There may be a need for appropriate mixing, control of flow dis-tribution and residence time, contacting between the reactants sometimes in the presence of a catalyst or biocatalyst, rem
Trang 2Copyright © 2008, 1997, 1984, 1973, 1963, 1950, 1941, 1934 by The McGraw-Hill Companies, Inc All rights reserved Manufactured in the UnitedStates of America Except as permitted under the United States Copyright Act of 1976, no part of this publication may be reproduced or distributed
in any form or by any means, or stored in a database or retrieval system, without the prior written permission of the publisher
0-07-154226-4
The material in this eBook also appears in the print version of this title: 0-07-151142-3
All trademarks are trademarks of their respective owners Rather than put a trademark symbol after every occurrence of a trademarked name, we usenames in an editorial fashion only, and to the benefit of the trademark owner, with no intention of infringement of the trademark Where such designations appear in this book, they have been printed with initial caps
McGraw-Hill eBooks are available at special quantity discounts to use as premiums and sales promotions, or for use in corporate training programs.For more information, please contact George Hoare, Special Sales, at george_hoare@mcgraw-hill.com or (212) 904-4069
TERMS OF USE
This is a copyrighted work and The McGraw-Hill Companies, Inc (“McGraw-Hill”) and its licensors reserve all rights in and to the work Use of thiswork is subject to these terms Except as permitted under the Copyright Act of 1976 and the right to store and retrieve one copy of the work, you maynot decompile, disassemble, reverse engineer, reproduce, modify, create derivative works based upon, transmit, distribute, disseminate, sell, publish
or sublicense the work or any part of it without McGraw-Hill’s prior consent You may use the work for your own noncommercial and personal use;any other use of the work is strictly prohibited Your right to use the work may be terminated if you fail to comply with these terms
THE WORK IS PROVIDED “AS IS.” McGRAW-HILL AND ITS LICENSORS MAKE NO GUARANTEES OR WARRANTIES AS TO THEACCURACY, ADEQUACY OR COMPLETENESS OF OR RESULTS TO BE OBTAINED FROM USING THE WORK, INCLUDING ANYINFORMATION THAT CAN BE ACCESSED THROUGH THE WORK VIA HYPERLINK OR OTHERWISE, AND EXPRESSLY DISCLAIMANY WARRANTY, EXPRESS OR IMPLIED, INCLUDING BUT NOT LIMITED TO IMPLIED WARRANTIES OF MERCHANTABILITY ORFITNESS FOR A PARTICULAR PURPOSE McGraw-Hill and its licensors do not warrant or guarantee that the functions contained in the work willmeet your requirements or that its operation will be uninterrupted or error free Neither McGraw-Hill nor its licensors shall be liable to you or anyone else for any inaccuracy, error or omission, regardless of cause, in the work or for any damages resulting therefrom McGraw-Hill has noresponsibility for the content of any information accessed through the work Under no circumstances shall McGraw-Hill and/or its licensors be liablefor any indirect, incidental, special, punitive, consequential or similar damages that result from the use of or inability to use the work, even if any ofthem has been advised of the possibility of such damages This limitation of liability shall apply to any claim or cause whatsoever whether such claim
or cause arises in contract, tort or otherwise
DOI: 10.1036/0071511423
Trang 4REACTOR CONCEPTS
Reactor Types 19-4
Classification by Mode of Operation 19-4
Classification by End Use 19-7
Reactor Tracer Responses 19-15
Understanding Reactor Flow Patterns 19-16
Connecting RTD to Conversion 19-17
Segregated Flow 19-18
Early versus Late Mixing—Maximum Mixedness 19-18
Reaction and Mixing Times 19-20
SINGLE-PHASE REACTORS
Liquid Phase 19-20 Homogeneous Catalysis 19-20 Gas Phase 19-21 Supercritical Conditions 19-21 Polymerization Reactors 19-21
FLUID-SOLID REACTORS
Heterogeneous Catalysts 19-25 Catalytic Reactors 19-27 Wire Gauzes 19-27 Monolith Catalysts 19-27 Fixed Beds 19-30 Moving Beds 19-33 Fluidized Beds 19-33 Slurry Reactors 19-36 Transport Reactors 19-36 Multifunctional Reactors 19-36 Noncatalytic Reactors 19-36 Rotary Kilns 19-36 Vertical Kilns 19-36
19-1
Section 19 Reactors*
Carmo J Pereira, Ph.D., MBA DuPont Fellow, DuPont Engineering Research and
Technology, E I du Pont de Nemours and Company; Fellow, American Institute of Chemical
Engineers
Tiberiu M Leib, Ph.D Principal Consultant, DuPont Engineering Research and
Technol-ogy, E I du Pont de Nemours and Company; Fellow, American Institute of Chemical Engineers
*The contributions of Stanley M Walas, Ph.D., Professor Emeritus, Department of Chemical and Petroleum Engineering, University of Kansas (Fellow, American Institute of Chemical Engineers), author of this section in the seventh edition, are acknowledged.
The authors of the present section would like to thank Dennie T Mah, M.S.Ch.E., Senior Consultant, DuPont Engineering Research and Technology, E I du Pont
de Nemours and Company (Senior Member, American Institute of Chemical Engineers; Member, Industrial Electrolysis and Electrochemical Engineering; Member, The Electrochemical Society), for his contributions to the “Electrochemical Reactors” subsection; and John Villadsen, Ph.D., Senior Professor, Department of Chem- ical Engineering, Technical University of Denmark, for his contributions to the “Bioreactors” subsection We acknowledge comments from Peter Harriott, Ph.D., Fred
H Rhodes Professor of Chemical Engineering (retired), School of Chemical and Biomolecular Engineering, Cornell University, on our original outline and on the ject of heat transfer in packed-bed reactors The authors also are grateful to the following colleagues for reading the manuscript and for thoughtful comments: Thomas
sub-R Keane, DuPont Fellow (retired), DuPont Engineering Research and Technology, E I du Pont de Nemours and Company (Senior Member, American Institute of Chemical Engineers); Güray Tosun, Ph.D., Senior Consultant, DuPont Engineering Research and Technology, E I du Pont de Nemours and Company (Senior Mem- ber, American Institute of Chemical Engineers); and Nitin H Kolhapure, Ph.D., Senior Consulting Engineer, DuPont Engineering Research and Technology, E I.
du Pont de Nemours and Company (Senior Member, American Institute of Chemical Engineers).
Copyright © 2008, 1997, 1984, 1973, 1963, 1950, 1941, 1934 by The McGraw-Hill Companies, Inc Click here for terms of use
Trang 5SOME CASE STUDIES
Trang 6Nomenclature and Units
In this section, the concentration is represented by C Mass balance accounting in terms of the number of moles and the fractional conversion is discussed in Sec 7 and can be very useful The rate of reaction is r; the flow rate in moles is N a ; the volumetric flow rate is V′; reactor volume is V r Several equations are presented with- out specification of units Use of any consistent unit set is appropriate.
Following is a listing of typical nomenclature expressed in SI and U.S Customary System units Specific definitions and units are stated at the place of application
in this section.
U.S Customary
Deff Effective diffusion coefficient m 2 /s ft 2 /s
D e Effective dispersion coefficient m 2 /s ft 2 /s
E(t) Residence time distribution
E(t r) Normalized residence time
distribution
f a Fraction of A remaining
unconverted, C a /C a0 or n a/a0
F(t) Age function of tracer
h Heat-transfer coefficient kJ(s⋅m 2 ⋅°C) Btu(h⋅ft 2 ⋅°F)
first-order reaction
Pe Peclet number for dispersion
PFR Plug flow reactor
q Heat flux, reaction order,
or impeller-induced flow
r Rate of reaction per
unit volume, radius
u(t) Unit step input
U Overall heat-transfer coefficient kJ(s⋅m 2 ⋅°C) Btu(h⋅ft 2 ⋅°F)
v Volumetric flow rate
v ij Stoichiometric coefficients
U.S Customary
ε Void fraction in a packed bed, particle porosity
η Effectiveness factor of porous catalyst
Trang 7A chemical reactor is a controlled volume in which a chemical reaction
can occur in a safe and controllable manner A reactor typically is a
piece of equipment; however, it can also be a product (such as a
coat-ing or a protective film) One or more reactants may react together at
a desired set of operating conditions, such as temperature and
pres-sure There may be a need for appropriate mixing, control of flow
dis-tribution and residence time, contacting between the reactants
(sometimes in the presence of a catalyst or biocatalyst), removal (or
addition) of heat, and integration of the reactor with the rest of the
downstream process Depending on the nature of the rate-limiting
step(s), a reactor may serve primarily as a holding tank, a heat
exchanger, or a mass-transfer device Chemical reactions generate
desired products and also by-products that have to be separated and
disposed A successful commercial unit is an economic balance of all
these factors A variety of reactor types are used in the chemical,
petrochemical, and pharmaceutical industries Some of these reactors
are listed in Table 19-1 They include gas, liquid, or multiphase batch
reactors, stirred tank reactors, and tubular rectors
There are a number of textbooks on chemical reaction engineering
Davis and Davis (Fundamentals of Chemical Reaction Engineering,
McGraw-Hill, 2003) provide a lucid discussion of kinetics and
princi-ples A more comprehensive treatment together with access to
Chemical Reaction Engineering, 3d ed., Prentice-Hall, 1999) A chemistry-oriented perspective is provided by Schmidt (The Engi- neering of Chemical Reactions, Oxford University Press, 1999) The
book by Froment and Bischoff provides a thorough discussion of tor analysis and design A practical manual on reactor design and
reac-scale-up is by Harriott (Chemical Reactor Design, Marcel Dekker, 2003) Levenspiel (Chemical Reaction Engineering, 3d ed., Wiley,
1999) was among the first to present a phenomenological discussion offundamentals The mathematical underpinnings of reactor modeling
are covered by Bird et al (Transport Phenomena, 2d ed., Wiley, 2002).
This section contains a number of illustrations and sketches from
books by Walas (Chemical Process Equipment Selection and Design, Butterworths, 1990) and Ullmann [Encyclopedia of Chemical Tech- nology (in German), vol 3, Verlag Chemie, 1973, pp 321–518].
Mathematical models may be used to design reactors and analyzetheir performance Detailed models have mainly been developed forlarge-scale commercial processes A number of software tools are nowavailable This chapter will discuss some of the reactors used commer-cially together with how mathematical models may be used For addi-tional details, a number of books on reactor analysis cited in this sectionare available The discussion will indicate that logical choices aimed atmaximizing reaction rate and selectivity for a given set of kinetics canlead to rational reactor selection While there has been progress inrecent years, reactor design and modeling are largely an art
REACTOR CONCEPTS
Since a primary purpose of a reactor is to provide desirable conditions
for reaction, the reaction rate per unit volume of reactor is important in
analyzing or sizing a reactor For a given production rate, it determines
the reactor volume required to effect the desired transformation The
residence time in a reactor is inversely related to the term space
veloc-ity (defined as volumetric feed rate/reactor volume) The fraction of
reactants converted to products and by-products is the conversion The
fraction of desired product in the material converted on a molar basis is
referred to as selectivity The product of conversion and the fractional
selectivity provides a measure of the fraction of reactants converted to
product, known as yield The product yield provides a direct measure of
the level of (atom) utilization of the raw materials and may be an
impor-tant component of operating cost A measure of reactor utilization
called space time yield (STY) is the ratio of product generation rate to
reactor volume When a catalyst is used, the reactor has to make
prod-uct without major process interruptions The catalyst may be
homoge-neous or heterogehomoge-neous, and the latter can be a living biological cell A
key aspect of catalyst performance is the durability of the active site.
Since a chemical or biochemical process has a number of unit
opera-tions around the reactor, it is often beneficial to minimize the variability
of reactant and product flows This typically means that the reactor is
operated at a steady state Interactions between kinetics, fluid flow,
transport resistances, and heat effects sometimes result in multiple
steady states and transient (dynamic) behavior Reactor dynamics can
also result in runaway behavior, where reactor temperature continues
to increase until the reactants are depleted, or wrong-way behavior,
where reducing inlet temperature (or reactant flow rate) can result in
temperature increases farther downstream and a possible runaway
Since such behavior can result in large perturbations in the process and
possibly safety issues, a reactor control strategy has to be implemented.
The need to operate safely under all conditions calls for a thorough
analysis to ensure that the reactor is inherently safe and that all possible
unsafe outcomes have been considered and addressed Since various
solvents may be used in chemical processes and reactors generate both
products and by-products, solvent and by-product emissions can cause
emission and environmental footprint issues that must be considered.
