You are on page 1of 37

Models of Imperfectly Mixed Reactors

MODELS OF IMPERFECTLY MIXED REACTORS

Not all tank reactors are perfectly mixed nor do all tubular reactors exhibit plug-flow behavior. In these situations,
some means must be used to allow for deviations from ideal behavior.

When constructing a flow model for a given chemical reactor, one starts by knowledge of the pattern of fluid passage
through the reactor. This flow behaviour could be determined by finding the complete history of each fluid element. It was
poited out that instead of this complexity of flow patterns, it is enough to know how long the fluid element stays in the reactor,
in other words, to determine the residence time distribution (RTD) of the fluid particles in the exit streams.

The residence time of a fluid element is the time that elapses from the time the element enters the vessel to the time
it leaves. The age of the fluid element at a given instant is the time that elapses between the elements entrance into the vessel
and the given instant. The residence distribution time (RDT) is the age distribution frequency of the fluid elements leaving the
vessel.

RTD is sufficient if the reaction is first order or if the fluid is either in a state of complete segregation or maximum
mixedness. We use the segregation and maximum mixedness models to bound the conversion when no adjustable parameters
are use. For non-first-order reactions in a fluid with good micromixing, more than just the RTD is needed. These situations
compose a number of reactor analysis problems and cannot be ignored. For example, we may have an existing reactor in
storage and want to carry out a new reaction in the reactor. To predict conversions and product distribution for such systems,
a model of reactor flow patterns and/or RTD is necessary.

Using the models, we will first measure the RTD to characterize the reactor at the new operating conditions of
temperature and flow rate. After selecting a model for the reactor, we use the RTD to evaluate the parameter(s) in the model
after which we calculate the conversion.

GUIDELINES FOR DEVELOPING MODELS

The choice of the particular model to be used depends largely on the engineering judgment of the person carrying out
the analysis. It is this person’s job to choose the model that best combines the conflicting goals of mathematical simplicity and
physical realism.

For a given real reactor, it is not uncommon to use all the models to predict conversion and then make comparisons. Usually,
the real conversion will be bounded by the model calculations.

The following guidelines are suggested when developing models for non-ideal reactors:

1. The model must be mathematically tractable. The equations used to describe a chemical reactor should be able to be
solved without an inordinate expenditure of human or computer time.

2. The model must realistically describe the characteristics of the non-ideal reactor. The phenomena occurring in the
nonideal reactor must be reasonably described physically, chemically, and mathematically.

3. The model should not have more than two adjustable parameters. This constraint is often used because an expression
with more than two adjustable parameters can be fitted to a great variety of experimental data, and the modeling process in
this circumstance is nothing more than an exercise in curve fitting. A one-parameter model is, of course, superior to a two-
parameter model if the one-parameter model is sufficiently realistic. To be fair, however, in complex systems (e.g., internal
diffusion and conduction, mass transfer limitations) where other parameters may be measured independently, then more
than two parameters are quite acceptable.

The reactors treated thus far—the perfectly mixed batch, the plug-flow tubular, the packed bed, and the perfectly
mixed continuous tank reactors have been modeled as ideal reactors. Unfortunately, in the real world we often observe
behavior very different from that expected from the exemplar; this behavior is true of students, engineers, college professors,
and chemical reactors. Just as we must learn to work with people who are not perfect, so the reactor analyst must learn to
diagnose and handle chemical reactors whose performance deviates from the ideal. Non-ideal reactors and the principles
behind their analysis form the subject of this chapter and the next.
Two types of ideal flow are commonly used as limits of flow patterns in process vessels; these are the "plug flow" and
the "perfectly mixed" flow.

The conditions for the physical realization of the plug flow are fulfilled in a piston-type flow, when it is assumed that
there is no mixing takes place in the direction of the flow. The model is employed to describe tubular apparatus with a large
length to diameter ratio. At the other extremes, perfect mixing assumes that the vessel contents are completely homogeneous
and the outlet stream properties are identical to the vessel-fluid properties. In chemical engineering, the usual tendency is to
come closer to conditions of perfect mixing by fitting apparatus with special mixers, baffles, etc.

Non-idealities of flow in industrial apparatus can be traced to the following most important reasons:

• Presence of dead spaces


• Channeling or by-passing
• Recycling or cross-flow streams
• Develop turbulence,etc.

Stagnant fluids or dead spaces - represents regions with extremely poor contacting. A dead space will contain fluid elements
for interval of time with an order of magnitude over the mean residence time.

Bypassing or channeling - is when some of the fluids slip or pass through the vessel considerably faster than others do.
Bypassing may be found in flow through poorly packed vessels, through heat exchanger in two-phase operations etc.

Recycling - a certain amount of fluid is recirculated or returned to the vessel inlet. This type of flow may be desirable for
example in auto-thermal reactions.

Overall three somewhat interrelated factors make up the contacting or flow pattern:

• the RTD or residence time distribution of material which is flowing through the vessel
• the state of aggregation of the flowing material, its tendency to clump and for a group of molecules to move about
together
• the earliness and lateness of mixing of material in the vessel.

