Sunteți pe pagina 1din 9

Thin-Film Reactors

Thin-Film Reactors
For other industrial reactors and their applications, see Stirred-Tanc and Loop Reactors, Tubular Reactors, Fixed-Bed Reactors, Fluidized-Bed Reactors, Bubble Columns, Three-Phase Trickle-Bed Reactors, Reaction Columns, Metallurgical Furnaces, and Biochemical Engineering. Bernhard Gutsche, Henkel KGaA, D sseldorf, Federal Republic of Germany u Christoph Breucker, Henkel KGaA, D sseldorf, Federal Republic of Germany u Gunter Panthel, Henkel KGaA, D sseldorf, Federal Republic of Germany u . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Applications of Thin-Film Reactors . Designs and Operating Modes of ThinFilm Reactors . . . . . . . . . . . . . . . . Thin-Film Reactors for Industrial Sulfonation . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . .

1. 1.1. 1.2. 1.3. 1.4. 1.5.

Design of Thin-Film Reactors Gas Liquid Interface . . . . . Fluid Dynamics . . . . . . . . . . Heat Transfer . . . . . . . . . . . Mass Transfer . . . . . . . . . . . Models . . . . . . . . . . . . . . .

2 2 3 4 5 5

2. 3. 4. 5.

6 6 7 9

Symbols (see also Principles of Chemical Design Engineering and Model Reactors and Their Design Equations) A a b cp d E ( ) L m n q T t v V Ha Nu Pr Re Sc Sh Ta transfer area interfacial area per unit reaction volume constant isobaric heat capacity diameter distribution function of residence time length mass ow number of tubes heat temperature time volume throughput Hatta number Nusselt number Prandtl number Reynolds number Schmidt number Sherwood number Taylor number heat-transfer coefcient mass-transfer coefcient lm thickness dimensionless residence time coefcient of thermal conduction kinematic viscosity residence time

angular velocity Subscripts

c G h i if l L t W

coolant gas hydraulic (diameter, see Eq. 17) internal (diameter of the inner cylinder of the reactor, see Eq. 17) interface laminar liquid tube, turbulent wall

A thin-lm apparatus, with or without rotating internals, is typically used for mild, nondestructive distillation, absorption, or desorption (stripping) in process engineering. Starting from this eld of application, it has also acquired an important role in gas liquid reactions.

1. Design of Thin-Film Reactors


Thin-lm reactors ( falling-lm reactors and lm reactors with mechanical wiping) are contacting devices for gas liquid systems. The liquid phase is in the form of a thin lm that runs down a vertical wall while gas ows over the liquid surface cocurrently or countercurrently. In the rising-lm reactor, gas owing upward at a high

c 2005 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 10.1002/14356007.b04 329

Thin-Film Reactors Entrainment or breakup of the lm cannot be permitted since heat transfer would be impaired and local overheating might degrade the product in the case of exothermic reactions. Options for calculating entrainment are summarized in [3]. To prevent entrainment, tubes must be installed precisely vertically, and the tube surface must be adequately nished. For highly viscous liquids, a rotating wiper system can be used to maintain a uniform lm, afford better mixing of the liquid, and promote heat transfer (wiped-lm reactor). Thin-lm reactors have a very small specic surface area for mass transfer between gas and liquid compared to other reactors. The mass transfer area A is governed only by reactor geometry (e.g., reactor length L, tube diameter d t , number of tubes n). For a laminar lm of thickness
A =n L (dt ) (1)

rate causes a thin liquid lm to ascend (gas and liquid cocurrent). Thin-lm reactors are used chiey when strongly endothermic or exothermic reactions require a very high rate of heat transfer. These reactors are operated continuously. Thin-lm reactors consist of one tube, or a number of tubes bundled as in a shell-and-tube heat exchanger, cooled or heated from the outside (Fig. 1). In another design [2], two large concentric pipes form an annular clearance. The liquid is distributed over both surfaces of the annular space and reacts with the gas owing in the free volume between the pipes.

