–organic phases is shown in the following:NaCN ðaqÞþ1-C 8 H 17 Cl ðorgÞ QCl! 1-C 8H 17 CN ðorgÞþNaCl ðaqÞ ð1ÞCopyright © 2003 by Taylor & Francis Group, LLC


ð2ÞThe PT catalyst QCl should first react with the cyanide anion to form the activeintermediate QCN, which is then transferred into the organic phase to react with theorganic reactant 1-C 8 H 17 Cl <strong>and</strong> is then regenerated back to QCl to conduct the nextcycle <strong>of</strong> reactions.2. Classification <strong>of</strong> PTC ReactionsPTC reactions can be classified into two types: soluble PTC <strong>and</strong> insoluble PTC. Each typecan be further divided into several categories. Figure 1 shows the classification <strong>of</strong> PTCreactions. Insoluble PTC consists <strong>of</strong> liquid–solid–liquid PTC (LSLPTC) <strong>and</strong> tri-liquidPTC (TLPTC), by which the catalyst can be recovered <strong>and</strong> reused, showing the greatpotential in large-scale production for industry. The catalyst used in LSLPTC is immobilizedon an organic or inorganic support, while in TLPTC it is concentrated within aviscous layer located between the organic <strong>and</strong> aqueous phases. Soluble PTC includesliquid–liquid PTC (LLPTC), solid–liquid PTC (SLPTC) <strong>and</strong> gas–liquid PTC (GLPTC).There are also nontypical PTC reactions termed inverse PTC (IPTC) <strong>and</strong> reverse PTC(RPTC), <strong>and</strong> these are different in catalyst type <strong>and</strong> transfer route, compared to normalPTC [2,3].PT catalysts commonly used are quaternary onium salts (ammonium <strong>and</strong> phosphonium),crown ethers, crypt<strong>and</strong>s, <strong>and</strong> polyethylene glycols. The essential characteristics <strong>of</strong> aPT catalyst are that the catalyst must have the ability to transfer the reactive anion into theorganic phase to conduct the nucleophilic attack on the organic substrate, <strong>and</strong> effect acation–anion bonding loose enough to allow a high reaction rate in the organic phase.FIG. 1Classification <strong>of</strong> PTC reactions.Copyright © 2003 by Taylor & Francis Group, LLC


Other factors in selecting asuitable catalyst that we should consider include the cost <strong>and</strong>structure <strong>of</strong> the catalyst, the toxicity <strong>of</strong> the catalyst <strong>and</strong> solvent, the ease <strong>of</strong> separation <strong>of</strong>the catalyst from the products, the energy requirement for reaction, the stability <strong>of</strong> thecatalyst in process conditions, <strong>and</strong> the ease <strong>of</strong> treatment <strong>of</strong> the waste streams, in order tolead to an efficient <strong>and</strong> economic PTC process.3. <strong>Interfacial</strong> CharacteristicsSince PTC reactions are carried out between immiscible phases, the nature <strong>of</strong> the interface<strong>and</strong> the physical properties <strong>of</strong> the reacting compounds at the interface become veryimportant in promoting the desired reaction rate at asatisfactory level. In aliquid–liquidsystem under agitation, one phase should be dispersed as small droplets in the secondphase in amanner such that alarge interfacial area between the two phases can beobtained. The nature <strong>of</strong> the interface includes interfacial tension, the presence <strong>of</strong> surfactants,<strong>and</strong> the degree <strong>of</strong> agitation rate [3]. These three factors determine the sharpness <strong>of</strong>the interface (or the thickness <strong>of</strong> interfacial film), the droplet size, <strong>and</strong> the interfacialarea available to transfer the reacting anion. The interfacial behaviors <strong>of</strong> the reactinganion include the surface equilibrium distribution <strong>of</strong> the active intermediate, the ease <strong>of</strong>penetration <strong>of</strong> the compounds into the other phase (the depth from the interface), <strong>and</strong>the mass transfer rate across the interface. Adding extra salts may induce achange in theproperties <strong>of</strong> the interface. For example, by adding more inorganic salts or bases, thecatalyst is salted out <strong>of</strong> the aqueous phase <strong>and</strong> an organic solvent with low polarity, <strong>and</strong>the interfacial film grows increasingly thick, finally becoming a separate observablephase. This situation alters the original reaction zone <strong>and</strong> the apparent reaction ratebecause the properties <strong>of</strong> the interface have been changed. Hence, the thickness <strong>of</strong> theinterfacial film (sharpness) is not only limited by the nature <strong>of</strong> the interface itself, butalso affected by the introduction <strong>of</strong> other ingredients. Figure 2shows the scheme <strong>of</strong> aconcentration gradient <strong>of</strong> the reacting compound within a dispersed organic dropletunder a slow or fast diffusion rate, which indicates that the organic reaction is conductedat the interface or in the whole droplet.4. Reaction at the Interface <strong>and</strong> in the Bulk SolutionIn PT catalysis, the reaction mechanisms that have been proposed are the Starks’ extractionmechanism <strong>and</strong> Mąkosza’s interfacial mechanism. These two mechanisms describethe zone where the organic reaction occurs or the phase where the rate-determining step islocated. However, in reality, it is realized that many PTC reactions are conducted both atthe interface <strong>and</strong> in the bulk solution, especially for a reaction controlled by the intrinsicorganic reaction [3]. The distinction between these two mechanisms is recognized as thedifference in the depth <strong>of</strong> the reaction zone penetrating the organic phase.Under the conditions <strong>of</strong> no agitation with a flat interface or slow agitation with aslow mass transfer rate, as the solubility <strong>of</strong> the transferred species in the organic phase issufficiently large, the rate <strong>of</strong> diffusion within the organic phase would not influence theobserved reaction rate significantly. Fast diffusion rates may exhibit extraction mechanismbehavior, while with a slow diffusion rate the system is suitably described by an interfacialmechanism. In other words, for the case <strong>of</strong> strong agitation with extreme low solubility <strong>of</strong>transferred species in the organic phase, the reaction should be mainly conducted near theinterface due to the short penetration depth <strong>of</strong> the transferred species, <strong>and</strong> so is describedby the interfacial mechanism. It is noted that, in general, increasing the agitation rateincreases the degree <strong>of</strong> dispersion <strong>of</strong> one phase <strong>and</strong> produces more tiny droplets, which inCopyright © 2003 by Taylor & Francis Group, LLC


FIG. 2 Concentration gradient in organic droplet: (a) slow diffusion rate (or low solubility); (b) fastdiffusion rate (or high solubility).turn generates a much larger interfacial surface area for transport. Hence, the mass transferrate between the phases, the diffusion rate, <strong>and</strong> the solubility in the organic phase (orthe distribution equilibrium) incorporated with the intrinsic organic reaction play importantroles in determining whether the PT reaction is dominated by an extraction mechanismor by an interfacial mechanism.In the following sections <strong>of</strong> this chapter, the interfacial mechanism <strong>and</strong> the kineticsconcerning LLPTC, LSLPTC, SLPTC, <strong>and</strong> TLPTC will be reviewed.B. Some ApplicationsThe vast literature on PT catalysis has demonstrated in past years the very broad <strong>and</strong>effective applications in organic synthesis [2,3]. Hundreds <strong>of</strong> articles are published per yearconcerning PTC. Hence, we do not intend to review the many uses <strong>of</strong> PTC that have beenreported, but just the typical later examples for illustration in this chapter.Copyright © 2003 by Taylor & Francis Group, LLC


1. Applications in BiologyOrsini et al. [4] synthesized biologically polyphenolic glycosides via Wittig reactions followedby glucosylation under PT conditions. These compounds include (E)-3-(-d-glucopyranosyloxy)-40 ,5-dihydroxystilbene (resveratrol 3--d-glucoside, piceid), (Z)-2 0 ,3 0 -dihydroxy-3,4,4 0 5-tetramethoxystilbene (combretastatin A-1), ;-dihydro-2 0 ,3 0 -dihydroxy-3,4,40 ,5-tetramethoxystilbene (combretastatin B-1), etc. Under PTC, the glucosylationis stereoselective <strong>and</strong> gives the best results for yields with benzyltriethylammoniumchloride <strong>and</strong> aqueous sodium hydroxide. The use <strong>of</strong> nonaqueous bases in dry solventsleads to a sluggish reaction at room temperature, probably due to the poor solubility <strong>of</strong>the phenolate ion in the solvents. Carrie` re et al. [5] synthesized O-, S-, Se-, <strong>and</strong> C-glycosidesby PTC. For the synthesis <strong>of</strong> O-glycosides under liquid–liquid conditions, usingdichloromethane as the organic solvent <strong>and</strong> aqueous NaOH as the base, the PT catalysttetrabutylammonium hydrogen sulfate is used to avoid the possibility <strong>of</strong> double halidedisplacement. PTC conditions are successfully applied in the synthesis <strong>of</strong> - <strong>and</strong> -naphthols to glycohydrolase substrate 7-hydroxy-4-methylcoumarin, to chromogenic substrateFat Brown B 1 , <strong>and</strong> to estrone prodrug. In the preparation <strong>of</strong> thio- <strong>and</strong> selenoglycosides,having saturated NAHCO 3 <strong>and</strong> 1 M Na 2 CO 3 as the aqueous base is sufficientwith thiols <strong>and</strong> selenols, <strong>and</strong> together with tetrabutylammonium hydrogen sulfate as thecatalyst, <strong>and</strong> ethyl acetate as the solvent instead <strong>of</strong> dichloromethane, whereby the sideproducts are produced.Albanese et al. [6] reported the synthesis <strong>of</strong> 2-substituted 3,4-dihydro-2H-1,4-benzoxazinesby ring opening <strong>of</strong> glycidols under solid–liquid PTC. They used N-(2-fluorophenyl)toluene-p-sulfonamideas the nitrogen nucleophile by incorporating the aromaticmoiety <strong>of</strong> benzoxazine as the leaving group, <strong>and</strong> performed the ring opening by stirring at90 C a heterogeneous mixture <strong>of</strong> 1,2-epoxy-3-phenoxypropane, sulfonamide, anhydrousK 2 CO 3 , the catalyst BzEt 3 NCl, <strong>and</strong> dioxane to produce a 95% yield <strong>of</strong> N-(2-fluorophenyl)-N-(2-hydroxy-propyl)toluene-p-sulfonamideafter 17 h <strong>of</strong> reaction. This method providesa straightforward <strong>and</strong> new approach to the synthesis <strong>of</strong> chiral 2-substituted 3,4-dihydro-2H-1,4-benzoxazines.Asymmetric PTC is an important method in the synthesis <strong>of</strong> -alkyl <strong>and</strong> -aminoacids. Belokon et al. [7] reported that the compound (4R,5R)-2,2-dimethyl-;; 0 ; 0 -tetraphenyl-1,3-dioxolane-4,5-dimethanol(TADDOL) was used to catalyze the C-alkylation <strong>of</strong>C–H acids with alkyl halides to the asymmetric synthesis <strong>of</strong> -methyl-substituted -aminoacids under PTC conditions. The alkylations <strong>of</strong> the substrate C–H acids with benzylbromide or allyl bromide were conducted in dry toluene at ambient temperature withNaH or solid NaOH as base <strong>and</strong> TADDOL as a chiral promoter. The type <strong>of</strong> base isimportant in the asymmetric C-alkylation <strong>of</strong> C–H acids.Lygo et al. [8] investigated the enantioselective synthesis <strong>of</strong> bis--amino acid estersvia asymmetric PTC. Under liquid–liquid conditions, the target amino acid esters wereobtained with high enantiometric excess ( 95% ee) from the alkylation reaction <strong>of</strong> benzophenone-derivedglycineimine with an appropriate dibromide. They reported that eitherthe mono- or di-alkylated product could be obtained, depending on the reaction conditions;the monoalkylated product was obtained in good yield with excess dibromide,whereas with stoichiometric quantities <strong>of</strong> dibromide this led to the dialkylated product.By controlling the stoichiometry <strong>of</strong> the reaction, the selectivity <strong>of</strong> the desired product canbe accessed at a high level. Lygo et al. [9] also reported the asymmetric synthesis <strong>of</strong> bis-aminoacids via alkylation <strong>of</strong> a benzophenone-derived glycineimine under PTC conditions.The target bisamino acids can be produced with high yields <strong>and</strong> high levels <strong>of</strong> stereoselectivityby applying chiral quaternary ammonium salts. The core structure <strong>of</strong> the chiralCopyright © 2003 by Taylor & Francis Group, LLC


quaternary ammonium salts closely related to the cinchona alkaloid cinchonine can beused in the benzoylation <strong>of</strong> a glycineimine [10].The indan-based -amino acid derivatives can be synthesized by PTC. Kotha <strong>and</strong>Brahmachary [11] indicated that solid–liquid PTC is an attractive method that <strong>of</strong>fered aneffective way <strong>of</strong> preparing optically active products by chiral PTC. They found that ethylisocyanoacetate can be easily bisalkylated in the presence <strong>of</strong> K 2 CO 3 as the base <strong>and</strong>tetrabutylammonium hydrogen sulfate as the catalyst. The advantage <strong>of</strong> isolating waterfrom the reaction medium is to avoid the formation <strong>of</strong> unwanted hydroxy compounds inthe nucleophilic substitution reaction. If liquid–liquid PTC is applied in the system withthe strong base NaOH <strong>and</strong> dichloromethane as the organic solvent, the formation <strong>of</strong>dihydroxy or cyclic ether can be observed.2. Other ApplicationsPTC incorporated with other methods usually greatly enhances the reaction rate. Masstransfer <strong>of</strong> the catalyst or the complex between different phases is an important effect thatinfluences the reaction rate. If the mass transfer resistance cannot be neglected, animprovement in the mass transfer rate will benefit the overall reaction rate. The application<strong>of</strong> ultrasound to these types <strong>of</strong> reactions can be very effective. Entezari <strong>and</strong>Keshavarzi [12] presented the utilization <strong>of</strong> ultrasound to cause efficient mixing <strong>of</strong> theliquid–liquid phases for the saponification <strong>of</strong> castor oil. They used cetyltrimethylammoniumbromide (CTAB), benzyltriethylammonium chloride (BTEAC), <strong>and</strong> tetrabutylammoniumbromide (TBAB) as the catalysts in aqueous alkaline solution. The more suitablePT catalyst CTAB can accumulate more at the liquid–liquid interface <strong>and</strong> produces anemulsion with smaller droplet size; this phenomenon makes the system have a high interfacialsurface area, but the degradation <strong>of</strong> CTAB is more severe than that <strong>of</strong> BTEAC orTBAB because <strong>of</strong> more accumulation at the interface <strong>of</strong> the cavity under ultrasound.Recently, electron-transfer catalysis by viologen compounds has attracted muchattention. The compounds function as mediators <strong>of</strong> electron transfer <strong>and</strong> have beenapplied in the reduction <strong>of</strong> aldehydes, ketones, quinines, azobenzene, acrylonitrile,nitroalkenes, etc., with zinc or sodium dithionite in a monophase or a two-liquid phasesystem [13]. Noguchi et al. [13] found that a redox-active macrocyclic ionene oligomer,cyclobis(paraquat-p-phenylene), acted as an electron phase-transfer catalyst for the reduction<strong>of</strong> quinines, as compared with acyclic benzyl viologen. The enhanced activity <strong>of</strong> thiscompound is due to the inclusion <strong>of</strong> the substrate into the catalyst cavity.One <strong>of</strong> the important applications <strong>of</strong> PTC is in the field <strong>of</strong> pollution control. Anearly utilization was to apply the PTC method to recover phenolic substances from aqueousalkaline waste streams [14]. The methodology is based on the reaction <strong>of</strong> phenolicsubstances in the aqueous solution with materials such as benzoyl chloride, p-toluenesulfonylchloride, etc., dissolved in the organic solvent in the presence <strong>of</strong> PT catalysts:ð3ÞCopyright © 2003 by Taylor & Francis Group, LLC


Tundo et al. [15] reported an efficient catalytic detoxification method for toxicpolychlorinated dibenzo-p-dioxins (PCDDs) <strong>and</strong> polychlorinated dibenz<strong>of</strong>urans(PCDFs) under mild conditions (50 C <strong>and</strong> 1 atm <strong>of</strong> hydrogen) with a supported metalcatalyst modified by the PT agent Aliquat 336. Their results show that the methodologyproved successful for hydrodechlorinating the toxic samples to yield mixtures containingconcentration <strong>of</strong> contaminants lower than the experimentally detectable limit by gas chromatography–high-resolutionmass spectrometry. This method has the potential to bepractically applied in the detoxification <strong>of</strong> PCDDs <strong>and</strong> PCDFs.PTC is also widely used in polymerization reactions. The main function <strong>of</strong> thequaternary ammonium salts is that they can transfer the diphenolate from the aqueousphase into the organic phase to react with the diacid chloride. Hodget et al. [16] presentedthe synthesis <strong>of</strong> polyesters by the reaction <strong>of</strong> dicarboxylic acid salts with bishalides ortosylates or by the self-condensation <strong>of</strong> salts <strong>of</strong> bromocarboxylic acids under liquid–liquidPTC. With benzyltrimethylammonium salts <strong>and</strong> halides in dry acetonitrile as solvent,using sodium or potassium salts, the yields <strong>of</strong> polyesters are, in degrees <strong>of</strong> polymerization(DP), in the range 17–47, <strong>and</strong> the rate <strong>of</strong> dissolution <strong>of</strong> salts is very slow <strong>and</strong> rate limiting;while in a liquid–liquid system, the DP is in the range 22–161. Liquid–liquid PTC is morefavorable in the synthesis <strong>of</strong> polyesters [16]:RCOX þ R 0 OH ! RCOOR 0 þ HXRCOO M þ þ R 0 X ! RCOOR 0 þ M þ Xð4Þð5Þwhere X ¼ -Cl, -Br, -I, -OSO 2 CH 3 , or -OSO 2 C 6 H 4 CH 3 .The applications <strong>of</strong> PTC in polymerization are gradually increasing. Tagle <strong>and</strong> coworkers[17,18] synthesized poly(amide ester)s from diphenols with the amide group in theside chain, using PT catalysts such as benzyltriethylammonium chloride, with good results.The use <strong>of</strong> anhydrous potassium carbonate as the base is to promote the organic reactionunder solid–liquid PTC. Albanese et al. [19] described some recent applications in thisarea, <strong>and</strong> the reactions <strong>of</strong> aza anions with 2-bromocarboxylic esters <strong>and</strong> expoxidesafforded protected -amino acids <strong>and</strong> -amido alcohols. Sirovski [20] described someexamples <strong>of</strong> PTC applications in organochlorine chemistry. Using a polymeric crownether the results <strong>of</strong> m-phenoxytoluene chlorination are also reported. Carboxylic acids<strong>and</strong> picric acid act as inhibitors, while benzyl alcohol behaves as a strong promoter. Inthe absence <strong>of</strong> the promoter, the reaction is conducted either at the interface or in the thirdphase that is a border liquid film between the organic <strong>and</strong> aqueous phases.The importance <strong>of</strong> triphase catalysis in industry grows continuously. The supportsfor immobilizing the triphase catalyst are mostly <strong>of</strong> organic type, i.e., copolymers <strong>of</strong>polystyrene. Yadav <strong>and</strong> Naik [21] reported that clay could be used as support for thePT catalyst; benzoic anhydride was prepared from benzoyl chloride <strong>and</strong> sodium benzoateusing a clay-supported quaternary ammonium salt at 30 C. The polymer-supportedcatalysts are less active than the clay-supported catalyst for this reaction system.Desikan <strong>and</strong> Doraiswamy [22] investigated the enhanced activity <strong>of</strong> polymer-supportedPT catalysts for the esterification <strong>of</strong> benzyl chloride with aqueous sodium acetate. Theyfound that the reactivity using a triphase catalyst is higher than that using a solubleone. They hypothesized that the enhancement due to increased lipophilicity <strong>of</strong> thepolymer-supported catalyst was more than compensated by the decreased diffusionalresistance.Jayach<strong>and</strong>ran <strong>and</strong> Wang [23] prepared a new PT catalyst, 2-benzilidine-N,N,N,N 0 ,N 0 ,N-hexaethylpropane-1,3-diammonium dibromide (Dq-Br), to investigateCopyright © 2003 by Taylor & Francis Group, LLC


the cycloalkylation <strong>of</strong> phenylacetonitrile (PAN) with an excess <strong>of</strong> 1,4-dibromobutaneusing aqueous sodium hydroxide as the base, <strong>and</strong> the following pseudo-first-order kineticswas observed:C 6 H 5 CH 2 CN þ BrðCH 2 Þ 4 Br Dq-Br !ð0:75 mol%Þ C 6 H 5 CðCH 2 Þ 4 CN ð6ÞHwang et al. [24] studied the Wittig reaction <strong>of</strong> benzyltriphenylphosphonium(BTPP) salts <strong>and</strong> benzaldehydes via ylide-mediated PTC. They concluded that the reaction<strong>of</strong> benzylidenetriphenyl phosphorane <strong>and</strong> the benzaldehyde in the organic phase is thedecisive step for stereoselectivity. The order <strong>of</strong> effectiveness <strong>of</strong> substituents isCF 3 > ðCl; BrÞ > MeO > F > NO 2 . Satrio <strong>and</strong> Doraiswamy [25] proposed a case studyfor the production <strong>of</strong> benzaldehyde in a possible industrial application <strong>of</strong> PTC. Thereaction between benzyl chloride <strong>and</strong> hypochlorite anion isC 6 H 5 CH 2 Cl ðorgÞþOCl ðaqÞ !C 6 H 5 CHO ðorgÞþHCl ðaqÞþCl ð7ÞThey show that the conventional route is the preferred one for a large-scale organicintermediate, <strong>and</strong> the improvements in merely one or two PTC steps can greatly enhancethe prospects <strong>of</strong> the PTC route.II.LIQUID–LIQUID PHASE TRANSFER CATALYSISA necessary condition for a reaction is to cause the collision <strong>of</strong> two reactant molecules. Itis obvious that the reaction rate <strong>of</strong> two immiscible reactants is low due to their lowsolubilities. A general method for overcoming this difficulty was to employ a protic oran aprotic solvent in order to improve their mutual solubilities. Nevertheless, thisimprovement was not very significant. The problems <strong>of</strong> two-phase reactions were notsolved until Jarrouse [26] discovered the catalyzing effect <strong>of</strong> quaternary ammonium saltin the aqueous–organic phase reaction system. PTC is an effective tool for synthesizingorganic chemicals from two immiscible reactants [27–32]. It has been extensively applied tothe synthesis <strong>of</strong> special organic chemicals by displacement, alkylation, arylation, condensation,elimination, oxidation, reduction, <strong>and</strong> polymerization. The advantage <strong>of</strong> PTC inthe synthesis <strong>of</strong> organic chemicals are fast reaction rate, high selectivity <strong>of</strong> product, moderateoperating temperature, <strong>and</strong> applicability to industrial-scale production.A. <strong>Mechanism</strong> <strong>of</strong> Liquid–Liquid <strong>Phase</strong> <strong>Transfer</strong> <strong>Catalysis</strong> (LLPTC)Quaternary salts, crown ethers, crypt<strong>and</strong>s, <strong>and</strong> polyethylene glycol (PEG) are the mostcommon agents used for LLPTC. Over the last few decades, the two reaction mechanismsused to describe the phenomenon <strong>of</strong> a two-phase PTC reaction were the Starks extractionmechanism <strong>and</strong> Mąkosza interfacial mechanism.1. Starks Extraction <strong>Mechanism</strong>This reaction mechanism described by Starks [28,33] is widely accepted for a catalysttransferring between the two phases. Reactions occurring in such systems involve: (1) thereactant reacting with catalyst in the normal phase to form an intermediate catalyticreactant, (2) transfer <strong>of</strong> the intermediate catalytic reactant from its normal phase intothe reaction phase, (3) transferred intermediate catalytic reactant reacting with untransformedreactant in the reaction phase to produce the product <strong>and</strong> catalyst, <strong>and</strong> (4)Copyright © 2003 by Taylor & Francis Group, LLC


transfer <strong>of</strong> catalyst from the reaction phase to the normal phase. The reaction mechanismcan be separated in three ways based on the reaction path, <strong>and</strong> can be described asfollows.(a) Normal Liquid–Liquid <strong>Phase</strong> <strong>Transfer</strong> <strong>Catalysis</strong> (N-LLPTC). Traditionally, moreapplications <strong>of</strong> PTC have been reported in N-LLPTC. The reaction mechanism (8) ismostly applied to alkylation, esterification etherification, <strong>and</strong> simple displacement reactionsin which a nucleophilic agent is transferred to the organic phase through the solublecatalyst therein:ð8Þ<strong>Mechanism</strong> (8) was first presented by Starks [33] for the reaction <strong>of</strong> 1-chloro-octane <strong>and</strong>aqueous sodium cyanide.(b) Inverse Liquid–Liquid <strong>Phase</strong> <strong>Transfer</strong> <strong>Catalysis</strong> (I-LLPTC). The organic reactantis converted, by means <strong>of</strong> a reagent (e.g., pyridine 1-oxide, PNO) partially soluble inthe organic phase, into a reactive ionic intermediate <strong>and</strong> transferred into the aqueousphase where reaction takes place to produce the desired product. The processes havebeen termed inverse phase-transfer catalysis [34–36]. The reaction mechanism can beexpressed as follows:ð9ÞThere are several examples where I-LLPTC has been used to synthesize acid anhydrides,by means <strong>of</strong> a substitution reaction, <strong>and</strong> ketones from oxidation <strong>of</strong> alcohols [37–40]. The reaction <strong>of</strong> an acid chloride (RX) with the carboylate ions (M þ R 0 ) catalyzed byPNO is to proceed through an intermediate 1-(acyloxy)pyridinium chloride formed in theorganic phase. PNO <strong>and</strong> N,N-dimethylaminopyridine (DMAP) are widely used as inversePT catalysts. The formation <strong>of</strong> hippuric acid was conducted in the presence <strong>of</strong> 4-dimethylaminopyridineas inverse PT catalyst [41].(c) Reverse Liquid–Liquid <strong>Phase</strong> <strong>Transfer</strong> <strong>Catalysis</strong> (R-LLPTC). This reactionmechanism was expressed as follows:Copyright © 2003 by Taylor & Francis Group, LLC


ð10ÞThe dehydrohalogenation reactions <strong>of</strong> alkyl halides take place in the presence <strong>of</strong>hydroxide ion <strong>and</strong> quaternary salts to form alkenes <strong>and</strong> alkynes [42–44]. The dehydrohalogenationis promoted by hydroxide ion. In general, two reaction conditions conducted inthis system were with highly lipophilic ammonium cation <strong>and</strong> 50% aqueous sodiumhydroxide. The reaction between 4-nitrobezenediazonium chloride <strong>and</strong> N-ethylcarbazolein aqueous media was accelerated by using a water–dichloromethane system containingsodium 4-dodecylbenzenesulfonate as a transfer catalyst for the diazonium ion [34].2. Mąkosza <strong>Interfacial</strong> <strong>Mechanism</strong>This reaction mechanism described by Mąkosza <strong>and</strong> Bialecka [45,46] is the acceptedcatalyst transport between the two phases. Reactions occurring in such systems involve:(1) transfer <strong>of</strong> ionic reactant from its normal phase <strong>and</strong> catalyst from the reaction phaseinto the interfacial region, (2) the ionic reactant reacting with catalyst in the interfacialregion to form intermediate catalytic reactant, (3) the intermediate catalytic reactanttransfer into the reaction phase to react with untransformed reactant to produce theproduct <strong>and</strong> catalyst. The reaction mechanism is expressed as follows:ð11ÞUsually, the aqueous salt could be too hydrophilic to allow the quaternary salt todissolve in the organic phase, <strong>and</strong> resided exclusively in the aqueous phase; anionexchange occured at or near the interface. The mechanism is applied to carbanion reactions,carbene reactions, condensation <strong>of</strong> polymerization, <strong>and</strong> C-alkylation <strong>of</strong> activemethylene compounds such as activated benzylic nitriles, activated hydrocarbons, <strong>and</strong>activated ketones under PTC=OH . In most cases, the reaction involves the Q þ OHcomplex because QOH is highly hydrophilic <strong>and</strong> has extremely low solubility in theorganic phase.A mechanism can also be applied when the quaternary salt is too lipophilic todissolve in the aqueous phase, <strong>and</strong> resides exclusively in the organic phase, anion exchangeoccuring at or near the interface. This parallel mechanism is called the Bra¨ ndstro¨ m–Montanari mechanism. The ion-exchange reaction existing at the interface was verifiedby L<strong>and</strong>ini et al. [47] <strong>and</strong> Bra¨ ndstro¨ m [48].Copyright © 2003 by Taylor & Francis Group, LLC