The Residence Time Distribution, RTD

Deviation from the two ideal flow patterns can be caused by channeling of fluid, by recycling of fluid, or by creation of
stagnant regions in the vessel. Figure 11.1 shows this behavior. In all types of process equipment, such as heat exchangers,
packed columns, and reactors, this type of flow should be avoided since it always lowers the performance of the unit. If we
know precisely what is happening within the vessel, thus if we have a complete velocity distribution map for the fluid in the
vessel, then we should, in principle, be able to predict the behavior of a vessel as a reactor. Unfortunately, this approach is
impractical, even in today's computer age.
Setting aside this goal of complete knowledge about the flow, let us be less ambitious and see what it is that we
actually need to know. In many cases we really do not need to know very much, simply how long the individual molecules stay
in the vessel, or more precisely, the distribution of residence times of the flowing fluid. This information can be determined
easily and directly by a widely used method of inquiry, the stimulus-response experiment.

State of Aggregation of the Flowing Stream

Flowing material is in some particular state of aggregation, depending on its nature. In the extremes these states can be called
microfluids and macrofluids, as sketched in Fig. 11.2.

Single-Phase Systems. These lie somewhere between the extremes of macro- and microfluids.

Two-Phase Systems. A stream of solids always behaves as a macrofluid, but for gas reacting with liquid, either phase can be a
macro- or microfluid depending on the contacting scheme being used. The sketches of Fig. 11.3 show completely opposite
behavior. We treat these two phase reactors in later chapters.

Earliness of Mixing

The fluid elements of a single flowing stream can mix with each other either early or late in their flow through the vessel. For
example, see Fig. 11.4. Usually this factor has little effect on overall behavior for a single flowing fluid. However, for a system
with two entering reactant streams it can be very important. For example, see Fig. 11.5.
17.2 Zero-Adjustable-Parameter Models

17.2.1 Segregation Model

In a “perfectly mixed” CSTR, the entering fluid is assumed to be distributed immediately and evenly throughout the
reacting mixture. This mixing is assumed to take place even on the microscale, and elements of different ages mix together
thoroughly to form a completely micromixed fluid (see Figure 17-1(b)). However, if fluid elements of different ages do not mix
together at all, the elements remain segregated from each other, and the fluid is termed completely segregated (see Figure
17-1(a)). The extremes of complete micromixing and complete segregation are the limits of the micromixing of a reacting
mixture.

17.2.2 Maximum Mixedness Model

In a reactor with a segregated fluid, mixing between particles of fluid does not occur until the fluid leaves the reactor.
The reactor exit is, of course, the latest possible point where mixing can occur, and any effect of mixing is postponed until after
all reaction has taken place. We can also think of a completely segregated flow as being in a state of minimum mixedness. We
now want to consider the other extreme, that of maximum mixedness consistent with a given residence time distribution. As
soon as the fluid enters the reactor, it is completely mixed radially (but not longitudinally) with the other fluid already in the
reactor. The entering fluid is fed into the reactor through the side entrances in such a manner that the RTD of the plug-flow
reactor with side entrances is identical to the RTD of the real reactor.

In the reactor with side entrances, mixing occurs at the earliest possible moment consistent with the RTD. This
situation is termed the condition of maximum mixedness.

18.1.1 One-Parameter Models

Here, we use a single parameter to account for the non-ideality of our reactor. This parameter is most always evaluated
by analyzing the RTD determined from a tracer test. Examples of one-parameter models for non-ideal CSTRs include either a
reactor dead volume, VD, where no reaction takes place, or volumetric flow rate with part of the fluid bypassing the reactor,
υb, thereby exiting unreacted. Examples of one-parameter models for tubular reactors include the tanks-in-series model and
the dispersion model. For the tanks-in-series model, the one parameter is the number of tanks, n, and for the dispersion
model, the one parameter is the dispersion coefficient, Da. Knowing the parameter values, we then proceed to determine the
conversion and/or effluent concentrations for the reactor.
18.1.1.1 The Tanks-in-Series (T-I-S) One-Parameter Model

The T-I-S model is a one-parameter model. We will analyze the RTD to determine the number of ideal tanks, n, in series that
will give approximately the same RTD as the non- ideal reactor.

The number of tank in series is

This expression represents the number of tanks necessary to model the real reactor as n ideal tanks in series. If the number of
reactors, n, turns out to be small, the reactor characteristics turn out to be those of a single CSTR or perhaps two CSTRs in
series. At the other extreme, when n turns out to be large, the reactor characteristics approach those of a PFR.

18.1.1.2 Dispersion One-Parameter Model

The dispersion model is also often used to describe non-ideal tubular reactors. In this model, there is an axial dispersion of the
material, which is governed by an analogy to Fick’s law of diffusion, superimposed on the flow as shown in Figure 18-4. So in
addition to transport by bulk flow, UAcC, every component in the mixture is transported through any cross section of the
reactor at a rate equal to [–DaAc(dC/dz)] resulting from molecular and convective diffusion. By convective diffusion (i.e.,
dispersion), we mean either Aris-Taylor dispersion in laminar-flow reactors or turbulent diffusion resulting from turbulent
eddies. Radial concentration profiles for plug flow (a) and a representative axial and radial profile for dispersive flow (b) are
shown in Figure 18-4. Some molecules will diffuse forward ahead of the molar average velocity, while others will lag behind.