The increase in lm surface area due to wave formation in the turbulent-ow regime is negligibly small. Since << d t in most cases, the geometric area of the tube can be used for practical calculations:
A =nL dt (2)

Hence the interfacial area per unit reaction volume is


a= n Ldt 4 = n Ld 2 dt t 4 (3)

If the geometric interfacial area is divided by the volume of liquid in the lm, the ratio is
aL = nLdt 1 = nLdt (4)

Figure 1. Schematic of a thin-lm reactor [1] a) Tube wall (heat transfer); b) Liquid lm; c) Gas liquid interface (reaction)

1.2. Fluid Dynamics


With regard to back mixing, the thin-lm reactor is similar to the tubular reactor. Neither the gas nor the liquid phase shows much back mixing, and both can be treated as plug ows. Thin-lm reactors are therefore also used when a narrow residence-time distribution enhances selectivity and yield. Figure 2 shows typical residence-time distributions for a number of reactors [1].

1.1. Gas Liquid Interface


The thin-lm reactor is the only type of gas liquid contact apparatus in which both gas and liquid are present as continuous phases and, in fact, the dispersion of either phase must be prevented.

Thin-Film Reactors

satisfactory agreement with experimental data [5]. The following equation generalizes the results:
=b1
2 3L g 1/3

Reb2

(7)

Table 1 gives the parameters for different ow regimes.

Table 1. Parameters from Equation (7) for different ow regimes ReL < 80 100 420 > 400 > 420 b1 1 0.93 0.435 0.369 b2 1/3 1/3 7/12 1/2 Author Nusselt Kapitza Zhivaikin Feind References [4] [6] [7] [8]

Figure 2. Residence-time distribution for different reactors [1] a) Laminar falling lm (ideal); b) Turbulent falling lm; c) Turbulent tubular reactor; d) Film reactor with mechanically produced lm

The residence times of the two phases are very short, and the liquid holdup is very small. Both quantities can be calculated easily from the lm thickness and the reactor geometry. The ratio of gas and liquid throughputs in this type of reactor is usually in the range
10 m3 m3 <VG /VL < 2000 3 m3 m (5)

1.3. Heat Transfer


The great advantage of thin-lm reactors is their large heat-transfer area per unit liquid volume. The heat-transfer area is roughly the same as the mass-transfer area (see Eqs. 3 and 4). Such reactors are employed, accordingly, for strongly exothermic and endothermic reactions. Because these reactions are often accompanied by the absorption of a gaseous reactant, the evolved heat of an exothermic reaction and heat of absorption raise the temperature of the liquid, especially at the gas liquid interface, thereby decreasing the solubility of the gas in the liquid. The absorption and the reaction rate therefore, strongly depend on the rate of heat removal. A compilation and comparison of literature data on heat transfer for liquid lms are given by Yuksel [9], who determined the heat-transfer and mass-transfer coefcients independently of each other and measured the lm surface temperature by a contactless method. He conrmed that the turbulence in the lm at the surface is damped, resulting in an additional thermal resistance near the lm surface besides the resistance in the boundary layer of liquid next to the wall (Fig. 3).

If constraints, such as residence time or optimal reaction conditions, make it necessary to alter this ratio, an inert gas component can be added to dilute the reactant gas. Falling lms can be analyzed in three characteristic ow regimes. If the Reynolds number for the lm is dened as
ReL = VL dt L (6)

the boundaries of the regimes are as follows: ReL 25 laminar, smooth lm 25 < ReL < 400 transition regime, laminar lm with waves ReL 400 turbulent lm For the laminar regime, the velocity distribution in the falling lm can be used in calculating lm thickness [4]. The calculated thickness is in

Thin-Film Reactors marked viscosity differences in the lm, the temperature dependence of the material parameters is also introduced.
N u = N u4 +N u4 t l
0.25

Pr P rW

0.25

(13)

Computational equations for heat transfer in mechanically produced lms can be found [1].

