Continuous stirred-tank reactor

Last updated
Diagram showing the setup of a continuous stirred-tank reactor Cstr.png
Diagram showing the setup of a continuous stirred-tank reactor

The continuous stirred-tank reactor (CSTR), also known as vat- or backmix reactor, mixed flow reactor (MFR), or a continuous-flow stirred-tank reactor (CFSTR), is a common model for a chemical reactor in chemical engineering and environmental 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.

Contents

The behavior of a CSTR is often approximated or modeled by that of an ideal CSTR, which assumes perfect mixing. In a perfectly mixed reactor, reagent is instantaneously and uniformly mixed throughout the reactor upon entry. Consequently, the output composition is identical to composition of the material inside the reactor, which is a function of residence time and reaction rate. The CSTR is the ideal limit of complete mixing in reactor design, which is the complete opposite of a plug flow reactor (PFR). In practice, no reactors behave ideally but instead fall somewhere in between the mixing limits of an ideal CSTR and PFR.

Ideal CSTR

Cross-sectional diagram of a CSTR Agitated vessel.svg
Cross-sectional diagram of a CSTR

Modeling

A continuous fluid flow containing non-conservative chemical reactant A enters an ideal CSTR of volume V.

Assumptions:

Integral mass balance on number of moles NA of species A in a reactor of volume V:

[1]

where,

Applying the assumptions of steady state and νA = −1, Equation 2 simplifies to:

The molar flow rates of species A can then be rewritten in terms of the concentration of A and the fluid flow rate (Q):

[2]

Equation 4 can then be rearranged to isolate rA and simplified:

[2]

where,

Residence time is the total amount of time a discrete quantity of reagent spends inside the reactor. For an ideal reactor, the theoretical residence time, , is always equal to the reactor volume divided by the fluid flow rate. [2] See the next section for a more in-depth discussion on the residence time distribution of a CSTR.

Depending on the order of the reaction, the reaction rate, rA, is generally dependent on the concentration of species A in the reactor and the rate constant. A key assumption when modeling a CSTR is that any reactant in the fluid is perfectly (i.e. uniformly) mixed in the reactor, implying that the concentration within the reactor is the same in the outlet stream. [3] The rate constant can be determined using a known empirical reaction rate that is adjusted for temperature using the Arrhenius temperature dependence. [2] Generally, as the temperature increases so does the rate at which the reaction occurs.

Equation 6 can be solved by integration after substituting the proper rate expression. The table below summarizes the outlet concentration of species A for an ideal CSTR. The values of the outlet concentration and residence time are major design criteria in the design of CSTRs for industrial applications.

Outlet Concentration for an Ideal CSTR
Reaction OrderCA
n=0
n=1 [1]
n=2
Other nNumerical solution required

Residence time distribution

Exit age distribution E(t) and cumulative age distribution F(t) functions for an ideal CSTR IdealCSTR.svg
Exit age distribution E(t) and cumulative age distribution F(t) functions for an ideal CSTR

An ideal CSTR will exhibit well-defined flow behavior that can be characterized by the reactor's residence time distribution, or exit age distribution. [4] Not all fluid particles will spend the same amount of time within the reactor. The exit age distribution (E(t)) defines the probability that a given fluid particle will spend time t in the reactor. Similarly, the cumulative age distribution (F(t)) gives the probability that a given fluid particle has an exit age less than time t. [3] One of the key takeaways from the exit age distribution is that a very small number of fluid particles will never exit the CSTR. [5] Depending on the application of the reactor, this may either be an asset or a drawback.

Non-ideal CSTR

While the ideal CSTR model is useful for predicting the fate of constituents during a chemical or biological process, CSTRs rarely exhibit ideal behavior in reality. [2] More commonly, the reactor hydraulics do not behave ideally or the system conditions do not obey the initial assumptions. Perfect mixing is a theoretical concept that is not achievable in practice. [6] For engineering purposes, however, if the residence time is 5–10 times the mixing time, the perfect mixing assumption generally holds true.