18.1.2 Two-Parameter Models

The premise for the two-parameter model is that we can use a combination of ideal reactors to model the real reactor. For
example, consider a packed bed reactor with channeling. Here, the response to a pulse tracer input would show two dispersed
pulses in the output as shown in Figure 16-1 and Figure 18-1.

Here, we could model the real reactor as two ideal PBRs in parallel, with the two parameters being the volumetric flow rate
that channels or by passes, υb , and the reactor dead volume, VD. The real reactor volume is V = VD + VS with entering
volumetric flow rate υ0 = υb + υS.

Non-Isothermal Reactor
Isothermal conditions are most useful for the measurement of kinetic data, real reactor operation is normally
nonisothennal. Within the limits of heat exchange, the reactor can operateisothermally (maximum heat exchange) or
adiabatically (no heat exchange).

The three types of reactor operations yield different temperature profiles within the reactor and are
illustrated in Figure 9.1.1 for an exothermic reaction.

(9.1.1)

Since the reaction rate expression now contains the independent variable T, the material balance cannot be
solved alone. The solution of the material balance equation is only possible by the simultaneous solution of the
energy balance. Thus, for nonisothennal reactor descriptions, an energy balance must accompany the material
balance.

Energy Balances
Applying the first law of thermodynamics to the reactor shown in Figure 9.2.1, the following is obtained:
“W" is "shaft work," that is, pump or turbine work, “U” is the internal energy of stream i, “m” is the mass of
stream i, “P” is the pressure of stream i, “V” is the volume of mass i, “U” is the velocity of stream i, “Z” is the
height of stream i above a datum plane, and gc is the gravitational constant, for i denoting either the outlet or inlet
stream. For most normal circumstances:

(9.2.2)

or

Using these assumptions to simplify Equation (9.2.1) yields:

(9.2.3)

Recall that the enthalpy, H, is:

(9.2.4)

SolvingThus, Equation (9.2.3) can be written as: for Tc and then substituting the expression back into the heat
transfer equation yields:

(9.2.5)

However, since it is more typical to deal with rates of energy transfer, Q, rather than energy, Q, when dealing with
reactors, Equation (9.2.5) can be differentiated with respect to time to give:
(9.2.6)

where Q is the rate of heat transfer, “h” is the enthalpy per unit mass of stream i, and “m” is the mass flow rate of
stream i. Generalizing Equation (9.2.6) to multi-input, multi-output reactors yields a generalized energy balance
subject to the assumptions stated above:

(9.2.7)

(9.2.8)

For a closed reactor (e.g., a batch reactor), the


potential and kinetic energy terms in Equation (9.2.1) are not relevant. Additionally, Ws ~ 0 for most cases (including the
work input from the stirring impellers). Since most reactions carried out in closed reactors involve condensed phases,
delta(PV) is small relative to delta(U), and for this case is:

(9.2.9)

(9.2.10)

Nonisothermal Batch Reactor

Consider the batch reactor schematically illustrated in Figure 9.3.1. Typically, reactants are charged into the
reactor from point (1), the temperature of the reactor is increased by elevating the temperature in the heat transfer
fluid, a temperature maximum is reached, the reactor is then cooled by decreasing the temperature of the heat
transfer fluid and products discharged via point (II). To describe this process, material and energy balances are
required. Recall that the mass balance on a batch reactor can be written as [refer to Equation (3.2.1)]:

(9.3.1)

(9.3.2)

To solve the mass balance, it must be accompanied by the simultaneous solution of the energy balance (i.e.,
the solution ofEquation (9.2.9». To do this, Equation (9.2.9) can be written in more convenient forms. Consider that
the enthalpy contains both sensible heat and heat of reaction effects. That is to say that Equation (9.2.10) can be
written as:

(9.3.2)

(9.3.2)
where <I> is the extent of reaction (see Chapter 1),delta(H) is the heat of reaction, MS is the total mass of the
system, and (Cp) is an average heat capacity per unit mass for the system. Since enthalpy is a state variable, the
solution of the integral in Equation (9.3.4) is path independent. Referring to Figure 9.3.2, a simple way to analyze
the integral [pathway (I)] is to allow the reaction to proceed isothermally [pathway (II)] and then evaluate the
sensible heat changes using the product composition [pathway (III)]. That is,

(9.3.7)

where U is an overall heat transfer coefficient, AH is the heat transfer area, and T* is the reference temperature.
Combining Equations (9.3.7), (9.3.6), and (9.3.5) and recalling the definition of the extent of reaction in terms of
the fractional conversion [Equation (1.2.10)] gives:
Non- Isothermal Plug Flow Reactor
Consider a PFR operating at non isothermal conditions (refer to Figure 9.4.1). To describe the reactor
performance, the material balance, Equation (9.1.1), must be solved simultaneously with the energy balance, Equation
(9.2.7). Assuming that the PFR is a tubular reactor of constant cross-sectional area and that T and Ci do not vary over
the radial direction of the tube, the heat transfer rate Q can be written for a differential section of reactor volume as
(see Figure 9.4.1):

where dr is the diameter of the tubular reactor. Recall again that the enthalpy contains both sensible heat and heat of
reaction effects. Thus, the energy balance Equation (9.2.7) can be written for the differential fluid element of the PFR
as:
where the heat of reaction is evaluated at a reference temperature TO and CPi is the molar heat capacity of species i.
Normally, TO is taken as the reactor entrance temperature.