Reactor design is often discussed in terms of independent and
dependent variables Independent variables are choices such as
reac-tor type and internals, catalyst type, inlet temperature, pressure, and
fresh feed composition Dependent variables result from independent
variable selection They may be constrained or unconstrained
Con-strained dependent variables often include pressure drop (limited due
to compressor cost), feed composition (dictated by the composition ofthe recycle streams), temperature rise (or decline), and local andeffluent composition The reactor design problem is often aimed atoptimizing independent variables (within constraints) to maximize anobjective function (such as conversion and selectivity)
Since the reactor feed may contain inert species (e.g., nitrogen andsolvents) and since there may be unconverted feed and by-products inthe reactor effluent, a number of unit operations (distillation, filtration,etc.) may be required to produce the desired product(s) In practice,the flow of mass and energy through the process is captured by aprocess flow sheet The flow sheet may require recycle (of unconvertedfeed, solvents, etc.) and purging that may affect reaction chemistry.Reactor design and operation influence the process and vice versa
REACTOR TYPES
Reactors may be classified according to the mode of operation, theend-use application, the number of phases present, whether (or not) acatalyst is used, and whether some other function (e.g., heat transfer,separations, etc.) is conducted in addition to the reaction
Classification by Mode of Operation
Batch Reactors A “batch” of reactants is introduced into the
reactor operated at the desired conditions until the target conversion
is reached Batch reactors are typically tanks in which stirring of thereactants is achieved using internal impellers, gas bubbles, or a pump-around loop where a fraction of the reactants is removed and exter-nally recirculated back to the reactor Temperature is regulated viainternal cooling surfaces (such as coils or tubes), jackets, reflux con-densers, or pump-around loop that passes through an exchanger.Batch processes are suited to small production rates, to long reactiontimes, to achieve desired selectivity, and for flexibility in campaigningdifferent products
Continuous Reactors Reactants are added and products removed
continuously at a constant mass flow rate Large daily production ratesare mostly conducted in continuous equipment
A continuous stirred tank reactor (CSTR) is a vessel to which
reac-tants are added and products removed while the contents within thevessel are vigorously stirred using internal agitation or by internally (orexternally) recycling the contents CSTRs may be employed in series or
in parallel An approach to employing CSTRs in series is to have a large
19-4
Trang 8TABLE 19-1 Residence Times and/or Space Velocities in Industrial Chemical Reactors*
Residence time or
(raw materials) Type phase Catalyst T,°C P, atm velocity page† Acetaldehyde (ethylene, air) FB L Cu and Pd chlorides 50–100 8 6–40 min [2] 1, [7] 3 Acetic anhydride (acetic acid) TO L Triethylphosphate 700–800 0.3 0.25–5 s [2]
Acrolein (formaldehyde, acetaldehyde) FL G MnO, silica gel 280–320 1 0.6 s [1] 1 384, [7] 33 Acrylonitrile (air, propylene, ammonia) FL G Bi phosphomolybdate 400 1 4.3 s [3] 684, [2] 47 Adipic acid (nitration of cyclohexanol) TO L Co naphthenate 125–160 4–20 2 h [2] 51, [7] 49 Adiponitrile (adipic acid) FB G H 3 BO 3 370–410 1 3.5–5 s [1] 2 152,
Alkylate (i-C4 , butenes) CST L H 2 SO 4 5–10 2–3 5–40 min [4] 223
Allyl chloride (propylene, Cl 2 ) TO G NA 500 3 0.3–1.5 s [1] 2 416, [7] 67
7,800 GHSV
10,000 GHSV
Aniline (nitrobenzene, H 2 ) B L FeCl 2 in H 2 O 95–100 1 8 h [1] 3 289
Aniline (nitrobenzene, H 2 ) FB G Cu on silica 250–300 1 0.5–100 s [7] 82
Aspirin (salicylic acid, acetic anhydride) B L None 90 1 >1 h [7] 89
815 GHSV [9] 109
Benzoic acid (toluene, air) SCST LG None 125–175 9–13 0.2–2 h [7] 101
34,000 GHSV Butadiene sulfone (butadiene, SO 2 ) CST L t-Butyl catechol 34 12 0.2 LHSV [1] 5 192
i-Butane (n-butane) FB L AlCl 3 on bauxite 40–120 18–36 0.5–1 LHSV [4] 239, [7] 683
Butanols (propylene hydroformylation) FB L PH 3 -modified 150–200 1,000 100 g L⋅h [1] 5 373
Co carbonyls Butanols (propylene hydroformylation) FB L Fe pentacarbonyl 110 10 1 h [7] 125
Caprolactam (cyclohexane oxime) CST L Polyphosphoric 80–110 1 0.25–2 h [1] 6 73, [7] 139
acid Carbon disulfide (methane, sulfur) Furn G None 500–700 1 1.0 s [1] 6 322, [7] 144 Carbon monoxide oxidation (shift) TU G Cu-Zn or Fe 2 O 3 390–220 26 4.5 s [6] 44
7,000 GHSV
Coking, delayed (drum, 100 ft max height) B LG None 500–440 4 0.3–0.5 ft/s [1] 10 8
vapor Cracking, fluid catalytic Riser G Zeolite 520–540 2–3 2–4 s (14) 353
Cracking, hydro (gas oils) FB LG Ni, SiO 2 , Al 2 O 3 350–420 100–150 1–2 LHSV [11]
Cracking (visbreaking residual oils) TU LG None 470–495 10–30 450 s, 8 LHSV [11]
Cumene hydroperoxide (cumene, air) CST L Metal porphyrins 95–120 2–15 1–3 h [7] 191
Cyclohexane (benzene, H 2 ) FB G Ni on Al 2 O 3 150–250 25–55 0.75–2 LHSV [7] 201
Cyclohexanol (cyclohexane, air) SCST LG None 185–200 48 2–10 min [7] 203
Cyclohexanone (cyclohexanol) MT G Cu on pumice 250–350 1 4–12 s [8] (1963) Cyclopentadiene (dicyclopentadiene) TJ G None 220–300 1–2 0.1–0.5 LHSV [7] 212
Dibutylphthalate (phthalic anhydride, butanol) B L H 2 SO 4 150–200 1 1–3 h [7] 227
Diethylketone (ethylene, CO) TO L Co oleate 150–300 200–500 0.1–10 h [7] 243
Dimethylsulfide (methanol, CS 2 ) FB G Al 2 O 3 375–535 5 150 GHSV [7] 266
3.3 LHSV [8] (1938) Dodecylbenzene (benzene, propylene tetramer) CST L AlCl 3 15–20 1 1–30 min [7] 283
Ethanol (ethylene, H 2 O) FB G H 3 PO 4 300 82 1,800 GHSV [2] 356, [7] 297 Ethyl acetate (ethanol, acetic acid) TU, CST L H 2 SO 4 100 1 0.5–0.8 LHSV [10] 45, 52, 58 Ethyl chloride (ethylene, HCl) TO G ZnCl 2 150–250 6–20 2 s [7] 305
1,880 GHSV [6] 13
Ethylene, propylene chlorohydrins (Cl 2 , H 2 O) CST LG None 30–40 3–10 0.5–5 min [7] 310, 580 Ethylene glycol (ethylene oxide, H 2 O) TO LG 1% H 2 SO 4 50–70 1 30 min [2] 398
Ethylene glycol (ethylene oxide, H 2 O) TO LG None 195 13 1 h [2] 398
Formaldehyde (methanol, air) FB G Ag gauze 450–600 1 0.01 s [2] 423
Glycerol (allyl alcohol, H 2 O 2 ) CST L H 2 WO 4 40–60 1 3 h [7] 347
19-5
Trang 9Residence time or
(raw materials) Type phase Catalyst T,°C P, atm velocity page†
3,000 GHSV Hydrodesulfurization of naphtha TO LG Co-MO 315–500 20–70 1.5–8 LHSV [4] 285,
125 WHSV [6] 179,
[9] 201
Isoprene (i-butene, formaldehyde) FB G HCl, silica gel 250–350 1 1 h [7] 389
Maleic anhydride (butenes, air) FL G V 2 O 5 300–450 2–10 0.1–5 s [7] 406
Methanol (CO, H 2 ) FB G ZnO, Cr 2 O 3 350–400 340 5,000 GHSV [7] 421
Methanol (CO, H 2 ) FB G ZnO, Cr 2 O 3 350–400 254 28,000 GHSV [3] 562
3.1 LHSV Methyl chloride (methanol, Cl 2 ) FB G Al 2 O 3 gel 340–350 1 275 GHSV [2] 533
Methyl ethyl ketone (2-butanol) FB G ZnO 425–475 2–4 0.5–10 min [7] 437
Methyl ethyl ketone (2-butanol) FB G Brass spheres 450 5 2.1 s [10] 284
13 LHSV Nitrobenzene (benzene, HNO 3 ) CST L H 2 SO 4 45–95 1 3–40 min [7] 468
Nitromethane (methane, HNO 3 ) TO G None 450–700 5–40 0.07–0.35 s [7] 474
Phenol (chlorobenzene, steam) FB G Cu, Ca phosphate 430–450 1–2 2 WHSV [7] 522
900 GHSV
Phthalic anhydride (o-xylene, air) MT G V 2 O 5 350 1 1.5 s [3] 482, 539, [7] 529 Phthalic anhydride (naphthalene, air) FL G V 2 O 5 350 1 5 s [9] 136, [10] 335 Polycarbonate resin (bisphenol-A, phosgene) B L Benzyltriethylammonium 30–40 1 0.25–4 h [7] 452
chloride Polyethylene TU L Organic peroxides 180–200 1,000–1,700 0.5–50 min [7] 547
Polyethylene TU L Cr 2 O 3 , Al 2 O 3 , SiO 2 70–200 20–50 0.1–1,000 s [7] 549
Propionitrile (propylene, NH 3 ) TU G CoO 350–425 70–200 0.3–2 LHSV [7] 578
Reforming of naphtha (H 2 /hydrocarbon = 6) FB G Pt 490 30–35 3 LHSV [6] 99
t-Butyl methacrylate (methacrylic acid, i-butene) CST L H 2 SO 4 25 3 0.3 LHSV [1] 5 328
Toluene diisocyanate (toluene diamine, phosgene) B LG None 200–210 1 7 h [7] 657
Tricresyl phosphate (cresyl, POCl 3 ) TO L MgCl 2 150–300 1 0.5–2.5 h [2] 850, [7] 673 Vinyl chloride (ethylene, Cl 2 ) FL G None 450–550 2–10 0.5–5 s [7] 699
Aldehydes (diisobutene, CO) CST LG Co Carbonyl 150 200 1.7 h [12] 173
Allyl alcohol (propylene oxide) FB G Li phosphate 250 1 1.0 LHSV [15] 23
NOxpollutant (with NH 3 ) FB G V 2 O 5 ⋅TiO 2 300–400 1–10 [14] 332
Automobile emission control M G Pt/Rh/Pd/Al 2 O 3 350–500 1 20,000 GHSV [16] 69
Nitrogen oxide emission control M G V 2 O 5 -WO 3 /TiO 2 300–400 1 4–10,000 [16] 306
GHSV Carbon monoxide and hydrocarbon emission M G Pt-Pd/Al 2 O 3 500–600 1 80–120,000 [16] 334
Ozone control from aircraft cabins M G Pd/Al 2 O 3 130–170 1 ~10 6 GHSV [16] 263
Vinyl acetate (ethylene + CO) MT LG Cu-Pd 130 30 1 h L, 10 s G [12] 140
*Abbreviations: reactors: batch (B), continuous stirred tank (CST), fixed bed of catalyst (FB), fluidized bed of catalyst (FL), furnace (Furn.), monolith (M), tubular (MT), semicontinuous stirred tank (SCST), tower (TO), tubular (TU) Phases: liquid (L), gas (G), both (LG) Space velocities (hourly): gas (GHSV), liquid (LHSV), weight (WHSV) Not available, NA To convert atm to kPa, multiply by 101.3.
multi-†1 J J McKetta, ed., Encyclopedia of Chemical Processing and Design, Marcel Dekker, 1976 to date (referenced by volume).
2 W L Faith, D B Keyes, and R L Clark, Industrial Chemicals, revised by F A Lowenstein and M K Moran, John Wiley & Sons, 1975.
3 G F Froment and K B Bischoff, Chemical Reactor Analysis and Design, John Wiley & Sons, 1979.
4 R J Hengstebeck, Petroleum Processing, McGraw-Hill, New York, 1959.
5 V G Jenson and G V Jeffreys, Mathematical Methods in Chemical Engineering, 2d ed., Academic Press, 1977.
6 H F Rase, Chemical Reactor Design for Process Plants, Vol 2: Case Studies, John Wiley & Sons, 1977.
7 M Sittig, Organic Chemical Process Encyclopedia, Noyes, 1969 (patent literature exclusively).
8 Student Contest Problems, published annually by AIChE, New York (referenced by year).
9 M O Tarhan, Catalytic Reactor Design, McGraw-Hill, 1983.
10 K R Westerterp, W P M van Swaaij, and A A C M Beenackers, Chemical Reactor Design and Operation, John Wiley & Sons, 1984.