A summary <strong>of</strong> characteristic kinetic criteria to distinguish between the operation <strong>of</strong>the extraction <strong>and</strong> interfacial mechanisms has been suggested [28,49]. The extractionmechanism is characterized by: (1) increased rates with increased lipophilicity <strong>of</strong> catalyst,(2) reaction rates that are independent <strong>of</strong> stirring speed above a certain value, (3) firstorderor fractional dependence <strong>of</strong> reaction rate on catalyst concentration, <strong>and</strong> (4) pseud<strong>of</strong>irstor second-order kinetics if the reaction in the organic phase reaction is rate controllingor zero-order kinetics if diffusion across the interface is rate controlling.The interfacial mechanism is characterized by: (1) increased rates with increasedelectrostaticity <strong>of</strong> catalyst, (2) reaction rates are dependent on agitation rate, (3) fractionalkinetic order with respect to the catalyst concentration, <strong>and</strong> (4) the value <strong>of</strong> substrateacidity pK a is in the range 16–23.B. <strong>Kinetics</strong> <strong>of</strong> a Liquid–Liquid <strong>Phase</strong> <strong>Transfer</strong> <strong>Catalysis</strong>1. Starks Extraction <strong>Mechanism</strong>A typical LLPTC cycle involves a nucleophilic substitution reaction, as shown in Eq. (8).A difficult problem in the kinetics <strong>of</strong> PT-catalyzed reactions is to sort out the rate effectsdue to equilibrium anion-transfer mechanism for transfer <strong>of</strong> anions from the aqueous tothe organic phase. The reactivity <strong>of</strong> the reaction by PTC is controlled by the rate <strong>of</strong>reaction in the organic phase, the rate <strong>of</strong> reaction in the aqueous phase, <strong>and</strong> the masstransfer steps between the organic <strong>and</strong> aqueous phases [27–29]. In general, one assumesthat the resistances <strong>of</strong> mass transfer <strong>and</strong> <strong>of</strong> chemical reaction in the aqueous phase can beneglected for a slow reaction in the organic phase by LLPTC.Although a large number <strong>of</strong> papers have been published on the synthetic applications<strong>of</strong> PTC in the last three decades, little mathematical analysis <strong>of</strong> the phenomenon hasbeen done, <strong>and</strong> such an analysis is especially desirable in a large-scale application. Evans<strong>and</strong> Palmer [50] considered a process <strong>of</strong> interphase mass transfer <strong>and</strong> chemical reaction.Melville <strong>and</strong> Goddard [51] <strong>and</strong> Melville <strong>and</strong> Yortsos [52] presented an analysis <strong>of</strong> masstransfer in solid–liquid PTC. Chen et al. [53] derived algebraic expressions for the interphaseflux <strong>of</strong> QY <strong>and</strong> QX. The reaction parameters were estimated from experimental datausing a two-stage method <strong>of</strong> optimal parameters. Wang <strong>and</strong> Chang [54–56] studied thekinetics <strong>of</strong> the allylation <strong>of</strong> phenoxide with allyl chloride in the presence <strong>of</strong> PEG asLLPTC. A simple mathematical model describing the liquid–liquid PT-catalyzed reactionwith the two-film theory was analyzed [57–59]. The results <strong>of</strong> the model’s prediction areconsistent with experimental data. Such mathematical analysis appears desirable <strong>and</strong>needed in view <strong>of</strong> the widespread interest in PTC in the chemical industry in whichtwo-phase transfer <strong>and</strong> triphase catalysis are the most common industrial processes.The reactivity in phase-transfer catalysis is controlled by: (1) the reaction rate in theorganic phase, (2) the mass transfer steps between the organic <strong>and</strong> aqueous phases, <strong>and</strong> (3)the distribution equilibrium <strong>of</strong> the quaternary salts between the two phases. The distribution<strong>of</strong> quaternary salts between two phases directly affects the entire system reactivity[60–62]. On the basis <strong>of</strong> the experimental data <strong>and</strong> earlier literature [27,28,63], a generalizedapproach describing a LLPTC reaction system uses a pseudo-first-order reaction. Therate expression is written asd½RXŠ¼ kdt int ½QYŠ½RXŠ¼ k app ½RXŠð12Þð13ÞCopyright © 2003 by Taylor & Francis Group, LLC


The fixed value <strong>of</strong> k app is called the apparent first-order reaction-rate constant. The overbardenotes the species in the organic phase. The reaction rate linearly increases withincreasing QY concentration. Equation (13) is established when the QY concentration isconstant. Most observed reaction rate would follow the pseudo-first-order kinetics for anexcess amount <strong>of</strong> aqueous reactant to that <strong>of</strong> organic reactant [37]. Wu [64] indicated thata pseudo-first-order hypothesis can be used to describe the PTC experiment data, eventhough the QY concentration is not kept constant. Wang <strong>and</strong> Wu [58] developed a comprehensivemodel in a sequential phosphazene reaction. Their experimental results wereconsistent with a first-order reaction rate; the pseudo-first-order reaction-rate constantwas not linearly related to the concentration <strong>of</strong> the catalyst, because the mass transfer<strong>of</strong> catalyst between the two phases influenced the reaction. Wang <strong>and</strong> Yang [57,65] <strong>and</strong>Wu [63] indicated that the QY concentration is constant over time when the molar ratio <strong>of</strong>nucleophile to catalyst is larger than unity. Therefore, in the general case, the QY concentrationcannot vary with time only when the ion-exchange rate in the aqueous phase ismore rapid than that in the organic phase [66], no mass transfer resistance <strong>of</strong> catalystbetween the two phases occurs, the molar ratio <strong>of</strong> nucleophile to catalyst is larger thanunity, <strong>and</strong> the ionic strength in the aqueous phase is high [67].The complicated nature <strong>of</strong> the LLPTC reaction system is attributed to two masstransfer steps <strong>and</strong> two reaction steps in the organic <strong>and</strong> aqueous phases. The equilibriumpartition <strong>of</strong> the catalysts between the two phases also affects the reaction rate. On the basis<strong>of</strong> the above factors <strong>and</strong> the steady-state two-film theory [60,63,64,68], a phase-planemodel to describe the dynamics <strong>of</strong> a liquid–liquid PTC reaction has been derived. Thismodel <strong>of</strong>fers physically meaningful parameters that demonstrate the complicated reactivecharacter <strong>of</strong> a liquid–liquid PT-catalyzed reaction. However, when the concentration <strong>of</strong>aqueous solution is dilute or the reactivity <strong>of</strong> aqueous reactant is weak, the onium cationhas to exist in the aqueous phase. The mathematical model cannot describe this completely.When the onium cation exists in the aqueous phase, several important phenomenainvolved in the liquid–liquid reaction need to be analyzed <strong>and</strong> discussed.ð14ÞOn the basis <strong>of</strong> Eq. (12), <strong>and</strong> mechanism (14) [64,68], the species balance equationswere solved by eliminating the time variable (phase-plane model). The relevant rate equationsaredy od ¼ y 1oy ody 1ody o¼ P 1 QYy oym 1aQY 1y 1oð15Þð16ÞCopyright © 2003 by Taylor & Francis Group, LLC


dy 1ady o¼ QYy ody 2ody o¼ QXy ody 2a¼ dy 2o y 1o y oym 1a QY 1 1 y 3a y 4a yþ 1ay 1o K d1 y 1o y 1o y 1o y oy 2o ym 2ay QX P1o y 11oy 2a 2 y 3a y 5ak d2 y 1o y o QXy oy 2oy 1om QXy 2ay 1ody 3ady o¼ 2y 3a y 5aK d2 y 1o y oþ 1y 3a y 4aK d1 y 1o y o 1y 1ay 1o y o 2y 2ay 1o y oð17Þð18Þð19Þð20ÞThe mass balances for Q i ,Y , <strong>and</strong> Xare given below:1 ¼ y 1o þ y 1a þ y 2o þ y 2a þ y 3a ð21Þy 4a ¼ P 2 y 1a y 1o þðy o 1ÞP 1 ð22Þy 5a ¼ P 3 þ 1 y 2a y 2o þð1 y o ÞP 1 ð23Þin which the dimensionless variables <strong>and</strong> parameters are defined asy o ¼ ½RXŠ o; y½RXŠ 1o ¼ V o½QYŠ oiy 3a ¼ V a½Q þ Š aQ iQ i; y 4a ¼ V a½Y Š aQ i QY ¼ K QYA=V o; Pk o Q i =V 1 ¼ V 0½RXŠ ioQ i; y 1a ¼ V a½QYŠ aQ i; y 2o ¼ V o½QXŠ o; yQ 2a ¼ V a½QXŠ a;iQ i; y 5a ¼ V a½X Š a; Q QX ¼ K QXA=V a;ik o Q i =V o; P 2 ¼ V a½MYŠ i; PQ 3 ¼ V a½MXŠ i;iQ i 1 ¼ k d1k o; 2 ¼ k d2k o; ¼ V oV a; ¼ tk oQ iV oð24Þ<strong>and</strong> ½MXŠ i , ½MYŠ i , <strong>and</strong> ½RXŠ i represent the initial concentrations <strong>of</strong> reactants MX, MY,<strong>and</strong> RX, respectively. By introducing the values <strong>of</strong> the parameters into Eqs (15)–(23), thedynamic phenomena <strong>of</strong> a liquid–liquid PT-catalytic reaction was obtained.Wang <strong>and</strong> Yang [57] reported that the ion-exchange reaction-rate constant wascalculated with three differential equations as below for the dynamics <strong>of</strong> QY in boththe aqueous <strong>and</strong> organic phases in a two-phase reaction without adding the organicreactant by the numerical shooting method <strong>and</strong> correlating it with the experimental data.d C QYdtdC QYdt¼ K QY AC QY¼ K da C QY C QXC QY =m QYK QY A VV C QYC QY =m QYð25Þð26ÞdC MYdt¼ K da C QY C QX ð27ÞThe intrinsic reaction-rate constant in the organic phase is obtained by reacting QYwith RX in a homogeneous solvent <strong>and</strong> using Eq. (12). According to the literature, Wang<strong>and</strong> Yang [57] <strong>and</strong> Wu <strong>and</strong> Meng [69] have found the intrinsic reaction-rate constant fromtheir systems. The equilibrium constant <strong>and</strong> mass transfer constant <strong>of</strong> the catalyst betweentwo phases obtained are discussed in the next section.Copyright © 2003 by Taylor & Francis Group, LLC


Wu [64] characterized the transfer <strong>of</strong> Q þ X from the organic phase to the aqueousphase <strong>and</strong> <strong>of</strong> Q þ Y from the aqueous to the organic phase by definingQY ¼ y 1am QYy 1o; QX ¼ y 2oy 2a m QXð28ÞIf the PT catalysts in the two phases are in extractive equilibrium <strong>and</strong> the mass transferresistance can be neglected completely, then QY <strong>and</strong> QX are each equal to 1.The dynamics for a slow PT reaction <strong>and</strong> a mass transfer controlled instantaneousreaction were studied. Wu [63] <strong>and</strong> Wu <strong>and</strong> Meng [69] indicated that the pseudo-steadystateLLPTC model could describe the complicated nature <strong>of</strong> the LLPTC reaction. Therate equation from the report <strong>of</strong> Wu [63] is expressed asd½RXŠdtk½RXŠQ¼1 = Vm QY þ 1m þ Da ð29ÞQYDaQY m þ Da QY QX þð1 þ m QX Þ QY þ 1m þ þ QYwhere Da QY ð¼ k½RXŠ=k QY A= VÞ <strong>and</strong> Da QX ð¼ k½RXÞ=K QX A= VÞ are the Damkohlernumbers for QY <strong>and</strong> QX, respectively; ð¼ k 2 ½MXŠ=k 2 ½MYŠÞ is the reaction ratio <strong>of</strong>the aqueous reverse reaction to the forward reaction for ion exchange; <strong>and</strong> ð¼ k½RXŠ=k 2 ½MYŠÞ is the reaction ratio <strong>of</strong> the organic phase to the aqueous forward ionexchangereaction.Wu [63] also derived an expression for the catalyst effectiveness, which is defined asthe ratio <strong>of</strong> the actual reaction rate to that with all the catalyst present as QY, in terms <strong>of</strong>seven physically meaningful dimensionless parameters: ¼ m QY þ 1þ Da QY Da QY þ 11þ Dam QY m QX þ 1 þ m QX þ ð30ÞQY m QYBefore evaluating Eq. [30], the parameters <strong>of</strong> kinetics, mass transfer, <strong>and</strong> thermodynamicequilibrium must be established. The aim <strong>of</strong> this work is to evaluate the equilibrium <strong>and</strong>extraction <strong>of</strong> a quaternary salt in an organic solvent/aqueous solution. The studies ondistribution equilibrium <strong>of</strong> the quaternary salts enable one to clarify the true mechanismthrough which the reactant anion is transferred.Models for LLPTC get even more complicated for special cases, e.g., reactions inboth aqueous <strong>and</strong> organic phases, systems involving a base reaction, or other complexseries–parallel multiple reactions. Wang <strong>and</strong> Wu [58] <strong>and</strong> Wu <strong>and</strong> Meng [69] studied thekinetics <strong>and</strong> mass transfer for a sequential reaction using LLPTC that involved a complexreaction with six sequential S N 2 reactions in the organic phase along with interphase masstransfer <strong>and</strong> ion exchange in the aqueous phase.Wang <strong>and</strong> Wu [70] analyzed the extraction equilibrium <strong>of</strong> the effects <strong>of</strong> catalyst,solvent, NaOH/organic substrate ratio, <strong>and</strong> temperature on the consecutive reactionbetween 2,2,2-trifluoroethanol with hexachlorocyclotriphosphazene in the presence <strong>of</strong>aqueous NaOH. Wu <strong>and</strong> Meng [69] reported the reaction between phenol with hexachlorocyclotriphosphazene.They first obtained the intrinsic reaction-rate constant <strong>and</strong> overallmass transfer coefficient simultaneously, <strong>and</strong> reported that the mass transfer resistance <strong>of</strong>QX from the organic to aqueous phase is larger than that <strong>of</strong> QY from the aqueous toorganic phase. The intrinsic reaction-rate constant <strong>and</strong> overall mass transfer coefficientswere obtained in three ways.Copyright © 2003 by Taylor & Francis Group, LLC


(a) Pseudo-Steady-State LLPTC model.The reaction relationship is given as1k app¼V K QX A þ V kQ ið31Þwhere denotes the reactivity <strong>of</strong> the phosphazene reaction. The plot <strong>of</strong> 1=k app versus , inwhich the data were measured at the initial time <strong>of</strong> different experimental runs, allows oneto obtain the mass transfer coefficient, K QX A, <strong>and</strong> the intrinsic reaction rate constant k,from the slope <strong>and</strong> intercept <strong>of</strong> the straight line.(b) Extrapolation Method. If mass transfer resistance influences the reaction, the concentration<strong>of</strong> the active catalyst QY cannot remain constant during the course <strong>of</strong> thereaction. Decreasing the concentration <strong>of</strong> organic reactant RX increases the apparentfirst-order reaction-rate constant. When the concentration <strong>of</strong> organic reactant decreases,both the reaction rate <strong>and</strong> the effect <strong>of</strong> mass transfer decrease. If the organic reactantconcentration extrapolates to zero ð½RXŠ !0Þ, the effect <strong>of</strong> mass transfer can beneglected. The intrinsic reaction-rate constant, k, is easily evaluated.(c) Half-Reaction in the Organic <strong>Phase</strong>. The organic reactant reacted with an intermediatecatalyst, tetra-n-butyl ammonium phenolate, in a homogeneous organic phase.The intrinsic reaction-rate constant was calculated from Eq. (12).Another LLPTC is usually performed in an agitated system, in which the organicphase is mostly dispersed. Several efforts have been made in developing the theory for atwo-liquid phase with chemical reactions. For an organic phase being the dispersed phase,several phenomena take place: (1) formation <strong>of</strong> a single droplet in the continuous phase bystirring, (2) free rise or fall <strong>of</strong> a droplet through the continuous phase, <strong>and</strong> (3) coalescence<strong>of</strong> a droplet at the end <strong>of</strong> the free-rise period. During the extraction <strong>of</strong> a catalytic intermediate,mass transfer from the bulk aqueous phase to the organic droplet surface influencesthe rate <strong>of</strong> PT reaction. Yang [71,72] studied the general analysis <strong>of</strong> the dynamics <strong>of</strong>a PT-catalytic reaction in a dispersed system <strong>of</strong> liquid–liquid phases, considering theirreversible <strong>and</strong> reversible reactions by solving the finite difference <strong>and</strong> Runge–Kuttafourth-order methods. The rates <strong>of</strong> change <strong>of</strong> RX, RY, QX, <strong>and</strong> QY in an organic dropletare described by the instantaneous equations <strong>of</strong> diffusion <strong>and</strong> reaction with the correspondinginitial <strong>and</strong> boundary conditions as follows:@ C i@t ¼ D i @r 2 @ C iþ @r @r iR; i ¼ RX; RY; QX; <strong>and</strong> QY ð32Þr 2where i is the stoichiometric coefficient <strong>of</strong> the i component.The kinetics <strong>of</strong> inverse PT-catalytic extraction <strong>of</strong> species into the water phase wascarried out with partially water-soluble pyridines or derivatives [36,38,40,59,73], as shownin mechanism (9). These reactions can be described by a pseudo-first-order hypothesis[38,40]:k app ¼ k h þ k c ½PNOŠ ið33ÞHowever, so far, the detailed kinetics <strong>of</strong> I-LLPTC are unclear.As mentioned above, the various approaches to LLPTC modeling have been taken,<strong>and</strong> a comprehensive general model for N-LLPTC reactions is widely held. However, akinetic model for I-LLPTC <strong>and</strong> R-LLPTC reactions is yet to be developed.Copyright © 2003 by Taylor & Francis Group, LLC


2. Mąkosza <strong>Interfacial</strong> <strong>Mechanism</strong>The interfacial mechanism is the most widely accepted mechanism for PTC reactions in thepresence <strong>of</strong> a base. However, although there are numerous industrially important applications,very few kinetic studies or mathematical models for this mechanism are reported. Ingeneral, the mechanism is also described by a pseudo-first-order hypothesis.Juang <strong>and</strong> Liu [74,75] proposed <strong>and</strong> discussed a possible mechanism based on amixed Mąkosza <strong>and</strong> modified interfacial mechanism. The reaction rate for the etherification<strong>of</strong> a substituted phenylacetic acid by PTC was measured using a constant interfacialarea cell, <strong>and</strong> expressed asR f ¼ k½R 0 XŠ 1=3 ½RHŠ½QXŠ½OH Š 5=21 þ k a ½QXŠ 1=2 ½OH Šþk b ½RHŠ 1=2 ½OH Šð34ÞC. Thermodynamic Equilibrium in LLPTCQuaternary salts are generally used as normal liquid–liquid PT catalysts. In general, thefunctional groups <strong>of</strong> the quaternary cation will affect the dissolution <strong>of</strong> the catalyst in theorganic phase. Further, the phase transfer <strong>of</strong> the anion will also affect the reaction rate inthe two-phase reaction. Therefore, a proper choice <strong>of</strong> PT catalyst is very important inpromoting the reaction rate. Unfortunately, a universal guideline is unavailable for selectingthe proper PT catalyst to enhance the reaction. The reactivity in PTC is controlled by:(1) the reaction rate in the organic phase, (2) the mass transfer steps between the organic<strong>and</strong> aqueous phases, <strong>and</strong> (3) the distribution equilibrium <strong>of</strong> the quaternary salts betweenthe two phases. The distribution <strong>of</strong> quaternary salts between two phases directly affects theentire system reactivity [60–62].In general, anion transfer <strong>and</strong> anion activation are the important steps involved intransferring anions from the aqueous phase to the organic phase where the reaction takesplace. Factors affecting the extraction ability <strong>of</strong> the anion from the aqueous to organicphase include cation–anion interaction energies, the ionic strength in the aqueous phase,ion-pair hydration, the lipophilicity <strong>of</strong> the catalyst, <strong>and</strong> the polarity <strong>of</strong> the organic phase.The extraction behavior <strong>and</strong> distribution coefficients <strong>of</strong> quaternary salts in various mediahave also been investigated [76–86].Bra¨ ndstro¨ m [48] indicated that the distribution <strong>of</strong> quaternary salt between two(liquid–liquid) phases exists as complicated multiequilibrium constants, which dependon the structure <strong>of</strong> the anion, cation, <strong>and</strong> solvent, as well as on pH, ionic strength, <strong>and</strong>concentrations in the aqueous solution. Such equilibrium properties have not yet beenevaluated completely. The relationship between quaternary salt <strong>and</strong> extraction constant isan important consideration for PTC work.The distribution coefficient <strong>of</strong> quaternary cation D Q was obtained by measuring theconcentrations <strong>of</strong> quaternary cation (Q) in the organic <strong>and</strong> aqueous phases, respectively.The distribution coefficient is highly dependent on the nature <strong>and</strong> concentration <strong>of</strong> thequaternary salts:D Q ¼ ½QŠ obs½QŠ obsð35ÞThe distribution coefficient <strong>of</strong> quaternary cations between both the phases not onlyprovides information on the phases to facilitate the modeling <strong>of</strong> the two-phase transfercatalysis system, but it can also give a criterion for evaluating the suitability <strong>of</strong> the catalyst.Copyright © 2003 by Taylor & Francis Group, LLC


The order <strong>of</strong> magnitude <strong>of</strong> D Q for quaternary salts is Aliquat 336 > TBA-TBPO >TBAI > TBPB > TBAB > TBAC. The sequence <strong>of</strong> D Q for solvents is CHCl 3 >CH 2 Cl 2 > 1;2-C 2 H 2 Cl 2 > C 6 H 5 Cl. The order <strong>of</strong> influence on the extraction capability<strong>of</strong> quaternary salts is Br 3 C 6 H 2 O > I > Br < Cl <strong>and</strong> P þ > N þ for the anion <strong>and</strong>central cation, respectively. Reasons for these behaviors have been discussed in previouswork [48,76,81,85,86]. The D Q value increased on increasing the temperature.The true extraction constants <strong>of</strong> quaternary salts QX corresponding to their infinitelydilute solutions in a two-phase system were calculated using the following equation:a QXEQX T ¼a Qþ; a X½QXŠ¼½Q þ Š½X Š2ð36Þwhere a <strong>and</strong> 2 are the activity <strong>and</strong> the mean ionic activity coefficient <strong>of</strong> the quaternarysalts, respectively.The distribution constant <strong>of</strong> quaternary salt at equilibrium between two phases ism ¼ ½Qþ X Š½Q þ X Šð37ÞThe dissolved Q þ Xin the aqueous <strong>and</strong> organic phase may dissociate toQ þ X Ð Q þ þ X ð38ÞQ þ X Ð Q þ þ X ð39ÞThus, the dissociation constants K da <strong>and</strong> K da <strong>of</strong> QX in the aqueous <strong>and</strong> organic phases arewritten asK da ¼ ½Qþ Š½X Š 2 ½Q þ X ŠK do ¼ ½Qþ Š½X Š 2 ½Q þ X Šð40Þð41ÞThe dissociation constant in aprotic organic solvents can be derived from fundamentalprinciples based on Bjerrum’s theory for ion pairs. In most organic media, thedissociation constant <strong>of</strong> ion pairs is very low (<strong>of</strong> the order <strong>of</strong> around 10 5 ) [48].Bra¨ ndstro¨ m [87] reported that the ionic aggregation states <strong>of</strong> quaternary salts existingin the organic phase were <strong>of</strong> various types, i.e., dissociated ions (Q þ þ X ), ion pairs(Q þ X ), quadruples ½ðQ þ X Þ 2 Š, etc. Hence, the total concentration <strong>of</strong> quaternary salt inan organic phase can be written asC Q ¼½Q þ Šþ½QXŠþ2½Q 2 X 2 Šþð42ÞSince the organic system is in electrical neutrality,½Q þ Š¼½X ŠEquation (42) can be transformed intoð43ÞC Q ¼ E T1=2Q þ ½Q þ Š½X Š 1=2þETQX 2 ½Q þ Š½X Š þ 2E T Q 2 X 22 4 ½Q þ Š½X Š 2þ ð44Þwhere EQ T þ, ET QX, <strong>and</strong> EQ T 2 X 2are the concentration quotients represented asCopyright © 2003 by Taylor & Francis Group, LLC


E T Q ¼ ½Qþ Š½X Š2 þ½Q þ Š½X Š2EQX T ½QXŠ¼½Q þ Š½X Š2E T Q 2 X 2¼ ½Q 2X 2 Š½Q þ Š½X Š 2 ð45Þð46Þð47ÞBy using Eqs (42)–(47), the values <strong>of</strong> E T Q þ, ET QX, EQ T 2 X 2, <strong>and</strong> the distribution constantm are evaluated. Corrections for the mean activity coefficient in the organic phase weremade using the Marshall <strong>and</strong> Grunwald expression, <strong>and</strong> the values <strong>of</strong> m, K da , K do , <strong>and</strong> were calculated by a numerical iteration method. Beronius <strong>and</strong> Bra¨ ndstro¨ m [91] evenclarified the identical value <strong>of</strong> K do at ½QXŠ ¼0 within the limits <strong>of</strong> experimental error<strong>and</strong> the conductance measurement. In view <strong>of</strong> past reports [87–92], most K da valueswere located in the range between 1 <strong>and</strong> 10; K do values were located in the range between10 1 <strong>and</strong> 10 5 . The dissociation ability <strong>of</strong> quaternary salt in the aqueous phase is greaterthan that in the organic phase.The quaternary salts QX can be completely dissociated to free ions (Q þ <strong>and</strong> X )inthe aqueous phase (, ½Q þ Š=½QXŠ > 100Þ <strong>and</strong> partially dissociated in the organic phasewhen the concentration <strong>of</strong> the quaternary salt is 0.0125 kmol/m 3 . The quaternary salts QXcan be partially dissociated to free ions in the aqueous <strong>and</strong> the organic phases when theconcentration <strong>of</strong> quaternary salt is 0:1 kmol=m 3 . The incremental rules <strong>of</strong> the dissociationdegree <strong>of</strong> the quaternary salts were obtained as follows: (1) increasing the charge-tovolumeratio <strong>of</strong> the central cation or counteranion (e.g., P þ > N þ or I > Br > Cl ),(2) increasing the electron-releasing groups on the quaternary cation (e.g., Aliquat336 > TBAC), <strong>and</strong> (3) increasing the electron-withdrawing groups on the quaternaryanion (e.g., TBA-TBPO > TBA-BPO > TBAC). Electron-releasing (or electron-withdrawing)groups apparently make the transition state more stable on the quaternarycation (or anion) while the ion-pair type <strong>of</strong> quaternary salts transferring through theinterface between two phases is a transition state. Bockries <strong>and</strong> Reddy [93] reportedthat the association constant decreased when the effective ionic radius <strong>of</strong> the ion pairwas increased.Quaternary salts in an organic phase must be determined experimentally to knowwhether the salts are dissociated or associated, <strong>and</strong>, if so, to what degree. The hydration <strong>of</strong>the anion plays an important role in dissociating the catalyst. Furthermore, the solvation<strong>of</strong> the anions increases the size <strong>of</strong> the ions, decreases their mobility <strong>and</strong> diffusion rate, <strong>and</strong>reduces the reactivity <strong>of</strong> the reactant. How many molecules <strong>of</strong> the coextracted water doeseach quaternary salt carry? Hence, the equation for the distribution <strong>of</strong> a tetralkylammoniumhalide into an organic phase can be written as [94,95]Q þ þ X þjH 2 O Ð Q þ þ X :jH 2 OQ þ þ X þjH 2 O Ð Q þ X :jH 2 Oð48Þð49ÞDepending on whether the species in the organic phase is dissociated as free ions [Eq. (48)]or associated as ion pairs [Eq. (49)], the corresponding equilibrium constants can bewritten asCopyright © 2003 by Taylor & Francis Group, LLC