Example:
Butadiene and ethylene can be reacted together to form cyclohexene as follows:

If equimolar butadiene and ethylene at 450°C and 1 atm are fed to a PFR operating adiabatically,what is the
space time necessary to reach a fractional conversion of 0.10?

Answer
Assume that each CPi is not a strong function of temperature over the temperature range obtained within the
PFR (i.e., each is not a function of T). The material and energy balance equations are:
The energy balance for the PFR can also be written as follows by combining Equations (9.4.1) and (9.4.2):

with Cl= Cl0 and T = TO at z = 0, where u is the superficial linear velocity, ρ is the average density of the fluid, and
Cp is the average heat capacity per unit mass. Equation (9.4.9) is only applicable when the density of the fluid is not
changing in the reactor and when Cp is not a function of temperature, since the following relationship is used to obtain
Equation (9.4.9):

Example:
Answer:

TEMPERATURE EFFECTS IN A CONTINUOUS STIRRED TANK REACTOR

Although the assumption of perfect mixing in the CSTR implies that the reactor contents will be at uniform
temperature (and thus the exit stream will be at this temperature), the reactor inlet may not be at the same temperature
as the reactor. If this is the case and/or it is necessary to determine the heat transferred to or from the reactor, then an
energy balance is required.

The energy balance for a CSTR:

Where the superscript f denotes the final or outlet conditions. For adiabatic operation, Q= O.

EXAMPLE PROBLEM 1: The nitration of aromatic compounds is a highly exothermic reaction that generally uses
catalysts that tend to be corrosive(e.g., HN03/H2S04). A less corrosive reaction employs N205 as the nitrating agent as
illustrated below:

If this reaction is conducted in an adiabatic CSTR, what is the reactor volume and space time necessary to achieve 35
percent conversion of N2O5? The reaction rate is first order in A and second order in B.

SOLUTION:

The reaction occurs in the liquid-phase and the concentrations are dilute so that mole change with reaction does not
change the overall density of the reacting fluid. Thus,

The material balance on the CSTR can be written as:


The energy balance for the adiabatic CSTR is:

• For fA =0.35,
the energy
balance
yields T=554K.

• At 554 K, the value of the rate constant is k=119.8 L2/(mol2-min)

• Using this value of k in the material balance gives V = 8,500 L and thus T = 8.5 min.

STABILITY AND SENSITIVITY OF REACTORS ACCOMPLISHING EXOTHERMIC REACTIONS

The reactor is accomplishing an exothermic reaction and therefore must transfer heat to a cooling fluid in order to
remain at temperature T. Assume that the heat transfer is sufficiently high to maintain the reactor wall temperature at
Tc. Therefore,

(9.6.1)

Where Vc is the volumetric flow rate of the coolant and Cpc is the heat capacity of the coolant and is not a function of temperature.

Solving for Tc and then substituting the expression back into the heat transfer equation yields:

(9.6.2)

The energy balance on the CSTR can be written as [from Equation (9.5.1) with a first-order reaction rate expression]:
Or since:

As:

Where vP is the volumetric flow rate of the product stream and ρ and Cp are the average density and heat capacity of
the outlet stream. Rearranging Equation (9.6.3) and substituting Equation(9.6.2) for Q gives (note that Q is heat
removed):

• If Qr and Qg are plotted versus the reaction temperature, T, the results are illustrated in Figure 9.6.2.
• A solution of Equation(9.6.6) occurs when Qr=Qg, and this can happen as shown in Figure9.6.3.
• For cases (I) and (III) a single solution exists.
• However, for case (II) three steady-states are possible.
• At steady-state 1, if T is increased then Qr > Qg so the reactor will return to point 1. Additionally, if T is
decreased, Qg > Qr so the reactor will also return to point1 in this case. Thus, steady-state 1is stable since small
perturbations from this position cause the reactor to return to the steady-state.
• Likewise steady-state 3 is a stable steady-state. However, for steady-state 2, if T is increased, Qg>Qr and the
reactor will move to position 3.
• If T is decreased below that of point 2, Qr>Qg and the reactor will move to point 1. Therefore, steady-state 2
is unstable.
• It is important to determine the stability of reactor operation since perturbations from steady-state always
occur in a real system.

• Finally, what determines whether the reactor achieves steady-state 1or 3 is the start-up of the reactor.
In addition to knowing the existence of multiple steady-states and that some may be unstable, it is important
to assess how stable the reactor operation is to variations in the processing parameters (i.e., the sensitivity). In
the above example for the CSTR, it is expected that the stable steady-states have low sensitivity to variations
in the processing parameters, while clearly the unstable steady-state would not.

To provide an example of how the sensitivity may be elucidated, consider a tubular reactor accomplishing an
exothermic reaction and operating at non-isothermal conditions. As described in Example 9.4.3, hotspots in the reactor
temperature profile can occur for this situation. Figure9.6.4 illustrates what a typical reactor temperature profile would
look like.

• Consider what could happen as the temperature of the coolant fluid, T0c increases. As T0c becomes warmer
T0c (i+1)> T0c (1),i = 1,2,3, then less heat is removed from the reactor and the temperature at the reactor hotspot
increases in value.

