You are on page 1of 69

Reactor of

Chemical Engineering
(Perrys Chemicial Engineerings, #8)

Presented by Nazarudin,PhD
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 velocity (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 important component of
Interactions between kinetics, fluid flow, transport resistances, and
operating cost. A measure of reactor utilization called space time yield (STY) is the
heat
ratio ofeffects
productsometimes result
generation rate in multiple
to reactor volume.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
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 reactor 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. Constrained dependent variables
often include pressure drop (limited due to compressor cost), feed composition (dictated by the
composition of the recycle streams), temperature rise (or decline), and local and effluent
The reactor design problem is often aimed at optimizing
composition.
independent variables (within constraints) to maximize an objective
function (such as conversion and selectivity).

Since the reactor feed may contain inert species (e.g., nitrogen and solvents) and since there may
be unconverted feed and by-products in the reactor effluent, a number of unit operations
(distillation, filtration,etc.) may be required to produce the desired product(s ). In practice,the
Reactor 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 polymers) applications.
Polymerization reactors convert raw materials to polymers having a specific molecular
weight and functionality. The difference
between polymerization and chemical reactors is artificially based on the size of the molecule
produced. Bioreactors utilize (often genetically manipulated) organisms to catalyze
biotransformations either aerobically (in the presence of air) or anaerobically (without air
present). Electrochemical reactors use electricity to drive desired reactions. Examples
include synthesis of Na metal from NaCl and Al from bauxite ore. A variety of reactor types are
employed for specialty materials synthesis applications (e.g., electronic, defense, and other).

Classification by Phase
Despite the generic classification by operating mode, reactors are designed to accommodate
the reactant phases
and provide optimal conditions for reaction. Reactants may be fluid(s) or solid(s), and as such,
several reactor types have been developed. Singlephase reactors are typically gas- (or
plasma- ) or liquid-phase reactors. Two-phase reactors may be gas-liquid, liquid-liquid, gas-
solid, or liquidsolid reactors. Multiphase reactors typically have more than two phases
present. The most common type of multiphase reactor is a gas-liquidsolid reactor; however,
liquid-liquid-solid reactors are also used. The classification by phases will be used to develop
the contents of this section.
Reactor Modeling
A key aspect of modeling is to derive the appropriate momentum, mass, or energy
conservation equations for the reactor. These balances may be used in lumped systems or
derived over a differential volume within the reactor and then integrated over the reactor
volume. Mass conservation equations have the following general form:

The model defines each of these terms. Solving the set of equations provides
outputs that can be validated against experimental observations and then used
for predictive purposes.
Reactor Types
Ideal
PFR
CSTR
Real
Unique design geometries and therefore RTD
Multiphase
Various regimes of momentum, mass and heat transfer
Reactor Cost
Reactor is
PRF
Pressure vessel
CSTR
Storage tank with mixer
Pressure vessel
Hydrostatic head gives the pressure to design for
Reactor Cost
PFR
Reactor Volume (various L and D) from reactor kinetics
hoop-stress formula for wall thickness:
PR
t tc
SE 0.6 P
t= vessel wall thickness, in.
P= design pressure difference between inside and outside of vessel,
psig
R= inside radius of steel vessel, in.
S= maximum allowable stress for the steel.
E= joint efficiency (0.9)
tc=corrosion allowance = 0.125 in.
Reactor Cost
Pressure Vessel Material of Construction gives metal
Mass of vessel = metal (VC+2VHead)
Vc = DL
VHead from tables that are based upon D

Cp= FMCv(W)
Reactors in Process Simulators
Stoichiometric Model
Specify reactant conversion and extents of
reaction for one or more reactions
Two Models for multiple phases in
chemical equilibrium
Kinetic model for a CSTR
Kinetic model for a PFR Used in early stages of design

Custom-made models (UDF)


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

where qAk is 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 heattransfer 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 reaction proceeds as the feed is added. The reactor
equations governing the feed addition portion of the process are

For a constant reactant flow rate v0,

Given an initial condition

For an isothermal first-order reaction, substitution of this relationship 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 concentration at the end of feed addition serving as the
initial concentration for the batch reactor.
Ideal Continuous Stirred Tank Reactor
In an ideal CSTR, reactants are fed into and removed from an ideally mixed tank. As a result, the
concentration within the tank is uniform and identical to the concentration of the effluent. The
mass and energy conservation equations for an ideal constant-volume or constant-density CSTR
with constant volumetric feed rate Vmay be written as

where Q(T) represents any addition or removal of heat from the reactor and mean values of
physical properties are used. For example, if heat is transferred through the reactor wall, Q(T) =
AkU(Tc T), where Ak is the heat-transfer area, U is the overall heat-transfer coefficient, and Tc is
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 ODE may be solved analytically. At steady state, the
accumulation term is
zero, and the solution for the effluent concentration becomes