1.4. Mass Transfer


Because of the limited transfer area and the short gas and liquid residence times, thin-lm reactors are best employed only for fast (0.3 < Ha < 3) or instantaneous (Ha > 3) reactions. Most of the reaction occurs in the boundary layer, while the contribution of the bulk liquid phase is negligible. The rate-determining step is accordingly gas-side mass transfer. For the gas-side local mass-transfer coefcient G , which is a function of the Sherwood number, Braun and Hiby give the following dimensionless formulas [11]: Cocurrent ow:
Sh = 0.18 Re0.4 Re0.16 Sc0.44 G L G 1 + 6.4 L dt
0.75

Figure 3. Temperature prole of a thin-lm reactor for an exothermic reaction a) Interface; b) Wall

On the basis of heat-transfer measurements in condensation on a falling lm with a variety of liquids (2.6 Pr < 52), plus literature data, Muller [10] reports the following practical formulas for the mean heat transfer over the lm length: Laminar lm:
N ul = 0.606Re2/9 (8)

(14)

Countercurrent ow:
Sh = 0.015Re0.75 Re0.16 Sc0.44 G L G 1 + 5.2 L dt
0.75

Turbulent lm:
N ut = 0.0105Re0.35 P r0.43 (9)

(15)

where
Nu= if /g3 if (10)

q Tif TW

(11)

The last term in Equations (14) and (15) takes care of the fact that the mass-transfer coefcient declines to a constant nal value as reactor length increases. For the wiped-lm reactor with rotating annular clearance that is operated with gas and liq uid ascending cocurrently [12], Brostrom gives the following empirical equation for gas-side mass transfer [13]:
Sh = 0.079Re0.67 T a0.16 Sc0.52 (16)

if cp Pr= if

(12)

where the Taylor number Ta is dened as


Ta= dh d
0.5

The fourth powers of the laminar and turbulent contributions can be added. Because poorer heat transfer in higher-viscosity liquids gives rise to

di dh 2 2

(17)

d h is the hydraulic diameter of the annulus and d i is the internal diameter of the inner cylinder.

Thin-Film Reactors Calculations for mass transfer in the liquid phase can be found, e.g., in [1] and [9]. The discussion above leads to the concentration curve of the gaseous component shown in Figure 4. The entire reaction takes place in the lm. The concentration of dissolved gas in the bulk liquid approaches zero. Absorption of the gaseous components should therefore be enhanced by the reaction. Jana and coworkers [14] nd correspondingly high enhancement factors, up to 12. Mann [15], on the other hand, measures enhancement factors < 1. This disagreement can be explained by strong thermal effects, leading to a temperature rise at the gas liquid interface (Fig. 3) and a correspondingly low solubility of the reactant gas in the reacting liquid.

sulfonation of alkylbenzenes) [18], [15]. Jana and coworkers present a general model for the countercurrent absorption of gas in a laminar lm with rst-order reaction [14]. For the absorption of carbon dioxide in dilute sodium hydroxide solution, Haimour and Sandall show that the chemosorption process can be precalculated [19]. Guti rrez-Gonzalez and coworkers proe pose a special model for the reactions of sulfur trioxide with alkylbenzenes in a cocurrent setup [20]. The model is compared with the authors experiments. A simple model for cocurrent ascending sulfonation in a special wiped-lm re actor is described by Brostrom [21].

2. Applications of Thin-Film Reactors


In industry, lm reactors are employed for a variety of processes involving either strongly exothermic reactions, where quasi-isothermal operation is achieved by cooling the reaction tube, or endothermic reactions, where evaporation or desorption can be achieved only by supplying heat. In the second case, thermodynamic constraints (vapor liquid or sorptive equilibrium) must be accounted for. For both applications a high selectivity can be attained only with precise temperature control avoiding side reactions and sequential reactions. The most important industrial use of the falling-lm reactor without rotating internals is the sulfonation of organic products with sulfur trioxide, but other applications are also mentioned frequently in publications and patents. Table 2 summarizes the elds in which these reactors are used. The reactions in Table 2 are divided up with respect to the direction of mass transfer: absorption with reaction (industrially the most important form), reaction with superimposed absorption/desorption (where a tower-type reactor is more often employed [22]), and reaction with desorption of a byproduct, e.g., water formed in esterications. In reaction-with-desorption processes used in polymer chemistry, lm reactors with rotating internals (i.e., vertical or horizontal thin-layer evaporators) are frequently used. A good survey of fast gas liquid reactions is found in [23].