• Eventually, heat is generated at a sufficiently high rate that it cannot be removed [illustrated for T0c (4)] such
that the hotspot temperature exceeds some physical limit (e.g., phase change of fluid, explosions or fire, the
catalyst melts, etc.) and this condition is called runaway.

• Thus, reactor operation close to a runaway point would not be prudent, and determining the sensitivity towards
the tendency of runaway a critical factor in the reactor analysis.

• Several criteria have been developed to assess the sensitivity of reactors; each involves the use of critical
assumptions.

Continuous-flow reactors are almost always operated at steady state. We will consider three types: the continuous-stirred
tank reactor (CSTR), the plug-flow reactor (PFR), and the packed-bed reactor (PBR). Detailed physical descriptions of these
reactors can be found in both the Professional Reference Shelf (PRS) for Chapter 1 and in the Visual Encyclopedia of Equipment,
encyclopedia.che.engin.umich.edu, and on the CRE Web site.

A. CONTINUOUS STIRRED TANK REACTOR (CSTR)


A type of reactor commonly used in industrial processing is the stirred tank operated continuously (Figure 1-7). It is
referred to as the continuous-stirred tank reactor (CSTR) or vat, or backmix reactor, and is primarily used for liquid-phase
reactions. It is normally operated at steady state and is assumed to be perfectly mixed; consequently, there is no time
dependence or position dependence of the temperature, concentration, or reaction rate inside the CSTR. That is, every
variable is the same at every point inside the reactor. Because the temperature and concentration are identical everywhere
within the reaction vessel, they are the same at the exit point as they are elsewhere in the tank. Thus, the temperature and
concentration in the exit stream are modeled as being the same as those inside the reactor. In systems where mixing is highly
non-ideal, the well-mixed model is inadequate, and we must resort to other modeling techniques, such as residence time
distributions to obtain meaningful results.

When the general mole balance equation

is applied to a CSTR operated at steady state (i.e., conditions do not change with time),

in which there are no spatial variations in the rate of reaction (i.e., perfect mixing),

it takes the familiar form known as the design equation for a CSTR

The CSTR design equation gives the reactor volume V necessary to reduce the entering flow rate of species j from Fj0 to the
exit flow rate Fj, when species j is disappearing at a rate of –rj. We note that the CSTR is modeled such that the conditions in
the exit stream (e.g., concentration and temperature) are identical to those in the tank. The molar flow rate Fj is just the
product of the concentration of species j and the volumetric flow rate
Similarly, for the entrance molar flow rate we have Fj0 = Cj0 · v0. Consequently, we can substitute for Fj0 and Fj into Equation
(1-7) to write a balance on species A as

The ideal CSTR mole balance equation is an algebraic equation, not a differential equation.

B. TUBULAR REACTOR
In addition to the CSTR and batch reactors, another type of reactor commonly used in industry is the tubular reactor.
It consists of a cylindrical pipe and is normally operated at steady state, as is the CSTR. Tubular reactors are used most often
for gas-phase reactions. A schematic and a photograph of industrial tubular reactors are shown in Figure 1-8.

In the tubular reactor, the reactants are continually consumed as they flow down the length of the reactor. In modeling
the tubular reactor, we assume that the concentration varies continuously in the axial direction through the reactor.
Consequently, the reaction rate, which is a function of concentration for all but zero-order reactions, will also vary axially. For
the purposes of the material presented here, we consider systems in which the flow field may be modeled by that of a plug-
flow profile (e.g., uniform velocity as in turbulent flow), as shown in Figure 1-9. That is, there is no radial variation in reaction
rate, and the reactor is referred to as a plug-flow reactor (PFR). (The laminar-flow reactor is discussed in Chapters 16 through
18 on non-ideal reactors.)

The equation we will use to design PFRs at steady state can be developed in two ways: (1) directly from Equation (1-4) by
differentiating with respect to volume V, and then rearranging the result or (2) from a mole balance on species j in a differential
segment of the reactor volume. Let’s choose the second way to arrive at the differential form of the PFR mole balance. The
differential volume, shown in Figure 1-10, will be chosen sufficiently small such that there are no spatial variations in reaction
rate within this volume. Thus the generation term, is
We could have made the cylindrical reactor on which we carried out our mole balance an irregularly shaped reactor, such as
the one shown in Figure 1-11 for reactant species A. However, we see that by applying Equation (1-10), the result would yield
the same equation (i.e., Equation (1-11)). For species A, the mole balance is

Consequently, we see that Equation (1-11) applies equally well to our model of tubular reactors of variable and
constant cross-sectional area, although it is doubtful that one would find a reactor of the shape shown in 1-11 unless it were
designed by Pablo Picasso.
The conclusion drawn from the application of the design equation to Picasso’s reactor is an important one: the degree
of completion of a reaction achieved in an ideal plug-flow reactor (PFR) does not depend on its shape, only on its total volume.
Again consider the isomerization A → B, this time in a PFR. As the reactants proceed down the reactor, A is consumed
by chemical reaction and B is produced. Consequently, the molar flow rate FA decreases, while FB increases as the reactor
volume V increases, as shown in Figure 1-12.