11 Personal communication (Walas, 1985).
12 B C Gates, J R Katzer, and G C A Schuit, Chemistry of Catalytic Processes, McGraw-Hill, 1979.
13 B E Leach, ed., Applied Industrial Catalysts, 3 vols., Academic Press, 1983.
14 C N Satterfield, Heterogeneous Catalysis in Industrial Practice, McGraw-Hill, 1991.
15 C L Thomas, Catalytic Processes and Proven Catalysts, Academic Press, 1970.
16 Heck, Farrauto, and Gulati, Catalytic Air Pollution Control: Commercial Technology, Wiley-Interscience, 2002.
Trang 10cylindrical tank with partitions: feed enters the first compartment and
over (or under) flows to the next compartment, and so on The
compo-sition is maintained as uniform as possible in each individual
compart-ment; however, a stepped concentration gradient exists from one
CSTR to the next When the reactants have limited solubility
(miscibil-ity) and a density difference, the vertical staged reactor with
counter-current operation may be used Alternatively, each CSTR in a series or
parallel configuration can be an independent vessel Examples of
stirred tank reactors with heat transfer are shown in Fig 19-1
A tubular flow reactor (TFR) is a tube (or pipe) through which
reac-tants flow and are converted to product The TFR may have a varying
diameter along the flow path In such a reactor, there is a continuous
gra-dient (in contrast to the stepped gragra-dient characteristic of a
CSTR-in-series battery) of concentration in the direction of flow Several tubular
reactors in series or in parallel may also be used Both horizontal and
ver-tical orientations are common When heat transfer is needed, individual
tubes are jacketed or a shell-and-tube construction is used The reaction
side may be filled with solid catalyst or internals such as static mixers (to
improve interphase contact in heterogeneous reactions or to improve
heat transfer by turbulence) Tubes that have 3- to 4-in diameter and are
several miles long may be used in polymerization service
Large-diame-ter vessels, with packing (or trays) used to regulate the residence time in
the reactor, may also be used Some of the configurations in use are axial
flow, radial flow, multishell with built-in heat exchangers, and so on
A reaction battery of CSTRs in series, although both mechanically
and operationally more complex and expensive than a tubular reactor,
provides flexibility Relatively slow reactions are best conducted in a
stirred tank reactor battery A tubular reactor is used when heat
trans-fer is needed, where high pressures and/or high (or low) temperatures
occur, and when relatively short reaction times suffice
Semibatch Reactors Some of the reactants are loaded into the
reactor, and the rest of the reactants are fed gradually Alternatively,
one reactant is loaded into the reactor, and the other reactant is fed
continuously Once the reactor is full, it may be operated in a batch
mode to complete the reaction Semibatch reactors are especially
favored when there are large heat effects and heat-transfer capability
is limited Exothermic reactions may be slowed down and
endother-mic reactions controlled by limiting reactant concentration In
biore-actors, the reactant concentration may be limited to minimize toxicity
Other situations that may call for semibatch reactors include control
of undesirable by-products or when one of the reactants is a gas of
lim-ited solubility that is fed continuously at the dissolution rate
Classification by End Use Chemical reactors are typically used for
the synthesis of chemical intermediates for a variety of specialty (e.g.,agricultural, pharmaceutical) or commodity (e.g., raw materials for poly-
mers) applications Polymerization reactors convert raw materials to
polymers having a specific molecular weight and functionality The ference between polymerization and chemical reactors is artificially
dif-based on the size of the molecule produced Bioreactors utilize (often
genetically manipulated) organisms to catalyze biotransformations eitheraerobically (in the presence of air) or anaerobically (without air present)
Electrochemical reactors use electricity to drive desired reactions
Exam-ples include synthesis of Na metal from NaCl and Al from bauxite ore Avariety of reactor types are employed for specialty materials synthesisapplications (e.g., electronic, defense, and other)
Classification by Phase Despite the generic classification by
oper-ating mode, reactors are designed to accommodate the reactant phasesand provide optimal conditions for reaction Reactants may be fluid(s) or
solid(s), and as such, several reactor types have been developed phase reactors are typically gas- (or plasma- ) or liquid-phase reactors Two-phase reactors may be gas-liquid, liquid-liquid, gas-solid, or liquid- solid reactors Multiphase reactors typically have more than two phases
Single-present The most common type of multiphase reactor is a solid reactor; however, liquid-liquid-solid reactors are also used The clas-sification by phases will be used to develop the contents of this section
gas-liquid-In addition, a reactor may perform a function other than reaction
alone Multifunctional reactors may provide both reaction and mass
transfer (e.g., reactive distillation, reactive crystallization, reactive branes, etc.), or reaction and heat transfer This coupling of functionswithin the reactor inevitably leads to additional operating constraints onone or the other function Multifunctional reactors are often discussed
mem-in the context of process mem-intensification The primary driver for functional reactors is functional synergy and equipment cost savings
multi-REACTOR MODELING
As discussed in Sec 7, chemical kinetics may be mathematicallydescribed by rate equations Reactor performance is also amenable toquantitative analysis The quantitative analysis of reaction systems isdealt with in the field of chemical reaction engineering
The level of mathematical detail that can be included in the analysisdepends on the level of understanding of the physical and chemicalprocesses that occur in a reactor As a practical matter, engineeringdata needed to build a detailed model for some new chemistry typicallyare unavailable early in the design phase Reactor designers may usesimilarity principles (e.g., dimensionless groups), rules of thumb, trendanalysis, design of experiments (DOE), and principal-componentanalysis (PCA) to scale up laboratory reactors For hazardous systems
in which compositional measurements are difficult, surrogate tors such as pressure or temperature may be used As more knowl-edge becomes available, however, a greater level of detail may beincluded in a mathematical model A detailed reactor model maycontain information on vessel configuration, stoichiometric relation-ships, kinetic rate equations, correlations for thermodynamic andtransport properties, contacting efficiency, residence time distribu-tion, and so on
indica-Models may be used for analyzing data, estimating performance,reactor scale-up, simulating start-up and shutdown behavior, and con-trol The level of detail in a model depends on the need, and this isoften a balance between value and cost Very elaborate models are jus-tifiable and have been developed for certain widely practiced andlarge-scale processes, or for processes where operating conditions areespecially critical
Modeling Considerations A useful reactor model allows the
user to predict performance or to explore uncertainties not easily orcost-effectively investigated through experimentation Uncertaintiesthat may be explored through modeling may include scale-up options,explosion hazards, runaway reactions, environmental emissions, reac-tor internals design, and so on As such, the model must contain an
optimal level of detail (principle of optimal sloppiness) required to
meet the desired objective(s) For example, if mixing is critical to formance, the model must include flow equations that reflect the role
per-of mixing If heat effects are small, an isothermal model may be used
(a)
FIG 19-1 Stirred tank reactors with heat transfer (a) Jacket (b) Internal coils.
(c) Internal tubes (d) External heat exchanger (e) External reflux condensor (f)
Fired heater (Walas, Reaction Kinetics for Chemical Engineers, McGraw-Hill,
1959.)
Trang 11A key aspect of modeling is to derive the appropriate momentum,
mass, or energy conservation equations for the reactor These
bal-ances may be used in lumped systems or derived over a differential
volume within the reactor and then integrated over the reactor
vol-ume Mass conservation equations have the following general form:
(19-1)The general form for the energy balance equation is
(19-2)The model defines each of these terms Solving the set of equations
provides outputs that can be validated against experimental
observa-tions and then used for predictive purposes Mathematical models for
ideal reactors that are generally useful in estimating reactor
perfor-mance will be presented Additional information on these reactors is
available also in Sec 7
Batch Reactor Since there is no addition or removal of reactants,
the mass and energy conservation equations for a batch reactor with a
constant reactor volume are
V r r(C,T) + V r = 0 (19-3)
−qA k − V r(−∆Hr )r(C,T) + V r ρc p = 0 (19-4)
where qA kis the addition (or removal) of heat from the reactor Mean
values of physical properties are used in Eqs (19-3) and (19-4) For an
isothermal first-order reaction r(C,T) = kC, the mass and energy
equations can be combined and the solution is
Typically batch reactors may have complex kinetics, mixing, and
heat-transfer issues In such cases, detailed momentum, mass, and energy
balance equations will be required
Semibatch Reactor Feed is added for a fixed time, and the
reac-tion proceeds as the feed is added The reactor equareac-tions governing
the feed addition portion of the process are
in Eq (19-6) yields
After feed addition is completed, the reactor may be operated in a
batch mode In this case, Eqs (19-3) and (19-4) may be used with the
concen-Ideal Continuous Stirred Tank Reactor In an ideal CSTR,
reactants are fed into and removed from an ideally mixed tank As aresult, the concentration within the tank is uniform and identical tothe concentration of the effluent The mass and energy conservationequations for an ideal constant-volume or constant-density CSTR
with constant volumetric feed rate V′ may be written as
V ′C0= V′C + V r r(C,T) + V r (19-11)
V ′ρc p T0= −Q(T) + V′ρc p T − V r(−∆H)r(C,T) + Vr ρc p (19-12)
where Q(T) represents any addition or removal of heat from the
reac-tor and mean values of physical properties are used For example, if
heat is transferred through the reactor wall, Q(T) = A k U(T c −T), where
A k is the heat-transfer area, U is the overall heat-transfer coefficient, and T cis the temperature of the heat-transfer fluid
The above ordinary differential equations (ODEs), Eqs (19-11)
and (19-12), can be solved with an initial condition For an isothermal
first-order reaction and an initial condition, C(0)= 0, the linear ODEmay be solved analytically At steady state, the accumulation term iszero, and the solution for the effluent concentration becomes
Since the contents of an ideal CSTR are perfectly mixed, the sion within the reactor is infinite In practice, CSTRs may not be ide-ally mixed In such cases, the reactor may be modeled as having afraction of the feed α in bypass and a fraction β of the reactor volumestagnant The material balance is
disper-C = αC0+ (1 − α)C1 (19-14)(1− α)V′C0= (1 − α)V′C1+ (1 − β)kV r C1 (19-15)
where C1is the concentration leaving the active zone of the tank
Elimination of C1will relate the input and overall output tions For a first-order reaction,
The two parameters α and β may be expected to depend on reactorinternals and the amount of agitation
Plug Flow Reactor A plug flow reactor (PFR) is an idealized
tubular reactor in which each reactant molecule enters and travelsthrough the reactor as a “plug,” i.e., each molecule enters the reactor
at the same velocity and has exactly the same residence time As aresult, the concentration of every molecule at a given distance down-stream of the inlet is the same The mass and energy balance for a dif-
ferential volume between position V r and V r + dV rfrom the inlet may
be written as partial differential equations (PDEs) for a
constant-density system:
+ V′c pρ − (−∆H)r(C,T) + c pρ = 0 (19-18)
where Q(T) represents any addition of heat to (or removal from) the
reactor wall and mean values of physical properties are used The
above PDEs can be solved with an initial condition, e.g., C(x,0)=
C t=0(x), and a boundary condition, e.g., C(0,t) = C0(t), which is the
concentration at the inlet At steady state, the accumulation termabove is zero, and the solution for an isothermal first-order reaction isthe same as that for a batch reactor, Eq (19-5):
per unit time
Energy generated per unit time
Accumulation
of energy per unit time[
Amount of A
accumulated per unit time[
Trang 12C = C0exp−k = C0e −kt (19-19)
A tubular reactor will likely deviate from plug flow in most practical
cases, e.g., due to backmixing in the direction of flow, reactor
inter-nals, etc A way of simulating axial backmixing is to represent the
reac-tor volume as a series of n stirred tanks in series The steady-state
solution for a single ideal CSTR may be extended to find the effluent
concentration after two ideal CSTRs and then to n ideal stages as
In this case, V ris the volume of each individual reactor in the battery
In modeling a reactor, n is empirically determined based on the extent
of reactor backmixing obtained from tracer studies or other
experi-mental data In general, the number of stages n required to approach
an ideal PFR depends on the rate of reaction (e.g., the magnitude of
the specific rate constant k for the first-order reaction above) As a
practical matter, the conversion for a series of stirred tanks approaches
a PFR for n> 6
An alternate way of generating backmixing is to recycle a fraction of
the product from a PFR back to the inlet This reactor, known as a
recy-cle reactor, has been described in Sec 7 of the Handbook As the recyrecy-cle
ratio (i.e., recycle flow to product flow) is increased, the effective
disper-sion is increased and the recycle reactor approaches an ideal CSTR
Tubular Reactor with Dispersion An alternative approach to
describe deviation from ideal plug flow due to backmixing is to
include a term that allows for axial dispersion D ein the plug flow
reac-tor equations The reacreac-tor mass balance equation now becomes
V′ − D e + r(C, T) + = 0 (19-21)
The model is referred to as a dispersion model, and the value of the
dispersion coefficient D eis determined empirically based on
correla-tions or experimental data In a case where Eq (19-21) is converted to
dimensionless variables, the coefficient of the second derivative is
referred to as the Peclet number (Pe = uL/D e ), where L is the reactor
length and u is the linear velocity For plug flow, D e= 0 (Pe 1 ∞) while
for a CSTR, D e= ∞ (Pe = 0) To solve Eq (19-21), one initial condition
and two boundary conditions are needed The “closed-ends” boundary
conditions are uC0= (uC − D e ∂C/∂L) L= 0and (∂C/∂L) L = L= 0 (e.g., see
Wen and Fan, Models for Flow Systems in Chemical Reactors, Marcel
Dekker, 1975) Figure 19-2 shows the performance of a tubular reactor
with dispersion compared to that of a plug flow reactor
Ideal chemical reactors typically may be modeled using a
combina-tion of ideal CSTR, PFR, and dispersion model equacombina-tions In the case
of a single phase, the approach is relatively straightforward In the case
of two-phase flow, a bubble column (fluidized-bed) reactor may be
modeled as containing an ideal CSTR liquid (emulsion) phase and a
plug flow (with dispersion) gas phase containing bubbles Given inlet
gas conditions, the concentration in the liquid (emulsion) may be
cal-culated using mass-transfer correlations from the bubbles to the liquid
(emulsion) along with reaction in the liquid (emulsion) phase along the
length of the reactor In flooded gas-liquid reactors where the gas and
liquid are countercurrent to each other, a plug flow (with dispersion)
model may be used for both phases The concentration of reactant in a
phase at each end of the reactor is known The concentration of the
other phase is assumed at one end, and mass-transfer correlations and
reaction kinetics are used together with a plug flow (with dispersion)
model to get to the other exit The iterative process continues until the
concentrations at each end match the feed conditions
Reactor Selection Ideal CSTR and PFR models are extreme cases
of complete axial dispersion (D e = ∞) and no axial dispersion (D e= 0),
respectively As discussed earlier, staged ideal CSTRs may be used to
represent intermediate axial dispersion Alternatively, within the context
of a PFR, the dispersion (or a PFR with recycle) model may be used to
represent increased dispersion Real reactors inevitably have a level of
dispersion in between that for a PFR or an ideal CSTR The level of
dis-persion may depend on fluid properties (e.g., is the fluid newtonian),
V′ fluid flow (e.g., the level of mixing), transport properties (e.g., the diffu-sivity of reactants in the fluid), and reactor geometry The effect of
dis-persion in a real reactor is discussed within the context of an ideal CSTRand PFR model in Fig 19-2
Figure 19-2a shows the effect of dispersion on the reactor volume
required to achieve a certain exit concentration (or conversion) As Penumber increases (i.e., dispersion decreases), the reactor begins toapproach plug flow and the reactor volume required to achieve a cer-tain conversion approaches the volume for a PFR At lower Pe num-bers, reactor performance approaches that of an ideal CSTR and thereactor volume required to achieve a certain concentration is much
higher than that of a PFR This behavior can be observed in Fig 19-2b
that shows the effect of exit concentration on reaction rate At a givenrate, an ideal CSTR has the highest exit concentration (lowest conver-sion) and a PFR has the lowest exit concentration (highest conver-
sion) As Fig 19-2c shows, since the concentration in an ideal CSTR is
the same as the exit concentration, there is a sharp drop in tion from the inlet to the bulk concentration In contrast, the concen-tration in the reactor drops continuously from the inlet to the outletfor a PFR At intermediate values of Pe, the “closed-ends” boundarycondition in the dispersion model causes a drop in concentration tolevels lower than for an ideal CSTR
concentra-As discussed in Fig 19-2, for a given conversion, the reactor dence time (or reactor volume required) for a positive order reactionwith dispersion will be greater than that of a PFR This need for alonger residence time is illustrated for a first-order isothermal reac-tion in a PFR versus an ideal CSTR using Eqs (19-13) and (19-19)
Equation (19-22) indicates that, for a nominal 90 percent conversion,
an ideal CSTR will need nearly 4 times the residence time (or volume)
of a PFR This result is also worth bearing in mind when batch reactorexperiments are converted to a battery of ideal CSTRs in series in thefield The performance of a completely mixed batch reactor and asteady-state PFR having the same residence time is the same [Eqs.(19-5) and (19-19)] At a given residence time, if a batch reactor pro-vides a nominal 90 percent conversion for a first-order reaction, a sin-gle ideal CSTR will only provide a conversion of 70 percent Theabove discussion addresses conversion Product selectivity in complexreaction networks may be profoundly affected by dispersion Thisaspect has been addressed from the standpoint of parallel and consec-utive reaction networks in Sec 7
Reactors may contain one or more fluid phases The level of sion in each phase may be represented mathematically by using some
disper-of the above thinking
In industrial practice, the laboratory equipment used in chemicalsynthesis can influence reaction selection As issues relating to kinet-ics, mass transfer, heat transfer, and thermodynamics are addressed,reactor design evolves to commercially viable equipment Often,more than one type of reactor may be suitable for a given reaction Forexample, in the partial oxidation of butane to maleic anhydride over avanadium pyrophosphate catalyst, heat-transfer considerations dictatereactor selection and choices may include fluidized beds or multi-tubular reactors Both types of reactors have been commercialized.Often, experience with a particular type of reactor within the organi-zation can play an important part in selection
There are several books on reactor analysis and modeling including
those by Froment and Bischoff (Chemical Reactor Analysis and Design, Wiley, 1990), Fogler (Elements of Chemical Reaction Engineering, Prentice-Hall International Series, 2005), Levenspeil (Chemical Reac- tion Engineering, Wiley, 1999), and Walas (Modeling with Differential Equations in Chemical Engineering, Butterworth-Heineman, 1991).