Figure 4. Concentration of gas in a thin-lm reactor for a fast reaction a) Interface; b) Wall

1.5. Models
Many publications have dealt with models of absorption and reaction in lms, especially for the thin-lm reactor. Villadsen and Nielsen [16], [17] give a survey of existing models (examples are the chlorination of decane and the

Thin-Film Reactors

Table 2. Practical applications of gas liquid reactions in thin lms without sulfonation/sulfation (see Chap. 4 for data on sulfonation with SO3 ) Absorption Chlorination of hydrocarbons [15] Absorption and desorption countercurrent esterication of fatty acids with short-chain alcohols [26] Desorption esterication, transesterication [26]

Oxidation [23] Nitration [24] Neutralization, (e.g., of CO2 ) [19] Alkylation of isobutane with butene in solvent containing catalyst [25]

polycondensation (polyesters) [1] Vilsmeier reaction [27] polymerization (polyamide 66) [28]

3. Designs and Operating Modes of Thin-Film Reactors


Many possible designs and modes of operation are described in the literature. Most of them nd use only in the laboratory. Figures 5,6,7,8 show examples of designs. The falling-lm reactor illustrated in Figure 5 as a cocurrent multitube apparatus [29], requires good liquid distribution to form the liquid lm. In addition, the viscosity must not increase much during the reaction. The surface area of the single-tube fallinglm apparatus is optimally utilized in the form of an annulus reactor [30], [31], since both the inner wall of the outer pipe and the outer wall of the inner pipe serve as reaction surfaces. Internals that swirl the gas stream give improved mass transfer, at least in the laboratory [32]. A corresponding thin-lm reactor design, operated with gas and liquid ascending cocurrently, is the rising-lm reactor [33]. For high-viscosity products, lm reactors with rotating internals are employed. The internals can be installed either vertically or horizontally. Figure 6 shows an annulus reactor (Votator) [12] in which the gas and the liquid are fed in from below cocurrently. To improve mass transfer inside the liquid lm, the rotating inner pipe is tted with many small pins.

Figure 5. Ballestra falling-lm sulfonation reactor [29]

Figure 6. Berol Kemi annulus reactor with rotating inner pipe (Votator) [12]

Thin-Film Reactors A wiped-lm reactor for countercurrent operation is shown in Figure 7. This device is used for reactions with superimposed absorption and possibly desorption, e.g., esterication of viscous fatty acids with short-chain alcohols. Figure 8 illustrates a wiped-lm reactor with specially designed wiper elements, intended for use with high-viscosity liquids (up to 80 Pa s) [34]. This reactor is employed for reactions with desorption in polymer production.

4. Thin-Film Reactors for Industrial Sulfonation


The literature contains many publications on the preparation of surfactants by sulfonating and sulfating organic feedstocks with sulfur trioxide. Along with the chlorination of hydrocarbons, the sulfonation of alkylbenzenes and the sulfation of fatty alcohols have served as test systems for the development of lm reactors. Selected publications on sulfonation of different organic feedstocks with sulfur trioxide in lm reactors of various types are listed below:
Petroleum distillates, crude oils -Olens Isoolens Alkylbenzene Ethoxylated alkylphenol Fatty alcohol Ethoxylated fatty alcohol Methyl esters [35] [30], [36], [37] [38] [12], [36], [37], [39] [36] [32], [33], [36], [37] [30], [33], [36], [37] [37], [40], [41]

Figure 7. Thin-lm reactor for countercurrent operation (schematic) a) Heating jacket; b) Rotor

The most important sulfonation feedstocks today are alkylbenzenes, -olens, and fatty alcohols and their derivatives: Alkylbenzenes

-Olens
RCH=CH2 + SO3 RCH=CHCH2 n SO3 H ( )

Fatty alcohols
ROH + SO3 ROSO3 H

Fatty alcohol ethoxylates


) ( ) ROC2 H4 n OH + SO3 ROC2 H4 n OSO3 H ( Figure 8. Buss-SMS wiped-lm reactor for high-viscosity liquids (polymer production) [34]

Thin-Film Reactors reaction. This effect, however, certainly results from further dilution of the sulfur trioxide gas by air, and the reduction in concentration could just as well be performed before the gas inlet to the reactor.