E T Q þ ;H 2 O ¼ ½Qþ Š½X :jH 2 OŠ 2 ½Q þ Š½X Š½H 2 OŠ j 2 EQX;H T ½QX:jH2 O ¼2 OŠ½Q þ Š½X Š½H 2 OŠ j 2ð50Þð51ÞThe j value can be calculated by dividing ðH 2 OÞ by the amount <strong>of</strong> quaternary salts in theorganic phase. The water content difference in the organic phase ððH 2 OÞÞ equals thedifference between the measured water content in the solvent <strong>and</strong> that in the solution atthe same temperature.The order <strong>of</strong> magnitude <strong>of</strong> H 2 O in the organic phase for quaternary salts is Aliquat336 > TBA-TBPO > TBAI > TBPB > TBAB > TBAC. The sequence <strong>of</strong> ½H 2 OŠ for solventsis 1,2-C 2 H 4 Cl 2 > CH 2 Cl 2 > CHCl 3 > C 6 H 5 Cl. This tendency <strong>of</strong> the sequence <strong>of</strong> thecoextracted water is identical to that <strong>of</strong> the solubility <strong>of</strong> water in the organic phase <strong>of</strong> 1,2-C 2 H 4 Cl 2 ð1:3Þ > CH 2 Cl 2 ð0:81Þ > CHCl 3 ð0:08Þ > C 6 H 5 Cl ð0:05Þ at 20 C. The orders <strong>of</strong>influencing extraction capability <strong>of</strong> H 2 O are Cl > Br 3 C 6 H 2 O > Br > I <strong>and</strong> N þ > P þfor the anion <strong>and</strong> central cation, respectively. The trend for water content in the organicphase varied with increasing temperature. L<strong>and</strong>ini et al. [96] indicated that the solvatingcapability between quaternary salt <strong>and</strong> water could reduce the quaternary salt’s reactivityin the organic phase in a PT-catalyzed reaction. This result was confirmed by previous work[61,76]. Hence, it is significant to study the liquid–liquid PT-catalyzed reaction <strong>and</strong> toevaluate how many molecules <strong>of</strong> the coextracted water are carried by each quaternarysalt. The water content in the organic phase increased with increasing temperature. The ½H 2 OŠ value increased when the charge-to-volume ratio <strong>of</strong> the anion increased <strong>and</strong> when thepolarity <strong>of</strong> the solvent increased, but decreased as the lipophilicity <strong>of</strong> the quaternary saltincreased. These tendencies correspond to those reported by L<strong>and</strong>ini <strong>and</strong> coworkers[97,98]. Kenjo <strong>and</strong> Diamond [95] reported that the average water contents in a nitrobenzene/watersystem at 23 C were 3.3, 1.8, <strong>and</strong> 1 (mol/mol quaternary salt) for Cl ,Br , <strong>and</strong>I , respectively. Starks <strong>and</strong> Owens [99] reported that the hydration numbers <strong>of</strong>C 16 H 33 Bu 3 P þ X were 0.4, 4, <strong>and</strong> 5 for NO 3 ,Cl , <strong>and</strong> CN , respectively. The averagewater content in the organic phase ð½H 2 OŠÞ was about 1–3 mol/mol quaternary salt, exceptfor TBAC. Because the hydration numbers for different anions were different when thequaternary salt was TBA þ [(n-C 4 H 9 Þ 4 N þ Š, the results demonstrate that the water <strong>of</strong> hydrationis primarily associated with the anion, rather than with the quaternary cation.Quaternary ammonium ions are used as PT catalysts because they are least likely tointerfere in chemical reactions. According to the experimental results <strong>of</strong> Bra¨ ndstro¨ m [48],Herriott <strong>and</strong> Picker [100], <strong>and</strong> L<strong>and</strong>ini et al. [97], the organophilic quaternary cationsserved as more effective PT catalysts than quaternary cations with small alkyl chains.Thus, the incremental number <strong>of</strong> C atoms surrounding the central atom (e.g., N) <strong>of</strong> aquaternary salt will increase its lipophilicity, thus raising the extraction constant.However, these researchers did not give the relationship between the extraction constant<strong>and</strong> the structure <strong>of</strong> quaternary salts. According to the literature, four relationships forquaternary cations have been reported.1. Gustavii [101] observed a linear relationship between log E QX <strong>and</strong> n, the number<strong>of</strong> C atoms in an ammonium ion. He extracted picrates into methylene chloride usingprimary amines as well as symmetrical secondary <strong>and</strong> tertiary amines <strong>and</strong> symmetricalquaternary ammonium salts. The relationships for quaternary ammonium salts islog E Q picrate ¼ 2:0 þ 0:54n.Copyright © 2003 by Taylor & Francis Group, LLC


2. A quantitative parameter for characterizing accessibility was suggested [28]based on the strong dependence <strong>of</strong> electrostatic interaction on the distance <strong>of</strong> closestapproach between the cation <strong>and</strong> anion (which is determined by steric factors). Thisparameter, termed q, is simply the sum <strong>of</strong> the reciprocals <strong>of</strong> the length <strong>of</strong> the linearalkyl chains attached to the central nitrogen <strong>of</strong> the quaternary cation;q ¼ 1=C# 1 þ 1=C# 2 þ 1=C# 3 þ 1=C# 4 , where C# is the number <strong>of</strong> carbon atoms in each<strong>of</strong> the four alkyl chains in the quaternary cation.3. Fukunaga et al. [102] had presented a correlation function based on hydrophile–lipophile balance (HLB) ideas to assess the efficiency <strong>of</strong> quaternary salts in the benzene–watersystem in terms <strong>of</strong> Hildebr<strong>and</strong> parameters ½Dð QX Þ¼ð QX Þ 2 =ð QX Þ 2 Š where QX , <strong>and</strong> are, respectively, the solubility parameters <strong>of</strong> the catalyst, water, <strong>and</strong> organic solvent.4. Sirovski [103] proposed that the structure–activity relationship for quaternarysalts can be described quantitatively using Hansch -hydrophobicity constants. Theseconstants are defined analogously to Hammett <strong>and</strong> Taft constants [104]: x ¼ log P x log P H , where P H is the distribution coefficient for the st<strong>and</strong>ard compound,<strong>and</strong> P x is the same from its derivative with the X substituent in the st<strong>and</strong>ard 1-octanol–water system, which has low ion selectivity in relation to halide <strong>and</strong> hydroxide ions.The former two relationships (paragraphs (1) <strong>and</strong> (2) above) were focused on to accessthe distribution <strong>of</strong> quaternary cations. The equilibrium property cannot reveal when thetotal carbon number for various quaternary salts is the same. In paragraph 3, theHildebr<strong>and</strong> parameter cannot be easily obtained for all quaternary salts. Hence, we tookthe results <strong>of</strong> paragraphs 1–3 <strong>and</strong> the concept <strong>of</strong> HLB for the surfactant to show that thedispersal efficiency <strong>of</strong> surfactant or emulsifier molecules is a function <strong>of</strong> the relative interactions<strong>of</strong> their polar, hydrophilic ‘‘heads’’ with the aqueous phase <strong>and</strong> <strong>of</strong> their nonpolar,lipophilic ‘‘tails’’ with the hydrocarbon phase [105,106]. We developed a new model as0:475HLB ¼ qðM T M H Þ þ 9:4 M TBABð52ÞM NXin which 0.475 <strong>and</strong> 9.4 are hydrophilic group numbers <strong>of</strong> CH 2 <strong>and</strong> N, respectively, whichwere defined by Davies [107]. The equation <strong>of</strong> the HLB was developed in respect <strong>of</strong> theextraction <strong>of</strong> quaternary salts between two phases based on molecular weights <strong>of</strong> hydrophilic<strong>and</strong> lipophilic groups. A linear relationship between extraction constant <strong>and</strong> HLBwas observed for ammonium cations. An average decrease in log E QX is about 10:5 2unts per HLB value for various counteranions. The free energies <strong>of</strong> transfer for ion pairs<strong>and</strong> dissociated ions were determined <strong>and</strong> were shown to correspond to the experimentaldata in the literature.It is <strong>of</strong> interest to determine the crude free energies <strong>of</strong> phase transfer between organic<strong>and</strong> aqueous phases for the quaternaries. This is combined with the free energies <strong>of</strong>transfer for halide ions to give the free energies for the tetrabutylammonium <strong>and</strong> tetrabutylphosphoniumions, which are not well established. Do different salts give the samevalues? Tseng [92] reports the free energy <strong>of</strong> transfer <strong>of</strong> some anions from water to variouskinds <strong>of</strong> solvents based on the distribution data for quaternary salts, <strong>and</strong> evaluates theextraction behavior <strong>of</strong> quaternary onium salts in order to underst<strong>and</strong> their performance ina PT-catalyzed reaction system.An extensive <strong>and</strong> self-consistent set <strong>of</strong> data on free energies <strong>of</strong> transfer <strong>of</strong> someinorganic salts has been reported [89]. The free energy <strong>of</strong> the extraction constant, distributionconstant, <strong>and</strong> dissociation constant are expressed asG i ¼ RT lnðiÞ; i ¼ E QX ; E T QX; m; K da ; or K do ð53ÞCopyright © 2003 by Taylor & Francis Group, LLC


Gustavii [101] <strong>and</strong> Bockries <strong>and</strong> Reddy [93] indicated that the dissociation constant<strong>of</strong> quaternary salts increased when the dielectric constant <strong>of</strong> the solvents was increased.Nagata [108] reported that the logarithmic value <strong>of</strong> the association constant <strong>of</strong> a quaternarysalt was proportional to the reciprocal <strong>of</strong> the dielectric constant <strong>of</strong> the mixed solvent.According to Eq. (53), the dissociation constant decreased slightly with increasing values<strong>of</strong> the reciprocal <strong>of</strong> the dielectric constant. Parker et al. [109] demonstrated that the freeenergies <strong>of</strong> transfer are very useful in correlation with the solvent effects on S N 2 effects inPTC. The free energies <strong>of</strong> transfer for the quaternary salts <strong>of</strong> dissociated ions from waterto the solvent can be written asG t Q þ þX ¼ RT ln a !Qþ a X¼ RT ln E T a Qþ aQXK do ð54ÞXThe free energies <strong>of</strong> transfer for free ions from water to the solvent can be written asG t i ¼ RT ln a i; i ¼ Q þ or X ð55Þa iAbraham [88], Czapkiewicz et al. [110], <strong>and</strong> Taft et al. [111] have reported the freeenergies <strong>of</strong> transfer <strong>of</strong> (C n H 2nþ1 Þ 4 NX (n ¼ 1–3) for ion pairs <strong>and</strong> dissociated ions. The freeenergies <strong>of</strong> transfer for quaternary salts <strong>of</strong> ion pairs <strong>and</strong> dissociated ions from water t<strong>of</strong>our kinds <strong>of</strong> organic solvents were determined in these studies. The free energies <strong>of</strong>transfer for ion pairs were less than those for dissociated ions, i.e., the transfer ability<strong>of</strong> ion pairs was greater than that <strong>of</strong> dissociated ions. The result <strong>of</strong> the stronger cation–anion attraction in ion pairs is to reduce significantly the magnitudes <strong>of</strong> the endoergicsolvent cavity terms, as well as the exoergic anion–solvent attractive terms. The stability <strong>of</strong>quaternary salts for ion pairs was greater than that for dissociated ions from water to theorganic phase. The result corresponds to that <strong>of</strong> Taft et al. [111]. The sequences <strong>of</strong> freeenergy <strong>of</strong> transfer for quaternary salts are <strong>of</strong> three sorts: (1) P þ > N þ , (2)TBPO < I < BPO < Br < Cl , <strong>and</strong> (3) the long chain <strong>of</strong> an alkyl group is <strong>of</strong> lowvalue (Aliquat 336 < TBAC). The stability <strong>of</strong> ion pairs in dichloromethane (or dissociatedions in chlor<strong>of</strong>orm) was the highest among the four kinds <strong>of</strong> solvents. These results revealthat the incremental charge localization in the anion <strong>and</strong> decrement in the cation increasesthe stability <strong>of</strong> quaternary salt in the organic phase.D. Mass Transport in LLPTCUsually, it is recognized that the rate-determining step is controlled by the chemicalreaction in the organic phase under LLPTC conditions. For a fast mass transfer rate <strong>of</strong>catalyst between the two phases, the influence <strong>of</strong> mass transfer on the reaction can beneglected. In the past, the reaction rate was assumed to be independent <strong>of</strong> agitation <strong>and</strong>the surface area <strong>of</strong> the interface beyond a minimum stirring rate ( 300 rpm). However,the reaction rates can increase with increased agitation in cases where the transfer rate <strong>of</strong>anion between both phases is slower than the organic reaction. The phenomenon <strong>of</strong> masstransfer <strong>of</strong> quaternary salt between the two phases has received little attention. The reactivity<strong>of</strong> the reaction by PTC is controlled by the rates <strong>of</strong> the organic <strong>and</strong> aqueous reactions,the partition equilibrium, <strong>and</strong> the mass transfer steps <strong>of</strong> the quaternary saltsbetween the organic <strong>and</strong> aqueous phases [27,28]. The partition equilibrium <strong>of</strong> quaternaryammonium salts was obtained in our previous work [85,86,92].Copyright © 2003 by Taylor & Francis Group, LLC


The mass transfer rates <strong>of</strong> catalysts between two phases are difficultly realized due tothe difficult identification <strong>of</strong> the active catalyst during the reaction [57,112–115]. Masstransfer coupled rapid reactions subjected to LLPTC have been studied extensively[58,63,69,115,116]. Mass transfer rates <strong>of</strong> catalysts in the reaction <strong>of</strong> 2,4,6-tribromophenol<strong>and</strong> tetra-n-butylammonium bromide in a solution <strong>of</strong> KOH were determined [57,114].Evans <strong>and</strong> Palmer [50] first consider theoretically the effect <strong>of</strong> diffusion <strong>and</strong> masstransfer in two well-mixed bulk phases <strong>of</strong> uniform composition separated by a uniformstagnant mass transfer layer at the interface. They studied the effect <strong>of</strong> the Damko¨ hlernumber, organic reaction equilibrium rate constant, reactant feed-rate ratio, flow rate <strong>of</strong>the organic phase, <strong>and</strong> the organic reaction reactivity on conversion. Chen et al. [53]derived algebraic expressions for the interphase flux <strong>of</strong> QY <strong>and</strong> QX. The reaction parameterswere estimated from experimental data using a two-stage method <strong>of</strong> optimal parameters.Naik <strong>and</strong> Doraiswamy [117] reported that future research should be directedtowards the use <strong>of</strong> a membrane module as a combination reactor <strong>and</strong> separator unitwith the membrane serving not merely to carry out the PT-catalyzed reaction, but alsosimultaneously <strong>and</strong> selectively to recover the organic product. Stanley <strong>and</strong> Quinn [118]reported the use <strong>of</strong> a membrane reactor for performing PT-catalytic reactions <strong>and</strong>included theoretical models <strong>and</strong> calculations to predict the kinetic behavior <strong>of</strong> the system.Matson [119] investigated the commercial feasibility <strong>of</strong> such membrane systems. However,the characterization <strong>of</strong> hydrodynamic phenomena in PT-catalyzed reactions has not beenattempted.Rushton et al. [120] developed a method for measuring the mass transfer coefficient.However, their method can only be used in systems with unity distribution ratio. Asai etal. [121] measured the liquid–liquid mass transfer coefficients in an agitated vessel with aflat interface. In their later work [122,123] on the alkaline hydrolysis <strong>of</strong> n-butyl acetate <strong>and</strong>oxidation <strong>of</strong> benzyl alcohol in an agitated vessel, the overall reaction rate <strong>of</strong> PTC withmass transfer at a flat interface was analyzed. The observed overall reaction rate wasconcluded to be proportional to the interfacial concentration <strong>of</strong> the actual reactant.Wang <strong>and</strong> Yang [57] investigated the dynamic behavior <strong>of</strong> PT-catalyzed reactions bydetermining the parameters accounting for mass transfer <strong>and</strong> the kinetics in a twophasesystem. The film theory was applied to interpret the behavior <strong>of</strong> PTC. The overallmass transfer coefficients <strong>of</strong> QX (or QY) from an agitated mixture <strong>of</strong> QX (or QY) werefirst calculated in known qualities <strong>of</strong> water <strong>and</strong> the organic solvent by using a simplecorrelation:"lnC QXþV C # QX1V!mQXC QX;i m QXV C QX;i V þ 1 ¼ K QX At ð56ÞThe overall mass transfer coefficient <strong>of</strong> QX was obtained by plotting the term on the lefth<strong>and</strong>side <strong>of</strong> Eq. (56) versus time. Yang et al. [124] developed a mathematical modelconcerning mass transfer in a single droplet to describe the dispersed phase system.They measured the distribution coefficient <strong>and</strong> the mass transfer coefficient <strong>of</strong> a PTcatalytic intermediate between two phases.Also, the diffusion boundary layer resistances on either side <strong>of</strong> the membrane filter inmembrane transport processes have been extensively examined [125,126]. Most <strong>of</strong> thesestudies deal with cases wherein solute diffuses across a membrane filter separating twoaqueous phases with different concentrations. However, the individual film mass transfercoefficients in both liquid phases are unavailable.Copyright © 2003 by Taylor & Francis Group, LLC


The mass transfer resistances strongly depend on the nature <strong>of</strong> the hydrodynamics inthe contacting device <strong>and</strong> the mode <strong>of</strong> operation. Many devices have been used to studytwo-phase mass transfer at or near the liquid–liquid interface. Hence, the hydrodynamiccharacteristics <strong>of</strong> ion transport through a membrane were presented to evaluate the feasibilitythat this permeation system can be calibrated as a st<strong>and</strong>ardized liquid–liquidsystem for studying the membrane-moderated PT-catalyzed reaction. The individualmass transfer coefficients <strong>and</strong> diffusivities for the aqueous phase, organic phase, <strong>and</strong>membrane phase were determined <strong>and</strong> then correlated in terms <strong>of</strong> the conventional Sh–Re–Sc relationship. The transfer time <strong>of</strong> quaternary salt across the membrane <strong>and</strong> thethickness <strong>of</strong> the hydrodynamic diffusion boundary layer are calculated <strong>and</strong> then the effect<strong>of</strong> environmental flow conditions on the rate <strong>of</strong> membrane permeation can be accuratelyinterpreted [127].The mass transfer <strong>of</strong> quaternary salt from the organic phase into the aqueous phasethrough a lipophilic membrane is indicated in Fig. 3. Assume that the solute activity in thislipophilic membrane is identical to that in the bulk organic solution, then the mass fluxvalues for the individual species are described byN ¼ k o ½QXŠ ½QXŠ i ð57ÞN ¼ k a ½QXŠ i ½QXŠ ð58ÞN ¼ k m ½QXŠ i ½QXŠ ið59ÞThe asterisk denotes the species in the organic phase <strong>and</strong> the membrane phase, respectively.The distribution coefficients <strong>of</strong> quaternary salts between membrane <strong>and</strong> aqueousphase or organic phase are defined asm ¼ ½QXŠ i½QXŠ ið60Þ<strong>and</strong>m ¼ ½QXŠ i½QXŠ ið61ÞFIG. 3Mass transfer <strong>of</strong> the catalyst between two phases <strong>and</strong> membrane.Copyright © 2003 by Taylor & Francis Group, LLC


According to Eqs. (57)–(61), <strong>and</strong> for organic volume V <strong>and</strong> interfacial area S, the rate <strong>of</strong>change <strong>of</strong> solute concentration can be expressed byV d½QXŠ¼ KS dt o m½QXŠ m½QXŠ ð62Þin whichK o ¼ m þ 1 þ m 1ð63Þk a k m k oAccording to the initial extractive concept that the content <strong>of</strong> quaternary salt is restrictedto less than 10% in the aqueous solution, the quaternary salt is completely dissociated, i.e.,[QX] approaches zero, <strong>and</strong> the magnitude <strong>of</strong> the distribution coefficient m is less than 10.By plotting ðV=SÞd½QXŠ=dt against ½QX], the overall mass transfer coefficient K o wasobtained by a least-squares regression. The regression factor is more than 0.99.If the extraction system is conducted in the absence <strong>of</strong> membrane, Eq. (63) is rewrittenasK o ¼m þ 1 1ð64Þk a k oThe values <strong>of</strong> diffusivities predicted for quaternary salts in the aqueous phase <strong>and</strong>the organic phase are in the following descending order: TBPB > TBAB > TBAI BTBAB <strong>and</strong> RBAB > TBAI > TBPB > BTBAB; respectively. The diffusivities <strong>of</strong> quaternarysalts increased with increasing temperature. The effects <strong>of</strong> solvents on diffusivitiesare ranked in the following descending order: CH 2 Cl 2 > C 6 H 5 CH 3 > CHCl 3 > C 6 H 6 >C 6 H 5 Cl > 1; 2-C 2 H 4 Cl 2 > H 2 O. The main influencing factor may be the viscosity <strong>of</strong> solvent.The overall mass transfer coefficients were determined by Lin [127]. The values <strong>of</strong> k o ,k a ,<strong>and</strong>k m were calculated by a numerical method for four types <strong>of</strong> quaternary salts inseven kinds <strong>of</strong> solvents. Assuming that the hydrodynamic characteristics <strong>of</strong> the diffusionboundary layer in the aqueous phase <strong>and</strong> the organic phase were similar in the presence orabsence <strong>of</strong> the membrane system if the agitation rate was kept below 100 rpm, the individualmass transfer coefficient <strong>of</strong> the membrane could then be calculated by subtractingEq. (64) from Eq. (63). The individual mass transfer coefficients increased with increasingagitation rates <strong>and</strong> temperatures. The sequence <strong>of</strong> mass transfer coefficient isk a k o > k m .Kiani et al. [125] <strong>and</strong> Prasad et al. [126] reported the following equation for theintrinsic mass transfer coefficient in the membrane, k m ¼ D"= m , where " <strong>and</strong> m are theporosity <strong>and</strong> thickness <strong>of</strong> the membrane, respectively, is the tortuosity factor <strong>of</strong> themembrane defined as the actual pore length divided by the membrane thickness, <strong>and</strong> Dis the diffusivity <strong>of</strong> species in the bulk liquid phase. The average tortuosities were calculated<strong>and</strong> found to reduce from 4.3 to 2.7 when the agitation rates increased from 90 to 600rpm. Because the individual mass transfer coefficient <strong>of</strong> a membrane is not a constant <strong>and</strong>increases with increasing agitation rate, the tortuosity decreases slightly with increasingagitation rate according to the equation <strong>of</strong> Kinai et al. [125].If the mixing is so vigorous that the diffusion boundary layer can be eliminated, Eq.(62) can be reduced toVSd½QXŠdt k m m½QXŠ m½QXŠka ;k o !1¼ ð65ÞCopyright © 2003 by Taylor & Francis Group, LLC


The extractive effectiveness factor , defined as the effect <strong>of</strong> the diffusion boundary layeron the extraction rate <strong>of</strong> quaternary salt can be characterized in terms <strong>of</strong> the ratio <strong>of</strong> Eq.(62) to Eq (65): ¼ K o m½QXŠ m½QXŠ k m m½QXŠ m½QXŠ ð66Þ¼ðBi o þ Bi a þ 1Þ 1in which Bi o ð¼ mk m =k o Þ <strong>and</strong> Bi a ð¼ mk m =k a Þ are Biot numbers for the organic phase <strong>and</strong>aqueous phase, respectively. Equation (66) represents the mass transfer ratio <strong>of</strong> conductionrate to convection rate <strong>of</strong> the quaternary salt at the interface. According to theexperimental data <strong>of</strong> Lin [127], the values <strong>of</strong> , Bi o ,<strong>and</strong>Bi a are calculated to be around0.96, 0.04, <strong>and</strong> 0.002, respectively, when the agitation rate is lower than 100 rpm. Hence, itclarified again that the membrane resistance at high agitation rates controls the masstransfer resistance <strong>of</strong> the membrane extraction.Usually, mass transfer coefficients can be correlated from the classical equation:Sh ¼ aRe b Sc cð67Þwhere Shð¼ k m d=DÞ is the Sherwood number; Re (¼ du=Þ is the Reynolds number, Sc(¼ =D) is the Schmidt number, D is the diffusivity in the bulk fluid, u is a characteristicvelocity <strong>of</strong> the fluid such as the mean fluid flow velocity, is the density, is the viscosity,<strong>and</strong> d is a characteristic dimension <strong>of</strong> the system.In Eq. (67), a is an experimental constant <strong>and</strong> c usually has a value <strong>of</strong> 1/3 [128–130].The value <strong>of</strong> b depends on the type <strong>of</strong> equipment <strong>and</strong> system, <strong>and</strong> most <strong>of</strong> the theoriespredict a one-half power on the Reynolds number [131]. The mass transfer from bulksolution to the surface <strong>of</strong> the membrane is mainly controlled by the turbulence <strong>of</strong> the fluidmotion created by stirring. The characteristic velocity is defined in terms <strong>of</strong> the stirringspeed ðu ¼ ndÞ. The values <strong>of</strong> a <strong>and</strong> b were determined from the intercept <strong>and</strong> slope <strong>of</strong> theline <strong>of</strong> Sh=Sc 1=3 against Re for the specified mass transfer coefficients <strong>of</strong> k a , k o , <strong>and</strong> K o .These parameters are different <strong>and</strong> are dependent on the system geometry <strong>and</strong> flow pattern.However, it can be concluded that the exponent value on Re varied from 0.2 to 1.0,depending on the design <strong>of</strong> the membrane permeation system.The correlating equation [67] established here can be used to evaluate the masstransfer coefficient <strong>and</strong> the thickness <strong>of</strong> the diffusion boundary layer, ð¼ d=shÞ. Thethickness <strong>of</strong> this layer calculated for an organic solvent <strong>and</strong> aqueous solution were10 3 –10 2 <strong>and</strong> 10 9 –10 7 cm, respectively, for the four types <strong>of</strong> quaternary salts studied.For a solute crossing a mass transfer resistance film, the transfer time can be approximatelyestimated by the following equation [131]:ðfilm thickness)2<strong>Transfer</strong> time ¼ð68Þdiffusion coefficientBased on the data presented here, the estimated transfer times for a solute crossing theorganic <strong>and</strong> aqueous mass transfer resistance film are about 1–10 <strong>and</strong> 10 11 –10 8 s, respectively.E. <strong>Interfacial</strong> Phenomena in LLPTCStarks [132] proposed that the transfer rate <strong>of</strong> an anion across the interface is largelygoverned by four factors: (1) interfacial area, (2) anion activity <strong>and</strong> hydration at theCopyright © 2003 by Taylor & Francis Group, LLC


interface, (3) bulkiness <strong>of</strong> the quaternary salt, <strong>and</strong> (4) sharpness <strong>of</strong> the interface. Starksindicated that three <strong>of</strong> the more important factors affecting the amount <strong>of</strong> interfacial areainclude interfacial tension, the presence <strong>of</strong> surfactants, <strong>and</strong> the degree <strong>of</strong> stirring or agitation.The interfacial area under steady-state stirring conditions will increase with decreasinginterfacial tension. The chemical natures <strong>of</strong> the organic <strong>and</strong> the aqueous phasesdetermine the interfacial tension that exists between these two phases. The quaternarysalt present in the reaction mixture may lower interfacial tension because <strong>of</strong> its surfactantproperties. Elegant ESCA studies [132,133] suggest that if the anion is highly hydrated itwill not be tightly bound at the interface with the quaternary cation, but rather tend to bemore dispersed in solution, removed from the interface.Reuben <strong>and</strong> Sjoberg [134] indicated that all boundaries are difficult to cross: political,legal, <strong>and</strong> geographical boundaries, <strong>and</strong> also phase boundaries in chemical systems.The interfacial mechanism is the most widely accepted mechanism for PTC reactions in thepresence <strong>of</strong> a base<strong>Interfacial</strong> tension is an important property in the process design <strong>of</strong> liquid–liquidprocesses. The decrement <strong>of</strong> interfacial tension between both phases leads to an increasedinterfacial area [135]. Because the volumetric rate <strong>of</strong> extraction was found to be dependenton the interfacial area, interfacial tension data are useful in underst<strong>and</strong>ing the effect <strong>of</strong>interfacial area on the volumetric rate <strong>of</strong> extraction <strong>and</strong> overall reaction rates for a PTcatalyzedreaction. Dutta <strong>and</strong> Patil [136] reported that the effect on the interfacial tension<strong>of</strong> the water/toluene system has been studied in the presence <strong>of</strong> four PT catalysts, i.e.,tricaprylmethyl ammonium chloride, hexadecyltrimethyl ammonium chloride, hexadecytrimethylammonium bromide, <strong>and</strong> hexadecyltributyl phosphonium bromide. Thedecrease in interfacial tension by surfactants increases the interfacial contact area, enhancingthe volumetric rate <strong>of</strong> extraction.Juang <strong>and</strong> Liu [74,75] presented that the interfacial tensions between water/n-hexane<strong>and</strong> water/toluene in the synthesis <strong>of</strong> ether–ester compounds by PTC could be measured.These two-phase systems contained PT catalyst, an aqueous phase reactant, <strong>and</strong>/or alkali.The interfacial data could be well described by the Gibbs adsorption equation coupledwith the Langmuir monolayer isotherm.III.LIQUID–SOLID–LIQUID PHASE TRANSFER CATALYSISLLPTC is the most widely synthesized method for solving the problem <strong>of</strong> the mutualinsolubility <strong>of</strong> nonpolar <strong>and</strong> ionic compounds [27–31]. Two compounds in immisciblephases are able to react because <strong>of</strong> the PT catalyst. However, processes using a twophasePT-catalytic reaction always encounter the separation problem <strong>of</strong> purifying thefinal product from the catalyst. Regen [137] first used a solid-phase catalyst [triphasecatalyst (TC) or polymer-support catalyst], in which a tertiary amine was immobilizedon a polymer support, in the reaction <strong>of</strong> an organic reactant <strong>and</strong> an aqueous reactant.From the industrial application point <strong>of</strong> view, the supported catalyst can be easily separatedfrom the final product <strong>and</strong> the unreacted reactants simply by filtration or centrifugation.In addition, either the plug flow reactor (PFR) or the continuous stirred tank reactor(CSTR) can be used to carry out the reaction. The most synthetic methods used for triphasecatalysis were studied by Regen <strong>and</strong> Beese [137–141] <strong>and</strong> Tomoi <strong>and</strong> coworkers [142–146].Another advantage <strong>of</strong> triphase catalysis is that it can be easily adapted to continuousprocesses [147–149]. Therefore, triphase catalysis possesses high potential in industrialscaleapplications for synthesizing organic chemicals from two immiscible reactants.Copyright © 2003 by Taylor & Francis Group, LLC