Chemical Kinetics Reactor models include chemical kinetics in
the mass and energy conservation equations The two basic laws of
kinetics are the law of mass action for the rate of a reaction and the Arrhenius equation for its dependence on temperature Both of these
strictly apply to elementary reactions More often, laboratory data are
Trang 13used to develop mathematical relationships that describe reaction
rates that are then used These relationships require analysis of the
laboratory reactor data, as discussed in Sec 7 Reactor models will
require that kinetic rate information be expressed on a unit reactor
volume basis Two-phase or multiphase reactors will require a level of
detail (e.g., heat and mass transport between phases) to capture the
relevant physical and chemical processes that affect rate
Pressure Drop, Mass and Heat Transfer Pressure drop is
more important in reactor design than in analysis or simulation The
size of the compressor is dictated by pressure drop across the reactor,
especially in the case of gas recycle Compressor costs can be
signifi-cant and can influence the aspect ratio of a packed or trickle bed
reac-tor Pressure drop correlations often may depend on the geometry, the
scale, and the fluids used in data generation Prior to using literature
correlations, it often is advisable to validate the correlation with
mea-surements on a similar system at a relevant scale
Depending on the type of reactor, appropriate mass-transfer
corre-lations may have to be used to connect intrinsic chemical kinetics to
the reaction rate per unit reactor volume A number of these tions have already been discussed in Sec 5 of the Handbook, “Heatand Mass Transfer.” The determination of intrinsic kinetics hasalready been discussed in Sec 7 of the Handbook In the absence of acorrelation validated for a specific use, the analogy between momen-tum, heat and mass transfer may often be invoked
correla-The local reactor temperature affects the rates of reaction, rium conversion, and catalyst deactivation As such, the local temper-ature has to be controlled to maximize reaction rate and to minimizedeactivation In the case of an exothermic (endothermic) reaction,higher (lower) local temperatures can cause suboptimal local concen-trations Heat will have to be removed (added) to maintain more uni-form temperature conditions The mode of heat removal (addition)will depend on the application and on the required heat-transfer rate.Examples of stirred tank reactors with heat transfer are shown inFig 19-1 If the heat of reaction is not significant, an adiabatic reactormay be used For modest heat addition (removal), a jacketed stirred
equilib-tank is adequate (Fig 19-1a) As the heat exchange requirements
(a)
(b)
FIG 19-2 Chemical conversion by the dispersion model (a) Volume relative to plug flow against residual concentration ratio for a first-order reaction (b) Residual concentration ratio against kC0t for a second-order reaction (c) Concentration profile at the inlet of a closed-ends vessel with disper- sion for a second-order reaction with kC0t= 5.
Trang 14increase, internal coils or internal tubes that contain a heat-transfer
fluid may be required (Fig 19-1b and c) In special cases, where the
peak temperature has to be tightly controlled (e.g., in bioreactors) or
where fouling may be an issue, the liquid may be withdrawn,
circu-lated through an external heat exchanger, and returned to the reactor
(Fig 19-1d) In some cases, the vapor above the liquid may be passed
through an external reflux condenser and returned to the reactor (Fig
19-1e) In highly endothermic reactors, the entire reactor may be
placed inside a fired heater (Fig 19-1f), or the reactor shell may be
heated to high temperatures by using induction heat
Several of the heat-transfer options for packed beds are illustrated
in Fig 19-3 Again, if heat requirements are modest, an adiabatic
reactor is adequate (Fig 19-3a) If pressure drop through the reactor
is an issue, a radial flow reactor may be used (Fig 19-3b) There are
few examples of radial flow reactors in industry Potential problems
include gas distribution in the case of catalyst attrition or settling A
common way of dealing with more exothermic (endothermic)
reac-tions is to split the reactor into several beds and then provide interbed
heat exchange (Fig 19-3c) For highly exothermic (endothermic)
reactors, a shell-and-tube multitubular reactor concept may be
uti-lized (Fig 19-3d) The reactor now begins to look more like a heat
exchanger If multiple beds are needed, rather than using interbed
heat exchangers, cold feed may be injected (also called cold shot) in
between beds (Fig 19-3e) In some cases, the heat exchanger may be
outside the reactor (Fig 19-3f) The concept of a reactor as a heat
exchanger may be extended to an autothermal multitubular reactor in
which, for example, the reactants are preheated on the shell side with
reaction occurring in the tubes (Fig 19-3g) Such reactors can have
control issues and are not widely used A common approach is to have
multiple adiabatic reactors with cooling in between reactors (Fig 19-3h).
If the reaction is endothermic, heat may be added by passing the
effluents from each reactor through tubes placed inside a common
process heater (as is the case for a petroleum reforming reactor shown
in Fig 19-3i) For highly endothermic reactions, a fuel-air mixture or
raw combustion gases may be introduced into the reactor In an
extreme situation, the entire reactor may be housed within a furnace
(as in the case of steam reforming for hydrogen synthesis or ethane
cracking for ethylene production)
At times, the reaction may be exothermic with conversion being
limited by thermodynamic equilibrium In such cases, packed beds in
series with interstage cooling may be used as well The performance
enhancement associated with this approach is shown for two cases in
Table 19-2 Such units can take advantage of initial high rates at high
temperatures and higher equilibrium conversions at lower
tempera-tures For SO2oxidation, the conversion attained in the fourth bed is97.5 percent, compared with an adiabatic single-bed value of 74.8 per-cent With the three-bed ammonia reactor, final ammonia concentra-tion is 18.0 percent, compared with the one-stage adiabatic value of15.4 percent
Since reactors come in a variety of configurations, use a variety ofoperating modes, and may handle mixed phases, design provisions fortemperature control may draw on a large body of heat-transfer theoryand data These extensive topics are treated in other sections of thisHandbook and in other references Some of the high points pertinent to
reactors are covered by Rase (Chemical Reactor Design for Process Plants, Wiley, 1977) Two encyclopedic references, Heat Exchanger Design Handbook (5 vols., Begell House, 1983–1998) and Cheremisi- noff (ed.) (Handbook of Heat and Mass Transfer, 4 vols., Gulf,
1986–1990), have several articles addressed specifically to reactors
Reactor Dynamics Continuous reactors are designed to operate
at or near a steady state by controlling the operating conditions Inaddition, process control systems are designed to minimize fluctua-tions from the target conditions and for safety Batch and semibatchreactors are designed to operate under predefined protocols based onthe best understanding of the process However, the potential forlarge and unexpected deviations from steady state as a result ofprocess variable fluctuations is significant due to the complexity andnonlinearity of reaction kinetics and of the relevant mass- and heat-transfer processes For a set of operating conditions (pressure, tem-perature, composition, and phases present), more than one steadystate can exist Which steady state is actually reached depends on theinitial condition Not all steady states are stable states, and only thosethat are stable can be reached without special control schemes Morecomplex behavior such as self-sustained oscillations and chaoticbehavior has also been observed with reacting systems Further, dur-ing start-up, shutdown, and abrupt changes in process conditions, thereactor dynamics may result in conditions that exceed reactor designlimits (e.g., of temperature, pressure, materials of construction, etc.)and can result in a temperature runaway, reactor blowout, and even anexplosion (or detonation) Parametric sensitivity deals with the analy-sis of reactor dynamics in response to abrupt changes
Steady-State Multiplicity and Stability A simple example of
steady-state multiplicity is due to the interaction between kinetics and
heat transport in an adiabatic CSTR For a first-order reaction atsteady state, Eq (19-13) gives
Trang 15Criteria for Chemically Reacting Systems,” in Dynamics and Modeling of Reactive Systems, Stewart et al (eds.), Academic Press, 1980], Schmitz
[Adv Chem Ser., 148: 156, ACS (1975)], and Razon and Schmitz [Chem Eng Sci., 42 (1987)] However, many of these criteria for specific reac-
tion and reactor systems have not been validated experimentally
Linearized or asymptotic stability analysis examines the stability of a
steady state to small perturbations from that state For example, when
heat generation is greater than heat removal (as at points A− and B+ in
Fig 19-4), the temperature will rise until the next stable steady-state
temperature is reached (for A − it is A, for B+ it is C) In contrast, when
heat generation is less than heat removal (as at points A+ and B− inFig 19-4), the temperature will fall to the next-lower stable steady-state
temperature (for A + and B− it is A) A similar analysis can be done around steady-state C, and the result indicates that A and C are stable
steady states since small perturbations from the vicinity of these return
the system to the corresponding stable points Point B is an unstable
steady state, since a small perturbation moves the system away to either
A or C, depending on the direction of the perturbation Similarly, at
conditions where a unique steady state exists, this steady state is alwaysstable for the adiabatic CSTR Hence, for the adiabatic CSTR consid-
ered in Fig 19-4, the slope condition dQ H /dT > dQ G /dT is a necessary
and sufficient condition for asymptotic stability of a steady state In eral (e.g., for an externally cooled CSTR), however, the slope condition
gen-is a necessary but not a sufficient condition for stability; i.e., violation ofthis condition leads to asymptotic instability, but its satisfaction does notensure asymptotic stability For example, in select reactor systems even
tures 502 ⇒ 433, 502 ⇒ 471, 502 ⇒ 496°C To convert feet to meters, multiply by 0.3048; BPSD to m 3 /h, multiply by 0.00662.
where C f is the feed concentration and a and b are constants related
to Arrhenius rate expression The energy balance equation at steady
state is given by
Q G (T) = −∆H r V r r(C,T) = V′ρC p (T − T f)= Q H (T) (19-24)
where Q G is the heat generation by reaction, Q H is the heat removal by
flow, T is the reactor temperature at steady state, and T fis the feed
temperature Plotting the heat generation and heat removal terms
ver-sus temperature gives the result shown in Fig 19-4 As shown, as many
as three steady states are possible at the intersection of Q G and Q H
Another example of multiplicity is shown in Fig 19-15 for an adiabatic
catalyst pellet, indicating that three effectiveness factor values can be
obtained for a given Thiele modulus for a range of Prater numbers and
Thiele modulus values, leading to three potential steady states Multiple
steady states can occur in different reactor types, including isothermal
systems with complex nonlinear kinetics and systems with interphase
transfer, the main requirement being the existence of a feedback
mech-anism—hence, a homogeneous PFR (without backmixing) will not
exhibit multiplicity Depending on the various physical and chemical
interactions in a reactor, oscillatory and chaotic behavior can also occur
There is a voluminous literature on steady-state multiplicity,
oscilla-tions (and chaos), and derivation of bifurcation points that define the
con-ditions that lead to onset of these phenomena For example, see
Morbidelli et al [“Reactor Steady-State Multiplicity and Stability,” in
Chemical Reaction and Reactor Engineering, Carberry and Varma (eds),
Marcel Dekker, 1987], Luss [“Steady State Multiplicity and Uniqueness
Trang 16a unique steady state can become unstable, leading to oscillatory or
chaotic behavior
Local asymptotic stability criteria may be obtained by first solving
the steady-state equations to obtain steady states and then linearizing
the transient mass and energy balance equations in terms of deviations
of variables around each steady state The determinant (or slope) and
trace conditions derived from the matrix A in the set of equations
obtained are necessary and sufficient for asymptotic stability
q r= Aq r x = C − Css y = T − Tss
∆ = det(A) > 0 σ = trace(A) < 0 (19-25)
where x and y are the deviation variables around the steady state (Css,
Tss) The approach may be extended to systems with multiple
concen-trations and complex nonlinear kinetics For additional references on
asymptotic stability analysis, see Denn (Process Modeling, Longman,
1986) and Morbidelli et al [“Reactor Steady-State Multiplicity and
Stability,” in Chemical Reaction and Reactor Engineering, Carberry
and Varma (eds.), Marcel Dekker, 1987]
Parametric Sensitivity and Dynamics The global stability and
sensitivity to abrupt changes in parameters cannot be determined
from an asymptotic analysis For instance, for the simple CSTR, a key
question is whether the temperature can run away from a lower stable
where T is the reactor temperature, T jis the cooling jacket
tempera-ture, E is the activation energy, and R is the universal gas constant.