The sulfonic acids or alkylsulfuric acids obtained are then neutralized to the desired alkali salts (mainly sodium salts). Organic feedstocks are very sensitive to formation of degradation products (oversulfonation); these byproducts lead to discoloration or impurities in the products. The feedstocks are therefore reacted with gaseous sulfur trioxide cocurrently, in proportions as close as possible to equimolar. Sulfonation reactions are strongly exothermic (H 150 kJ/mol), and the heat of reaction must be removed quickly from the liquid phase. Not all products of the reaction with sulfur trioxide are stable so they must be neutralized immediately after the sulfonation/sulfation reaction. Examples of such products are alkylsulfuric acids formed from fatty alcohols and sulfur trioxide. Thin-lm reactors ideally satisfy the two principal requirements for sulfonation reactions: short residence time of reactants in the reactor and good removal of heat of reaction from the reactor. Both single-tube and multitube reactors are used for this process. In single-tube reactors, diluted sulfur trioxide gas is led into the annular clearance formed by two concentric pipes. The liquid lm ows down the respective inner and outer surfaces of the pipes. Each pipe has intensive cooling on the side away from the liquid lm. A typical representative of such single-tube reactors is the Chemithon reactor [42]. Multitube reactors have frequently been used since the 1980s and have up to 120 individual reaction tubes, the organic feedstock being fed to the inner surface of each. After the liquid lm has formed, diluted sulfur trioxide gas is fed to each reaction tube. The heat of reaction is removed just as if the reactor were a vertical shell-and-tube heat exchanger. Figure 5 shows the Ballestra multitube reactor [29]. The Mazzoni multitube reactor has a special feature [43]: After the feedstock is admitted, air is blown into each reaction tube to spread out the liquid, thus promoting the formation of a uniform lm of organic liquid. Only when this step is complete is sulfur trioxide gas admitted. As shown in Figure 9, each reaction tube has a double wall so that it can be cooled separately. It is sometimes argued that the equalizing air would retard the transport of sulfur trioxide to the reaction site and thus reduce local overheating due to the fast

Figure 9. Schematic diagram of one tube in a Mazzoni multitube reactor [43]

Along with the development of hardware, more stringent product quality standards have also led to changes in reaction conditions. For example, the sulfur trioxide concentration in the process gas has been decreased to allow better handling of the heat of reaction. The dewpoint of the process air has also been lowered so that the formation of oleum from residual moisture can be minimized. Table 3 lists the ranges of the most important reaction parameters.
Table 3. Operating parameters of sulfonation reactors Parameter Number of tubes Tube diameter, mm Tube length, mm V G , m3 /h (STP) mL , kg/h Dewpoint of air, C SO3 , vol % Single-tube reactor 1 150 1000 ca. 3000 2000 7000 150 3500 60 to 80 27 Multitube reactor 4 200 25.4 ca. 6000 20 70 per tube 18 36 per tube 60 to 80 27

Thin-Film Reactors

5. References
1. P. Trambouze, H. van Landeghem, J. P. Wauquier: Les r acteurs chimiques, Editions e Technip, Paris 1984. 2. Lion Corp., DE 2 629 009, 1976 (T. Ogoshi, Y. Miyawaki, F. Kondo, S. Sakurai). 3. R. G. Krebs, E. U. Schl nder, Chem. Eng. u Process. 18 (1984) 341 356. 4. W. Nusselt, VDI Z. 60 (1916) 541 546. 5. H. Brauer: Grundlagen der Einphasen- und Mehrphasenstr mungen, Sauerl nder, o a Aarau-Frankfurt/M. 1971. 6. P. L. Kapitza, Z. Exp. Theor. Phys. 18 (1948) 3 18. 7. L. Y. Zhivaikin, Int. Chem. Eng. 2 (1962) 337 341. 8. K. Feind: Str mungsuntersuchungen bei o Gegenstrom von Riesellmen in lotrechten Rohren, VDI-Forschungsh. 481 (1960). 9. M. L. Y ksel: W rme- und Stoff bergang bei u a u der nicht isothermen Absorption am Riesellm, Fortschr. Ber., VDI Reihe 3, no. 133, VDI-Verlag, D sseldorf 1987. u 10. J. M ller: Einu der Stoffwerte auf den u W rme- bergang bei der Kondensation am a u Riesellm, lecture held at GVC Fachausschu W rme- und Stoff- bertragung, a u April 25, 1991, Freudenstadt, to be published in Chem. Eng. Process. 11. D. Braun, J. W. Hiby, Chem. Ing. Tech. 42 (1970) 345 349. 12. Berol Kemi, DE-OS 2 523 875, 1975 (A. Brostr m). o 13. A. Brostr m, Trans. Inst. Chem. Eng. 53 o (1975) 26 28. 14. S. C. Jana et al., Chem. Eng. Commun. 79 (1989) 27 37. 15. R. Mann, H. Moyes, AIChE J. 23 (1977) 17 23. 16. P. H. Nielsen, J. Villadsen, Chem. Eng. Sci. 38 (1983) 1439 1453. 17. J. Villadsen, P. H. Nielsen, Chem. Eng. Sci. 41 (1986) 1655 1671. 18. G. R. Johnson, B. L. Crynes, Ind. Eng. Chem. Process Des. Dev. 13 (1974) 6 14. 19. N. Haimour, O. C. Sandall, AIChE J. 29 (1983) 277 281. 20. J. Guti` rrez-Gonz` lez, C. Mans-Teixid , J. e a o Costa L` pez, J. Dispersion Sci. Technol. 6 o (1985) 303 315.