Since the contents of an ideal CSTR are perfectly mixed, the dispersion within the reactor is
infinite. In practice, CSTRs may not be ideally mixed. In such cases, the reactor may be modeled as
having a fraction of the feed in bypass and a fraction of the reactor volume stagnant. The
material balance is
Ideal Continuous Stirred Tank Reactor (Continued)

where C1 is the concentration leaving the active zone (coniof the tank. Elimination of C1 will relate
the input and overall output concentrations. For a first-order reaction,

The two parameters and may be expected to depend on reactor internals 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
travels through the reactor as a plug, i.e., each molecule enters the reactor at the same velocity
and has exactly the same residence time. As a
result, the concentration of every molecule at a given distance downstream of the inlet is the same.
The mass and energy balance for a differential volume between position Vr and Vr dVr from the inlet
may be written as partial differential equations (PDEs) for a constantdensity system:

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)
=Ct=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 term above is zero, and the solution for an isothermal first-order
reaction is the same as that for a batch reactor, Eq. (19-5)

A tubular reactor will likely deviate from plug flow in most practical cases, e.g., due to backmixing in
the direction of flow, reactor internals, etc. A way of simulating axial backmixing is to represent the
reactor volume as a series of n stirred tanks in series. The steady-state solution for a single ideal
Plug Flow Reactor (continued)

In this case, Vr is 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 experimental 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 recycle reactor, has been described in Sec. 7 of the Handbook. As
the recycle ratio (i.e., recycle flow to product flow) is increased, the effective dispersion is increased
and the recycle reactor approaches an ideal CSTR.
Reactor Selection
Ideal CSTR and PFR models are extreme cases of complete axial dispersion (De ) and no axial
dispersion (De 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 dispersion may depend on fluid properties
(e.g., is the fluid newtonian), fluid flow (e.g., the level of mixing), transport properties (e.g., the
diffusivity
of reactants in the fluid), and reactor geometry. The effect of dispersion in a real reactor is discussed
within the context of an ideal CSTR and 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 Pe number increases (i.e., dispersion decreases), the reactor
begins to
approach plug flow and the reactor volume required to achieve a certain conversion approaches the
volume for a PFR. At lower Pe numbers, reactor performance approaches that of an ideal CSTR and
the reactor 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 given rate, an ideal CSTR has the highest exit concentration (lowest conversion)
and a PFR has the lowest exit concentration (highest conversion). 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
concentration from the inlet to the bulk concentration. In contrast, the concentration in the reactor
drops continuously from the inlet to the outlet for a PFR. At intermediate values of Pe, the closed-
ends boundary condition in the dispersion model causes a drop in concentration to levels lower
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 reactor experiments are converted to a battery of ideal CSTRs in series in the field. The
performance of a completely mixed batch reactor and a steady-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 provides
a nominal 90 percent conversion for a first-order reaction, a single ideal CSTR will only provide a
conversion of 70 percent. The above discussion addresses conversion. Product selectivity in complex
reaction networks may be profoundly affected by dispersion. This aspect has been addressed from
the standpoint of parallel and consecutive reaction networks in Sec. 7.
Reactors may contain one or more fluid phases. The level of dispersion in each phase may be
represented mathematically by using some of the above thinking.
In industrial practice, the laboratory equipment used in chemical synthesis can influence reaction
selection. As issues relating to kinetics, 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. For example, in the partial oxidation of butane to
maleic anhydride over a vanadium pyrophosphate catalyst, heat-transfer considerations dictate
reactor selection and choices may include fluidized beds or multitubular reactors. Both types of
reactors have been commercialized. Often, experience with a particular type of reactor within the
organization 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 Reaction Engineering,
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 used
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
Examples of stirred tank
reactors with heat transfer are
shown in Fig. 19-1. If the heat of
reaction is not significant, an
adiabatic reactor may be used.
For modest heat addition
(removal), a jacketed stirred
tank is adequate (Fig. 19-1a). As
the heat exchange requirements
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, circulated
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
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) reactions 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
utilized (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
If the reaction is endothermic, heat may be added by passing
may be injected (also called cold shot) in
the effluents from each reactor through tubes placed inside a
between beds (Fig. 19-3e). In some cases, the
common process heater (as is the case for a petroleum
heat exchanger may be outside the reactor
reforming reactor shown in Fig. 19-3i). For highly endothermic
(Fig. 19-3f). The concept of a reactor as a heat
reactions, a fuel-air mixture or raw combustion gases may be
exchanger may be extended to an autothermal
introduced into the reactor. In an extreme situation, the entire
multitubular reactor in which, for example, the
reactor may be housed within a furnace (as in the case of
reactants are preheated on the shell side with
steam reforming for hydrogen synthesis or ethane cracking for
Reactor Models As discussed earlier, reactor models
attempt to
strike a balance between the level of detail included and the
usefulness
of the model. Too many details in the model may require a
larger
number of adjustable model parameters, increase
computational
requirements, and limit how widely the model may be used.
Too few
details, on the other hand, increase ease of implementation
but may
compromise the predictive or design capabilities of the model.
Figure
19-5 is a schematic of the inherent tradeoff between ease of
implementation
and the insight that may be obtained from the model.
Increases in computational power are allowing a more cost-
effective
inclusion of a greater number of details. Computational fluid
dynamics
(CFD) models provide detailed flow information by solving the
Navier-Stokes transport equations for mass, momentum, and
Kinetic Reactors - CSTR & PFR
Used to Size the Reactor
Used to determine the reactor dynamics
Reaction Kinetics