Similarly, for a jacketed PFR, a conservative criterion for stability is
Tmax− T j < ∆T c , where Tmaxis the temperature of the hot spot
Another example of sensitivity to abrupt changes is the wrong-way effect, exhibited, for instance, in packed-bed reactors, where an abrupt
reduction in feed rate or in feed temperature results in a dramaticincrease in reactor peak temperature for exothermic reactions Eitherthe reactor may eventually return to the original steady state or, if ahigher-temperature steady state exists, the reactor may establish a tem-perature profile corresponding to the new high steady state Such adynamic excursion can result in an increase of undesirable by-productsconcentration, catalyst deactivation, permanent reactor damage, andsafety issues; e.g., see work by Luss and coworkers [“Wrong-WayBehavior of Packed-Bed Reactors: I The Pseudo-homogeneous
Model,” AIChE J 27: 234–246 (1981)] For more complex systems, the
transient model equations are solved numerically A more detailed
dis-cussion of parametric sensitivity is provided by Varma et al (Parametric Sensitivity in Chemical Systems, Cambridge University Press, 1999).
Reactor Models As discussed earlier, reactor models attempt to
strike a balance between the level of detail included and the ness of the model Too many details in the model may require a largernumber of adjustable model parameters, increase computationalrequirements, and limit how widely the model may be used Too fewdetails, on the other hand, increase ease of implementation but maycompromise the predictive or design capabilities of the model Figure19-5 is a schematic of the inherent tradeoff between ease of imple-mentation and the insight that may be obtained from the model.Increases in computational power are allowing a more cost-effectiveinclusion of a greater number of details Computational fluid dynamics(CFD) models provide detailed flow information by solving theNavier-Stokes transport equations for mass, momentum, and heat bal-ances The user will, however, need to be familiar with the basic ele-ments of the software and may need a license A typical numericalsolution of the governing transport equations is obtained within theeulerian framework, using a large number of computational cells (orfinite volumes that represent reactor geometry) Current capabilities incommercial CFD software can be used to resolve the flow, concentra-tion, and temperature patterns in a single phase with sufficient detailand reasonable accuracy for all length and time scales The ability tovisualize flow, concentration, and temperature inside a reactor is useful
useful-in understanduseful-ing performance and useful-in designuseful-ing reactor useful-internals
RT2
E
TABLE 19-2 Multibed Reactors, Adiabatic Temperature Rises
and Approaches to Equilibrium*
Oxidation of SO 2 at atmospheric pressure in a four-bed reactor Feed 6.26%
*To convert atm to kPa multiply by 101.3.
SOURCE: Plant data and calculated design values from Rase, Chemical
Reac-tor Design for Process Plants, Wiley, 1977.
FIG 19-4 Multiple steady states of CSTRs, stable and unstable, adiabatic (a) First-order reaction, A and C stable, B unstable, the dashed line is for a reversible reaction (b) One, two, or three steady states depending on the combination (C , T).
Trang 17Addition of transport properties and more than one phase (as is thecase with solid catalysts) within a CFD framework complicates theproblem in that the other phase(s) also may have to be included in thecalculations This may require additional transport equations toaddress a range of complexities associated with the dynamics andphysics of each phase, the interaction between and within phases, sub-grid-scale heterogeneities (such as size distributions within eachphase), and coupling with kinetics at the molecular level For exam-ple, one needs the bubble size distribution in a bubble column reactor
to correctly model interfacial area and local mass-transfer coefficients,which can further affect the chemical kinetics Although phenomeno-logical models describing such physical effects have greatly improvedover the years, this area still lacks reliable multiphase turbulence clo-sures, or experimentally validated intraphase and interphase transportmodels Mathematical modeling in industrial practice will continue toinvolve compromises between computational complexity, experimen-tal data needs, ability to validate the model, cost, and the time frame
in which the work may be useful to the organization
FIG 19-5 Hierarchy of reactor models.
RESIDENCE TIME DISTRIBUTION AND MIXING
The time spent by reactants and intermediates at reaction conditions
determines conversion (and perhaps selectivity) It is therefore often
important to understand the residence time distribution (RTD) of
reaction species in the reactor This RTD could be considerably
dif-ferent from what is expected Reasons for the deviation could be
channeling of fluid, recycling of fluid, or creation of stagnant regions
in the reactor, as illustrated in Fig 19-6
This section introduces how tracers are used to establish the RTD
in a reactor and to contrast against RTDs of ideal reactors The section
ends with a discussion of how reactor performance may be connected
to RTD information
TRACERS
Tracers are typically nonreactive substances used in small concentrationthat can be easily detected The tracer is injected at the inlet of the reac-tor along with the feed or by using a carrier fluid, according to some def-inite time sequence The inlet and outlet concentrations of the tracer
Short-circuiting
Stagnant regions
Packed bed
Extreme short-circuitingand bypassChanneling, especially
serious in countercurrenttwo-phase operations
FIG 19-6 Some examples of nonideal flow in reactors (Fig 11.1 in Levenspiel, Chemical tion Engineering, John Wiley & Sons, 1999.)
Trang 18Reac-are recorded as a function of time These data Reac-are converted to a
resi-dence time distribution of feed in the reactor vessel Tracer studies may
be used to detect and define regions of nonideal behavior, develop
phe-nomenological zone models, calculate reactor performance
(conver-sion, selectivity), and synthesize optimal reactor configurations for a
given process The RTD does not represent the mixing behavior in a
vessel uniquely Several arrangements of reactors or internals within a
vessel may provide the same tracer response For example, any series
arrangement of the same number of CSTR and plug flow reactor
ele-ments will provide the same RTD This lack of uniqueness may limit
direct application of tracer studies to first-order reactions with constant
specific rates For other reactions, the tracer curve may determine the
upper and lower limits of reactor performance When this range is not
too broad, or when the purpose of the tracer test is to diagnose
maldis-tribution or bypassing in the reactor, the result can be useful Tracer
data also may be taken at several representative positions in the vessel in
order to develop a better understanding for the flow behavior
Inputs Although some arbitrary variation of input concentration
with time may be employed, five mathematically simple tracer input
signals meet most needs These are impulse, step, square pulse (started
at time a, kept constant for an interval, then reduced to the original
value), ramp (increased at a constant rate for a period of interest), and
sinusoidal Sinusoidal inputs are difficult to generate experimentally.
Types of Responses The key relationships associated with
trac-ers are provided in Table 19-3 Effluent concentrations resulting from
impulse and step inputs are designated Cδand C u, respectively The
mean concentration resulting from an impulse of magnitude m into a
vessel of volume V r is C0= m/V r The mean residence time is the ratio
of the vessel volume to the volumetric flow rate:
t
⎯=
or t⎯=
(19-27)
The reduced time is t r = tt⎯ Residence time distributions are
used in two forms: normalized, E(t r)= CδC0; or plain, E(t)=
E(t)dt = 1, and the relation between them is E(t r)= t⎯E(t) The area
between the ordinates at t1and t2is the fraction of the total effluent
that has spent the period between those times in the vessel The age
function is defined in terms of the step input as
F(t)= =t
0
Reactor Tracer Responses
Continuous Stirred Tank Reactor (CSTR) With a step input of
magnitude C f, the unsteady material balance of tracer
can be integrated to yield
= F(t r)= 1 − exp (−t r) (19-30)
With an impulse input of magnitude m or an initial mean
concentra-tion C0= m/V r, the material balance is
+ C = 0 with C = C0, t= 0 (19-31)And integration gives
= E(t r)= exp(−t r) (19-32)These results show that
t Cδdt
∞ 0
Cδdt
V r
V′
Multistage CSTR Since tubular reactor performance can be
simulated by a series of CSTRs, multistage CSTR tracer models areuseful in analyzing data from empty tubular and packed-bed reactors
The solution for a tracer through n CSTRs in series is found by
induc-tion from the soluinduc-tion of one stage, two stages, and so on
TABLE 19-3 Tracer Response Functions
Mean residence time:
Variance:
σ 2(t)=∞ 0
(t − t)2E(t) dt = −t2 +
Variance, normalized:
σ 2(t r) = = −1 +
=1 0
(t r− 1) 2dF(tr) Skewness, third moment:
γ 3(t r) =∞
0 (t r− 1) 3E(tr ) dt r
∞ 0
t2Cδdt
∞ 0
t2Cδdt
∞ 0
Cδdt
E(t)
I(t) E(t)
tCδdt
∞ 0
Cδdt
Trang 19The solution for a step response can be obtained by integration
where E(t r ) and F(t r ) for various values of n are shown in Fig 19-7.
The theoretical RTD responses in Fig 19-7a are similar in shape to
the experimental responses from pilot and commercial reactors shown
in Fig 19-8 The value of n in Fig 19-8 represents the number of
CSTRs in series that provide a similar RTD to that observed
commer-cially Although not shown in the figure, a commercial reactor having
a similar space velocity as a pilot reactor and a longer length typically
has a higher n value than a pilot reactor due to greater linear velocity.
The variance of the RTD of a series of CSTRs, σ2, is the inverse of n.
σ2=∞ 0
(t r− 1)2E(t r ) dt r= (19-36)
Plug Flow Reactor The tracer material balance over a
differen-tial reactor volume dV ris
by t r = 1, or t = t⎯
Tubular Reactor with Dispersion As discussed earlier, a
multi-stage CSTR model can be used to simulate the RTD in pilot and
com-mercial reactors The dispersion model, similar to Fick’s molecular
diffusion law with an empirical dispersion coefficient D ereplacing the
diffusion coefficient, may also be used
the literature [Otake and Kunigata, Kagaku Kogaku, 22: 144 (1958)].
The plots of E(t r ) versus t rare bell-shaped, similar to the response for
a series of n CSTRs model (Fig 19-7) A relation between σ2(t r ), n,
and Pe (for the closed-ends condition) is
Examples of values of Pe are provided in Fig 19-8 When Pe is
large, n1 Pe2 and the dispersion model reduces to the PFRmodel For small values of Pe, the above equation breaks down
since the lower limit on n is n= 1 for a single CSTR To better resent dispersion behavior, a series of CSTRs with backmixing may
rep-be used; e.g., see Froment and Bischoff (Chemical Reactor sis and Design, Wiley, 1990) A model analogous to the dispersion
Analy-model may be used when there are velocity profiles across the tor cross-section (e.g., for laminar flow) In this case, the equationabove will contain terms associated with the radial position in thereactor
reac-Understanding Reactor Flow Patterns As discussed above, a
RTD obtained using a nonreactive tracer may not uniquely representthe flow behavior within a reactor For diagnostic and simulation pur-poses, however, tracer results may be explained by combining theexpected tracer responses of ideal reactors combined in series, in par-allel, or both, to provide an RTD that matches the observed reactorresponse The most commonly used ideal models for matching anactual RTD are PRF and CSTR models Figure 19-9 illustrates theresponses of CSTRs and PFRs to impulse or step inputs of tracers.Since the tracer equations are linear differential equations, a
Laplace transform L{f(t)}=
0
∞
f(t)e −st dt may be used to relate tracer
inputs to responses The concept of a transfer function facilitates thecombination of linear elements
C
⎯
output(s) = (transfer function) C⎯input(s) = G(s)C⎯input(s) (19-43)Some common Laplace transfer functions are listed in Table 19-4.The Laplace transform may be inverted to provide a tracerresponse in the time domain In many cases, the overall transferfunction cannot be analytically inverted Even in this case,moments of the RTD may be derived from the overall transfer
function For instance, if G′and G″are the limits of the first and
Trang 20second derivatives of the transfer function G(s) as s1 0, the mean
residence time and variance are
t
⎯= G′
0 and σ2(t) = G″0− (G′0)2 (19-44)
In addition to understanding the flow distribution, tracer
experi-ments may be conducted to predict or explain reactor performance
based on a particular RTD To do this, a mathematical expression for
the RTD is needed A PFR, or a dispersion model with a small value
of the dispersion coefficient, may be used to simulate an empty
tubu-lar reactor Stirred tank performance often is nearly completely mixed
(CSTR) In some cases, to fit the measured RTD, the model may have
to be modified by taking account of bypass zones, stagnant zones, or
other parameters associated with the geometry and operation of the
reactor Sometimes the vessel can be visualized as a zone of complete
mixing in the vicinity of impellers followed by plug flow zones
else-where, e.g., CSTRs followed by PFRs Packed beds usually deviate
substantially from plug flow The dispersion model and some
combi-nation of PFRs and CSTRs or multiple CSTRs in series may
approxi-mate their behavior Fluidized beds in small sizes approxiapproxi-mate CSTR
behavior, but large ones exhibit bypassing, stagnancy,
nonhomoge-neous regions, and several varieties of contact between particles and
fluid The additional parameters required to simulate such mixing
behavior can increase the mathematical complexity of the model
The characteristic bell shape of many RTDs can be fit to
well-known statistical distributions Hahn and Shapiro (Statistical Models
in Engineering, Wiley, 1967) discuss many of the standard
distribu-tions and condidistribu-tions for their use The most useful distribudistribu-tions arethe gamma (or Erlang) and the gaussian together with its Gram-Charlier extension These distributions are represented by only a fewparameters that can be used to determine, for instance, the mean andthe variance
Qualitative inspection of the tracer response can go a long waytoward identifying flow distribution problems Additional references
on tracers are Wen and Fan (Models for Flow Systems in Chemical Reactors, Marcel Dekker, 1975) and Levenspiel (Chemical Reaction Engineering, 3d ed., Wiley, 1999).