We now ask what is the reactor volume V1 necessary to reduce the entering molar flow rate of A from FA0 to FA1. Rearranging
Equation (1-12) in the form

V1 is the volume necessary to reduce the entering molar flow rate FA0 to some specified value FA1 and also the volume
necessary to produce a molar flow rate of B of FB1.
C. PACKED-BED REACTOR
The principal difference between reactor design calculations involving homogeneous reactions and those involving fluid-solid
heterogeneous reactions is that for the latter, the reaction takes place on the surface of the catalyst (see Chapter 10). The
greater the mass of a given catalyst, the greater the reactive surface area. Consequently, the reaction rate is based on mass of
solid catalyst, W, rather than on reactor volume, V. For a fluid–solid heterogeneous system, the rate of reaction of a species A
is defined as

The reactor volume that contains the catalyst is of secondary significance. Figure 1-13 shows a schematic of an industrial
catalytic reactor with vertical tubes packed with solid catalyst.
In the three idealized types of reactors just discussed (the perfectly mixed batch reactor, the plug-flow tubular reactor [PFR]),
and the perfectly mixed continuous-stirred tank reactor [CSTR]), the design equations (i.e., mole balances) were developed
based on reactor volume. The derivation of the design equation for a packed-bed catalytic reactor (PBR) will be carried out in
a manner analogous to the development of the tubular design equation. To accomplish this derivation, we simply replace the
volume coordinate in Equation (1-10) with the catalyst mass (i.e., weight) coordinate W (Figure 1-14).

As with the PFR, the PBR is assumed to have no radial gradients in concentration, temperature, or reaction rate. The
generalized mole balance on species A over catalyst weight ΔW results in the equation

which are, as expected, the same dimensions of the molar flow rate FA. After dividing by ΔW and taking the limit as ΔW → 0,
we arrive at the differential form of the mole balance for a packed-bed reactor

When pressure drop through the reactor (see Section 5.5) and catalyst decay (see Section 10.7 on the CRE Web site Chapter
10) are neglected, the integral form of the packed-catalyst-bed design equation can be used to calculate the catalyst weight
W is the catalyst weight necessary to reduce the entering molar flow rate of species A, FA0, down to a flow rate FA.

SAMPLE PROBLEM

1. Sketch the concentration profile.


2. Derive an equation relating the reactor volume to the entering and exiting concentrations of A, the rate constant k, and the
volumetric flow rate
3. Determine the reactor volume, V1, necessary to reduce the exiting concentration to 10% of the entering concentration, i.e.,
CA = 0.1CA0, when the volumetric flow rate is 10 dm3/min (i.e., liters/min) and the specific reaction rate, k, is 0.23

Solution
1. Sketch CA as a function of V. Species A is consumed as we move down the reactor, and as a result, both the molar flow rate
of A and the concentration of A will decrease as we move. Because the volumetric flow rate is constant, = , one can use
Equation (1-8) to obtain the concentration of A, CA = FA/ , and then by comparison with the Figure 1-12 plot, obtain the
concentration of A as a function of reactor volume, as shown in Figure E1-2.1.
REFERENCE
Fogler, H. Scott. Elements of Chemical Reaction Engineering. Fifth Edition. Pp 12-22.

The continuous stirred-tank reactor (CSTR), also known as vat- or backmix reactor, or a
continuous-flow stirred-tank reactor (CFSTR), is a common model for a chemical reactor in chemical
engineering. A CSTR often refers to a model used to estimate the key unit operation variables when
using a continuous agitated-tank reactor to reach a specified output. The mathematical model works
for all fluids: liquids, gases, and slurries.

Multiple reactions have both the size requirement and the distribution of reaction products
which are affected by the pattern of flow within the vessel. We may recall at this point that the
distinction between a single reaction and multiple reactions is that the single reaction requires only
one rate expression to describe its kinetic behavior whereas multiple reactions require more than one
rate expression.

Since multiple reactions are so varied in type and seem to have so little in common, we may
despair of finding general guiding principles for design.

Fortunately, this is not so because many multiple reactions can be considered to be


combinations of two primary types: parallel reactions and series reactions.

Let us consider the general approach and nomenclature. First of all, we find it more
convenient to deal with concentrations rather than conversions. Second, in examining product
distribution the procedure is to eliminate the time variable by dividing one rate equation by another.
We end up then with equations relating the rates of change of certain components with respect to
other components of the systems. Such relationships are relatively easy to treat. Thus, we use two
distinct analyses, one for determination of reactor size and the other for the study of product
distribution. The two requirements, small reactor size and maximization of desired product, may run
counter to each other. In such a situation an economic analysis will yield the best compromise. In
general, however, product distribution controls; consequently, this concerns primarily optimization
with respect to product distribution, a factor which plays no role in single reactions.

Advantages:

1. These are commonly used in industrial processing primarily in homogenous liquid phase flow
reactions where constant agitation is required they may be used in parallel,

2. When agitation and good mixing is required.

Disadvantages:

1. It is not recommended for high pressure reactions because of cost consideration. For high
pressure reactions it requires complex sealing arrangements for the agitator which increase the
initial as well as maintenance cost.

2. Conversion of these reactors is low due to this they are not preferred.

3. These reactors are not suited for high heat effect since availability of both heat transfer
coefficient and heat transfer per unit area is low.