Quaternary onium salts, crown ethers, crypt<strong>and</strong>s, <strong>and</strong> polyethylene glycol have allbeen immobilized on various kinds <strong>of</strong> supports, including polymers (most commonlymethylstyrene-co-styrene resin cross-linked with divinylbenzene), alumina, silica gel,clays, <strong>and</strong> zeolites [137–156]. Because <strong>of</strong> diffusional limitations <strong>and</strong> high cost, the industrialapplications <strong>of</strong> immobilized catalysis (triphase catalysis) are not fully utilized. Thisunfortunate lack <strong>of</strong> technology for industrial scale-up <strong>of</strong> triphase catalysis is mainly due toa lack <strong>of</strong> underst<strong>and</strong>ing <strong>of</strong> the complex interactions between the three phases involved insuch a system. In addition to the support macrostructure, the support microenvironment isalso crucial in triphase catalysis since it determines the interactions <strong>of</strong> the aqueous <strong>and</strong> theorganic phases with the PT catalyst immobilized on the support surface [117]. However, todate, few papers have discussed the microenvironment. The effect <strong>of</strong> the internal molecularstructure <strong>of</strong> the polymer support, which plays an important role in the imbibed composition,on the reaction rate has seldom been discussed. In addition to the reactivity, for a TCin an organic <strong>and</strong> aqueous solution the volume swelling, imbibed different solvent ratio,amount <strong>of</strong> active site, <strong>and</strong> mechanical structure <strong>of</strong> the catalyst must be considered. Hence,these complex interactions in the microenvironment must be solved in order to obtain ahigh reactivity <strong>of</strong> TC.A. Characterization <strong>and</strong> <strong>Mechanism</strong> <strong>of</strong> LSLPTC1. <strong>Mechanism</strong> <strong>of</strong> LSLPTCIn general, the reaction mechanism <strong>of</strong> the fluid–solid reactions involves: (1) mass transfer<strong>of</strong> reactants from the bulk solution to the surface <strong>of</strong> the catalyst pellet, (2) diffusion <strong>of</strong>reactant to the interior <strong>of</strong> the catalyst pellet (active site) through pores, <strong>and</strong> (3) intrinsicreaction <strong>of</strong> reactant with active sites. Triphase catalysis is more complicated than traditionalheterogeneous catalysis, because it involves not merely diffusion <strong>of</strong> a single gaseousor liquid phase into the solid catalyst. Both organic reactant <strong>and</strong> aqueous reactant existwithin the pores <strong>of</strong> the polymer pellet. For step (3), a substitution reaction in the organicphase <strong>and</strong> an ion-exchange reaction in the aqueous phase occurred. Diffusion <strong>of</strong> both theaqueous <strong>and</strong> organic phases within the solid support is important <strong>and</strong> various mechanismshave been proposed for triphase catalysis. However, each mechanism can only explain asingle reaction system. Naik <strong>and</strong> Doraiswamy [117] discussed these mechanism in theirreview paper.Tundo <strong>and</strong> Venturello [155,157] proposed a mechanism for a TC system using silicagel as support to account for the active participation <strong>of</strong> the gel by adsorption <strong>of</strong> reagents.Telford et al. [158] suggested an alternation shell model that requires periodical changes inthe liquid phase filling the pores <strong>of</strong> the catalyst. Schlunt <strong>and</strong> Chau [150] from the sameresearch group tried to validate this model using a novel cyclic slurry reactor, <strong>and</strong> indicatedthat only the catalyst in a thin shell near the particle surface was utilized. Tomoi <strong>and</strong> Ford[142] <strong>and</strong> Hradil et al. [159] reported that a realistic mechanism involves the collision <strong>of</strong>droplets <strong>of</strong> the organic phase with solid catalyst particles dispersed in a continuous aqueousphase. Svec’s model [160] for transport <strong>of</strong> the organic reagent from the bulk phase throughwater to the catalyst particle has been developed in terms <strong>of</strong> emulsion polymerization.Because the triphase reaction involves not merely diffusion <strong>of</strong> a single phase into thesolid support, the organic reaction take places in the organic phase, <strong>and</strong> the ion-exchangereaction occurs in the aqueous phase. The catalyst support is usually lipophilic. Theorganic phase <strong>and</strong> aqueous phase fill the catalyst pores to form the continuous phase<strong>and</strong> the disperse phase, respectively. The interaction between quaternary salts as well asCopyright © 2003 by Taylor & Francis Group, LLC


the organic phase <strong>and</strong> aqueous phase play a crucial role in promoting the triphase reactionrate. However, this information is unclear.2. Characterization <strong>of</strong> LSLPTCPoly(styrene-co-chloromethylstyrene) crosslinked with divinylbenzene, which is immobilizedwith quaternary ammonium salts, was investigated for the synthesis <strong>of</strong> the finechemicals in our previous studies [161–166]. The microenvironment <strong>of</strong> the polymer supportplayed a crucial role in enhancing the reaction rate. More information about characterization<strong>of</strong> the polymer structure, the interaction between organic solvent, resin, <strong>and</strong>aqueous solution, <strong>and</strong> the reuse <strong>of</strong> the catalyst is required to encourage application.Wu <strong>and</strong> Lee [166] report that 24 kinds <strong>of</strong> ion-exchange resin were used to clarify thischaracter <strong>of</strong> the resin, including six kinds <strong>of</strong> commercial ion-exchange microresin, fivekinds <strong>of</strong> laboratory-produced macroresin, <strong>and</strong> 13 kinds <strong>of</strong> laboratory-produced microresin,using instrumental analysis by TGA, EA, <strong>and</strong> SEM-EDS, <strong>and</strong> the reaction method.The densities <strong>of</strong> active sites in the resin, titrated using the Volhard method for commercialanion exchangers, were higher than those for laboratory-produced resins.ð69ÞScanning electron microscopy (SEM) analyzes electrons that are scattered from thesample’s surface, <strong>and</strong> monitors the morphological observation <strong>of</strong> the polymer resin. Theelemental analysis (EA) is effected by means <strong>of</strong> energy-dispersive X-ray spectrometer(EDS) methods. The chloride density was shown to be well distributed on the resin surfaceby X-ray images <strong>of</strong> chloride. It was also demonstrated that the active sites (-NR 4 Cl) in theresin were completely dispersed. Some other chemical compounds used for synthesizingthe polymer resin were also detected. Although the pretreatment <strong>of</strong> the resin was conductedby washing with water, NaOH solution, <strong>and</strong> acetone, the salts (Al, Si, <strong>and</strong> Ca) usedas reactants in the suspension method were slightly retained in the resin.The immobilized content <strong>of</strong> tri-n-butylamine in the resin was determined by theTGA, EA, <strong>and</strong> Volhard methods. The polymer backbone formed in a one-stage processwhere the decomposing temperature range was 300 –450 C. The immobilized resin (mi4-20) was formed in a two-stage process, where the ranges <strong>of</strong> decomposing temperature forthe two stages were 160 –200 C <strong>and</strong> 350 –450 C. Although it is tempting to divide the twostages into two distinctive units, the correlation between quaternary salt content <strong>and</strong>weight loss in the first was qualitative. The weight loss in the first step is equal to theimmobilized amount <strong>of</strong> the functional group <strong>of</strong> -NðC 4 H 9 Þ 3 . The accuracy <strong>of</strong> the analyticaltechnique was within 10%. The commercial ion-exchange resins were revealed in a threestageprocess. The decomposed compound <strong>and</strong> temperature for each decomposition stepare: imbibed water ( 100 C), functional group (160 –300 C), <strong>and</strong> polymer backbone(350 –450 C). The sequence <strong>of</strong> the imbibed capability <strong>of</strong> water is: IRA-900 ð20%Þ > A-Copyright © 2003 by Taylor & Francis Group, LLC


26 > Dowex 1 2 > A-27 IRA-410 > IRA-904 > mi4-20 (4%). Most commercial ionexchangeresins are <strong>of</strong> the hydrophilic functional group type.In addition, the immobilized amount <strong>of</strong> the functional group <strong>of</strong> -NðC 4 H 9 Þ 3 in theresin was determined from the mass fraction <strong>of</strong> nitrogen by EA for C, H, <strong>and</strong> N, <strong>and</strong> fromthe chloride ion density titrated by the Volhard method. The sequence <strong>of</strong> determiningmethod for the immobilized content <strong>of</strong> tri-n-butylamine in the resin wasTGA > EA > Volhard. The analyzed result <strong>of</strong> the TGA (or EA) method was based onthe elemental weight, <strong>and</strong> it revealed the real immobilized content. However, the analyzedresult <strong>of</strong> the Volhard method determined the free chloride ion in the solution by theAgNO 3 titration method. The immobilized content <strong>of</strong> tri-n-butylamine in the resin bythe TGA (or EA) method was > 20% larger than that determined by the Volhard method.The immobilized content <strong>of</strong> tri-n-butylamine in the resin by the TGA (or EA) method wasindependent <strong>of</strong> the number <strong>of</strong> cross-linkages, <strong>and</strong> only dependent <strong>of</strong> the number <strong>of</strong> thering substitution.These experimental results demonstrate that tri-n-butylamine could be immobilizedcompletely with the active site on the resin for an immobilizion duration <strong>of</strong> 6 days.However, the immobilized content <strong>of</strong> tri-n-butylamine by the Volhard method was dependenton both the number <strong>of</strong> cross-linkages <strong>and</strong> the number <strong>of</strong> ring substitutions. Theimmobilized contents for the Volhard method are about 50–70% that for TGA (orEA). Since the analyzed results <strong>of</strong> the Volhard method determined the free chlorideions in the solution by the AgNO 3 titration method, the free chloride ion <strong>of</strong> the activesite were only measured at 50–70% <strong>of</strong> the amount <strong>of</strong> immobilized content. The trend <strong>of</strong>the varied content for microresin is larger than that for macroresin. This result indicatesthat the analysis by the Volhard method may be influenced by the diffusion problem, <strong>and</strong>may be because the resin did not swell completely in the aqueous solution. On the otherh<strong>and</strong>, if the resin is used as a TC to react in an actual reaction system, <strong>and</strong> the resin couldnot swell completely to release all free chloride ions, then the reaction environment wouldbe influenced by the mass transfer <strong>of</strong> the reactant.As indicated by Ohtani et al. [32] both organic reactant <strong>and</strong> aqueous reactant existwithin the pores <strong>of</strong> the polymer pellet. The HLB <strong>of</strong> the support structure determines thedistribution <strong>of</strong> the two phases within the catalyst support [167,168]. Therefore, the distribution<strong>of</strong> the organic reactant <strong>and</strong> aqueous reactant within the pores <strong>of</strong> the polymerpellet will directly influence the reaction. The swollen capability <strong>of</strong> the resin is used toestimate the validity <strong>of</strong> the resin. The effect factor <strong>of</strong> the swollen capability <strong>of</strong> the resinincludes the cross-linkage, the number <strong>of</strong> ring substitutions (total exchange capability), theelectronic charge <strong>and</strong> diameter <strong>of</strong> the counterion, the polarity <strong>of</strong> the organic solvent, thecomposition <strong>of</strong> the functional group, the chemical bonding type between both exchangeions, <strong>and</strong> the electrolyte concentration in the aqueous solution.Wu <strong>and</strong> Lee [166] <strong>and</strong> Tang [169] reported the amount <strong>of</strong> imbibed solvent, volumeratio, <strong>and</strong> porosity <strong>of</strong> 12 kinds <strong>of</strong> ion-exchange resin for seven kinds <strong>of</strong> solvents (dichloromethane,chlor<strong>of</strong>orm, 1,2-dichloroethane, benzene, toluene, chlorobenzene, <strong>and</strong> water)when 1 g <strong>of</strong> the resin was placed in 25 mL <strong>of</strong> the pure solvent. The experimental results forthe commercial ion-exchange resin were as follows:1. The amounts <strong>of</strong> the imbibed solvent for the aromatic solvents (benzene,toluene, <strong>and</strong> chlorobenzene) were larger than those for halide aliphatic solvents(dichloromethane, chlor<strong>of</strong>orm, <strong>and</strong> 1,2-dichloroethane) since the resin was <strong>of</strong>the styrene type; the sequence <strong>of</strong> the imbibed amount for the aromatic solventswas benzene > toluene > chlorobenzene.Copyright © 2003 by Taylor & Francis Group, LLC


2. The imbibed amounts for water <strong>and</strong> organic solvent were around 1 g <strong>and</strong> < 1g,respectively.3. The volume ratios were mostly located between 1 <strong>and</strong> 2.4. The porosities were located between 0.5 <strong>and</strong> 0; the porosity <strong>of</strong> most ionexchangeresins was about 0.5 [170].The experimental results for the laboratory-produced resins were as follows:1. The amounts <strong>of</strong> the imbibed solvent were different, depending on the structure<strong>of</strong> the resins, <strong>and</strong> the amount for water was less than that for organic solvent.2. The amounts <strong>of</strong> imbibed solvent were in the range 0–3g.3. The volume ratios were almost all located between 1 <strong>and</strong> 3, <strong>and</strong> decreased withincreasing cross-linkage <strong>of</strong> the resin.4. The porosities were located between 0.25 <strong>and</strong> 0.75.The porosity <strong>and</strong> imbibed amount decreased for the solvents benzene, toluene, <strong>and</strong>chlorobenzene, <strong>and</strong> increased for the solvents chlor<strong>of</strong>orm, 1,2-dichloroethane, dichloromethane,<strong>and</strong> water, with increasing number <strong>of</strong> ring substitutions. These results indicatethat the solubility <strong>of</strong> water in chlor<strong>of</strong>orm, 1,2-dichloroethane, <strong>and</strong> dichloromethane isgreater than the solubility in benzene, toluene, <strong>and</strong> chlorobenzene. The imbibed amountfor aromatic solvents was larger than that for halide aliphatic solvents when the number <strong>of</strong>ring substitutions was small, <strong>and</strong> the trend was opposite when the number <strong>of</strong> the ringsubstitutions was large.Because the functional group <strong>of</strong> the laboratory-produced resin (tetrabutylammoniumchloride) is more lipophilic than that <strong>of</strong> the commercial ion-exchange resin [tetramethyl-(or ethyl-) ammonium chloride], the amount <strong>of</strong> imbibed water was larger than that<strong>of</strong> the organic solvent for commercial resin; on the other h<strong>and</strong>, for laboratory-producedresin, the amount <strong>of</strong> water was less than that <strong>of</strong> organic solvent. The imbibed amount <strong>of</strong>organic solvent for laboratory-produced resin was larger than that for commercial resin.Since the swollen A-27 <strong>and</strong> IRA-904 was high in the commercial resin in this study, theothers (IRA-900, A26, IRA410, Dowex IX2) were not good for swelling. Hence, they areimproperly used as PT catalysts in an organic phase/aqueous solution reaction system.Tang [169] reported the amount <strong>of</strong> the imbibed solvent for commercial resin <strong>and</strong>laboratory-produced resin in an organic solvent <strong>and</strong> in an aqueous solution in the presence<strong>of</strong> KOH, NaOH, KCl, <strong>and</strong> NaCl. Four kinds <strong>of</strong> salts were used to investigate the swellingphenomenon since the KOH <strong>and</strong> NaOH were usually used as reactants <strong>and</strong> the chlorideion was a byproduct in the PTC reaction. Chlorobenzene was chosen as solvent because <strong>of</strong>its high boiling point. The imbibed amounts <strong>of</strong> chlorobenzene <strong>and</strong> water increased for thecommercial resin, <strong>and</strong> decreased for the laboratory-produced resin when the salt wasadded. The imbibed amounts <strong>of</strong> chlorobenzene <strong>and</strong> water for NaOH were less than thatfor KOH, <strong>and</strong> that for NaCl was also less than that for KCl since the diameter <strong>of</strong> the aquaion for Na is larger than that for K. The aqua interaction between metal <strong>and</strong> waterincreased to increase the swelling capability <strong>of</strong> the resin when the diameter <strong>of</strong> the aquaion increased. Also, the imbibed amounts <strong>of</strong> chlorobenzene <strong>and</strong> water for KCl were lessthan that for KOH, <strong>and</strong> that for NaCl was also less than that for NaOH since the diameter<strong>of</strong> the aqua ion for Cl is larger than that for OH. The imbibed amounts <strong>of</strong> chlorobenzene<strong>and</strong> water for microresin was larger than that for macroresin.In general, the reaction rate increases with augmentation <strong>of</strong> the polarity <strong>of</strong> thesolvent. The apparent reaction-rate constant increased with a rise in temperature. Thesequence <strong>of</strong> the reactivity for macroresin was CH 2 Cl 2 > CHCl 3 > C 6 H 5 Cl > C 6 H 6 >Copyright © 2003 by Taylor & Francis Group, LLC


C 6 H 6 > C 6 H 5 CH 3 <strong>and</strong> the trend for microresin was similar, except for chlor<strong>of</strong>orm. Theincrement <strong>of</strong> the reactivity <strong>of</strong> this triphase reaction corresponds to the polarity <strong>of</strong> thesolvent. Dichloromethane has the highest reactivity among the solvents. However, theboiling point <strong>of</strong> dichloromethane is 39 C, <strong>and</strong> it is unsuitable for reaction in the highertemperature system. The activity energy varied with the structure <strong>of</strong> the resin. Mostactivity energy levels for microresins were greater than those for macroresins, except forchlor<strong>of</strong>orm.The four functions <strong>of</strong> a base in a liquid–solid–liquid triphase catalytic reaction werereported [165]: (1) reactant; (2) deprotonation <strong>of</strong> acidic organic compound to become thereactive form; (3) improving the reactive environment in the catalytic pellets, such asswelling volume, imbibed composition, solubility between two phases, etc; <strong>and</strong> (4) reducingthe solvation <strong>of</strong> catalyst <strong>and</strong> water to upgrade the reactivity <strong>of</strong> active catalyst in theorganic phase. Wu <strong>and</strong> Lee [166] showed the effect <strong>of</strong> base concentration for the reactivity<strong>of</strong> 4-methoxyphenylacetic acid. The apparent reaction-rate constant was maintainedalmost constant when the NaOH (or KOH) concentration was greater than 1 kmol=m 3 .The increment <strong>of</strong> the deprotonation <strong>of</strong> 4-methoxyphenylacetic acid dramatically increasedthe reactivity <strong>of</strong> the reaction when the base concentration was below 1 kmol=m 3 . When thesalt concentration was increased to change the reactivity environment, the reactivity <strong>of</strong> thereaction was slightly increased with increasing base concentration. The reactivity for KOHwas greater than that for NaOH. The result corresponded with the imbibed composition<strong>of</strong> the resin.The advantage <strong>of</strong> using a triphase catalytic reaction is that it easily recovers thecatalyst <strong>and</strong> purifies the product <strong>and</strong> reactant. Hence, the reuse, stability, <strong>and</strong> degradation<strong>of</strong> the catalyst must always be considered. Resins with onium groups may be used forextended periods or repeated cycles only if the catalyzed reactions occur under sufficientlymild conditions to avoid degradation. The degradation <strong>of</strong> the triphase catalyst may havethree factors: high temperature, strong base, <strong>and</strong> mechanical degradation. In the pastliterature [143,147,148,166,171–173], the reactivity <strong>of</strong> the triphase reaction was slightlyinfluenced by the degradation <strong>of</strong> the catalyst (polymer-supported resin). The active sitewas seen to decrease slightly with increasing base concentration up to 9 kmol=m 3 . Thenumber <strong>of</strong> active sites remained constant up to 60 C <strong>and</strong> then decreased dramatically asthe temperature increased. The degradation <strong>of</strong> the catalyst with temperature is moresensible than that for base concentration.In addition to the diffusion resistance <strong>of</strong> reactants affected by the particle size, it isalso influenced by the characterization <strong>of</strong> the polymer pellet, i.e., the degree <strong>of</strong> crosslinkage.In principle, the cross-linkage is related to the covalent bonds between two ormore linear polymer chains. For this reason, the degree <strong>of</strong> cross-linkage <strong>of</strong> the polymerwill affect the pore size <strong>and</strong> the amount <strong>of</strong> swelling [142,161]. The structure <strong>of</strong> the polymeris compact for a higher degree <strong>of</strong> cross-linkage. The pore size <strong>of</strong> the pellet is increasedwhen a polymer with a low degree <strong>of</strong> cross-linkage is swollen in an organic solvent. Thus, alower degree <strong>of</strong> swelling for a higher cross-linking polymer in an organic solvent is adisadvantage, i.e., a large diffusion resistance is obtained for using a higher degree <strong>of</strong>cross-linking <strong>of</strong> the polymer. Hence, the application <strong>of</strong> a highly cross-linked polymer islimited because <strong>of</strong> low reactivity in the triphase catalytic reaction.B. <strong>Kinetics</strong> <strong>and</strong> Modeling in LSLPTCThe reaction <strong>of</strong> triphase catalysis is carried out in a three-phase liquid (organic)–solid(catalyst)–liquid (aqueous) medium. In general, the reaction mechanism <strong>of</strong> the triphaseCopyright © 2003 by Taylor & Francis Group, LLC


catalysis is: (1) mass transfer <strong>of</strong> reactants from the bulk solution to the surface <strong>of</strong> thecatalyst pellet, (2) diffusion <strong>of</strong> reactants to the interior <strong>of</strong> the catalyst pellet (active sites)through pores, <strong>and</strong> (3) surface or intrinsic reaction <strong>of</strong> reactants with active sites. For step(3), the substitution reaction in the organic phase <strong>and</strong> ion-exchange reaction in the aqueousphase occurrs.r RX0¼R c V 03V cal k IRXþ V 0kM c c¼ 1 C RX0 ¼ k obs C RX0Wang <strong>and</strong> Yang [174–176] have proposed a general dynamic model for triphasecatalysis in a batch reactor. The mass transfer <strong>of</strong> reactants in the bulk aqueous <strong>and</strong> organicphases, diffusion <strong>of</strong> reactants within the pores <strong>of</strong> the solid catalyst particle, <strong>and</strong> intrinsicreactivities <strong>of</strong> the ion-exchange <strong>and</strong> organic reactions at the active sites within the solidcatalyst were investigated. Desikan <strong>and</strong> Doraiswamy [151] account for the effect <strong>of</strong> thereversibility <strong>of</strong> the ion-exchange reaction. The concentration <strong>of</strong> the catalytic active siteswithin the catalyst is given as@q QX¼ k@t 1 C Y q QX þ k 1 C X q QY þ k 2 C RX q QY ð71ÞMass balances <strong>of</strong> organic substrate <strong>and</strong> inorganic species within the catalyst are written as <strong>and</strong>" @ C RX@t" @C Y@t¼ D RXr 2¼ D Yr 2@ @ C RX@r r2 @r @ @C Y@r r2 @r s k 2C RX q QY s k 1 C Y q RXk 1 q QY C Xrespectively.In a heterogeneous catalytic reaction, the intraparticle effectiveness, c , for a firstorderreaction within a spherical catalyst at steady state is [177] c ¼ 3cothð3Þ 13 2where is the Thiele modulus:sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ R c k 2 M c c3 V cat D eAn apparent overall effectiveness factor <strong>of</strong> the catalyst is obtained by applying the pseudosteady-stateassumption to the mass balance equations within the catalyst, as" 0 ¼ 3 # app coth app 1 2 app 1 þ app coth app 1 ð76Þ=Bi mð70Þð72Þð73Þð74Þð75Þwhere app is the apparent Thiele modulus, <strong>and</strong> Bi m is the Biot number.231=2R c ðk 2 c q 0 =D RX Þ 0:5app6 1 þ D 745QYk 2D RX k 1ð77ÞCopyright © 2003 by Taylor & Francis Group, LLC


Many experimental studies on three-phase catalytic reactions indicated that the reactionrates for the organic phase <strong>and</strong> the aqueous phase follow pseudo-first-order kinetics[65,69,164,165]. The different type <strong>of</strong> the reaction expression can be written asd C RXdt¼ k appM resinV TC RX ¼ k obsC RX ¼ k 0 a V CVC RXð78Þ<strong>and</strong>dC MYdtM¼ k resinapp CV MY ¼ k obs C MY ¼ k 0 a V CTV C MYð79Þwhere k app <strong>and</strong> k app [cm 3 ðmin g resinÞ 1 ] represent the apparent reaction rate constant inthe organic phase <strong>and</strong> aqueous phase, respectively.C. Mass <strong>Transfer</strong> Problem in LSLPTCThe reactivity <strong>of</strong> a liquid–solid–liquid triphase reaction (i.e., polymer-supported catalyticreaction) is influenced by the structure <strong>of</strong> the active sites, particle size, degree <strong>of</strong> crosslinkage,degree <strong>of</strong> ring substitution, swollen volume, <strong>and</strong> spacer chain <strong>of</strong> a catalyst pellet.In the past, the characteristics <strong>of</strong> a triphase reaction, subjected to the mass transferlimitation <strong>of</strong> the reactants <strong>and</strong> ion-exchange rate in the aqueous phase, have been discussed[146,158,162,178,179]. The ion-exchange rate in the aqueous phase affects thereactivity <strong>of</strong> the triphase reaction.Past efforts have carried out this investigation macroscopically. The planar phaseboundary in a classical two-phase system cannot be described for the triphase system.Telford et al. [158] suggested an alternating shell model that requires periodical changes inthe liquid phase filling the pores <strong>of</strong> the catalyst. Schlunt <strong>and</strong> Chau [150] indicated that thereaction occurred in a thin shell near the particle surface. Tomoi <strong>and</strong> Ford [142] <strong>and</strong>Hradil et al. [159] proposed that the droplet <strong>of</strong> organic (or aqueous) phase collided withthe solid catalyst. However, the mechanism <strong>and</strong> effects <strong>of</strong> the internal molecular structure<strong>of</strong> the polymer support with the reaction are seldom discussed. Although some rules werelisted in the text <strong>and</strong> clarified by the experimental results [27,28], the relationship betweenthe reaction mechanism <strong>and</strong> polymer resin in a liquid–solid–liquid triphase reaction hasnot been understood completely. Hence, this study aims to discuss the mechanism <strong>of</strong> apolymer-supported triphase reaction.Among the vast scope <strong>of</strong> PTC application [27,28], approximately 40% <strong>of</strong> PTCpatents involve the hydroxide ion <strong>and</strong> it has been estimated that approximately 60% <strong>of</strong>commercial PTC applications involve the hydroxide ion [28]. Many papers[61,76,96,116,164,165] have proposed that the reactivity <strong>of</strong> a reactant in an organic reactionis influenced by the base concentration. The base concentration plays a crucial role ina PT-catalyzed reaction. However, the base effect for the reactivity <strong>of</strong> reactant in a triphasereaction was rarely paid attention to.Most PTC reactions are carried out on an industrial scale in the batch mode inmixer–settler arrangements. In view <strong>of</strong> the reactor design in the liquid–solid–liquid PTcatalyzedreaction, Ragaini <strong>and</strong> coworkers [147–149] reported the use <strong>of</strong> fixed-bed reactorswith a recycling pump or with a recycling pump <strong>and</strong> an ultrasonic mixer, <strong>and</strong> emphasizedthe importance <strong>of</strong> effluent recycle concept. Schlunt <strong>and</strong> Chau [150] reported the use <strong>of</strong> acyclic slurry reactor, which allowed the immiscible reactants to contact the catalyst sites incontrolled sequential steps. However, for triphase reactions in liquid–liquid systems whereCopyright © 2003 by Taylor & Francis Group, LLC


the catalyst is asolid phase, which reactor type should be properly used in this reactionsystem is not clear.The substitution reaction <strong>of</strong> (NPCl 2 Þ 3 with phenol is asequential reaction [166,169].The reaction type is different from the common one-stage reaction. The experimentalresults can easily demonstrate the relationship between the reaction kinetic limitation<strong>and</strong> the particle diffusion limitation. In atriphase reaction, the overall kinetic cycle canbe broken up into two steps by virtue <strong>of</strong> the presence <strong>of</strong> two practically insoluble liquidphases:achemicalconversionstepinwhichtheactivecatalystsites(Resin þ withphenolateions) react with the hexachlorocyclotriphosphazene in the organic solvent, <strong>and</strong> an ionexchangestep in which the attached catalyst sites are in contact with the aqueous phase:ð80ÞThe function <strong>of</strong> base in aliquid–solid–liquid triphasic reaction has four roles asmentioned above. In previous studies, it was observed that the reactivity <strong>of</strong> organic reactantvaried with the concentration <strong>of</strong> the base concentration in aliquid–liquid PT-catalyzedreaction [96,116,180]. In the liquid–solid–liquid triphase reaction, the effect <strong>of</strong> baseonthe reactivity <strong>of</strong> reactant (or reactive environment) was rarely paid attention to the k app<strong>and</strong> k app values dramatically increased with increasing concentration <strong>of</strong> NaOH when theratio <strong>of</strong> NaOH to C 6 H 5 OHwas in the range 1–1.5. This trend corresponds to that <strong>of</strong>liquid–liquid phase-transfer catalysis [165]. The apparent activity energies for an organicreaction decreased, <strong>and</strong> for an aqueous reaction increased when the NaOH concentrationwas increased. The reactive behavior <strong>of</strong> the organic reaction changed from reaction chemicalcontrol to diffusion control E a