CONNECTING RTD TO CONVERSION
When the flow pattern is known, the conversion for a given reactionmechanism may be evaluated from the appropriate material andenergy balances When only the RTD is known (or can be calculatedfrom tracer response data), however, different networks of reactorelements can match the observed RTD In reality, reactor perfor-mance for a given reactor network will be unique The conversionobtained by matching the RTD is, however, unique only for linearkinetics For nonlinear kinetics, two additional factors have to be
FIG 19-8 Residence time distributions of pilot and commercial reactors σ 2 = variance of
the residence time distribution, n= number of stirred tanks with the same variance, Pe =
Peclet number (Walas, Chemical Process Equipment, Butterworths, 1990.)
1 䊊 Aldolization of butyraldehyde 0.050 20.0 39.0
2 䊉 Olefin oxonation pilot plant 0.663 1.5 1.4
3 ⵧ Hydrodesulfurization pilot plant 0.181 5.5 9.9
Trang 21accounted for to fully describe the contacting or flow pattern: the
degree of segregation of the fluid and the earliness of mixing of the
reactants
Segregated Flow The degree of segregation relates to the
tendency of fluid particles to move together as aggregates or
clumps (e.g., bubbles in gas-liquid reactors, particle clumps in
flu-idized beds, polymer striations in high-viscosity polymerization
reactors) rather than each molecule behaving independently (e.g.,
homogeneous gas, low-viscosity liquid) A system with no
aggre-gates may be called a microfluid, and the system with aggreaggre-gates a
macrofluid (e.g., see Levenspiel, Chemical Reaction Engineering,
3d ed., Wiley, 1999) In an ideal plug flow or in an ideal batch
reac-tor, the segregated particles in each clump spend an equal time in
the reactor and therefore the behavior is no different from that of a
microfluid that has individual molecules acting independently The
reactor performance is therefore unaffected by the degree of
segre-gation, and the PFR or ideal batch model equations may be used to
estimate performance As shown below, however, this is not the
case for a CSTR where the performance equation for a microfluid
is the same as that of an ideal CSTR, while that of a CSTR with
seg-regated flow is not
In segregated flow the molecules travel as distinct groups All
mol-ecules that enter the vessel together leave together The groups are
small enough that the RTD of the whole system is represented by a
smooth curve Each group of molecules reacts independently of anyother group, that is, as a batch reactor For a batch reactor with apower law kinetics,
When a conversion and an RTD are known, a value of k may be
esti-mated by trial and error so the segregated integral is equal to theknown value If a series of conversions are known at several residencetimes, the order of the reaction that matches the data may be esti-mated by trial and error One has to realize, however, that the RTDmay change with residence time Alternatively, for known intrinsickinetics, a combination of ideal reactors that reasonably match bothRTD and performance may be considered
Early versus Late Mixing—Maximum Mixedness The
con-cept of early versus late mixing may be illustrated using a plug flowreactor and an ideal CSTR in series In one case, the ideal CSTR pre-cedes the plug flow reactor, a case of early mixing In the other case,the plug flow reactor precedes the CSTR, and this is a case of late mix-ing Each of the two arrangements has the same RTD
In maximum mixedness (or earliest possible mixing), the feed is mately mixed with elements of fluid of different ages, for instance, using
inti-multiple side inlets at various points along a plug flow reactor The
TABLE 19-4 Some Common Laplace Transform Functions
Element Transfer function G(s)
Trang 22amount and location of the inlet flows match the RTD This means that
each portion of fresh material is mixed with all the material that has the
same life expectation, regardless of the actual residence time in the
ves-sel up to the time of mixing The life expectation under plug flow
con-ditions is related to the distance remaining to be traveled before leaving
the vessel The concept of maximum mixedness and completely
segre-gated flow is illustrated in Fig 19-10 Segresegre-gated flow is represented as
a plug flow reactor with multiple side outlets and has the same RTD.
In contrast to segregated flow, in which the mixing occurs only after
each side stream leaves the vessel, under maximum mixedness flow,
mixing of all molecules having a certain life expectancy occurs at the
time of introduction of fresh material These two mixing extremes—as
late as possible and as soon as possible, both having the same RTD—
correspond to extremes of reactor performance
The mathematical model for maximum mixedness has been
pro-vided by Zwietering [Chem Eng Sci 11: 1 (1959)].
= r c− (C0− C) (19-47)
where r c is the chemical reaction rate; e.g., for an order q, r c = kC q
The above differential equation in dimensionless variable form
(where f = C/C0and t r = t/t⎯) becomes
= kt⎯ C0q−1 f q− (1− f) (19-48)with boundary condition
which makes
kt⎯
C0q−1 f q∞− (1− f∞)= 0 (19-50)The conversion achieved in the vessel is obtained by the solution of the
differential equation at the exit of the vessel where the life expectation
is t = 0 The starting point for the integration is (f∞,t∞) When
integrat-ing numerically, however, the RTD becomes essentially 0 by the time t r
approaches 3 or 4 Accordingly, the integration interval is from (f∞, t r≤ 3
or 4) to (feffluent,t r = 0) with f∞obtained from Eq (19-50)
The conversion is a maximum in segregated flow and a minimum
under maximum mixedness conditions, for a given RTD and reaction
orders>1 A few comparisons are made in Fig 19-11 In some ranges
of the parameters n or r c, the differences in reactor volume for a given
conversion, when segregated or maximum mixedness flow is assumed,
< segregated flow (and the opposite is the case for orders <1).Increased deviation from ideal plug flow increases the effect of segre-gation on conversion At low conversion, the conversion is insensitive
to the RTD and to the extent of segregation
Novosad and Thyn [Coll Czech Chem Comm 31: 3,710–3,720
(1966)] solved the maximum mixedness and segregated flow tions (fit with the Erlang model) numerically There are few experi-mental confirmations of these mixing extremes One study with a
FIG 19-10 Two limiting flow patterns with the same RTD (a) Segregated flow (b) Maximum mixedness flow.
FIG 19-11 Ratio of reactor volume for maximum mixedness and segregated
flow models as a function of the variance (or n), for several reaction orders.
(a)
(b)
Trang 2350-gal stirred tank reactor found segregation at low agitation and was
able to correlate complete mixing and maximum mixedness in terms
of the power input and recirculation within the vessel [Worrell and
Eagleton, Can J Chem Eng pp 254–258 (Dec 1964)].
REACTION AND MIXING TIMES
Reactants may be premixed or fed directly into the reactor To the
extent that the kinetics are limiting (i.e., reaction rate is slow), the rate
of mixing plays a minor role in determining conversion or selectivity
If the time to mix reactants is comparable to the reaction rate,
how-ever, mixing can have a significant impact
The characteristic chemical reaction time t ror characteristic time
scale of the chemistry may be calculated from the reaction rate
expression For a single reaction,
where C0is a reference concentration of the limiting reactant and T0
is a reference temperature For a first-order reaction, t r=1⁄k , where k
(s−1) is the rate constant
Mixing may occur on several scales: on the reactor scale (macro), on
the scale of dispersion from a feed nozzle or pipe (meso), and on a
molecular level (micro) Examples of reactions where mixing is
impor-tant include fast consecutive-parallel reactions where reacimpor-tant
con-centrations at the boundaries between zones rich in one or the other
reactant being mixed can determine selectivity
Much of the literature around mixing times has been developed
around the mixing of two liquids in agitated stirred tanks The
macromix-ing time t macan be defined as the time for the concentration to settle
within, say,±2 percent of its final value (98 percent homogeneity) With
a standard turbine in a baffled tank and Re (= nD2aρ/µ) > 5000,
t ma≅ 2
where n is the stirrer speed, D t is the tank diameter, D ais the agitator
diameter, and H is the height of the tank; t mavaries inversely with the
stirrer speed In a case of a tank with an aspect ratio of unity and
D a /D t=13, nt ma≅ 36 For a stirrer speed of 120 rpm, the
where q is the flow induced by the impeller The induced flow is about
2 times the direct discharge from the turbine, creating uncertainty in
estimating q; tciris roughly one-fourth of the macromixing time
The micromixing time t miis the time required for equilibration ofthe smallest eddies by molecular diffusion, engulfment, and stretch-ing For liquid-liquid mixing, stretching and engulfment are limiting
factors and t midepends on the kinematic viscosity (µ/ρ) and the localrate of energy dissipation φε⎯:
t mi= 17 1/2
(19-54)For a kinematic viscosity of 10−6m2/s and an energy dissipation of 1.0
W/kg, t mi= 0.017 s The local energy dissipation will vary greatly withposition in the tank with its greatest value near the tip of theimpeller Injection of reactant at the point of greatest turbulence
minimizes t mi
The mesomixing time t meis the time for “significant mixing” of anincoming jet of feed liquid with the surrounding fluid A formula for
estimating t meis the time for turbulent diffusion to transport liquid
over a distance equal to the feed pipe diameter d0
If the diameter of the pipe is proportional to the agitator diameter, t me
increases as d0
2/3
Since t medepends on the local energy dissipation, it is
sensitive to location Typically, t me(> t mi) is a fraction of a second or so
A parameter used to diagnose mixing issues for reactive systems is
the Damköhler number Da which is the ratio of the mixing time to the
reaction time, Da= tmixing/t r Small Da numbers (Da<< 1) indicate atively rapid mixing compared to the reaction, so mixing is less impor-tant In contrast, large Da numbers (Da>>1) indicate a need toconsider mixing issues A more complete discussion of the topic is pro-vided in the appropriate section of the Handbook, in Baldyga and
rel-Bourne (Turbulent Mixing and Chemical Reactions, Wiley, 1998), and
in Harriott (Chemical Reactor Design, Marcel Dekker, 2003).
V r
q
SINGLE-PHASE REACTORS
Section 7 of this Handbook presents the theory of reaction kinetics
that deals with homogeneous reactions in batch and continuous
equipment Single-phase reactors typically contain a liquid or a gas
with (or without) a homogeneous catalyst that is processed in a
reac-tor at conditions required to complete the desired chemical
transfor-mation
LIQUID PHASE
Batch reactions of single or miscible liquids are often done in stirred
or pump-around tanks The agitation is needed to mix multiple feeds
and to enhance heat exchange with cooling (or heating media) during
the process Topics that acquire special importance on an industrial
scale are the quality of mixing in tanks and the residence time
distrib-ution in vessels where plug flow may be the goal A special case is that
of laminar and related flow distributions characteristic of
nonnewton-ian fluids, which often occurs in polymerization reactors The
infor-mation about agitation and heat transfer in tanks is described in the
relevant Handbook section
Homogeneous Catalysis A catalyst is a substance, usually used
in small amounts relative to the reactants, that increases the rate of
a reaction without being consumed in the process Liquid-phase
reactions are often conducted in the presence of homogeneous
cat-alysts Typically, homogeneous catalysts are ions or metal coordination
complexes or enzymes in aqueous solution The specific action of aparticular metal complex can be altered by varying the ligands (orcoordination number) of the complex or the oxidation state of thecentral metal atom Some examples of homogeneous catalysts inindustrial practice include hydrolysis of esters by hydronium(H3O+) or hydroxyl (OH−) ions, hydroformylation of olefins using
Rh or Co carbonyls, decomposition of hydrogen peroxide by ferrousions, decomposition of nitramides catalyzed by acetate ion, inver-sion of sucrose by HCl, halogenation of acetone by H+and OH−,and hydration of isobutene by acids A characteristic of homoge-neous catalysis is that, compared to solid catalysis, the reaction(s)proceeds under relatively mild conditions A key issue associatedwith homogeneous catalysis is the difficulty of separating productand catalyst
In stirred tanks, the power input to agitate the tank will depend onthe physical properties of the liquid In tubular reactors, the axial dis-persion in empty tubes may be estimated [e.g., Wen in Petho and
Noble (eds.), Residence Time Distribution Theory in Chemical neering, Verlag Chemie, 1982] as
Engi-= + 1≤ Re ≤ 2000 and 0.2 ≤ Sc ≤ 1000
(19-56)
(Re)(Sc)
192
1
(Re)(Sc)
1
Pe
Trang 24= + Re≥ 2000
In a general case, the velocity may also be a function of radius One
such case is that of laminar flow which is characterized by a parabolic
velocity profile The velocity at the wall is zero while that at the
cen-terline is twice the average velocity In such cases, a momentum
bal-ance equation is solved along with the equations for heat and mass
transfer, and each equation contains terms for the radial
contribu-tion Laminar flow can be avoided by mixing over the cross-seccontribu-tion
For this purpose, in-line static mixers can be provided For very
vis-cous materials and pastes, screws of the type used for pumping and
extrusion are used as reactors When the temperature of the
reac-tants changes during the course of the reaction (due to either the
heat of reaction or the work required to keep the contents well
mixed), material and energy balance equations have to be solved
simultaneously
Examples
• Crude oil is heated to temperatures at which it thermally cracks into
gasoline and distillate products and lower-molecular-weight gases
This liquid cracking process is referred to as visbreaking A
schematic of the process and the effect of operating variables on
performance is shown in Fig 19-12
• The Wacker process for the oxidation of ethylene to acetaldehyde
with PdCl2/CuCl2at 100°C (212°F) with 95 percent yield and 95 to
99 percent conversion per pass
• The OXO process for higher alcohols: CO+ H2+ C3H6 1
n-butanal1 further processing The catalyst is a rhodium
triph-enylphosphine coordination compound at 100°C (212°F), 30 atm
(441 psi)
• Acetic acid from methanol by the Monsanto process, CH3OH+
CO1 CH3COOH, rhodium iodide catalyst, 3 atm (44 psi), 150°C
(302°F), 99 percent selectivity
See a review of industrial processes that employ homogeneous
cata-lysts by Jennings (ed.), Selected Developments in Catalysis, Blackwell
Scientific, 1985
GAS PHASE
There are few examples of industrial processes with pure gas-phase
reactions The most common and oldest example is combustion
Although termed homogeneous, most gas-phase reactions take place
in contact with solids, either the vessel wall or particles as heat
carri-ers With inert solids, the only complication is with heat transfer
Sev-eral of these reactions are listed in Table 19-1 Whenever possible,
liquefaction of gas-phase systems is considered to take advantage of
the higher rates of liquid reactions, to utilize liquid homogeneous
cat-alysts, or to keep equipment size down
The specific type of equipment used for gas-phase reactions
depends on the conditions required for undertaking the reaction
Examples of noncatalytic gas-phase reactions are shown in Fig 19-13
In general, mixing of feed gases and temperature control are major
process requirements Gases are usually mixed by injecting one of the
streams into the rest of the gases using a high-speed nozzle, as in the
flame reactor (Fig 19-13d).