Exit age distribution E(t) and cumulative age distribution F(t) functions for a CSTR with dead space CSTR deadvolume.svg
Exit age distribution E(t) and cumulative age distribution F(t) functions for a CSTR with dead space

Non-ideal hydraulic behavior is commonly classified by either dead space or short-circuiting. These phenomena occur when some fluid spends less time in the reactor than the theoretical residence time, . The presence of corners or baffles in a reactor often results in some dead space where the fluid is poorly mixed. [6] Similarly, a jet of fluid in the reactor can cause short-circuiting, in which a portion of the flow exits the reactor much quicker than the bulk fluid. If dead space or short-circuiting occur in a CSTR, the relevant chemical or biological reactions may not finish before the fluid exits the reactor. [2] Any deviation from ideal flow will result in a residence time distribution different from the ideal distribution, as seen at right.

Modeling non-ideal flow

Although ideal flow reactors are seldom found in practice, they are useful tools for modeling non-ideal flow reactors. Any flow regime can be achieved by modeling a reactor as a combination of ideal CSTRs and plug flow reactors (PFRs) either in series or in parallel. [6] For examples, an infinite series of ideal CSTRs is hydraulically equivalent to an ideal PFR. [2] Reactor models combining a number of CSTRs in series are often termed tanks-in-series (TIS) models. [7]

To model systems that do not obey the assumptions of constant temperature and a single reaction, additional dependent variables must be considered. If the system is considered to be in unsteady-state, a differential equation or a system of coupled differential equations must be solved. Deviations of the CSTR behavior can be considered by the dispersion model. CSTRs are known to be one of the systems which exhibit complex behavior such as steady-state multiplicity, limit cycles, and chaos.

Cascades of CSTRs

A series of three CSTRs CSTR Cascade Schematic.jpg
A series of three CSTRs

Cascades of CSTRs, also known as a series of CSTRs, are used to decrease the volume of a system. [8]

Minimizing Volume

As the number of CSTRs in series increases, the total reactor volume decreases. Volume considerations for a cascade of CSTRs.jpg
As the number of CSTRs in series increases, the total reactor volume decreases.

As seen in the graph with one CSTR, where the inverse rate is plotted as a function of fractional conversion, the area in the box is equal to where V is the total reactor volume and is the molar flow rate of the feed. When the same process is applied to a cascade of CSTRs as seen in the graph with three CSTRs, the volume of each reactor is calculated from each inlet and outlet fractional conversion, therefore resulting in a decrease in total reactor volume. Optimum size is achieved when the area above the rectangles from the CSTRs in series that was previously covered by a single CSTR is maximized. For a first order reaction with two CSTRs, equal volumes should be used. As the number of ideal CSTRs (n) approaches infinity, the total reactor volume approaches that of an ideal PFR for the same reaction and fractional conversion.

Ideal Cascade of CSTRs

From the design equation of a single CSTR where , we can determine that for a single CSTR in series that

where is the space time of the reactor, is the feed concentration of A, is the outlet concentration of A, and is the rate of reaction of A.

First order

For an isothermal first order, constant density reaction in a cascade of identical CSTRs operating at steady state

For one CSTR: , where k is the rate constant and is the outlet concentration of A from the first CSTR

Two CSTRs: and

Plugging in the first CSTR equation to the second:

Therefore for m identical CSTRs in series:

When the volumes of the individual CSTRs in series vary, the order of the CSTRs does not change the overall conversion for a first order reaction as long as the CSTRs are run at the same temperature.

Zeroth order

At steady state, the general equation for an isothermal zeroth order reaction at in a cascade of CSTRs is given by

When the cascade of CSTRs is isothermal with identical reactors, the concentration is given by

Second order

For an isothermal second order reaction at steady state in a cascade of CSTRs, the general design equation is

Non-ideal cascade of CSTRs

With non-ideal reactors, residence time distributions can be calculated. At the concentration at the jth reactor in series is given by

where n is the total number of CSTRs in series, and is the average residence time of the cascade given by where Q is the volumetric flow rate.