FIG. 4 Yields <strong>of</strong> products <strong>and</strong> conversion <strong>of</strong> reactant as a function <strong>of</strong> reactant C 6 H 5 OH=ðNPCl 2 Þ 3consumption ratio at different NaOH concentrations: (abcd) 0.5 kmol=m 3 , (efgh) 0.9 kmol=m 3 , (ijkl)1.8 kmol=m 3 ;(*) (NPCl 2 Þ 3 ,(*)N 3 P 3 Cl 5 ðOC 6 H 5 Þ 1 ,(!)N 3 P 3 Cl 4 ðOC 6 H 5 Þ 2 ,(&)N 3 P 3 Cl 3 ðOC 6 H 5 Þ 3 ,(^) N 3 P 3 Cl 2 ðOC 6 H 5 Þ 4 ; ðÞ N 3 P 3 CLðOC 6 H 5 Þ 5 .reactant in the particle. In Fig. 4, the maximum yield <strong>of</strong> monophenolated product shifts tothe right by more than 0.2 unit, <strong>and</strong> the maximum yield <strong>of</strong> diphenolated product shifts tothe right by more than 0.1 unit. This reveals that the effect <strong>of</strong> intraparticle diffusion on theorganic reaction influences the reaction rate. This trend <strong>of</strong> shifting to the right <strong>of</strong> themaximum yield was increased with increasing concentration <strong>of</strong> NaOH.The reactivity <strong>of</strong> a triphase reaction is influenced by the structure <strong>of</strong> the active sites,particle size, degree <strong>of</strong> cross-linkage, degree <strong>of</strong> ring substitution, swollen volume, <strong>and</strong>spacer chain <strong>of</strong> a catalyst pellet. All these make the triphase reaction a complicatedone. Past efforts have carried out this investigation macroscopically. However, themechanism <strong>and</strong> effects <strong>of</strong> the internal molecular structure <strong>of</strong> the polymer support haveseldom been discussed.According to the steric effect <strong>of</strong> phenolate ion reacting with hexachlorocyclotriphosphazene<strong>and</strong> the reports <strong>of</strong> Wu <strong>and</strong> Meng two-phase catalysis [69]; triphase catalysis[165]), the maximum yield <strong>of</strong> partially substituted phenolated product was increasedwith increasing degree <strong>of</strong> substitution reaction. Figure 4 shows that the maximum yield<strong>of</strong> monophenolated product was larger than that <strong>of</strong> the diphenolated product, <strong>and</strong> themaximum yield <strong>of</strong> partially phenolated product decreased when the NaOH concentrationincreased (i.e., reactivity <strong>of</strong> the active site increased). This result reveals that the reactionrate <strong>of</strong> phenolate reacting with monophenolated (or diphenolated) product was greaterthan the diffusion rate <strong>of</strong> monophenolated (or diphenolated) product from active site tobulk solution <strong>and</strong> hexachlorocyclotriphosphazene from bulk solution to active site. Mostmonophenolated (or diphenolated) product reacted in situ with Resin þ OC 6 H 5 in theCopyright © 2003 by Taylor & Francis Group, LLC


neighborhood <strong>of</strong> the active site. Meanwhile, when the reactivity <strong>of</strong> Resin þ OC 6 H 5 increasedas the NaOH concentration increased, the diffusion resistance <strong>of</strong> reactants was obvious.In the present reaction, the overall reaction includes organic substitution <strong>and</strong> anaqueous ion-exchange reaction, Eq. (80). Two rate-controlling steps influence the reactionrate simultaneously. The reaction is complicated. Hence, from the literature[142,146,158,164,165,181,182], four special relationships are established between the ionexchangereaction <strong>and</strong> organic reaction with increasing concentration <strong>of</strong> organic reactant(NPCl 2 Þ 3 , according to Eqs (78) <strong>and</strong> (79), <strong>and</strong> these are listed below:1. It is assumed that the ion-exchange rate in the aqueous phase is much higherthan the substitution reaction rate in the organic phase. The effect <strong>of</strong> the ionexchangereaction could be eliminated from the reaction controlling steps.a. The intrinsic organic reaction is the rate-controlling step. Hence, the concentration<strong>of</strong> the active-site <strong>of</strong> triphase catalyst Resin þ C 6 H 5 O remainsconstant. The value <strong>of</strong> k app is constant. The value <strong>of</strong> k app increases withhigher concentration <strong>of</strong> (NPCl 2 Þ 3 due to an increase in the consumptionrate <strong>of</strong> phenolate ion. Similar results were obtained by Wu <strong>and</strong> Tang [164].b. The organic reaction rate is limited by both reaction kinetics <strong>and</strong> particlediffusion. The values <strong>of</strong> k app increase, <strong>and</strong> the values <strong>of</strong> k app decrease withincreasing concentration <strong>of</strong> (NPCl 2 Þ 3 [181].c. The organic reaction rate is only limited by film diffusion <strong>of</strong> reactant fromthe bulk organic solution to the surface <strong>of</strong> the catalyst pellet; this is the ratecontrollingstep. The values <strong>of</strong> kapp are constant, <strong>and</strong> the values <strong>of</strong> k appdramatically increase with increasing concentration <strong>of</strong> (NPCl 2 Þ 3 .Similarresults were obtained by Tomoi <strong>and</strong> Ford [142].2. It is assumed that the organic reaction rate in the organic phase is much higherthan the substitution reaction rate in the organic phase. Controlling the reactioncould eliminate the effect <strong>of</strong> the organic reaction.a. If the film diffusion <strong>of</strong> ion from the bulk aqueous solution to the surface <strong>of</strong>the catalyst pellet is the rate-controlling step, the value <strong>of</strong> k app remainsconstant because the initial concentration <strong>of</strong> phenolate ion is kept constant,<strong>and</strong> the value <strong>of</strong> k app dramatically decreases [181].b. If the ion exchange rate is limited by both particle diffusion <strong>and</strong> filmdiffusion, the value <strong>of</strong> k app decreases with increasing (NPCl 2 Þ 3 concentration,<strong>and</strong> the value <strong>of</strong> k app dramatically decreases. With low-percentage ringsubstitution, the ion-exchange process is the rate-limiting step[146,153,158,182].c. If the ion-exchange rate is limited by the intrinsic ion-exchange rate, thevalue <strong>of</strong> k app remains constant with increasing (NPCl 2 Þ 3 concentration,<strong>and</strong> the value <strong>of</strong> k app dramatically decreases.3. If the organic reaction rate is limited by both reaction kinetics <strong>and</strong> particlediffusion, <strong>and</strong> the ion-exchange rate is also limited by film (or particle) diffusion,the value <strong>of</strong> k app decreases, <strong>and</strong> the value <strong>of</strong> k app also decreases [165].4. When the organic reaction rate competes with the ion-exchange rate, the values<strong>of</strong> k app <strong>and</strong> k app remain almost constant with increasing concentration <strong>of</strong>(NPCl 2 Þ 3 .If mass transfer resistance influences the reaction, the concentration <strong>of</strong> the activecatalyst cannot remain constant during the course <strong>of</strong> the reaction. Also, when the concentration<strong>of</strong> organic reactant decreases, both the reaction rate <strong>and</strong> the effect <strong>of</strong> massCopyright © 2003 by Taylor & Francis Group, LLC


transfer <strong>of</strong> organic or aqueous reactant between solid<strong>and</strong> liquid phase decrease. However,theapparent first-order reaction-rate constant isincreasedby decreasingtheconcentration<strong>of</strong> organic reactant [165,181].Elemental analysis is studied by means <strong>of</strong> energy-dispersive X-ray spectrometer(EDS) methods. Ahigh Cl peak was detected due to the active site. Some chemicalcompounds (Si, Ca) added in the procedure <strong>of</strong> synthesizing the polymer resin were alsodetected. Although the pretreatment <strong>of</strong> the resin was conducted by washing with water,NaOH solution, <strong>and</strong> acetone, the salts (Si, Ca) used as reaction agents by the suspensionmethod were slightly retained in the resin. Alow Cl peak was detected due to the activesite. The peak height for the Cl atom was decreased <strong>and</strong> was increased for the Oatombetween, before, <strong>and</strong> after the reaction. This finding demonstrates that the phenoxide ionexchanged the chloride ion as counterion on the polymer-supported catalyst during thecourse <strong>of</strong> the reaction, <strong>and</strong> did not, however, occupy all the active sites in the catalyst.Hence, the result reveals that the mass transfer resistance <strong>of</strong> the ion-exchange step influencedthe concentration <strong>of</strong> anion on the active site.The volume <strong>and</strong> wet porosity <strong>of</strong> catalyst was increased about three times when thecatalyst imbibed the organic solvent <strong>and</strong> water. Different catalyst interacts differentlywith the organic phase <strong>and</strong> aqueous phase. Wu <strong>and</strong> Lee [166] indicated that the imbibedamount <strong>of</strong> organic solvent was larger than that <strong>of</strong> water because the catalyst supportwas lipophilic. The imbibed amount <strong>of</strong> water was dependent <strong>of</strong> the amount <strong>of</strong> ammoniumcation (i.e., active site). Hence, the imbibed amount <strong>of</strong> water increases withincreasing number <strong>of</strong> ring substitutions. If the structure <strong>of</strong> the resin is rigid (higherdegree <strong>of</strong> cross-linkage) or <strong>of</strong> larger particle size, the organic <strong>and</strong> aqueous phasesremains quiescent in the interior <strong>of</strong> the resin. The organic <strong>and</strong> aqueous reactants shouldnot diffuse simultaneously to the active site. The reaction occurs at a shell near thesurface <strong>of</strong> the resin. When the degree <strong>of</strong> cross-linkage <strong>of</strong> the resin is low, the structure <strong>of</strong>the resin is not solid. The flow rate <strong>of</strong> the organic <strong>and</strong> aqueous solutions in the interior<strong>of</strong> the resin increases with increasing agitation rate. The number <strong>of</strong> the effective activesites in the resin is increased.Wu<strong>and</strong>Lee[166]indicatedthatthefree chlorideionsontheactivesite(measured byVolhard analysis) were at only 50–70% <strong>of</strong> the amount <strong>of</strong> immobilized content (measuredby element analysis). The results <strong>of</strong> the Volhard analysis method determined the freechloride ions in the bulk solution measured by the AgNO 3 titration method. Their resultsimpliedthat theactivesiteintheresincouldnotreactcompletely withtheorganic reactantin durating the triphase reaction. According to the experimental results, this reaction is atwo-zone model (or shell–core model). The reaction occurs in ashell zone, <strong>and</strong> does notoccur in acore zone. The triphasic reaction mechanism <strong>and</strong> the swollen type <strong>of</strong> resin areshowninFig.5.Thismechanismcan<strong>of</strong>ferusanunderst<strong>and</strong>ing<strong>of</strong>thereactionphenomenain triphase reactions.IV.SOLID–LIQUID PHASE TRANSFER CATALYSISThe function <strong>of</strong> solid–liquid phase transfer catalysis (SLPTC) is to conduct the reaction <strong>of</strong>a solid salt <strong>and</strong> the organic reactant using a PT catalyst that is easily dissolved in theorganic phase in the absence <strong>of</strong> water. These catalysts can be tertiary amines, quaternaryammonium salts, diamines, crown ethers <strong>and</strong> crypt<strong>and</strong>s, among which crown ethers, act asthe catalysts because <strong>of</strong> their specific molecular structures [183–186]. Starks et al. [183]indicated that 100% <strong>of</strong> the yield <strong>of</strong> product benzyl acetate was obtained at 258C in 2 h forCopyright © 2003 by Taylor & Francis Group, LLC


FIG. 5<strong>Mechanism</strong> <strong>of</strong> the triphasic reaction (a) <strong>and</strong> the swollen type <strong>of</strong> resin (b).the substitution reaction <strong>of</strong> potassium acetate <strong>and</strong> benzyl bromide using 18-crown-6 asthe catalyst under solid–liquid PT conditions. This phenomenon <strong>of</strong> high conversion <strong>and</strong>product yield using SLPTC promotes more research work in investigating this type <strong>of</strong>reaction.The most important step in PT-catalyzed reactions is that the catalyst must have theability to transfer the reacting anion into the organic phase to react with the organicsubstrate. In an aqueous–organic two-phase system, the reacting nucleophile is locatedin the aqueous phase <strong>and</strong> is usually insoluble or slightly soluble in the organic phase underthe operating conditions. In the situation <strong>of</strong> the absence <strong>of</strong> water, the anion nucleophileshould be given by the solid salt reactant, such that the unfavorable side reaction isprobably inhibited. In addition, SLPTC can promote the weak nucleophiles, such assalts <strong>of</strong> acetate, to have much higher reactivity by eliminating the hydrolysis effect.Hence, for SLPTC, it has the advantages <strong>of</strong> easy separation <strong>of</strong> products from reactants,easy selection <strong>of</strong> organic solvents, easy recovery <strong>of</strong> catalysts, the inhibition or prevention<strong>of</strong> unfavorable side reactions, etc., <strong>and</strong> shows great potential for commercial applicationsCopyright © 2003 by Taylor & Francis Group, LLC


[187–196]. Several reactions that cannot be performed in liquid–liquid phases can becarried out efficiently in solid–liquid systems.Starks et al. [183] have addressed several questions regarding the mechanistic details<strong>of</strong> SLPTC, <strong>and</strong> those include what are the mechanisms <strong>of</strong> transport <strong>of</strong> anions from thesolid phase to the organic phase, the mechanisms <strong>of</strong> formation <strong>of</strong> reactive ion pairs, themechanisms <strong>of</strong> exchange <strong>of</strong> product anions located in the organic phase with reactantanions located in the solid phase, the effects <strong>of</strong> particle size on the rates <strong>of</strong> reaction, themechanistic differences between quaternary cation <strong>and</strong> crown ethers as PT catalysts, <strong>and</strong>the mechanistic role <strong>of</strong> small quantities <strong>of</strong> water in SLPTC. Obviously, the behavior <strong>of</strong> theactive ion pairs or catalytic intermediates is important in realizing the mechanism <strong>of</strong>SLPTC.A. <strong>Interfacial</strong> Phenomena1. The Omega <strong>Phase</strong>For solid–liquid PT-catalyzed reactions using crown ethers as the catalyst, the correspondingcation <strong>of</strong> the solid reactant has some limitations, e.g., a potassium salt system can onlyuse 18-crown-6 as the catalyst, while 15-crown-5 can only catalyze the reaction <strong>of</strong> asodium salt. This is because metal salts carried by crown ethers depend on their molecularstructures with the cation size just fitting into the cage <strong>of</strong> the crown ether; the activecomplex is then transported into the organic phase. Moreover, the solubility <strong>of</strong> this activecomplex is related to its lipophilicity in the organic solvent [184,185].In many solid–liquid systems using crown ethers as the catalyst, adding smallamounts <strong>of</strong> water enhances the reaction rate greatly. A trace amount <strong>of</strong> water inSLPTC obviously plays an important role. When small quantities <strong>of</strong> water are added,the solid particles are surrounded by water molecule to form a thin layer. This interfaciallayer between the solid <strong>and</strong> the organic phases is termed the omega phase, whereby thesolubility <strong>of</strong> solid reactant in the solution is enhanced to produce easily the active intermediate.Liotta et al. [186] indicated that, using 18-crown-6 as the catalyst for the solid–liquid reaction <strong>of</strong> benzyl halide <strong>and</strong> potassium cyanide, 92% <strong>of</strong> the 18-crown-6 (as asolution in toluene) <strong>and</strong> inorganic salts KCN <strong>and</strong> KCl resided in the toluene phase;however, about 1–2% <strong>of</strong> the crown ether was transferred on to the surface <strong>of</strong> the salt<strong>and</strong> coated the surface <strong>of</strong> the salt particles to form a third phase when adding smallamounts <strong>of</strong> water.When the omega phase is formed, the overall reaction rate can be described bypseudo-first-order kinetics with respect to the organic reactant. While the reaction followspseudo-zero-order kinetics as the substitution reaction is conducted in the presence <strong>of</strong>crown ether <strong>and</strong> in the absence <strong>of</strong> water, it is independent <strong>of</strong> the benzyl halide concentration.Crown ether directly dissociates the cation <strong>of</strong> the reacting salt. A reaction mechanismwas proposed for the esterification reaction <strong>of</strong> solid potassium 4-nitrobenzoate <strong>and</strong> benzylbromide by using crown ether [197]. The overall reaction isO 2 N C 6 H 4 COO K þ CHCl 3 ;25 Cþ C 6 H 5 CH 2 Br !O 2 N C 6 H 4 COOCH 2 C 6 H 5 þ KBr ð81ÞCopyright © 2003 by Taylor & Francis Group, LLC


The reaction steps involve [197]:1. Dissolution <strong>of</strong> solid potassium nitrobenzoate:CE org þ KNB solid Ð CE KNB org ð82Þ2. Intrinsic reaction in the organic phase:CE KNB org þ PhCH 2 Br org ! PhCH 2 Br org þ CE KBr org ð83Þ3. Release <strong>of</strong> crown ether:CE KBr org Ð CE org þ KBr solid ð84ÞWith further additions <strong>of</strong> water, the overall reaction rate does not inevitably increase, butreaches a maximum with an optimal amount <strong>of</strong> water added.2. Solubilization <strong>of</strong> Solid Salt by Quaternary Ammonium SaltsSLPTC can also be conducted by using quaternary ammonium salt as the catalyst. Thisphenomenon is somewhat different from using crown ether. V<strong>and</strong>er Zwain <strong>and</strong> Hartner[198] concluded that, for the reaction <strong>of</strong> acetate <strong>and</strong> adeninyl anions in the solid–liquidPT-catalyzed reaction using tricaprylmethylammonium chloride, showed better efficiencythan crown ether. Yadav <strong>and</strong> Sharma [199] investigated the kinetics <strong>of</strong> the reaction forbenzyl chloride <strong>and</strong> sodium acetate/benzoate by SLPTC. They found that cetyldimethylbenzylammoniumchloride was the most efficient catalyst among those studied in thetemperature range 90–139 C, <strong>and</strong> the rate <strong>of</strong> reaction in the presence <strong>of</strong> water was lessthan that in the absence <strong>of</strong> water. The solubilities <strong>of</strong> NaOAc <strong>and</strong> NaCl in toluene assolvent at 101 C are 3:85 10 5 <strong>and</strong> 3:24 10 5 gmol/mL, respectively, while being 2:1910 5 gmol=mL for the former in the presence <strong>of</strong> dimethylhexadecylbenzyl chloride. Theconcentrations <strong>of</strong> chlorides <strong>and</strong> acetates are 6:25 10 5 <strong>and</strong> 5:6 10 5 gmol=mL.Obviously, the solubilities <strong>of</strong> these two salts are affected by the reaction with the PTcatalyst. Yee et al. [200] showed that the slower reactions catalyzed by quaternary saltsin a well-mixed batch reactor were caused by the limited effectiveness <strong>of</strong> quaternary saltsin solubilizing the solid reactant.Yang <strong>and</strong> Wu [201] investigated the esterification <strong>of</strong> dipotassium phthalate withbenzyl bromide in a solid–liquid system. We found that the catalytic intermediate, formedby the solid reactant with tetrabutylammonium bromide, was the key-reacting componentin SLPTC. Yang <strong>and</strong> Wu [202] explored the kinetics <strong>of</strong> the O-allylation <strong>of</strong> sodium phenoxidewith allyl bromide in the presence <strong>of</strong> quaternary ammonium salt catalyst in a solid–liquid system. The behaviors <strong>of</strong> the catalytic intermediate tetrabutylammonium phenoxide,formed from the reaction <strong>of</strong> solid sodium phenoxide <strong>and</strong> tetrabutylammonium bromidein the solid–liquid phases, are important in conducting the etherification, <strong>and</strong>pseudo-first-order kinetics are observed.The past efforts in SLPTC show that not only can the reactions be catalyzed byquaternary ammonium salt, but the interfacial reaction <strong>of</strong> the solid reactant with thequaternary ammonium salt is also important in this type <strong>of</strong> reaction. Moreover, thebehaviors <strong>of</strong> the active intermediate are also influenced by the addition <strong>of</strong> water in conductingthe quaternary salts catalyzed reactions. A conceptual scheme describing thereaction mechanism for SLPTC was proposed by Melville <strong>and</strong> Goddard [203,204], i.e.,heterogeneous solubilization <strong>and</strong> homogeneous solubilization, by considering the solubility<strong>of</strong> solid salts in the organic phase. For the heterogeneous solubilization mechanism,Copyright © 2003 by Taylor & Francis Group, LLC


the solid salt directly reacts with the quaternary catalyst at the solid–liquid interface toproduce the intermediate, which then transfers into the solvent <strong>and</strong> reacts with the organicsubstrate to form the product. For the homogeneous solubilization mechanism, the solidreactant can be dissolved in an organic solvent <strong>of</strong> generally higher polarity, <strong>and</strong> then reactswith the catalyst to form the intermediate. Melville <strong>and</strong> Yortsos [205] performed a theoreticalstudy regarding rapid homogeneous reactions based on a simple stagnant filmmodel in the system <strong>of</strong> SLPTC.Naik <strong>and</strong> Doraiswamy [206] reported that the homogeneous solubilization could befurther subdivided into four types, models A to D, for the following reactions:QX org þ MY s=aq Ð QY org þ MX s=aqQY org þ RX org ! QX org þ RY orgð85Þð86ÞModel A assumes that the solid dissolution <strong>and</strong> mass transfer steps are very fast comparedwith the organic reaction <strong>and</strong> that the solid particles MY <strong>and</strong> MX are present at theirequilibrium solubility levels in the organic phase. The concentrations <strong>of</strong> QY <strong>and</strong> QX in theorganic phase are both constant, i.e.,C QYo ¼ Kq 0K þ ; C QXo ¼ q 0K þ ð87ÞModel B assumes that both the ion-exchange reaction in Eq. (85) <strong>and</strong> the organic reactionin Eq. (86) are under kinetic control with the solid dissolution <strong>and</strong> mass transfer steps stillfast, <strong>and</strong> a differential equation describing the variation <strong>of</strong> QY with reaction time in theorganic phase is required. Model C assumes that MY is no longer at saturation concentrationin the organic phase, but is at some finite value. The rate <strong>of</strong> dissolution is governedby the interfacial area per unit volume <strong>of</strong> the organic phase, the dissolution rate constant,<strong>and</strong> the driving force between the saturation <strong>and</strong> the instant concentrations. Both the ionexchange<strong>and</strong> the organic reactions take place in the bulk organic phase, <strong>and</strong> the transport<strong>of</strong> species from the solid surface to the bulk liquid is assumed to be fast; in addition, thevariation <strong>of</strong> the interfacial area according to the progress <strong>of</strong> the reaction should also beaccounted for. Model D accounts for the effect <strong>of</strong> transport <strong>of</strong> QY from the thin filmoutside the solid surface to the bulk liquid, <strong>and</strong> incorporates the rate <strong>of</strong> the organicreaction. The ion-exchange reaction is assumed to be fast <strong>and</strong> completed within the film.In order to describe the solubilization <strong>of</strong> solid reactant in the organic phase, Yang<strong>and</strong> Wu [200] performed the ion-exchange reaction <strong>of</strong> sodium phenoxide with tetrabutylammoniumbromide in a solid–liquid system. The interfacial reaction <strong>and</strong> mass transfersteps are shown as follows. The independent ion-exchange reaction isPhONa ðsÞþQBr ðorgÞ !PhOQ ðorgÞþNaBr ðsÞð88ÞThis reaction involves several steps:(a) Dissolution <strong>of</strong> PhONa. Traces <strong>of</strong> water are present in the solid reactantPhONa:3H 2 O, <strong>and</strong> the omega phase around the solid particle is formed to enhance thesolubilization <strong>of</strong> PhONa in the solution. The rate <strong>of</strong> dissolution <strong>of</strong> PhONa is very fast,leading to the solid part <strong>of</strong> PhONa readily in equilibrium with its soluble part. The concentration<strong>of</strong> PhONa at the interface is thus kept at its saturation state.(b) Reaction <strong>of</strong> PhONa with QBr. PT catalyst QBr reacts with the soluble part <strong>of</strong>PhONa to form PhOQ at the solid–liquid interface. The film reaction is assumed to bereversible with the equilibrium constant K 1 :Copyright © 2003 by Taylor & Francis Group, LLC