Examples
• In the cracking of light hydrocarbons and naphtha to olefins, heat is
supplied from combustion gases through tubes in fired heaters at
800°C (1472°F) and sufficiently above atmospheric pressure to
overcome pressure drop Superheated steam is injected to bring the
temperature up quickly and retard coke deposition The reaction
time is 0.5 to 3.0 s, followed by rapid quenching The total tube
length of an industrial furnace may be more than 1000 m Some
other important gas-phase cracking processes include conversion of
toluene to benzene, diphenyl to benzene, dicyclopentadiene to
cyclopentadiene, and 1-butene to butadiene Figure 19-13a shows a
cracking furnace
• The Wulf process for acetylene by pyrolysis of natural gas utilizes a
heated brick checkerwork on a 4-min cycle of heating and reacting
Heat is transferred by direct contact with solids that have been
pre-1.35
(Re)0.125
which 0.03 s is near the peak (Faith, Keyes, and Clark, Industrial Chemicals, vol 27, Wiley, 1975).
• The Wisconsin process for the fixation of nitrogen from air operates
at 2200°C (3992°F), followed by extremely rapid quenching tofreeze the small equilibrium content of nitrogen oxide that is made
[Ermenc, Chem Eng Prog 52: 149 (1956)] A pebble heater
recir-culates refractory pebbles continuously through heating and tion zones Such moving-bed units have been proposed for cracking
reac-to olefins but have been obsolesced like most moving-bed reacreac-tors
• Acetylene may be produced from light hydrocarbons and thas by injecting inert combustion gases directly into the reacting
naph-stream in a flame reactor Figure 19-13a and d shows two such devices; Fig 19-13e shows a temperature profile (with reaction
times in milliseconds)
• Oxidative pyrolysis of light hydrocarbons to acetylene is conducted
in a special burner, at 0.001- to 0.01-s reaction time, peak at 1400°C(2552°F), followed by rapid quenching with oil or water A portion
of a combustible reactant is burned by adding a small amount of air
or oxygen to generate the reaction temperatures needed
• Chlorination reactions of methane and other hydrocarbons cally result in a mixture of products whose relative amounts can becontrolled by varying the Cl/hydrocarbon ratio and recyclingunwanted derivatives For example, one can recycle the mono and
typi-di derivatives when only the tri and tetra derivatives are of value
or keep the chlorine ratio low when emphasizing the lower atives Temperatures are normally kept in the range of 230 to
deriv-400°C (446 to 752°F) to limit carbon formation but may be raised
to 500°C (932°F) when favoring CCl4 Exothermic processes lize cooling through heat-transfer surfaces or cold shots Shell-and-tube reactors with small-diameter tubes, towers with internalrecirculation of gases, or multiple stages with intercooling may beused for these reactions
uti-SUPERCRITICAL CONDITIONS
At near-critical or supercritical conditions, a heterogeneous reactionmixture (e.g., of water, organic compounds, and oxygen) becomeshomogeneous and has some liquid and gaseous properties The rate ofreaction may be considerably accelerated because of (1) the highergas-phase diffusivity, (2) increase of concentration due to liquidlikedensity, (3) enhanced solubility, and (4) increase of the specific rate ofreaction by pressure The mole fraction solubility of naphthalene inethylene at 35°C (95°F) goes from 0.004 at 20 atm (294 psi) to 0.02 at
100 atm (1470 psi) and 0.05 at 300 atm (4410 psi) High destructiveefficiencies (above 99.99 percent) of complex organic pollutant com-pounds in water can be achieved with residence times of under 5 min
at near-critical conditions The critical properties of water are 374°C(705°F) and 218 atm (3205 psi)
We are not aware of any industrial implementation of supercriticalconditions in reactors Two areas of potential interest are wastewatertreatment (for instance, removal of phenol or organic compounds)and reduction of coke on refining catalysts by keeping heavy oildecomposition products in solution A pertinent reference is by Kohn-
stam (“The Kinetic Effects of Pressure,” in Progress in Reaction Kinetics, Pergamon, 1970) More recent reviews of research progress are by Bruno and Ely (eds.), Supercritical Fluid Technology, CRC Press, 1991; Kiran and Brennecke (eds.), Supercritical Engineering Science, ACS, 1992.
POLYMERIZATION REACTORS
Polymerization reactors contain one or more phases There areexamples using solvents in which the reactants and products are inthe liquid phase, the reactants are fed as a liquid (gas) but the prod-ucts are solid, or the reactants are a slurry and the products are sol-uble Phase transformations can occur, and polymers that formfrom the liquid phase may remain dissolved in the remainingmonomer or solvent, or they may precipitate Sometimes beads are
Trang 25FIG 19-12 (a) Visbreaking flow sketch, feed 160,000 lbm/h, k800= 0.000248/s, tubes 5.05-in ID by 40 ft (b) Q/A = 10,000 Btu(ft2⋅h), Pout= 250 psig (c) Q/A = 10,000
Btu(ft 2⋅h), Pout= 150 or 250 psig (d) Three different heat fluxes, Pout= 250 psig (e) Variation of heat flux, average 10,000 Btu(ft2⋅h), Pout= 250 psig ( f ) Halving the specific rate T in°F To convert psi to kPa, multiply by 6.895; ft to m, multiply by 0.3048; in to cm, multiply by 2.54.
Trang 26formed and remain in suspension; sometimes emulsions form In
some processes, solid polymers precipitate from a gas phase into a
fluidized bed containing product solids Polymers are thought of as
organic materials; however, inorganic polymers may be also
synthe-sized (e.g., using crystallization and precipitation) Examples of
inorganic polymers are zeolites
The structure of the polymer determines its physical properties,
e.g., crystallinity, refractive index, tensile strength, glass transition
temperature (at which the specific volume changes slope), and
processability The average molecular weight can cover a wide range
between 104to 107 Given the change in molecular weight, the
vis-cosity can change dramatically as conversion increases For example,
in styrene polymerization, the viscosity increases by a factor of 106asconversion increases from 0 to 60 percent Initiators of chain poly-merization reactions have concentration as low as 10−8g⋅mol/L sothey are highly sensitive to small concentrations of poisons and impu-rities The reaction time can also vary Reaction times for butadiene-styrene rubbers are 8 to 12 h; polyethylene molecules continue togrow for 30 min, whereas ethyl acrylate in 20 percent emulsion reacts
in less than 1 min, so monomer must be added gradually to keep thetemperature within limits In some cases, the adiabatic temperaturerise may be very high For example, in polymerization of ethylene, ahigh adiabatic temperature rise may lead to reactor safety issues byinitiating runaway ethylene decomposition reactions The reactor
FIG 19-13 Noncatalytic gas-phase reactions (a) Steam cracking of light hydrocarbons in a tubular fired heater (b) Pebble heater for the fixation of nitrogen from
air (c) Flame reactor for the production of acetylene from hydrocarbon gases or naphthas [Patton, Grubb, and Stephenson, Pet Ref 37(11): 180 (1958).] (d) Flame
reactor for acetylene from light hydrocarbons (BASF) (e) Temperature profiles in a flame reactor for acetylene (Ullmann Encyclopadie der Technischen Chemie, vol.
3, Verlag Chemie, 1973, p 335).
Trang 27operating conditions have to be controlled such that the possibility of
ethylene decomposition is eliminated.
Since it is impractical to fractionate the products and reformulate
them into desirable ranges of molecular weights, immediate
attain-ment of desired properties must be achieved through the correct
choice of reactor type and operating conditions, notably of
distribu-tions of residence time and temperature Reactor selection may be
made on rational grounds, for historical reasons, or to obtain a
propri-etary position
Each reactor is designed based on the need for mass transfer, heattransfer, and reaction Stirred batch (autoclave) and continuous tubularreactors are widely used because of their flexibility In stirred tanks, idealmixing is typically not achieved, wide variations in temperatures mayresult, and stagnant zones and bypassing may exist Devices that coun-teract these unfavorable characteristics include inserts that cause radialmixing, scraping impellers, screw feeders, hollow-shaft impellers (withcoolant flow through them), recirculation using internal and externaldraft tubes, and so on The high viscosity of bulk and melt polymerization
Trang 28reactions is avoided with solution, bead, or emulsion polymerization, and
more favorable RTDs are obtained In tubular reactors, such as for
low-density polyethylene production, there are strong temperature gradients
in the radial direction and cooling may become an issue These reactors
are operated in a single phase, often with multiple catalyst injection
points, and the reactor can be several miles in length Examples of
poly-merization reactors are illustrated in Fig 19-14
A number of terms unique to polymerization are discussed in Sec 7
of this Handbook A general reference on polymerization is Rodriguez
(Principles of Polymer Systems, McGraw-Hill, 1989) and a reference guide on polymerization reactors is available by Gerrens [German
Chem Eng 4: 1–13 (1981); ChemTech, pp 380–383, 434–443
(1982)] and Meyer and Keurentjes (Handbook of Polymer Reaction Engineering, Wiley VCH, 2005).
FLUID-SOLID REACTORS
A number of industrial reactors involve contact between a fluid (either
a gas or a liquid) and solids In these reactors, the fluid phase contacts
the solid catalyst which may be either stationary (in a fixed bed) or in
motion (particles in a fluidized bed, moving bed, or a slurry) The
solids may be a catalyst or a reactant (product) Catalyst and reactor
selection and design largely depend upon issues related to heat
trans-fer, pressure drop and contacting of the phases In many cases,
con-tinuous regeneration or periodic replacement of deteriorated or
deactivated catalyst may be needed
HETEROGENEOUS CATALYSTS
Solid catalysts may have a homogeneous catalyst (or enzyme) or
cat-alytic ingredients dispersed on a support The support may be
organic or inorganic in nature For example, a catalyst metal atom
may be anchored to the polymer (e.g., polystyrene) through a group
that is chemically bound to the polymer with a coordinating site such
as −P(C6H5)2 or −C5H4 (cyclopentadienyl) Immobilized catalysts
have applications in hydrogenation, hydroformylation, and
polymer-ization reactions [Lieto and Gates, ChemTech, pp 46–53 (Jan.