APPLICATION
Continuous flow stirred-tank reactors are usually applied in waste water treatment processes. CSTRs
facilitate rapid dilution rates which make them resistant to both high pH and low pH volatile fatty
acid wastes. CSTRs are less efficient compared to other types of reactors as they require larger reactor
volumes to achieve the same reaction rate as other reactor models such as Plug Flow Reactors.
PARALLEL REACTIONS

The first reaction type is when the reactants form, not just the desired products, but also other
undesired products in parallel with the main reaction. We want to show here the implications of
parallel reactions, so we consider a simple batch isothermal reactor at constant volume:

Assuming first-order kinetics, we can express the change with time in the concentrations of
reactant A (CA) and products B (𝐶𝐵) and C (CC):

The kinetic rate constants are kB and kC. We can analytically solve these differential equations,
assuming that we start at time zero with only reactant A (𝐶𝐴𝑂):

The normalized concentrations as functions of time during the batch for different values of the
two rate constants kB and kC. The higher the value of kB compared

with kC, the more of product B is generated.

SELECTIVITY

• Selectivity quantifies the formation of desired wrt undesired products

SD /U = SD /U

• Instantaneous selectivity
S =
rate of formation desired = rD
of

D /U
rate formation undesired rU
of of
• Overall selectivity

S Exit molar rate of desired FD


= =

D /U
Exit molar rate of undesired FU
SIMULTANEOUS REACTIONS
Reactors in which simultaneous reactions occur are quite similar to the consecutive reactors case
in that yield of the desired product is important. But since the same reactants are involved in both
reactions, reactant concentrations cannot be as readily adjusted to affect yield. Yield of the
desired product depends primarily on reactor temperature and on the kinetics of the two reactions.

If the desired product is C, reactor temperature should be selected so that the specific reaction
rate k1 of the desired reaction is large compared to the specific reaction rate k2 of the undesired
reaction. If the orders of reactions are different in terms of their dependence on the two reactants,
the concentrations can also be adjusted to favor the desired reaction:

Equation 2.4

To illustrate some of these issues, we take a numerical example with the parameters given in
Equation 2.4. There are four components, so four component balances can be written as
follows:

Component A balance (kmol A/s):

Component B balance (kmol B/s):

Component C balance (kmol C/s):

Component D balance (kmol D/s):


Figure 2.20 gives results for a 200-m3 reactor with volumetric flowrates of 4.377 1023 m3 /s of
each fresh feed. The maximum concentration of C occurs at 343 K, but there is a significant
amount of D produced at this temperature. The optimum design may be a larger reactor operating
at a lower temperature so that the production of D is kept very small.
Continuous Stirred-Tank Reactor (CSTR) in Parallel Reactor

• Suppose we got n CSTR


• Same Size (Volume)
• Same Temperature of Operation
• Same “k” or constant rate
• Parallel Arrangement (independent of each other)

• For i reactors in parallel (see diagram),

• The volume of each individual reactor is related to the total volume of all reactors, you
need N tanks to get the total volume by

• Similarly,

• Therefore,

The conversion in any one reactor in parallel is identical to what would be achieved if the
reactant was fed to one large reactor of volume V.

FT= FAi + FAJ + FAJ = 3FAN


FT= nFAN

VT= VAi + VAJ + VAJ = 3VAN


VT= nVAN
SAMPLE PROBLEM

If 80% conversion is to be achieved, determine the necessary CSTR volume.

CONCLUSION

We have proved that the parallel arrangement would be the same if we would actually
have one large reactor of that volume.

CONSECUTIVE REACTIONS
Reactor design becomes more challenging when yield as well as conversion must be
considered. One common situation in which this arises is when there are consecutive irreversible
reactions such as the following:
A+B C
C+B D

where,
C = desired product
D = undesired product
Important industrial examples of this type:

▪ Chlorination
▪ Oxidation
▪ Nitration of a variety of hydrocarbons

The specific reaction rates for the first and second reactions are k1 and k2 respectively. The
“conversion” of reactant A is defined as
𝒎𝒐𝒍𝒆𝒔 𝑨 𝒇𝒆𝒅 − 𝒎𝒐𝒍𝒆𝒔 𝑨 𝒍𝒆𝒂𝒗𝒊𝒏𝒈 𝒓𝒆𝒂𝒄𝒕𝒐𝒓
𝑪𝒐𝒏𝒗𝒆𝒓𝒔𝒊𝒐 𝒏 𝒐𝒇 𝑨 =
𝒎𝒐𝒍𝒆𝒔 𝑨 𝒇𝒆𝒅
There are several ways to define “yield” or “selectivity” of the desired product C. One is the
basis of the amount of A fed. The other is on the basis of the amount of A that has reacted.
𝒎𝒐𝒍𝒆𝒔 𝑪 𝒑𝒓𝒐𝒅𝒖𝒄𝒆𝒅
𝒀𝒊𝒆𝒍𝒅 𝟏 =
𝒎𝒐𝒍𝒆𝒔 𝑨 𝒇𝒆𝒅
𝒎𝒐𝒍𝒆𝒔 𝑪 𝒑𝒓𝒐𝒅𝒖𝒄𝒆𝒅
𝒀𝒊𝒆𝒍𝒅 𝟐 =
𝒎𝒐𝒍𝒆𝒔 𝑨 𝒓𝒆𝒂𝒄𝒕𝒆𝒅