PhONa ðorgÞþQBr ðorgÞk 1!PhOQ ðorgÞþNaBr ðorgÞð89ÞK 1 ¼ C PhOQC NaBrC PhONa C QBrð90ÞIn Eq. (90) the asterisk represents the component concentration in the layer adjacent to thesurface <strong>of</strong> the solid particle.(c) Mass <strong>Transfer</strong> <strong>of</strong> PhOQ to the Bulk Organic <strong>Phase</strong>. The intermediate PhOQ transfersfrom the solid–liquid interface to the organic phase, wherein PhOQ has limitingsolubility. The equation for the rate <strong>of</strong> change is given asdC PhOQV org ¼ Kdt m A s CPhOQ C PhOQ ð91ÞThe term A s denotes the surface area <strong>of</strong> the solid particle, which gradually reduces duringthe progress <strong>of</strong> the reaction, <strong>and</strong> is derived from the mass balance <strong>of</strong> PhONa used, i.e.,with 2=3¼ A s0 1 q C PhOQ 2=3ð92ÞA s ¼ A s0N PhONa;0 N PhOQN PhONa;0q ¼N QBr;0N PhONa;0<strong>and</strong> C PhOQ ¼ C PhOQC QBr;0The mass transfer coefficient k m , which is also dependent on the particle size <strong>and</strong> theoreticallyinversely proportional to the n power <strong>of</strong> the particle size where n is in the range0.25–1.0 (from high to low Reynolds number), can be expressed as a function <strong>of</strong> saltconversion:N PhONa;0 N PhOQk m ¼ k m0N PhONa;0By combining Eqs (91)–(93), the rate <strong>of</strong> change <strong>of</strong> PhOQ in the organic phase is deduced,d C PhOQdtwhereð93Þ n=3¼ k m0 1 q C PhOQ n=3ð94Þ¼ 1 C PhOQ 1 q C PhOQ 2 nÞ=3ð95ÞC QBr;0C NaBr ¼K 1 CPhONa C QBr<strong>and</strong> ¼ k m0A s 0V orgð96ÞYang <strong>and</strong> Wu [202] showed that near-saturated concentrations <strong>of</strong> PhOQ in chlorobenzeneas solvent at different temperatures were achieved after about 20 min <strong>of</strong> operation.The difference in PhOQ concentraitons at various reaction temperatures was notsignificant in this case. This shows that the catalytic intermediate PhOQ can be formedfrom tetra-n-butylammonium salt reacted with PhONa in a solid-liquid system.Polyethylene glycols (PEGs) can also be used as the catalyst in SLPTC. Chu [207]reported the kinetics for etherification <strong>of</strong> sodium phenoxide with benzyl bromide usingquaternary ammonium salts <strong>and</strong> PEG as the catalyst in SLPTC. When PEG is used as thecatalyst, formation <strong>of</strong> the complex PEG–Na þ PhO mainly occurs at the solid–liquidinterface. The phenoxide anion carried by PEG can dissolve much more than its originalCopyright © 2003 by Taylor & Francis Group, LLC


solubility in the organic phase, thus enhancing the overall reaction. The reaction scheme isshown in Fig. 6. For the same reaction system, but using quaternary ammonium saltinsteadPEGs,theactiveintermediatebecomesPhOQproducedfromsolidNaOPhreactedwith catalyst QBr. The solubility <strong>of</strong> PhOQ varies in different kinds <strong>of</strong> solvent <strong>and</strong> leads todifferent reaction rates. The variations in the catalytic intermediate PhOQ with respect totimeforchlorobenzene, dichlorobenzene, <strong>and</strong>heptaneare shown in Fig. 7,from whichtheconcentration PhOQ in heptane is the least. However, the overall reaction rate in heptaneis still at a high level; this shows that the interfacial reaction is dominant in this case [207].B. Adsorption Effect on the Solid Surface1. Formation <strong>of</strong> the Active ComplexIn contrast with the reaction mechanism <strong>of</strong> heterogeneous <strong>and</strong> homogeneous solubilization,Yufit et al. [208] proposed a new mechanism for SLPTC that included step-by-stepformation <strong>of</strong> a cyclic ternary complex [208]. This mechanism is based on the formation <strong>of</strong>two pairs <strong>of</strong> binary complexes (BCs) <strong>and</strong> ternary complexes (TCs) obtained from theorganic reactant RX, the solid reactant MY, <strong>and</strong> the PT catalyst QX adsorbed on asolid salt surface as follows [208]:RX þ MY þ QX Ð TC1 Ð TC2 Ð RY þ MX þ QXð97Þð98ÞThey also analyzed the energetics <strong>of</strong> the substitutions with solid salts <strong>of</strong> different strength<strong>of</strong> M—Y bonds <strong>and</strong> concluded that the rate-determining step was the rearrangement <strong>of</strong>FIG. 6 Reaction scheme for benzyl bromide reacted with sodium phenoxide using PEG as thecatalyst in SLPTC.Copyright © 2003 by Taylor & Francis Group, LLC


FIG. 7 Effect <strong>of</strong> solvent on formation <strong>of</strong> PhOQ in SLPTC: benzyl bromide 0.005 mol, sodiumphenoxide 0.005 mol, TBAB 0.001 mol, solvent 50 mL, agitation 350 rpm, temperature 70 C; (*)C 6 H 5 Cl, (~) C 6 H 4 Cl 2 ,(&) n-C 7 H 16 .the TCs. From the kinetic analysis <strong>of</strong> different reaction mechanisms, the observed reactionrate constant k obs was determined by the equation:k obs ¼k½MYŠ 0 ½QXŠ 01 þ K½RXŠ 0 ½MYŠ 0 þ½QXŠ 0 ð99ÞIn Eq. (99), k is the combined rate constant <strong>and</strong> K is the equilibrium constant for thereversible reaction <strong>of</strong> TC formation.Generally, many experimental results can be described by applying pseudo-firstorderor pseudo-second-order kinetics successfully. Sometimes, however, using confinedkinetic data to elucidate exactly the reaction mechanism is indeed difficult. Hence, severalsimplified reaction mechanisms are usually employed to describe the kinetic behaviors <strong>of</strong>the reaction systems successfully. The technique <strong>of</strong> topochemistry is an effective methodfor achieving an approximate <strong>and</strong> quite precise interpretation <strong>of</strong> the kinetic data. Sirovskiet al. [209] discussed the applicability <strong>of</strong> the models developed for the topochemical reactionsin SLPTC. They considered that the simplest kinetic equation, called the Er<strong>of</strong>eevequation [210,211]:x ¼ 1 exp k n ð Þ ð100Þwith the rate constant k, the conversion degree x, <strong>and</strong> the parameter n depending on thegeometry <strong>of</strong> the nuclei, is more appropriate for a description <strong>of</strong> SLPTC than more complicatedones recommended in the literature. Such a kinetic equation is widely used for thedescription <strong>of</strong> topochemical processes. Sirovski et al. [209] investigated an S N 2 reaction <strong>of</strong>benzyl chloride with sodium acetate under SLPTC conditions:PhCH 2 Cl þ AcON Aliquat !336 PhCH 2 OAcð101ÞThey observed that the reaction rate did not follow simple kinetic laws under their operatingconditions. A possible reaction scheme was thus proposed [209]:Copyright © 2003 by Taylor & Francis Group, LLC


QCl þ AcONa ðsÞ ÐQCl:AcONa ðsÞQCl:AcONa ðsÞþPhCH 2 Cl Ð QCl:AcONa:PhCH 2 ClQCl:AcONa:PhCH 2 Cl ! QCl:NaCl:PhCH 2 OAcQCl:NaCl:PhCH 2 OAc Ð QCl:NaCl ðsÞþPhCH 2 OAcQCl:NaCl ðsÞ ÐQCl þ NaCl ðsÞð102Þð103Þð104Þð105Þð106ÞThey also found that the Er<strong>of</strong>eev equation described the observed kinetics much betterthan other simple kinetic equations. Yufit <strong>and</strong> Zinovyev [212] compared the kinetic study <strong>of</strong>nucleophilic substitution under PTC conditions in liquid–liquid <strong>and</strong> solid–liquid systems.They observed the effect <strong>of</strong> initial exponential burst (IB) on the kinetic curve in the reactionwith solid salts for the S N 2 reaction <strong>of</strong> 2-octylmesylate with potassium halides under PTCconditions. In their study, they assumed that the active sites on which the reaction occuredwere present on the solid surface through the formation <strong>of</strong> complexes <strong>of</strong> salts, catalysts, <strong>and</strong>substrate [212–215]. They also concluded that the phenomenon <strong>of</strong> IB was characterized bythe first-order dependence on the initial stage <strong>of</strong> conversion <strong>and</strong> by zero-order dependenceup to high conversion. Therefore, the kinetic equation for the reaction becomes a sum <strong>of</strong>linear <strong>and</strong> exponential terms with correlated parameters A <strong>and</strong> B½PŠ ¼At þ½XŠð1 exp BtÞ ð107Þwhere P represents the key product, [X] is the concentration <strong>of</strong> product formed by the firstorderlaw, <strong>and</strong> t is the reaction time. They also proposed a reaction mechanism includingthe adsorption on the solid surface for the solid–liquid system [212]:KBr ðsÞþQCl ð1Þ !KCl ðsÞþQBr ð1Þð108ÞQBr ð1ÞþROMs ð1Þ !QOMs ð1ÞþRBr ð1ÞQOMs ð1ÞþKBr ðsÞ ÐQOMs:KBr ðadsÞQOMs:KBr ðadsÞþROMs ð1Þ ÐQOMs:KBr:ROMs ðadsÞQOMs:KBr:ROMs ðadsÞ !RBr ð1ÞþQOMs:KOMs ðadsÞQOMs:KOMs ðadsÞþKBr ðsÞ ÐQOMs ð1ÞþKOMs:KBr ðadsÞð109Þð110Þð111Þð112Þð113ÞKOMs:KBr ðadsÞ !KOMs ðsÞþKBr ðsÞð114ÞAnother mechanism based on the concept <strong>of</strong> topochemical reaction, which meansthat the reaction rate is dependent on the characteristics or properties <strong>of</strong> the interface, hasbeen proposed by Yang <strong>and</strong> Wu [216] in investigating the esterification <strong>of</strong> linear dicarboxylateusing SLPTC for solid dipotassium sebacate (SeK 2 ) reacted with benzyl bromide(RBr). The overall reaction isKOOCC 8 H 16 COOK ðsÞþ2C 6 H 5 CH 2 Br ðorgÞ QBr!C 6 H 5 CH 2 OOCC 8 H 16 COOC H 2 C 6 H 5 ðorgÞþ2KBrðsÞð115ÞThe reaction steps involves [216]:1. Dissolution <strong>of</strong> SeK 2 from the solid surface to the organic film.SeK 2 ðsolidÞ !SeK 2 ðorgÞð116ÞCopyright © 2003 by Taylor & Francis Group, LLC


With the saturation solubility CSeK 2in the organic solvent, the dissolution rate coefficientk S is dependent on the surface property <strong>and</strong> the degree <strong>of</strong> agitation.2. Formation <strong>of</strong> a transition complex [R–Br–Q–Br] (TC) in the organic phase withequilibrium constant K 1 .RBr ðorgÞþQBr ðorgÞ ÐTC ðorgÞð117ÞFormation <strong>of</strong> the active transition complex TC from the reaction <strong>of</strong> QBr <strong>and</strong> RBr isassumed to be fast <strong>and</strong> reversible.3. The substitution reaction <strong>of</strong> TC <strong>and</strong> dissolved SeK 2 in the organic film with thethird-order rate constant k.SeK 2 ðorgÞþ2TCðorgÞ !SeR 2 ðorgÞþ2 KBr ðorgÞþ2 QBr ðorgÞð118ÞThe adsorption <strong>of</strong> QBr on the solid surface <strong>of</strong> SeK 2 plays an insignificant role in thekinetic description. The solid–liquid equilibrium <strong>of</strong> KBr between its soluble parts <strong>and</strong>solid parts is still existed. Wu [217] reported that the kinetic data for S-shape curveswere found in this system, as shown in Fig. 8 for different amounts <strong>of</strong> potassium sebacateused. This revealed that the catalytic transition complex [R–Br–Q–Br] in the organic phasewould lead to a long induction period for the reaction <strong>of</strong> SeK 2 with TC.The concentration <strong>of</strong> TC in the organic phase <strong>and</strong> the rate <strong>of</strong> change <strong>of</strong> componentsare derived as follows [216]:C TC;org ¼ K 1 C RBr;org C QBr;orgdC SeK2 ;orgdtdC SeR2 ;orgdt¼ k S C SeK 2¼ kC SeK2 ;orgC 2 TC;orgC SeK2 ;orgkC SeK2 ;orgC 2 TC;orgð119Þð120Þð121ÞEquation (120) is simplified by neglecting the term kC SeK2 ;orgC 2 TC;org to obtain the concentration<strong>of</strong> SeK 2 in the organic film, C SeK2 ;org ¼ C SeK 2ð1 e k St Þ, which can be furthertransformed into the equation C SeK2 ;org ¼ C SeK 2ðk s tÞ n for ease <strong>of</strong> interpreting the kineticFIG. 8 Yields <strong>of</strong> product dibenzyl sebacate for different molar ratios (r) <strong>of</strong> dipotassium sebacate tobenzyl bromide: chlorobenzene 50 mL, benzyl bromide 0.01 mol, TBAB 0.0025 mol, agitation 350rpm, temperature 80 C; r values: (*) 0.125, (^) 0.25, (&) 0.5,()1.0, (*) 2.0.Copyright © 2003 by Taylor & Francis Group, LLC


curve <strong>of</strong> the product SeK 2 . The parameters <strong>and</strong> n are determined for a definite set <strong>of</strong>data <strong>and</strong> have the physical meaning <strong>of</strong> characterizing the surface properties <strong>of</strong> solidreactant SeK 2 . Different sets <strong>of</strong> data in ð1 e kst ) give different sets <strong>of</strong> <strong>and</strong> n for bestfitting with a precision greater than 0.99. The apparent reaction rate constant is thendeduced from the equation:wheredYdt ¼ 2kK 2 1 C 2 QBr;orgC RBr;0 C SeK 2k n S1 þ K 1 C QBr;org 2t n ð1 YÞ 2 ¼ k app;0 C RBr;0 t n ð1 YÞ 2 ð122Þk app;0 ¼ 2k n SkC SeK 2K 2 1 C 2 QBr;org= 1 þ K 1 C QBr;org 2<strong>and</strong> Y ¼ 2C SeR2 ;org=C RBr;0It it noted that k app;0 is subjected to the effects <strong>of</strong> rate <strong>of</strong> dissolution, intrinsicreactivity, rate <strong>of</strong> formation <strong>of</strong> transition complex, catalyst amounts, the solubility <strong>of</strong>solid reactant in the organic phase, <strong>and</strong> the characteristics <strong>of</strong> the solid surface, <strong>and</strong> hasthe dimensions <strong>of</strong> [ðtimeÞ 1 n ðconcentrationÞ 1 ]. The resultant equation from integratingEq. (122) is similar to the conversion equation deduced from topochemistry theory. Bytaking the natural logarithm on both sides, one can obtain a rather simplified equationused to correlate the kinetic behaviors, i.e., Yln ¼ ln k app;0C RBr;0þðn þ 1Þ½ lnðtÞŠ ð123Þ1 Yn þ 1Moreover, further simultaneous generation <strong>of</strong> KBr during the reaction <strong>of</strong> SeK 2 with TCwould make the organic solution much more slushy, which in turn would reduce the filmreaction rate due to the steric hindrance when a much higher catalyst amount was used.2. Deactivation Behavior <strong>of</strong> CatalystIn SLPTC, the effect <strong>of</strong> the side-product salt on the overall reaction rate is sometimesobserved to be severe after a specific reaction time. The kinetic curve shows that thepseudo-first-order reaction initially obeyed is no longer followed at later reaction times,behaving in a diminished kinetic order, <strong>and</strong> the whole reaction is finally stopped. This sideproduct is produced from the anion-exchange reaction at the solid–liquid interface.Depending on the polarity <strong>of</strong> the solvent, this generated metal salt is usually difficult todissolve in the organic phase <strong>and</strong> has a tendency to adsorb on to the surface <strong>of</strong> the solidparticle. The adsorbed salts thus strongly influence the subsequent anion-exchange reaction,from which the active intermediate is formed. Such an effect is more significant forthe fast anion-exchange reaction or at a higher reaction temperature.Yang et al. [218] investigated the substitution reaction <strong>of</strong> sodium phenoxide withethyl 2-bromoisobutyrate in a solid–liquid PT-catalyzed system. The deactivation <strong>of</strong> catalystactivity on the apparent pseudo-first-order reaction rate was observed. Such a phenomenonresults from the salts deposited on the surface <strong>of</strong> solid particles during theprogress <strong>of</strong> a reaction. The deposition <strong>of</strong> salts decreases the rate <strong>of</strong> formation <strong>of</strong> the activeintermediate, leading to the observed reaction order change [218]. We applied the pseud<strong>of</strong>irst-orderreaction with catalytic deactivation kinetics to show that the initial reaction ratewas not influenced by the agitation rate when exceeding 350 rpm, but the deactivation rateincreased with increasing stirring speed <strong>and</strong> the amount <strong>of</strong> catalyst used. Adding a smallamount <strong>of</strong> water resulted in a reduction in the apparent reaction rate. A more lipophilicCopyright © 2003 by Taylor & Francis Group, LLC


quaternary cation solvates the solid reactant anion much more easily, <strong>and</strong> leads to a fasterinitial reaction rate. The order <strong>of</strong> reactivity for different PT catalysts is determined asTBPB > TBAB > TBAI TBAHS Aliquat 336 for the reaction.Yang et al. [219] investigated the kinetics <strong>of</strong> the etherification <strong>of</strong> ethyl 2-bromoisobutyrate(RX) with potassium 4-benzyloxyphenoxide in the presence <strong>of</strong> potassium iodidein SLPTC. In that work, we found that for various molar ratios <strong>of</strong> TBAB to RX (denotedas f ) the yield <strong>of</strong> product ArOR increased with increasing catalyst amounts up to f ¼ 0:60.Too much catalyst employed in the presence <strong>of</strong> KI results in the reduction <strong>of</strong> catalyticefficiency. This effect is due to two major reasons: first, the solubility <strong>of</strong> the catalyticintermediate in the organic solvent is limited; second, the formation <strong>of</strong> the catalytic intermediatein chlorobenzene is retarded because use <strong>of</strong> a higher amount <strong>of</strong> catalyst inducedrapid deposition <strong>of</strong> the generated potassium salts on the solid surface. Adding the extrasalt KI enhances the reactivity <strong>of</strong> PT catalyst, but the active intermediate in the organicphase is diminished when much KI is present. Small amounts <strong>of</strong> KI promote the conversion<strong>of</strong> RX into RI, which is more reactive in the organic reaction. The reaction stepsconcerning the deactivation <strong>of</strong> the catalyst are shown below [219].The overall reaction isKI;QXArOK ðsÞþRX ðorgÞ ! ArOR ðorgÞþKX ðsÞ ð124ÞThe reaction mechanism for the overall reaction is as the following steps:ArOKðsÞ ! ArOK ðorgÞKI ðsÞ ! KI ðorgÞð125Þð126ÞK 1!KI ðorgÞþQX ðorgÞ KX ðorgÞþQI ðorgÞð127ÞK 2!KI ðorgÞþRX ðorgÞ KX ðorgÞþRI ðorgÞð128ÞKXðorgÞ ! KX ðsÞð129ÞK 3!ArOK ðorgÞþQX ðorgÞ ArOQ ðorgÞþKX ðorgÞð130ÞK 4ArOK ðorgÞþQI ðorgÞ! ArOQ ðorgÞþKI ðorgÞð131ÞRX ðorgÞþArOQ ðorgÞk b! ArOR ðorgÞþQX ðorgÞð132ÞRI ðorgÞþArOQ ðorgÞk b! ArOR ðorgÞþQI ðorgÞð133ÞThe rate <strong>of</strong> change <strong>of</strong> ArOR is then expressed asdC orgArOR¼ kdt a C orgRX þ k bC org RI CorgArOQð134ÞCopyright © 2003 by Taylor & Francis Group, LLC


Taking the mass balance for the cation Q <strong>of</strong> the PT catalyst in the system givesC orgQX;0 ¼ Corg QX þ Corg QI þ Corg QrOQThe expression for the active intermediate isC orgArOQ ¼1 þ 1C org ArOKC orgQX;0C orgKXþ Corg KIK 3 K 4ð135Þ ð136ÞThe concentration <strong>of</strong> ArOQ depends on the amounts <strong>of</strong> ArOK, KX, <strong>and</strong> KI, <strong>and</strong> theinitial usage <strong>of</strong> the catalyst QX. Combining the mass balance equation for initial RX in theorganic phase with Eq. (134), a deactivation function can be introduced in the situationunder decline <strong>of</strong> catalytic efficiency, leading to the following equations:withdC orgArORdt¼ k app;0C orgRX;0C orgArORk app;0 ¼ðk a þ k b K 2 C orgKI =Corg KX ÞCorg QX;0ð137Þ<strong>and</strong>1 ¼ 1 þ K 2C org ð138ÞKIC org 1 þ 1 C orgKXC org KXþ Corg KIArOK K 3 K 4In Eq. (137), k app;0 is the initial apparent reaction rate constant <strong>and</strong> is dependent on theamounts <strong>of</strong> KI, KX, <strong>and</strong> QX in the organic phase. If the rate <strong>of</strong> change <strong>of</strong> ArOR follows apseudo-first-order reaction, then would be approximately a constant. In such cases, nodeactivation effect appears. If the reaction rate behaves as a diminished first order, then decreases with the progress <strong>of</strong> the overall reaction.To evaluate the exact variation <strong>of</strong> with time is to measure the rate <strong>of</strong> deposition <strong>of</strong>KX <strong>and</strong> the other parameters directly or to apply an empirical correlation relating to theeffect <strong>of</strong> the decrease in C orgArOKon the overall reaction. The expression <strong>of</strong> the deactivationfunction <strong>and</strong> the kinetic data would determine the form <strong>of</strong> , such as1 ¼ð1 þ k d tÞ 2 or ¼ 1ð139Þ1 þ k d tLepertit <strong>and</strong> Che [220] discussed the definitions <strong>of</strong> interfacial co-ordination chemistry(ICC) <strong>and</strong> surface organometallic chemistry (SOMC) <strong>and</strong> compared their maincharacteristics <strong>and</strong> applications. The concepts <strong>of</strong> ICC applied to catalyst preparation,adsorption, <strong>and</strong> relations with catalysis are also useful in the development <strong>of</strong> interfacialmechanisms.C. Mass <strong>Transfer</strong> EffectsSufficient kinetic information should be collected to proceed the process design for aspecific reaction system. The factors affecting the performance <strong>of</strong> SLPTC includes agitationrate, particle size <strong>of</strong> solid salt, reaction temperature, the amount <strong>of</strong> solid reactant, thekinds <strong>and</strong> amount <strong>of</strong> PT catalyst, the solubility <strong>and</strong> the dissolution rate <strong>of</strong> solid reactant inthe organic solvent, extra addition <strong>of</strong> other metal salts, the polarity, surface tension, <strong>and</strong>Copyright © 2003 by Taylor & Francis Group, LLC


viscosity <strong>of</strong> the solvent, the amount <strong>of</strong> organic reactant, <strong>and</strong> the presence or absence <strong>of</strong>water. The reactor employed in SLPTC is usually an agitated batch type with the solid saltsuspended in the solution. The mass transfer rate between the solid <strong>and</strong> the liquid phases isimportant in designing a SLPTC reactor.In an agitated reactor, the effect <strong>of</strong> mass transfer resistance can be reduced to aminimum by adjusting the stirring speed. The mass transfer coefficient is also a function <strong>of</strong>the size <strong>of</strong> suspended particles. From the point <strong>of</strong> view <strong>of</strong> reactor design, to maintain theuniformity <strong>of</strong> the desired product from batch to batch the particle size distribution <strong>of</strong> thesolid reactant should be in a rather narrow range to render the mass transfer resistanceunimportant.1. Mass <strong>Transfer</strong> CoefficientMelville <strong>and</strong> Goddard [204] used rotating disk flow to measure the mass transfer coefficientbetween the solid <strong>and</strong> liquid phases in SLPTC for the reaction <strong>of</strong> benzyl chloride <strong>and</strong>solid potassium acetate using Aliquat 336 as the catalyst in acetonitrile as solvent. Theconcentration <strong>of</strong> quaternary ammonium acetate is expressed in the following equations:C QOAc ¼ 1 e t ð140Þ ¼ K 01C KOAc; ¼kð1 þ SK01Þ1 þ K 01ð141ÞK1 0 ¼ K 1C QCl; k ¼ A D kC KCl V ;S ¼ D QD Kð142Þwhere K 1 is the equilibrium constant for the reaction, KOAc þ QCl ! KCl þ QOAc; D i isthe diffusivity coefficient <strong>of</strong> component i; is the film thickness <strong>of</strong> mass transfer; A is thesurface area <strong>of</strong> solid potassium acetate; <strong>and</strong> V is the liquid volume. They concluded thatthe solid–liquid reaction was effected after the solid potassium acetate dissolved in thehigh-polarity solvent. Yee et al. [200] also applied rotating disk flow to carry out the masstransfer experiments for solid benzoate. The mass transfer coefficient K is obtained asK ¼ 0:6205D1=3 ! 1=2 1=6 f ðScÞf ðScÞ ¼1 þ 0:2980Sc 1=3 þ 0:01451Sc 2=3ð143Þð144Þwhere ! is the angular velocity (rad/s), is the kinematic viscosity (cm 2 =s), D is thediffusivity (cm 2 =sÞ, <strong>and</strong> SC is the Schmidt number (=DÞ:2. Pseudo-First-Order <strong>Kinetics</strong> Neglecting Mass <strong>Transfer</strong> EffectPseudo-first-order kinetics are usually observed in many solid–liquid PT-catalyzed reactionswhen the mass transfer effect is insignificant. For the reaction between the organicsubstrate RX <strong>and</strong> the nucleophile MY, the equation isRX ðorgÞþMY ðsÞþQ þ X ðorgÞ !RY ðorgÞþMX ðsÞþQ þ X ðorgÞð145ÞIn the case <strong>of</strong> some solid MY dissolved in the organic phase, the equilibrium state isachieved in a short reaction time:MY ðsÞ $M þ Y ðorgÞð146ÞCopyright © 2003 by Taylor & Francis Group, LLC


The ion-exchange reaction occurs at the interfacial zone to form QY, then conducting theintrinsic reaction:M þ Y ðorgÞþQ þ X ðorgÞ $Q þ Y ðorgÞþM þ X ðorgÞQ þ Y ðorgÞþRX ðorgÞ !RY ðorgÞþQ þ X ðorgÞð147Þð148ÞThe regenerated Q þ X ðorgÞ continues to catalyze the formation <strong>of</strong> Q þ Y ðorgÞ, <strong>and</strong>M + X (org) is always in equilibrium with MX(s):M þ X ðorgÞ $MX ðsÞð149ÞThe rate <strong>of</strong> equation is derived asd½RXŠdt¼ k org ½RXŠ½Q þ Y Š¼ k org K e½M þ Y Š½M þ X Š ½RXŠ½Qþ X ŠK e ¼ ½Qþ Y Š½M þ X Š½Q þ X Š½M þ Y Šð150Þð151ÞIf the value f ¼½M þ Y Š=½M þ X Š is approximately a constant, the rate <strong>of</strong> the equation canbe expressed in a pseudo-first-order form:d½RXŠ¼ kdt app ½RXŠk app ¼ k org K e f ½Q þ X Šð152Þð153ÞDuring the course <strong>of</strong> reaction, when the value <strong>of</strong> f is gradually diminished, the initialreaction rate <strong>and</strong> deactivation rate should be applied. With or without adding waterinfluences the mass transfer rate from the interface into the organic phase.D. Kinetic Modeling <strong>of</strong> Heterogeneous SolubilizationNaik <strong>and</strong> Doraiswamy [206] developed a mathematical model for the case <strong>of</strong> heterogeneoussolubilization that involves the steps <strong>of</strong> ion exchange in the solid phase, interphasetransport <strong>of</strong> the catalyst <strong>and</strong> the intermediate, <strong>and</strong> the organic reaction. In this model, ionexchange occurring within the solid phase is assumed due to the possible deposition <strong>of</strong> theproduct salt MX on the solid surface retarding formation <strong>of</strong> the catalytic intermediate.The controlling step can either be the liquid-phase transfer steps, the diffusion within thereactive solid, the adsorption–desorption steps, the surface ion-exchange reaction, or theliquid organic reaction. This treatment is similar to that in gas–solid catalytic reaction.The controlling step may shift to another step continuously with time. The reaction <strong>of</strong>organic RX with solid MY in the presence <strong>of</strong> PT catalyst QX is considered, <strong>and</strong> a poroussolid wherein the ion-exchange reaction takes place throughout the whole pellet, ratherthan at a sharp interface due to the liquid penetrating into it is assumed. The governingequations are derived as follows [206]:RX ðorgÞþMY ðsÞ ! QX RY ðorgÞ MX ðsÞ ð154ÞWithin the solid phase:Copyright © 2003 by Taylor & Francis Group, LLC