1983)] Metal or mixed metal oxides may be dispersed on amorphous
materials (such as carbon, silica, or alumina) or exchanged into the
cages of a zeolite Expensive catalytic metal ingredients, such as Pt or
Pd, may be < 1 percent of catalyst weight Catalysts may be shaped as
monoliths, shaped pellets, spheres, or powders Some exceptions are
bulk catalysts such as platinum gauzes for the oxidation of ammonia
and synthesis of hydrogen cyanide, which are in the form of several
layers of fine-mesh catalyst gauze
The catalyst support may either be inert or play a role in catalysis
Supports typically have a high internal surface area Special shapes
(e.g., trilobed particles) are often used to maximize the geometric
sur-face area of the catalyst per reactor volume (and thereby increase the
reaction rate per unit volume for diffusion-limited reactions) or to
min-imize pressure drop Smaller particles may be used instead of shaped
catalysts; however, the pressure drop increases and compressor costs
become an issue For fixed beds, the catalyst size range is 1 to 5 mm
(0.04 to 0.197 in) In reactors where pressure drop is not an issue, such
as fluidized and transport reactors, particle diameters can average less
than 0.1 mm (0.0039 in) Smaller particles improve fluidization;
how-ever, they are entrained and have to be recovered In slurry beds the
diameters can be from about 1.0 mm (0.039 in) down to 10 µm or less
The support has an internal pore structure (i.e., pore volume and
pore size distribution) that facilitates transport of reactants (products)
into (out of) the particle Low pore volume and small pores limit the
accessibility of the internal surface because of increased diffusion
resistance Diffusion of products outward also is decreased, and this
may cause product degradation or catalyst fouling within the catalyst
particle As discussed in Sec 7, the effectiveness factor η is the ratio
of the actual reaction rate to the rate in the absence of any diffusion
limitations When the rate of reaction greatly exceeds the rate of
dif-fusion, the effectiveness factor is low and the internal volume of the
catalyst pellet is not utilized for catalysis In such cases, expensive
cat-alytic metals are best placed as a shell around the pellet The rate of
diffusion may be increased by optimizing the pore structure to
pro-vide larger pores (or macropores) that transport the reactants
(prod-ucts) into (out of) the pellet and smaller pores (micropores) that
provide the internal surface area needed for effective catalyst
disper-sion Micropores typically have volume-averaged diameters of 50 to
200 Å with macropore diameters of 1000 to 5000 Å The pore volumeand the pore size distribution within a porous support determine itssurface area The surface area of supports can range from 0.06 m2/mL(18,300 ft2/ft3) to 600 m2/mL (1.83 × 108ft2/ft3) and above Higherpore volume catalysts have higher diffusion rate at the expense ofreduced crush strength and increased particle attrition
The effective diffusion coefficient Deffdetermines the rate of sion and therefore the volume of the catalyst utilized The coefficient
diffu-is determined by the nature of the diffusing species and the porestructure of the catalyst It has been found to be directly proportional
to the product of diffusivity and porosity ε and inversely proportional
to the tortuosity τ (that is empirically determined) In large pores of
>1000 Å, where molecules collide with one another and the tion with the pore walls is minimal, molecular (or bulk) diffusion isimportant For pore diameters in the range of 50 to 200 Å, collisionwith the pore walls becomes more important, and this regime is calledthe Knudsen diffusion regime In an extreme case where the size ofthe molecule is comparable to the size of the pore, the size and con-figuration of the pores themselves affect diffusivity This happenswhen the diffusing molecule is very large (as in transporting largeorganometallitic molecules through catalyst pores in heavy oilhydrotreating) or the pore is very small (as in diffusion in zeolites), orboth (e.g., see Sec 7 for diffusion regimes) ε ranges from 0.1 to 0.5andτ ranges from 1 to 7 In the absence of other information, a τ value
interac-of 3 to 4 may be used; however, it is best measured for the catalyst interac-ofinterest Expressions for estimating the effective diffusion coefficient
are available in textbooks such as Satterfield (Heterogeneous Catalysis
in Practice, McGraw-Hill, 1991).
The effectiveness factor η is the ratio of the rate of reaction in aporous catalyst to the rate in the absence of diffusion (i.e., under bulkconditions) The theoretical basis for η in a porous catalyst has beendiscussed in Sec 7 For example, for an isothermal first-order reaction
where C iis the bulk concentration of the reactant As discussed ously, η is a function of the ratio of the rate of reaction to diffusion,also called the Thiele modulus φ As the rate constant increases, ηdecreases and eventually reaches an asymptotic value (that depends
previ-onφ) Under these conditions, kη increases as k1 >2 The role of sion and reaction in porous catalysts, however, is more complicated in
diffu-a cdiffu-ase where hediffu-at effects diffu-are present In diffu-addition to the mdiffu-ass vation equation around the pellet, an energy balance equation isrequired Two additional dimensionless parameters are needed forestimating an effectiveness factor:
where∆H ris the heat of reaction, λ is the thermal conductivity of the
catalyst, E is the activation energy, and R is the universal gas constant.
The dimensionless parameter β, known as the Prater number, is theratio of the heat generation to heat conduction within the pellet and is
a measure of the intra-particle temperature increase; γ is the sionless activation energy for the reaction For an exothermic reac-tion, the temperature inside the catalyst pellet is greater than or equal
dimen-to the surface temperature The maximum steady-state temperature
inside the pellet is T s(1+ β) Figure 19-15 is one of several cases ined by Weisz and Hicks for a first-order reaction in an adiabatic
Trang 29catalyst pellet [Chem Eng Sci 17: 263 (1962)] Although this
pre-dicts some very large values of η in some ranges of the parameters,
these values are often not realized in commercial reactors (see Table
19-5) The modified Lewis number defined as Lw′= λs/ρs C ps Deffcan
determine the transient temperature inside the pellet, which can be
much larger than the steady-state temperature
The concept of an effectiveness factor is useful in estimating the tion rate per catalyst pellet (volume or mass) It is, however, mainly use-ful for simple reactions and simple kinetics When there are complexreaction pathways, the concept of effectiveness factor is no longer easilyapplicable, and species and energy balance equations inside the particlemay have to be solved to obtain the reaction rates per unit volume of
reac-FIG 19-15 Effectiveness factors versus Thiele modulus for a first-order reaction in spheres
under adiabatic conditions [Weisz and Hicks, Chem Eng Sci., 17: 265 (1962).]
TABLE 19-5 Parameters of Some Exothermic Catalytic Reactions
Trang 30catalyst Dumesic et al (The Microkinetics of Heterogeneous Catalysis,
American Chemical Society, 1993) use microkinetic analysis to
eluci-date reaction pathways of several commercial catalysts
Another complication is the fact that Fig 19-15 was developed for
the constant-concentration boundary condition, C⏐ r =R = C0 In a more
general case, external mass-transfer limitations will need to be
included
k m a(C0− C i)= r c (C i)= kηC i (19-59)
where k mis the external mass-transfer coefficient obtained from
litera-ture correlations and a is the external surface area per unit pellet
vol-ume The above equation will have to be solved for C i, the
concentration of the reactant on the external surface of the catalyst, so
that the rate per pellet can be obtained The reaction rate per unit
reac-tor volume then becomes r c(1− εb), where εbis the bed void fraction
A further complication is that catalyst activity declines with time
Catalysts may deactivate chemically (via poisons and masking agents),
thermally (via support sintering), or mechanically (through attrition)
Commercial catalyst life can range from a second to several years For
example, in refinery fluid catalytic cracking, the catalyst may lose most
of its activity in less than 10 s, and a transport bed reactor coupled with
a fluidized-bed regenerator is used to circulate catalyst In contrast, a
refinery hydroprocessing catalyst deactivates very slowly and a
fixed-bed reactor may be used without catalyst replacement for one or more
years The deactivation rate expression may often be inferred from
aging experiments undertaken under pilot-plant conditions of constant
temperature or conversion Since accelerated-aging experiments are
often difficult (especially when the concentration of reactant or
prod-ucts affects the deactivation rate), reactor designs where the catalyst
charge provides the required performance between regeneration
cycles is typically based on good basic data and experience The
litera-ture describes approaches aimed at managing deactivation In the case
of platinum reforming with fixed beds, a large recycle of hydrogen
pre-vents coke deposition while a high temperature compensates for the
retarding effect of hydrogen on this essentially dehydrogenating
process Fluidized beds are largely isothermal and can be designed for
continuous regeneration; however, they are more difficult to operate,
require provisions for dust recovery, suffer from backmixing, and are
more expensive Catalyst deactivation mechanisms and kinetics are
dis-cussed in detail in Sec 7 of the Handbook
A catalyst for a particular chemical transformation is selected using
knowledge of similar chemistry and some level on empirical
experi-mentation Solid catalysts are widely used due to lower cost and ease
of separation from the reaction medium Their drawbacks include a
possible lack of specificity and deactivation that can require reactor
shutdown for catalyst regeneration or replacement
There are number of useful books on catalysis Information on
cata-lysts and processes is presented by Thomas (Catalytic Processes and
Proven Catalysts, Academic Press, 1970), Pines (Chemistry of
Cat-alytic Conversions of Hydrocarbons, Academic Press, 1981), Gates
et al (Chemistry of Catalytic Processes, McGraw-Hill, 1979), Matar et
al (Catalysis in Petrochemical Processes, Kluwer Academic Publishers,
1989), and Satterfield (Heterogeneous Catalysis in IndustrialPractice,
McGraw-Hill, 1991) The books by Thomas (Catalytic Processes and
Proven Catalysts, Academic Press, 1970), Butt and Petersen
(Activa-tion, Deactivation and Poisoning of Catalyst, Academic Press, 1988),
and Delmon and Froment (Catalyst Deactivation, Elsevier, 1980)
pro-vide several examples of catalyst deactivation Catalyst design is
dis-cussed by Trimm (Design of Industrial Catalysts, Elsevier, 1980),
Hegedus et al (Catalyst Design Progress and Perspectives, Wiley,
1987), and Becker and Pereira (Catalyst Design, Marcel Dekker,
1993) A thorough review of catalytic reactions and catalysts arranged
according to the periodic table is in a series by Roiter (ed.) (Handbook
of Catalytic Properties of Substances, in Russian, 1968) Stiles
(Cata-lyst Manufacture, Dekker, 1983) discusses cata(Cata-lyst manufacture.
CATALYTIC REACTORS
Due to the considerations noted above, reactor selection will depend
on the type of catalyst chosen and its activity, selectivity, and
deactiva-tion behavior Some reactors with solid catalysts are represented inFig 19-16
Wire Gauzes Wire screens are used for very fast catalytic
reac-tions or reacreac-tions that require a bulk noble metal surface for reactionand must be quenched rapidly The nature and morphology of thegauze or the finely divided catalyst are important in reactor design.Reaction temperatures are typically high, and the residence times are
on the order of milliseconds
Since noble metals are expensive, the catalyst cost is typically high.The physical properties of the gauze pack are important to determineperformance, selectivity, and catalyst replacement strategy The gauze
is typically mounted over the top of a heat exchanger tube sheet orover porous ceramic bricks that are laid over the tube sheet Thegauze pack may be covered with a ceramic blanket to minimize radia-tion losses From a modeling standpoint, the external surface area pergauze volume and the external mass-transfer coefficient for each com-ponent are important parameters, and the reaction rate per unit vol-ume of catalyst may be limited by the rate of external mass transfer.The reaction rate can then be included into a corresponding PFR ordispersion model to obtain estimates of conversion and selectivity
• In hydrogen cyanide synthesis using the Andrussow process, air,methane, and ammonia are fed over 15 to 50 layers of noble metalgauze at 1050 to 1150°C at near atmospheric pressure
Monolith Catalysts For fast reactions that may require a slightly
higher residence time than gauzes or that do not benefit from the bulknoble metal gauze structure, monoliths may be used Most often, themonolith catalyst is an extruded ceramic honeycomb structure thathas discrete channels that traverse its length The catalytic ingredientsmay be dispersed on a high surface area support and coated on aninert honeycomb In some cases, the catalyst paste itself may beextruded into a monolith catalyst Monoliths may also be made ofmetallic supports Stainless steel plates (or wire mesh) with ridges may
be coated with catalysts and stacked one against the other in a reactor.Corrugated stainless steel layers may alternate in between flat sheets
to form the structure A variant is a stainless steel sheet that is gated in a herringbone pattern, coated with catalyst and then rolled(or folded back and forth onto itself) into a reactor module Examples
corru-of cross-sections corru-of the types corru-of monoliths used in industry are shown
in Fig 19-17
The thickness of monolith walls is adjusted according to the als of construction (ceramic honeycombs have thicker walls to providemechanical strength) The size of the channels is selected according tothe application For example, for particulate-laden gases, a largerchannel size ceramic monolith and a higher linear velocity allow theparticles to pass through the catalyst without plugging the channel Incontrast, for feed that does not contain particles, smaller channelmonoliths may be used The cell density of the monolith may varybetween 9 and 600 cells per square inch
materi-A monolith catalyst has a much higher void fraction (between 65and 91 percent) than does a packed bed (which is between 36 and
45 percent) In the case of small channels, monoliths have a highgeometric surface area per unit volume and may be preferred formass-transfer-limited reactions The higher void fraction providesthe monolith catalyst with a pressure drop advantage compared tofixed beds
A schematic of a monolith catalyst is shown in Fig 19-18a In cases
where pressure drop is limiting, such as for CO oxidation in cogenerationpower plant exhausts, monolith catalyst panels may be stacked to form athin (3- to 4-in-thick) wall The other dimensions of the wall can be onthe order of 35× 40 ft CO conversion is over 90 percent with a pressuredrop across the catalyst of 1.5 in of water Alternatively, the monolith may
be used as a catalyst and filter, as is the case for a diesel particulate filter
In this case, monolith channels are blocked and the exhaust gases from a
diesel truck are forced through the walls (Fig 19-18b) The filter is a
crit-ical component in a continuous regenerable trap NO in the exhaust
Trang 3119-28
Trang 32gases is oxidized into NO2that reacts with the soot trapped in the walls of
the filter to regenerate it in situ
Modeling considerations for monoliths are similar to those of gauze
catalysts; however, since the flow and temperature in each channel
may be assumed to be identical to those in the next channel, the
solu-tion for a single channel may reflect the performance of the reactor
For an application in which the reaction rate is mass-transfer-limited,
the reactant concentration at the wall of the catalyst is much lower
than in the bulk and may be neglected In such a case, the fractional
conversionξ is
ξ = 1 − e −k m at= 1 − exp−Sh aL (19-60)
Sc Re
where Sh (= k m d ch /D) is the Sherwood number, Sc ( = µ/ρD) is the
Schmidt number, and Re (= ud chρ/µ) is the channel Reynolds
num-ber; a is the geometric surface area per unit volume of monolith A
number of correlations for Sh are available for various types of liths For example, in the case of extruded ceramic monoliths, a corre-lation for estimating the external mass-transfer coefficient is provided
mono-by Uberoi and Pereira (Ind Eng Chem Res 35: 113–116 (1996)]:
Sh= 2.6961+ 0.139 ScRe 0.81
(19-61)Since typical monolith catalysts have a thin coating of catalyticingredients on the channel walls, they can be susceptible to poisoning
FIG 19-18 Monolith catalysts (a) Schematic of an automobile catalytic converter for the three-way removal of CO, hydrocarbons, and NO x (b) Schematic
of a diesel trap (Figs 7.10 and 9.6 in Heck, Farrauto, and Gulati, Catalytic Air Pollution Control: Commercial Technology, Wiley-Interscience, 2002.)