dC j C
rj k (T ) Ci i
dt i 1

EA
k (T ) ko exp( /)
RT
PFR no backmixing
Used to Size the Reactor
Xk
dX
V Fko
0
rk

Space Time = Vol./Q


Outlet Conversion is used for flow sheet mass and heat
balances
CSTR complete backmixing
Used to Size the Reactor

Fko X k
V
rk
Outlet Conversion is used for flow sheet mass and heat
balances
Review : Catalytic Reactors Brief Introduction
Major Steps

A B
Bulk FluidCAb

1. External 7 . Diffusion of
Diffusion products
Rate = kC(CAb from pore
CAS) mouth to
External Surface bulk
C
of Catalyst Pellet As
6 . Diffusion of
2. Defined by an products
Effectiveness from interior to
Factor
Internal pore
Surface mouth
of Catalyst
3. Surface
Pellet
Adsorption A B
A + S <-> A.S 5. Surface Desorption
Catalyst 4. Surface Reaction
B. S <-> B + S
Surface
Catalytic Reactors
Various Mechanisms depending on rate limiting step
Surface Reaction Limiting
Surface Adsorption Limiting
Surface Desorption Limiting
Combinations
Langmuir-Hinschelwood Mechanism (SR Limiting)
H2 + C7H8 (T) CH4 + C6H6(B)

k pT p H 2
rT
1 1.39 p B 1.04 pT
Catalytic Reactors Implications on design

1. What effects do the particle diameter and the fluid velocity


above the catalyst surface play?
2. What is the effect of particle diameter on pore diffusion ?
3. How the surface adsorption and surface desorption influence the
rate law?
4. Whether the surface reaction occurs by a single-site/dual site /
reaction between adsorbed molecule and molecular gas?
5. How does the reaction heat generated get dissipated by reactor
design?
Enzyme Catalysis
Enzyme Kinetics

k1k3C H 2O C E CS
rs
k1CS k 2 k3C H 2O

S= substrate (reactant)
E= Enzyme (catalyst)
Problems
Managing Heat effects
Optimization
Make the most product from the least reactant
Optimization of Desired Product
Reaction Networks
Maximize yield,
moles of product formed per mole of reactant consumed
Maximize Selectivity
Number of moles of desired product formed per mole of undesirable
product formed
Maximum Attainable Region see discussion in Chapt. 7.
Reactors (pfrs &cstrs in series) and bypass
Reactor sequences
Which come first
Managing Heat Effects
Reaction Run Away
Exothermic
Reaction Dies
Endothermic

Preventing Explosions
Preventing Stalling
Temperature Effects
On Equilibrium
On Kinetics
Equilibrium Reactor-
Temperature Effects
Single Equilibrium
aA +bB rR + sS

ai activity of component I
Gas Phase, ai = iyiP,
Vant Hoff eq.
i== fugacity coefficient of i
Liquid Phase, ai= i xi exp[Vi (P-Pis) /RT]

aRr aSs Grxn


o
d ln K eq H rxn o
i = activity coefficient of i
Vi =Partial Molar Volume of i

K eq a a exp , 2
a A aB RT dT RT
Overview of CRE Aspects related to Process
Design

Le Chateliers
Principle
1. Levenspiel , O. (1999), Chemical Reaction Engineering, John Wiley and Sons , 3 rd ed.
Unfavorable Equilibrium

Increasing Temperature Increases the


Rate
Equilibrium Limits Conversion
Overview of CRE Aspects related to Process Design