@CQXS ¼ D e@t@C s QY@th i@ r 2 @CQXS =@r r 2 k@rs¼ D e @ r 2 @C s QY =@rr 2 þ k@rsIn the organic bulk solution:dC orgQX¼ kdt 2 C orgRX Corg QYdC orgRX¼dtC orgQY ¼ q 0k 2 C orgRX Corg QYhk q aC oegQXC s QXC s QXC s QXiC s QYKCQYs Kð155Þð156Þð157Þð158ÞC orgQXC s QX;a C s QY;a ð159ÞC s QX;a ¼ 3ð 10 2 C s QXdThe initial <strong>and</strong> boundary conditions are:IC: t ¼ 0; C orgRX ¼ C0 RX; C s QX ¼ 0; C s QY ¼ 0; C orgQX ¼ q 0; C orgQY ¼ 0ð160Þð161ÞBC: r ¼ 0; dCs QXdrr ¼ R; D qdC s QXdrr ¼ R; D qdC s QYdr¼ dCs QYdr¼ 0h¼ k q C orgQXh¼ k q C orgQYC s QX;RC s QY;Riið162Þð163Þð164ÞThe above equations can be rendered dimensionless in terms <strong>of</strong> Thiele’s modulus, Biotnumber for mass transfer, <strong>and</strong> nondimensional time <strong>and</strong> distance, which are defined as 2 ¼ k sR 2; BiD m ¼ k qR; ¼ D ete D q R 2 ; ¼ r Rð165ÞIn the analysis <strong>of</strong> heterogeneous solubilization, the role <strong>of</strong> the solid-phase reaction ininfluencing the overall reaction is different from that for the usual gas–solid catalyticreaction. The most important situation is that the film <strong>and</strong> internal diffusion effects withinthe solid <strong>and</strong> at the solid–liquid interface are significant.V. TRI-LIQUID PHASE TRANSFER CATALYSISNeumann <strong>and</strong> Sasson [221] investigated the isomerization <strong>of</strong> allylanisole using PEG as thecatalyst in a toluene <strong>and</strong> aqueous KOH solution. They observed that a third-liquid phasewas formed between the aqueous <strong>and</strong> the organic phases. This was the first report regardingtri-liquid PTC. In 1987, Wang <strong>and</strong> Weng [222] performed the reaction <strong>of</strong> benzylchloride <strong>and</strong> sodium bromide using tetra-n-butylammonium bromide as the PT catalystin liquid–liquid phases. They found that the overall reaction rate rapidly increased whenthe amount <strong>of</strong> catalyst used exceeded some critical value. In such reaction conditions, thePT catalyst was found to be concentrated within a viscous liquid phase that was insolubleCopyright © 2003 by Taylor & Francis Group, LLC


in both aqueous <strong>and</strong> organic phases [222]. This liquid phase enhanced the overall reactionrate as much as several fold that in two-liquid phase systems with PTC, <strong>and</strong> was called thethird-liquid phase. The third-liquid phase was found to contain little <strong>of</strong> the organic <strong>and</strong>aqueous reactants, but mainly the highly concentrated catalyst, which exhibited hydrophilic<strong>and</strong> lipophilic properties. In the bromide–chloride exchange reaction system <strong>of</strong>Wang <strong>and</strong> Weng [222], the third liquid phase was found to consist mainly <strong>of</strong> Bu 4 NBr,small amounts <strong>of</strong> toluene, water, <strong>and</strong> sodium bromide. Above about 70% <strong>of</strong> the tetrabutylammoniumbromide was forced to form a separate liquid phase. The organic <strong>and</strong>aqueous reactants readily reacted with the concentrated catalyst to yield a high reactionrate. The PTC reaction in this situation was termed as tri-liquid PTC (TLPTC).From the point <strong>of</strong> view <strong>of</strong> industrial practice, the formation <strong>of</strong> a third phaseprovides not only enhancement <strong>of</strong> the reaction rate, but also easier separation <strong>of</strong> thePT catalyst from the product stream than that in a two-liquid phase. However, in someparticular reaction systems, the catalyst could lose as much as approximately 25% <strong>of</strong>the initial amount used. <strong>Catalysis</strong> by TLPTC was briefly reviewed by Naik <strong>and</strong>Doraiswamy in 1998 [223]. The key results from the previous publications are discussedas follows.A. Formation <strong>of</strong> the Third Liquid <strong>Phase</strong>Tetrabutylammonium salts are found to be able to form a third liquid phase underappropriate conditions. In principle, the formation <strong>of</strong> a third catalyst phase can beobtained by using a PT catalyst having limiting solubility both in the aqueous phase<strong>and</strong> organic phase under the interaction <strong>of</strong> other concentrated ingredients. Ido et al.[224] effected the elimination reaction <strong>of</strong> 2-bromo-octane with aqueous sodium hydroxideusing PEG as the catalyst [224]. By adding small quantities <strong>of</strong> methanol the solubility<strong>of</strong> PEG in the organic phase was greatly reduced, leading to the formation <strong>of</strong> athird liquid phase. Mason et al. [225] investigated the elimination <strong>of</strong> phenethyl bromideto styrene using tetrabutylammonium bromide under PT-catalytic conditions. Theyfound that the rate <strong>of</strong> reaction was accelerated rapidly due to the addition <strong>of</strong> morethan the critical amount <strong>of</strong> PT catalyst, <strong>and</strong> the third phase was rich in catalyst. Whenthe PT catalyst used was replaced by the tetrapropyl- or tetrapentyl-ammonium salts,the third liquid phase was not formed, <strong>and</strong> the precipitation <strong>of</strong> excess catalyst wassimply induced.Wang <strong>and</strong> Weng [226] explored the effects <strong>of</strong> solvents <strong>and</strong> salts on the formation <strong>of</strong> athird liquid phase for the reaction between n-butyl bromide <strong>and</strong> sodium phenolate. Theyconcluded that the polarity <strong>of</strong> the solvent <strong>and</strong> the amount <strong>of</strong> NaOH are two importantfactors in influencing the formation <strong>of</strong> a third liquid phase, the distribution <strong>of</strong> catalyst, <strong>and</strong>the reaction rate. The aqueous reactant NaOPh also exhibitis significant behavior incertain conditions. With the catalytic intermediate QOPh produced by the reaction <strong>of</strong>NaOPh <strong>and</strong> the catalyst QBr, NaOH has the ability to extract QOPh from the organicphase or the third liquid phase into the aqueous phase. For example, when the amount <strong>of</strong>NaOH added was 2 g, the amount <strong>of</strong> catalyst in chlorobenzene decreased to less than 10%<strong>of</strong> the original content, while the concentration <strong>of</strong> the catalyst in the aqueous phaseincreased with increasing NaOH added. In addition, when hexane was used as the solvent,adding a small amount <strong>of</strong> NaOH caused the disappearance <strong>of</strong> the third liquid phase, whichhad been formed before the addition <strong>of</strong> NaOH. This phenomenon is due to the dissolution<strong>of</strong> QOPh in the aqueous phase.Copyright © 2003 by Taylor & Francis Group, LLC


Ido et al. (1997) reported a halogen substitution between benzyl chloride in theorganic phase <strong>and</strong> potassium bromide in the aqueous phase catalyzed under the thirdliquid phase that was formed by changing the concentration <strong>of</strong> KBr, types <strong>of</strong> PT catalysts,<strong>and</strong> the organic solvents [227]. Yadav <strong>and</strong> Reddy (1999) investigated the n-butoxylation <strong>of</strong>p-chloronitrobenzene (PCNB) using the base NaOH under tri-liquid phase conditions.The typical composition <strong>of</strong> the third liquid phase was 55.12% <strong>of</strong> toluene, 22.52% <strong>of</strong>tetrabutylammonium bromide, 4.96% <strong>of</strong> p-chloronotrobenzene, 14.51% <strong>of</strong> water, <strong>and</strong>2.89% <strong>of</strong> n-butyl alcohol by weight. Distribution <strong>of</strong> the catalyst between the organicphase <strong>and</strong> the third liquid phase indicates about 89% <strong>of</strong> the total catalyst residing inthe third phase, <strong>and</strong> the overall reaction rate is attributed to the contribution <strong>of</strong> thereaction occurring in both the organic <strong>and</strong> the third liquid phases.Jin et al. [229] further performed the dehydrohalogenation <strong>of</strong> 2-bromo-octane withdodecane as the organic solvent <strong>and</strong> potassium hydroxide in the aqueous solution toinvestigate the synergistic effect <strong>of</strong> two PT catalysts in the situation <strong>of</strong> a third liquidphase using a combination <strong>of</strong> tetrahexylammonium bromide <strong>and</strong> PEG. They concludedthat a molecule <strong>of</strong> tetrahexylammonium bromide surrounded by many molecules <strong>of</strong> water<strong>and</strong> some PEG 200 led to the effect <strong>of</strong> water on the catalytic activity <strong>of</strong> tetrahexylammoniumbromide becoming weaker when the amount <strong>of</strong> PEG was increased.In summary, the operating conditions influencing the formation <strong>of</strong> the third liquidphase are: (1) type <strong>and</strong> quantity <strong>of</strong> the aqueous reactant, (2) type <strong>and</strong> quantity <strong>of</strong> PTcatalyst, (3) reactant <strong>and</strong> product in the organic phase, (4) the addition <strong>of</strong> otherinorganic salts, (5) lower polarity <strong>of</strong> the organic solvent, <strong>and</strong> (6) the reaction temperature.Increasing the reaction temperature benefits the formation <strong>of</strong> the third liquidphase due to the breakage <strong>of</strong> hydrogen bonding between the PT catalyst <strong>and</strong> thewater molecule.B. <strong>Interfacial</strong> <strong>Mechanism</strong> <strong>of</strong> TLPTC1. Reaction <strong>Mechanism</strong>The typical reaction mechanism for tri-liquid PTC in a batch reactor under agitation isillustrated in the schematic diagram <strong>of</strong> Fig. 9. Three types <strong>of</strong> reaction scheme consideringthe partition <strong>of</strong> the catalyst in the different phases <strong>and</strong> the place where the inherentreaction occurred have been proposed [226,227]. For the substitution reaction <strong>of</strong> alkylhalide (RX) <strong>and</strong> aqueous reactant metal salt (MY) using quaternary ammonium salt (QX)as the catalyst, the different types <strong>of</strong> reaction are addressed as follows [226].FIG. 9Reaction scheme for benzyl bromide reacted with sodium bromide in TLPTC.Copyright © 2003 by Taylor & Francis Group, LLC


Type I.The catalyst resides in the organic phase in a significant quantity:ð166ÞFor type I, the partition <strong>of</strong> catalyst between the organic <strong>and</strong> aqueous phases is importantin determining the intrinsic reaction rate <strong>and</strong> the utilization <strong>of</strong> the catalyst in the organicphase.Type II. The catalyst resides in the aqueous phase in a significant quantity:ð167ÞFor type II, the catalyst mostly stays in the aqueous phase, <strong>and</strong> the transferred RBr fromthe organic phase to the aqueous–organic interface reacts with the catalytic intermediateQOPh that is transported from the aqueous to the interface. The intrinsic reaction ismainly conducted at the interface between aqueous <strong>and</strong> organic phases.Type III. The third liquid phase appears with the catalyst <strong>and</strong> active intermediate allresiding in this viscous phase:ð168ÞBy adding more NaOPh to the reaction system the catalyst is salted out to form the thirdliquid phase. The active intermediate is then formed at the interface <strong>of</strong> the aqueous <strong>and</strong> thethird liquid phases by the reaction <strong>of</strong> QBr <strong>and</strong> NaOPh, which is transferred from theCopyright © 2003 by Taylor & Francis Group, LLC


aqueous bulk phase. The main intrinsic reaction <strong>of</strong> QOPh <strong>and</strong> RBr is conducted in thethird liquid phase. The product ROPh is then transferred into the organic phase.In reality, not all the catalyst would exist in the third liquid phase especially for thathaving high solubility in the aqueous phase. A distribution <strong>of</strong> the catalyst between theaqueous phase <strong>and</strong> the third liquid phase is probably retained. Hence, some parts <strong>of</strong> QOPhwould be produced in the aqueous phase, resulting in the distribution <strong>of</strong> QOPh betweenthe aqueous <strong>and</strong> the third liquid phases.2. Factors Affecting Catalyst Activity in TLPTCIn TLPTC, the essential step is to form the third liquid phase by adjusting the contents <strong>of</strong>inorganic salts <strong>and</strong> PT catalyst, <strong>and</strong> the interaction <strong>of</strong> the strong bases added. The overallreaction rates catalyzed by applying the third liquid phase are commonly enhanced tremendously,compared with the same reaction proceeding in liquid–liquid phases. Thevariables influencing the reaction rate can be summarized as follows:(a) Agitation Speed. Agitation plays an important role in a multiphase reaction system.Increasing the agitation rate increases the mass transfer rate <strong>of</strong> the componentbetween the immiscible phases <strong>and</strong> reduces the droplet size <strong>of</strong> the dispersed phase.Under agitated conditions, the mass transfer resistance at the interface between thethird liquid layer <strong>and</strong> the organic phase is affected by the droplet size. When the agitationrate increases to a critical value, the limiting step is dominated by the reactionwithin the catalyst-rich phase [224–227,230–233]. Yadav <strong>and</strong> Reddy [228] reported thatwith a speed <strong>of</strong> agitation from 650 to 1400 rpm, the rate <strong>of</strong> reaction increased withincreasing stirring, <strong>and</strong> after 1400 rpm the rate was independent <strong>of</strong> the interfacial masstransfer resistance.(b) Amount <strong>of</strong> <strong>Phase</strong> <strong>Transfer</strong> Catalyst. Different types <strong>of</strong> PT catalysts including quaternaryammonium salts <strong>and</strong> PEGs have been observed to enable the formation <strong>of</strong> thethird liquid phase, but under different conditions. Their common behaviors show thesharp discontinuity <strong>of</strong> the reaction rate before <strong>and</strong> after the formation <strong>of</strong> the thirdliquid phase. The observed reaction rate in the case <strong>of</strong> the tri-liquid phase increases linearlywith the total moles <strong>of</strong> quaternary ammonium bromide [228]. However, the quaternaryammonium salts with shorter alkyl chains show less tendency to form the thirdliquid phase, e.g., tetrapropylammonium bromide is ineffective for use as a catalyst in atri-liquid system. Mason et al. [225] indicated that, when a reaction mixture forming athree-liquid system was reconstructed by separating the middle third liquid phase, thereaction rate dropped by over half.Ido et al. [227] investigated the kinetics <strong>of</strong> a halogen exchange reaction in a threeliquidphase system <strong>and</strong> applied first-order kinetics to describe the overall reaction rate.They observed that the reaction rate constant includes the contributions <strong>of</strong> reactions in thethird-liquid phase <strong>and</strong> in the organic phase, <strong>and</strong> is a first order proportional to the totalcatalyst moles m cat . The reaction rate k inter occur at the interface between the aqueousphase <strong>and</strong> the organic phase is also important. Their results are shown as the followingequations [227]:dm A¼ kdt org x org þ k inter x aq þ k third D A x third mcat C A;org V orgmy A;org ¼exp k catV org þ D A V obs tthird V orgð169Þð170ÞCopyright © 2003 by Taylor & Francis Group, LLC


V orgk obs ¼V org þD A V thirdk org x org þk inter x aq þk third D A x thirdð171Þwhere x A denotes the mole fraction <strong>of</strong> the catalyst existing in the different phases, <strong>and</strong> K Arepresents the distribution <strong>of</strong> A(benzyl chloride) between an organic phase <strong>and</strong> athirdphase in equilibrium <strong>and</strong> is defined asD A C A;thirdC A;org¼ m A;thirdV orgm A;org V thirdð172ÞTaking the logarithm for Eq. (170), one can determine the observed rate constant k obsfrom the experimental data by plotting ½ lnðy A;org ÞŠversus time t:V orgmlnðy A;org Þ ln¼k catV org þD A V obs tð173Þthird V org(c) Reactant <strong>and</strong> Alkali Salt in the Aqueous <strong>Phase</strong>. The overall reaction rate inTLPTC usually increases with the increase in amount <strong>of</strong> strong base reactant in theaqueous phase. In contrast with abase-catalyzed elimination reaction, the third liquidphase already formed will be precipitated under the excess base to dehydrate the catalystphase. In the presence <strong>of</strong> 49% <strong>of</strong> NaOH, two liquids <strong>and</strong> one solid are observedinstead <strong>of</strong> three liquids at asomewhat lower base concentration.Ido et al. [227] found that increasing the aqueous reactantKBrincreases the reactionrateinTLPTC.Theionicstrengthintheaqueousphasealsoaffectstheease<strong>of</strong>formingthethird liquid phase, since adding extra salts tends to salt out ion pairs produced from theaqueous reactant with the quaternary salt. In the system <strong>of</strong> n-butyl bromide reacted withsodium phenolate [225], the water molecules form hydrogen bonds with NaOPh as well aswith QOPh, leading to the amount <strong>of</strong> tetrabutylammonium salts in the organic phase <strong>and</strong>in the third liquid phase increasing with the amount <strong>of</strong> NaOPh added, which in turnenhances the overall reaction rate.(d) Organic Solvent <strong>and</strong> the Reaction Temperature. In general, the more polar theorganic solvent the faster is the overall reaction rate in LLPTC due to the increasingsolubility <strong>of</strong> the catalytic intermediate in the organic phase, <strong>and</strong> leading to much easiertransport <strong>of</strong> ion pairs into the solvent to react with the organic substrate. In contrast,in TLPTC, the solubility <strong>of</strong> the catalytic ion pairs in the organic solvent should be lowenough to push the catalyst to form aseparate phase. Thus, asolvent with low polarityor anonpolar one is favorable. Under the same conditions <strong>of</strong> using KBr <strong>and</strong> acatalyst,the reaction rate in dodecane was observed to be much faster than in toluene [227].Increasing the reaction temperature accelerates the reaction rate [221–226,230–233].However, the catalyst existing in the third liquid phase as well as in the organic phaseshould still be maintained. Under strong base conditions in TLPTC, the catalyst <strong>and</strong> theactive intermediate have the tendency to decompose at ahigh temperature, hence, alimitingreaction temperature should be kept in maintaining the third liquid phase.C. Kinetic Modeling for TLPTCYang[234]hasdevelopedatheoreticalmodeltoinvestigatetheeffects<strong>of</strong>masstransfer<strong>and</strong>distribution <strong>of</strong> the catalyst within the third liquid phase <strong>and</strong> organic or aqueous phase ontheoverallreactionrate.Themodelingconsidersthedispersedorganicdropletsurroundedbyaninterfacialcatalystlayerunderagitationconditions,asshowninFig.10.Thistype<strong>of</strong>droplet is similar to some oil/water emulsions in the presence <strong>of</strong> surfactants. The reactantCopyright © 2003 by Taylor & Francis Group, LLC


FIG. 10Conceptual scheme for dispersed droplet <strong>and</strong> third liquid layer in TLPTC.MY in the aqueous phase undergoes a substitution reaction with the organic reactant RX t<strong>of</strong>orm the product RY. Under some appropriate conditions, by introducing extra inorganicsalts or reactants into the system, a separate liquid phase appears <strong>and</strong> is composed <strong>of</strong> PTcatalyst (QX), active intermediate (QY), a little water, <strong>and</strong> organic solvent. The third liquidlayer at the aqueous/organic interface exists if the solubility <strong>of</strong> QY in both the organic <strong>and</strong>aqueous phases is limited. This reaction system is shown in the following scheme:ð174ÞIn the third liquid phase system, the organic phase is considered as the dispersedphase with spherical <strong>and</strong> rigid droplets, <strong>and</strong> with a high distribution coefficient <strong>of</strong> catalystbetween aqueous <strong>and</strong> third phases. The rates <strong>of</strong> change <strong>of</strong> RX in the organic phase <strong>and</strong>QY/QX in the third phase are formulated as follows [234]:Copyright © 2003 by Taylor & Francis Group, LLC


@C orgRX@t@C catQY@t@C catQX@t¼ D RXr 2¼ D QYr 2¼ D QXr 2@r 2 @C org RX@r @r @r 2 @Ccat QY@r @r @r 2 @Ccat QX@r @rfor r r dfor r d r r cfor r D r r cð175Þð176Þð177ÞIn which r d is the radius <strong>of</strong> the organic droplet <strong>and</strong> r c is the radius <strong>of</strong> the organic dropletplus the thickness <strong>of</strong> the catalyst layer. The relationship between r c <strong>and</strong> r d isr c ¼ 1 þ V 1=3catrV d ð178ÞorgThe initial <strong>and</strong> boundary conditions areat t ¼ 0;at r ¼ 0;at r ¼ r d ;C orgRX ¼ C RX;0for 0 r r dC catQX ¼ 0; C catQY ¼ 0 for r D r r c ð179Þ@C orgRX¼ 0@r3 @CD orgRX ¼r dRX@r3 @CD catQYr dr dQY@r3 @CD catQX ¼QX@rk 2 C orgRX Ccat QY¼ k 2 C orgRX Ccat QYk 2 C orgRX Ccat QYThe distribution coefficients <strong>of</strong> QY <strong>and</strong> QX, m QY <strong>and</strong> m QX , are defined asm QY ¼ CcatðsÞ QYC aqðsÞQYAt r ¼ r c ;<strong>and</strong> m QX ¼ CcatðsÞ QXC aqðsÞQXD QY@C catQY@r@CQXcatD QX ¼@r¼ K QYC aqQYK QX ðC catQX1CQYcatm QYm QY C aqQX Þð180Þð181Þð182Þð183Þwhere K QY <strong>and</strong> K QX denote the mass transfer coefficients <strong>of</strong> QY <strong>and</strong> QX, respectively. Inthe aqueous phase, the rates <strong>of</strong> change <strong>of</strong> MY, QY, <strong>and</strong> QX aredC aqMY¼ kdt 1 C aqMY Caq QXdC aq QY¼ kdt 1 C aq3 VcatMY Caq QXK QYr c V aqdC aq QXVcatC catðsÞQXmdtQX C aqQX¼ K QX3r cV aqC aqQY1C catðsÞQYm QYk 1 C aqMY Caq QXð184Þð185Þð186ÞCopyright © 2003 by Taylor & Francis Group, LLC


The initial conditions areAt t ¼ 0;C aqMY ¼ C MY;0C aqQX ¼ C QX;0;C aqQY ¼ 0ð187ÞThe average concentration <strong>and</strong> the conservation <strong>of</strong> PT catalyst Q at any reaction time areC orgRX ¼ 3 r 3 dð rd0V aq C QX;0 ¼ V catr 2 C orgRX drQY þ C QXcat þ Vaq C aqQY þ Caq QXC catð188Þð189ÞEquations (175)–(189) constitute the system <strong>of</strong> PT-catalyzed reactions with the third liquidphase, <strong>and</strong> can be further tendered in dimensionless form <strong>and</strong> solved by finite difference<strong>and</strong> Runge–Kutta methods.Recently, Krueger et al. [235] developed a theoretical model, based on the dispersedorganic phase, for modeling the mass transfer <strong>and</strong> interfacial reactions <strong>of</strong> the bromination<strong>of</strong> benzyl chloride in three-liquid PTC. The reaction occurring at the interface between theinner organic droplet <strong>and</strong> outer shell (or layer) <strong>of</strong> the third phase isr ¼ R: BzCl þ QBr k 1BzBr þ QCl ð190Þ!They assumed that the ion-exchange reaction occurs at the interface between the aqueous<strong>and</strong> the third liquid phase according tor ¼ R þ : Br þ Q þ Ð QBr ð191ÞQCl Ð Q þ þ ClThe governing equations in the dimensionless form for the system are@C BC¼ 1 @@ 2 2 @C BC@ @ð0


tively. When the amount <strong>of</strong> the third liquid phase has little effect on the overall reactionrate, the contribution <strong>of</strong> greater amounts <strong>of</strong> third liquid phase is to increase the masstransfer resistance. This case is similar to the situation <strong>of</strong> a very thin film <strong>of</strong> the third phasewith =R 1. They also proposed an equation [235] for the concentration <strong>of</strong> benzylchloride using the analogous analytical solution <strong>of</strong> heat transfer in terms <strong>of</strong> a mass transferproblem asC BC ¼ X1n¼1A n exp 2 n sinð n Þ n ð199Þwhere the eigenvalues n <strong>and</strong> the coefficients A n are given by1 n cotð n Þ¼N D <strong>and</strong> A n 4sinð ½ nÞ n cosð n ÞŠ2 n sinð2 n Þð200ÞVI.CONCLUSIONIn recent years, researches in PTC have achieved great progress in the development <strong>of</strong> thebasic theory <strong>and</strong> applications; the number <strong>of</strong> publications <strong>and</strong> patents in PTC has grownsteadily during the past decade. Halpern [236] estimated that the 5-year average <strong>of</strong> TBABpatents issued per year was 38 in 1985, 46 in 1990, <strong>and</strong> 57 in 1995. The PTC publicationsfor a 5-year average issued per year are 389 in 1985, 467 in 1990, <strong>and</strong> 484 in 1995. Thisinformation indicates the potential applications <strong>of</strong> PTC in industries either revamping ordeveloping new processes, including the applications <strong>of</strong> PTC in biology.The aspects <strong>of</strong> PTC publications were mostly concentrated in chemistry-based investigationsin past years. On viewing the nature <strong>of</strong> PTC, the presence <strong>of</strong> immiscible phases<strong>and</strong> the transport <strong>of</strong> reacting species between the phases are the basic phenomena; therefore,the mass transfer resistances at the interface <strong>and</strong> within the intraphase should betaken into account in accompanying the chemical reactions for the purpose <strong>of</strong> reactordesign. The engineering analysis in various types <strong>of</strong> PTC, together with other techniquesfor enhancing the overall reactivity, has the advantage <strong>of</strong> realizing the factors influencingthe observed reaction rate, which makes the process design closer to the inherent results. Itis hoped that the review in this chapter could serve to generate more applications <strong>of</strong> PTCin the future.Copyright © 2003 by Taylor & Francis Group, LLC


SYMBOLSaactivity <strong>of</strong> quaternary salt in a solutionA interfacial area between the organic <strong>and</strong> aqueous phase ðm 2 )A s surface area <strong>of</strong> solid particles ðdm 2 )Aliquat 336 tricaprylmethylammonium chlorideCconcentration <strong>of</strong> quaternary saltC j iconcentration <strong>of</strong> component i in phase j (j ¼ org:, aq., cat.)C PhOQ dimensionless concentration <strong>of</strong> PhOQC# number <strong>of</strong> carbon atoms in each <strong>of</strong> the four alkyl chains in the quaternarycationD Q distribution coefficient <strong>of</strong> quaternary cation definedEQX T true extraction constant ðkmol=m 3 Þ 1jhydration number per quaternary salt in the organic phasek app apparent first-order reaction-rate constant (s 1 )k app;0 initial apparent reaction rate constant (1/min)k 2 forward reaction rate constant <strong>of</strong> the aqueous phase (kmol=m 3 minÞ 1k 2 reverse reaction rate constant <strong>of</strong> the aqueous phase ðkmol=m 3 minÞ 1k ireaction rate constantK iequilibrium constantK da dissociation constant <strong>of</strong> QX in the aqueous phase (kmol=m 3 )K do dissociation constant <strong>of</strong> QX in the organic phase (kmol=m 3 )K QY mass transfer coefficient <strong>of</strong> QY (m/min)K QX overall mass transfer coefficient <strong>of</strong> QX (kmol=m 3 min 1 )mdistribution coefficient <strong>of</strong> onium saltM H molecular weight <strong>of</strong> hydrophilic group in quaternary salt (e.g., N þ X )M T molecular weight <strong>of</strong> quaternary saltM TBAB molecular weight <strong>of</strong> TBABMX side product in the aqueous solutionMY aqueous reactantQ þ quaternary cationQX quaternary saltrspatial co-ordinate in radial directionr cmean radius <strong>of</strong> droplet containing organic phase <strong>and</strong> catalyst phase (m)r dmean radius <strong>of</strong> organic droplet (m)R freaction rate per unit area (mol=m 2 sÞRH reactantRX organic reactantR 0 X organic reactantttimeTabsolute temperature (K)TBAB tetra-n-butylammonium bromideTBAC tetra-n-butylammonium chlorideTBAI tetra-n-butylammonium iodideTBA-TBPO tetra-n-butylammonium 2,4,6-tribromophenoxideTBPB tetra-n-butylphosphonium bromideV volume <strong>of</strong> aqueous phase (m 3 )V aq volume <strong>of</strong> aqueous phase (L)volume <strong>of</strong> catalyst phase (L)V catCopyright © 2003 by Taylor & Francis Group, LLC