21. A. Brostr m, Trans. Inst. Chem. Eng. 53 o (1975) 29 33. 22. B. Schleper, B. Gutsche, J. Wnuck, L. Jeromin, Chem. Ing. Tech. 62 (1990) 226 227. 23. Y. T. Shah: Gas-Liquid-Solid Reactor Design, McGraw-Hill, New York 1979. 24. A. Beenackers, W. van Swaaij, Chem. Eng. J. (Lausanne) 15 (1978) 25 38. 25. I. Pervez, C. A. Karagiozov, C. B. Boyadjev, Chem. Eng. Sci. 46 (1991) 1589 1594. 26. VEB Deutsches Hydrierwerk Rodleben, FR 1 534 622, 1967. 27. B. Covelli, U. Lattmann, W rme-Stoff bertrag. 12 (1979) 233 241. a u 28. D. D. Steppan, M. F. Doherty, M. F. Malone, Ind. Eng. Chem. Res. 29 (1990) 2012 2020. 29. Ballestra Chimica, DE 3 006 791, 1980 (G. Moretti, S. Noe). 30. I. Yamane, J. Am. Oil Chem. Soc. 55 (1978) 81 86. 31. Lion Corp., US 3 839 391, 1971 (R. Suzuki, S. Tanimori, S. Toyoda, T. Ogoshi); US 3 925 441, 1973 (S. Toyoda, T. Ogoshi, M. Maruyama, Y. Miyawaki). 32. Yu. D. Panaev, D. I. Zemenkov, V. G. Pravdin, V. N. Sokolov, Int. Chem. Eng. 17 (1977) 337 338. 33. K. Takei, K. Tsuto, S. Miyamoto, J. Wakatsuki, JAOCS J. Am. Oil Chem. Soc. 62 (1985) 341 347. 34. Buss-SMS, High Viscosity Technology, company brochure, Butzbach 1976. 35. T. L. Ashcraft, R. K. Saunders, JCPT J. Can. Pet. Technol. 19 (1980) 47 53. 36. W. Fricke, Tenside 4 (1967) 317 320. 37. M. Ballestra, G. Moretti, Chim. Oggi 7 8 (1984) 41 43. 38. D. W. Roberts, D. L. Williams, Tenside Deterg. 22 (1985) 193 195. 39. A. Ujhidi, B. Babos, L. Far dy, Chem. Tech. a (Leipzig) 18 (1966) 652 654. 40. K. H. Schmid, H. Baumann, W. Stein, H. Dolhaine, Henkel Ref. 21 (1985) 11 19. 41. T. Ogoshi, Y. Miyawaki, JAOCS J. Am. Oil Chem. Soc. 62 (1985) 331 335. 42. The Chemithon Corp., Sulfonation Process Equipment, company brochure, Seattle. 43. Mazzoni, DE 2 107 968, 1971 (A. Lanteri).

S-ar putea să vă placă și