The desirable product C is produced by the first reaction whose rate depends on the
concentrations of A and B in the reactor. But C is consumed by the second reaction whose rate
depends on the concentrations of C and B in the reactor.
𝑹𝑪 = −𝒌𝟏 𝑪𝑨𝑪𝑩 + 𝒌𝟐 𝑪𝑪𝑪𝑩
𝑹𝑪 = −𝑪𝑨𝑪𝑩 𝒌𝟎𝟏𝒆−𝑬𝟏/𝑹𝑻𝑹 + 𝑪𝑪 𝑪𝑩 𝒌𝟎𝟐𝒆−𝑬𝟐/𝑹𝑻𝑹

𝑹𝑫 = −𝒌𝟐 𝑪𝑪𝑪𝑩
𝑹𝑫 = −𝑪𝑪 𝑪𝑩 𝒌𝟎𝟐𝒆−𝑬𝟐/𝑹𝑻𝑹

where the reaction rates RC and RD are rates of consumption.


With this design, the per-pass conversion of A will be small, but the yield of C per mole of A reacted
will be large. The downside of this design is that the excess A must be recovered and recycled, which
means high capital and energy costs. However, the resulting improvement in the yield of C is
typically well worth the added cost.
Environment and safety concerns have pushed the designs of many chemical processes to include
several large recycle streams so that the yields of desirable products are increased and the yields of
undesirable products are decreased. These recycle streams increase the difficulty of the plantwide
control problem.
There are four components, so four component balances can be written.
Component A balance (kmol A/s):

𝑭𝑨𝑶𝑪𝑨𝑶 = 𝑭𝑪𝑨 + 𝑽𝑹 𝒌𝟏𝑪𝑨𝑪𝑩


Component B balance (kmol B/s):

𝑭𝑩𝑶𝑪𝑩𝑶 = 𝑭𝑪𝑩 + 𝑽𝑹 (𝒌𝟏𝑪𝑨𝑪𝑩 + 𝒌𝟐 𝑪𝑪𝑪𝑩)


Component C balance (kmol C/s):

𝟎 = 𝑭𝑪𝑪 + 𝑽𝑹 (−𝒌𝟏𝑪𝑨𝑪𝑩 + 𝒌𝟐 𝑪𝑪𝑪𝑩)


Component D balance (kmol D/s):

𝟎 = 𝑭𝑪𝑫 − 𝑽𝑹 𝒌𝟐𝑪𝑪𝑪𝑩
The assumption of constant liquid densities means the volumetric flowrate of the reactor
effluent F is the sum of the two volumetric feed flowrates.
The steady-state economic design of a process with this type of reaction requires
consideration of the effects of reactor size and temperature on the entire plant. High recycle
flowrates of A and a large reactor operating at a low temperature will suppress the production
of D. but this will require a large capital investment in the reactor and separation sections of the
plant and consume significant energy.
SIMULTANEOUS REACTIONS
Reactors in which simultaneous reactions occur are quite similar to the consecutive reactors
case in that yield of the desired product is important. But since the same reactants are involved
in both reactions, reactant concentrations cannot be as readily adjusted to affect yield. Yield of
the desired product depends primarily on reactor temperature and on the kinetics of the two
reactions.
A+B C
C+B D

If the desired product is C, reactor temperature should be selected so that the specific reaction
rate k1 of the desired reaction is large compared to the specific reaction rate k2 of the undesired
reaction. If the orders of reactions are different in terms of their dependence on the two
reactants, the concentrations can also be adjusted to favor the desired reaction:
𝑹𝑨 = 𝑹𝑩 = (𝒌𝟏 + 𝒌𝟐 )𝑪𝑨𝑪𝑩
𝑬 𝑬
− 𝟏 − 𝟐

𝑹𝑨 = 𝑹𝑩 = 𝑪𝑨𝑪𝑩[𝒌𝟎𝟏𝒆 𝑹𝑻𝑹 + 𝒌
𝟎𝟐 𝒆 𝑹𝑻𝑹 ]

𝑹𝑪 = −𝒌𝟏 𝑪𝑨𝑪𝑩
𝑹𝑫 = −𝒌𝟐 𝑪𝑨𝑪𝑩

To illustrate some of these issues, we take a numerical example with the parameters given in
Table 2.4. There are four components, so four component balances can be written as follows:
Component A balance (kmol A/s):

𝑭𝑨𝑶𝑪𝑨𝑶 = 𝑭𝑪𝑨 + 𝑽𝑹 (𝒌𝟏 + 𝒌𝟐 )𝑪𝑨𝑪𝑩


Component B balance (kmol B/s):

𝑭𝑩𝑶𝑪𝑩𝑶 = 𝑭𝑪𝑩 + 𝑽𝑹 (𝒌𝟏 + 𝒌𝟐 )𝑪𝑨𝑪𝑩


Component C balance (kmol C/s):

𝟎 = 𝑭𝑪𝑪 − 𝑽𝑹𝒌𝟏𝑪𝑨𝑪𝑩
Component D balance (kmol D/s):
𝟎 = 𝑭𝑪𝑫 − 𝑽𝑹 𝒌𝟐𝑪𝑨𝑪𝑩

REFERENCE
Chemical Reactor Design. Retrieved from: www.academia.edu

You might also like