V org volume <strong>of</strong> organic phase (L)XanionXconversion <strong>of</strong> RX in the organic phaseYproduct yield[ ] molar concentration <strong>of</strong> species in brackets (kmol=m 3 )GreekG nvolume ratio (organic phase/aqueous phase)Hildebr<strong>and</strong> parametersGibbs free energymean activity coefficienteigenvaluesdimensionless spatial co-ordinate in radial directiondimensionless timeSubscripts0, i initial valueicompound iapp apparent valueobs observed valueSuperscriptsaqcatIjorgSðoverbarÞaqueous phasecatalyst phaseinterface between catalyst <strong>and</strong> organic phasesj phase (j ¼ org:, aq., cat).organic phasedroplet surface between catalyst <strong>and</strong> aqueous phasesspecies in the organic phaseCopyright © 2003 by Taylor & Francis Group, LLC


REFERENCES1. CM Starks. J Am Chem Soc 93:195–199, 1971.2. CM Starks, CL Liotta, ME Halpern. <strong>Phase</strong>-<strong>Transfer</strong> <strong>Catalysis</strong>: Fundamental Application <strong>and</strong>Industrial Perspectives. 1st ed. New York: Chapman & Hall, 1994, pp 1–22.3. CM Starks. Modern Perspectives on the <strong>Mechanism</strong>s <strong>of</strong> <strong>Phase</strong>-<strong>Transfer</strong> <strong>Catalysis</strong>. ACSSymposium Series 659. Washington, DC: American Chemical Society, 1997, pp 10–28.4. F Orsini, F Pelizzoni, B Bellini, G Miglierini. Carbohydr Res 301:95–109, 1997.5. D Carrie` re, SJ Meunier, FD Tropper, S Cao, R Roy. J Mol Catal A: Chem 154:9–22, 2000.6. D Albanese, D L<strong>and</strong>ini, M Penso. Chem Commun 2095–2096, 1999.7. YN Belokon, KA Kochetkov, TD Churkina, NS Ikonnikov, AA Chesnokov, OV Larionov,VS Parmar, R Kumar, HB Kagan. Tetrahedron: Asym 9:851–857, 1998.8. B Lygo, J Crosby, JA Peterson. Tetrahedron 40:1385–1388, 1999.9. B Lygo, J Crosby, JA Peterson. Tetrahedron 57:6447–6453, 2001.10. B Lygo, J Crosby, TR Lowdon, PG Wainwright. Tetrahedron 57:2391–2402, 2001.<strong>11.</strong> S Kotha, E Brahmachary. J Org Chem 65:1359–1365, 2000.12. MH Entezari, A Keshavarzi. Ultrasoni Sonochem 8:213–216, 2001.13. H Noguchi, H Tsutsumi, M Komiyama. Chem Commun 2455–2456, 2000.14. VK Krishnakumar, MM Sharma. Ind Eng Chem Proc Des Dev 23:410–413, 1984.15. P Tundo, A Perosa, M Selva, SS Zinovyev. Appl Catal B: Environ 32: L1–L7, 2001.16. P Hodget, R O’Dell, MSK Lee. Polymer 37:1267–1271, 1996.17. LH Tagle, FR Diaz, G Cerda, M Oyarzo, G Pen˜ afiel. Polym Bull 40:35–42, 1998.18. LH Tagle, FR Diaz, C Cares, A Brito. Polym Bull 42:627–634, 1999.19. D Albanese, D L<strong>and</strong>ini, A Maia, M Penso. J Mol Catal A: Chem 150:113–131, 1999.20. FS Sirovski. Org Proc Res Dev 3:437–441, 1999.21. GD Yadav, SS Naik. Org Proc Res Dev 4:141–146, 2000.22. S Desikan, LK Doraiswamy. Chem Eng Sci 55:6119–6127, 2000.23. JP Jayach<strong>and</strong>ran, ML Wang. Appl Catal A: Gen 198:127–137, 2000.24. JJ Hwang, RL Lin, RL Shieh, JJ Jwo. J Mol Catal A: Chem 142:125–139, 1999.25. JAB Satrio, LK Doraiswamy. Chem Eng J 82:43–56, 2001.26. JC Jarrouse. CR Hebd Seances Acad Sci Ser C 1424–1434, 1951.27. EV Dehmlow, SS Dehmlow. <strong>Phase</strong> <strong>Transfer</strong> <strong>Catalysis</strong>. Weinheim: Verlag Chemie, 1993.28. CM Starks, CL Liotta, ME Halpern. <strong>Phase</strong> <strong>Transfer</strong> <strong>Catalysis</strong>, Fundamental, Application<strong>and</strong> Industrial Perspectives. 1st ed. New York: Chapman & Hall, 1994, pp 1–17.29. WP Weber, GW Gokel. <strong>Phase</strong> <strong>Transfer</strong> <strong>Catalysis</strong>: Organic Synthesis. New York: Springer-Verlag, 1977.30. ME Halpern. <strong>Phase</strong>-<strong>Transfer</strong> <strong>Catalysis</strong>: <strong>Mechanism</strong>s <strong>and</strong> Syntheses. ACS Symposium Series659. Washington, DC: American Chemical Society, 1997.31. G. Goldberg. <strong>Phase</strong> <strong>Transfer</strong> <strong>Catalysis</strong>: Selected Problems <strong>and</strong> Applications. Netherl<strong>and</strong>s:Gordon & Breach, 1992.32. N Ohtani, CA Wilkio, A Nigam, SL Regen. Macromolecules 14:516–520, 1981.33. CM Starks. J Am Chem Soc 93: 195–199, 1971.34. M Ellwood, J Griffiths, P Gregory. J Chem Soc, Chem Commun 181, 1980.35. H Iwamoto, T Sonoda, H Kobayash. Tetrahedron Lett 4703–4706, 1983.36. LJ Mathias, RA Vaidya. J Am Chem Soc 108:1093, 1986.37. ML Wang. In: Y Sasson, R Neumann, eds. H<strong>and</strong>book <strong>of</strong> <strong>Phase</strong> <strong>Transfer</strong> <strong>Catalysis</strong>. NewYork: Chapman & Hall, 1997, pp 36–109.38. CS Kuo, JJ Jwo. J Org Chem 57:1991–1995, 1992.39. ML Wang, CC Ou, JJ Jwo. Bull Chem Soc Jpn 67:2949, 1994.40. ML Wang, CC Ou, JJ Jwo. Ind Eng Chem Res 33:2034, 1994.41. S Asai, H Nakamura, W Okada, M Yamada. Chem Eng Sci 50:943–949, 1995.42. J Dockx. Synthesis 441–456, 1973.43. A Gorgoes, A Le Coq. Tetrahedron Lett 4521, 1976.Copyright © 2003 by Taylor & Francis Group, LLC


44. Halpern, HA Zahalka, Y Sasson, M Rabinovitz. J Org Chem 50:5088, 1985.45. M Mąkosza. Pure Appl Chem 43:439–462, 1975.46. M Mąkosza, E Bialecka. Tetrahedron Lett 2:183–186, 1977.47. D L<strong>and</strong>ini, A Maia, F Montanari. J Chem Soc, Chem Comm 112–113, 1977.48. A Bra¨ ndstro¨ m. Adv Phys Org Chem 15:267–330, 1977.49. M Rabinovitz, Y Cohen, M Halpern. Angew Chem, Int Ed Engl 25:960, 1980.50. KJ Evan, HJ Palmer. AIChE Symp Ser 77:104–113, 1981.51. JB Melville, JD Goddard. Chem Eng Sci 40:2207–2215, 1986.52. JB Melville, YC Yortsos, Chem Eng Sci 41:2873–2882, 1986.53. CT Chen, C Hwang, MY Yen. J Chem Eng Jpn 24:284–290, 1991.54. ML Wang, SW Chang. Ind Eng Chem Res 33:1606–1611, 1994.55. ML Wang, SW Chang. Can J Chem Eng 69:340–346, 199156. ML Wang, SW Chang. Ind Eng Chem Res 30:2378–2383, 199157. ML Wang, HM Yang. Chem Eng Sci 46:619–627, 1991.58. ML Wang, HS Wu. Chem Eng Sci 46:509–517, 1991.59. A Bhattacharya. Ind Eng Chem Res 35:645–652, 1996.60. HS Wu. J Chin Inst Chem Eng 25:183–189, 1994.61. HS Wu, SH Jou. J Chem Tech Biotechnol 64:325–330, 1995.62. HS Wu, JJ Lai. J Chin Inst Chem Eng 26:277–283, 1995.63. HS Wu. Chem Eng Sci 51:827–830, 1996.64. HS Wu. Ind Eng Chem. Res 32:1323–1327, 1993.65. ML Wang, HM Yang. Ind Eng Chem Res 30:631–635, 1991.66. ML Wang <strong>and</strong> HM Yang. Ind Eng Chem Res 29:522–526, 1990.67. CM Starks, CL Liotta. <strong>Phase</strong> <strong>Transfer</strong> <strong>Catalysis</strong>: Principles <strong>and</strong> Techniques. New York:Academic Press, 1978, p 40.68. HS Wu. J Chin Inst Chem Eng 27:185–193, 1996.69. HS Wu, SS Meng. AIChE J 43:1309–1318, 1997.70. ML Wang, HS Wu. J Org Chem 55:2344–2350, 1990.71. HM Yang. Chem Eng Commun 179:117–132, 2000.72. HM Yang. Ind Eng Chem Res 37:398–404, 1998.73. WK Fife, Y Xin. J Am Chem Soc 109:1278, 1987.74. RS Juang, SC Liu. Ind Eng Chem Res 36:5296–5301, 1997.75. RS Juang, SC Liu. Ind Eng Chem Res 37:4625–4630, 1998.76. BR Agarwal, RN Diamond. J Phys Chem 67:2785–2792, 1963.77. S Asai, H Nakamura, Y Furuichi. J Chem Eng Jpn 24:653–658, 1991.78. M Cerna, V Bizek, J Stastova, V Rov. Chem Eng Sci 48:99–103, 1993.79. EV Dehmlow, B Vehre. J Chem Res Symp 10:350–351, 1987.80. EV Dehmlow, B Vehre. New J Chem 13:117–119, 1989.81. NA Gibson, DC Weatherburn. Anal Chim Acta 58:149–157, 1972.82. NA Gibson, DC Weatherburn. Anal Chim Acta 58:159–165, 1972.83. HMNH Irvine, A D Damodaran. Anal Chim Acta 53:267–275, 1971.84. HMNH Irvine, A D Damodaran. Anal Chim Acta 53:277–285, 1971.85. HS Wu, RR Fang, SS Feng, KH Hu. J Chin Int Chem Eng 29:99–108, 1998.86. HS Wu, RR Fang, SS Feng, KH Hu. J Mol Catal A: Chem 136:135–146, 1998.87. A Bra¨ ndstro¨ m. Pure Appl Chem 54:1769–1782, 1982.88. MH Abraham. J Phys Org Chem 299–308, 1971.89. MH Abraham, AF Danil de Namor. J Chem Soc, Faraday I 72:955–962, 1976.90. AA Ansari, MR Islam. Can J Chem 66:1720–1727, 1988.91. P Beronius, A Bra¨ ndstro¨ m. Acta Chem Sc<strong>and</strong> A 30:687–691, 197692. MS Tseng. Thermodynamic Properties <strong>of</strong> Quaternary Salts in an Organic Solvent <strong>and</strong>Alkaline Solution. MS thesis, Yuan-Ze University, Taoyuan, Taiwan, 2000.93. JO Bockries, KN Reddy. Modern Electrochemistry. Plenum Press, 1980, pp 261–265.94. T Kenjo, RM Diamond. J Phys Chem 76:2454–2459, 1972.Copyright © 2003 by Taylor & Francis Group, LLC


95. T. Kenjo, RM Diamond. J Inorg Nucl Chem 36:183–188, 1974.96. D L<strong>and</strong>ini, A Maia, G Podda. J Org Chem 47:2264–2268, 1982.97. D L<strong>and</strong>ini, A Maia, F Montanari. J Am Chem Soc 100:2796–2801, 1978.98. D L<strong>and</strong>ini, A Maia, A Rampoldi. J Org Chem 54:328–332, 1989.99. CM Starks, RM Owens. J Am Chem Soc 95:3613–3617, 1973.100. AW Herriott, D Picker. J Am Chem Soc 97:2345–2349, 1975.101. K Gustavii. Acta Pharm Suec 4:233, 1967.102. K Fukunaga, H Shirai, S Ide, M Kimura. Nippon Kagaku Kaishi 1148–1153, 1980.103. FS Sirovski. <strong>Phase</strong>-<strong>Transfer</strong> <strong>and</strong> Micellar <strong>Catalysis</strong> in Two-<strong>Phase</strong> System. ACS SymposiumSeries 659. Washington, DC: American Chemical Society, 1997.104. A Leo, C Hansch, D Elkins. Chem Rev 71:525–563, 1971.105. WC Griffin. J Soc Cosmet Chem 1:311, 1949.106. WC Griffin. Am Perfum Essent Oil Rev 65:26, 1952.107. JT Davies. Proceedings <strong>of</strong> the 2nd International Congress on Surface Activity, 1957, vol. 1,p 426.108. S Nagata. Mixing Principles <strong>and</strong> Applications. New York: Halsted Press, 1975.109. AJ Parker, U Mayer, R Schmid, V Gutmann. J Org Chem 43:1843–1854, 1978.110. J Czapkiewicz, T Czapkiewicz, D Struck. Pol J Chem 52, 2203, 1978.1<strong>11.</strong> RW Taft, MH Abraham, RM Doherty, MJ Kamlet. J Am Chem Soc 107:3105–3110, 1985.112. TB Lin, MY Yeh, YP Shih. Proc PAC Chem Eng Congr 3:178, 1983.113. JB Melville, JD Goddard. Ind Eng Chem Res 27:551, 1988.114. ML Wang, YM Hsieh. J Chem Eng Jpn 26:374–381, 1993.115. SS Leie, RR Bhave, MM Sharma. Chem Eng Sci 38:765–773, 1983.116. ML Wang, HS Wu. J Org Chem 55:2344–2350, 1990.117. SD Naik, LK Doraiswamy. AIChE J 612–645, 1998.118. T Stanley, J Quinn. Chem Eng Sci 42:2313–2324, 1987.119. S Matson. Advances in Catalytic Technologies Seminars, Catalystica. Mountain View, CA,USA, 1988.120. JH Rushton, S Nagata, TB Rooney. AIChE J 10:298–302, 1964.121. S Asai, J Hatanaka, Y Uekawa. J Chem Eng Jpn 16:463–469, 1983.122. S Asai, H Nakamura, Y Furuichi. AIChE J 38:397–403, 1992.123. S Asai, H Nakamura, T Sumita. AIChE J 40:2028, 1994.124. HM Yang, CK Yu, TM Chen. J Chin Inst Chem Eng 29:161–169, 1998.125. A Kiani, RR Bahave, KK Shirkar. J Membr Sci 20:125–145, 1984.126. R Prasad, A Kiani, RR Bhave, KK Sirkar. J Membr Sci 26:79–97, 1986.127. YH Lin. Mass–<strong>Transfer</strong> Behavior <strong>of</strong> Quaternary Salts in Membrane Extractor, MS thesis,Yuan-Ze University, Taoyuan, Taiwan, 2000.128. SV Save, SS Zanwar, VG Pangarkar. Chem Eng Sci 44:1591, 1989.129. N Wakao, S Kaguei. Heat <strong>and</strong> Mass <strong>Transfer</strong> in Packed Beds. New York: Gordon & Breach,1982, p 156.130. WJ McManamey, DKS Multani, JT Davies. Chem Eng Sci 30:1536, 1975.131. EL Cussler. Diffusion Mass <strong>Transfer</strong> in Fluid System. 2nd ed. New York: CambridgeUniversity Press, 1997, pp 48, 355.132. CM Starks. Modern Perspective on the <strong>Mechanism</strong>s <strong>of</strong> <strong>Phase</strong>-<strong>Transfer</strong> <strong>Catalysis</strong>. ACSSymposium Series 659. Washington, DC: American Chemical Society, 1997.133. F Moberg, O Bokman, H Bokman, OG Siegbahn. J Am Chem Soc 113:3663–3667, 1991.134. B Reuben, K Sjoberg. Chemtech 315–320, 1981.135. LL Travlarides, M Stamatoudis. In: TB Drew, GR Cokelet, JW Hooper Jr, T Vermeulen, eds.Advances in Chemical Engineering, vol. <strong>11.</strong> New York: Academic Press, 1981, pp 119–273.136. NN Dutta, GS Patil, Can J Chem Eng 71:802–804, 1993.137. SL Regen. J Am Chem Soc 97:5956–5958, 1975.138. SL Regen. J Am Chem Soc 98:6270–6274, 1976.139. SL Regen. J Org Chem 42:875–879, 1977.Copyright © 2003 by Taylor & Francis Group, LLC


140. SL Regen. Angew Chem, Int Ed Engl 18:421–429, 1979.141. SL Regen, JJ Besse. J Am Chem Soc 101:4059–4063, 1979.142. M Tomoi, WT Ford. J Am Chem Soc 103:3821–3828, 1981.143. M Tomoi, WT Ford. J Am Chem Soc 103:3828–3829, 1981.144. M Tomoi, E Ogawa, Y Hosokawa, H Kakjuchi. J Polym Sci, Polym Chem Ed 20:3421, 1982.145. M Tomoi, E Nakamura, Y Hosokawa, H Kakjuchi. J Polym Sci, Polym Chem Ed 23:49–61,1985.146. M Tomoi, Y Hosokawa, H Kakjuchi. Makromol Chem Rapid Commun 4:227–230, 1983.147. B Ragaini, G Verzella, A Ghigone, G Colombo. In Eng Chem Proc Des Dev 25:878–885,1986.148. V Ragaini, G Colombo, P. Barzhagi. Ind Eng Chem Res 27:1382–1387, 1988.149. V Ragaini, G Colombo, P Barzhagi, E Chiellini, S D’Antone. Ind Eng Chem Res 29:924–928,1990.150. P Schlunt, PC Chau. J Catal 102:348–356, 1986.151. S Desikan, LK Doraiswany. Ind Eng Chem Res 34:3524–3537, 1995.152. WT Ford, M Tomoi. Adv Polym Sci 55:49–104, 1984.153. WT Ford, J Lee, M Tomoi. Macromolecules 15:1246–1251, 1982.154. O Arrad, Y Sasson. J Org Chem 55:2252, 1990.155. P Tundo, P Venturello. J Am Chem Soc 103:856–861, 1981.156. P Tundo, P Venturello, E Angeletti. J Am Chem Soc 104:6551–6555, 1982.157. P Tundo, P Venturello. J Am Chem Soc 101:6606–6613, 1979.158. S Telford, P Schlunt, PC Chau. Macromolecules 19:2435–2439, 1986.159. JF Hradil, C Konak, K Jurek. React Polym 9:81–89, 1988.160. F Svec. Pure Appl Chem 60:377–386, 1988.161. ML Wang, HS Wu. Ind Eng Chem Res 31:490–496, 1992.162. ML Wang, HS Wu. J Polym Sci, Polym Chem 30:1393–1399, 1992.163. ML Wang, HS Wu. Ind Eng Chem Res 31:2238–2242, 1992.164. HS Wu, JF Tang. J Mol Cata, A: Chem 145:95–105, 1999.165. HS Wu, SS Meng S. Can J Chem Eng 77:1146–1153, 1999.166. HS Wu, CS Lee. J Catal 199:217–223, 2001.167. E Ruckenstein, L Hong. J Catal 136:378, 1992.168. E Ruckenstein, JS Park. J Poly Sci C, Polym Lett 26:529, 1988.169. JF Tang. Application <strong>and</strong> <strong>Kinetics</strong> <strong>of</strong> Liquid–Solid–Liquid Triphase Reaction Ether–EsterCompound. MS thesis, Yuan-Ze University, Taoyan, Taiwan, 1998.170. DH Freeman. In: JA Marinsky, ed. Ion Exchange. vol. I. New York: Marcel Dekker, 1966,ch. 5.171. JP Idoux, R Wysocki, S Young, J Turcot, C Ohlman, R Leonard. Synth Commun 13:139,1983.172. Y Okahata, K Ariga, T Seki. J Chem Soc, Chem Commun 920, 1985.173. Y Shan, R Kang, W Li. Ind Eng Chem Res 28:1289, 1989.174. ML Wang, HM Yang. Ind Eng Chem Res 30:2384–2390, 1991.175. ML Wang, HM Yang. Ind Eng Chem Res 30:1868–1875, 1992.176. ML Wang, HM Yang. J Chin Inst Chem Eng 23:200–208, 1992.177. R Aris. The Mathematical Theory <strong>of</strong> Diffusion <strong>and</strong> Reaction in Permeable <strong>Catalysis</strong>. vol. I.London: Oxford University Press, 1975, pp 119–126.178. PF Marconi, WT Ford. J Catal 83:160–167, 1983.179. DY Cha. React Polym 5:269–279, 1987.180. HS Wu, SS Meng. Chem Eng Sci 53:4073–4084, 1998.181. ML Wang, HS Wu. Ind Eng Chem Res 29:2137–2142, 1990.182. M Tomoi, Y Hosokawa, H Kakjuchi. J Polym Sci, Polym Chem Ed 22:1243–1250, 1984.183. CM Starks, CL Liotta, ME Halpern. <strong>Phase</strong>-<strong>Transfer</strong> <strong>Catalysis</strong>: Fundamental Application <strong>and</strong>Industrial Perspectives. 1st ed. New York: Chapman & Hall, 1994, pp 108–1<strong>11.</strong>Copyright © 2003 by Taylor & Francis Group, LLC


184. Y Sasson. In: Y Sasson, R Neumann, ed. H<strong>and</strong>book <strong>of</strong> <strong>Phase</strong> <strong>Transfer</strong> <strong>Catalysis</strong>. 1st ed.London: Blackie Academic & Pr<strong>of</strong>essional, 1997, pp 111–134.185. A Esikova. In: Y Sasson, R Neumann, ed. H<strong>and</strong>book <strong>of</strong> <strong>Phase</strong> <strong>Transfer</strong> <strong>Catalysis</strong>. 1st ed.London: Blackie Academic & Pr<strong>of</strong>essional, 1997, pp 1–35.186. CL Liotta, J Berkner, J Wright, B Fair. <strong>Mechanism</strong>s <strong>and</strong> Applications <strong>of</strong> Solid–Liquid <strong>Phase</strong>-<strong>Transfer</strong> <strong>Catalysis</strong>. ACS Symposium Series 659. Washington, DC: American ChemicalSociety, 1997, pp 29–40.187. A Lo´ pez, MM Man˜ as, R Pleixats, A Roglans, J Ezquerra, C Pedregal. Tetrahedron 52:8365–8386, 1996.188. A Lo´ pez, R Pleixats. Tetrahedron: Asym 9:1967–1977, 1998.189. FS Sirovski, ER Berlin, SA Mulyashov, EA Bobrova, ZI Batrakova, DF Dankovskaya. OrgProc Res Dev 1:253–256, 1997.190. T Perrad, JC Plaquevent, JR Desmurs, D He´ brault. Org Lett 2:2959–2962, 2000.191. HM Yang, TM Chen. J Chin Inst Chem Eng 29:367–374, 1998.192. S Cohen, A Zoran, Y Sasson. Tetrahedron Lett 39:9815–9818, 1998.193. R Guilet, J Berlan, O Louisnard, J Schwartzentruber. Ultrason Sonochem 5:21–25, 1998.194. A Knochel, G Rudolph. Tetrahedron Lett 3739–3740, 1974.195. AW Herriot, D Picker. J Am Chem Soc 97:2345–2349, 1975.196. BP Czech, MJ Pugia, RA Bartsch. Tetrahedron 41:5439–5444, 1985.197. KH Wong, APW Wai. J Chem Soc, Perkin Trans II 317–321, 1983.198. MC V<strong>and</strong>er Zwan, FW Hartner. J Org Chem 43:2655–2657, 1978.199. GD Yadav, MM Sharma. Ind Eng Chem Proc Des Dev 20:385–390, 1981.200. HA Yee, HJ Palmer, SH Chen. Chem Eng Progr 83:33–39, 1987.201. HM Yang, HE Wu. Ind Eng Chem Res 37:4536–4541, 1998.202. HM Yang, CM Wu. J Mol Catal A: Chem 153:83–91, 1999.203. JB Melville, JD Goddard. Chem Eng Sci 40:2207–2215, 1985.204. JB Melville, JD Goddard. Ind Eng Chem Res 27:551–555, 1988.205. JB Melville, YC Yortsos. Chem Eng Sci 41:2873–2882, 1986.206. SD Naik, LK Doraiswamy. Chem Eng Sci 52:4533–4546, 1997.207. CS Chu. <strong>Kinetics</strong> for Synthesizing Benzyl Phenyl Ether via Solid–Liquid <strong>Phase</strong>–<strong>Transfer</strong><strong>Catalysis</strong> over Polyethylene Glycol <strong>and</strong> Quaternary Ammonium Salts. MS thesis, NationalChung Hsing University, Taichung, Taiwan, 2001.208. SS Yufit, GV Kryshtal, IA Esikova. Adsorption on Interfaces <strong>and</strong> Solvation in <strong>Phase</strong>-<strong>Transfer</strong> <strong>Catalysis</strong>. ACS Symposium Series 659. Washington, DC: American ChemicalSociety, 1997, pp 52–67.209. F Sirovski, C Reichardt, M Gorokhova, S Ruban, E Stoikova. Tetrahedron 55:6363–6374,1999.210. BV Er<strong>of</strong>eev. Dokl Akad Nauk SSSR 52:515–518, 1946.2<strong>11.</strong> AY Rozovskii. Heterogeneous Chemical Reactions: <strong>Kinetics</strong> <strong>and</strong> Macrokinetics. Moscow:Nauka, 1980, pp. 180–300.212. SS Yuift, SS Zinovyev. Tetrahedron 55:6319–6328, 1999.213. SS Yufit, IA Esikova. J Phys Org Chem 4:336–340, 1991.214. IA Esikova, SS Yufit. J Phys Org Chem 4:149–157, 1991.215. IA Esikova, SS Yufit. J Phys Org Chem 4:341–345, 1991.216. HM Yang, PI Wu. Appl Catal A: Gen 209:17–26, 2001.217. PI Wu. <strong>Kinetics</strong> for Solid–Liquid <strong>and</strong> Solid–Liquid–Liquid <strong>Phase</strong>-<strong>Transfer</strong> <strong>Catalysis</strong>: I.Etherification <strong>of</strong> Halo-ester II. Esterification <strong>of</strong> Aliphatic Dicarboxylate. MS thesis,National Chung Hsing University, Taichung, Taiwan, 2000.218. HM Yang, CM Wu, HE Wu. J Chem Technol Biotechnol 75:387–393, 2000.219. HM Yang, PI Wu, CM Li. Appl Catal A: Gen 193:129–137, 2000.220. C Lepetit, M Che. J Mol Catal A: Chem 100:147–160, 1995.221. R Neumann, Y Sasson. J Org Chem 49:3448–3451, 1984.222. DH Wang, HS Weng. Chem Eng Sci 43:2019–2024, 1988.Copyright © 2003 by Taylor & Francis Group, LLC


223. SD Naik, LK Doraiswamy. AIChE J 44:612–646, 1998.224. T Ido, M Saiki, S Goto. Kagaku Kogaku Ronbunshu 15:403–410, 1989.225. D Mason, S Magdassi, Y Sasson. J Org Chem 56:7229–7232, 1991.226. DH Wang, HS Weng. Chem Eng Sci 50:3477–3486, 1995.227. T Ido, T Yamamoto, G Jin, S Goto. Chem Eng Sci 52:3511–3520, 1997.228. GD Yadav, CA Reddy. Ind Eng Chem Res 38:2245–2253, 1999.229. G Jin, T. Ido, S. Goto. J Chem Eng Jpn 32:417–423, 1999.230. HC Hsiao, HS Weng. Ind Eng Chem Res 38:2911–2918, 1999.231. HC Hsiao, SM Kao, HS Weng. Ind Eng Chem Res 39:2772–2778, 2000.232. HS Weng, CM Wang, DH Wang. Ind Eng Chem Res 36:3613–3618, 1997.233. T Ido, T Susaki, G Jin, S Goto. Appl Catal A:Gen 201:139–143, 2000.234. HM Yang. J Chin Inst Eng 21:399–408, 1998.235. JJ Krueger, MD Amiridis, HJ Ploehn. Ind Eng Chem Res 40:3158–3163, 2001.236. ME Halpern. Recent Trends in Industrial <strong>and</strong> Academic <strong>Phase</strong>-<strong>Transfer</strong> <strong>Catalysis</strong>. ACSSymposium Series 659. Washington, DC: American Chemical Society, 1997, pp 1–9.Copyright © 2003 by Taylor & Francis Group, LLC

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!