From this, the cumulative residence time distribution (F(t)) can be calculated as

As n → ∞, F(t) approaches the ideal PFR response. The variance associated with F(t) for a pulse stimulus into a cascade of CSTRs is .

Cost

Cost initially decreases with the number of CSTRs as volume decreases but as operational costs increase, the total cost eventually begins to increase. Relationship between cost and the number of CSTRs.jpg
Cost initially decreases with the number of CSTRs as volume decreases but as operational costs increase, the total cost eventually begins to increase.

When determining the cost of a series of CSTRs, capital and operating costs must be taken into account. As seen above, an increase in the number of CSTRs in series will decrease the total reactor volume. Since cost scales with volume, capital costs are lowered by increasing the number of CSTRs. The largest decrease in cost, and therefore volume, occurs between a single CSTR and having two CSTRs in series. When considering operating cost, operating cost scales with the number of pumps and controls, construction, installation, and maintenance that accompany larger cascades. Therefore as the number of CSTRs increases, the operating cost increases. Therefore, there is a minimum cost associated with a cascade of CSTRs.

Zeroth order reactions

From a rearrangement of the equation given for identical isothermal CSTRs running a zeroth order reaction: , the volume of each individual CSTR will scale by . Therefore the total reactor volume is independent of the number of CSTRs for a zeroth order reaction. Therefore, cost is not a function of the number of reactors for a zeroth order reaction and does not decrease as the number of CSTRs increases.

Selectivity of parallel reactions

When considering parallel reactions, utilizing a cascade of CSTRs can achieve greater selectivity for a desired product.

For a given parallel reaction and with constants and and rate equations and , respectively, we can obtain a relationship between the two by dividing by . Therefore . In the case where and B is the desired product, the cascade of CSTRs is favored with a fresh secondary feed of in order to maximize the concentration of .


For a parallel reaction with two or more reactants such as and with constants and and rate equations and , respectively, we can obtain a relationship between the two by dividing by . Therefore . In the case where and and B is the desired product, a cascade of CSTRs with an inlet stream of high and is favored. In the case where and and B is the desired product, a cascade of CSTRs with a high concentration of in the feed and small secondary streams of is favored. [9]

Series reactions such as also have selectivity between and but CSTRs in general are typically not chosen when the desired product is as the back mixing from the CSTR favors . Typically a batch reactor or PFR is chosen for these reactions.

Applications

CSTRs facilitate rapid dilution of reagents through mixing. Therefore, for non-zero-order reactions, the low concentration of reagent in the reactor means a CSTR will be less efficient at removing the reagent compared to a PFR with the same residence time. [3] Therefore, CSTRs are typically larger than PFRs, which may be a challenge in applications where space is limited. However, one of the added benefits of dilution in CSTRs is the ability to neutralize shocks to the system. As opposed to PFRs, the performance of CSTRs is less susceptible to changes in the influent composition, which makes it ideal for a variety of industrial applications:

Anaerobic digesters at Newtown Creek Wastewater Treatment Plant in Greenpoint, Brooklyn Kingsland Avenue 3 digester tanks jeh.jpg
Anaerobic digesters at Newtown Creek Wastewater Treatment Plant in Greenpoint, Brooklyn

Environmental engineering

Chemical engineering

See also

Notes

    Related Research Articles

    In a chemical reaction, chemical equilibrium is the state in which both the reactants and products are present in concentrations which have no further tendency to change with time, so that there is no observable change in the properties of the system. This state results when the forward reaction proceeds at the same rate as the reverse reaction. The reaction rates of the forward and backward reactions are generally not zero, but they are equal. Thus, there are no net changes in the concentrations of the reactants and products. Such a state is known as dynamic equilibrium.