1. Levenspiel , O. (1999), Chemical Reaction Engineering, John Wiley and


Sons , 3rd ed.
Feed Temperature, Hrxn
Adiabatic
Adiabatic
Cooling

Heat Balance over Reactor

Q = UA Tlm
Reactor with Heating or Cooling

Q = UA T
Kinetic Reactors - CSTR & PFR
Temperature Effects
Used to Size the Reactor
Used to determine the reactor dynamics
Reaction Kinetics

dC j C
rj k (T ) Cii
dt i 1

EA
k (T ) ko exp
RT
PFR no backmixing
Used to Size the Reactor
Xk
dX
V Fko
0
rk

Space Time = Vol./Q


Outlet Conversion is used for flow sheet mass and heat
balances
CSTR complete backmixing
Used to Size the Reactor
Fko X k
V
rk
Outlet Conversion is used for flow sheet mass and heat
balances
Unfavorable Equilibrium

Increasing Temperature Increases the


Rate
Equilibrium Limits Conversion
Various Reactors, Various Reactions
X
k
dX
V Fko
0
rk

Fko X k
V
rk
Reactor with Heating or Cooling

Q = UA T
Temperature Profiles in a Reactor

Exothermic Reaction

Recycle
Best Temperature Path
Optimum Inlet Temperature
Exothermic Rxn

CSTR
PFR
Managing Heat Effects
Reaction Run Away
Exothermic
Reaction Dies
Endothermic

Preventing Explosions
Preventing Stalling
Inter-stage Cooler

Lowers Temp.

Exothermic Equilibria
Inter-stage Cold Feed

Lowers Temp
Lowers Conversion

Exothermic Equilibria
Optimization of Desired Product
Reaction Networks
Maximize yield,
moles of product formed per mole of reactant consumed
Maximize Selectivity
Number of moles of desired product formed per mole of undesirable
product formed
Maximum Attainable Region see discussion in Chapt. 6.
Reactors and bypass
Reactor sequences
Reactor Design for Selective
Product Distribution
S,S&L Chapt. 7
Overview
Parallel Reactions
A+BR (desired)
AS
Series Reactions
ABC(desired)D
Independent Reactions
AB (desired)
CD+E
Series Parallel Reactions
A+BC+D
A+CE(desired)
Mixing, Temperature and Pressure Effects
Examples
Ethylene Oxide Synthesis
CH2=CH2 + 3O22CO2 + 2H2O

CH2=CH2 + O2CH2-CH2(desired)
Examples
Diethanolamine Synthesis

/ O\
CH 2 CH 2 NH 3 HOCH 2CH 2 NH 2
/ O\
CH 2 CH 2 HOCH 2CH 2 NH 2 ( HOCH 2CH 2 ) 2 NH (desired )
/ O\
CH 2 CH 2 ( HOCH 2CH 2 ) 2 NH ( HOCH 2CH 2 ) 3 N
Examples
Butadiene Synthesis, C4H6, from Ethanol

C2 H 5OH C2 H 4 H 2O
C2 H 5OH CH 3CHO H 2
C2 H 4 CH 3CHO C4 H 6 H 2O
Rate Selectivity
Parallel Reactions
A+BR (desired)
A+BS k D ( D U ) ( D U )
SD/U rD
rU CA CB
Rate Selectivity kU

(D- U) >1 make CA as large as possible


(D U)>1 make CB as large as possible

(kD/kU)= (koD/koU)exp[-(EA-D-EA-U)/(RT)]
EA-D > EA-U T
EA-D < EA-U T
Reactor Design to Maximize Desired
Product for Parallel Rxns.
Maximize Desired Product
Series Reactions
AB(desired)CD
Plug Flow Reactor
Optimum Time in Reactor
Fractional Yield
1
CH 3CH 2OH ( g ) O2 CH 3CHO H 2O
2
5
CH 3CHO O2 2CO2 2 H 2O
2

(k2/k1)=f(T)
Real Reaction Systems
More complicated than either
Series Reactions
Parallel Reactions
Effects of equilibrium must be considered
Confounding heat effects
All have Reactor Design Implications
Engineering Tricks
Reactor types
Multiple Reactors
Mixtures of Reactors
Bypass
Recycle after Separation
Split Feed Points/ Multiple Feed Points
Diluents
Temperature Management with interstage
Cooling/Heating
A few words about simulators
Aspen
Kinetics
Must put in with Aspen
ProMax
Units Reactor type and Kinetics must match!!

Equilibrium constants Kinetics

Selectable units
Must put in in the form
Equilibrium constants
lnK=A+B/T+CT+DT2

You might also like