    <span class="mw-page-title-main">Fick's laws of diffusion</span> Mathematical descriptions of molecular diffusion

    Fick's laws of diffusion describe diffusion and were first posited by Adolf Fick in 1855 on the basis of largely experimental results. They can be used to solve for the diffusion coefficient, D. Fick's first law can be used to derive his second law which in turn is identical to the diffusion equation.

    <span class="mw-page-title-main">Reaction rate</span> Speed at which a chemical reaction takes place

    The reaction rate or rate of reaction is the speed at which a chemical reaction takes place, defined as proportional to the increase in the concentration of a product per unit time and to the decrease in the concentration of a reactant per unit time. Reaction rates can vary dramatically. For example, the oxidative rusting of iron under Earth's atmosphere is a slow reaction that can take many years, but the combustion of cellulose in a fire is a reaction that takes place in fractions of a second. For most reactions, the rate decreases as the reaction proceeds. A reaction's rate can be determined by measuring the changes in concentration over time.

    <span class="mw-page-title-main">Dispersity</span> Measure of heterogeneity of particle or molecular sizes

    In chemistry, the dispersity is a measure of the heterogeneity of sizes of molecules or particles in a mixture. A collection of objects is called uniform if the objects have the same size, shape, or mass. A sample of objects that have an inconsistent size, shape and mass distribution is called non-uniform. The objects can be in any form of chemical dispersion, such as particles in a colloid, droplets in a cloud, crystals in a rock, or polymer macromolecules in a solution or a solid polymer mass. Polymers can be described by molecular mass distribution; a population of particles can be described by size, surface area, and/or mass distribution; and thin films can be described by film thickness distribution.

    The Damköhler numbers (Da) are dimensionless numbers used in chemical engineering to relate the chemical reaction timescale to the transport phenomena rate occurring in a system. It is named after German chemist Gerhard Damköhler. The Karlovitz number (Ka) is related to the Damköhler number by Da = 1/Ka.

    Hemorheology, also spelled haemorheology, or blood rheology, is the study of flow properties of blood and its elements of plasma and cells. Proper tissue perfusion can occur only when blood's rheological properties are within certain levels. Alterations of these properties play significant roles in disease processes. Blood viscosity is determined by plasma viscosity, hematocrit and mechanical properties of red blood cells. Red blood cells have unique mechanical behavior, which can be discussed under the terms erythrocyte deformability and erythrocyte aggregation. Because of that, blood behaves as a non-Newtonian fluid. As such, the viscosity of blood varies with shear rate. Blood becomes less viscous at high shear rates like those experienced with increased flow such as during exercise or in peak-systole. Therefore, blood is a shear-thinning fluid. Contrarily, blood viscosity increases when shear rate goes down with increased vessel diameters or with low flow, such as downstream from an obstruction or in diastole. Blood viscosity also increases with increases in red cell aggregability.

    <span class="mw-page-title-main">Lawson criterion</span> Criterion for igniting a nuclear fusion chain reaction

    The Lawson criterion is a figure of merit used in nuclear fusion research. It compares the rate of energy being generated by fusion reactions within the fusion fuel to the rate of energy losses to the environment. When the rate of production is higher than the rate of loss, the system will produce net energy. If enough of that energy is captured by the fuel, the system will become self-sustaining and is said to be ignited.

    The equilibrium constant of a chemical reaction is the value of its reaction quotient at chemical equilibrium, a state approached by a dynamic chemical system after sufficient time has elapsed at which its composition has no measurable tendency towards further change. For a given set of reaction conditions, the equilibrium constant is independent of the initial analytical concentrations of the reactant and product species in the mixture. Thus, given the initial composition of a system, known equilibrium constant values can be used to determine the composition of the system at equilibrium. However, reaction parameters like temperature, solvent, and ionic strength may all influence the value of the equilibrium constant.

    <span class="mw-page-title-main">Chemical reactor</span> Enclosed volume where interconversion of compounds takes place

    A chemical reactor is an enclosed volume in which a chemical reaction takes place. In chemical engineering, it is generally understood to be a process vessel used to carry out a chemical reaction, which is one of the classic unit operations in chemical process analysis. The design of a chemical reactor deals with multiple aspects of chemical engineering. Chemical engineers design reactors to maximize net present value for the given reaction. Designers ensure that the reaction proceeds with the highest efficiency towards the desired output product, producing the highest yield of product while requiring the least amount of money to purchase and operate. Normal operating expenses include energy input, energy removal, raw material costs, labor, etc. Energy changes can come in the form of heating or cooling, pumping to increase pressure, frictional pressure loss or agitation.

    Conversion and its related terms yield and selectivity are important terms in chemical reaction engineering. They are described as ratios of how much of a reactant has reacted (X — conversion, normally between zero and one), how much of a desired product was formed (Y — yield, normally also between zero and one) and how much desired product was formed in ratio to the undesired product(s) (S — selectivity).

    In the physical sciences, relaxation usually means the return of a perturbed system into equilibrium. Each relaxation process can be categorized by a relaxation time τ. The simplest theoretical description of relaxation as function of time t is an exponential law exp(−t/τ).

    The diffusion of plasma across a magnetic field was conjectured to follow the Bohm diffusion scaling as indicated from the early plasma experiments of very lossy machines. This predicted that the rate of diffusion was linear with temperature and inversely linear with the strength of the confining magnetic field.

    In physics, a mass balance, also called a material balance, is an application of conservation of mass to the analysis of physical systems. By accounting for material entering and leaving a system, mass flows can be identified which might have been unknown, or difficult to measure without this technique. The exact conservation law used in the analysis of the system depends on the context of the problem, but all revolve around mass conservation, i.e., that matter cannot disappear or be created spontaneously.

    <span class="mw-page-title-main">Plug flow reactor model</span> Reactor simulation model

    The plug flow reactor model is a model used to describe chemical reactions in continuous, flowing systems of cylindrical geometry. The PFR model is used to predict the behavior of chemical reactors of such design, so that key reactor variables, such as the dimensions of the reactor, can be estimated.

    <span class="mw-page-title-main">Sediment transport</span> Movement of solid particles, typically by gravity and fluid entrainment

    Sediment transport is the movement of solid particles (sediment), typically due to a combination of gravity acting on the sediment, and the movement of the fluid in which the sediment is entrained. Sediment transport occurs in natural systems where the particles are clastic rocks, mud, or clay; the fluid is air, water, or ice; and the force of gravity acts to move the particles along the sloping surface on which they are resting. Sediment transport due to fluid motion occurs in rivers, oceans, lakes, seas, and other bodies of water due to currents and tides. Transport is also caused by glaciers as they flow, and on terrestrial surfaces under the influence of wind. Sediment transport due only to gravity can occur on sloping surfaces in general, including hillslopes, scarps, cliffs, and the continental shelf—continental slope boundary.

    The Herschel–Bulkley fluid is a generalized model of a non-Newtonian fluid, in which the strain experienced by the fluid is related to the stress in a complicated, non-linear way. Three parameters characterize this relationship: the consistency k, the flow index n, and the yield shear stress . The consistency is a simple constant of proportionality, while the flow index measures the degree to which the fluid is shear-thinning or shear-thickening. Ordinary paint is one example of a shear-thinning fluid, while oobleck provides one realization of a shear-thickening fluid. Finally, the yield stress quantifies the amount of stress that the fluid may experience before it yields and begins to flow.

    <span class="mw-page-title-main">Bubble column reactor</span>

    A bubble column reactor is a chemical reactor that belongs to the general class of multiphase reactors, which consists of three main categories: trickle bed reactor, fluidized bed reactor, and bubble column reactor. A bubble column reactor is a very simple device consisting of a vertical vessel filled with water with a gas distributor at the inlet. Due to the ease of design and operation, which does not involve moving parts, they are widely used in the chemical, biochemical, petrochemical, and pharmaceutical industries to generate and control gas-liquid chemical reactions.

    For both chemical and biological engineering, Semibatch (semiflow) reactors operate much like batch reactors in that they take place in a single stirred tank with similar equipment. However, they are modified to allow reactant addition and/or product removal in time.

    A Levenspiel plot is a plot used in chemical reaction engineering to determine the required volume of a chemical reactor given experimental data on the chemical reaction taking place in it. It is named after the late chemical engineering professor Octave Levenspiel.

    The residence time of a fluid parcel is the total time that the parcel has spent inside a control volume (e.g.: a chemical reactor, a lake, a human body). The residence time of a set of parcels is quantified in terms of the frequency distribution of the residence time in the set, which is known as residence time distribution (RTD), or in terms of its average, known as mean residence time.

    References

    1. 1 2 Schmidt, Lanny D. (1998). The Engineering of Chemical Reactions. New York: Oxford University Press. ISBN   0-19-510588-5.
    2. 1 2 3 4 5 6 7 8 9 Metcalf & Eddy (2013-09-03). Wastewater engineering : treatment and resource recovery. Tchobanoglous, George,, Stensel, H. David,, Tsuchihashi, Ryujiro,, Burton, Franklin L. (Franklin Louis), 1927-, Abu-Orf, Mohammad,, Bowden, Gregory (Fifth ed.). New York, NY. ISBN   978-0-07-340118-8. OCLC   858915999.{{cite book}}: CS1 maint: location missing publisher (link)
    3. 1 2 3 Benjamin, Mark M. (2013-06-13). Water quality engineering : physical/chemical treatment processes. Lawler, Desmond F. Hoboken, New Jersey. ISBN   978-1-118-63227-7. OCLC   856567226.{{cite book}}: CS1 maint: location missing publisher (link)
    4. Bolin, Bert; Rodhe, Henning (January 1973). "A note on the concepts of age distribution and transit time in natural reservoirs". Tellus. 25 (1): 58–62. Bibcode:1973Tell...25...58B. doi:10.3402/tellusa.v25i1.9644. ISSN   0040-2826.
    5. Monsen, Nancy E.; Cloern, James E.; Lucas, Lisa V.; Monismith, Stephen G. (September 2002). "A comment on the use of flushing time, residence time, and age as transport time scales". Limnology and Oceanography. 47 (5): 1545–1553. Bibcode:2002LimOc..47.1545M. doi: 10.4319/lo.2002.47.5.1545 . S2CID   11505988.
    6. 1 2 3 Davis, Mark E. (2003). Fundamentals of chemical reaction engineering. Davis, Robert J. (International ed.). Boston: McGraw-Hill. ISBN   978-1-62870-437-2. OCLC   880604539.
    7. Stokes, R. L.; Nauman, E. Bruce (1970). "Residence Time Distribution Functions for Stirred Tanks in Series". Canadian Journal of Chemical Engineering. 48 (6): 723–725.
    8. Hill, Charles G.; Root, Thatcher W. (2014). Introduction to Chemical Engineering Kinetics and Reactor Design, Second Edition. Hoboken, new Jersey: Wiley. pp. 241–253, 349–358. ISBN   9781118368251.
    9. Levenspiel, Octave (1998). Chemical Reaction Engineering, 3rd Edition. Wiley. ISBN   978-0-471-25424-9.
    10. Hurtado, F.J.; Kaiser, A.S.; Zamora, B. (March 2015). "Fluid dynamic analysis of a continuous stirred tank reactor for technical optimization of wastewater digestion". Water Research. 71: 282–293. Bibcode:2015WatRe..71..282H. doi:10.1016/j.watres.2014.11.053. ISSN   0043-1354. PMID   25635665.
    11. Kadlec, Robert H.; Wallace, Scott D. (2009). Treatment Wetlands (second ed.). Boca Raton, FL, USA: CRC Press. p. 181. ISBN   978-1-56670-526-4.
    12. 1 2 "Visual Encyclopedia of Chemical Engineering". encyclopedia.che.engin.umich.edu. Archived from the original on 2013-12-14. Retrieved 2020-04-30.